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

Boundedness of log Fano cone singularities and discreteness of local volumes

Chenyang Xu Department of Mathematics, Princeton University, Princeton, NJ 08544, USA [email protected]  and  Ziquan Zhuang Department of Mathematics, Johns Hopkins University, Baltimore, MD 21218, USA [email protected]
Abstract.

We prove that in any fixed dimension, K-semistable log Fano cone singularities whose volumes are bounded from below by a fixed positive number form a bounded set. As a consequence, we show that the set of local volumes of klt singularities of a fixed dimension has zero as the only accumulation point.

1. Introduction

In recent years, there has been remarkable progress in the algebro-geometric study of K-stability. Besides the global theory for Fano varieties, a local stability theory has also been introduced for Kawamata log terminal (klt) singularities, see [LLX-nv-survey, Z-SDC-survey] for an overview. Notably, the stable degeneration conjecture is settled in [XZ-SDC] (see also [Blu-minimizer-exist, LX-stability-higher-rank, Xu-quasi-monomial, XZ-minimizer-unique]). It provides for any klt singularity x(X=Spec(R),Δ)x\in(X={\rm Spec}(R),\Delta), a canonical degeneration to a K-semistable log Fano cone singularity x0(X0,Δ0;ξv)x_{0}\in(X_{0},\Delta_{0};\xi_{v}). More precisely, let vv be a valuation minimizing the normalized volume function

vol^X,Δ:ValX,x>0{+}\widehat{\rm vol}_{X,\Delta}\colon\mathrm{Val}_{X,x}\to\mathbb{R}_{>0}\cup\{+\infty\}

defined as in [Li-normalized-volume], then the corresponding degeneration is obtained as X0=Spec(GrvR)X_{0}={\rm Spec}({\rm Gr}_{v}R) with a cone vertex x0X0x_{0}\in X_{0}, Δ0\Delta_{0} is the corresponding degeneration of Δ\Delta and ξv\xi_{v} the Reeb vector induced by the valuation vv. This degeneration is volume-preserving, i.e. it satisfies

(1.1) vol^(x,X,Δ)=vol^X,Δ(v)=vol^X0,Δ0(wtξv)=vol^(x0,X0,Δ0).\widehat{\rm vol}(x,X,\Delta)=\widehat{\rm vol}_{X,\Delta}(v)=\widehat{\rm vol}_{X_{0},\Delta_{0}}(\mathrm{wt}_{\xi_{v}})=\widehat{\rm vol}(x_{0},X_{0},\Delta_{0})\,.

To complete the picture of the local stability theory, one central open problem (see e.g. [XZ-SDC, Conjecture 1.7]) is the boundedness of K-semistable nn-dimensional log Fano cone singularities (X,Δ;ξ)(X,\Delta;\xi), assuming we fix a lower bound of the local volume

vol^(x,X,Δ)=vol^X,Δ(wtξ).\widehat{\rm vol}(x,X,\Delta)=\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi})\,.

The aim of this paper is to settle this local boundedness question. Our main theorem is the following. The relevant definitions are recalled in Section 2.

Theorem 1.1.

Let nn\in\mathbb{N} and let I[0,1]I\subseteq[0,1] be a finite set. Let ε>0\varepsilon>0 be a positive real number. Let 𝒦\mathcal{K} be the set of K-semistable nn-dimensional polarized log Fano cone singularities x(X,Δ;ξ)x\in(X,\Delta;\xi) with coefficients in II such that

vol^(x,X,Δ)ε.\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon.

Then 𝒦\mathcal{K} is bounded.

By the volume-preserving property of the degeneration (1.1), Theorem 1.1 implies that all nn-dimensional klt singularities with a positive lower bound on their volumes, are bounded up to degeneration. In particular, we have the following consequence which gives a positive answer to [LLX-nv-survey, Question 6.12].

Theorem 1.2.

Fix nn\in\mathbb{N} and a finite set I[0,1]I\subseteq[0,1]. Then the set

Vol^n,I={vol^(x,X,Δ)x(X,Δ) is klt, dim(X)=n,Coeff(Δ)I}\widehat{\rm Vol}_{n,I}=\left\{\widehat{\rm vol}(x,X,\Delta)\mid x\in(X,\Delta)\mbox{ is klt, }\dim(X)=n,\,{\rm Coeff}(\Delta)\subseteq I\right\}

has 0 as the only accumulation point.

As another application of Theorem 1.1, we show that many classical invariants of klt singularities are controlled by their local volume.

Theorem 1.3.

Fix nn\in\mathbb{N}, ε>0\varepsilon>0 and a finite set I[0,1]I\subseteq[0,1]. Then there exists some constant M>0M>0 depending only on n,ε,In,\varepsilon,I such that for any nn-dimensional klt singularity x(X,Δ)x\in(X,\Delta) with Coeff(Δ)I{\rm Coeff}(\Delta)\subseteq I and vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon, we have

dimκ(x)𝔪x/𝔪x2M,multxXM,mldK(x,X,Δ)M.\dim_{\kappa(x)}\mathfrak{m}_{x}/\mathfrak{m}_{x}^{2}\leq M,\quad\mathrm{mult}_{x}X\leq M,\quad\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq M\,.

Here mldK(x,X,Δ)\mathrm{mld}^{K}(x,X,\Delta) denotes the minimal log discrepancy of the Kollár components of the singularity, see Definition 4.2. For the embedded dimension dimκ(x)𝔪x/𝔪x2\dim_{\kappa(x)}\mathfrak{m}_{x}/\mathfrak{m}_{x}^{2} and the multiplicity multxX\mathrm{mult}_{x}X, the existence of the uniform upper bound MM directly follows from (1.1), Theorem 1.1 and the upper semi-continuity of these invariants. The mldK\mathrm{mld}^{K} case is more subtle, see Theorem 4.3. Note that the converse of Theorem 1.3 is false, i.e. even if the embedded dimension, multiplicity and mldK\mathrm{mld}^{K} of the singularities are all bounded from above, the local volume of the singularity can still be arbitrarily small. This can be seen already in the case of ADE surface singularities.

Philosophically, one can compare Theorem 1.2 with [HMX-ACC, Theorem 1.3] which deals with the global case of log general type pairs, where volumes might accumulate even when the coefficients belong to a finite set, though they still satisfy the descend chain condition (DCC). Special cases of Theorem 1.2 have been previously established, including when XX is bounded [HLQ-vol-ACC], (X,Δ)(X,\Delta) is of complexity at most one [MS-bdd-toric, LMS-bdd-dim-3], or when XX is of dimension at most three [LMS-bdd-dim-3, Z-mld^K-2].

To prove Theorem 1.1, we need to sort out a birational geometric condition that is more flexible than K-semistability so that it is preserved under small perturbations of the polarization. In particular, we want to include the case when ξ\xi is a rational perturbation of the Reeb vector coming from the K-semistable log Fano cone structure. This is why instead of proving Theorem 1.1 directly, we aim to prove the following more general statement.

Theorem 1.4.

Let nn\in\mathbb{N} and let I[0,1]I\subseteq[0,1] be a finite set. Let ε,θ>0\varepsilon,\theta>0. Let 𝒮\mathcal{S} be the set of nn-dimensional polarized log Fano cone singularities x(X,Δ;ξ)x\in(X,\Delta;\xi) with coefficients in II such that

vol^(x,X,Δ)ε,andΘ(X,Δ;ξ)θ.\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon,\quad\mathrm{and}\quad\Theta(X,\Delta;\xi)\geq\theta.

Then 𝒮\mathcal{S} is bounded.

Here Θ(X,Δ;ξ)\Theta(X,\Delta;\xi) is the volume ratio of the log Fano cone singularity, see Definition 2.7. It serves as a local analog of the stability invariant of Fano varieties. A polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) is K-semistable if and only if Θ(X,Δ;ξ)=1\Theta(X,\Delta;\xi)=1, thus Theorem 1.4 implies Theorem 1.1. An observation from [Z-mld^K-2] is that the condition of volume ratio having a lower bound should be the right generalization of the K-semistable condition to guarantee boundedness.

Under the assumption that Θ(X,Δ;ξ)θ\Theta(X,\Delta;\xi)\geq\theta, the condition that vol^(x,X,Δ)\widehat{\rm vol}(x,X,\Delta) has a uniform positive lower bound is equivalent to vol^X,Δ(wtξ)\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi}) being uniformly bounded from below by a positive number. When ξ\xi is rational, the latter condition is equivalent to a global condition on the orbifold base. This leads to a special case for this version of Theorem 1.4.

Corollary 1.5.

Fix a positive integer nn, and two positive numbers α0,ε\alpha_{0},\varepsilon. Consider all (n1)(n-1)-dimensional log Fano pairs (V,ΔV)(V,\Delta_{V}) such that there exist some r>0r>0 and some Weil divisor LL on VV which satisfy

α(V,ΔV)α0,(KV+ΔV)rL and r(KVΔV)n1ε.\alpha(V,\Delta_{V})\geq\alpha_{0},\,-(K_{V}+\Delta_{V})\sim_{\mathbb{Q}}rL\mbox{ \ \ and \ \ }r(-K_{V}-\Delta_{V})^{n-1}\geq\varepsilon\,.

Let 𝒮\mathcal{S} be the set of log Fano cone singularities given by

X=SpecmH0(V,mL)X={\rm Spec}\bigoplus_{m}H^{0}(V,mL)

for all possible (V,Δ,L)(V,\Delta,L) as above and Δ\Delta the closure of the pull back of ΔV\Delta_{V} on XX. Then 𝒮\mathcal{S} is bounded.

We note that, somewhat surprisingly, in Corollary 1.5 all such (V,Δ,L)(V,\Delta,L) themselves are not bounded.

1.6Strategy of the proof.

The corresponding global result of Theorem 1.4 has been proved in [Jia-Kss-Fano-bdd], where it is shown that Fano varieties whose volume and alpha invariant are bounded away from zero form a bounded set. To prove the boundedness of the log Fano cone singularities x(X,Δ;ξ)x\in(X,\Delta;\xi), one wants to reduce it to some boundedness question for projective Fano type varieties. A natural candidate would be the quotient

(V,ΔV):=((X,Δ){x})/ξ(V,\Delta_{V}):=((X,\Delta)\setminus\{x\})/\langle\xi\rangle

(we may assume ξ\xi is rational after a perturbation and hence it generates a 𝔾m\mathbb{G}_{m}-action). While one can show that the alpha invariant of the log Fano pair (V,ΔV)(V,\Delta_{V}) is bounded from below (see [Z-mld^K-1, Z-mld^K-2]), however, the volume (KVΔV)n1(-K_{V}-\Delta_{V})^{n-1} could be arbitrarily small (see Example 2.8). In particular, (V,ΔV)(V,\Delta_{V}) is not necessarily bounded, which posts a major challenge in the proof.

A better candidate, first proposed in [Z-mld^K-2], is the projective orbifold cone compactification (X¯,Δ¯)(\overline{X},\overline{\Delta}) of (X,Δ)(X,\Delta), which adds a divisor isomorphic to VV at infinity. One piece of evidence from [Z-mld^K-2] is that if the local volume of the singularity x(X,Δ)x\in(X,\Delta) is bounded from below, then so is the volume of log Fano pair (X¯,Δ¯)(\overline{X},\overline{\Delta}). On the other hand, the alpha invariant of (X¯,Δ¯)(\overline{X},\overline{\Delta}) can be arbitrarily small and the log Fano pairs (X¯,Δ¯)(\overline{X},\overline{\Delta}) still do not form a bounded family as we vary the Reeb vector ξ\xi. To get around this problem, the arguments in [Z-mld^K-2] rely on an additional subtle assumption on the Kollár components of the singularities, and it is not clear if this extra assumption is always satisfied. In this paper, we take a different path and follow a strategy which in spirit is closer to [Birkar-bab-1, Jia-Kss-Fano-bdd].

The first step is to prove a birational version of [Jia-Kss-Fano-bdd]. More specifically, we first show (see Section 2.4) that if the volume ratio Θ(X,Δ;ξ)\Theta(X,\Delta;\xi) is bounded from below, then away from the divisor at infinity, the alpha invariant of (X¯,Δ¯)(\overline{X},\overline{\Delta}) is also bounded from below, i.e. there is some uniform α0>0\alpha_{0}>0 such that αx1(X¯,Δ¯)α0\alpha_{x_{1}}(\overline{X},\overline{\Delta})\geq\alpha_{0} for all x1Xx_{1}\in X. By adapting the arguments of [Jia-Kss-Fano-bdd], we then show (see Section 3.1) that this together with boundedness of the volume imply that the set of projective orbifold cones (X¯,Δ¯)(\overline{X},\overline{\Delta}) is log birationally bounded (Definition 2.19). In particular, they are birational to a bounded set of pairs.

The next step is to improve the log birational boundedness to boundedness in codimension one, see Section 3.2. Inspired by [Birkar-bab-1], and using the fact that αx1(X¯,Δ¯)α0\alpha_{x_{1}}(\overline{X},\overline{\Delta})\geq\alpha_{0} away from the divisor at infinity, we show that there is a uniform way to modify the bounded birational model YY obtained from the previous step, so that the only exceptional divisor of the induced birational map X¯Y\overline{X}\dashrightarrow Y is the divisor at infinity. In other words, YX¯Y\dashrightarrow\overline{X} is close to a birational contraction except possibly over one divisor. This is the best we can hope for, as the divisor at infinity depends on the Reeb vector, and in general cannot be extracted on a bounded model. The main ingredient for this step is the construction of sub-klt bounded complements of (X¯,Δ¯)(\overline{X},\overline{\Delta}) with certain control on the negative part. This in turn relies on the boundedness of complements proved in [Birkar-bab-1, Theorem 1.7], as well as the birationally bounded model constructed in the previous step.

Finally, to finish the argument, we recover (X,Δ)(X,\Delta) by running a carefully chosen minimal model program on YY and using the affineness of XX to show that XX is embedded as an open set of the ample model obtained from the minimal model program sequence. See Section 3.3.

Acknowledgement. CX is partially supported by NSF DMS-2139613, DMS-2201349 and a Simons Investigator grant. ZZ is partially supported by the NSF Grants DMS-2240926, DMS-2234736, a Clay research fellowship, as well as a Sloan fellowship.

2. Preliminaries

2.1. Notation and conventions

We work over an algebraically closed field 𝕜\mathbbm{k} of characteristic 0. We follow the standard terminology from [KM98, Kol13].

A sub-pair (X,Δ)(X,\Delta) consists of a normal variety XX together with an \mathbb{R}-divisor Δ\Delta on XX (a priori, we do not require that KX+ΔK_{X}+\Delta is \mathbb{R}-Cartier). It is called a pair if Δ\Delta is effective. A log smooth pair (Y,Σ)(Y,\Sigma) consists of a smooth variety YY and a simple normal crossing divisor Σ\Sigma on YY. We say that a pair (X,Δ)(X,\Delta) is log Fano if (X,Δ)(X,\Delta) is klt ([KM98, Definition 2.34]) and (KX+Δ)-(K_{X}+\Delta) is \mathbb{R}-Cartier and ample.

A singularity x(X,Δ)x\in(X,\Delta) consists of a pair (X,Δ)(X,\Delta) and a closed point xXx\in X. We will always assume that XX is affine and xSupp(Δ)x\in\mathrm{Supp}(\Delta) (whenever Δ0\Delta\neq 0). We say that the singularity is klt if (X,Δ)(X,\Delta) is klt in a neighbourhood of xx.

Given an \mathbb{R}-divisor Δ\Delta on XX and a birational map φ:YX\varphi\colon Y\dashrightarrow X, we denote the strict transform of Δ\Delta on the birational model YY by ΔY\Delta_{Y}, i.e. ΔY=φ1Δ\Delta_{Y}=\varphi_{*}^{-1}\Delta. If Δ\Delta is \mathbb{R}-Cartier, the birational pullback φΔ\varphi^{*}\Delta is defined as the \mathbb{R}-divisor fgΔf_{*}g^{*}\Delta where f:WYf\colon W\to Y, g:WXg\colon W\to X is a common resolution.

When we refer to a constant CC as C=C(n,ε,)C=C(n,\varepsilon,\cdots) it means CC only depends on n,ε,n,\varepsilon,\cdots, etc.

2.2. Local volumes

We first briefly recall the definition of the local volumes of klt singularities [Li-normalized-volume]. For this we need the notion of valuations. A valuation over a singularity xXx\in X is an \mathbb{R}-valued valuation v:K(X)v:K(X)^{*}\to\mathbb{R} (where K(X)K(X) denotes the function field of XX) such that vv is centered at xx (i.e. if f𝒪X,xf\in\mathcal{O}_{X,x}, then v(f)>0v(f)>0 if and only if f𝔪xf\in\mathfrak{m}_{x}) and v|𝕜=0v|_{\mathbbm{k}^{*}}=0. The set of such valuations is denoted as ValX,x\mathrm{Val}_{X,x}. Let x(X,Δ)x\in(X,\Delta) be a singularity and assume that KX+ΔK_{X}+\Delta is \mathbb{R}-Cartier. The log discrepancy function

AX,Δ:ValX,x{+},A_{X,\Delta}\colon\mathrm{Val}_{X,x}\to\mathbb{R}\cup\{+\infty\},

is defined as in [JM-val-ideal-seq] and [BdFFU-log-discrepancy, Theorem 3.1]. It generalizes the usual log discrepancies of divisors; in particular, for divisorial valuations, i.e. valuations of the form λordF\lambda\cdot\mathrm{ord}_{F} where λ>0\lambda>0 and FF is a prime divisor on some proper birational model π:YX\pi\colon Y\to X, we have

AX,Δ(λordF)=λAX,Δ(F)=λ(1+ordF(KYπ(KX+Δ))).A_{X,\Delta}(\lambda\cdot\mathrm{ord}_{F})=\lambda\cdot A_{X,\Delta}(F)=\lambda\cdot(1+\mathrm{ord}_{F}(K_{Y}-\pi^{*}(K_{X}+\Delta))).

