Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-25T08:11:12.336Z Has data issue: false hasContentIssue false

Existence of valuations with smallest normalized volume

Published online by Cambridge University Press:  08 March 2018

Harold Blum*
Affiliation:
Department of Mathematics, University of Michigan, Ann Arbor, MI 48109-1043, USA email blum@umich.eduhttp://www-personal.umich.edu/∼blum/
Rights & Permissions [Opens in a new window]

Abstract

Li introduced the normalized volume of a valuation due to its relation to K-semistability. He conjectured that over a Kawamata log terminal (klt) singularity there exists a valuation with smallest normalized volume. We prove this conjecture and give an explicit example to show that such a valuation need not be divisorial.

MSC classification

Type
Research Article
Copyright
© The Author 2018 

1 Introduction

Fix a variety $X$ of dimension $n$ and $x\in X$ a closed point. Let $\operatorname{Val}_{X,x}$ denote the set of real valuations on $X$ with center equal to $x$ . An element of $\operatorname{Val}_{X,x}$ is an $\mathbf{R}$ -valued valuation of the function field $K(X)$ that takes nonnegative values on ${\mathcal{O}}_{X,x}\subseteq K(X)$ and strictly positive values on the maximal ideal of ${\mathcal{O}}_{X,x}$ . For examples, divisorial valuations centered at $x$ form an important class inside $\operatorname{Val}_{X,x}$ . These valuations are determined by the order of vanishing along a prime divisor $E\subset Y$ where $Y$ is normal and there is a proper birational morphism $f:Y\rightarrow X$ contracting $E$ to  $x$ . We denote such a valuation by $\operatorname{ord}_{E}\in \operatorname{Val}_{X,x}$ .

Li introduced the normalized volume function

$$\begin{eqnarray}\widehat{\operatorname{vol}}_{X,x}:\operatorname{Val}_{X,x}\longrightarrow \mathbf{R}_{{>}0}\cup \{+\infty \}\end{eqnarray}$$

that sends a valuation $v$ to its normalized volume, denoted $\widehat{\operatorname{vol}}(v)$ [Reference LiLi15]. To define the normalized volume, we recall the following. Given a valuation $v\in \operatorname{Val}_{X,x}$ , we have valuation ideals

$$\begin{eqnarray}\mathfrak{a}_{m}(v)_{x}:=\{f\in {\mathcal{O}}_{X,x}\mid v(f)\geqslant m\}\subseteq {\mathcal{O}}_{X,x}\end{eqnarray}$$

for all positive integers $m$ . The volume of $v$ is given by

$$\begin{eqnarray}\operatorname{vol}(v):=\underset{m\rightarrow \infty }{\limsup }\frac{\operatorname{length}({\mathcal{O}}_{X,x}/\mathfrak{a}_{m}(v)_{x})}{m^{n}/n!}.\end{eqnarray}$$

The normalized volume of $v$ is

$$\begin{eqnarray}\widehat{\operatorname{vol}}(v):=A_{X}(v)^{n}\operatorname{vol}(v),\end{eqnarray}$$

where $A_{X}(v)$ is the log discrepancy of $v$ (see § 2.5). When $X$ has Kawamata log terminal (klt) singularities, $A_{X}(v)>0$ , and, thus, $\widehat{\operatorname{vol}}(v)>0$ for all $v\in \operatorname{Val}_{X,x}$ . Li conjectured the following.

Conjecture 1.1 [Reference LiLi15].

If $X$ has klt singularities at $x$ , there exists a valuation $v^{\ast }\in \operatorname{Val}_{X,x}$ that minimizes $\widehat{\operatorname{vol}}_{X,x}$ . Furthermore, such a minimizer $v^{\ast }$ is unique (up to scaling) and quasimonomial.

The above conjecture holds when $x\in X$ is a smooth point. As observed in [Reference LiLi15], if $x$ is a smooth point, then $\widehat{\operatorname{vol}}_{X,x}$ is minimized at $\operatorname{ord}_{x}$ , the valuation that measures order of vanishing at $x$ . Thus,

$$\begin{eqnarray}n^{n}=\widehat{\operatorname{vol}}(\operatorname{ord}_{x})\leqslant \widehat{\operatorname{vol}}(v)\end{eqnarray}$$

for all $v\in \operatorname{Val}_{X,x}$ . The above observation follows from the work of de Fernex et al.

Theorem 1.2 [Reference de Fernex, Ein and MustaţăFEM04].

Let $X$ be a variety of dimension $n$ and $x\in X$ a smooth point. If $\mathfrak{a}\subseteq {\mathcal{O}}_{X,x}$ is an ideal that vanishes precisely at $x$ , then

$$\begin{eqnarray}n^{n}=\operatorname{lct}(\mathfrak{m}_{x})^{n}\operatorname{e}(\mathfrak{m})\leqslant \operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a})\end{eqnarray}$$

where $\mathfrak{m}_{x}$ is the maximal ideal of ${\mathcal{O}}_{X,x}$ .

The authors of the previous theorem were motivated by their interest in singularity theory, as well as applications to birational rigidity [Reference de Fernex, Ein and MustaţăFEM03, Reference de Fernex, Ein and MustaţăFEM04, Reference de FernexFer13]. Li’s interest in volume minimization stems from questions concerning K-semistability of Fano varieties. Let $V$ be a smooth Fano variety and $C(V,-K_{V}):=\operatorname{Spec}(\bigoplus _{m\geqslant 0}H^{0}(V,-mK_{V}))$ the affine cone over $V$ with cone point $0\in C(V,-K_{V})$ . The blowup of $C(V,-K_{V})$ at $0$ has a unique exceptional divisor, which we denote by $\tilde{V}$ .

Theorem 1.3 [Reference LiLi17, Reference Li and LiuLL16, Reference Li and XuLX16].

Let $V$ be a smooth Fano variety. The following are equivalent.

  1. (a) The Fano variety $V$ is $K$ -semistable.

  2. (b) The function $\widehat{\operatorname{vol}}_{C,0}$ is minimized at $\operatorname{ord}_{\tilde{V}}$ .

Thus, if $V$ is K-semistable, there exists a valuation centered at $0\in C(V,-K_{V})$ with smallest normalized volume. If $V$ is not K-semistable, Conjecture 1.1 implies the existence of such a valuation. We prove the following.

Main Theorem. If $x\in X$ is a closed point on a klt variety, then there exists a valuation $v^{\ast }\in \operatorname{Val}_{X,x}$ that is a minimizer of $\widehat{\operatorname{vol}}_{X,x}$ .

In practice, it is rather difficult to pinpoint such a valuation $v^{\ast }$ satisfying the conclusion of this theorem. For a good source of computable examples, we consider the toric setting.

Theorem 1.4. If $X$ is a klt toric variety and $x\in X$ a torus invariant point, then

$$\begin{eqnarray}\inf _{v\in \operatorname{Val}_{X,x}^{\text{toric}}}\widehat{\operatorname{vol}}(v)=\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v),\end{eqnarray}$$

where $\operatorname{Val}_{X,x}^{\text{toric}}$ denotes the set of toric valuations of $X$ with center equal to  $x$ .

