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

On boundedness of singularities and minimal log discrepancies of Kollár components, II

Ziquan Zhuang Department of Mathematics, Johns Hopkins University, Baltimore, MD 21218, USA [email protected]
Abstract.

We show that a set of K-semistable log Fano cone singularities is bounded if and only if their local volumes are bounded away from zero, and their minimal log discrepancies of Kollár components are bounded from above. As corollaries, we confirm the boundedness conjecture for K-semistable log Fano cone singularities in dimension three, and show that local volumes of 33-dimensional klt singularities only accumulate at zero.

1. Introduction

Following the recent study on K-stability of Fano varieties (see [Xu-K-stability-survey] for a comprehensive survey), there has been growing interest in establishing a parallel stability theory for klt singularities, which are the local analog of Fano varieties. In the global theory, boundedness of Fano varieties plays an important role. Building on the seminal work of Birkar [Birkar-bab-1, Birkar-bab-2], Jiang [Jia-Kss-Fano-bdd] proved that (in any fixed dimension) K-semistable Fano varieties with anti-canonical volumes bounded away from zero form a bounded family.111Several different proofs were later found in [LLX-nv-survey, XZ-minimizer-unique]. This is the first step in the general construction of the K-moduli space of Fano varieties; it is also a key ingredient in the proof [LXZ-HRFG] of the global version of the Higher Rank Finite Generation Conjecture.

To advance the local stability theory, it is therefore natural to investigate the boundedness of klt singularities. Several years ago, Li [Li-normalized-volume] introduced an interesting invariant of klt singularities called the local volume. It has become clear that the stability theory of klt singularities should be built around this invariant. In particular, it is speculated that (in any fixed dimension) klt singularities with local volumes bounded away from zero are specially bounded, i.e. they isotrivially degenerate to a bounded family. This is known in some special cases [HLQ-vol-ACC, MS-bdd-toric, MS-bdd-complexity-one, Z-mld^K-1], but the general situation is still quite mysterious.

By the recent solution of the Stable Degeneration Conjecture [Blu-minimizer-exist, LX-stability-higher-rank, Xu-quasi-monomial, XZ-minimizer-unique, LWX-metric-tangent-cone, XZ-SDC] (see also [LX-stability-kc, BLQ-convexity]), every klt singularity has a canonical “stable degeneration” to a K-semistable log Fano cone singularity (see Section 2.4 for the precise definition; roughly speaking, K-semistable log Fano cone singularities are generalizations of cones over K-semistable Fano varieties). This suggests a more precise boundedness conjecture ([XZ-SDC, Conjecture 1.7], see also [SZ-no-semistability, Problem 6.9]):

Conjecture 1.1.

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

{(X,Δ)|x(X,Δ) is a K-semistable log Fano cone singularity,dimX=n,Coef(Δ)I,and vol^(x,X,Δ)ε}\Big{\{}(X,\Delta)\ \Big{|}\begin{array}[]{l}x\in(X,\Delta)\mbox{ is a K-semistable log Fano cone singularity},\\ \dim X=n,\,{\rm Coef}(\Delta)\subseteq I,\,\mbox{and }\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon\end{array}\Big{\}}

is bounded.

Here vol^(x,X,Δ)\widehat{\rm vol}(x,X,\Delta) denotes the local volume of the singularity x(X,Δ)x\in(X,\Delta). This conjecture has been verified for toric singularities [MS-bdd-toric, Z-mld^K-1], for hypersurface singularities, and for singularities with torus actions of complexity one [MS-bdd-complexity-one].

In this paper, we study Conjecture 1.1 using the minimal log discrepancies of Kollár components (or simply mldK\mathrm{mld}^{\mathrm{K}}), see Section 2.2. Our main result gives a boundedness criterion in terms of mldK\mathrm{mld}^{\mathrm{K}} and the local volume.

Theorem 1.2 (=Corollary 3.10).

Fix nn\in\mathbb{N} and consider a set 𝒮\mathcal{S} of nn-dimensional K-semistable log Fano cone singularities with coefficients in a fixed finite set I[0,1]I\subseteq[0,1]. Then 𝒮\mathcal{S} is bounded if and only if there exist some ε,A>0\varepsilon,A>0 such that

vol^(x,X,Δ)ε𝑎𝑛𝑑mldK(x,X,Δ)A\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon\quad\mathit{and}\quad\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq A

for all x(X,Δ)x\in(X,\Delta) in 𝒮\mathcal{S}.

This upgrades the special boundedness result from our previous work [Z-mld^K-1] to actual boundedness. We also prove boundedness results for more general log Fano cone singularities, replacing the K-semistability requirement by lower bounds on stability thresholds (as introduced in [Hua-thesis]), see Corollary 3.14.

Comparing Conjecture 1.1 and Theorem 1.2, we are naturally led to the following conjecture, already raised in [Z-mld^K-1]:

Conjecture 1.3 ([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 AA 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 Coef(Δ)I\mathrm{Coef}(\Delta)\subseteq I and vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon.

This is known in dimension up to three [Z-mld^K-1]. By Theorem 1.2, we then get a complete solution of Conjecture 1.1 in dimension three (the surface case was already treated in [HLQ-vol-ACC]). The same result is also independently proved in [LMS-bdd-dim-3] using a different method.

Corollary 1.4 (=Corollaries 3.16+3.17).

Conjecture 1.1 holds in dimension 33. Moreover, for any fixed finite coefficient set I[0,1]I\subseteq[0,1], the set of possible local volumes of 33-dimensional klt singularities is discrete away from zero.

1.1. Strategy of the proof

In addition to our previous work [Z-mld^K-1], the proof of Theorem 1.2 relies on several new ingredients. For simplicity, we assume Δ=0\Delta=0, so that we are dealing with Fano cone singularities. Every such singularity xXx\in X is an orbifold cone over some Fano variety VV, so a natural idea is to prove Theorem 1.2 by showing the boundedness of the associated Fano variety VV.

There are two main reasons why this naïve approach does not directly work. First, the orbifold base is highly non-unique; in fact, for a fixed Fano cone singularity the possible orbifold bases can be unbounded. For example, the simplest Fano cone singularity 0𝔸n0\in\mathbb{A}^{n} can be realized as an orbifold cone over any weighted projective space of dimension n1n-1, but without further constraint, weighted projective spaces do not form a bounded family.

Moreover, even when there is a canonical choice of the orbifold base (e.g. when the singularity has a unique 𝔾m\mathbb{G}_{m}-action), the anti-canonical volume of VV is only a fraction of the local volume of the singularity, where the factor is related to the Weil index of VV (defined as the largest number qq such that KVqL-K_{V}\sim_{\mathbb{Q}}qL for some Weil divisor LL). In particular, the volume of VV can a priori be arbitrarily small, which needs to be ruled out if we want any boundedness of this sort.

Our solution is to turn the local boundedness question into a global one by considering the projective orbifold cone X¯\overline{X} over VV. A key observation is that by choosing the appropriate orbifold base VV, the anti-canonical volume of X¯\overline{X} approximates the local volume of xXx\in X and in particular is bounded from both above and below. This takes care of the second issue mentioned above.

Still, as we make different choices of VV, the corresponding projective orbifold cones X¯\overline{X} can be unbounded. To circumvent this issue, we prove an effective birationality result for X¯\overline{X}. More precisely, we show that regardless of the choice of VV, there is some positive integer mm depending only on vol^(x,X)\widehat{\rm vol}(x,X) and mldK(x,X)\mathrm{mld}^{\mathrm{K}}(x,X) such that |mKX¯||-mK_{\overline{X}}| induces a birational map that restricts to an embedding on XX. This is the main technical part of the proof, and ultimately reduces to the construction of certain isolated non-klt centers and some careful analysis of certain Izumi type constants, see Section 3.2. Once the effective birationality is established, it is fairly straightforward to conclude the boundedness of XX.

1.2. Structure of the article

In Sections 2.22.5, we give the necessary background on mldK\mathrm{mld}^{\mathrm{K}}, local volumes, log Fano cone singularities and their boundedness. In sections 2.62.7, we collect some useful results in previous works; these include results from [HLS-epsilon-plt-blowup] that tackles pairs with real coefficients, as well as modifications of some results from the prequel [Z-mld^K-1] of the present work. In Section 3, we present and prove a more general boundedness statement for polarized log Fano cone singularities, Theorem 3.1. Our main Theorem 1.2 will follow as an application of Theorem 3.1.

Acknowledgement

The author is partially supported by the NSF Grants DMS-2240926, DMS-2234736, a Clay research fellowship, as well as a Sloan fellowship. He would like to thank Harold Blum, János Kollár, Yuchen Liu and Chenyang Xu for helpful discussions and comments. He also likes to thank the referees for careful reading of the manuscript and several helpful suggestions.

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 pair (X,Δ)(X,\Delta) consists of a normal variety XX together with an effective \mathbb{R}-divisor Δ\Delta on XX (a priori, we do not require that KX+ΔK_{X}+\Delta is \mathbb{R}-Cartier). 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).

Suppose that XX is a normal variety. A prime divisor FF on some birational model π:YX\pi\colon Y\to X (where YY is normal and π\pi is proper) of XX is called a divisor over XX. Given an \mathbb{R}-divisor Δ\Delta on XX, we denote its strict transform on the birational model YY by ΔY\Delta_{Y}. If KX+ΔK_{X}+\Delta is \mathbb{R}-Cartier, the log discrepancy AX,Δ(F)A_{X,\Delta}(F) is defined to be

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

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. 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}. Given λ\lambda\in\mathbb{R}, the corresponding valuation ideal 𝔞λ(v)\mathfrak{a}_{\lambda}(v) is

𝔞λ(v):={f𝒪X,xv(f)λ}.\mathfrak{a}_{\lambda}(v):=\{f\in\mathcal{O}_{X,x}\mid v(f)\geq\lambda\}.

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.

2.2. Kollár components

We first recall some definitions related to klt singularities and Kollár components.

Definition 2.1 ([KM98, Definition 2.34]).

We say a pair (X,Δ)(X,\Delta) is klt if KX+ΔK_{X}+\Delta is \mathbb{R}-Cartier, and for any prime divisor FF over XX we have AX,Δ(F)>0A_{X,\Delta}(F)>0. We say x(X,Δ)x\in(X,\Delta) is a klt singularity if (X,Δ)(X,\Delta) is klt.

Definition 2.2 ([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.

By adjunction, we may write

(KY+ΔY+E)|E=KE+ΔE(K_{Y}+\Delta_{Y}+E)|_{E}=K_{E}+\Delta_{E}

for some effective divisor ΔE=DiffE(ΔY)\Delta_{E}=\mathrm{Diff}_{E}(\Delta_{Y}) (called the different) on EE, and (E,ΔE)(E,\Delta_{E}) is a klt log Fano pair.

Definition 2.3 ([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).

If x(X,Δ)x\in(X,\Delta) is a klt singularity, then we can write Δ=i=1raiΔi\Delta=\sum_{i=1}^{r}a_{i}\Delta_{i} as a convex combination of \mathbb{Q}-divisors Δi\Delta_{i} such that each (X,Δi)(X,\Delta_{i}) is klt. Let r>0r>0 be an integer such that r(KX+Δi)r(K_{X}+\Delta_{i}) is Cartier for all ii. Then as AX,Δ(F)=i=1raiAX,Δi(F)A_{X,\Delta}(F)=\sum_{i=1}^{r}a_{i}A_{X,\Delta_{i}}(F), the possible values of log discrepancies AX,Δ(F)A_{X,\Delta}(F) belong to the discrete set {1ri=1raimi|mi}\left\{\left.\frac{1}{r}\sum_{i=1}^{r}a_{i}m_{i}\,\right|\,m_{i}\in\mathbb{N}\right\}. This implies that the infimum in the above definition is also a minimum.

The following result will be useful later. Recall that the log canonical threshold lct(X,Δ;D)\mathrm{lct}(X,\Delta;D) of an effective \mathbb{R}-Cartier divisor DD with respect to a klt pair (X,Δ)(X,\Delta) is the largest number t0t\geq 0 such that (X,Δ+tD)(X,\Delta+tD) is klt, and the α\alpha-invariant of a log Fano pair (X,Δ)(X,\Delta) is defined as the infimum of the log canonical thresholds lct(X,Δ;D)\mathrm{lct}(X,\Delta;D) where 0D(KX+Δ)0\leq D\sim_{\mathbb{R}}-(K_{X}+\Delta). The log canonical threshold lctx(X,Δ;D)\mathrm{lct}_{x}(X,\Delta;D) at a closed point xXx\in X is defined analogously.