For klt singularities, one has AX,Δ(v)>0A_{X,\Delta}(v)>0 for all vValX,xv\in\mathrm{Val}_{X,x}. We denote by ValX,x\mathrm{Val}^{*}_{X,x} the set of valuations vValXv\in\mathrm{Val}_{X} with center xx and AX,Δ(v)<+A_{X,\Delta}(v)<+\infty. The volume of a valuation vValX,xv\in\mathrm{Val}_{X,x} is defined as

vol(v)=volX,x(v)=lim supm(𝒪X,x/𝔞m(v))mn/n!,\mathrm{vol}(v)=\mathrm{vol}_{X,x}(v)=\limsup_{m\to\infty}\frac{\ell(\mathcal{O}_{X,x}/\mathfrak{a}_{m}(v))}{m^{n}/n!},

where n=dimXn=\dim X and 𝔞m(v)\mathfrak{a}_{m}(v) denotes the valuation ideal, i.e.

𝔞m(v):={f𝒪X,xv(f)m}.\mathfrak{a}_{m}(v):=\{f\in\mathcal{O}_{X,x}\mid v(f)\geq m\}.
Definition 2.1.

Let x(X,Δ)x\in(X,\Delta) be an nn-dimensional klt singularity. For any vValX,xv\in\mathrm{Val}_{X,x}, we define the normalized volume of vv as

vol^X,Δ(v):={AX,Δ(v)nvolX,x(v) if AX,Δ(v)<++ if AX,Δ(v)=+.\widehat{\rm vol}_{X,\Delta}(v):=\begin{cases}A_{X,\Delta}(v)^{n}\cdot\mathrm{vol}_{X,x}(v)\text{ \ \ \ \ if $A_{X,\Delta}(v)<+\infty$}\\ +\infty\text{ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ if $A_{X,\Delta}(v)=+\infty$}\end{cases}\,.

The local volume of x(X,Δ)x\in(X,\Delta) is defined as

vol^(x,X,Δ):=infvValX,xvol^X,Δ(v).\widehat{\rm vol}(x,X,\Delta):=\inf_{v\in\mathrm{Val}^{*}_{X,x}}\widehat{\rm vol}_{X,\Delta}(v)\,.

By [Li-normalized-volume, Theorem 1.2], the local volume of a klt singularity is always positive.

2.3. Log Fano cone singularities

Next we recall the definition of log Fano cone singularities and their K-semistability.

Definition 2.2.

Let X=Spec(R)X=\mathrm{Spec}(R) be a normal affine variety and 𝕋\mathbb{T} an algebraic torus (i.e. 𝕋𝔾mr\mathbb{T}\cong\mathbb{G}_{m}^{r} for some r>0r>0). We say that a 𝕋\mathbb{T}-action on XX is good if it is effective and there is a unique closed point xXx\in X that is in the orbit closure of any 𝕋\mathbb{T}-orbit. We call xx the vertex of the 𝕋\mathbb{T}-variety XX.

Let N:=N(𝕋)=Hom(𝔾m,𝕋)N:=N(\mathbb{T})=\mathrm{Hom}(\mathbb{G}_{m},\mathbb{T}) be the co-weight lattice and M=NM=N^{*} the weight lattice. We have a weight decomposition

R=αMRα,R=\oplus_{\alpha\in M}R_{\alpha},

and the action being good implies that R0=𝕜R_{0}=\mathbbm{k} and every RαR_{\alpha} is finite dimensional. For fRf\in R, we denote by fαf_{\alpha} the corresponding component in the above weight decomposition.

Definition 2.3.

A Reeb vector on XX is a vector ξN\xi\in N_{\mathbb{R}} such that ξ,α>0\langle\xi,\alpha\rangle>0 for all 0αM0\neq\alpha\in M with Rα0R_{\alpha}\neq 0. The set 𝔱+\mathfrak{t}^{+}_{\mathbb{R}} of Reeb vectors is called the Reeb cone.

For any ξ𝔱+\xi\in\mathfrak{t}^{+}_{\mathbb{R}}, we can define a valuation wtξ\mathrm{wt}_{\xi} by setting

wtξ(f):=min{ξ,ααM,fα0}\mathrm{wt}_{\xi}(f):=\min\{\langle\xi,\alpha\rangle\mid\alpha\in M,f_{\alpha}\neq 0\}

where fRf\in R. It is not hard to verify that wtξValX,x\mathrm{wt}_{\xi}\in\mathrm{Val}_{X,x}.

Definition 2.4.

A log Fano cone singularity is a klt singularity that admits a nontrivial good torus action. A polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) consists of a log Fano cone singularity x(X,Δ)x\in(X,\Delta) together with a Reeb vector ξ\xi (called a polarization).

By abuse of convention, a good 𝕋\mathbb{T}-action on a klt singularity x(X,Δ)x\in(X,\Delta) means a good 𝕋\mathbb{T}-action on XX such that xx is the vertex and Δ\Delta is 𝕋\mathbb{T}-invariant. Using terminology from Sasakian geometry, we say a polarized log Fano cone x(X,Δ;ξ)x\in(X,\Delta;\xi) is quasi-regular if ξ\xi generates a 𝔾m\mathbb{G}_{m}-action (i.e. ξ\xi is a real multiple of some element of NN); otherwise, we say that x(X,Δ;ξ)x\in(X,\Delta;\xi) is irregular. The torus generated by ξ\xi will be denoted by ξ\langle\xi\rangle.

Definition 2.5 ([CS-Kss-Sasaki, CS-Sasaki-Einstein], [LX-stability-higher-rank, Theorem 2.34]).

We say that a polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) is K-semistable if

vol^(x,X,Δ)=vol^X,Δ(wtξ).\widehat{\rm vol}(x,X,\Delta)=\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi}).

Log Fano cone singularities play a special role in the local stability theory of klt singularities, due to the following statement. It was originally known as the Stable Degeneration Conjecture.

Theorem 2.6.

Every klt singularity x(X=Spec(R),Δ)x\in(X={\rm Spec}(R),\Delta) has a special degeneration to a K-semistable log Fano cone singularity x0(X0,Δ0;ξv)x_{0}\in(X_{0},\Delta_{0};\xi_{v}) with

vol^(x,X,Δ)=vol^(x0,X0,Δ0).\widehat{\rm vol}(x,X,\Delta)=\widehat{\rm vol}(x_{0},X_{0},\Delta_{0}).

More precisely, up to rescaling, there is a unique valuation vv minimizing vol^X,Δ\widehat{\rm vol}_{X,\Delta}, and X0=Spec(Grv(R))X_{0}={\rm Spec}({\rm Gr}_{v}(R)). In addition, denote by Δ0\Delta_{0} the degeneration of Δ\Delta, and ξv\xi_{v} the Reeb vector induced by vv, then (X0,Δ0;ξv)(X_{0},\Delta_{0};\xi_{v}) is a K-semistable log Fano cone.

We call x0(X0,Δ0;ξv)x_{0}\in(X_{0},\Delta_{0};\xi_{v}) the K-semistable log Fano cone degeneration of x(X,Δ)x\in(X,\Delta). It is unique up to the rescaling of the Reeb vector ξv\xi_{v}.

Proof.

See [Blu-minimizer-exist, LX-stability-higher-rank, Xu-quasi-monomial, XZ-minimizer-unique, XZ-SDC] for the case of rational coefficients and the extension to real coefficients in [Z-mld^K-2]. ∎

Definition 2.7.

The volume ratio of a polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) is defined to be

Θ(X,Δ;ξ):=vol^(x,X,Δ)vol^X,Δ(wtξ).\Theta(X,\Delta;\xi):=\frac{\widehat{\rm vol}(x,X,\Delta)}{\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi})}.

By definition, 0<Θ(X,Δ;ξ)10<\Theta(X,\Delta;\xi)\leq 1 and Θ(X,Δ;ξ)=1\Theta(X,\Delta;\xi)=1 if and only if x(X,Δ;ξ)x\in(X,\Delta;\xi) is K-semistable.

2.4. Orbifold cones

Every quasi-regular polarized log Fano cone singularity has a natural projective orbifold cone compactification. This provides a convenient way to think about these singularities. In this subsection, we recall this construction and relate some invariants of the singularities with those of the projective orbifold cones. For more background, see [Kol-Seifert-bundle] or [Z-mld^K-2, Section 3.1].

Let x(X,Δ;ξ)x\in(X,\Delta;\xi) be a quasi-regular log Fano cone singularity. By definition, the ξ𝔾m\langle\xi\rangle\cong\mathbb{G}_{m}-action on X{x}X\setminus\{x\} has finite stabilizers, hence the quotient map

X{x}V:=(X{x})/ξX\setminus\{x\}\to V:=(X\setminus\{x\})/\langle\xi\rangle

is a Seifert 𝔾m\mathbb{G}_{m}-bundle (in the sense of [Kol-Seifert-bundle]), and there is an ample \mathbb{Q}-divisor LL on VV such that (see [Kol-Seifert-bundle, Theorem 7])

X=SpecmH0(V,mL).X={\rm Spec}\bigoplus_{m\in\mathbb{N}}H^{0}(V,\lfloor mL\rfloor)\,.

The zero section V0V_{0} of this Seifert 𝔾m\mathbb{G}_{m}-bundle gets contracted to the closed point xXx\in X (as a valuation, ordV0\mathrm{ord}_{V_{0}} is proportional to wtξ\mathrm{wt}_{\xi}), thus we can compactify XX to X¯\overline{X} by adding the infinity section VV_{\infty}. Let Δ¯\overline{\Delta} be the closure of Δ\Delta on X¯\overline{X}. We call (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) the (projective) orbifold cone compactification of x(X,Δ;ξ)x\in(X,\Delta;\xi) (cf. [Z-mld^K-2, Section 3.1]).

Note that (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) is plt and (KX¯+Δ¯+V)-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}) is ample, see [Z-mld^K-2, Lemma 3.3] or [Kol-Seifert-bundle]. By adjunction along VVV_{\infty}\cong V, we may write

(KX¯+Δ¯+V)|V=KV+ΔV(K_{\overline{X}}+\overline{\Delta}+V_{\infty})|_{V_{\infty}}=K_{V}+\Delta_{V}

for some effective divisor ΔV\Delta_{V}. Then (V,ΔV)(V,\Delta_{V}) is a klt log Fano pair. We call (V,ΔV)(V,\Delta_{V}) the orbifold base of the singularity x(X,Δ;ξ)x\in(X,\Delta;\xi). There exists some r>0r>0 such that (KV+ΔV)rL-(K_{V}+\Delta_{V})\sim_{\mathbb{R}}rL, and we have

(2.1) (KX¯+Δ¯+V)rV.-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})\sim_{\mathbb{R}}rV_{\infty}\,.

A subtle feature of the local boundedness problem is that the orbifold bases do not belong to a bounded set; already their volumes can be arbitrarily small when we fix the singularity x(X,Δ)x\in(X,\Delta) and vary the Reeb vector ξ\xi.

Example 2.8.

Let (xX)=(0𝔸n)(x\in X)=(0\in\mathbb{A}^{n}) and ξ=(ξ1,,ξn)\xi=(\xi_{1},\dots,\xi_{n}) for some pairwise coprime positive integers ξi\xi_{i} (1in)(1\leq i\leq n). Assume that n3n\geq 3 and ξ1ξn\xi_{1}\leq\cdots\leq\xi_{n}. Then

X¯=(1,ξ1,,ξn),V(ξ1,,ξn),ΔV=0,andL=𝒪(1).\overline{X}=\mathbb{P}(1,\xi_{1},\dots,\xi_{n}),\quad V\cong\mathbb{P}(\xi_{1},\dots,\xi_{n}),\quad\Delta_{V}=0,\quad\mathrm{and}\quad L=\mathcal{O}(1).

We can easily compute

Θ(𝔸n;ξ)=nnvol^𝔸n(wt(ξ))=ξ1ξnnn(ξ1++ξn)n.\Theta(\mathbb{A}^{n};\xi)=\frac{n^{n}}{\widehat{\rm vol}_{\mathbb{A}^{n}}({\rm wt}(\xi))}=\frac{\xi_{1}\cdots\xi_{n}\cdot n^{n}}{(\xi_{1}+\cdots+\xi_{n})^{n}}\,.

So Θ(𝔸n,ξ)\Theta(\mathbb{A}^{n},\xi) has a positive lower bound if and only if ξnξ1\frac{\xi_{n}}{\xi_{1}} has an upper bound. Using [BJ-delta, Corollary 7.16], one can show that this is also equivalent to the condition that the α\alpha-invariant α(V)\alpha(V) defined below in Definition 2.10 has a positive lower bound.

On the other hand,

vol(KV)=(ξ1++ξn)n1ξ1ξn.\mathrm{vol}(-K_{V})=\frac{(\xi_{1}+\cdots+\xi_{n})^{n-1}}{\xi_{1}\cdots\xi_{n}}\,.

So if ξnξ1\frac{\xi_{n}}{\xi_{1}} is bounded from above, then vol(KV)\mathrm{vol}(-K_{V}) is bounded away from zero if and only if all the weights ξi\xi_{i} are bounded from above.

A key observation from [Z-mld^K-2], following a direct calculation using (2.1), is that the volume of a log Fano cone singularity is more closely related to the global volume of its projective orbifold cone compactification (rather than the orbifold base).

Lemma 2.9.

Notation as above. Then we have

vol^X,Δ(wtξ)=vol((KX¯+Δ¯+V)).\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi})=\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})).

In particular, if Θ(X,Δ;ξ)θ>0\Theta(X,\Delta;\xi)\geq\theta>0, then vol((KX¯+Δ¯+V))nnθ1\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}))\leq n^{n}\theta^{-1}.

Proof.

The equality is [Z-mld^K-2, Lemma 3.4]. The other implication then follows from [LX-cubic-3fold, Theorem 1.6]. ∎

We next relate the volume ratio with the α\alpha-invariants of the orbifold base or the projective orbifold cone. First we recall some definitions.

Definition 2.10.

Let (X,Δ)(X,\Delta) be a projective klt pair and let LL be a big \mathbb{R}-Cartier \mathbb{R}-divisor. We define the α\alpha-invariant α(X,Δ;L)\alpha(X,\Delta;L) as

α(X,Δ;L):=inf{lct(X,Δ;D)0DL},\alpha(X,\Delta;L):=\inf\left\{\mathrm{lct}(X,\Delta;D)\mid 0\leq D\sim_{\mathbb{R}}L\right\},

where lct(X,Δ;D)\mathrm{lct}(X,\Delta;D) denotes the log canonical threshold, i.e. the largest number λ\lambda such that (X,Δ+λD)(X,\Delta+\lambda D) is log canonical. For any projective pair (X,Δ)(X,\Delta) that is klt at a closed point xXx\in X, we can similarly define the log canonical threshold lctx(X,Δ;D)\mathrm{lct}_{x}(X,\Delta;D) at xx and the local α\alpha-invariant

αx(X,Δ;L):=inf{lctx(X,Δ;D)0DL}.\alpha_{x}(X,\Delta;L):=\inf\left\{\mathrm{lct}_{x}(X,\Delta;D)\mid 0\leq D\sim_{\mathbb{R}}L\right\}.

When the pair (X,Δ)(X,\Delta) is clear from the context, we will just write α(L)\alpha(L) and αx(L)\alpha_{x}(L). For a log Fano pair (X,Δ)(X,\Delta), we define α(X,Δ):=α(X,Δ;KXΔ)\alpha(X,\Delta):=\alpha(X,\Delta;-K_{X}-\Delta) and similarly αx(X,Δ)\alpha_{x}(X,\Delta).

While at a point on the infinity divisor VV_{\infty}, the α\alpha-invariant of the projective orbifold cone (X¯,Δ¯)(\overline{X},\overline{\Delta}) could be very small when rr is large in (2.1), the following result roughly says that for any point outside the infinity divisor, the local α\alpha-invariant of (X¯,Δ¯)(\overline{X},\overline{\Delta}) is bounded by the (global) α\alpha-invariant of the orbifold base.

Lemma 2.11.