In § 8.3, we look at a concrete example, the cone over $\mathbb{P}^{2}$ blown up at a point. In this example, we find a quasimonomial valuation that minimizes the normalized volume function. Additionally, we show that there does not exist a divisorial volume minimizer. While this example is not new, our computation is unique in that it relies on purely algebraic methods. As explained in [Reference Li and XuLX16, Example 6.2], examples from Sasakian geometry with irregular Sasaki–Einstein metrics will provide similar examples. This example was looked at in [Reference Martelli and SparksMS06, § 7].

Sketch of the proof of the main theorem

In order to prove the Main Theorem we first take a sequence of valuations $(v_{i})_{i\in \mathbf{N}}$ such that

$$\begin{eqnarray}\lim _{i\rightarrow \infty }\widehat{\operatorname{vol}}(v_{i})=\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v).\end{eqnarray}$$

Ideally, we would like to find a valuation $v^{\ast }$ that is a limit point of the sequence $(v_{i})_{i\in \mathbf{N}}$ and then argue that $v^{\ast }$ is a minimizer of $\widehat{\operatorname{vol}}_{X,x}$ . To proceed with such an argument, one would likely need to show that $\widehat{\operatorname{vol}}_{X,x}$ is a lower semicontinuous function on $\operatorname{Val}_{X,x}$ . It is unclear how to prove such a statement.Footnote 1

We proceed by shifting our focus. Instead of studying valuations $v\in \operatorname{Val}_{X,x}$ , we may consider ideals $\mathfrak{a}\subseteq {\mathcal{O}}_{X}$ that are $\mathfrak{m}_{x}$ -primary. For an $\mathfrak{m}_{x}$ -primary ideal, the normalized multiplicity of $\mathfrak{a}$ is given by $\operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a})$ , where

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}):=\min _{v\in \operatorname{Val}_{X,x}}\frac{A_{X}(v)}{v(\mathfrak{a})}\quad \text{and}\quad \operatorname{e}(\mathfrak{a}):=\lim _{m\rightarrow \infty }\frac{\operatorname{length}({\mathcal{O}}_{X}/\mathfrak{a}^{m})}{m^{n}/n!},\end{eqnarray}$$

and the above invariants are the log canonical threshold and Hilbert–Samuel multiplicity.

We can also define similar invariants for graded sequences of $\mathfrak{m}_{x}$ -primary ideals. Note that a graded sequence of ideals on $X$ is a sequence of ideals $\mathfrak{a}_{\bullet }=\{\mathfrak{a}_{m}\}_{m\in \mathbf{N}}$ such that $\mathfrak{a}_{m}\cdot \mathfrak{a}_{n}\subseteq \mathfrak{a}_{m+n}$ for all $m,n\in \mathbf{N}$ . The following proposition relates minimizing the normalized volume function to minimizing the normalized multiplicity.

Proposition 4.3 [Reference LiuLiu16]. If $x\in X$ is a closed point on a klt variety, then

(1.1) $$\begin{eqnarray}\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)=\inf _{\mathfrak{a}_{\bullet }\,\mathfrak{m}_{x}\text{-}\text{primary}}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet })=\inf _{\mathfrak{a}\,\mathfrak{m}_{x}\text{-}\text{primary}}\operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a}).\end{eqnarray}$$

While our goal is to find $v^{\ast }\in \operatorname{Val}_{X,x}$ that achieves the first infimum of (1.1), we will instead find a graded sequence of $\mathfrak{m}_{x}$ -primary ideals $\tilde{\mathfrak{a}}_{\bullet }$ that achieves the second infimum of the equation. After having constructed such a graded sequence $\tilde{\mathfrak{a}}_{\bullet }$ , a valuation $v^{\ast }$ that computes $\operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })$ (see § 2.9) will be a minimizer of $\widehat{\operatorname{vol}}_{X,x}$ .

To construct such a graded sequence, we will take our previously mentioned sequence of valuations $(v_{i})_{i\in \mathbf{N}}$ . This gives us a collection of graded sequences of ideals $(\mathfrak{a}_{\bullet }(v_{i}))_{i\in \mathbf{N}}$ . Our goal will be to find a graded sequence $\tilde{\mathfrak{a}}_{\bullet }$ that is a ‘limit point’ of the previous collection.

We recall the work of de Fernex and Mustaţă [Reference de Fernex and MustaţăFM09], Kollár [Reference KollárKol08], and de Fernex et al. [Reference de Fernex, Ein and MustaţăFEM10, Reference de Fernex, Ein and MustaţăFEM11] on generic limits. Given a collection of ideals $\{\mathfrak{a}_{i}\}_{i\in \mathbf{N}}$ where $\mathfrak{a}_{i}\subset k[x_{1},\ldots ,x_{r}]$ , there exists a field extension $k\subseteq K$ and an ideal $\tilde{\mathfrak{a}}\subset K[[x_{1},\ldots ,x_{r}]]$ that encodes information on infinitely many members of $\{\mathfrak{a}_{i}\}_{i\in \mathbf{N}}$ . We extend previous work on generic limits to find a ‘limit point’ of a collection of graded sequences of ideals.

Along the way, we will need a technical result on the rate of convergence of $(\operatorname{e}(\mathfrak{a}_{m}(v))/m^{n})_{m\in \mathbf{N}}$ for a valuation $v\in \operatorname{Val}_{X,x}$ . To perform this task, we extend the work of Ein et al. on approximation of valuation ideals [Reference Ein, Lazarsfeld and SmithELS03] and prove a technical, but also surprising, uniform convergence type result for the volume function.

Proposition 3.7 . Let $X$ be a klt variety of dimension $n$ and $x\in X$ a closed point. For $\unicode[STIX]{x1D716}>0$ and constants $B,E,r\in \mathbf{Z}_{{>}0}$ , there exists $N=N(\unicode[STIX]{x1D716},B,E,r)\in \mathbf{Z}_{{>}0}$ such that for every valuation $v\in \operatorname{Val}_{X,x}$ with $\operatorname{vol}(v)\leqslant B$ , $A_{X}(v)\leqslant E$ , and $v(\mathfrak{m}_{x})\geqslant 1/r$ , we have

$$\begin{eqnarray}\operatorname{vol}(v)\leqslant \frac{\operatorname{e}(\mathfrak{a}_{m}(v))}{m^{n}}<\operatorname{vol}(v)+\unicode[STIX]{x1D716}\quad \text{for all}~m\geqslant N.\end{eqnarray}$$

Structure of the paper

In § 2 we provide preliminary information on valuations, graded sequences of ideals, and log canonical thresholds. Section 3 extends [Reference Ein, Lazarsfeld and SmithELS03] to klt varieties and gives a proof of the previous proposition on the volume of a valuation. Section 4 provides information on Li’s normalized volume function. Section 5 extends the theory of generic limits from ideals to graded sequences of ideals. Section 6 provides a proof of the Main Theorem. In § 7, we explain that the arguments in this paper extend to the setting of log pairs. Lastly, § 8 provides a proof of Theorem 1.4 and a computation of an example of a nondivisorial volume minimizer.

The paper also has two appendices that collect known statements that do not explicitly appear in the literature. Appendix A provides information on the behavior of the Hilbert–Samuel multiplicity and log canonical threshold in families. Appendix B provides a proof of the existence of valuations computing log canonical thresholds on klt varieties.

2 Preliminaries

Conventions

For the purpose of this paper, a variety is an irreducible, reduced, separated scheme of finite type over a field $k$ . Furthermore, we will always assume that $k$ is of characteristic zero, algebraically closed, and uncountable. We use the convention that $\mathbf{N}=\{1,2,3,\ldots \}$ .