Lemma 2.4.

Let EE be a Kollár component over a klt singularity x(X,Δ)x\in(X,\Delta). Then for any effective \mathbb{R}-Cartier divisor DD on XX, we have

lctx(X,Δ;D)min{1,α(E,ΔE)}AX,Δ(E)ordE(D).\mathrm{lct}_{x}(X,\Delta;D)\geq\min\{1,\alpha(E,\Delta_{E})\}\cdot\frac{A_{X,\Delta}(E)}{\mathrm{ord}_{E}(D)}.
Proof.

Let α=min{1,α(E,ΔE)}\alpha=\min\{1,\alpha(E,\Delta_{E})\}, let t=AX,Δ(E)ordE(D)t=\frac{A_{X,\Delta}(E)}{\mathrm{ord}_{E}(D)} and let π:YX\pi\colon Y\to X be the plt blowup of EE. Then we have

π(KX+Δ+tD)=KY+ΔY+tDY+E,\pi^{*}(K_{X}+\Delta+tD)=K_{Y}+\Delta_{Y}+tD_{Y}+E,

which gives tDY|E(KY+ΔY+E)|E(KE+ΔE)tD_{Y}|_{E}\sim_{\mathbb{R}}-(K_{Y}+\Delta_{Y}+E)|_{E}\sim_{\mathbb{R}}-(K_{E}+\Delta_{E}). By the definition of alpha invariants, the pair (E,ΔE+αtDY|E)(E,\Delta_{E}+\alpha tD_{Y}|_{E}) is log canonical, hence by inversion of adjunction [KM98, Theorem 5.50] we know that (Y,ΔY+αtDY+E)(Y,\Delta_{Y}+\alpha tD_{Y}+E) is lc around EE. As α1\alpha\leq 1, we also have

π(KX+Δ+αtD)=KY+ΔE+αtDY+sE\pi^{*}(K_{X}+\Delta+\alpha tD)=K_{Y}+\Delta_{E}+\alpha tD_{Y}+sE

for some s1s\leq 1, thus by the above discussion we deduce that (Y,ΔY+αtDY+sE)(Y,\Delta_{Y}+\alpha tD_{Y}+sE) is sub-lc around EE, and hence (X,Δ+αtD)(X,\Delta+\alpha tD) is lc at xx. In other words, lctx(X,Δ;D)αt\mathrm{lct}_{x}(X,\Delta;D)\geq\alpha t as desired. ∎

2.3. Local volumes

We next briefly recall the definition of the local volumes of klt singularities [Li-normalized-volume]. Let x(X,Δ)x\in(X,\Delta) be a klt singularity and let n=dimXn=\dim X. 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 divisor over XX, we have

AX,Δ(λordF)=λAX,Δ(F).A_{X,\Delta}(\lambda\cdot\mathrm{ord}_{F})=\lambda\cdot A_{X,\Delta}(F).

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!}.
Definition 2.5.

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).\widehat{\rm vol}_{X,\Delta}(v):=A_{X,\Delta}(v)^{n}\cdot\mathrm{vol}_{X,x}(v).

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. We will frequently use the following properties of local volumes.

Lemma 2.6 ([LX-cubic-3fold, Theorem 1.6]).

Let x(X,Δ)x\in(X,\Delta) be a klt singularity of dimension nn. Then vol^(x,X,Δ)nn\widehat{\rm vol}(x,X,\Delta)\leq n^{n}.

Lemma 2.7 ([XZ-minimizer-unique, Corollary 1.4]).

Let x(X,Δ)x\in(X,\Delta) be a klt singularity of dimension nn and let DD be a \mathbb{Q}-Cartier Weil divisor on XX. Then the Cartier index of DD is at most nnvol^(x,X,Δ)\frac{n^{n}}{\widehat{\rm vol}(x,X,\Delta)}.

2.4. Log Fano cone singularities

In this subsection we recall the definition of log Fano cone singularities and their K-semistability. These notions originally appear in the study of Sasaki-Einstein metrics [CS-Kss-Sasaki, CS-Sasaki-Einstein] and is further explored in works related to the Stable Degeneration Conjecture [LX-stability-higher-rank, LWX-metric-tangent-cone].

Definition 2.8.

Let X=𝐒𝐩𝐞𝐜(R)X=\mathbf{Spec}(R) be a normal affine variety and 𝕋=𝔾mr\mathbb{T}=\mathbb{G}_{m}^{r} (r>0r>0) an algebraic torus. 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, and call the corresponding singularity xXx\in X a 𝕋\mathbb{T}-singularity.

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 2.9.

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} (called a toric valuation) 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 2.10.

A log Fano cone singularities is a klt singularities 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.

We will often use the following result to perturb an irregular polarization to a quasi-regular one.

Lemma 2.11.

The function ξvol^X,Δ(wtξ)\xi\mapsto\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi}) defined on the Reeb cone is continuous and has a minimum.

Proof.

This follows from [LX-stability-higher-rank, Theorem 2.15(3) and Proposition 2.39]. ∎

Definition 2.12.

We say 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}).

This definition differs from the original ones from [CS-Kss-Sasaki, CS-Sasaki-Einstein], but they are equivalent by [LX-stability-higher-rank, Theorem 2.34]. For our purpose, the above definition is more convenient. Since the minimizer of the normalized volume function is unique up to rescaling [XZ-minimizer-unique], the polarization ξ\xi is essentially determined by the K-semistability condition and hence we often omit the polarization and simply say x(X,Δ)x\in(X,\Delta) is K-semistable.

Definition 2.13.

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})}.

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

Θ(x,X,Δ):=supξΘ(X,Δ;ξ),\Theta(x,X,\Delta):=\sup_{\xi}\Theta(X,\Delta;\xi),

where the supremum runs over all polarizations on XX.

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. By Lemma 2.11, similar statement holds in the unpolarized case.

2.5. Bounded family of singularities

In this subsection we define boundedness of singularities and recall some properties of singularity invariants in bounded families. They will be useful in proving the easier direction of Theorem 1.2.

Definition 2.14.

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\mathcal{X} is flat over BB, and B𝒳B\subseteq\mathcal{X} is a section of the projection,

  2. (2)

    For any closed point bBb\in B, 𝒳b\mathcal{X}_{b} is connected, normal and is not contained in Supp(𝒟)\mathrm{Supp}(\mathcal{D}),

  3. (3)

    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 any bBb\in B.

Lemma 2.15.

Let B(𝒳,𝒟)BB\subseteq(\mathcal{X},\mathcal{D})\to B be an \mathbb{R}-Gorenstein family of klt singularities. Then there exist constants ε,A>0\varepsilon,A>0 such that

vol^(b,𝒳b,𝒟b)εandmldK(b,𝒳b,𝒟b)A\widehat{\rm vol}(b,\mathcal{X}_{b},\mathcal{D}_{b})\geq\varepsilon\quad\mathrm{and}\quad\mathrm{mld}^{\mathrm{K}}(b,\mathcal{X}_{b},\mathcal{D}_{b})\leq A

for all closed point bBb\in B.

Proof.

The uniform lower bound of the local volumes follows from their lower semi-continuity [BL-vol-lsc] in families together with Noetherian induction. Note that although the main result of [BL-uks-openness] is stated for families with \mathbb{Q}-coefficients, it remains valid for real coefficients. This is because there exist \mathbb{Q}-divisors 𝒟𝒟\mathcal{D}^{\prime}\leq\mathcal{D} such that K𝒳/B+𝒟K_{\mathcal{X}/B}+\mathcal{D}^{\prime} is \mathbb{Q}-Cartier and the coefficients of 𝒟𝒟\mathcal{D}-\mathcal{D}^{\prime} are arbitrarily small; in particular, vol^(b,𝒳b,𝒟b)vol^(b,𝒳b,𝒟b)\widehat{\rm vol}(b,\mathcal{X}_{b},\mathcal{D}_{b})\geq\widehat{\rm vol}(b,\mathcal{X}_{b},\mathcal{D}^{\prime}_{b}) and their difference can be made arbitrarily small for any fixed bBb\in B. Thus the lower semicontinuity of vol^(b,𝒳b,𝒟b)\widehat{\rm vol}(b,\mathcal{X}_{b},\mathcal{D}_{b}) in bBb\in B follows from the lower semicontinuity of vol^(b,𝒳b,𝒟b)\widehat{\rm vol}(b,\mathcal{X}_{b},\mathcal{D}^{\prime}_{b}). Alternatively, the local volume lower bound follows from the constructibility of the volume function [Xu-quasi-monomial, Theorem 1.3], which also holds for families with real coefficients by the discussions in Theorem 2.19.

The uniform upper bound of mldK\mathrm{mld}^{\mathrm{K}} is a direct consequence of [HLQ-vol-ACC, Theorem 2.34]: after a (quasi-finite and surjective) base change, the family admits a flat family of Kollár components (in the sense of [HLQ-vol-ACC, Definition 2.32]), whose log discrepancy is locally constant and hence uniformly bounded. ∎

We next define the boundedness of log Fano cone singularities. Note that our definition is a priori stronger than their boundedness as klt singularities (it’s not clear to us whether they are equivalent).

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, 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}}. We say 𝒮\mathcal{S} is log bounded if we only ask x(X;ξ)x\in(X;\xi) to be isomorphic to b(𝒳i,b;ξb)b\in(\mathcal{X}_{i,b};\xi_{b}) and that Supp(Δ)Supp(𝒟i,b)\mathrm{Supp}(\Delta)\subseteq\mathrm{Supp}(\mathcal{D}_{i,b}) under this isomorphism.

We say that a set 𝒞\mathcal{C} of log Fano cone singularities is bounded if it’s the underlying set of singularities from a bounded set of polarized log Fano cone singularities.

Lemma 2.17.

Let 𝒞\mathcal{C} be a bounded set of log Fano cone singularities. Then there exists some constant θ>0\theta>0 such that Θ(x,X,Δ)θ\Theta(x,X,\Delta)\geq\theta for all x(X,Δ)x\in(X,\Delta) in 𝒞\mathcal{C}.

Proof.

By definition, it suffices to show that for any \mathbb{R}-Gorenstein family of log Fano cone singularities B(𝒳,𝒟)BB\subseteq(\mathcal{X},\mathcal{D})\to B with a fiberwise good torus 𝕋\mathbb{T}-action, we have a uniform positive lower bound of Θ(b,𝒳b,𝒟b)\Theta(b,\mathcal{X}_{b},\mathcal{D}_{b}). We may assume that BB is connected. Replacing 𝒳\mathcal{X} by 𝕋\mathbb{T}-invariant affine subset, We may also assume that 𝒳=𝐒𝐩𝐞𝐜()\mathcal{X}=\mathbf{Spec}(\mathcal{R}) is affine. We have a weight decomposition =αMα\mathcal{R}=\oplus_{\alpha\in M}\mathcal{R}_{\alpha} which reduces to the weight decomposition on the fibers. By flatness of 𝒳B\mathcal{X}\to B, each α\mathcal{R}_{\alpha} is flat over BB. It follows that the Reeb cone 𝔱,b+N\mathfrak{t}^{+}_{\mathbb{R},b}\subseteq N_{\mathbb{R}} is locally constant in bBb\in B, hence is constant as BB is connected. Choose any primitive ξN\xi\in N that lies in the (constant) Reeb cone. It suffices to show that Θ(b,𝒳b,𝒟b;ξ)\Theta(b,\mathcal{X}_{b},\mathcal{D}_{b};\xi) is uniformly bounded from below. Since the 𝕋\mathbb{T}-action is good, the α\mathcal{R}_{\alpha}’s are also finite over BB, hence locally free by flatness. From the defition of volume, this easily implies that vol(wtξ)\mathrm{vol}(\mathrm{wt}_{\xi}) is constant as bBb\in B varies. Since ξN\xi\in N comes from a 𝔾m\mathbb{G}_{m}-action, there is a divisor \mathcal{E} over 𝒳\mathcal{X} such that ord=wtξ\mathrm{ord}_{\mathcal{E}}=\mathrm{wt}_{\xi} as valuations over 𝒳\mathcal{X} and hence A𝒳b,𝒟b(wtξ)=A𝒳b,𝒟b(ordb)=A𝒳,𝒟()A_{\mathcal{X}_{b},\mathcal{D}_{b}}(\mathrm{wt}_{\xi})=A_{\mathcal{X}_{b},\mathcal{D}_{b}}(\mathrm{ord}_{\mathcal{E}_{b}})=A_{\mathcal{X},\mathcal{D}}(\mathcal{E}) for general bBb\in B. By Noetherian induction, this implies that A𝒳b,𝒟b(wtξ)A_{\mathcal{X}_{b},\mathcal{D}_{b}}(\mathrm{wt}_{\xi}) is constructible in bBb\in B and thus uniformly bounded from above. Finally, by Lemma 2.15 we know that vol^(b,𝒳b,𝒟b)ε\widehat{\rm vol}(b,\mathcal{X}_{b},\mathcal{D}_{b})\geq\varepsilon for some constant ε>0\varepsilon>0. Putting these together we deduce from the definition that Θ(b,𝒳b,𝒟b;ξ)\Theta(b,\mathcal{X}_{b},\mathcal{D}_{b};\xi) is uniformly bounded from below. ∎

2.6. Real coefficients

The following result will help us reduce many questions about pairs with real coefficients to ones with rational coefficients. We will often use it without explicit mention when we refer to results that are originally stated for \mathbb{Q}-coefficients.