Let x(X,Δ;ξ)x\in(X,\Delta;\xi) be a quasi-regular polarized log Fano cone singularity. Let (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) be its projective orbifold cone compactification, and let (V,ΔV)(V,\Delta_{V}) be the orbifold base. Then we have

αx1(X¯,Δ¯+V)min{1,α(V,ΔV)}\alpha_{x_{1}}(\overline{X},\overline{\Delta}+V_{\infty})\geq\min\{1,\alpha(V,\Delta_{V})\}

for all closed point x1Xx_{1}\in X. In particular, αx(X¯,Δ¯+V)=min{1,α(V,ΔV)}\alpha_{x}(\overline{X},\overline{\Delta}+V_{\infty})=\min\{1,\alpha(V,\Delta_{V})\}.

Proof.

Let DV(KV+ΔV)D_{V}\sim_{\mathbb{R}}-(K_{V}+\Delta_{V}) be an effective \mathbb{R}-divisor and let DD be the closure in X¯\overline{X} of its pullback to X{x}X\setminus\{x\}. Then we have D(KX¯+Δ¯+V)D\sim_{\mathbb{R}}-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}) and ordV0(D)=AX,Δ(V0)\mathrm{ord}_{V_{0}}(D)=A_{X,\Delta}(V_{0}) (for usual cones see [Kol13, Proposition 3.14], the general case follows from the computations in [Kol-Seifert-bundle, Section 4]). In particular, αx(X¯,Δ¯+V)1\alpha_{x}(\overline{X},\overline{\Delta}+V_{\infty})\leq 1. Since X{x}VX\setminus\{x\}\to V is a Seifert 𝔾m\mathbb{G}_{m}-bundle, the pair (V,ΔV+tDV)(V,\Delta_{V}+tD_{V}) is log canonical if and only if (X¯,Δ¯+tD)(\overline{X},\overline{\Delta}+tD) is log canonical on X{x}X\setminus\{x\}. Thus we also get αx(X¯,Δ¯+V)α(V,ΔV)\alpha_{x}(\overline{X},\overline{\Delta}+V_{\infty})\leq\alpha(V,\Delta_{V}).

Suppose that there exists some effective \mathbb{R}-divisor D(KX¯+Δ¯+V)D\sim_{\mathbb{R}}-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}) such that t:=lctx1(X¯,Δ¯+V;D)<min{1,α(V,ΔV)}t:=\mathrm{lct}_{x_{1}}(\overline{X},\overline{\Delta}+V_{\infty};D)<\min\{1,\alpha(V,\Delta_{V})\} for some x1Xx_{1}\in X. Since x1Vx_{1}\not\in V_{\infty} and VV_{\infty} is ample, we may assume that VSupp(D)V_{\infty}\not\in\mathrm{Supp}(D). The non-klt locus of the pair (X¯,Δ¯+V+tD)(\overline{X},\overline{\Delta}+V_{\infty}+tD) thus contains at least x1x_{1} and VV_{\infty}. Since (KX¯+Δ¯+V+tD)-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}+tD) is ample, Kollár-Shokurov’s connectedness lemma implies that (X¯,Δ¯+V+tD)(\overline{X},\overline{\Delta}+V_{\infty}+tD) is not plt along VV_{\infty}. It then follows from adjunction that (V,ΔV+tD|V)(V,\Delta_{V}+tD|_{V_{\infty}}) (we identify VV with VV_{\infty}) is not klt, and hence α(V,ΔV)t\alpha(V,\Delta_{V})\leq t, a contradiction. In other words, we have

αx1(X¯,Δ¯+V)min{1,α(V,ΔV)}.\alpha_{x_{1}}(\overline{X},\overline{\Delta}+V_{\infty})\geq\min\{1,\alpha(V,\Delta_{V})\}.

Combined with the upper bounds of αx(X¯,Δ¯+V)\alpha_{x}(\overline{X},\overline{\Delta}+V_{\infty}) we obtain above, this also gives αx(X¯,Δ¯+V)=min{1,α(V,ΔV)}\alpha_{x}(\overline{X},\overline{\Delta}+V_{\infty})=\min\{1,\alpha(V,\Delta_{V})\}. ∎

We now relate the volume ratio with the α\alpha-invariant of the orbifold base.

Lemma 2.12.

There exists some constant c>0c>0 depending only on the dimension such that for any quasi-regular polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) of dimension nn with orbifold base (V,ΔV)(V,\Delta_{V}), we have

cα(V,ΔV)Θ(X,Δ;ξ)min{1,α(V,ΔV)}n.c\cdot\alpha(V,\Delta_{V})\geq\Theta(X,\Delta;\xi)\geq\min\{1,\alpha(V,\Delta_{V})\}^{n}.
Proof.

This essentially follows from [Z-mld^K-2, Lemma 3.12 and Remark 3.13]. We provide a (slightly different) proof for the reader’s convenience. Let (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) be the associated projective orbifold cone as before. Let DV(KV+ΔV)D_{V}\sim_{\mathbb{R}}-(K_{V}+\Delta_{V}) be an effective \mathbb{R}-divisor and let DD be the closure in XX of its pullback to X{x}X\setminus\{x\}. Then wtξ(D)=AX,Δ(wtξ)\mathrm{wt}_{\xi}(D)=A_{X,\Delta}(\mathrm{wt}_{\xi}). The uniform Izumi inequality in [Z-mld^K-1, Lemma 3.4] thus implies that

lctx(X,Δ;D)c0Θ(X,Δ;ξ)\mathrm{lct}_{x}(X,\Delta;D)\geq c_{0}\cdot\Theta(X,\Delta;\xi)

for some constant c0=c0(n)>0c_{0}=c_{0}(n)>0. But we also have lct(V,ΔV;DV)lctx(X,Δ;D)\mathrm{lct}(V,\Delta_{V};D_{V})\geq\mathrm{lct}_{x}(X,\Delta;D) as in the proof of Lemma 2.11. As DVD_{V} is arbitrary, this gives the first inequality with c=c01c=c_{0}^{-1}.

By Lemma 2.11 and the following Lemma 2.13, we have

vol^(x,X,Δ)\displaystyle\widehat{\rm vol}(x,X,\Delta) \displaystyle\geq αx(X¯,Δ¯+V)nvol((KX¯+Δ¯+V))\displaystyle\alpha_{x}(\overline{X},\overline{\Delta}+V_{\infty})^{n}\cdot\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}))
=\displaystyle= min{1,α(V,ΔV)}nvol((KX¯+Δ¯+V)).\displaystyle\min\{1,\alpha(V,\Delta_{V})\}^{n}\cdot\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}))\,.

On the other hand, we have vol^X,Δ(wtξ)=vol((KX¯+Δ¯+V))\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi})=\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})) by Lemma 2.9. This gives the second inequality. ∎

We have used the following statement, which is well-known to experts.

Lemma 2.13.

Let (X,Δ)(X,\Delta) be a pair of dimension nn that is klt at a closed point xx, and let LL be a big \mathbb{R}-Cartier \mathbb{R}-divisor on XX. Then we have

vol^(x,X,Δ)αx(X,Δ;L)nvol(L).\widehat{\rm vol}(x,X,\Delta)\geq\alpha_{x}(X,\Delta;L)^{n}\cdot\mathrm{vol}(L).
Proof.

Let t=α(X,Δ;L)t=\alpha(X,\Delta;L). Suppose that xXx\in X is a smooth point and vol(L)>nntn\mathrm{vol}(L)>\frac{n^{n}}{t^{n}}. Then it is well-known, by a simple dimension count, that there exists some effective \mathbb{Q}-divisor DLD\sim_{\mathbb{Q}}L such that multxD>nt\mathrm{mult}_{x}D>\frac{n}{t}; in particular,

αx(X,Δ;L)lctx(X¯,Δ¯;D)<t,\alpha_{x}(X,\Delta;L)\leq\mathrm{lct}_{x}(\overline{X},\overline{\Delta};D)<t\,,

a contradiction. We can apply the same dimension counting argument at a singular point xXx\in X, as long as we replace multx\mathrm{mult}_{x} by the minimizing valuation of the normalized volume function, and nnn^{n} by the local volume vol^(x,X,Δ)\widehat{\rm vol}(x,X,\Delta). ∎

2.5. Bounded family

In this subsection we define various notions of boundedness.

Definition 2.14.

We call (𝒳,𝒟)B(\mathcal{X},\mathcal{D})\to B a family of pairs if 𝒳\mathcal{X} is flat over BB, the fibers 𝒳b\mathcal{X}_{b} are connected, normal and not contained in Supp(𝒟)\mathrm{Supp}(\mathcal{D}).

We call B(𝒳,𝒟)BB\subseteq(\mathcal{X},\mathcal{D})\to B an \mathbb{R}-Gorenstein family of klt singularities (over a normal but possibly disconnected base BB) if

  1. (1)

    (𝒳,𝒟)B(\mathcal{X},\mathcal{D})\to B is a family of pairs, and B𝒳B\subseteq\mathcal{X} is a section,

  2. (2)

    K𝒳/B+𝒟K_{\mathcal{X}/B}+\mathcal{D} is \mathbb{R}-Cartier and b(𝒳b,𝒟b)b\in(\mathcal{X}_{b},\mathcal{D}_{b}) is a klt singularity for all bBb\in B.

Definition 2.15.

We say that a set 𝒞\mathcal{C} of sub-pairs is bounded if there exists a family (𝒳,𝒟)B(\mathcal{X},\mathcal{D})\to B of pairs over a finite type base BB such that for any (X,D)𝒞(X,D)\in\mathcal{C}, there exists a closed point bBb\in B and an isomorphism (X,Supp(D))(𝒳b,Supp(𝒟b))(X,\mathrm{Supp}(D))\cong(\mathcal{X}_{b},\mathrm{Supp}(\mathcal{D}_{b})).

Definition 2.16 ([Z-mld^K-2, Definition 2.16]).

We say that a set 𝒮\mathcal{S} of polarized log Fano cone singularities is bounded if there exists finitely many \mathbb{R}-Gorenstein families Bi(𝒳i,𝒟i)BiB_{i}\subseteq(\mathcal{X}_{i},\mathcal{D}_{i})\to B_{i} of klt singularities over finite type bases, each with a fiberwise good 𝕋i\mathbb{T}_{i}-action for some nontrivial algebraic torus 𝕋i\mathbb{T}_{i}, such that every x(X,Δ;ξ)x\in(X,\Delta;\xi) in 𝒮\mathcal{S} is isomorphic to b(𝒳i,b,𝒟i,b;ξb)b\in(\mathcal{X}_{i,b},\mathcal{D}_{i,b};\xi_{b}) for some ii, some bBib\in B_{i} and some ξbN(𝕋i)\xi_{b}\in N(\mathbb{T}_{i})_{\mathbb{R}}.

A priori, it may happen that a set of log Fano cone singularities is bounded as a set of sub-pairs, but becomes unbounded when we take into account the log Fano cone structure. Nonetheless, the two boundedness notions coincide if the volume ratios are bounded away from zero and the coefficients belong to a fixed finite set.

Lemma 2.17.

Let θ>0\theta>0 and let I[0,1]I\subseteq[0,1] be a finite set. Let 𝒮\mathcal{S} be a set of polarized log Fano cone singularities x(X,Δ;ξ)x\in(X,\Delta;\xi) with coefficients in II and Θ(X,Δ;ξ)θ\Theta(X,\Delta;\xi)\geq\theta. Assume that the underlying set of pairs is bounded. Then 𝒮\mathcal{S} is bounded.

Proof.

Let f:(𝒳,𝒟)Bf\colon(\mathcal{X},\mathcal{D})\to B be a family of pairs over a finite type base such that for any x(X,Δ;ξ)x\in(X,\Delta;\xi) in 𝒮\mathcal{S}, we have an isomorphism (X,Supp(Δ))(𝒳b,Supp(𝒟b))(X,\mathrm{Supp}(\Delta))\cong(\mathcal{X}_{b},\mathrm{Supp}(\mathcal{D}_{b})) for some bBb\in B. After base change along 𝒳B\mathcal{X}\to B and possibly stratifying BB, we may assume that ff admits a section σ:B𝒳\sigma\colon B\to\mathcal{X} so that the above isomorphism induces an isomorphism

(x(X,Supp(Δ)))(σ(b)(𝒳b,Supp(𝒟b))).\left(x\in(X,\mathrm{Supp}(\Delta))\right)\cong\left(\sigma(b)\in(\mathcal{X}_{b},\mathrm{Supp}(\mathcal{D}_{b}))\right).

Since the coefficients belong to the finite set II, we may also assign coefficients to 𝒟\mathcal{D} and assume that (X,Δ)(𝒳b,𝒟b)(X,\Delta)\cong(\mathcal{X}_{b},\mathcal{D}_{b}). After these reductions, by [Kol-moduli-book, Lemma 4.44] (or rather its proof) and inversion of adjunction, we know that there exists a finite collection of locally closed subset BiB_{i} of BB such that the family (𝒳,𝒟)(\mathcal{X},\mathcal{D}) becomes \mathbb{R}-Gorenstein after base change to iBi\sqcup_{i}B_{i} and enumerates exactly all the klt fibers of B(𝒳,𝒟)BB\subseteq(\mathcal{X},\mathcal{D})\to B (cf. the last part of the proof of [Z-mld^K-2, Theorem 3.1]). Thus by replacing BB with iBi\sqcup_{i}B_{i}, we may assume that B(𝒳,𝒟)BB\subseteq(\mathcal{X},\mathcal{D})\to B is an \mathbb{R}-Gorenstein family of klt singularities to begin with. Since Θ(X,Δ;ξ)θ\Theta(X,\Delta;\xi)\geq\theta by assumption, together with [Z-mld^K-2, Lemma 2.15 and Theorem 3.1], we then see that the set 𝒮\mathcal{S} is bounded as a set of log Fano cone singularities. ∎

We also recall the definition of log birational boundedness. For an \mathbb{R}-divisor GG, we denote its positive part by G+G^{+} and negative part by GG^{-}, i.e. G=G+GG=G^{+}-G^{-} where G+G^{+}, GG^{-} are effective without common components.

Definition 2.18.

Let (X,G)(X,G) and (Y,Σ)(Y,\Sigma) be projective sub-pairs. We say that (Y,Σ)(Y,\Sigma) log birationally dominates (X,G)(X,G) if there exist a birational map φ:YX\varphi\colon Y\dashrightarrow X such that Supp(Σ)\mathrm{Supp}(\Sigma) contains the birational transform of Supp(G)\mathrm{Supp}(G) and the exceptional divisors of φ\varphi, i.e. Supp(Σ)Supp(φ1G)+Ex(φ){\rm Supp}(\Sigma)\supseteq{\rm Supp}(\varphi^{-1}_{*}G)+{\rm Ex}(\varphi). We say that (Y,Σ)(Y,\Sigma) log birationally dominates (X,G)(X,G) effectively if in addition the φ1\varphi^{-1}-exceptional divisors are contained in Supp(G)\mathrm{Supp}(G^{-}).

We will say (Y,Σ)(Y,\Sigma) log birationally dominates (X,G)(X,G) (effectively) through φ\varphi if we want to specify the birational map φ:XY\varphi\colon X\dashrightarrow Y.

Note that if (Y,Σ)(Y,\Sigma) log birationally dominates (X,G)(X,G) with GG being \mathbb{R}-Cartier, and GG^{\prime} is the birational pullback of GG, then Supp(G)Supp(Σ)\mathrm{Supp}(G^{\prime})\subseteq\mathrm{Supp}(\Sigma).

Definition 2.19.

Let 𝒞\mathcal{C} be a set of projective sub-pairs and let 𝒫\mathcal{P} be a set of projective pairs. We say that 𝒫\mathcal{P} log birationally dominates 𝒞\mathcal{C} (resp. log birationally dominates 𝒞\mathcal{C} effectively) if any (X,G)𝒞(X,G)\in\mathcal{C} is log birationally (resp. log birationally and effectively) dominated by some (Y,Σ)𝒫(Y,\Sigma)\in\mathcal{P}.

We say that 𝒞\mathcal{C} is log birationally bounded if there exists a bounded set 𝒫\mathcal{P} of pairs that log birationally dominates 𝒞\mathcal{C} (cf. [HMX-BirAut, Definition 2.4.1]).

The following criterion for log birational boundedness is a special case of [HMX-BirAut, Lemma 3.2] or [Birkar-bab-1, Proposition 4.4].

Proposition 2.20.

Let nn be a positive integer and let c0,c1>0c_{0},c_{1}>0. Let 𝒞\mathcal{C} be the set of pairs (X,Δ+Γ)(X,\Delta+\Gamma) of dimension nn such that

  • (KX+Δ)-(K_{X}+\Delta) is ample,

  • the non-zero coefficients of Δ\Delta are at least c0c_{0},

  • Γ\Gamma is a \mathbb{Q}-Cartier, effective, nef Weil divisor,

  • |Γ||\Gamma| defines a birational map and vol(Γ)c1\mathrm{vol}(\Gamma)\leq c_{1}.

Then 𝒞\mathcal{C} is log birationally bounded. More precisely, there exists a bounded set 𝒫\mathcal{P} of projective log smooth pairs (Y,Σ)(Y,\Sigma) depending only on n,c0,c1n,c_{0},c_{1} such that the following are satisfied: For any (X,Δ+Γ)𝒞(X,\Delta+\Gamma)\in\mathcal{C}, there exist some log smooth pair (Y,Σ)𝒫(Y,\Sigma)\in\mathcal{P} and a birational map φ:YX\varphi\colon Y\dashrightarrow X such that:

  1. (1)

    (Y,Σ)(Y,\Sigma) log birationally dominates (X,Δ+Γ)(X,\Delta+\Gamma) through φ\varphi.