2.1 Real valuations

Let $X$ be a variety and $K(X)$ denote its function field. A real valuation of $K(X)$ is a group homomorphism

$$\begin{eqnarray}v:K(X)^{\times }\rightarrow \mathbf{R}\end{eqnarray}$$

such that $v$ is trivial on $k$ (the base field) and $v(f+g)\geqslant \min \{v(f),v(g)\}$ . We use the convention that $v(0)=+\infty$ .

A valuation $v$ gives rise to a valuation ring ${\mathcal{O}}_{v}\subset K(X)$ , where ${\mathcal{O}}_{v}:=\{f\in K(X)\mid v(f)\geqslant 0\}$ . Note that if $v$ is a valuation of $K(X)$ and $\unicode[STIX]{x1D706}\in \mathbf{R}_{{>}0}$ , scaling the outputs of $v$ by $\unicode[STIX]{x1D706}$ gives a new valuation $\unicode[STIX]{x1D706}v$ .

We say that $v$ has a center on $X$ if there exists a map $\unicode[STIX]{x1D70B}:\operatorname{Spec}({\mathcal{O}}_{v})\rightarrow X$ as below.

By [Reference HartshorneHar77, Theorem II.4.3], if such a map $\unicode[STIX]{x1D70B}$ exists, it is necessarily unique. Let $\unicode[STIX]{x1D701}$ denote the unique closed point of $\operatorname{Spec}({\mathcal{O}}_{v})$ . If such a $\unicode[STIX]{x1D70B}$ exists, we define the center of $v$ on $X$ , denoted $c_{X}(v)$ , to be $\unicode[STIX]{x1D70B}(\unicode[STIX]{x1D701})$ . We let $\operatorname{Val}_{X}$ (respectively, $\operatorname{Val}_{X,x}$ ) denote the set of nontrivial real valuations of $K(X)$ with center on $X$ (respectively, center equal to  $x$ ).

Given a valuation $v\in \operatorname{Val}_{X}$ and a nonzero ideal $\mathfrak{a}\subseteq {\mathcal{O}}_{X}$ , we may evaluate $\mathfrak{a}$ along $v$ by setting

$$\begin{eqnarray}v(\mathfrak{a}):=\min \{v(f)\mid f\in \mathfrak{a}\cdot {\mathcal{O}}_{X,c_{X}(v)}\}.\end{eqnarray}$$

When $X$ is affine,

$$\begin{eqnarray}v(\mathfrak{a})=\min \{v(f)\mid f\in \mathfrak{a}(X)\}.\end{eqnarray}$$

It follows from the above definition that if $\mathfrak{a}\subseteq \mathfrak{b}\subset {\mathcal{O}}_{X}$ are nonzero ideals, then $v(\mathfrak{a})\geqslant v(\mathfrak{b})$ . In addition, $v(\mathfrak{a})>0$ if and only if $c_{x}(v)\in \operatorname{Cosupp}(\mathfrak{a})$ .Footnote 2

We endow $\operatorname{Val}_{X}$ with the weakest topology such that, for every ideal $\mathfrak{a}$ on $X$ , the map $\operatorname{Val}_{X}\rightarrow \mathbf{R}\cup \{+\infty \}$ defined by $v\mapsto v(\mathfrak{a})$ is continuous. For information on the space of valuations, see [Reference Jonsson and MustaţǎJM12] and [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15].

2.2 Divisorial valuations

Let $f:Y\rightarrow X$ be a proper birational morphism with $Y$ normal and $E\subset Y$ a prime divisor. The discrete valuation ring ${\mathcal{O}}_{Y,E}$ gives rise to a valuation $\operatorname{ord}_{E}\in \operatorname{Val}_{X}$ that sends $a\in K(X)^{\times }$ to the order of vanishing of $a$ along  $E$ . Note that $c_{X}(\operatorname{ord}_{E})$ is the generic point of $f(E)$ .

We say $v\in \operatorname{Val}_{X}$ is a divisorial valuation if there exists $E$ as above and $\unicode[STIX]{x1D706}\in \mathbf{R}_{{>}0}$ such that $v=\unicode[STIX]{x1D706}\operatorname{ord}_{E}$ . We write $\operatorname{DVal}_{X}\subset \operatorname{Val}_{X}$ for the set of divisorial valuations on  $X$ .

2.3 Quasimonomial valuations

A quasimonomial valuation is a valuation that becomes monomial on some birational model over $X$ . Specifically, let $f:Y\rightarrow X$ be a proper birational morphism and $p\in Y$ a closed point such that $Y$ is regular at $p$ . Given a regular system of parameters $y_{1},\ldots ,y_{n}\in {\mathcal{O}}_{Y,p}$ at $p$ and $\unicode[STIX]{x1D736}=(\unicode[STIX]{x1D6FC}_{1},\ldots ,\unicode[STIX]{x1D6FC}_{n})\in \mathbf{R}_{{\geqslant}0}^{n}\setminus \{\mathbf{0}\}$ , we define a valuation $v_{\unicode[STIX]{x1D736}}$ as follows. For $r\in {\mathcal{O}}_{Y,p}$ we can write $r$ in $\widehat{{\mathcal{O}}_{Y,p}}$ as $r=\sum _{\unicode[STIX]{x1D6FD}\in \mathbf{Z}_{{\geqslant}0}^{n}}c_{\unicode[STIX]{x1D6FD}}y^{\unicode[STIX]{x1D6FD}}$ , with $c_{\unicode[STIX]{x1D6FD}}\in \widehat{{\mathcal{O}}_{Y,p}}$ either zero or unit. We set

$$\begin{eqnarray}v_{\unicode[STIX]{x1D6FC}}(r)=\min \{\langle \unicode[STIX]{x1D6FC},\unicode[STIX]{x1D6FD}\rangle \mid c_{\unicode[STIX]{x1D6FD}}\neq 0\}.\end{eqnarray}$$

A quasimonomial valuation is a valuation that can be written in the above form. Note that in the above notation, if there exists $\unicode[STIX]{x1D706}\in \mathbf{R}_{{>}0}$ such that $\unicode[STIX]{x1D706}\cdot \unicode[STIX]{x1D736}\in \mathbf{Z}_{{\geqslant}0}^{r}$ , then $v_{\unicode[STIX]{x1D6FC}}$ is a divisorial valuation. Indeed, take a weighted blowup of $Y$ at $p$ to find the correct exceptional divisor.

2.4 The relative canonical divisor

Let $f:Y\rightarrow X$ be a proper birational morphism of normal varieties. If $X$ is $\mathbf{Q}$ -Gorenstein, that is $K_{X}$ is $\mathbf{Q}$ -Cartier, we define the relative canonical divisor of $f$ to be

$$\begin{eqnarray}K_{Y/X}:=K_{Y}-f^{\ast }(K_{X}),\end{eqnarray}$$

where $K_{Y}$ and $K_{X}$ are chosen so that $f_{\ast }K_{Y}=K_{X}$ . While $K_{Y}$ and $K_{X}$ are only defined up to linear equivalence, $K_{Y/X}$ is a well-defined divisor.