Lemma 2.18.

Let nn\in\mathbb{N} and let I[0,1]I\subseteq[0,1] be a finite set. Then there exists a finite set I[0,1]I^{\prime}\subseteq[0,1]\cap\mathbb{Q} depending only on nn and II such that the following holds.

For any nn-dimensional klt singularity x(X,Δ)x\in(X,\Delta) with coefficients in II, there exists some effective \mathbb{Q}-divisor ΔΔ\Delta^{\prime}\geq\Delta on XX with coefficients in II^{\prime}, such that Supp(Δ)=Supp(Δ)\mathrm{Supp}(\Delta)=\mathrm{Supp}(\Delta^{\prime}), x(X,Δ)x\in(X,\Delta^{\prime}) is klt, vol^(x,X,Δ)2nvol^(x,X,Δ)\widehat{\rm vol}(x,X,\Delta^{\prime})\geq 2^{-n}\widehat{\rm vol}(x,X,\Delta) and mldK(x,X,Δ)2mldK(x,X,Δ)\mathrm{mld}^{\mathrm{K}}(x,X,\Delta^{\prime})\leq 2\cdot\mathrm{mld}^{\mathrm{K}}(x,X,\Delta). If in addition x(X,Δ;ξ)x\in(X,\Delta;\xi) is a log Fano cone singularity, then Θ(x,X,Δ;ξ)4nΘ(x,X,Δ;ξ)\Theta(x,X,\Delta^{\prime};\xi)\geq 4^{-n}\Theta(x,X,\Delta;\xi).

Proof.

This is essentially a consequence of [HLS-epsilon-plt-blowup, Theorem 5.6]. Without loss of generality we may assume 1I1\in I. Write I={a1,,am}I=\{a_{1},\dots,a_{m}\} and let 𝐚=(a1,,am)m\mathbf{a}=(a_{1},\dots,a_{m})\in\mathbb{R}^{m}. Let VmV\subseteq\mathbb{R}^{m} be the rational envelope of 𝐚\mathbf{a}, i.e. the smallest affine subspace defined over \mathbb{Q} that contains 𝐚\mathbf{a}. By [HLS-epsilon-plt-blowup, Theorem 5.6], there exists an open neighbourhood UU of 𝐚V\mathbf{a}\in V depending only on nn and II such that for any nn-dimensional singularity xXx\in X and any Weil divisors Δ1,,Δm0\Delta_{1},\dots,\Delta_{m}\geq 0, if x(X,i=1maiΔi)x\in(X,\sum_{i=1}^{m}a_{i}\Delta_{i}) is klt (resp. plt) then x(X,i=1maiΔi)x\in(X,\sum_{i=1}^{m}a^{\prime}_{i}\Delta_{i}) is klt (resp. plt) for any 𝐚=(a1,,am)U\mathbf{a}^{\prime}=(a^{\prime}_{1},\dots,a^{\prime}_{m})\in U. Note that the plt case implies that if EE is a Kollár component over x(X,i=1maiΔi)x\in(X,\sum_{i=1}^{m}a_{i}\Delta_{i}) then it is also a Kollár component over x(X,i=1maiΔi)x\in(X,\sum_{i=1}^{m}a^{\prime}_{i}\Delta_{i}).

Choose some 𝐚=(a1,,am)Um\mathbf{a}^{\prime}=(a^{\prime}_{1},\dots,a^{\prime}_{m})\in U\cap\mathbb{Q}^{m} such that 2𝐚𝐚,2𝐚𝐚U2\mathbf{a}^{\prime}-\mathbf{a},2\mathbf{a}-\mathbf{a}^{\prime}\in U. We claim that I={a1,,am}I^{\prime}=\{a^{\prime}_{1},\dots,a^{\prime}_{m}\} satisfies the required conditions. Indeed, for any nn-dimensional klt singularity x(X,Δ=i=1maiΔi)x\in(X,\Delta=\sum_{i=1}^{m}a_{i}\Delta_{i}) (where the Δi\Delta_{i}’s are Weil divisors), if we set Δ=i=1maiΔi\Delta^{\prime}=\sum_{i=1}^{m}a^{\prime}_{i}\Delta_{i} and let Δ1=2ΔΔ\Delta_{1}=2\Delta^{\prime}-\Delta, then by our choice of UU and 𝐚\mathbf{a}^{\prime} we know that Supp(Δ)=Supp(Δ)\mathrm{Supp}(\Delta^{\prime})=\mathrm{Supp}(\Delta), x(X,Δ)x\in(X,\Delta^{\prime}) and x(X,Δ1)x\in(X,\Delta_{1}) are both klt, and any Kollár component over x(X,Δ)x\in(X,\Delta) is also a Kollár component over x(X,Δ)x\in(X,\Delta^{\prime}). Moreover, since Δ=12(Δ+Δ1)\Delta^{\prime}=\frac{1}{2}(\Delta+\Delta_{1}), we have AX,Δ(v)=12(AX,Δ(v)+AX,Δ1(v))12AX,Δ(v)A_{X,\Delta^{\prime}}(v)=\frac{1}{2}(A_{X,\Delta}(v)+A_{X,\Delta_{1}}(v))\geq\frac{1}{2}A_{X,\Delta}(v) for any valuation vValX,xv\in\mathrm{Val}_{X,x}^{*}. Similarly using 2𝐚𝐚U2\mathbf{a}-\mathbf{a}^{\prime}\in U we get AX,Δ(v)2AX,Δ(v)A_{X,\Delta^{\prime}}(v)\leq 2A_{X,\Delta}(v). In particular we see that 2nvol^X,Δ(v)vol^X,Δ(v)2nvol^X,Δ(v)2^{-n}\widehat{\rm vol}_{X,\Delta}(v)\leq\widehat{\rm vol}_{X,\Delta^{\prime}}(v)\leq 2^{n}\widehat{\rm vol}_{X,\Delta}(v). These imply the inequalities about vol^\widehat{\rm vol}, mldK\mathrm{mld}^{\mathrm{K}} and the volume ratio (all by definition). ∎

We also note the Stable Degeneration Conjecture holds for pairs with real coefficients.

Theorem 2.19 ([Blu-minimizer-exist, LX-stability-higher-rank, Xu-quasi-monomial, XZ-minimizer-unique, XZ-SDC]).

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

Proof.

The arguments in [Blu-minimizer-exist, LX-stability-higher-rank, XZ-minimizer-unique, XZ-SDC] extend directly to real coefficients, since they are not sensitive to the coefficients. The proof in [Xu-quasi-monomial] uses the existence of monotonic NN-complement which requires \mathbb{Q}-coefficients, but the main results in loc. cit. can be extended to \mathbb{R}-coefficients by [HLQ-vol-ACC, Theorems 3.3 and 3.4]. Thus [XZ-SDC, Theorem 1.2] holds for \mathbb{R}-pairs. In particular, the minimizer vv of the normalized volume function vol^X,Δ\widehat{\rm vol}_{X,\Delta} induces a special degeneration of x(X=𝐒𝐩𝐞𝐜(R),Δ)x\in(X=\mathbf{Spec}(R),\Delta) to a K-semistable log Fano cone x0(X0=𝐒𝐩𝐞𝐜(grvR),Δ0;ξv)x_{0}\in(X_{0}=\mathbf{Spec}(\mathrm{gr}_{v}R),\Delta_{0};\xi_{v}). By [LX-stability-higher-rank, Lemma 2.58], we have

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

Since vol^(x,X,Δ)=vol^X,Δ(v)\widehat{\rm vol}(x,X,\Delta)=\widehat{\rm vol}_{X,\Delta}(v) (as vv is the minimizer of vol^X,Δ\widehat{\rm vol}_{X,\Delta}) and vol^(x0,X0,Δ0)=vol^X0,Δ0(wtξv)\widehat{\rm vol}(x_{0},X_{0},\Delta_{0})=\widehat{\rm vol}_{X_{0},\Delta_{0}}(\mathrm{wt}_{\xi_{v}}) (as x0(X0,Δ0;ξv)x_{0}\in(X_{0},\Delta_{0};\xi_{v}) is K-semistable), we see that the degeneration preserves the local volume. ∎

2.7. Results from Part I

In this subsection we collect slight modification of several results from [Z-mld^K-1] that we need in later proofs. For the first result, recall that we say a singularity x(X,Δ)x\in(X,\Delta) is of klt type if there exists some effective \mathbb{R}-divisor DD on XX such that x(X,Δ+D)x\in(X,\Delta+D) is klt.

Lemma 2.20.

Let nn\in\mathbb{N} and let ε>0\varepsilon>0. Then there exists some constant c=c(n,ε)>0c=c(n,\varepsilon)>0 such that for any nn-dimensional klt singularity x(X,Δ)x\in(X,\Delta) with vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon, we have that x(X,(1+c)Δ)x\in(X,(1+c)\Delta) is of klt type.

Proof.

Since x(X,Δ)x\in(X,\Delta) is klt, by [BCHM] (cf. [Z-mld^K-1, Lemma 4.7]) there exists a small birational morphism π:YX\pi\colon Y\to X such that KYK_{Y} is \mathbb{Q}-Cartier and relatively ample. Since π\pi is small, we have KY+ΔY=π(KX+Δ)K_{Y}+\Delta_{Y}=\pi^{*}(K_{X}+\Delta), hence by [LX-cubic-3fold, Lemma 2.9(2)] (cf. the proof of [Z-mld^K-1, Lemma 2.10]) we have

vol^(y,Y,ΔY)vol^(x,X,Δ)ε\widehat{\rm vol}(y,Y,\Delta_{Y})\geq\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon

for all yπ1(x)y\in\pi^{-1}(x). By [Z-mld^K-1, Theorem 3.1], there exists some constant c=c(n,ε)>0c=c(n,\varepsilon)>0 such that the pair (Y,(1+c)ΔY)(Y,(1+c)\Delta_{Y}) is klt for all yπ1(x)y\in\pi^{-1}(x). Note that (KY+(1+c)ΔY),πcKY-(K_{Y}+(1+c)\Delta_{Y})\sim_{\mathbb{Q},\pi}cK_{Y}, hence π\pi is also the ample model of (KX+(1+c)Δ)-(K_{X}+(1+c)\Delta). By [Z-direct-summand, Lemma 2.4], this implies that x(X,(1+c)Δ)x\in(X,(1+c)\Delta) is of klt type. ∎

The second result is extracted from the proof of [Z-mld^K-1, Theorem 4.1].

Lemma 2.21.

Let nn\in\mathbb{N}, let ε,A>0\varepsilon,A>0, and let I[0,1]I\subseteq[0,1]\cap\mathbb{Q} be a finite set. Then there exists some integer N=N(n,ε,A,I)>0N=N(n,\varepsilon,A,I)>0 such that for any nn-dimensional klt singularity x(X,Δ)x\in(X,\Delta) with coefficients in II and vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon, and for any Kollár component EE over x(X,Δ)x\in(X,\Delta) with AX,Δ(E)AA_{X,\Delta}(E)\leq A, we have that NENE and N(KY+ΔY+E)N(K_{Y}+\Delta_{Y}+E) are Cartier on the plt blowup YXY\to X of EE.