  2. (2)

    There exists some effective and big Cartier divisor AΣA\leq\Sigma on YY such that |A||A| is base point free and |Γ(φ1)A||\Gamma-(\varphi^{-1})^{*}A|\neq\emptyset.

Proof.

Log birational boundedness follows from [Birkar-bab-1, Proposition 4.4(1)], which also gives the property (1). Property (2) follows from [Birkar-bab-1, Proposition 4.4(3)] (or from the construction of the bounded set 𝒫\mathcal{P} in loc. cit., as AA is simply the birational transform of the movable part of |Γ||\Gamma|). ∎

3. Boundedness

In this section, we give the proof of our main theorems. The main statement is Theorem 1.4, and we divide its proof into three parts, as outlined in 1.6.

3.1. Log birational boundedness

To prove Theorem 1.4, we first aim to show that the log Fano cone singularities have log birationally bounded projective orbifold cone compactifications. From Section 2.4, we have seen that the local alpha invariants of the projective orbifold cones are bounded from below away from the divisor at infinity, and their volumes are also bounded. The situation is thus somewhat similar to those of [Jia-Kss-Fano-bdd]. Our first step is to refine some of the arguments in [Jia-Kss-Fano-bdd] to prove an effective birationality result. Log birationally boundedness is then an immediate consequence.

In the global (Fano) setting, [Jia-Kss-Fano-bdd] proceeds as follow. In order to show that |mKX||-mK_{X}| defines a birational map for some fixed integer mm, one aims to create isolated non-klt centers on the Fano variety XX. The main observation from [Jia-Kss-Fano-bdd] is that if both the alpha invariant and the volume are bounded from below, then the volumes of any covering family of subvarieties are also bounded from below, and this allows one to cut down the dimension of the non-klt centers. The next two lemmas show that this strategy still work if we replace the global alpha invariant by the local one.

Lemma 3.1 (cf. [Jia-Kss-Fano-bdd, Lemma 3.1]).

Let XX be a normal projective variety of dimension nn and LL a big \mathbb{R}-Cartier \mathbb{R}-divisor on XX. Let f:YTf\colon Y\to T be a projective morphism and μ:YX\mu\colon Y\to X a surjective morphism. Assume that a general fiber FF of ff is of dimension kk and is mapped birationally onto its image GG in XX. Then for any general smooth point xXx\in X, we have

vol(L|G)αx(L)nk(nk)(nk)nkvol(L).\mathrm{vol}(L|_{G})\geq\frac{\alpha_{x}(L)^{n-k}}{\binom{n}{k}(n-k)^{n-k}}\mathrm{vol}(L).
Proof.

This follows from [Jia-Kss-Fano-bdd, Lemma 3.1] with some small modifications. We sketch the argument for the reader’s convenience. By perturbing the coefficients and rescaling, we may assume that LL is Cartier. Replacing ff by its Stein factorization, we may assume that FF is connected. We also assume that YY and TT are smooth by taking log resolution. Moreover, by the Bertini Theorem we may replace TT by a general complete intersection subvariety and assume that μ\mu is generically finite. In particular, it is étale at the generic point of FF (since FF is a general fiber). We may also choose FF so that xGx\in G. Clearly it suffices to consider the case when k<nk<n.

Let t=f(F)Tt=f(F)\in T and l+l\in\mathbb{Q}_{+}. By a direct calculation (using that FF has trivial normal bundle in YY), we have

h0(Y,μLm𝒪Y/Flm)h0(F,μLm)h0(𝒪T/𝔪tlm)+O(mn1)h^{0}(Y,\mu^{*}L^{\otimes m}\otimes\mathcal{O}_{Y}/\mathcal{I}_{F}^{lm})\leq h^{0}(F,\mu^{*}L^{\otimes m})\cdot h^{0}(\mathcal{O}_{T}/\mathfrak{m}_{t}^{lm})+O(m^{n-1})

for sufficiently large and divisible integers mm. Hence if

(3.1) vol(L)n!>vol(L|G)lnkk!(nk)!,\frac{\mathrm{vol}(L)}{n!}>\frac{\mathrm{vol}(L|_{G})\cdot l^{n-k}}{k!\cdot(n-k)!},

then h0(X,mL)>h0(Y,μLm𝒪Y/Flm)h^{0}(X,mL)>h^{0}(Y,\mu^{*}L^{\otimes m}\otimes\mathcal{O}_{Y}/\mathcal{I}_{F}^{lm}) for m0m\gg 0. It follows that there exists some effective divisor DLD\sim_{\mathbb{Q}}L such that multF(μD)l\mathrm{mult}_{F}(\mu^{*}D)\geq l; as μ\mu is étale at the generic point of FF, this also implies that multGDl\mathrm{mult}_{G}D\geq l and therefore

αx(L)lctx(D)nkl\alpha_{x}(L)\leq\mathrm{lct}_{x}(D)\leq\frac{n-k}{l}

as GG has codimension nkn-k in XX. This holds for every ll that satisfies (3.1); the lemma then follows. ∎

Lemma 3.2 (cf. [Jia-Kss-Fano-bdd, Theorem 1.5]).

Let ε,α>0\varepsilon,\alpha>0. Let XX be a normal projective variety of dimension nn, and let LL be an ample \mathbb{Q}-Cartier \mathbb{Q}-divisor on XX such that (Ln)ε(L^{n})\geq\varepsilon and αx(L)α\alpha_{x}(L)\geq\alpha for a general point xXx\in X. Then there exists some positive integer m0=m0(n,ε,α)m_{0}=m_{0}(n,\varepsilon,\alpha) such that |KX+mL||K_{X}+\lceil mL\rceil| defines a birational map for all mm0m\geq m_{0}.

Proof.

The assumptions and Lemma 3.1 imply that there exists some m0=m0(n,ε,α)>0m_{0}=m_{0}(n,\varepsilon,\alpha)>0 such that vol(mL|G)>(2k)k\mathrm{vol}(mL|_{G})>(2k)^{k} (where k=dimGk=\dim G) for any general member GG of a covering family of positive dimensional subvarieties of XX and all mm0m\geq m_{0}. The argument in [Birkar-bab-1, 2.31(2)] implies that mLmL is potentially birational ([HMX-ACC, Definition 3.5.3]), and then the lemma follows from [HMX-BirAut, Lemma 2.3.4]. ∎

We can now prove the effective birationality of the orbifold cone compactifications.

Proposition 3.3.

Fix some positive integer nn, a finite coefficient set I[0,1]I\subseteq[0,1]\cap\mathbb{Q}, and some positive real numbers ε,θ>0\varepsilon,\theta>0. Then there exist some positive integer m=m(n,ε,θ,I)m=m(n,\varepsilon,\theta,I) such that mIm\cdot I\subseteq\mathbb{N} and for any quasi-regular polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) with

(3.2) dimX=n, Coeff(Δ)I, vol^(X,Δ;ξ)ε and Θ(X,Δ;ξ)θ,\dim X=n,\mbox{\ \ }\mathrm{Coeff}(\Delta)\subseteq I,\mbox{\ \ }\widehat{\rm vol}(X,\Delta;\xi)\geq\varepsilon\mbox{\ \ and \ \ }\Theta(X,\Delta;\xi)\geq\theta,

the following statements hold for its orbifold cone compactification (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}):

  1. (1)

    The pair (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) has an mm-complement.

  2. (2)

    The linear system |m(KX¯+Δ¯+V)||-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})| defines a birational map.

Here we define an mm-complement of a pair (X,D)(X,D) as an effective \mathbb{Q}-divisor Γ\Gamma such that (X,D+Γ)(X,D+\Gamma) is log canonical and m(KX+D+Γ)0m(K_{X}+D+\Gamma)\sim 0.

Proof.

Item (1) is the boundedness of complements proved in [Birkar-bab-1, Theorem 1.7]. Let us prove (2). Let L=(KX¯+Δ¯+V)L=-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}). By Lemma 2.9 and our assumption on the local volume, we have vol(L)ε\mathrm{vol}(L)\geq\varepsilon. By Lemmas 2.11 and 2.12, there exists some positive number α=α(n,θ)>0\alpha=\alpha(n,\theta)>0 such that αx1(L)=αx1(X¯,Δ¯+V)α\alpha_{x_{1}}(L)=\alpha_{x_{1}}(\overline{X},\overline{\Delta}+V_{\infty})\geq\alpha for all x1Xx_{1}\in X. Thus Lemma 3.2 guarantees the existence of some positive integer m=m(n,ε,θ,I)m=m(n,\varepsilon,\theta,I) such that mΔm\Delta has integer coefficients and

|KX¯+(m+1)L|=|mLV||K_{\overline{X}}+\lceil(m+1)L\rceil|=|mL-V_{\infty}|

defines a birational map. It follows that |mL||mL| also defines a birational map. By taking common multiples, we get a positive integer m=m(n,ε,θ,I)>0m=m(n,\varepsilon,\theta,I)>0 such that (1) and (2) simultaneously hold. ∎

From Lemmas 2.9, we know that

vol(M)mnvol((KX¯+Δ¯+V))(mn)nθ1.\mathrm{vol}(M)\leq m^{n}\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}))\leq(mn)^{n}\theta^{-1}.

Thus by Proposition 2.20, this immediately implies that the set of orbifold cone compactifications of quasi-regular log Fano cone singularities satisfying (3.2) is log birationally bounded. Choose some (log bounded) birational model φ:(Y,Σ)X¯\varphi\colon(Y,\Sigma)\dashrightarrow\overline{X} of the projective orbifold cone X¯\overline{X}. Our next task is to reconstruct XX from YY.

3.2. Boundedness in codimension one

To reconstruct XX, we need to first understand the exceptional divisors of the birational map φ1:X¯Y\varphi^{-1}\colon\overline{X}\dashrightarrow Y. Note that the infinity divisor VV_{\infty} is typically φ1\varphi^{-1}-exceptional, and since it depends on the choice of the Reeb vector ξ\xi, we will not have much control over it. The next result shows that other than VV_{\infty}, the remaining φ1\varphi^{-1}-exceptional divisors are essentially “bounded”. To state it precisely let us make one more definition.

Definition 3.4.

Let (X,Δ)(X,\Delta) be a pair and let NN be a positive integer. A sub-klt NN-complement of (X,Δ)(X,\Delta) is a (not necessarily effective) \mathbb{Q}-divisor GG on XX such that N(KX+Δ+G)0N(K_{X}+\Delta+G)\sim 0 and (X,Δ+G)(X,\Delta+G) is sub-klt.

Proposition 3.5.

Fix some positive integer nn, a finite coefficient set I[0,1]I\subseteq[0,1]\cap\mathbb{Q}, and some positive real numbers ε,θ>0\varepsilon,\theta>0. There exist a bounded set 𝒫\mathcal{P} of projective log smooth pairs (Y,Σ)(Y,\Sigma) and a positive integer N=N(n,ε,θ,I)N=N(n,\varepsilon,\theta,I), such that the following holds for any quasi-regular polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) satisfying (3.2):

  1. (1)

    (X¯,Δ¯)(\overline{X},\overline{\Delta}) has a sub-klt NN-complement GG such that Supp(G)VSupp(G)\mathrm{Supp}(G^{-})\subseteq V_{\infty}\subseteq\mathrm{Supp}(G).

  2. (2)

    There exists some (Y,Σ)𝒫(Y,\Sigma)\in\mathcal{P} that log birationally dominates (X¯,Δ¯+G)(\overline{X},\overline{\Delta}+G) effectively ((Definition 2.18)).

Informally, the implication (2) means that the log Fano cone singularities are bounded in codimension one. The existence of a sub-klt bounded complement will later be used to ensure that certain MMPs exist and terminate.

The proof of the proposition is inspired by the proof of [HMX-ACC, Theorem 1.6] and [Birkar-bab-1]*Proposition 7.13. The main technical part is to construct a bounded sub-klt complement of (X¯,Δ¯)(\overline{X},\overline{\Delta}) satisfying certain conditions. We first discuss how the existence of such a complement affects boundedness in codimension one.

Lemma 3.6.

Let NN be a positive integer and let 𝒞\mathcal{C} be a set of projective sub-klt sub-pairs (X,G)(X,G) satisfying N(KX+G)0N(K_{X}+G)\sim 0. Assume that 𝒞\mathcal{C} is log birationally bounded. Then there exists a bounded set 𝒫\mathcal{P} of projective log smooth pairs that log birationally dominates 𝒞\mathcal{C} effectively ((Definition 2.19)).

Proof.

By assumption, we may choose a bounded set 𝒫\mathcal{P} of projective pairs that log birationally dominates 𝒞\mathcal{C}. After passing to a log resolution, we may assume that 𝒫\mathcal{P} is bounded set of log smooth pairs. For any (X,G)𝒞(X,G)\in\mathcal{C}, let (Y,Σ)𝒫(Y,\Sigma)\in\mathcal{P} be a log smooth pair that log birationally dominates (X,G)(X,G) through a birational map φ:YX\varphi\colon Y\dashrightarrow X. Write φ(KX+G)=KY+GY\varphi^{*}(K_{X}+G)=K_{Y}+G_{Y}. Then GYG_{Y} is a sub-klt NN-complement of YY supported on Σ\Sigma (see the remark after Definition 2.18). In particular, GY(11N)ΣG_{Y}\leq(1-\frac{1}{N})\Sigma. The discrepancy of any φ1\varphi^{-1}-exceptional divisor FF must satisfy

a(F;Y,(11N)Σ)a(F;Y,GY)=a(F;X,G)0,a(F;Y,(1-\frac{1}{N})\Sigma)\leq a(F;Y,G_{Y})=a(F;X,G)\leq 0,

unless FSupp(G)F\subseteq\mathrm{Supp}(G^{-}). Since (Y,Σ)(Y,\Sigma) is log smooth, the pair (Y,(11N)Σ)(Y,(1-\frac{1}{N})\Sigma) is klt, hence there are only finitely many exceptional divisors with discrepancy at most 0, and these can be extracted via successive blowups along the strata of Σ\Sigma. In other words, up to replacing the bounded set 𝒫\mathcal{P} of log smooth pairs, we may assume that the only φ1\varphi^{-1}-exceptional divisors are among the components of GG^{-}, thus 𝒫\mathcal{P} also dominates 𝒞\mathcal{C} effectively. ∎

We next construct the sub-klt bounded complements on the projective orbifold cones.

Lemma 3.7.

There exists a positive integer N=N(n,ε,θ,I)N=N(n,\varepsilon,\theta,I) such that for any quasi-regular polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) satisfying (3.2), there exists a sub-klt NN-complement GG of (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) such that

  1. (1)

    Supp(G)VSupp(G+V)\mathrm{Supp}(G^{-})\subseteq V_{\infty}\subseteq\mathrm{Supp}(G+V_{\infty}), and

  2. (2)

    (KX¯+Δ¯+V)12G+-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})-\frac{1}{2}G^{+} is ample.

Proof.

We follow the argument of [Birkar-bab-1]*Proposition 7.13. Let m=m(n,ε,θ,I)>0m=m(n,\varepsilon,\theta,I)>0 be the integer given by Proposition 3.3. In particular, there exists an mm-complement Γ1m|M|\Gamma\in\frac{1}{m}|M| where M=m(KX¯+Δ¯+V)M=-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty}). By Proposition 2.20 as in the remark right after Proposition 3.3, we find a bounded set 𝒫\mathcal{P} of projective log smooth pairs (Y,Σ)(Y,\Sigma) depending only on n,ε,θ,In,\varepsilon,\theta,I, such that for any x(X,Δ;ξ)x\in(X,\Delta;\xi) satisfying (3.2) and any mm-complement Γ\Gamma as above, there exists some log smooth pair (Y,Σ)𝒫(Y,\Sigma)\in\mathcal{P} and some birational map φ:X¯Y\varphi\colon\overline{X}\dashrightarrow Y such that:

  1. (1)

    (Y,Σ)(Y,\Sigma) log birationally dominates (X¯,Δ¯+V+Γ)(\overline{X},\overline{\Delta}+V_{\infty}+\Gamma) through φ\varphi.

  2. (2)

    There exists some effective and big Cartier divisor AΣA\leq\Sigma on YY such that |A||A| is base point free and |M(φ1)A||M-(\varphi^{-1})^{*}A|\neq\emptyset.

Define ΓY\Gamma_{Y} by the crepant pullback formula

KY+ΓY=φ(KX¯+Δ¯+V+Γ)0.K_{Y}+\Gamma_{Y}=\varphi^{*}(K_{\overline{X}}+\overline{\Delta}+V_{\infty}+\Gamma)\sim_{\mathbb{Q}}0.