We say that a variety $X$ is klt if $X$ is normal, $\mathbf{Q}$ -Gorenstein, and for any proper birational morphism of normal varieties $Y\rightarrow X$ the coefficients of $K_{Y/X}$ are ${>}-1$ . For further details on klt varieties and the relative canonical divisor, see [Reference Kollár and MoriKM98, § 2.3].

2.5 The log discrepancy of a valuation

Let $X$ be a normal $\mathbf{Q}$ -Gorenstein variety. If $Y\rightarrow X$ is a proper birational morphism with $Y$ normal, and $E\subset Y$ a prime divisor, then the log discrepancy of $\operatorname{ord}_{E}$ is defined by

$$\begin{eqnarray}A_{X}(\operatorname{ord}_{E}):=1+(\text{coefficient of}~E~\text{in}~K_{Y/X}).\end{eqnarray}$$

As explained in [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15] (building upon [Reference Boucksom, Favre and JonssonBFJ08, Reference Jonsson and MustaţǎJM12]), the log discrepancy can be extended to a lower semicontinuous function $A_{X}:\operatorname{Val}_{X}\rightarrow \mathbf{R}\cup \{+\infty \}$ that is homogeneous of order 1, (i.e. $A_{X}(\unicode[STIX]{x1D706}v)=\unicode[STIX]{x1D706}A_{X}(v)$ for $v\in \operatorname{Val}_{X}$ and $\unicode[STIX]{x1D706}\in \mathbf{R}_{{>}0}$ ). Additionally, $X$ is klt if and only if $A_{X}(v)>0$ for all $v\in \operatorname{Val}_{X}$ .

Note that [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15] defines the log discrepancy function $A_{X}:\operatorname{Val}_{X}\rightarrow \mathbf{R}\cup \{+\infty \}$ when $X$ is normal and $K_{X}$ is not necessarily assumed to be $\mathbf{Q}$ -Cartier. We will not work in this level of generality.

2.6 Graded sequences of ideals

A graded sequence of ideals on a variety $X$ is a sequence of ideals $\mathfrak{a}_{\bullet }=\{\mathfrak{a}_{m}\}_{m\in \mathbf{N}}$ such that $\mathfrak{a}_{m}\cdot \mathfrak{a}_{n}\subseteq \mathfrak{a}_{m+n}$ for all $m,n\in \mathbf{N}$ . By convention, we put $\mathfrak{a}_{0}={\mathcal{O}}_{X}$ . To simplify exposition, we always assume that $\mathfrak{a}_{m}$ is not equal to the zero ideal for all $m\in \mathbf{N}$ .

We provide two examples of graded sequences of ideals.

  1. (a) Let $\mathfrak{b}$ be a nonzero ideal on $X$ . We may define a graded sequence $\mathfrak{a}_{\bullet }$ by setting $\mathfrak{a}_{m}:=\mathfrak{b}^{m}$ for all $m\in \mathbf{N}$ . This example is trivial.

  2. (b) We fix $v\in \operatorname{Val}_{X}$ and define $\mathfrak{a}_{\bullet }(v)=\{\mathfrak{a}_{m}(v)\}_{m\in \mathbf{N}}$ as follows. If $U\subseteq X$ is an open affine set such that $c_{X}(v)\in U$ , then

    $$\begin{eqnarray}\mathfrak{a}_{m}(v)(U):=\{f\in {\mathcal{O}}_{X}(U)\mid v(f)\geqslant m\}.\end{eqnarray}$$
    If $c_{x}(v)\notin U$ , then $\mathfrak{a}_{m}(v)(U):={\mathcal{O}}_{X}(U)$ . If $c_{X}(v)$ is a closed point $x$ , we have that each ideal $\mathfrak{a}_{m}(v)$ is $\mathfrak{m}_{x}$ -primary,Footnote 3 where $\mathfrak{m}_{x}\subseteq {\mathcal{O}}_{X}$ denotes the ideal of functions vanishing at  $x$ .

Given $v\in \operatorname{Val}_{X}$ and a graded sequence $\mathfrak{a}_{\bullet }$ , we may evaluate $\mathfrak{a}_{\bullet }$ along $v$ by setting

$$\begin{eqnarray}v(\mathfrak{a}_{\bullet }):=\inf _{m\in \mathbf{N}}\frac{v(\mathfrak{a}_{m})}{m}=\lim _{m\rightarrow \infty }\frac{v(\mathfrak{a}_{m})}{m}.\end{eqnarray}$$

See [Reference Jonsson and MustaţǎJM12, Lemma 2.3] for a proof of the previous equality.

2.7 Multiplicities

Let $X$ be a variety of dimension $n$ and $x\in X$ a closed point. Let $\mathfrak{m}_{x}\subseteq {\mathcal{O}}_{X}$ denote the ideal of functions vanishing at  $x$ . We recall that for an $\mathfrak{m}_{x}$ -primary ideal $\mathfrak{a}$ , the Hilbert–Samuel multiplicity of $\mathfrak{a}$ is

$$\begin{eqnarray}\operatorname{e}(\mathfrak{a}):=\lim _{m\rightarrow \infty }\frac{\operatorname{length}({\mathcal{O}}_{X,x}/\mathfrak{a}^{m})}{m^{n}/n!}.\end{eqnarray}$$

If $\mathfrak{a}\subseteq \mathfrak{b}\subseteq {\mathcal{O}}_{X}$ are $\mathfrak{m}_{x}$ -primary ideals on $X$ , then $\operatorname{e}(\mathfrak{a})\geqslant \operatorname{e}(\mathfrak{b})$ . In addition, $\operatorname{e}(\mathfrak{a})=\operatorname{e}(\overline{\mathfrak{a}})$ where $\overline{\mathfrak{a}}$ denotes the integral closure of  $\mathfrak{a}$ .

We recall the valuative definition of the integral closure of an ideal $\mathfrak{a}$ on a normal variety $X$ [Reference LazarsfeldLaz04, Example 9.6.8]. Let $U\subset X$ affine open subset. We have

$$\begin{eqnarray}\overline{\mathfrak{a}}(U):=\{f\in {\mathcal{O}}_{X}(U)\mid w(f)\geqslant w(\mathfrak{a})~\text{for all}~w\in \operatorname{Val}_{U}~\text{divisorial}\}.\end{eqnarray}$$

2.8 Volumes

Let $\mathfrak{a}_{\bullet }$ be a graded sequence of ideals with the property that each $\mathfrak{a}_{m}$ is $\mathfrak{m}_{x}$ -primary. The volume of $\mathfrak{a}_{\bullet }$ is defined as

$$\begin{eqnarray}\operatorname{vol}(\mathfrak{a}_{\bullet }):=\limsup _{m\rightarrow \infty }\frac{\operatorname{length}({\mathcal{O}}_{X,x}/\mathfrak{a}_{m})}{m^{n}/n!}.\end{eqnarray}$$

A similar invariant is the multiplicity of $\mathfrak{a}_{\bullet }$ , which is defined as

$$\begin{eqnarray}\operatorname{e}(\mathfrak{a}_{\bullet })=\lim _{m\rightarrow \infty }\frac{\operatorname{e}(\mathfrak{a}_{m})}{m^{n}}.\end{eqnarray}$$

In various degrees of generality, it has been proven that

$$\begin{eqnarray}\operatorname{e}(\mathfrak{a}_{\bullet })=\operatorname{vol}(\mathfrak{a}_{\bullet })\end{eqnarray}$$