Proof.

By [Z-mld^K-1, (4.1)], the local volumes vol^(y,Y,ΔY)\widehat{\rm vol}(y,Y,\Delta_{Y}) (where yEy\in E) are bounded from below by some constants that only relies on the given constants n,ε,An,\varepsilon,A. Thus by Lemma 2.7 we get the desired Cartier index bound. ∎

3. Boundedness

In this section, we prove the following statement on boundedness of log Fano cone singularities, which will imply the main results of this paper.

Theorem 3.1.

Let nn\in\mathbb{N} and let I[0,1]I\subseteq[0,1] be a finite set. Let ε,θ,A>0\varepsilon,\theta,A>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,Θ(X,Δ;ξ)θ,andmldK(x,X,Δ)A.\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon,\quad\Theta(X,\Delta;\xi)\geq\theta,\quad\mathrm{and}\quad\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq A.

Then 𝒮\mathcal{S} is bounded.

Recall that Θ(X,Δ;ξ)\Theta(X,\Delta;\xi) is the volume ratio of the log Fano cone singularity (Definition 2.13). We refer to Sections 2.22.5 for the other relevant definitions.

3.1. Orbifold cones

Since the volume ratio is continuous in the Reeb vector ξ\xi (Lemma 2.11), by perturbing the polarization, we see that it suffices to prove Theorem 3.1 when 𝒮\mathcal{S} consists of quasi-regular log Fano cones. Every quasi-regular log Fano cone singularity has a natural affine orbifold cone structure induced by the polarization. The proof of Theorem 3.1 relies on the associated projective orbifold cone construction. In this subsection, we first fix some notation and recall some basic properties of orbifold cones from [Kol-Seifert-bundle].

Definition 3.2.

Let VV be a normal projective variety. Let LL be an ample \mathbb{Q}-Cartier \mathbb{Q}-divisor on VV.

  1. (1)

    The affine orbifold cone Ca(V,L)C_{a}(V,L) is defined as

    Ca(V,L):=𝐒𝐩𝐞𝐜m=0H0(V,𝒪V(mL)).C_{a}(V,L):=\mathbf{Spec}\bigoplus_{m=0}^{\infty}H^{0}(V,\mathcal{O}_{V}(\lfloor mL\rfloor)).
  2. (2)

    The projective orbifold cone Cp(V,L)C_{p}(V,L) is defined as

    Cp(V,L):=𝐏𝐫𝐨𝐣m=0i=0H0(V,𝒪V(mL)si,C_{p}(V,L):=\mathbf{Proj}\bigoplus_{m=0}^{\infty}\bigoplus_{i=0}^{\infty}H^{0}(V,\mathcal{O}_{V}(\lfloor mL\rfloor)\cdot s^{i},

    where the grading of H0(V,𝒪V(mL))H^{0}(V,\mathcal{O}_{V}(\lfloor mL\rfloor)) and ss are mm and 11, respectively.

For ease of notation, denote X=Ca(V,L)X=C_{a}(V,L), X¯=Cp(V,L)\overline{X}=C_{p}(V,L) and let xXx\in X be the vertex of the orbifold cone. On the projective orbifold cone, we also denote by VV_{\infty} the divisor at infinity, i.e., the divisor corresponding to (s=0)(s=0). We have XX¯VX\cong\overline{X}\setminus V_{\infty}. Note that X{x}X\setminus\{x\} is a Seifert 𝔾m\mathbb{G}_{m}-bundle over VV (in the sense of [Kol-Seifert-bundle]). Thus for any effective \mathbb{R}-divisor ΔV\Delta_{V} on VV, we can define the affine (resp. projective) orbifold cone over the polarized pair (V,ΔV;L)(V,\Delta_{V};L) as the pair (X,Δ)(X,\Delta) (resp. (X¯,Δ¯)(\overline{X},\overline{\Delta})), where Δ\Delta (resp. Δ¯\overline{\Delta}) is the closure of the pullback of ΔV\Delta_{V} to X{x}X\setminus\{x\} (since the projection X{x}VX\setminus\{x\}\to V is equidimensional, the pullback of a Weil divisor is well-defined). Every quasi-regular polarized log Fano cone singularity x(X,Δ;ξ)x\in(X,\Delta;\xi) arises in this way, so we can talk about its associated projective orbifold cone (X¯,Δ¯)(\overline{X},\overline{\Delta}).

Alternatively, the Seifert 𝔾m\mathbb{G}_{m}-bundle X{x}X\setminus\{x\} can be compactified by adding VV_{\infty} and the missing zero section V0V_{0}; the resulting space Y¯\overline{Y} is also the orbifold blowup of X¯\overline{X} at xx. Similarly, let YY be the orbifold blowup of XX at xx (i.e. Y=Y¯VY=\overline{Y}\setminus V_{\infty}). Write the fractional part of LL as {L}=i=1kaibiDi\{L\}=\sum_{i=1}^{k}\frac{a_{i}}{b_{i}}D_{i} for some prime divisors DiD_{i} on VV and 0<ai<bi0<a_{i}<b_{i} coprime integers. We denote ΔL:=i=1kbi1biDi\Delta_{L}:=\sum_{i=1}^{k}\frac{b_{i}-1}{b_{i}}D_{i}. Then the pairs (V0,DiffV0(0))(V_{0},\mathrm{Diff}_{V_{0}}(0)) and (V,DiffV(0))(V_{\infty},\mathrm{Diff}_{V_{\infty}}(0)) obtained by taking adjunction over some big open set of VV from the pair (X¯,V0+V)(\overline{X},V_{0}+V_{\infty}) are both isomorphic to (V,ΔL)(V,\Delta_{L}) (this is a local computation, see [Kol-Seifert-bundle, Section 4]).

Under these notation, we have the following well known result [Kol-Seifert-bundle] (cf. [Kol13, Section 3.1] for analogous statement for usual cones).

Lemma 3.3.

The following conditions are equivalent:

  1. (1)

    x(X,Δ)x\in(X,\Delta) is klt;

  2. (2)

    (X¯,Δ¯+V)(\overline{X},\overline{\Delta}+V_{\infty}) is plt;

  3. (3)

    (V,ΔV+ΔL)(V,\Delta_{V}+\Delta_{L}) is a log Fano pair, and (KV+ΔV+ΔL)rL-(K_{V}+\Delta_{V}+\Delta_{L})\sim_{\mathbb{R}}rL for some r>0r>0.

Moreover, when the above conditions are satisfied, we have KX¯+Δ¯(1+r)VK_{\overline{X}}+\overline{\Delta}\sim_{\mathbb{R}}-(1+r)V_{\infty} and AX,Δ(V0)=rA_{X,\Delta}(V_{0})=r.

The next result is a key observation in the proof of Theorem 3.1. It expresses the normalized volume of a polarized log Fano cone singularities as the global volume of the associated projective orbifold cone.

Lemma 3.4.

Assume that x(X,Δ)x\in(X,\Delta) is klt. Then under the notation of Lemma 3.3 we have

vol^X,Δ(ordV0)=vol((KX¯+Δ¯+V))=rvol((KV+ΔV+ΔL)).\widehat{\rm vol}_{X,\Delta}(\mathrm{ord}_{V_{0}})=\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}))=r\cdot\mathrm{vol}(-(K_{V}+\Delta_{V}+\Delta_{L})).
Proof.

This follows from a direct calculation. First we have AX,Δ(V0)=rA_{X,\Delta}(V_{0})=r by Lemma 3.3. We also have V0|V0L-V_{0}|_{V_{0}}\cong L, hence

vol^X,Δ(ordV0)=AX,Δ(V0)nvolX,Δ(ordV0)=rnvol(V0|V0)=rnLn1\widehat{\rm vol}_{X,\Delta}(\mathrm{ord}_{V_{0}})=A_{X,\Delta}(V_{0})^{n}\cdot\mathrm{vol}_{X,\Delta}(\mathrm{ord}_{V_{0}})=r^{n}\mathrm{vol}(-V_{0}|_{V_{0}})=r^{n}L^{n-1}

where n=dimXn=\dim X. On the other hand, by Lemma 3.3 we get (KX¯+Δ¯+V)rV-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})\sim_{\mathbb{R}}rV_{\infty}; recall also that (KX¯+Δ¯+V)|V(KV+ΔV+ΔL)rL-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})|_{V_{\infty}}\sim_{\mathbb{R}}-(K_{V}+\Delta_{V}+\Delta_{L})\sim_{\mathbb{Q}}rL (where we identify VV_{\infty} with VV), thus

vol((KX¯+Δ¯+V))\displaystyle\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})) =((KX¯+Δ¯+V))n1(rV)\displaystyle=\left(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})\right)^{n-1}\cdot(rV_{\infty})
=rvol((KV+ΔV+ΔL))=r(rL)n1=rnLn1,\displaystyle=r\cdot\mathrm{vol}(-(K_{V}+\Delta_{V}+\Delta_{L}))=r\cdot(rL)^{n-1}=r^{n}L^{n-1},

which proves the desired equality. ∎

We also need a slight generalization of Lemma 3.3.

Lemma 3.5.

Notation as before. Assume that x(X,Δ)x\in(X,\Delta) is of klt type. Then (KV+ΔV+ΔL)-(K_{V}+\Delta_{V}+\Delta_{L}) is big.

Proof.

By assumption, there exists some effective \mathbb{R}-divisor DD such that x(X,Δ+D)x\in(X,\Delta+D) is klt (note that DD is not necessarily invariant under the 𝔾m\mathbb{G}_{m}-action). In particular, AX,Δ+D(V0)>0A_{X,\Delta+D}(V_{0})>0 and we have

KY+ΔY+DY+V0AX,Δ+D(V0)V0K_{Y}+\Delta_{Y}+D_{Y}+V_{0}\sim_{\mathbb{R}}A_{X,\Delta+D}(V_{0})\cdot V_{0}

over XX. It follows that (KY+ΔY+DY+V0)|V0AX,Δ+D(V0)L-(K_{Y}+\Delta_{Y}+D_{Y}+V_{0})|_{V_{0}}\sim_{\mathbb{R}}A_{X,\Delta+D}(V_{0})\cdot L is ample. By adjunction, this implies that

(KV+ΔV+ΔL)=(KY+ΔY+DY+V0)|V0+DY|V0-(K_{V}+\Delta_{V}+\Delta_{L})=-(K_{Y}+\Delta_{Y}+D_{Y}+V_{0})|_{V_{0}}+D_{Y}|_{V_{0}}

is the sum of an ample divisor and an effective divisor, hence is big. ∎

3.2. Effective birationality

Given Lemma 3.4, a naïve idea to prove Theorem 3.1 is to associate to each log Fano cone singularity a projective orbifold cone and show that the corresponding set of projective orbifold cones is bounded. In general, this is too much to hope for, as the projective orbifold cone depends on the (auxiliary) choice of a quasi-regular polarization. The following example shows that even for a fixed singularity we can get an unbounded family of projective orbifold cones by choosing different polarizations.

Example 3.6.

Let a1,,ana_{1},\dots,a_{n}\in\mathbb{N} be pairwise coprime positive integers. Then ξ=(a1,,an)n\xi=(a_{1},\dots,a_{n})\in\mathbb{N}^{n} gives a polarization of the Fano cone singularity 0𝔸n0\in\mathbb{A}^{n}; it generates the 𝔾m\mathbb{G}_{m}-action with weights a1,,ana_{1},\dots,a_{n} on the coordinates. This endows 𝔸n\mathbb{A}^{n} with an affine orbifold cone structure Ca(V,L)C_{a}(V,L) where V=(a1,,an)V=\mathbb{P}(a_{1},\dots,a_{n}) and L=𝒪V(1)L=\mathcal{O}_{V}(1). The associated projective orbifold cone is X¯=(1,a1,,an)\overline{X}=\mathbb{P}(1,a_{1},\dots,a_{n}), which is clearly unbounded as the weights aia_{i}’s vary. By choosing appropriate weights, we can even ensure that the normalized volume vol^(wtξ)=(a1++an)na1an\widehat{\rm vol}(\mathrm{wt}_{\xi})=\frac{(a_{1}+\dots+a_{n})^{n}}{a_{1}\dots a_{n}} is fixed.