Since (Y,Σ)(Y,\Sigma) belongs to a bounded family, We can also choose some positive integer m0m_{0} depending only on 𝒫\mathcal{P}, and some \mathbb{Q}-divisor B=B+BB=B^{+}-B^{-} in a bounded family where

B+|A|andm0B|m0A|,B^{+}\in|A|\quad\mathrm{and}\quad m_{0}B^{-}\in|m_{0}A|,

such that B+B^{+} is in a general position (by Bertini theorem) and ΣSupp(B)\Sigma\subseteq\mathrm{Supp}(B^{-}) (this is possible since AA is big). By construction, Supp(ΓY)Σ\mathrm{Supp}(\Gamma_{Y})\subseteq\Sigma and (Y,ΓY)(Y,\Gamma_{Y}) is sub-lc. Hence the pair (Y,ΓY+B)(Y,\Gamma_{Y}+B) is sub-klt and KY+ΓY+B0K_{Y}+\Gamma_{Y}+B\sim_{\mathbb{Q}}0. Its crepant pullback to X¯\overline{X} is (X¯,Δ¯+V+Γ+BX)(\overline{X},\overline{\Delta}+V_{\infty}+\Gamma+B_{X}) where

BX+=(φ1)B+ and BX=(φ1)B.B_{X}^{+}=(\varphi^{-1})^{*}B^{+}\mbox{\ \ and \ \ }B_{X}^{-}=(\varphi^{-1})^{*}B^{-}\,.

In particular, as B+B^{+} and m0Bm_{0}B^{-} are both Cartier, the coefficients of BXB_{X} belongs to 1m0\frac{1}{m_{0}}\mathbb{Z}.

Choose some R|M(φ1)A|R\in|M-(\varphi^{-1})^{*}A|. We may write

BX+R=λV+mCB_{X}^{-}+R=\lambda V_{\infty}+mC

where VSupp(C)V_{\infty}\not\subseteq\mathrm{Supp}(C). Note that BX(φ1)AB_{X}^{-}\sim_{\mathbb{Q}}(\varphi^{-1})^{*}A and thus

BX+RM=m(KX¯+Δ¯+V),B_{X}^{-}+R\sim_{\mathbb{Q}}M=-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})\,,

hence Cμ(KX¯+Δ¯+V)C\sim_{\mathbb{Q}}-\mu(K_{\overline{X}}+\overline{\Delta}+V_{\infty}) for some μ1\mu\leq 1. We also note that the coefficients of CC are contained in 1mm0\frac{1}{mm_{0}}\mathbb{Z}.

By Lemma 2.12, the orbifold base satisfies α(V,ΔV)α0\alpha(V,\Delta_{V})\geq\alpha_{0} for some positive constant α0=α0(n,ε,θ,I)>0\alpha_{0}=\alpha_{0}(n,\varepsilon,\theta,I)>0. We may assume that α0<1\alpha_{0}<1. By adjunction, this implies that (X¯,Δ¯+V+α0C)(\overline{X},\overline{\Delta}+V_{\infty}+\alpha_{0}C) is log canonical in a neighbourhood of VV_{\infty}. By Lemma 2.11, we also know that (X¯,Δ¯+V+α0C)(\overline{X},\overline{\Delta}+V_{\infty}+\alpha_{0}C) is log canonical away from VV_{\infty}. Hence the pair (X¯,Δ¯+V+α0C)(\overline{X},\overline{\Delta}+V_{\infty}+\alpha_{0}C) is log canonical everywhere. By [Birkar-bab-1]*Theorem 1.7, it has an NN-complement C0C^{\prime}\geq 0 for some positive integer NN that only depends on the dimension and the coefficients; tracing through the construction above, this in turn means that NN only depends on n,ε,θn,\varepsilon,\theta and the finite set II. Now consider the linear combination

G:=t(Γ+BX)+(1t)(α0C+C).G:=t(\Gamma+B_{X})+(1-t)(\alpha_{0}C+C^{\prime}).

for some fixed rational number t(0,1)t\in(0,1) such that mt(1t)α0<1mt\leq(1-t)\alpha_{0}<1 and m0t\frac{m_{0}}{t}\not\in\mathbb{Z}. As multVBX1m0\mathrm{mult}_{V_{\infty}}B_{X}\in\frac{1}{m_{0}}\mathbb{Z} and VSupp(Γ+C+C)V_{\infty}\not\subseteq\mathrm{Supp}(\Gamma+C+C^{\prime}), the second condition on tt simply guarantees that multVG1\mathrm{mult}_{V_{\infty}}G\neq-1 and hence VSupp(G+V)V_{\infty}\subseteq\mathrm{Supp}(G+V_{\infty}). Since GG is a convex combination of bounded complements of (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) and (X,V+Γ+BX)(X,V_{\infty}+\Gamma+B_{X}) is sub-klt, we see that GG is a sub-klt NN-complement of (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) after possibly enlarging NN. Moreover, as mt(1t)α0mt\leq(1-t)\alpha_{0} by our choice of tt and BXmCB_{X}^{-}\leq mC away from VV_{\infty}, we have tBX(1t)α0CtB_{X}^{-}\leq(1-t)\alpha_{0}C away from VV_{\infty} and therefore Supp(G)V\mathrm{Supp}(G^{-})\subseteq V_{\infty}. In particular, the resulting sub-klt complement GG satisfies (1).

By construction, GtBXG^{-}\leq tB_{X}^{-} and MBXM-B_{X}^{-} is pseudo-effective. Thus

(KX¯+Δ¯+V)G(1mt)M+t(MBX)+(tBXG)-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})-G^{-}\sim_{\mathbb{Q}}\left(\frac{1}{m}-t\right)M+t(M-B_{X}^{-})+(tB_{X}^{-}-G^{-})

is big. But since both KX¯+Δ¯K_{\overline{X}}+\overline{\Delta} and GG^{-} are proportional to the ample divisor VV_{\infty}, this implies the left hand side above is in fact ample. As

G=G+G(KX¯+Δ¯+V),G=G^{+}-G^{-}\sim_{\mathbb{Q}}-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})\,,

it follows that (KX¯+Δ¯+V)12G+-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})-\frac{1}{2}G^{+} is also ample, proving (2). ∎

We may now return to the proof of Proposition 3.5.

Proof of Proposition 3.5.

Let N=N(n,ε,θ,I)N=N(n,\varepsilon,\theta,I) be the positive integer from Lemma 3.7. Then for any x(X,Δ;ξ)x\in(X,\Delta;\xi) satisfying (3.2), there exists a sub-klt NN-complement G0G_{0} of (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) such that

  1. (a)

    Supp(G0)VSupp(G0+V)\mathrm{Supp}(G_{0}^{-})\subseteq V_{\infty}\subseteq\mathrm{Supp}(G_{0}+V_{\infty}), and

  2. (b)

    (KX¯+Δ¯+V)12G0+-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})-\frac{1}{2}G_{0}^{+} is ample.

In particular, part (1) of the proposition is satisfied by G=G0+VG=G_{0}+V_{\infty}. Possibly replacing NN by a larger multiple, we may assume, by Proposition 3.3, that |N(KX¯+Δ¯+V)||-N(K_{\overline{X}}+\overline{\Delta}+V_{\infty})| defines a birational map. It follows that

|NG0+|=|N(KX¯+Δ¯+V)+NG0||NG_{0}^{+}|=|-N(K_{\overline{X}}+\overline{\Delta}+V_{\infty})+NG_{0}^{-}|

also defines a birational map.

Condition (b) above together with Lemma 2.9 implies that

vol(G0+)2nvol((KX¯+Δ¯+V))(2n)nθ1.\mathrm{vol}(G_{0}^{+})\leq 2^{n}\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}))\leq(2n)^{n}\theta^{-1}.

By Proposition 2.20, we deduce that there exists a bounded set 𝒫\mathcal{P} of projective log smooth pairs, such that for any quasi-regular log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) satisfying (3.2), there exists some (Y,Σ)𝒫(Y,\Sigma)\in\mathcal{P} that log birationally dominates (X¯,Δ¯+V+G0+)(\overline{X},\overline{\Delta}+V_{\infty}+G_{0}^{+}). Using condition (a), we see that (Y,Σ)(Y,\Sigma) also log birationally dominates (X¯,Δ¯+G)(\overline{X},\overline{\Delta}+G). In particular, the sub-klt pair (X¯,Δ¯+G)(\overline{X},\overline{\Delta}+G) belongs to a log birationally bounded set. But then by Lemma 3.6, after possibly replacing the bounded set 𝒫\mathcal{P}, we may further assume that (Y,Σ)(Y,\Sigma) log birationally dominates (X¯,Δ¯+G)(\overline{X},\overline{\Delta}+G) effectively. This implies part (2) of the proposition. ∎

3.3. From boundedness in codimension one to boundedness

Finally, we shall recover the log Fano cone singularity XX from its (modified) birational model YY given by the previous subsection (it is important to note that we will not attempt to recover the projective orbifold cone X¯\overline{X}, which does not belong to a bounded family). The basic strategy is as follows. Since XX is affine, it suffices to recover its section ring from YY. Using the birational model YY, we will identify a big open subset of XX with an open subset UU of YY, and the question is to find the section ring Γ(𝒪U)\Gamma(\mathcal{O}_{U}). If DD is an effective divisor with support YUY\setminus U, we may try to run a DD-MMP on YY and construct its ample model (Y¯,D¯)(\overline{Y},\overline{D}). Then U=Y¯D¯U^{\prime}=\overline{Y}\setminus\overline{D} is affine since D¯\overline{D} is ample, and Γ(𝒪U)\Gamma(\mathcal{O}_{U}) is simply the section ring of UU^{\prime}.

Turning to more details, we begin with some general setup. Let XX be an affine normal variety, let X¯\overline{X} be a normal projective compactification, and let VV_{\infty} be the divisorial part of X¯X\overline{X}\setminus X (a typical example is the orbifold cone compactifications we consider in previous sections). Let φ:YX¯\varphi\colon Y\dashrightarrow\overline{X} be a birational map with YY proper. Let Σ0\Sigma_{0} be the sum of φ1V\varphi^{-1}_{*}V_{\infty} and the exceptional divisors of φ\varphi, and let DD be an effective \mathbb{Q}-Cartier \mathbb{Q}-divisor on YY such that Supp(D)=Σ0\mathrm{Supp}(D)=\Sigma_{0}. A birational contraction g:YY¯g\colon Y\dashrightarrow\overline{Y} is called an ample model of DD if Y¯\overline{Y} is proper, gDg_{*}D is \mathbb{Q}-Cartier ample, and DggDD\geq g^{*}g_{*}D.

Lemma 3.8.

Assume that all the φ1\varphi^{-1}-exceptional divisors are contained in VV_{\infty}, and the ample model g:YY¯g\colon Y\dashrightarrow\overline{Y} of DD exists. Then the composition ψ=gφ1:X¯Y¯\psi=g\circ\varphi^{-1}\colon\overline{X}\dashrightarrow\overline{Y} induces an isomorphism XY¯gΣ0X\cong\overline{Y}\setminus g_{*}\Sigma_{0}.

Proof.

Let ψ:X¯Y¯\psi\colon\overline{X}\dasharrow\overline{Y}, U=Y¯gΣ0U^{\prime}=\overline{Y}\setminus g_{*}\Sigma_{0} and let U=YΣ0U=Y\setminus\Sigma_{0}. Since gDg_{*}D is ample, its complement UU^{\prime} is affine. In order to prove the lemma, it suffices to show that ψ\psi induces an isomorphism Γ(𝒪X)Γ(𝒪U)\Gamma(\mathcal{O}_{X})\cong\Gamma(\mathcal{O}_{U^{\prime}}). For this we first show that the induced birational map (φ1)|X:XU(\varphi^{-1})|_{X}\colon X\dashrightarrow U is an isomorphism over some big open sets of both XX and UU.

To see this, note that the exceptional divisors of φ\varphi are contained in Σ0\Sigma_{0}, while the exceptional divisors of φ1\varphi^{-1} are contained in VV_{\infty}. We also have Supp(φΣ0)V\mathrm{Supp}(\varphi_{*}\Sigma_{0})\subseteq V_{\infty} and Supp(φ1V)Σ0\mathrm{Supp}(\varphi^{-1}_{*}V_{\infty})\subseteq\Sigma_{0} by construction. Thus the complement of all the exceptional locus contains a big open subset of both XX and UU, and φ\varphi is an isomorphism over this open set. In particular, we have Γ(𝒪X)Γ(𝒪U)\Gamma(\mathcal{O}_{X})\cong\Gamma(\mathcal{O}_{U}).

Let f:WYf\colon W\to Y, h:WY¯h\colon W\to\overline{Y} be a common resolution. Since gD=hfDg_{*}D=h_{*}f^{*}D is ample and DggDD\geq g^{*}g_{*}D by assumption, we have fDhgDf^{*}D\geq h^{*}g_{*}D by the negativity lemma. This implies that Supp(fD)=h1(Supp(gD))\mathrm{Supp}(f^{*}D)=h^{-1}(\mathrm{Supp}(g_{*}D)) and therefore the induced morphism

WSupp(fD)Y¯Supp(gD)=UW\setminus\mathrm{Supp}(f^{*}D)\to\overline{Y}\setminus\mathrm{Supp}(g_{*}D)=U^{\prime}

is proper, hence they have the same global sections. Similarly the morphism

WSupp(fD)YSupp(D)=UW\setminus\mathrm{Supp}(f^{*}D)\to Y\setminus\mathrm{Supp}(D)=U

is proper as well. Thus they induce isomorphisms

Γ(𝒪U)Γ(𝒪WSupp(fD))Γ(𝒪U).\Gamma(\mathcal{O}_{U})\cong\Gamma(\mathcal{O}_{W\setminus\mathrm{Supp}(f^{*}D)})\cong\Gamma(\mathcal{O}_{U^{\prime}}).

Combined with the previous established isomorphism Γ(𝒪X)Γ(𝒪U)\Gamma(\mathcal{O}_{X})\cong\Gamma(\mathcal{O}_{U}), this proves that ψ\psi induces an isomorphism Γ(𝒪X)Γ(𝒪U)\Gamma(\mathcal{O}_{X})\cong\Gamma(\mathcal{O}_{U^{\prime}}) and hence XY¯gΣ0X\cong\overline{Y}\setminus g_{*}\Sigma_{0}. ∎

Now we can put things together to prove the main theorems.

Proof of Theorem 1.4.

By [Z-mld^K-2, Lemma 2.18], after possibly replacing the positive constants ε,θ\varepsilon,\theta and the finite set II, we may assume that I[0,1]I\subseteq[0,1]\cap\mathbb{Q}. By [Z-mld^K-2, Lemma 2.11], after perturbing the Reeb vector ξ\xi and decreasing ε,θ\varepsilon,\theta, we may further assume that the all the polarized log Fano cone singularities in 𝒮\mathcal{S} are quasi-regular.

By Proposition 3.5, there exist a bounded set 𝒫\mathcal{P} of projective log smooth pairs (Y,Σ)(Y,\Sigma) and a positive integer NN, depending only on n,ε,θn,\varepsilon,\theta and II, such that the following holds for the projective orbifold cone compactification (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) of any log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) in 𝒮\mathcal{S}:

  1. (1)

    (X¯,Δ¯)(\overline{X},\overline{\Delta}) has a sub-klt NN-complement GG such that Supp(G)VSupp(G)\mathrm{Supp}(G^{-})\subseteq V_{\infty}\subseteq\mathrm{Supp}(G).

  2. (2)

    There exists some (Y,Σ)𝒫(Y,\Sigma)\in\mathcal{P} that log birationally dominates (X¯,Δ¯+G)(\overline{X},\overline{\Delta}+G) effectively.

Let Σ0Y\Sigma_{0}\subseteq Y be the sum of the birational transform of VV_{\infty} and the exceptional divisors of the birational map φ:YX¯\varphi\colon Y\dashrightarrow\overline{X}. Note that Σ0Σ\Sigma_{0}\subseteq\Sigma. In order to apply Lemma 3.8, let us show that there exists some effective divisor DD with Supp(D)=Σ0\mathrm{Supp}(D)=\Sigma_{0} such that the ample model of DD exists. Once this is achieved, the remaining step is to run a DD-MMP in the bounded family 𝒫\mathcal{P} for some uniform choice of DD.

Let (Y,GY)(Y,G_{Y}) be the crepant pull back of (X,Δ¯+G)(X,\overline{\Delta}+G). Then GYG_{Y} is a sub-klt NN-complement of YY which satisfies Supp(GY)Σ\mathrm{Supp}(G_{Y})\subseteq\Sigma and Supp(GY)Σ0\mathrm{Supp}(G_{Y}^{-})\subseteq\Sigma_{0}. In particular, the coefficients of GYG_{Y} are at most 11N1-\frac{1}{N}.

Consider a new boundary divisor Γ\Gamma on YY as follows: if FF is a prime divisor on YY but is not a component of Σ0\Sigma_{0}, then we set

multF(Γ):=multF(GY)0;\mathrm{mult}_{F}(\Gamma):=\mathrm{mult}_{F}(G_{Y})\geq 0;

if FF is an irreducible component of Σ0\Sigma_{0}, then set

multF(Γ):=max{0,multF(GY)+12N}.\mathrm{mult}_{F}(\Gamma):=\max\left\{0,\mathrm{mult}_{F}(G_{Y})+\frac{1}{2N}\right\}.