[Reference Ein, Lazarsfeld and SmithELS03, Corollary C], [Reference MustaţǎMus02, Theorem 1.7], [Reference Lazarsfeld and MustaţǎLM09, Theorem 3.8], [Reference CutkoskyCut13, Theorem 6.5]. In our setting, the above equality will always hold. In addition, by [Reference CutkoskyCut13, Theorem 1.1], we also have that

$$\begin{eqnarray}\operatorname{vol}(\mathfrak{a}_{\bullet }):=\lim _{m\rightarrow \infty }\frac{\operatorname{length}({\mathcal{O}}_{X,x}/\mathfrak{a}_{m})}{m^{n}/n!}.\end{eqnarray}$$

For a valuation $v\in \operatorname{Val}_{X,x}$ , the volume of $v$ is given by

$$\begin{eqnarray}\operatorname{vol}(v):=\operatorname{e}(\mathfrak{a}_{\bullet }(v)).\end{eqnarray}$$

Note that if $\unicode[STIX]{x1D706}\in \mathbf{R}_{{>}0}$ , then $\operatorname{vol}(\unicode[STIX]{x1D706}v)=\operatorname{vol}(v)/\unicode[STIX]{x1D706}^{n}$ .

2.9 Log canonical thresholds

The log canonical threshold is an invariant of singularities that has received considerable interest in the field of birational geometry [Reference KollárKol97, § 8]. For a nonzero ideal $\mathfrak{a}$ on a klt variety  $X$ , the log canonical threshold of $\mathfrak{a}$ is given by

(2.1) $$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}):=\inf _{v\in \operatorname{DVal}_{X}}\frac{A_{X}(v)}{v(\mathfrak{a})}.\end{eqnarray}$$

By [Reference Jonsson and MustaţǎJM12, Lemma 6.7] in the case when $X$ is smooth and a similar argument in the klt case,

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a})=\inf _{v\in \operatorname{Val}_{X}}\frac{A_{X}(v)}{v(\mathfrak{a})}.\end{eqnarray}$$

In the previous expression, we are using the convention that if $v(\mathfrak{a})=0$ or $A(v)<+\infty$ , then $A(v)/v(\mathfrak{a})=+\infty$ . Thus, $\operatorname{lct}({\mathcal{O}}_{X})=+\infty$ . We say that a valuation $v^{\ast }$ computes $\operatorname{lct}(\mathfrak{a})$ if $A_{X}(v)<+\infty$ and $\operatorname{lct}(\mathfrak{a})=A(v^{\ast })/v^{\ast }(\mathfrak{a})$ .

Note the following elementary properties of this invariant. If $m\in \mathbf{Z}_{{>}0}$ , then

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}^{m})=\operatorname{lct}(\mathfrak{a})/m.\end{eqnarray}$$

If $\mathfrak{a}\subseteq \mathfrak{b}$ , then

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a})\leqslant \operatorname{lct}(\mathfrak{b}).\end{eqnarray}$$

The log canonical threshold may be understood in terms of a log resolution of $\mathfrak{a}$ . Recall that $\unicode[STIX]{x1D707}:Y\rightarrow X$ is a log resolution of $\mathfrak{a}$ if the following hold:

  1. (a) $\unicode[STIX]{x1D707}$ is a projective birational morphism;

  2. (b) $Y$ is smooth and $\operatorname{Exc}(\unicode[STIX]{x1D707})$ is pure codimension 1;

  3. (c) $\mathfrak{a}\cdot {\mathcal{O}}_{Y}={\mathcal{O}}_{Y}(-D)$ for an effective divisor $D$ on $Y$ ; and

  4. (d) $D_{\text{red}}+\operatorname{Exc}(\unicode[STIX]{x1D707})$ has simple normal crossing.

In this case, we have

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a})=\min _{i=1,\ldots ,r}\frac{A_{X}(\operatorname{ord}_{E_{i}})}{\operatorname{ord}_{E_{i}}(\mathfrak{a})},\end{eqnarray}$$

where $D=\sum _{i=1}^{r}a_{i}E_{i}$ . (Note that $\operatorname{ord}_{E_{i}}(\mathfrak{a})=a_{i}$ .) Thus, there always exists a divisorial valuation that computes $\operatorname{lct}(\mathfrak{a})$ .

For a graded sequence of ideals $\mathfrak{a}_{\bullet }$ on $X$ , the log canonical threshold of $\mathfrak{a}_{\bullet }$ is given by

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet }):=\lim _{m\rightarrow \infty }m\cdot \operatorname{lct}(\mathfrak{a}_{m})=\sup _{m}m\cdot \operatorname{lct}(\mathfrak{a}_{m}).\end{eqnarray}$$

By [Reference Jonsson and MustaţǎJM12, Corollary 6.9] when $X$ is smooth or as a consequence of [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15, Theorem 1.2] when $X$ is klt, we have

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet })=\inf _{v\in \operatorname{Val}_{X}}\frac{A_{X}(v)}{v(\mathfrak{a}_{\bullet })}.\end{eqnarray}$$

As before, we are using the convention that if either $v(\mathfrak{a}_{\bullet })=0$ or $A_{X}(v)=+\infty$ , then $A(v)/v(\mathfrak{a}_{\bullet })=+\infty$ . We say $v^{\ast }\in \operatorname{Val}_{X}$ computes $\operatorname{lct}(\mathfrak{a}_{\bullet })$ if $A(v)<+\infty$ and $\operatorname{lct}(\mathfrak{a}_{\bullet })=A_{X}(v^{\ast })/v^{\ast }(\mathfrak{a}_{\bullet })$ . Such valuations $v^{\ast }$ always exist (see Appendix B). When $X$ is smooth, this is precisely [Reference Jonsson and MustaţǎJM12, Theorem A].

3 Approximation of valuation ideals

In this section we extend the arguments of [Reference Ein, Lazarsfeld and SmithELS03] to approximate valuation ideals on singular varieties. We will use this approximation to determine the speed of convergence of $(\operatorname{e}(\mathfrak{a}_{m}(v))/m^{n})_{m\in \mathbf{N}}$ for a fixed valuation  $v$ . The main technical tool is the asymptotic multiplier ideal of a graded family of ideals. For an excellent reference on multiplier ideals, see [Reference LazarsfeldLaz04, ch. 9].

3.1 Multiplier ideals

Let $X$ be a normal $\mathbf{Q}$ -Gorenstein variety. Fix a nonzero ideal $\mathfrak{a}\subseteq {\mathcal{O}}_{X}$ and $f:Y\rightarrow X$ a log resolution of $\mathfrak{a}$ such that $\mathfrak{a}\cdot {\mathcal{O}}_{Y}={\mathcal{O}}_{Y}(-D)$ . For a rational number $c>0$ , the multiplier ideal ${\mathcal{J}}(X,c\cdot \mathfrak{a})$ is defined by

$$\begin{eqnarray}{\mathcal{J}}(X,c\cdot \mathfrak{a}):=f_{\ast }{\mathcal{O}}_{Y}(\lceil K_{Y/X}-cD\rceil )\subseteq {\mathcal{O}}_{X}.\end{eqnarray}$$

Note that if $c$ is an integer, then ${\mathcal{J}}(X,c\cdot \mathfrak{a})={\mathcal{J}}(X,\mathfrak{a}^{c})$ . It is a basic fact that ${\mathcal{J}}(X,c\cdot \mathfrak{a})$ is independent of the choice of  $f$ .