Nonetheless, we observe that in the above example the projective orbifold cones we get satisfy the following interesting property: the linear system |KX¯||-K_{\overline{X}}| always defines a birational map that is an embedding at the vertex [1:0::0][1:0:\dots:0]. In fact, if [s:x1::xn][s:x_{1}:\dots:x_{n}] are the weighted homogeneous coordinates of X¯\overline{X}, then for every i{1,,n}i\in\{1,\dots,n\} there exists some kik_{i}\in\mathbb{N} such that skixiH0(KX¯)s^{k_{i}}x_{i}\in H^{0}(-K_{\overline{X}}) (this is possible because ss has weight 11); it is not hard to see that the sub-linear system spanned by skixis^{k_{i}}x_{i} (i=0,,ni=0,\dots,n) is base point free and restricts to an embedding on the affine chart 𝔸n=X¯(s=0)\mathbb{A}^{n}=\overline{X}\setminus(s=0).

This motivates us to raise the following question.

Question 3.7.

Let nn\in\mathbb{N} and ε>0\varepsilon>0. Let I[0,1]I\subseteq[0,1]\cap\mathbb{Q} be a finite set. Is there some integer m=m(n,ε,I)>0m=m(n,\varepsilon,I)>0 such that for any nn-dimensional quasi-regular polarized log Fano cone x(X,Δ;ξ)x\in(X,\Delta;\xi) with vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon, we have that |m(KX¯+Δ¯+V)||-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})| defines a birational map that restricts to an embedding on XX?

As before, (X¯,Δ¯)(\overline{X},\overline{\Delta}) denotes the associated projective orbifold cone, and V=X¯XV_{\infty}=\overline{X}\setminus X is the divisor at infinity.

The technical core of the proof of Theorem 3.1 is the answer to this question after imposing upper bounds on the minimal log discrepancies of Kollár components.

Proposition 3.8.

Let nn\in\mathbb{N} and ε,A>0\varepsilon,A>0. Let I[0,1]I\subseteq[0,1]\cap\mathbb{Q} be a finite set. Then there is some integer m=m(n,ε,A,I)>0m=m(n,\varepsilon,A,I)>0 such that for any nn-dimensional quasi-regular polarized log Fano cone x(X,Δ;ξ)x\in(X,\Delta;\xi) with vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon and mldK(x,X,Δ)A\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq A, we have that |m(KX¯+Δ¯+V)||-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})| defines a birational map that restricts to an embedding on XX.

We remark that the integer mm is independent of the polarization ξ\xi.

Proof.

We start with some reductions. Let H:=(KX¯+Δ¯+V)H:=-(K_{\overline{X}}+\overline{\Delta}+V_{\infty}). We view XX as an affine orbifold cone Ca(V,L)C_{a}(V,L) and keep the notation (e.g. ΔV,ΔL\Delta_{V},\Delta_{L}, etc) from Section 3.1. First note that since the coefficients of Δ¯\overline{\Delta} belong to a fixed finite set II of rational numbers, we can choose some m0m_{0} depending only on II such that m0Hm_{0}H has integer coefficients. Since vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon, by Lemma 2.7 we know that the Cartier index of m0Hm_{0}H at xx is bounded from above by nnε\frac{n^{n}}{\varepsilon}. Thus replacing m0m_{0} by a sufficiently large fixed multiple (e.g. nnε!\lfloor\frac{n^{n}}{\varepsilon}\rfloor!) we may further assume that m0Hm_{0}H is Cartier in a neighbourhood of xx. By 𝔾m\mathbb{G}_{m}-translation, this implies that m0Hm_{0}H is Cartier on XX.

The next step is to produce enough sections (of some multiple of m0Hm_{0}H) that do not vanish at xx. We do this by creating isolated non-klt centers at xx and apply Nadel vanishing. Let φ:Y¯X¯\varphi\colon\overline{Y}\to\overline{X} be the orbifold blowup of the vertex xx, and let ΔY¯\Delta_{\overline{Y}} be the strict transform of Δ¯\overline{\Delta}. Let V0V_{0} be the exceptional divisor as before. Recall that Y¯\overline{Y} has an orbifold 1\mathbb{P}^{1}-bundle structure π:Y¯V\pi\colon\overline{Y}\to V. Then we have (see [Kol-Seifert-bundle, Proposition 40 and Corollary 41])

(KY¯+ΔY¯+V0+V)π(KV+ΔV+ΔL).-(K_{\overline{Y}}+\Delta_{\overline{Y}}+V_{0}+V_{\infty})\sim_{\mathbb{Q}}-\pi^{*}(K_{V}+\Delta_{V}+\Delta_{L}).

The left hand side is φHaV0\sim_{\mathbb{Q}}\varphi^{*}H-aV_{0} where a=AX,Δ(V0)>0a=A_{X,\Delta}(V_{0})>0, while the right hand side is semiample since (KV+ΔV+ΔL)-(K_{V}+\Delta_{V}+\Delta_{L}) is ample on VV by Lemma 3.3. From here we deduce that φHaV0\varphi^{*}H-aV_{0} is semiample. For any positive integer mm, if we take DD to be a general member of the \mathbb{Q}-linear system φ|φHaV0|\varphi_{*}|\varphi^{*}H-aV_{0}|_{\mathbb{Q}}, then DHD\sim_{\mathbb{Q}}H and (X¯,Δ¯+mD)(\overline{X},\overline{\Delta}+mD) is klt away from xx (since this is the only base point of the \mathbb{Q}-linear system). In particular, the multiplier ideal 𝒥(X¯,Δ¯+mD)\mathcal{J}(\overline{X},\overline{\Delta}+mD) is co-supported at xx. Furthermore, we have

ordV0𝒥(X¯,Δ¯+mD)>mordV0(D)AX¯,Δ¯(V0)=(m1)a\mathrm{ord}_{V_{0}}\mathcal{J}(\overline{X},\overline{\Delta}+mD)>m\cdot\mathrm{ord}_{V_{0}}(D)-A_{\overline{X},\overline{\Delta}}(V_{0})=(m-1)a

by the definition of multiplier ideals. Thus we obtain

(3.1) 𝒥(X¯,Δ¯+mD)𝔞(m1)a,\mathcal{J}(\overline{X},\overline{\Delta}+mD)\subseteq\mathfrak{a}_{(m-1)a},

where 𝔞(m1)a:=𝔞(m1)a(ordV0)\mathfrak{a}_{(m-1)a}:=\mathfrak{a}_{(m-1)a}(\mathrm{ord}_{V_{0}}) denotes the valuation ideals. Recall that m0Hm_{0}H is Cartier at xx, hence as

mH(KX¯+Δ¯+mD)(KX¯+Δ¯)mH-(K_{\overline{X}}+\overline{\Delta}+mD)\sim_{\mathbb{Q}}-(K_{\overline{X}}+\overline{\Delta})

is ample, by the following Lemma 3.9 (Nadel vanishing for Weil divisor), we have

H1(X¯,𝒥(X¯,Δ¯+mD)𝒪X¯(mH))=0H^{1}(\overline{X},\mathcal{J}(\overline{X},\overline{\Delta}+mD)\otimes\mathcal{O}_{\overline{X}}(mH))=0

for all mm\in\mathbb{N} that’s divisible by m0m_{0}. From the long exact sequence we then deduce that the natural map

H0(X¯,𝒪X¯(mH))H0(X¯,𝒪X¯/𝒥(X¯,Δ¯+mD))H^{0}(\overline{X},\mathcal{O}_{\overline{X}}(mH))\to H^{0}(\overline{X},\mathcal{O}_{\overline{X}}/\mathcal{J}(\overline{X},\overline{\Delta}+mD))

is surjective. Combined with (3.1), we get a surjection

(3.2) H0(X¯,𝒪X¯(mH))H0(X¯,𝒪X¯/𝔞(m1)a)H^{0}(\overline{X},\mathcal{O}_{\overline{X}}(mH))\to H^{0}(\overline{X},\mathcal{O}_{\overline{X}}/\mathfrak{a}_{(m-1)a})

for all mm\in\mathbb{N} that’s divisible by m0m_{0}.

At this point, we have produced sections of mHmH that separate the jets in 𝒪X¯/𝔞(m1)a\mathcal{O}_{\overline{X}}/\mathfrak{a}_{(m-1)a}. Our goal is to make |mH||mH| into a birational map that restricts to an embedding on XX. Clearly, a necessary condition is that |mH||mH| separates tangent directions at xx. Let us first show that this latter condition can be achieved. In view of (3.2), it suffices to show that

(3.3) 𝔞(m1)a𝔪x2,\mathfrak{a}_{(m-1)a}\subseteq\mathfrak{m}_{x}^{2},

which becomes a local question. If XX is fixed, this is certainly true for m0m\gg 0, so the main point is to make sure that the bound on mm depend only on the given data n,ε,An,\varepsilon,A rather than on XX. To this end, note that one way to interpret the opposite condition 𝔞(m1)a𝔪x2\mathfrak{a}_{(m-1)a}\not\subseteq\mathfrak{m}_{x}^{2} is that there exists some f𝔪x𝔪x2f\in\mathfrak{m}_{x}\setminus\mathfrak{m}_{x}^{2} such that ordV0(f)(m1)a\mathrm{ord}_{V_{0}}(f)\geq(m-1)a; in particular, lct(f):=lct(X,Δ;(f=0))1m1\mathrm{lct}(f):=\mathrm{lct}(X,\Delta;(f=0))\leq\frac{1}{m-1}. This suggests to us that we should try to analyze the Izumi type constant

(3.4) inf{lct(f)f𝔪x𝔪x2}\inf\{\mathrm{lct}(f)\mid f\in\mathfrak{m}_{x}\setminus\mathfrak{m}_{x}^{2}\}

for the singularities in question.

Let EE be a Kollár component over x(X,Δ)x\in(X,\Delta) such that AX,Δ(E)AA_{X,\Delta}(E)\leq A, whose existence is guaranteed by our assumption that mldK(x,X,Δ)A\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq A. Let f:ZXf\colon Z\to X be the plt blowup that extract EE and let ΔE=DiffE(ΔZ)\Delta_{E}=\mathrm{Diff}_{E}(\Delta_{Z}) be the different. By Lemma 2.4, we have

(3.5) lct(f)min{1,α(E,ΔE)}AX,Δ(E)ordE(f).\mathrm{lct}(f)\geq\min\{1,\alpha(E,\Delta_{E})\}\cdot\frac{A_{X,\Delta}(E)}{\mathrm{ord}_{E}(f)}.

In order to give uniform estimate of (3.4) using (3.5), we need further information about the log discrepancy AX,Δ(E)A_{X,\Delta}(E), the α\alpha-invariant α(E,ΔE)\alpha(E,\Delta_{E}), and the value of

dE:=sup{ordE(f)f𝔪x𝔪x2}d_{E}:=\sup\{\mathrm{ord}_{E}(f)\mid f\in\mathfrak{m}_{x}\setminus\mathfrak{m}_{x}^{2}\}

(the last one can be thought of as measuring how well the valuation ideals of ordE\mathrm{ord}_{E} approximate the first two powers of the maximal ideal 𝔪x\mathfrak{m}_{x}).

For the log discrepancy, recall from the previous discussion that m0(KX+Δ)m_{0}(K_{X}+\Delta) is Cartier. Since (X,Δ)(X,\Delta) is klt, this implies that AX,Δ(E)1m0A_{X,\Delta}(E)\geq\frac{1}{m_{0}}. For the α\alpha-invariant, we know by Lemma 2.21 that there exists some positive integer N1=N1(n,ε,A,I)N_{1}=N_{1}(n,\varepsilon,A,I) divisible by m0m_{0} such that N1EN_{1}E and N1(KZ+ΔZ+E)N_{1}(K_{Z}+\Delta_{Z}+E) are Cartier. By adjunction, we see that N1(KE+ΔE)N_{1}(K_{E}+\Delta_{E}) is also Cartier. By [HMX-ACC, Corollary 1.8], the log Fano pair (E,ΔE)(E,\Delta_{E}) belongs to a bounded family (which relies only on N1N_{1}). This implies that α(E,ΔE)\alpha(E,\Delta_{E}) is uniformly bounded below. Finally, to give an upper bound for dEd_{E}, let L=E|EL=-E|_{E} be the (ample) \mathbb{Q}-divisor defined as in [HLS-epsilon-plt-blowup, Definition A.4] and let

𝔟k:=𝔞k(ordE)=f𝒪Z(kE)\mathfrak{b}_{k}:=\mathfrak{a}_{k}(\mathrm{ord}_{E})=f_{*}\mathcal{O}_{Z}(-kE)

where k=1,2,k=1,2,\dots. Then N1LN_{1}L is Cartier as N1EN_{1}E is Cartier. Since (KE+ΔE)AX,Δ(E)L-(K_{E}+\Delta_{E})\sim_{\mathbb{Q}}A_{X,\Delta}(E)\cdot L and AX,Δ(E)1m0A_{X,\Delta}(E)\geq\frac{1}{m_{0}}, we see that the coefficient and degree of LL are bounded, hence the triple (E,Δ,L)(E,\Delta,L) belongs to a bounded family. Thus there exists an integer N=N(n,ε,A,I)>0N=N(n,\varepsilon,A,I)>0 such that the section ring kH0(E,kL)\oplus_{k\in\mathbb{N}}H^{0}(E,\lfloor kL\rfloor) is generated in degree N\leq N. We claim that

(3.6) 𝔟N+1𝔪x2.\mathfrak{b}_{N+1}\subseteq\mathfrak{m}_{x}^{2}.