Let D:=ΓGYD:=\Gamma-G_{Y}. By construction, both \mathbb{Q}-divisors DD and Γ\Gamma are effective, Supp(Γ)Σ\mathrm{Supp}(\Gamma)\subseteq\Sigma, Supp(D)=Σ0\mathrm{Supp}(D)=\Sigma_{0}, the coefficients of Γ\Gamma are contained in

Λ:={0,12N,22N,,112N}\Lambda:=\left\{0,\frac{1}{2N},\frac{2}{2N},\dots,1-\frac{1}{2N}\right\}

(in particular, as (Y,Σ)(Y,\Sigma) is log smooth, the pair (Y,Γ)(Y,\Gamma) is klt), and we have

KY+Γ(KY+Γ)(KY+GY)D.K_{Y}+\Gamma\sim_{\mathbb{Q}}(K_{Y}+\Gamma)-(K_{Y}+G_{Y})\sim_{\mathbb{Q}}D.

As Supp(φV)Σ0\mathrm{Supp}(\varphi^{*}V_{\infty})\subseteq\Sigma_{0} and VV_{\infty} is ample on X¯\overline{X}, we know that Σ0\Sigma_{0} is big and the same holds for DD as Supp(D)=Σ0\mathrm{Supp}(D)=\Sigma_{0}. In particular, KY+ΓK_{Y}+\Gamma is big. Thus by [BCHM, Theorem 1.2], the ample model g:YY¯g\colon Y\dashrightarrow\overline{Y} of KY+ΓDK_{Y}+\Gamma\sim_{\mathbb{Q}}D exists. By Lemma 3.8, the composition ψ:X¯Y¯\psi\colon\overline{X}\dashrightarrow\overline{Y} induces an isomorphism XY¯gΣ0X\cong\overline{Y}\setminus g_{*}\Sigma_{0}. Since Σ\Sigma contains the birational transform of Δ¯\overline{\Delta} in its support, we also see that Supp(Δ)=(gΣ1)|X\mathrm{Supp}(\Delta)=(g_{*}\Sigma_{1})|_{X} for some reduced divisor Σ1Σ\Sigma_{1}\leq\Sigma.

To summarize, we have proved the following. For any log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) in 𝒮\mathcal{S}, there exist a log smooth pair (Y,Σ)(Y,\Sigma) from the bounded set 𝒫\mathcal{P}, two reduced divisors Σ0,Σ1Σ\Sigma_{0},\Sigma_{1}\leq\Sigma, and an effective divisor Γ\Gamma supported on Σ\Sigma with coefficients in Λ\Lambda, such that KY+ΓK_{Y}+\Gamma is big and

(X,Supp(Δ))(Y¯gΣ0,(gΣ1)|Y¯gΣ0),(X,\mathrm{Supp}(\Delta))\cong(\overline{Y}\setminus g_{*}\Sigma_{0},(g_{*}\Sigma_{1})|_{\overline{Y}\setminus g_{*}\Sigma_{0}}),

where g:YY¯g\colon Y\dashrightarrow\overline{Y} is the ample model of KY+ΓK_{Y}+\Gamma.

Since 𝒫\mathcal{P} is bounded, all the pairs (Y,Σ)(Y,\Sigma) in 𝒫\mathcal{P} arise as the fibers of some bounded family (𝒴,Σ𝒴)B(\mathcal{Y},\Sigma_{\mathcal{Y}})\to B of pairs over a finite type base BB, i.e. there exists bBb\in B such that

(Y,Σ)(𝒴b,Σ𝒴b):=(𝒴,Σ𝒴)×Bb.(Y,\Sigma)\cong(\mathcal{Y}_{b},\Sigma_{\mathcal{Y}_{b}}):=(\mathcal{Y},\Sigma_{\mathcal{Y}})\times_{B}b.

After stratifying BB and performing a base change, we may assume that BB is smooth, (𝒴,Σ𝒴)B(\mathcal{Y},\Sigma_{\mathcal{Y}})\to B is log smooth, and every irreducible component of Σ\Sigma is the restriction of some component of Σ𝒴\Sigma_{\mathcal{Y}}. In particular, there are \mathbb{Q}-divisors Σ0,𝒴,Σ1,𝒴\Sigma_{0,\mathcal{Y}},\Sigma_{1,\mathcal{Y}} and Γ𝒴\Gamma_{\mathcal{Y}} supported on Σ𝒴\Sigma_{\mathcal{Y}} that restricts to Σ0,Σ1\Sigma_{0},\Sigma_{1} and Γ\Gamma on the fiber (Y,Σ)(Y,\Sigma).

To conclude the proof of the boundedness, we note that by [HMX-BirAut, Theorem 1.8], the volume of K𝒴b+Γ𝒴bK_{\mathcal{Y}_{b}}+\Gamma_{\mathcal{Y}_{b}} is locally constant in bBb\in B; moreover, over the components of BB where K𝒴b+Γ𝒴bK_{\mathcal{Y}_{b}}+\Gamma_{\mathcal{Y}_{b}} is big, the relative ample model h:𝒴𝒴¯h\colon\mathcal{Y}\dashrightarrow\overline{\mathcal{Y}} of K𝒴+Γ𝒴K_{\mathcal{Y}}+\Gamma_{\mathcal{Y}} over BB exists, whose restriction over bb yields the ample model gb:𝒴b𝒴¯bg_{b}\colon\mathcal{Y}_{b}\dashrightarrow\overline{\mathcal{Y}}_{b} of K𝒴b+Γ𝒴bK_{\mathcal{Y}_{b}}+\Gamma_{\mathcal{Y}_{b}}. By Noetherian induction, after possibly stratifying BB again, we may assume that the restriction of hΣ0,𝒴h_{*}\Sigma_{0,\mathcal{Y}} (resp. hΣ1,𝒴h_{*}\Sigma_{1,\mathcal{Y}}) to the fiber 𝒴¯b\overline{\mathcal{Y}}_{b} is exactly (gb)Σ0,𝒴b(g_{b})_{*}\Sigma_{0,\mathcal{Y}_{b}} (resp. (gb)Σ1,𝒴b(g_{b})_{*}\Sigma_{1,\mathcal{Y}_{b}}). Set

𝒳:=𝒴¯hΣ0,𝒴andΔ𝒳:=hΣ1,𝒴|𝒳.\mathcal{X}:=\overline{\mathcal{Y}}\setminus h_{*}\Sigma_{0,\mathcal{Y}}\quad\mathrm{and}\quad\Delta_{\mathcal{X}}:=h_{*}\Sigma_{1,\mathcal{Y}}|_{\mathcal{X}}.

There are only finitely many choices of Σ0,𝒴,Σ1,𝒴\Sigma_{0,\mathcal{Y}},\Sigma_{1,\mathcal{Y}} and Γ𝒴\Gamma_{\mathcal{Y}}, as their coefficients belong to the finite set Λ{1}\Lambda\cup\{1\}. This leads to finitely many families (𝒳,Δ𝒳)B(\mathcal{X},\Delta_{\mathcal{X}})\to B as above. By the previous discussion, for any log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) in 𝒮\mathcal{S}, the pair (X,Supp(Δ))(X,\mathrm{Supp}(\Delta)) appears as a fiber of one of the families (𝒳,Δ𝒳)B(\mathcal{X},\Delta_{\mathcal{X}})\to B; therefore, the set of pairs underlying 𝒮\mathcal{S} is bounded. By Lemma 2.17, this implies that 𝒮\mathcal{S} is also a bounded set of log Fano cone singularities. ∎

Proof of Corollary 1.5.

It suffices to prove x(X,Δ;ξ)x\in(X,\Delta;\xi) satisfies the condition of Theorem 1.4, where xx is the vertex of the cone, and ξ\xi corresponds to the 𝔾m\mathbb{G}_{m}-action given by the cone structure.

By Lemma 2.12, Θ(X,Δ;ξ)(min{α0,1})n\Theta(X,\Delta;\xi)\geq(\min\{\alpha_{0},1\})^{n}. One can directly calculate

vol^X,Δ(wtξ)=r(KVΔV)n1ε\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi})=r(-K_{V}-\Delta_{V})^{n-1}\geq\varepsilon

(see e.g. [Z-mld^K-2, Lemma 3.4]), which implies that vol^(x,X,Δ)ε(min{α0,1})n\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon\cdot(\min\{\alpha_{0},1\})^{n}. Thus the set 𝒮\mathcal{S} is bounded by Theorem 1.4. ∎

Proof of Theorem 1.1.

This directly follows from Theorem 1.4 when Θ(X,Δ;ξ)=1\Theta(X,\Delta;\xi)=1. ∎

Proof of Theorem 1.2.

By Theorem 2.6, every klt singularity x(X,Δ)x\in(X,\Delta) has a special degeneration to a K-semistable log Fano cone singularity x0(X0,Δ0;ξv)x_{0}\in(X_{0},\Delta_{0};\xi_{v}) with vol^(x,X,Δ)=vol^(x0,X0,Δ0)\widehat{\rm vol}(x,X,\Delta)=\widehat{\rm vol}(x_{0},X_{0},\Delta_{0}). Moreover, the coefficients of Δ0\Delta_{0} belong to the finite set

I+:={imiaimi,aiI}[0,1].I^{+}:=\{\sum_{i}m_{i}a_{i}\mid m_{i}\in\mathbb{N},a_{i}\in I\}\cap[0,1].

Therefore, Theorem 1.2 follows from Theorem 1.1 and the constructibility of the local volume function in bounded families [Xu-quasi-monomial, Theorem 1.3] (see also [HLQ-vol-ACC, Theorem 3.5] for the real coefficient case). ∎

Remark 3.9.

It is important to understand more about the set Vol^n,I\widehat{\rm Vol}_{n,I}. When I={0}I=\{0\}, it is proved in [LX-cubic-3fold] that the maximal number in Vol^n:=Vol^n,{0}\widehat{\rm Vol}_{n}:=\widehat{\rm Vol}_{n,\{0\}} is nnn^{n}, the local volume of a smooth point. It is conjectured that the second largest number in Vol^n\widehat{\rm Vol}_{n} is 2(n1)n2(n-1)^{n} (the local volume of an ordinary double point), but for now this is known only when n3n\leq 3.

4. Minimal log discrepancy of Kollár components

It is observed in [Z-mld^K-1, Z-mld^K-2] that the boundedness of log Fano cone singularities is closely related to the boundedness of mldK\mathrm{mld}^{\mathrm{K}}, the minimal log discrepancies of Kollár component. The goal of this section is to use the boundedness result from the previous section to prove some upper bounds of mldK\mathrm{mld}^{\mathrm{K}} that only depend on the local volume.

Definition 4.1 ([Xu-pi_1-finite]).

Let x(X,Δ)x\in(X,\Delta) be a klt singularity and let EE be a prime divisor over XX. If there exists a proper birational morphism π:YX\pi\colon Y\to X such that E=π1(x)E=\pi^{-1}(x) is the unique exceptional divisor, (Y,E+ΔY)(Y,E+\Delta_{Y}) is plt and (KY+ΔY+E)-(K_{Y}+\Delta_{Y}+E) is π\pi-ample, we call EE a Kollár component over x(X,Δ)x\in(X,\Delta) and π:YX\pi\colon Y\to X the plt blowup of EE.

Definition 4.2 ([Z-mld^K-1]).

Let x(X,Δ)x\in(X,\Delta) be a klt singularity. The minimal log discrepancy of Kollár components, denoted mldK(x,X,Δ)\mathrm{mld}^{\mathrm{K}}(x,X,\Delta), is the infimum of the log discrepancies AX,Δ(E)A_{X,\Delta}(E) as EE varies among all Kollár components over x(X,Δ)x\in(X,\Delta).

Theorem 4.3 (cf. [Z-mld^K-1, Conjecture 1.7]).

Let nn\in\mathbb{N}, ε>0\varepsilon>0 and let I[0,1]I\subseteq[0,1] be a finite set. Then there exists some constant A>0A>0 depending only on n,ε,In,\varepsilon,I such that

mldK(x,X,Δ)A\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq A

for any nn-dimensional klt singularity x(X,Δ)x\in(X,\Delta) with Coeff(Δ)I{\rm Coeff}(\Delta)\subseteq I and vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon.

The idea of the proof is quite straightforward: by the Stable Degeneration Theorem 2.6 and the Boundedness Theorem 1.1, we know that klt singularities with Coeff(Δ)I{\rm Coeff}(\Delta)\subseteq I and vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon degenerate into a bounded family of log Fano cone singularities (X0,Δ0;ξ)(X_{0},\Delta_{0};\xi). Then it suffices to show that there is a uniform choice of Kollár components over this family that deforms through the stable degeneration to the original singularities. We achieve this by analyzing the syzygies of the singularities, and show that there is a uniform rational perturbation ξ0\xi_{0} of ξ\xi such that (X,Δ)(X,\Delta) also degenerates to (X0,Δ0;ξ0)(X_{0},\Delta_{0};\xi_{0}), yielded by a Kollár component EE over (X,Δ)(X,\Delta) by [XZ-SDC, Theorem 4.1]. We conclude by the fact that the log discrepancy of EE is the same as the one of the Kollár component induced by (X0,Δ0;ξ0)(X_{0},\Delta_{0};\xi_{0}).

4.1. Filtered resolution

Before we prove Theorem 4.3, we need to recall a few results on syzygies of filtered algebras. These results should be well known but we haven’t found a good reference. Throughout this subsection, let (S,𝔪)(S,\mathfrak{m}) be a complete local Noetherian 𝕜\mathbbm{k}-algebra endowed with an 𝔪\mathfrak{m}-adic filtration111All filtrations in this section are exhaustive, decreasing, left-continuous, multiplicative and indexed by \mathbb{R}. Here we say a filtration \mathcal{F} on MM is exhaustive if λλM=0\cap_{\lambda}\mathcal{F}^{\lambda}M=0 and λλM=M\cup_{\lambda}\mathcal{F}^{\lambda}M=M. \mathcal{F}, i.e. there exists some positive constants c0,c1c_{0},c_{1} such that

𝔪c0λλS𝔪c1λ\mathfrak{m}^{\lceil c_{0}\lambda\rceil}\subseteq\mathcal{F}^{\lambda}S\subseteq\mathfrak{m}^{\lfloor c_{1}\lambda\rfloor}

for all λ0\lambda\geq 0. This implies that SS is also complete with respect to the filtration \mathcal{F}. Assume that the associated graded ring grS\mathrm{gr}_{\mathcal{F}}S is finitely generated.

Lemma 4.4.

Consider an SS-module MM with a compatible filtration also denoted by \mathcal{F}. Then the following conditions are equivalent:

  1. (1)

    grM\mathrm{gr}_{\mathcal{F}}M is finitely generated over grS\mathrm{gr}_{\mathcal{F}}S.

  2. (2)

    MM is finitely generated, and there exists some uiμiMu_{i}\in\mathcal{F}^{\mu_{i}}M (i=1,,Ni=1,\dots,N) such that λM=i=1NλμiSui\mathcal{F}^{\lambda}M=\sum_{i=1}^{N}\mathcal{F}^{\lambda-\mu_{i}}S\cdot u_{i} for all λ\lambda\in\mathbb{R}.

Proof.

Clearly (2) implies (1). For the other direction, note that (1) implies the existence of some μ\mu\in\mathbb{R} and some uiμiMu_{i}\in\mathcal{F}^{\mu_{i}}M (i=1,,Ni=1,\dots,N) such that μM=M\mathcal{F}^{\mu}M=M and that

λM=>λM+i=1NλμiSui\mathcal{F}^{\lambda}M=\mathcal{F}^{>\lambda}M+\sum_{i=1}^{N}\mathcal{F}^{\lambda-\mu_{i}}S\cdot u_{i}

for all λ\lambda\in\mathbb{R}. By the finite generation of grS\mathrm{gr}_{\mathcal{F}}S and grM\mathrm{gr}_{\mathcal{F}}M, we also know that the set {λ|grλM0}\{\lambda\,|\,\mathrm{gr}^{\lambda}_{\mathcal{F}}M\neq 0\} is discrete. Thus as SS is complete, (2) follows from a repeated application of the above equality. ∎

We say that the filtration \mathcal{F} on MM is good if any of the equivalent conditions in the above lemma holds. In particular, if \mathcal{F} is a good filtration then there exists some μ0\mu\gg 0 such that

λ+μM=λSμM\mathcal{F}^{\lambda+\mu}M=\mathcal{F}^{\lambda}S\cdot\mathcal{F}^{\mu}M

for all λ0\lambda\geq 0. Moreover, MM is complete with respect to the good filtration \mathcal{F}; in other words, if ukλkMu_{k}\in\mathcal{F}^{\lambda_{k}}M (k=1,2,k=1,2,\dots) is an infinite sequence such that λk+\lambda_{k}\to+\infty, then the formal series k=1uk\sum_{k=1}^{\infty}u_{k} converges in MM (this follows from the equivalent condition (2) and the completeness of SS).

Lemma 4.5.