Alternatively, the multiplier ideal can be understood valuatively. If $X$ is an affine variety, then [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15, Theorem 1.2] implies

$$\begin{eqnarray}{\mathcal{J}}(X,c\cdot \mathfrak{a})(X)=\{f\in {\mathcal{O}}_{X}(X)\mid v(f)>cv(\mathfrak{a})-A_{X}(v)~\text{for all}~v\in \operatorname{Val}_{X}\}.\end{eqnarray}$$

When $X$ is not necessarily affine, the above criterion allows us to understand the multiplier ideal locally.

It is important to note the relationship between the log canonical threshold and the multiplier ideal. If $X$ is klt, then

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a})=\sup \{c\mid {\mathcal{J}}(X,c\cdot \mathfrak{a})={\mathcal{O}}_{X}\}.\end{eqnarray}$$

The following lemma provides basic properties of multiplier ideals. The proof is left to the reader. See [Reference LazarsfeldLaz04, Proposition 9.2.32] for the case when $X$ is smooth.

Lemma 3.1. Let $\mathfrak{a},\mathfrak{b}$ be nonzero ideals on $X$ .

  1. (a) If $X$ is klt, then

    $$\begin{eqnarray}\mathfrak{a}\subseteq {\mathcal{J}}(X,\mathfrak{a}).\end{eqnarray}$$
  2. (b) If $\mathfrak{a}\subseteq \mathfrak{b}$ and $c>0$ is a rational number, then

    $$\begin{eqnarray}{\mathcal{J}}(X,c\cdot \mathfrak{a})\subseteq {\mathcal{J}}(X,c\cdot \mathfrak{b}).\end{eqnarray}$$
  3. (c) For rational numbers $c\geqslant d>0$ , we have

    $$\begin{eqnarray}{\mathcal{J}}(X,c\cdot \mathfrak{a})\subseteq {\mathcal{J}}(X,d\cdot \mathfrak{a}).\end{eqnarray}$$

Multiplier ideals satisfy the following ‘subadditivity property’. The property was first observed and proved by Demailly et al. in the smooth case [Reference Demailly, Ein and LazarsfeldDEL00]. The statement was extended to the singular case in [Reference TakagiTak06, Theorem 2.3] and [Reference EisensteinEis11, Theorem 7.3.4].

Theorem 3.2 (Subadditivity).

If $\mathfrak{a},\mathfrak{b}$ are nonzero ideals on $X$ and $c>0$ a rational number, then

$$\begin{eqnarray}\operatorname{Jac}_{X}\cdot {\mathcal{J}}(X,c\cdot (\mathfrak{a}\cdot \mathfrak{b}))\subseteq {\mathcal{J}}(X,c\cdot \mathfrak{a})\cdot {\mathcal{J}}(X,c\cdot \mathfrak{b}),\end{eqnarray}$$

where $\operatorname{Jac}_{X}$ denotes the Jacobian ideal of $X$ .

We recall that the Jacobian ideal of a variety $X$ is $\operatorname{Jac}_{X}:=\operatorname{Fitt}_{n}(X)$ , where $n$ is the dimension of $X$ and $\operatorname{Fitt}_{n}$ denotes the $n$ th fitting ideal as in [Reference EisenbudEis95, 20.2]. The singular locus of $X$ is equal to $\operatorname{Cosupp}(\operatorname{Jac}_{X})$ .

3.2 Asymptotic multiplier ideals

Let $\mathfrak{a}_{\bullet }$ be a graded sequence of ideals on a normal $\mathbf{Q}$ -Gorenstein variety $X$ and $c>0$ a rational number. We recall the definition of the asymptotic multiplier ideal ${\mathcal{J}}(c\cdot \mathfrak{a}_{\bullet })$ . By Lemma 3.1, we have that

$$\begin{eqnarray}{\mathcal{J}}\biggl(X,\frac{1}{p}\cdot \mathfrak{a}_{mp}\biggr)\subseteq {\mathcal{J}}\biggl(X,\frac{1}{pq}\cdot \mathfrak{a}_{pqm}\biggr)\end{eqnarray}$$

for all positive integers $p,q$ . From the above inclusion and Noetherianity, we conclude

$$\begin{eqnarray}\biggl\{{\mathcal{J}}\biggl(X,\frac{1}{p}\cdot \mathfrak{a}_{pm}\biggr)\biggr\}_{p\in \mathbf{N}}\end{eqnarray}$$

has a unique maximal element. The $m$ th asymptotic multiplier ideal ${\mathcal{J}}(X,m\cdot \mathfrak{a}_{\bullet })$ is defined to be this element. Like the standard multiplier ideal, the asymptotic multiplier ideal can also be understood valuatively.

Proposition 3.3 [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15, Theorem 1.2].

If $X$ is affine, $\mathfrak{a}_{\bullet }$ is a graded sequence of ideals on $X$ , and $c>0$ a rational number, then

$$\begin{eqnarray}{\mathcal{J}}(X,c\cdot \mathfrak{a}_{\bullet })=\{f\in {\mathcal{O}}_{X}(X)\mid v(f)>cv(\mathfrak{a}_{\bullet })-A_{X}(v)~\text{for all}~v\in \operatorname{Val}_{X}\}.\end{eqnarray}$$

The asymptotic multiplier ideals satisfy the following property. This property will allow us to approximate valuation ideals.

Proposition 3.4. If $\mathfrak{a}_{\bullet }$ is a graded sequence of ideals on a klt variety $X$ and $m,\ell \in \mathbf{N}$ , then

$$\begin{eqnarray}(\operatorname{Jac}_{X})^{\ell -1}\mathfrak{a}_{m}^{\ell }\subseteq (\operatorname{Jac}_{X})^{\ell -1}\mathfrak{a}_{m\ell }\subseteq {\mathcal{J}}(m\cdot \mathfrak{a}_{\bullet })^{\ell }.\end{eqnarray}$$

Proof. The proof is the same as the proof of [Reference Ein, Lazarsfeld and SmithELS03, Theorem 1.7] and is a consequence of Lemma 3.1(a) and Theorem 3.2. ◻

3.3 The case of valuation ideals

For a valuation $v\in \operatorname{Val}_{X}$ , we examine the asymptotic multiplier ideals of $\mathfrak{a}_{\bullet }(v)$ . We first prove the following elementary lemma.

Lemma 3.5. If $v$ is a valuation on a variety $X$ , then $v(\mathfrak{a}_{\bullet }(v))=1$ .