Taking this for granted, it follows immediately that dENd_{E}\leq N. Putting these information together we deduce from (3.5) that (3.4) is bounded below by some positive constant that only rely on n,ε,A,In,\varepsilon,A,I. From the discussion right above (3.4) we also know that (3.4) is bounded from above by 1m1\frac{1}{m-1} where mm is the largest integer such that 𝔞(m1)a𝔪x2\mathfrak{a}_{(m-1)a}\not\subseteq\mathfrak{m}_{x}^{2}. Thus we see that there is some fixed mm depending only on n,ε,A,In,\varepsilon,A,I such that mHmH is Cartier and (3.3) holds. Combined with (3.2), we deduce that

(3.7) H0(X¯,𝒪X¯(mH))H0(X¯,𝒪X¯/𝔪x2)H^{0}(\overline{X},\mathcal{O}_{\overline{X}}(mH))\to H^{0}(\overline{X},\mathcal{O}_{\overline{X}}/\mathfrak{m}_{x}^{2})

is surjective, i.e. |mH||mH| separates tangent direction at xx.

Before we proceed to show that |mH||mH| also defines a birational map, let us finish the proof of the claim (3.6). In fact, we shall prove by descending induction that 𝔟k𝔪x2\mathfrak{b}_{k}\subseteq\mathfrak{m}_{x}^{2} for all kN+1k\geq N+1. This is clear when k0k\gg 0. Suppose that 𝔟k+1𝔪x2\mathfrak{b}_{k+1}\subseteq\mathfrak{m}_{x}^{2} and kN+1k\geq N+1, then since 𝔟k+1𝔟k\mathfrak{b}_{k+1}\subseteq\mathfrak{b}_{k} for all kk and since k𝔟k/𝔟k+1kH0(E,kL)\oplus_{k\in\mathbb{N}}\mathfrak{b}_{k}/\mathfrak{b}_{k+1}\cong\oplus_{k\in\mathbb{N}}H^{0}(E,\lfloor kL\rfloor) (see [LX-stability-kc, Section 2.4] or [LZ-Tian-sharp, Proposition 2.10]) is generated in degree N\leq N, we see that

𝔟k/𝔟k+1i=1N(𝔟i/𝔟i+1)(𝔟ki/𝔟ki+1)𝔟12/𝔟k+1;\mathfrak{b}_{k}/\mathfrak{b}_{k+1}\subseteq\sum_{i=1}^{N}(\mathfrak{b}_{i}/\mathfrak{b}_{i+1})\cdot(\mathfrak{b}_{k-i}/\mathfrak{b}_{k-i+1})\subseteq\mathfrak{b}_{1}^{2}/\mathfrak{b}_{k+1};

in other words, 𝔟k𝔟k+1+𝔟12\mathfrak{b}_{k}\subseteq\mathfrak{b}_{k+1}+\mathfrak{b}_{1}^{2}. As 𝔟k+1𝔪x2\mathfrak{b}_{k+1}\subseteq\mathfrak{m}_{x}^{2} by induction hypothesis and clearly 𝔟1𝔪x\mathfrak{b}_{1}\subseteq\mathfrak{m}_{x}, we obtain 𝔟k𝔪x2\mathfrak{b}_{k}\subseteq\mathfrak{m}_{x}^{2} as desired.

We are finally in a position to show that |mH||mH| defines a birational map that restricts to an embedding on XX. By (3.7), we already know that in a neighbourhood of xx, the linear system |mH||mH| is base point free and the induced map is unramified. Since the base locus of |mH||mH| is closed and invariant under the 𝔾m\mathbb{G}_{m}-action (coming from the orbifold cone structure), we see that |mH||mH| has no base point in XX. Similarly, as the ramification locus of the induced map φ\varphi on XX is a closed 𝔾m\mathbb{G}_{m}-invariant subset, we see that φ\varphi is unramified on XX and thus it is quasi-finite. This implies that φ(x)φ(x)\varphi(x^{\prime})\neq\varphi(x) for all xxXx^{\prime}\neq x\in X, otherwise by 𝔾m\mathbb{G}_{m}-translation we deduce that φ\varphi contracts the closed orbit 𝔾mx¯\overline{\mathbb{G}_{m}\cdot x^{\prime}}. In particular, φ|X1(φ(x))\varphi|_{X}^{-1}(\varphi(x)) is supported at xx; as φ\varphi is also unramified, the scheme-theoretic pre-image φ|X1(φ(x))\varphi|_{X}^{-1}(\varphi(x)) equals {x}\{x\}. By upper semi-continuity and the 𝔾m\mathbb{G}_{m}-action, this implies that φ|X1(φ(x))\varphi|_{X}^{-1}(\varphi(x^{\prime})) has length 11 and thus consists of a single point for all xXx^{\prime}\in X. It follows that φ|X\varphi|_{X} is an embedding on XX and we finish the proof. ∎

We have used the following vanishing result in the above proof. This should be well-known to experts, but we are unable to find a suitable reference.

Lemma 3.9.

Let (X,Δ)(X,\Delta) be a pair such that KX+ΔK_{X}+\Delta is \mathbb{Q}-Cartier. Let LL be a \mathbb{Q}-Cartier Weil divisor such that L(KX+Δ)L-(K_{X}+\Delta) is nef and big. Assume that (X,Δ)(X,\Delta) is klt along the non-Cartier locus of LL. Then

Hi(X,𝒥(X,Δ)𝒪X(L))=0H^{i}(X,\mathcal{J}(X,\Delta)\otimes\mathcal{O}_{X}(L))=0

for all i>0i>0.

Proof.

If LL is a line bundle this is just the usual Nadel vanishing; in the general case we follow the proof of Nadel vanishing. Let f:YXf\colon Y\to X be a log resolution. We may write

KY+ΔY\displaystyle K_{Y}+\Delta_{Y} =f(KX+Δ)+aiEi,\displaystyle=f^{*}(K_{X}+\Delta)+\sum a_{i}E_{i},
fL\displaystyle f^{*}L =fL+biEi,\displaystyle=\lfloor f^{*}L\rfloor+\sum b_{i}E_{i},

where the EiE_{i}’s are the exceptional divisors, and 0bi<10\leq b_{i}<1. Let

L\displaystyle L^{\prime} =fLΔY+ai+biEi,\displaystyle=\lfloor f^{*}L\rfloor-\lfloor\Delta_{Y}\rfloor+\sum\lceil a_{i}+b_{i}\rceil E_{i},
D\displaystyle D ={ΔY}+{aibi}Ei.\displaystyle=\{\Delta_{Y}\}+\sum\{-a_{i}-b_{i}\}E_{i}.

Then it’s easy to check that (Y,D)(Y,D) is klt (i.e. {D}=0\{D\}=0 since YY is a log resolution) and L(KY+D)f(LKXΔ)L^{\prime}-(K_{Y}+D)\sim_{\mathbb{Q}}f^{*}(L-K_{X}-\Delta) is nef and big. By Kawamata-Viehweg vanishing we have Hi(Y,𝒪Y(L))=0H^{i}(Y,\mathcal{O}_{Y}(L^{\prime}))=0 and Rif𝒪Y(L)=0R^{i}f_{*}\mathcal{O}_{Y}(L^{\prime})=0 for all i>0i>0. It follows that Rf𝒪Y(L)=f𝒪Y(L)Rf_{*}\mathcal{O}_{Y}(L^{\prime})=f_{*}\mathcal{O}_{Y}(L^{\prime}) and hence

Hi(X,f𝒪Y(L))=Hi(X,Rf𝒪Y(L))=Hi(Y,𝒪Y(L))=0.H^{i}(X,f_{*}\mathcal{O}_{Y}(L^{\prime}))=H^{i}(X,Rf_{*}\mathcal{O}_{Y}(L^{\prime}))=H^{i}(Y,\mathcal{O}_{Y}(L^{\prime}))=0.

It remains to show that

(3.8) f𝒪Y(L)=𝒥(X,Δ)𝒪X(L).f_{*}\mathcal{O}_{Y}(L^{\prime})=\mathcal{J}(X,\Delta)\otimes\mathcal{O}_{X}(L).

This is a local question, so it suffices to check the equality at any xXx\in X. If (X,Δ)(X,\Delta) is klt at xx, then locally 𝒥(X,Δ)=𝒪X\mathcal{J}(X,\Delta)=\mathcal{O}_{X}, ΔY=0\lfloor\Delta_{Y}\rfloor=0, and ai>1a_{i}>-1 which implies ai+bi0\lceil a_{i}+b_{i}\rceil\geq 0. It follows that

f𝒪Y(L)=𝒪X(L)=𝒥(X,Δ)𝒪X(L).f_{*}\mathcal{O}_{Y}(L^{\prime})=\mathcal{O}_{X}(L)=\mathcal{J}(X,\Delta)\otimes\mathcal{O}_{X}(L).

If (X,Δ)(X,\Delta) is not klt at xx, then LL is Cartier around xx, which gives bi=0b_{i}=0 for every EiE_{i} whose image contains xx, thus locally

f𝒪Y(L)=f𝒪Y(fL+ΔY+aiEi)=𝒥(X,Δ)𝒪X(L)f_{*}\mathcal{O}_{Y}(L^{\prime})=f_{*}\mathcal{O}_{Y}(f^{*}L+\lceil-\Delta_{Y}+\sum a_{i}E_{i}\rceil)=\mathcal{J}(X,\Delta)\otimes\mathcal{O}_{X}(L)

by projection formula and the definition of the multiplier ideal. Thus we see that (3.8) always holds. This completes the proof. ∎

3.3. Proof of Theorem 3.1

We are now in a position to prove Theorem 3.1.

Proof.

First assume that I[0,1]I\subseteq[0,1]\cap\mathbb{Q}. Let x(X,Δ;ξ)x\in(X,\Delta;\xi) be a polarized log Fano cone singularity in 𝒮\mathcal{S} and let 𝕋\mathbb{T} be the torus generated by ξ\xi. By Lemma 2.11, there exists some quasi-regular polarization ξ\xi^{\prime} that is sufficiently close to ξ\xi in the Reeb cone such that Θ(X,Δ;ξ)θ2\Theta(X,\Delta;\xi^{\prime})\geq\frac{\theta}{2}. Using ξ\xi^{\prime}, we may realize (X,Δ)(X,\Delta) as an orbifold cone over some polarized pair (V,ΔV;L)(V,\Delta_{V};L). Moreover, if V0V_{0} is the exceptional divisor of the orbifold vertex blowup as in Section 3.1, then ordV0\mathrm{ord}_{V_{0}} is proportional to wtξ\mathrm{wt}_{\xi^{\prime}}. By assumption and Lemma 2.6, we obtain

(3.9) vol^X,Δ(ordV0)=vol^X,Δ(wtξ)2θ1vol^(x,X,Δ)2θ1nn.\widehat{\rm vol}_{X,\Delta}(\mathrm{ord}_{V_{0}})=\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi^{\prime}})\leq 2\theta^{-1}\widehat{\rm vol}(x,X,\Delta)\leq 2\theta^{-1}n^{n}.