Let MψMφM′′M^{\prime}\stackrel{{\scriptstyle\psi}}{{\to}}M\stackrel{{\scriptstyle\varphi}}{{\to}}M^{\prime\prime} be a complex of filtered SS-modules. Assume that the filtration on MM is good and the induced complex

(4.1) gr(M)\textstyle{\mathrm{gr}(M^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ¯\scriptstyle{\bar{\psi}}gr(M)\textstyle{\mathrm{gr}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ¯\scriptstyle{\bar{\varphi}}gr(M′′)\textstyle{\mathrm{gr}(M^{\prime\prime})}

is exact. Then MψMφM′′M^{\prime}\stackrel{{\scriptstyle\psi}}{{\to}}M\stackrel{{\scriptstyle\varphi}}{{\to}}M^{\prime\prime} is filtered exact, i.e.,

(4.2) λM\textstyle{\mathcal{F}^{\lambda}M^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}λM\textstyle{\mathcal{F}^{\lambda}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}λM′′\textstyle{\mathcal{F}^{\lambda}M^{\prime\prime}}

is exact for all λ\lambda\in\mathbb{R}.

Proof.

Since gr(M)\mathrm{gr}(M) is finitely generated over grS\mathrm{gr}_{\mathcal{F}}S by assumption, we may replace MM^{\prime} by a submodule so that (4.1) is still exact and gr(M)\mathrm{gr}(M^{\prime}) is finitely generated. If (4.2) is filtered exact for this submodule, then the same is true for MM^{\prime}. Thus it suffices to prove the lemma when the filtration on MM^{\prime} is also good.

Exactness of (4.1) implies that

ker(φ)λMψ(λM)+ker(φ)>λM\ker(\varphi)\cap\mathcal{F}^{\lambda}M\subseteq\psi(\mathcal{F}^{\lambda}M^{\prime})+\ker(\varphi)\cap\mathcal{F}^{>\lambda}M

for any λ\lambda\in\mathbb{R}. By induction on the filtration index λ\lambda (which belongs to a discrete set) this in turn implies that for any uker(φ)λMu\in\ker(\varphi)\cap\mathcal{F}^{\lambda}M there exists an infinite sequence ukλkMu_{k}\in\mathcal{F}^{\lambda_{k}}M^{\prime} with λλ1<λ2<\lambda\leq\lambda_{1}<\lambda_{2}<\dots and λk\lambda_{k}\to\infty such that uψ(u)μMu-\psi(u^{\prime})\in\mathcal{F}^{\mu}M for any μ\mu\in\mathbb{R}, where u=kukλMu^{\prime}=\sum_{k}u_{k}\in\mathcal{F}^{\lambda}M^{\prime} (the series converges since the filtration on MM^{\prime} is good). Since the filtration on MM is assumed to be exhaustive and in particular μμM=0\cap_{\mu}\mathcal{F}^{\mu}M=0, we deduce that u=ψ(u)u=\psi(u^{\prime}) and hence (4.2) is filtered exact. ∎

The following is the main result of this subsection. We denote by S(μ)S(\mu) the filtered free SS-module that is isomorphic to SS as an SS-module but with a filtration given by λS(μ)=λ+μS\mathcal{F}^{\lambda}S(\mu)=\mathcal{F}^{\lambda+\mu}S.

Lemma 4.6.

Let MM be an SS-module with a compatible good filtration and let

(4.3) \textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ¯2\scriptstyle{\bar{\varphi}_{2}}M¯2\textstyle{\overline{M}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ¯1\scriptstyle{\bar{\varphi}_{1}}M¯1\textstyle{\overline{M}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ¯0\scriptstyle{\bar{\varphi}_{0}}gr(M)\textstyle{\mathrm{gr}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

be a graded free resolution of gr(M)\mathrm{gr}(M), i.e. for each ii there exists some λij\lambda_{ij}\in\mathbb{R} such that M¯ijgr(S)(λij)\overline{M}_{i}\cong\oplus_{j}\mathrm{gr}_{\mathcal{F}}(S)(\lambda_{ij}). Then it can be lifted to a filtered exact free resolution

\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ2\scriptstyle{\varphi_{2}}M2\textstyle{M_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ1\scriptstyle{\varphi_{1}}M1\textstyle{M_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ0\scriptstyle{\varphi_{0}}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

of MM, i.e. the associated graded complex is (4.3).

Proof.

First set M1=jS(λ1j)M_{1}=\oplus_{j}S(\lambda_{1j}) and lift φ¯0\bar{\varphi}_{0} to a map φ0:M1M\varphi_{0}\colon M_{1}\to M. By Lemma 4.5, the exactness of M¯1gr(M)0\overline{M}_{1}\to\mathrm{gr}(M)\to 0 implies the filtered exactness of M1M0M_{1}\to M\to 0; in particular, φ0\varphi_{0} is surjective. Let M=ker(φ0)M^{\prime}=\ker(\varphi_{0}) and equip it with the filtration induced from M1M_{1}. We claim that gr(M)=ker(φ0¯)\mathrm{gr}(M^{\prime})=\ker(\bar{\varphi_{0}}). Indeed, gr(M)\mathrm{gr}(M^{\prime}) injects into gr(M1)=M¯1\mathrm{gr}(M_{1})=\overline{M}_{1} as a general property of induced filtrations on submodules. Clearly we also have gr(M)ker(φ0¯)\mathrm{gr}(M^{\prime})\subseteq\ker(\bar{\varphi_{0}}). On the other hand, any homogeneous element u¯ker(φ0¯)\bar{u}\in\ker(\bar{\varphi_{0}}) is represented by some uλM1u\in\mathcal{F}^{\lambda}M_{1} such that φ0(u)>λM\varphi_{0}(u)\in\mathcal{F}^{>\lambda}M. As M1M0M_{1}\to M\to 0 is filtered exact, this implies that φ0(u)=φ0(u1)\varphi_{0}(u)=\varphi_{0}(u_{1}) for some u1>λM1u_{1}\in\mathcal{F}^{>\lambda}M_{1}. But then u=uu1ker(φ0)=Mu^{\prime}=u-u_{1}\in\ker(\varphi_{0})=M^{\prime}; moreover, we have uλMu^{\prime}\in\mathcal{F}^{\lambda}M^{\prime} and it is also a lift of u¯\bar{u}. It follows that ker(φ0¯)gr(M)\ker(\bar{\varphi_{0}})\subseteq\mathrm{gr}(M^{\prime}). This proves the claim. From here we deduce that M¯2gr(M)\cdots\to\overline{M}_{2}\to\mathrm{gr}(M^{\prime}) is also a graded free resolution. By construction, the sequence 0MM1M00\to M^{\prime}\to M_{1}\to M\to 0 is also filtered exact. Thus we may repeat the same argument above with MM^{\prime} in place of MM and the lemma follows. ∎

4.2. Lifting valuations

Next consider a singularity xX=Spec(R)x\in X=\mathrm{Spec}(R) and a quasi-monomial valuation vValX,xv\in\mathrm{Val}_{X,x} such that R0:=grvRR_{0}:=\mathrm{gr}_{v}R is finitely generated. As in [XZ-SDC, paragraph after Theorem 4.1], every \mathbb{R}-divisor Δ\Delta has an induced degeneration Δ0\Delta_{0} on X0:=Spec(R0)X_{0}:=\mathrm{Spec}(R_{0}). The natural grading on R0R_{0} induces a Reeb vector field ξv\xi_{v} on X0X_{0} generating a torus 𝕋=ξv\mathbb{T}=\langle\xi_{v}\rangle. In this setup, using terminologies from K-stability, we say that x(X,Δ)x\in(X,\Delta) degenerates into (X0,Δ0;ξ0)(X_{0},\Delta_{0};\xi_{0}) via the \mathbb{R}-test configuration induced by vv. For the application to Theorem 4.3, x(X,Δ)x\in(X,\Delta) will be a klt singularity whose local volume is bounded from below, and (X0,Δ0;ξ0)(X_{0},\Delta_{0};\xi_{0}) is its K-semistable log Fano cone degeneration. Denote by

Φξv(R0):=wtξv(R0{0})=v(R{0})0\Phi_{\xi_{v}}(R_{0}):=\mathrm{wt}_{\xi_{v}}(R_{0}\setminus\{0\})=v(R\setminus\{0\})\subseteq\mathbb{R}_{\geq 0}

the value semigroup which is a discrete set that only depends on the graded ring R0R_{0} and the Reeb vector ξv\xi_{v}. We shall identify the Reeb cone 𝔱+\mathfrak{t}^{+}_{\mathbb{R}} with the set of semigroup homomorphisms Φξv(R0){0}+\Phi_{\xi_{v}}(R_{0})\setminus\{0\}\to\mathbb{R}_{+} (in particular, ξv\xi_{v} is identified with the natural inclusion Φξv(R0)0\Phi_{\xi_{v}}(R_{0})\subseteq\mathbb{R}_{\geq 0}). The question we need to address is which (rational) perturbation of ξv\xi_{v} inside the Reeb cone can be lifted to a (divisorial) valuation on XX, as such a lift will help us control the mldK\mathrm{mld}^{\mathrm{K}} of the singularity xXx\in X.

For any finite set {u¯1,,u¯N}\{\bar{u}_{1},\dots,\bar{u}_{N}\} of homogeneous generators of R0R_{0}, we have a graded surjection

(4.4) φ¯0:S:=𝕜[x1,,xN]R0\bar{\varphi}_{0}\colon S:=\mathbbm{k}[x_{1},\dots,x_{N}]\to R_{0}

sending xix_{i} to u¯i\bar{u}_{i}. Here the grading of SS is defined by setting deg(xi)=deg(u¯i)\deg(x_{i})=\deg(\bar{u}_{i}). This also induces a filtration on S^:=𝕜[[x1,,xN]]\hat{S}:=\mathbbm{k}[\![x_{1},\dots,x_{N}]\!] such that gr(S^)S\mathrm{gr}(\hat{S})\cong S.

Definition 4.7.

Let c>0c>0 be a positive number. We say that the syzygy complexity of the graded algebra R0R_{0} is at most cc, if there exist homogeneous generators u¯1,,u¯N\bar{u}_{1},\dots,\bar{u}_{N} of R0R_{0} (for some NcN\leq c) such that deg(u¯i)c\deg(\bar{u}_{i})\leq c for all ii, and for the induced map φ¯0\bar{\varphi}_{0} in (4.4), there exists a graded free resolution of R0R_{0} (viewed as a graded SS-module)

(4.5) M¯2\textstyle{\overline{M}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ¯2\scriptstyle{\bar{\varphi}_{2}}M¯1\textstyle{\overline{M}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ¯1\scriptstyle{\bar{\varphi}_{1}}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ¯0\scriptstyle{\bar{\varphi}_{0}}R0\textstyle{R_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

such that all the matrix entries of φ¯1\bar{\varphi}_{1} and φ¯2\bar{\varphi}_{2} have degrees at most cc.

The technical result of this subsection is that the liftability of the perturbed Reeb vector only depends on the syzygy complexity of R0R_{0} and the value semigroup Φξv(R0)\Phi_{\xi_{v}}(R_{0}). To prepare for such a statement, fix some c>0c>0 such that the syzygy complexity of R0R_{0} is at most cc, and choose some generators u¯i\bar{u}_{i} and a graded free resolution of R0R_{0} as in (4.5) that satisfy the conditions in Definition 4.7. Let λi:=deg(u¯i)\lambda_{i}:=\deg(\bar{u}_{i}). By Lemma 4.6, we can lift (4.5) to a filtered free resolution

(4.6) M2\textstyle{M_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ2\scriptstyle{\varphi_{2}}M1\textstyle{M_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ1\scriptstyle{\varphi_{1}}S^\textstyle{\hat{S}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ0\scriptstyle{\varphi_{0}}R^\textstyle{\hat{R}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

of the S^\hat{S}-module R^\hat{R} (the completion of RR). In concrete terms, this means that the matrix entries of the homomorphism φ¯i\bar{\varphi}_{i} are given by the initial terms of the matrix entries of φi\varphi_{i}. For example, to lift φ¯0\bar{\varphi}_{0} to φ0\varphi_{0}, we choose a lift uiRu_{i}\in R of u¯i\bar{u}_{i} for each ii, namely, ui𝔞λi(v)u_{i}\in\mathfrak{a}_{\lambda_{i}}(v) and its reduction in 𝔞λi(v)/𝔞>λi(v)\mathfrak{a}_{\lambda_{i}}(v)/\mathfrak{a}_{>\lambda_{i}}(v) is u¯i\bar{u}_{i}. We then lift φ¯0\bar{\varphi}_{0} to a map

φ0:𝕜[x1,,xN]R\varphi_{0}\colon\mathbbm{k}[x_{1},\dots,x_{N}]\to R

sending xix_{i} to uiu_{i}, which is surjective after passing to the completion.

Each Reeb vector ξ𝔱+\xi\in\mathfrak{t}^{+}_{\mathbb{R}} induces a valuation of R0R_{0}, and there is a natural candidate for its lift to a valuation of RR. Indeed, since we identify Reeb vectors with homomorphism Φξv(R0){0}+\Phi_{\xi_{v}}(R_{0})\setminus\{0\}\to\mathbb{R}_{+}, there is a natural map

ι:𝔱++N,ξ(ξ(λ1),,ξ(λN)),\iota\colon\mathfrak{t}^{+}_{\mathbb{R}}\to\mathbb{R}^{N}_{+},\quad\xi\mapsto(\xi(\lambda_{1}),\dots,\xi(\lambda_{N})),

whose image ι(𝔱+)\iota(\mathfrak{t}_{\mathbb{R}}^{+}) is an open subset of the rational envelope ΛN\Lambda\subseteq\mathbb{R}^{N} of λ=ι(ξv)=(λ1,,λN)\lambda=\iota(\xi_{v})=(\lambda_{1},\dots,\lambda_{N}) (i.e. the smallest rational affine linear subspaces in N\mathbb{R}^{N} containing λ\lambda). For any μ+N\mu\in\mathbb{R}^{N}_{+} we get a monomial valuation wtμ\mathrm{wt}_{\mu} on S^\hat{S} which induces a filtration μ\mathcal{F}_{\mu} on S^\hat{S}. Through the surjection φ0:S^R^\varphi_{0}\colon\hat{S}\to\hat{R}, it also induces a filtration (also denoted by μ\mathcal{F}_{\mu}) on R^\hat{R}. For any uR^{0}u\in\hat{R}\setminus\{0\} set

(4.7) vμ(u):=max{t|uμtR^};v_{\mu}(u):=\max\{t\in\mathbb{R}\,|\,u\in\mathcal{F}_{\mu}^{t}\hat{R}\};

when μ=ι(ξ)\mu=\iota(\xi) this is our candidate lift of ξ\xi. Note that vμv_{\mu} need not be a valuation, and our next task is find conditions that make it into a valuation.

For this we need an auxiliary result. For any μ+N\mu\in\mathbb{R}^{N}_{+} and any uS^u\in\hat{S}, we denote by inμ(u)\mathrm{in}_{\mu}(u) the initial term of uu with respect to wtμ\mathrm{wt}_{\mu}.

Lemma 4.8.

For any uS^u\in\hat{S}, there exists an open neighbourhood U𝔱+U\subseteq\mathfrak{t}^{+}_{\mathbb{R}} of ξv\xi_{v} depending only on λ\lambda and wtλ(u)\mathrm{wt}_{\lambda}(u) such that inμ(u)=inλ(u)\mathrm{in}_{\mu}(u)=\mathrm{in}_{\lambda}(u) for all μι(U)\mu\in\iota(U).

Proof.

Every monomial in S^\hat{S} is of the form x1a1xNaNx_{1}^{a_{1}}\cdots x_{N}^{a_{N}} for some (a1,,aN)N(a_{1},\dots,a_{N})\in\mathbb{N}^{N}. Observe that (both facts are elementary to verify):

  1. (1)

    if a,bNa,b\in\mathbb{N}^{N} and λ,a=λ,b\langle\lambda,a\rangle=\langle\lambda,b\rangle, then for any ξ𝔱+\xi\in\mathfrak{t}^{+}_{\mathbb{R}} it also holds that ι(ξ),a=ι(ξ),b\langle\iota(\xi),a\rangle=\langle\iota(\xi),b\rangle;

  2. (2)

    if a,bNa,b\in\mathbb{N}^{N} is such that λ,a<λ,b\langle\lambda,a\rangle<\langle\lambda,b\rangle, then there exists an open neighbourhood UNU\subseteq\mathbb{R}^{N} of λ\lambda depending only on aa and λ\lambda such that μ,a<μ,b\langle\mu,a\rangle<\langle\mu,b\rangle for all μU\mu\in U.