Proof. Note that $v(\mathfrak{a}_{m}(v))\geqslant m$ , since $\mathfrak{a}_{m}(v)$ is the ideal of functions vanishing to at least order $m$ along  $v$ . Next, set $\unicode[STIX]{x1D6FC}:=v(\mathfrak{a}_{1}(v))$ . We have $\mathfrak{a}_{1}(v)^{\lceil m/\unicode[STIX]{x1D6FC}\rceil }\subseteq \mathfrak{a}_{m}(v)$ , since $v(\mathfrak{a}_{1}(v)^{\lceil m/\unicode[STIX]{x1D6FC}\rceil })=\unicode[STIX]{x1D6FC}\lceil m/\unicode[STIX]{x1D6FC}\rceil \geqslant m$ . Thus,

$$\begin{eqnarray}v(\mathfrak{a}_{m}(v))\leqslant v(\mathfrak{a}_{1}(v)^{\lceil m/\unicode[STIX]{x1D6FC}\rceil })=\unicode[STIX]{x1D6FC}\lceil m/\unicode[STIX]{x1D6FC}\rceil .\end{eqnarray}$$

The previous two bounds combine to show

$$\begin{eqnarray}1\leqslant \frac{v(\mathfrak{a}_{m}(v))}{m}\leqslant \frac{\unicode[STIX]{x1D6FC}\cdot \lceil m/\unicode[STIX]{x1D6FC}\rceil }{m},\end{eqnarray}$$

and the result follows. ◻

The following results allows us to approximate valuation ideals. In the case when $X$ is smooth and $v$ is an Abhyankhar valuation, the theorem below is a slight strengthening of [Reference Ein, Lazarsfeld and SmithELS03, Theorem A].

Theorem 3.6. If $X$ is a klt variety and $v\in \operatorname{Val}_{X}$ satisfies $A_{X}(v)<+\infty$ , then

$$\begin{eqnarray}(\operatorname{Jac}_{X})^{\ell -1}\cdot \mathfrak{a}_{m}^{\ell }\subseteq (\operatorname{Jac}_{X})^{\ell -1}\cdot \mathfrak{a}_{m\ell }\subseteq \mathfrak{a}_{m-e}^{\ell }\end{eqnarray}$$

for every $m\geqslant e$ and $\ell \in \mathbf{Z}_{{>}0}$ , where $\mathfrak{a}_{\bullet }:=\mathfrak{a}_{\bullet }(v)$ and $e:=\lceil A_{X}(v)\rceil$ .

Proof. By Proposition 3.4, we have that

$$\begin{eqnarray}(\operatorname{Jac}_{X})^{\ell -1}\cdot \mathfrak{a}_{m}^{\ell }\subseteq (\operatorname{Jac}_{X})^{\ell -1}\cdot \mathfrak{a}_{m\ell }\subseteq {\mathcal{J}}(X,m\cdot \mathfrak{a}_{\bullet })^{\ell }.\end{eqnarray}$$

Applying Proposition 3.3 and Lemma 3.5 to $\mathfrak{a}_{\bullet }$ gives that

$$\begin{eqnarray}{\mathcal{J}}(X,m\cdot \mathfrak{a}_{\bullet })\subseteq \mathfrak{a}_{m-e},\end{eqnarray}$$

and the result follows. ◻

3.4 Uniform approximation of volumes

Given a valuation $v\in \operatorname{Val}_{X}$ centered at a closed point on a $n$ -dimensional variety $X$ , we recall

$$\begin{eqnarray}\operatorname{vol}(v)=\lim _{m\rightarrow \infty }\frac{\operatorname{e}(\mathfrak{a}_{m}(v))}{m^{n}},\end{eqnarray}$$

where $n$ is the dimension of $X$ . The following proposition provides a uniform rate of convergence for the terms in the above limit.

Proposition 3.7. Let $X$ be a klt $n$ -dimensional variety and $x\in X$ a closed point. For $\unicode[STIX]{x1D716}>0$ and constants $B,E,r\in \mathbf{Z}_{{>}0}$ , there exists $N=N(\unicode[STIX]{x1D716},B,E,r)$ such that for every valuation $v\in \operatorname{Val}_{X,x}$ with $\operatorname{vol}(v)\leqslant B$ , $A_{X}(v)\leqslant E$ , and $v(\mathfrak{m}_{x})\geqslant 1/r$ , we have

$$\begin{eqnarray}\operatorname{vol}(v)\leqslant \frac{\operatorname{e}(\mathfrak{a}_{m}(v))}{m^{n}}<\operatorname{vol}(v)+\unicode[STIX]{x1D716}\quad \text{for all}~m\geqslant N.\end{eqnarray}$$

Remark 3.8. In an earlier version of this paper, we proved the following statement with the additional assumption that $x\in X$ is an isolated singularity. We are grateful to Mircea Mustaţă for noticing that a modification of the original proof allows us to prove the more general statement.

Proof of Proposition 3.7.

For any valuation $v\in \operatorname{Val}_{X,x}$ , the first inequality is well known. Indeed, the inclusion $\mathfrak{a}_{m}(v)^{p}\subseteq \mathfrak{a}_{mp}(v)$ for $m,p\in \mathbf{N}$ implies that

$$\begin{eqnarray}\frac{\operatorname{e}(\mathfrak{a}_{mp}(v))}{(mp)^{n}}\leqslant \frac{\operatorname{e}(\mathfrak{a}_{m}(v))}{(m)^{n}}.\end{eqnarray}$$

Fixing $m$ and sending $p\rightarrow \infty$ gives

$$\begin{eqnarray}\operatorname{vol}(v)\leqslant \frac{\operatorname{e}(\mathfrak{a}_{m}(v))}{(m)^{n}}.\end{eqnarray}$$

Next, fix $v\in \operatorname{Val}_{X,x}$ satisfying the hypotheses in the statement of Proposition 3.7. We have

$$\begin{eqnarray}(\operatorname{Jac}_{X})^{\ell -1}\mathfrak{a}_{m\ell }(v)\subseteq (\mathfrak{a}_{m-e}(v))^{\ell }\subseteq (\mathfrak{a}_{m-E}(v))^{\ell },\end{eqnarray}$$

where $e=\lceil A_{X}(v)\rceil$ . The first inclusion is the statement in Theorem 3.6, and the second follows from the assumption that $e\leqslant E$ . After replacing $m$ by $m+E$ , we obtain

(3.1) $$\begin{eqnarray}(\operatorname{Jac}_{X})^{\ell }\cdot \mathfrak{a}_{(m+E)\ell }(v)\subseteq (\operatorname{Jac}_{X})^{\ell -1}\cdot \mathfrak{a}_{(m+E)\ell }(v)\subseteq \mathfrak{a}_{m}(v)^{\ell }.\end{eqnarray}$$

On the other hand, the assumption that $v(\mathfrak{m}_{x})\geqslant 1/r$ implies that

(3.2) $$\begin{eqnarray}\mathfrak{m}_{x}^{mr}\subseteq \mathfrak{a}_{m}(v).\end{eqnarray}$$

It follows from (3.1) and (3.2) and the valuative criterion for integral closure (§ 2.7) that

(3.3) $$\begin{eqnarray}(\operatorname{Jac}_{X}+\mathfrak{m}_{x}^{mr})^{\ell }\mathfrak{a}_{(m+E)\ell }(v)\subseteq \overline{\mathfrak{a}_{m}(v)^{\ell }}.\end{eqnarray}$$

Indeed, let $w\in \operatorname{Val}_{X}$ be a divisorial valuation and $f$ and $g$ local sections of $\operatorname{Jac}_{X}^{i}$ and $\mathfrak{m}_{x}^{mrj}$ , respectively, with $i+j=\ell$ . We have $\ell \cdot w(f)+i\cdot w(\mathfrak{a}_{(m+E)\ell }(v))\geqslant i\cdot w(\mathfrak{a}_{m}(v)^{\ell })$ and $w(g)\geqslant j\cdot w(\mathfrak{a}_{m}(v))$ by the two inclusions. Thus,

$$\begin{eqnarray}\displaystyle w(fg)=w(f)+w(g) & {\geqslant} & \displaystyle \frac{i}{\ell }(w(\mathfrak{a}_{m}(v)^{\ell })-w(\mathfrak{a}_{(m+E)\ell }(v)))+\frac{j}{\ell }w(\mathfrak{a}_{m}(v)^{\ell })\nonumber\\ \displaystyle & = & \displaystyle w(\mathfrak{a}_{m}(v)^{\ell })-w(\mathfrak{a}_{(m+E)\ell }(v)).\nonumber\end{eqnarray}$$