Let (X¯,Δ¯)(\overline{X},\overline{\Delta}) be the projective orbifold cone over (V,ΔV;L)(V,\Delta_{V};L) and VV_{\infty} the divisor at infinity. By Proposition 3.8, there exists some integer mm depending only on n,ε,A,In,\varepsilon,A,I such that |m(KX¯+Δ¯+V)||-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})| defines a birational map φ\varphi that restricts to an embedding on XX. By (3.9) and Lemma 3.4, we see that

vol(m(KX¯+Δ¯+V))=mnvol^X,Δ(ordV0)2θ1(mn)n\mathrm{vol}(-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty}))=m^{n}\cdot\widehat{\rm vol}_{X,\Delta}(\mathrm{ord}_{V_{0}})\leq 2\theta^{-1}(mn)^{n}

is bounded above.

Let us show that vol(m(KX¯+Δ¯+V)|Δ¯)\mathrm{vol}(-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})|_{\overline{\Delta}}) is also bounded from above, so that the pair (X¯,Δ¯)(\overline{X},\overline{\Delta}) belongs to a birationally bounded family. By Lemma 2.20, we know that x(X,(1+c)Δ)x\in(X,(1+c)\Delta) is of klt type for some constant c=c(n,ε)>0c=c(n,\varepsilon)>0. By Lemma 3.5, this implies that (KV+(1+c)ΔV+ΔL)-(K_{V}+(1+c)\Delta_{V}+\Delta_{L}) is big. By a similar calculation as in Lemma 3.4, it follows that

vol((KX¯+Δ¯+V)|Δ¯)\displaystyle\mathrm{vol}(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})|_{\overline{\Delta}}) =((KX¯+Δ¯+V))n2rVΔ¯\displaystyle=\left(-(K_{\overline{X}}+\overline{\Delta}+V_{\infty})\right)^{n-2}\cdot rV_{\infty}\cdot\overline{\Delta}
=r((KV+ΔV+ΔL))n2ΔV\displaystyle=r\cdot\left(-(K_{V}+\Delta_{V}+\Delta_{L})\right)^{n-2}\cdot\Delta_{V}
c1r((KV+ΔV+ΔL))n1\displaystyle\leq c^{-1}r\cdot\left(-(K_{V}+\Delta_{V}+\Delta_{L})\right)^{n-1}
=c1vol^X,Δ(ordV0)2(cθ)1nn,\displaystyle=c^{-1}\widehat{\rm vol}_{X,\Delta}(\mathrm{ord}_{V_{0}})\leq 2(c\theta)^{-1}n^{n},

where the first equality is by Lemma 3.3, the second by adjunction along VVV_{\infty}\cong V, the next inequality by the bigness of (KV+(1+c)ΔV+ΔL)-(K_{V}+(1+c)\Delta_{V}+\Delta_{L}), and the last equality by Lemma 3.4. Thus vol(m(KX¯+Δ¯+V)|Δ¯)\mathrm{vol}(-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})|_{\overline{\Delta}}) is also bounded from above as desired. Note that φ|Supp(Δ¯)\varphi|_{\mathrm{Supp}(\overline{\Delta})} is also birational since φ\varphi is an embedding at xx. Therefore, the image (W,ΔW)(W,\Delta_{W}) of (X¯,Δ¯)(\overline{X},\overline{\Delta}) under the birational map induced by |m(KX¯+Δ¯+V)||-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})| belongs to a fixed bounded family (𝒲,𝒟¯)B(\mathcal{W},\overline{\mathcal{D}})\to B.

By construction, (X¯,Δ¯)(\overline{X},\overline{\Delta}) carries an effective 𝕋\mathbb{T}-action and VV_{\infty} is 𝕋\mathbb{T}-invariant. It follows that |m(KX¯+Δ¯+V)||-m(K_{\overline{X}}+\overline{\Delta}+V_{\infty})| is a 𝕋\mathbb{T}-invariant linear system and the induced birational map φ\varphi is 𝕋\mathbb{T}-equivariant. In particular, the image (W,ΔW)(W,\Delta_{W}) also carries an effective 𝕋\mathbb{T}-action. We claim that there exists finitely many morphisms BiBB_{i}\to B (depending only on the family (𝒲,𝒟¯)B(\mathcal{W},\overline{\mathcal{D}})\to B) such that

  1. (1)

    after base change, each family (𝒲i,𝒟¯i)=(𝒲,𝒟¯)×BBiBi(\mathcal{W}_{i},\overline{\mathcal{D}}_{i})=(\mathcal{W},\overline{\mathcal{D}})\times_{B}B_{i}\to B_{i} admits an effective fiberwise 𝕋i\mathbb{T}_{i}-action for some torus 𝕋i\mathbb{T}_{i},

  2. (2)

    (W,ΔW)(𝒲bi,𝒟¯bi)(W,\Delta_{W})\cong(\mathcal{W}_{b_{i}},\overline{\mathcal{D}}_{b_{i}}) for some biBib_{i}\in B_{i}, and under this isomorphism, the 𝕋\mathbb{T}-action on WW is induced by some group homomorphism 𝕋𝕋i\mathbb{T}\to\mathbb{T}_{i} (in other words, the 𝕋\mathbb{T}-action on WW is induced by the 𝕋i\mathbb{T}_{i}-action on 𝒲bi\mathcal{W}_{b_{i}}).

To see this, first observe that for any torus action on a projective variety, the induced action on the Picard scheme is trivial. This is because both Pic0\mathrm{Pic}^{0} (an abelian variety) and Pic/Pic0\mathrm{Pic}/\mathrm{Pic}^{0} (a discrete group) have no 𝔾m\mathbb{G}_{m}-action. Thus if \mathcal{L} is a relatively ample line bundle on 𝒲\mathcal{W}, then it is invariant under any fiberwise torus action. Consider the relative automorphism group scheme 𝔾\mathbb{G} over BB, which parametrizes the automorphisms of the polarized fibers (𝒲b,𝒟¯b;b)(\mathcal{W}_{b},\overline{\mathcal{D}}_{b};\mathcal{L}_{b}). Note that 𝔾\mathbb{G} is affine over BB. Possibly after stratifying the base BB, we may also assume that 𝔾\mathbb{G} is smooth over BB. By [SGA3-II, Exposé XII, Théorèm 1.7(a)], after a further stratification of BB, we may assume that the dimension of the maximal torus of 𝔾b\mathbb{G}_{b} is locally constant in bBb\in B. We may discard the components of BB where this torus dimension is zero. By [SGA3-II, Exposé XII, Théorèm 1.7(b)], it then follows that there exists an (étale) cover BiB\cup B_{i}\to B of the remaining components of BB and a subgroup scheme 𝔾i𝔾×BBi\mathbb{G}_{i}\subseteq\mathbb{G}\times_{B}B_{i} that reduces to the maximal tori on the fibers. Passing to a further finite cover of the BiB_{i}’s, we may assume that the 𝔾i\mathbb{G}_{i}’s are split, i.e. 𝔾i=𝕋i×Bi\mathbb{G}_{i}=\mathbb{T}_{i}\times B_{i} for some torus 𝕋i\mathbb{T}_{i}. In particular, the 𝔾\mathbb{G}-action on (𝒲,𝒟¯)(\mathcal{W},\overline{\mathcal{D}}) induces a fiberwise 𝕋i\mathbb{T}_{i}-action on (𝒲i,𝒟¯i):=(𝒲,𝒟¯)×BBi(\mathcal{W}_{i},\overline{\mathcal{D}}_{i}):=(\mathcal{W},\overline{\mathcal{D}})\times_{B}B_{i}. This gives the family in (1). Since (W,ΔW)(W,\Delta_{W}) has a non-trivial torus action, we see that (W,ΔW)(W,\Delta_{W}) appears as a fiber of (𝒲i,𝒟¯i)Bi(\mathcal{W}_{i},\overline{\mathcal{D}}_{i})\to B_{i} for some ii. Since all maximal tori in Aut(W,ΔW)\mathrm{Aut}(W,\Delta_{W}) are conjugate to each other, we see that 𝕋\mathbb{T} is conjugate to a subtorus of the maximal torus 𝕋i\mathbb{T}_{i}. In other words, there exists an isomorphism (W,ΔW)(𝒲bi,𝒟¯bi)(W,\Delta_{W})\cong(\mathcal{W}_{b_{i}},\overline{\mathcal{D}}_{b_{i}}) such that (2) holds. This proves the claim.

Taking the fiberwise isolated 𝕋i\mathbb{T}_{i}-fixed points (𝒲i,𝒟¯i)Bi(\mathcal{W}_{i},\overline{\mathcal{D}}_{i})\to B_{i} for all ii, we get families Bi(𝒳i,𝒟i)BiB_{i}\subseteq(\mathcal{X}_{i},\mathcal{D}_{i})\to B_{i} of singularities (possibly after a refinement of the BiB_{i}’s), each with an effective fiberwise torus 𝕋i\mathbb{T}_{i}-action. Since φ\varphi restricts to an embedding on XX, by the second part of the above claim we see that x(X,Δ;ξ)x\in(X,\Delta;\xi) is isomorphic to bi(𝒲bi,𝒟bi;ξbi)b_{i}\in(\mathcal{W}_{b_{i}},\mathcal{D}_{b_{i}};\xi_{b_{i}}) for some biBib_{i}\in B_{i} and some ξbiN(𝕋i)\xi_{b_{i}}\in N(\mathbb{T}_{i})_{\mathbb{R}}. In particular, this gives log boundedness. Note that when II\not\subseteq\mathbb{Q}, we can still apply the above argument to the singularities and the coefficient set constructed from Lemma 2.18, thus the same conclusion holds.

To get boundedness, we need to further stratify the family (𝒲i,𝒟¯i)Bi(\mathcal{W}_{i},\overline{\mathcal{D}}_{i})\to B_{i} so that it becomes \mathbb{R}-Gorenstein klt. By [Kol-moduli-book, Lemma 4.44] and inversion of adjunction, there exists a finite collection of locally closed subset SjS_{j} of iBi\cup_{i}B_{i} such that the family becomes \mathbb{R}-Gorenstein after base change to jSj\cup_{j}S_{j} and enumerates exactly all the klt fibers of i(𝒲i,𝒟¯i)iBi\cup_{i}(\mathcal{W}_{i},\overline{\mathcal{D}}_{i})\to\cup_{i}B_{i} (note that [Kol-moduli-book, Lemma 4.44] requires the family to be proper but the proof applies to our situation, essentially because the section Bi𝒳iB_{i}\subseteq\mathcal{X}_{i} is proper over the base). Replacing the BiB_{i}’s by the SjS_{j}’s, we obtain the desired family. The proof is now complete. ∎

3.4. Applications

We now explain how to deduce the other main results of this paper from Theorem 3.1. First we prove the boundedness criterion for K-semistable log Fano cone singularities.

Corollary 3.10.

Let 𝒮\mathcal{S} be a set of nn-dimensional K-semistable log Fano cone singularities with coefficients in a fixed finite set I[0,1]I\subseteq[0,1]. Then 𝒮\mathcal{S} is bounded if and only if there exist some ε,A>0\varepsilon,A>0 such that

vol^(x,X,Δ)ε𝑎𝑛𝑑mldK(x,X,Δ)A\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon\quad\mathit{and}\quad\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq A

for all x(X,Δ)x\in(X,\Delta) in 𝒮\mathcal{S}.

Proof.

One direction follows from Lemma 2.15, while the other direction is implied by Theorem 3.1 as the volume ratio of a K-semistable log Fano cone singularity (with the K-semistable polarization) is equal to 11. ∎

Similarly, we have an unpolarized version of Theorem 3.1.

Corollary 3.11.

Let 𝒮\mathcal{S} be a set of nn-dimensional log Fano cone singularities with coefficients in a fixed finite set I[0,1]I\subseteq[0,1]. Then 𝒮\mathcal{S} is bounded if and only if there exist positive constants ε,θ,A>0\varepsilon,\theta,A>0 such that

vol^(x,X,Δ)ε,Θ(x,X,Δ)θ,andmldK(x,X,Δ)A.\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon,\quad\Theta(x,X,\Delta)\geq\theta,\quad\mathrm{and}\quad\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq A.

for all x(X,Δ)x\in(X,\Delta) in 𝒮\mathcal{S}.

Proof.