Note that there are only finitely many aNa\in\mathbb{N}^{N} such that λ,a=wtλ(u)\langle\lambda,a\rangle=\mathrm{wt}_{\lambda}(u). The lemma then follows by translating these facts into statements about weights of monomials in S^\hat{S}. ∎

By Lemma 4.8, we can choose a small neighbourhood U𝔱+U\subseteq\mathfrak{t}_{\mathbb{R}}^{+} of ξv\xi_{v}, depending only on λ\lambda and the weights of the matrix entries of φ1\varphi_{1} and φ2\varphi_{2} (with respect to λ\lambda), such that inμ(φi)=inλ(φi)\mathrm{in}_{\mu}(\varphi_{i})=\mathrm{in}_{\lambda}(\varphi_{i}) for all i=1,2i=1,2 and all μι(U)+N\mu\in\iota(U)\subseteq\mathbb{R}^{N}_{+}. In particular, the first three terms M¯2M¯1S\overline{M}_{2}\to\overline{M}_{1}\to S in (4.5) is the reduction of M2M1S^M_{2}\to M_{1}\to\hat{S} with respect to wtμ\mathrm{wt}_{\mu} for any μι(U)\mu\in\iota(U). As there are only finitely many values in the semigroup Φξv(R0)\Phi_{\xi_{v}}(R_{0}) that are bounded from above, we see that in fact UU only depends on Φξv(R0)\Phi_{\xi_{v}}(R_{0}) together with the syzygy complexity of R0R_{0} (that is, if R0R_{0} has syzygy complexity at most cc, then UU can be chosen in terms of Φξv(R0)\Phi_{\xi_{v}}(R_{0}) and cc). We now show that every ξU\xi\in U lifts to a valuation of RR through (4.7).

Lemma 4.9.

For any ξU\xi\in U, there exists a quasi-monomial valuation vξValX,xv_{\xi}\in\mathrm{Val}_{X,x} such that grvξRR0\mathrm{gr}_{v_{\xi}}R\cong R_{0} and its induced Reeb vector on X0X_{0} is ξ\xi.

Proof.

Let μ=ι(ξ)\mu=\iota(\xi). Recall that we have a filtration μ\mathcal{F}_{\mu} of R^\hat{R} induced by the valuation wtμ\mathrm{wt}_{\mu} on S^\hat{S}. Similarly, we have an induced filtration (still denoted by μ\mathcal{F}_{\mu}) on I^:=ker(φ0)S^\hat{I}:=\ker(\varphi_{0})\subseteq\hat{S}. Note that I^=φ1(M1)\hat{I}=\varphi_{1}(M_{1}) since (4.6) is exact. As M1M_{1}, M2M_{2} are free S^\hat{S}-modules, there are also natural filtrations (again denoted μ\mathcal{F}_{\mu}) on them so that (4.6) is compatible with these new filtrations (on each free summand of MiM_{i} the filtration is simply given by an appropriate degree shift of the filtration μ\mathcal{F}_{\mu} of S^\hat{S}).

As in (4.7), for M=MiM=M_{i} or S^\hat{S} and any uM{0}u\in M\setminus\{0\}, we set vμ(u):=max{t|uμtM}v_{\mu}(u):=\max\{t\in\mathbb{R}\,|\,u\in\mathcal{F}_{\mu}^{t}M\}. For ease of notation, we shall write grμ(R^)\mathrm{gr}_{\mu}(\hat{R}) instead of grμ(R^)\mathrm{gr}_{\mathcal{F}_{\mu}}(\hat{R}) etc. We claim that

(4.8) grμR^R0.\mathrm{gr}_{\mu}\hat{R}\cong R_{0}.

Note that the kernel of the induced map Sgrμ(S^)grμ(R^)S\cong\mathrm{gr}_{\mu}(\hat{S})\to\mathrm{gr}_{\mu}(\hat{R}) is grμ(I^)\mathrm{gr}_{\mu}(\hat{I}). Thus in view of the exact sequence (4.5), to prove the claim it suffices to show that M¯1grμ(M1)\overline{M}_{1}\cong\mathrm{gr}_{\mu}(M_{1}) surjects onto grμ(I^)\mathrm{gr}_{\mu}(\hat{I}). Equivalently, we need to show that the map φ1:M1S^\varphi_{1}\colon M_{1}\to\hat{S} is strict with respect to the filtration μ\mathcal{F}_{\mu}, namely, φ1(μM1)=φ1(M1)μS^=μI^\varphi_{1}(\mathcal{F}_{\mu}M_{1})=\varphi_{1}(M_{1})\cap\mathcal{F}_{\mu}\hat{S}=\mathcal{F}_{\mu}\hat{I}.

To this end, let uI^u\in\hat{I}; among its preimages in M1M_{1}, choose one (denoted by rr) that lies in the smallest filtered subspaces μtM1\mathcal{F}_{\mu}^{t}M_{1}, namely, let rφ11(u)r\in\varphi^{-1}_{1}(u) be such that vμ(r)vμ(r)v_{\mu}(r^{\prime})\leq v_{\mu}(r) for all rφ11(u)r^{\prime}\in\varphi^{-1}_{1}(u). This is possible since by our choice of the filtration μ\mathcal{F}_{\mu} on the free S^\hat{S}-module M1M_{1}, the set vμ(M1{0})v_{\mu}(M_{1}\setminus\{0\}) of jumping numbers is a discrete set that is bounded from below, and for any rφ11(u)r\in\varphi^{-1}_{1}(u) we have vμ(u)vμ(r)v_{\mu}(u)\geq v_{\mu}(r) as φ1\varphi_{1} is compatible with the filtration μ\mathcal{F}_{\mu}, thus vμ(φ11(u))v_{\mu}(\varphi^{-1}_{1}(u)) is also bounded from above. Let us prove that vμ(u)=vμ(r)v_{\mu}(u)=v_{\mu}(r). Suppose not, then vμ(u)>vμ(r)v_{\mu}(u)>v_{\mu}(r), hence the reduction r¯M¯1\bar{r}\in\overline{M}_{1} of rM1r\in M_{1} is in the kernel of φ¯1\bar{\varphi}_{1}, hence by the exactness of (4.5) we see that r¯=φ¯2(s¯)\bar{r}=\bar{\varphi}_{2}(\bar{s}) for some s¯M¯2\bar{s}\in\overline{M}_{2}. Choose some sM2s\in M_{2} whose reduction is s¯\bar{s}. Then we have vμ(rφ2(s))>vμ(r)v_{\mu}(r-\varphi_{2}(s))>v_{\mu}(r). But since (4.6) is exact, we also have rφ2(s)φ11(u)r-\varphi_{2}(s)\in\varphi^{-1}_{1}(u). This contradicts our initial choice of rr. Thus the equality vμ(u)=vμ(r)v_{\mu}(u)=v_{\mu}(r) holds and we conclude that φ1(μM1)=μI^\varphi_{1}(\mathcal{F}_{\mu}M_{1})=\mathcal{F}_{\mu}\hat{I}. As discussed at the beginning of the proof, this proves the claim (4.8).

The lemma is now a formal consequence of (4.8). Since R0R_{0} is an integral domain, we deduce that vμv_{\mu} is a valuation (see e.g. [Z-SDC-survey, Lemma 2.4]; it is quasi-monomial since R0R_{0} is finitely generated (see e.g. [ELS03, Discussion 3.16]). By construction, we have vμ(u)>0v_{\mu}(u)>0 for all u𝔪R^u\in\mathfrak{m}\subseteq\hat{R}. Hence the restriction of vμv_{\mu} to RR gives the desired valuation. ∎

Specializing to the stable degeneration setting, we obtain the following.

Corollary 4.10.

Let c1,c2>0c_{1},c_{2}>0 be two positive constants, let mm be a positive integer and let x0(X0=Spec(R0),Δ0;ξ0)x_{0}\in(X_{0}=\mathrm{Spec}(R_{0}),\Delta_{0};\xi_{0}) be a log Fano cone singularity. Assume that mΔ0m\Delta_{0} is Cartier, R0R_{0} has syzygy complexity at most c1c_{1}, and wtξ0(mΔ)c2\mathrm{wt}_{\xi_{0}}(m\Delta)\leq c_{2}. Then there exists a rational perturbation ξ1𝔱+\xi_{1}\in\mathfrak{t}^{+}_{\mathbb{R}} of the Reeb vector ξ0\xi_{0} depending only on c1,c2,mc_{1},c_{2},m and the value semigroup Φξ0(R0)\Phi_{\xi_{0}}(R_{0}) such that for any klt singularity x(X=Spec(R),Δ)x\in(X=\mathrm{Spec}(R),\Delta) which degenerates into x0(X0,Δ0;ξ0)x_{0}\in(X_{0},\Delta_{0};\xi_{0}) through some \mathbb{R}-test configuration and such that mΔm\Delta is a \mathbb{Q}-Cartier Weil divisor, there exists a divisorial valuation v1ValX,xv_{1}\in\mathrm{Val}_{X,x} that induces a degeneration of x(X,Δ)x\in(X,\Delta) to the log Fano cone x0(X0,Δ0;ξ1)x_{0}\in(X_{0},\Delta_{0};\xi_{1}).

Proof.

By the above construction and Lemma 4.9, we see that there exists an open neighbourhood U𝔱+U\subseteq\mathfrak{t}_{\mathbb{R}}^{+} of ξ0\xi_{0}, depending only on c1c_{1} and Φξ0(R0)\Phi_{\xi_{0}}(R_{0}), such that for any ξ1U\xi_{1}\in U there exists a quasi-monomial valuation v1v_{1} of RR whose associated graded ring grv1R\mathrm{gr}_{v_{1}}R is isomorphic to R0R_{0}, and the induced Reeb vector on X0=Spec(R0)X_{0}=\mathrm{Spec}(R_{0}) is exactly ξ1\xi_{1}. If ξ1\xi_{1} is rational, then v1v_{1} has rational rank one and hence is divisorial. Since mΔ0m\Delta_{0} is Cartier, the same holds for mΔm\Delta at xx. As wtξ0(mΔ)\mathrm{wt}_{\xi_{0}}(m\Delta) belongs to the bounded discrete set Φξ0(R0)[0,c2]\Phi_{\xi_{0}}(R_{0})\cap[0,c_{2}], there are only finitely possible values of wtξ0(mΔ)\mathrm{wt}_{\xi_{0}}(m\Delta). By Lemma 4.8, this implies that after possibly shrinking UU (in terms of c1,c2,mc_{1},c_{2},m and Φξ0(R0)\Phi_{\xi_{0}}(R_{0})), we can further assume that the Cartier divisor mΔm\Delta still specializes to mΔ0m\Delta_{0} under the degeneration induced by v1v_{1}. The desired statement then holds for any rational Reeb vector ξ1U\xi_{1}\in U. ∎

4.3. Mld of Kollár components

We now put the construction from the previous subsection in families to prove Theorem 4.3. We need another auxiliary lemma.

Lemma 4.11.

Let x(X=Spec(R),Δ)x\in(X=\mathrm{Spec}(R),\Delta) be a klt singularity such that KXK_{X} is \mathbb{Q}-Cartier. Then the same is true for its K-semistable log Fano cone degeneration.

Proof.

This is a direct consequence of the stable degeneration theory. Let vValX,xv\in\mathrm{Val}_{X,x} be the minimizer of the normalized volume function. By [XZ-SDC, Lemma 3.4], there exists a log smooth model π:(Y,E)(X,Δ)\pi\colon(Y,E)\to(X,\Delta) such that vv is a monomial lc place of some special \mathbb{Q}-complement Γ\Gamma with respect to (Y,E)(Y,E) (we refer to Definition 3.1 of loc. cit. for the relevant definitions). Since KXK_{X} is \mathbb{Q}-Cartier, Γ\Gamma is also a special \mathbb{Q}-complement of the klt singularity xXx\in X. Since the underlying singularity of the K-semistable log Fano cone is X0=Spec(grvR)X_{0}=\mathrm{Spec}(\mathrm{gr}_{v}R), we deduce from [XZ-SDC, Theorem 4.1] that it is also \mathbb{Q}-Gorenstein (and in fact klt). ∎

Proof of Theorem 4.3.

By [Z-mld^K-2, Lemma 2.18], it suffices to prove the theorem when II\subseteq\mathbb{Q} and thus we may assume that I=1m0[0,1]I=\frac{1}{m_{0}}\mathbb{N}\cap[0,1] for some positive integer m0m_{0}. By [Z-mld^K-1, Proposition 5.8], we also know that it is enough to consider singularities x(X,Δ)x\in(X,\Delta) for which KXK_{X} is \mathbb{Q}-Cartier. These in particular imply that m0Δm_{0}\Delta is a \mathbb{Q}-Cartier Weil divisor.

There exists a bounded set 𝒮\mathcal{S} of \mathbb{Q}-Gorenstein log Fano cone singularities with coefficients in I=1m0I=\frac{1}{m_{0}}\mathbb{N} and local volume at least ε\varepsilon. This follows from Theorem 1.1 and the fact that in any bounded family B𝒳BB\subseteq\mathcal{X}\to B of singularities the locus of bBb\in B for which the fiber b𝒳bb\in\mathcal{X}_{b} is \mathbb{Q}-Gorenstein klt is constructible (cf. the proof of Lemma 2.17). By Lemma 4.11, the K-semistable log Fano cone degeneration of any nn-dimensional \mathbb{Q}-Gorenstein klt singularity x(X,Δ)x\in(X,\Delta) with Coeff(Δ)I{\rm Coeff}(\Delta)\subseteq I and vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon belongs to 𝒮\mathcal{S}.

Without loss of generality, we shall assume that 𝒮\mathcal{S} consists of just one flat family B(𝒳=Spec(),𝒟)BB\subseteq(\mathcal{X}=\mathrm{Spec}(\mathcal{R}),\mathcal{D})\to B over a smooth base where K𝒳K_{\mathcal{X}} is \mathbb{Q}-Cartier (in practice this means we freely stratify the base BB and analyze one deformation family at a time). Let 𝕋\mathbb{T} be the torus that acts fiberwise on the family. Note that the Reeb cone and the volume function on it is independent of bBb\in B. This is because if =α\mathcal{R}=\oplus\mathcal{R}_{\alpha} is the weight decomposition, then each α\mathcal{R}_{\alpha} is free over 𝒪B\mathcal{O}_{B} by the flatness of 𝒳B\mathcal{X}\to B, hence their ranks are constant in bBb\in B. After possibly stratifying the base BB, we may assume (by e.g. [LX-stability-higher-rank, Theorem 2.15(3)]) that the log discrepancy function A𝒳b,𝒟bA_{\mathcal{X}_{b},\mathcal{D}_{b}} on the Reeb cone is independent of bBb\in B as well. This implies that there exists some ξN(𝕋)\xi\in N(\mathbb{T})_{\mathbb{R}}, independent of bBb\in B, that minimizes the normalized volume function on the Reeb cone. In particular, if b(𝒳b,𝒟b)b\in(\mathcal{X}_{b},\mathcal{D}_{b}) is a K-semistable log Fano cone, then (up to rescaling) ξ\xi is the associated Reeb vector. The value semigroup Φ:=Φξ(b)\Phi:=\Phi_{\xi}(\mathcal{R}_{b}) is thus also constant in bb.

Since 𝒮\mathcal{S} is bounded, we may choose some c>0c>0 such that for all bBb\in B, the syzygy complexity of b\mathcal{R}_{b} is at most cc. By construction, the Weil divisor m0𝒟m_{0}\mathcal{D} is \mathbb{Q}-Cartier (since both K𝒳+𝒟K_{\mathcal{X}}+\mathcal{D} and K𝒳K_{\mathcal{X}} are), thus by [XZ-minimizer-unique, Corollary 1.4] we see that there exists some positive integer m1m_{1} depending only on ε\varepsilon such that m𝒟bm\mathcal{D}_{b} is Cartier for all bb, where m=m0m1m=m_{0}m_{1}. Again since 𝒮\mathcal{S} is bounded, we may further assume that wtξ(m𝒟b)c\mathrm{wt}_{\xi}(m\mathcal{D}_{b})\leq c for all bBb\in B after possibly enlarging cc.

By Corollary 4.10, we deduce that there exists a (rational) Reeb vector ξ1\xi_{1}, independent of bBb\in B, such that for any klt singularity x(X,Δ)x\in(X,\Delta) whose K-semistable log Fano cone degeneration (X0,Δ0;ξ0)(X_{0},\Delta_{0};\xi_{0}) belongs to 𝒮\mathcal{S}, there exists a divisorial valuation v1ValX,xv_{1}\in\mathrm{Val}_{X,x} that induces a special degeneration of x(X,Δ)x\in(X,\Delta) to the log Fano cone (X0,Δ0;ξ1)(X_{0},\Delta_{0};\xi_{1}). Replacing ξ1\xi_{1} by a fixed multiple, we may also assume that its value group equals \mathbb{Z}. In particular, we have v1=ordEv_{1}=\mathrm{ord}_{E} for some divisor EE over x(X,Δ)x\in(X,\Delta). Since its induced degeneration (X0,Δ0)(X_{0},\Delta_{0}) is klt, by [XZ-SDC, Theorem 4.1] we see that EE is necessarily a Kollár component over x(X,Δ)x\in(X,\Delta). On the other hand, by [LX-stability-higher-rank, Lemma 2.58] we also have AX,Δ(E)=AX0,Δ0(wtξ1)A_{X,\Delta}(E)=A_{X_{0},\Delta_{0}}(\mathrm{wt}_{\xi_{1}}), and the right hand side is bounded from above since (X0,Δ0)(X_{0},\Delta_{0}) belongs to a bounded family and the Reeb vector ξ1\xi_{1} is constant in this family. Hence we conclude that mldK(x,X,Δ)\mathrm{mld}^{K}(x,X,\Delta) is bounded from above and this finishes the proof. ∎

References