One direction follows from Lemmas 2.15 and 2.17, while the other direction is immediate by Theorem 3.1. ∎

We next prove a version of Theorem 3.1 that replaces the volume ratios with the stability thresholds introduced in [Hua-thesis]. First we recall the definition. Let x(X=𝐒𝐩𝐞𝐜(R),Δ;ξ)x\in(X=\mathbf{Spec}(R),\Delta;\xi) be a polarized log Fano cone singularity and let 𝕋\mathbb{T} be the torus generated by ξ\xi. For this definition it would be more convenient to rescale the polarization so that AX,Δ(wtξ)=1A_{X,\Delta}(\mathrm{wt}_{\xi})=1, which we will assume in what follows. Using the weight decomposition R=αRαR=\oplus_{\alpha}R_{\alpha}, we set

Rm:=α,m1<α,ξmRα.R_{m}:=\bigoplus_{\alpha,\,m-1<\langle\alpha,\xi\rangle\leq m}R_{\alpha}.

An mm-basis type \mathbb{Q}-divisor of x(X=𝐒𝐩𝐞𝐜(R),Δ;ξ)x\in(X=\mathbf{Spec}(R),\Delta;\xi) is defined to be a \mathbb{Q}-divisor of the form

D=1mNmi=1Nm{si=0}D=\frac{1}{mN_{m}}\sum_{i=1}^{N_{m}}\{s_{i}=0\}

where Nm=dimRmN_{m}=\dim R_{m} and s1,,sNms_{1},\dots,s_{N_{m}} form a basis of RmR_{m}. Set δm=infDlctx(X,Δ;D)\delta_{m}=\inf_{D}\mathrm{lct}_{x}(X,\Delta;D) where the infimum runs over all mm-basis type \mathbb{Q}-divisors DD. The stability threshold of x(X=𝐒𝐩𝐞𝐜(R),Δ;ξ)x\in(X=\mathbf{Spec}(R),\Delta;\xi) is defined as

δ(X,Δ;ξ):=limmδm.\delta(X,\Delta;\xi):=\lim_{m\to\infty}\delta_{m}.

If (X,Δ;ξ)(X,\Delta;\xi) is the cone over a log Fano pair (V,ΔV)(V,\Delta_{V}), then this definition is closely related to the stability threshold of (V,ΔV)(V,\Delta_{V}) introduced in [FO-delta] (see also [BJ-delta]). In fact, using inversion of adjunction it is not hard to show that δ(X,Δ;ξ)=min{1,δ(V,ΔV)}\delta(X,\Delta;\xi)=\min\{1,\delta(V,\Delta_{V})\}, cf. [XZ-minimizer-unique, Theorem 3.6].

The stability threshold version of Theorem 3.1 is a direct consequence of the following result.

Lemma 3.12.

Let x(X=𝐒𝐩𝐞𝐜(R),Δ;ξ)x\in(X=\mathbf{Spec}(R),\Delta;\xi) be a polarized log Fano cone singularity of dimension nn. Then

Θ(X,Δ;ξ)δ(X,Δ;ξ)n.\Theta(X,\Delta;\xi)\geq\delta(X,\Delta;\xi)^{n}.
Proof.

Let δ:=δ(X,Δ;ξ)\delta:=\delta(X,\Delta;\xi). We first recall the valuative interpretation of δ(X,Δ;ξ)\delta(X,\Delta;\xi) as explained in [Hua-thesis]. Fix some 𝕋\mathbb{T}-invariant valuation vValX,xv\in\mathrm{Val}_{X,x}^{*}. Set Sm(v):=supDv(D)S_{m}(v):=\sup_{D}v(D) where DD varies among mm-basis type \mathbb{Q}-divisors. By an Okounkov body argument [Hua-thesis, Section 4], we know that S(v):=limmSm(v)S(v):=\lim_{m\to\infty}S_{m}(v) exists and

(3.10) AX,Δ(v)δS(v)A_{X,\Delta}(v)\geq\delta\cdot S(v)

by [Hua-thesis, Theorem 4.3.5]. We next relate S(v)S(v) to the “relative” SS-invariant S(wtξ;v)S(\mathrm{wt}_{\xi};v) defined in [XZ-minimizer-unique, Section 3.1]. In our notation (and under our assumption that AX,Δ(wtξ)=1)A_{X,\Delta}(\mathrm{wt}_{\xi})=1), we have S(wtξ;v)=limmSm(wtξ;v)S(\mathrm{wt}_{\xi};v)=\lim_{m\to\infty}S_{m}(\mathrm{wt}_{\xi};v) where

Sm(wtξ;v):=k=0mkNkSk(v)k=0mkNk.S_{m}(\mathrm{wt}_{\xi};v):=\frac{\sum_{k=0}^{m}kN_{k}S_{k}(v)}{\sum_{k=0}^{m}kN_{k}}.

Intuitively, the previous Sm(v)S_{m}(v) is defined using basis type divisors for RmR_{m} while Sm(wtξ;v)S_{m}(\mathrm{wt}_{\xi};v) is defined via basis type divisors for kmRk\oplus_{k\leq m}R_{k} (the main reason for doing so in [XZ-minimizer-unique] is that in the more general situation, the dimension of the analogous space kmRk\oplus_{k\leq m}R_{k} has an asymptotic expression, while the individual RmR_{m} does not). Note that our Sm(wtξ;v)S_{m}(\mathrm{wt}_{\xi};v) differs slightly from the one in [XZ-minimizer-unique, Section 3.1] by some round-downs, but after taking the mm\to\infty limit we get the same value of S(wtξ;v)S(\mathrm{wt}_{\xi};v). Since Sm(v)S(v)S_{m}(v)\to S(v) as mm\to\infty, from the above expression we get S(v)=S(wtξ;v)S(v)=S(\mathrm{wt}_{\xi};v). From the proof of [XZ-minimizer-unique, Theorem 3.7] (especially the inequality after (3.8) in loc. cit.), we then obtain

S(v)(vol(wtξ)vol(v))1n.S(v)\geq\left(\frac{\mathrm{vol}(\mathrm{wt}_{\xi})}{\mathrm{vol}(v)}\right)^{\frac{1}{n}}.

Combined with (3.10) and recall that AX,Δ(wtξ)=1A_{X,\Delta}(\mathrm{wt}_{\xi})=1, we deduce

vol^X,Δ(v)=AX,Δ(v)nvol(v)δnS(v)nvol(v)δnvol(wtξ)=δnvol^X,Δ(wtξ)\widehat{\rm vol}_{X,\Delta}(v)=A_{X,\Delta}(v)^{n}\cdot\mathrm{vol}(v)\geq\delta^{n}S(v)^{n}\cdot\mathrm{vol}(v)\geq\delta^{n}\mathrm{vol}(\mathrm{wt}_{\xi})=\delta^{n}\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi})

for all 𝕋\mathbb{T}-invariant valuation vValX,xv\in\mathrm{Val}_{X,x}^{*}. Since the local volume of x(X,Δ)x\in(X,\Delta) is computed by some 𝕋\mathbb{T}-invariant valuation ([Blu-minimizer-exist] and [XZ-minimizer-unique, Corollary 1.2]), this implies that vol^(x,X,Δ)δnvol^X,Δ(wtξ)\widehat{\rm vol}(x,X,\Delta)\geq\delta^{n}\widehat{\rm vol}_{X,\Delta}(\mathrm{wt}_{\xi}) and hence Θ(X,Δ;ξ)δn\Theta(X,\Delta;\xi)\geq\delta^{n}. ∎

Remark 3.13.

We also have an inequality in the reverse direction, namely, there exists some positive constant c>0c>0 that only depends on the dimension such that

(3.11) cδ(X,Δ;ξ)Θ(X,Δ;ξ).c\cdot\delta(X,\Delta;\xi)\geq\Theta(X,\Delta;\xi).

To see this, let c=c01c=c_{0}^{-1} where c0c_{0} is the constant from [Z-mld^K-1, Lemma 3.4] and let DD be any mm-basis type \mathbb{Q}-divisor of (X,Δ;ξ)(X,\Delta;\xi). Then by definition we have wtξ(D)1=AX,Δ(wtξ)\mathrm{wt}_{\xi}(D)\leq 1=A_{X,\Delta}(\mathrm{wt}_{\xi}), hence by loc. cit. we see that

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

which gives (3.11). Note that the upper bound on the volume ratio is at least linear in the stability threshold, as can be seen on the smooth toric singularities (X,Δ)=(𝔸n,0)(X,\Delta)=(\mathbb{A}^{n},0): if ξ=(a1,,an)𝔱+=>0n\xi=(a_{1},\dots,a_{n})\in\mathfrak{t}_{\mathbb{R}}^{+}=\mathbb{R}^{n}_{>0} and a1a2==ana_{1}\ll a_{2}=\dots=a_{n}, then we have that Θ(X,Δ;ξ)=nna1an(a1++an)n\Theta(X,\Delta;\xi)=\frac{n^{n}a_{1}\cdots a_{n}}{(a_{1}+\dots+a_{n})^{n}} is roughly linear in δ(X,Δ;ξ)=na1a1++an\delta(X,\Delta;\xi)=\frac{na_{1}}{a_{1}+\dots+a_{n}}.

Corollary 3.14.

Let nn\in\mathbb{N} and let I[0,1]I\subseteq[0,1] be a finite set. Let ε,δ,A>0\varepsilon,\delta,A>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,δ(X,Δ;ξ)δ,andmldK(x,X,Δ)A.\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon,\quad\delta(X,\Delta;\xi)\geq\delta,\quad\mathrm{and}\quad\mathrm{mld}^{\mathrm{K}}(x,X,\Delta)\leq A.

Then 𝒮\mathcal{S} is bounded.

Proof.

This is a direct consequence of Theorem 3.1 and Lemma 3.12. ∎

Finally we specialize our results to dimension three. For this we need the following result from [Z-mld^K-1].

Proposition 3.15.

Let ε>0\varepsilon>0 and let I[0,1]I\subseteq[0,1] be a finite set. Then there exists some constant A=A(ε,I)A=A(\varepsilon,I) such that

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

for all 33-dimensional klt singularity x(X,Δ)x\in(X,\Delta) with coefficients in II and vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon.

Proof.

If II\subseteq\mathbb{Q} this is [Z-mld^K-1, Corollary 6.11]; in general we apply Lemma 2.18 to reduce to the rational coefficient case. ∎

Corollary 3.16.

For any finite set I[0,1]I\subseteq[0,1] and any ε>0\varepsilon>0, the set of 33-dimensional K-semistable log Fano cone singularities with coefficients in II and with local volume at least ε\varepsilon is bounded.

Proof.

Immediate from Corollary 3.10 and Proposition 3.15. ∎

The last application concerns the distribution of local volumes in dimension 33. For any nn\in\mathbb{N} and I[0,1]I\subseteq[0,1], consider the set Voln,Iloc\mathrm{Vol}^{\mathrm{loc}}_{n,I} of all possible local volumes of nn-dimensional klt singularities x(X,Δ)x\in(X,\Delta) with coefficients in II.

Corollary 3.17.

Let I[0,1]I\subseteq[0,1] be a finite set. Then Vol3,Iloc\mathrm{Vol}^{\mathrm{loc}}_{3,I} is discrete away from zero.

Proof.

By Theorem 2.19, it suffices to consider local volumes of K-semistable log Fano cone singularities. Let ε>0\varepsilon>0. By Corollary 3.16, the set of 33-dimensional K-semistable log Fano cone singularities x(X,Δ)x\in(X,\Delta) with coefficients in II and vol^(x,X,Δ)ε\widehat{\rm vol}(x,X,\Delta)\geq\varepsilon is bounded. On the other hand, the local volume function is constructible in \mathbb{R}-Gorenstein families by [Xu-quasi-monomial, Theorem 1.3] (see [HLQ-vol-ACC, Theorem 3.5] for the real coefficient case). In particular, the local volumes only take finitely many possible values in a bounded family. This implies Vol3,Iloc[ε,+)\mathrm{Vol}^{\mathrm{loc}}_{3,I}\cap[\varepsilon,+\infty) is a finite set and we are done. ∎

Remark 3.18.

By combining the ideas in this work with some generalization of the proof of [HLQ-vol-ACC, Theorem 1.2(2)], one should be able to show that if the set II in Corollary 3.17 is not finite but satisfies DCC (descending chain condition), then Vol3,Iloc\mathrm{Vol}^{\mathrm{loc}}_{3,I} satisfies ACC (ascending chain condition). We leave the details to the interested readers.

References