From Inclusion (3.3) and Teissier’s Minkowski inequality [Reference TeissierTei77], we see

$$\begin{eqnarray}\operatorname{e}(\mathfrak{a}_{m}(v)^{\ell })^{1/n}\leqslant \operatorname{e}((\operatorname{Jac}_{X}+\mathfrak{m}_{x}^{mr})^{\ell })^{1/n}+\operatorname{e}(\mathfrak{a}_{(m+E)\ell }(v))^{1/n}.\end{eqnarray}$$

Next, note that if $\mathfrak{a}$ is an $\mathfrak{m}_{x}$ -primary ideal, then $\operatorname{e}(\mathfrak{a}^{m})=m^{n}\operatorname{e}(\mathfrak{a})$ . Applying this property and dividing by $m\cdot \ell$ , gives that

$$\begin{eqnarray}\frac{\operatorname{e}(\mathfrak{a}_{m}(v))^{1/n}}{m}\leqslant \frac{\operatorname{e}(\operatorname{Jac}_{X}+\mathfrak{m}_{x}^{mr})^{1/n}}{m}+\frac{m+E}{m}\cdot \frac{\operatorname{e}(\mathfrak{a}_{(m+E)\ell }(v))^{1/n}}{(m+E)\ell }.\end{eqnarray}$$

After letting $\ell \rightarrow \infty$ , we obtain

$$\begin{eqnarray}\frac{\operatorname{e}(\mathfrak{a}_{m}(v))^{1/n}}{m}\leqslant \frac{\operatorname{e}(\operatorname{Jac}_{X}+\mathfrak{m}_{x}^{mr})^{1/n}}{m}+\frac{m+E}{m}\operatorname{vol}(v)^{1/n}.\end{eqnarray}$$

Since $\operatorname{vol}(v)^{1/n}\leqslant B^{1/n}$ , the assertion will follow if we show that

$$\begin{eqnarray}\lim _{m\rightarrow \infty }\frac{\operatorname{e}(\operatorname{Jac}_{X}+\mathfrak{m}_{x}^{mr})^{1/n}}{m}=0.\end{eqnarray}$$

Choose $h\in \operatorname{Jac}_{X}\cdot {\mathcal{O}}_{X,x}$ that is nonzero and set $R:={\mathcal{O}}_{X,x}/(h)$ and $\tilde{\mathfrak{m}}_{x}:=\mathfrak{m}_{x}\cdot R$ . We have

$$\begin{eqnarray}\lim _{m\rightarrow \infty }\frac{\operatorname{e}(\operatorname{Jac}_{X}+\mathfrak{m}_{x}^{mr})^{1/n}}{m}=\lim _{m\rightarrow \infty }\frac{\operatorname{length}({\mathcal{O}}_{X,x}/(\operatorname{Jac}_{X}+\mathfrak{m}_{x}^{mr}))^{1/n}}{m}\leqslant \lim _{m\rightarrow \infty }\frac{\operatorname{length}(R/\tilde{\mathfrak{m}}_{x}^{mr})^{1/n}}{m}.\end{eqnarray}$$

The last limit is 0, since

$$\begin{eqnarray}\lim _{m\rightarrow \infty }\frac{\operatorname{length}(R/{\tilde{\mathfrak{m}}_{x}}^{mr})}{m^{n-1}/(n-1)!}=\operatorname{e}({\tilde{\mathfrak{m}}_{x}}^{r})<\infty .\Box\end{eqnarray}$$

4 Normalized volumes

For this section, we fix $X$ an $n$ -dimensional klt variety and $x\in X$ a closed point. As introduced in [Reference LiLi15], the normalized volume of a valuation $v\in \operatorname{Val}_{X,x}$ is defined as

$$\begin{eqnarray}\widehat{\operatorname{vol}}(v):=A_{X}(v)^{n}\operatorname{vol}(v).\end{eqnarray}$$

In the case when $A_{X}(v)=+\infty$ and $\operatorname{vol}(v)=0$ , we set $\widehat{\operatorname{vol}}(v):=+\infty$ . The word ‘normalized’ refers to the property that $\widehat{\operatorname{vol}}(\unicode[STIX]{x1D706}v)=\widehat{\operatorname{vol}}(v)$ for $\unicode[STIX]{x1D706}\in \mathbf{R}_{{>}0}$ .

Given a graded sequence $\mathfrak{a}_{\bullet }$ of $\mathfrak{m}_{x}$ -primary ideals on $X$ , we define a similar invariant. We refer to

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet })\end{eqnarray}$$

as the normalized multiplicity of $\mathfrak{a}_{\bullet }$ . Similar to the normalized volume, when $\operatorname{lct}(\mathfrak{a}_{\bullet })=+\infty$ and $\operatorname{e}(\mathfrak{a}_{\bullet })=0$ , we set $\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet }):=+\infty$ . The above invariant was looked at in [Reference de Fernex, Ein and MustaţăFEM04] and [Reference MustaţǎMus02].

The following lemma provides elementary information on the normalized multiplicity. The proof is left to the reader.

Lemma 4.1. Let $\mathfrak{a}$ be an $\mathfrak{m}_{x}$ -primary ideal and $\mathfrak{a}_{\bullet }$ a graded sequence of $\mathfrak{m}_{x}$ -primary ideals on  $X$ .

  1. (a) If $\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet })<+\infty$ , then

    $$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet })=\lim _{m\rightarrow \infty }\operatorname{lct}(\mathfrak{a}_{m})^{n}\operatorname{e}(\mathfrak{a}_{m}).\end{eqnarray}$$
  2. (b) If $\mathfrak{b}_{\bullet }$ is a graded sequence given by $\mathfrak{b}_{m}:=\mathfrak{a}^{m}$ , then

    $$\begin{eqnarray}\operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a})=\operatorname{lct}(\mathfrak{b}_{\bullet })^{n}\operatorname{e}(\mathfrak{b}_{\bullet }).\end{eqnarray}$$
  3. (c) If $\mathfrak{a}_{N\bullet }$ is the graded sequence whose $m$ th term is $\mathfrak{a}_{N\cdot m}$ , then

    $$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet })=\operatorname{lct}(\mathfrak{a}_{N\bullet })^{n}\operatorname{e}(\mathfrak{a}_{N\bullet }).\end{eqnarray}$$

Remark 4.2. Fix $\unicode[STIX]{x1D6FF}>0$ . If $\mathfrak{a}_{\bullet }$ a graded sequence of $\mathfrak{m}_{x}$ -primary ideals such that $\mathfrak{a}_{m}\subseteq \mathfrak{m}_{x}^{\lfloor \unicode[STIX]{x1D6FF}m\rfloor }$ for all $m$ , then

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet })<+\infty .\end{eqnarray}$$

It is always the case that $\operatorname{e}(\mathfrak{a}_{\bullet })<+\infty$ , since $\operatorname{e}(\mathfrak{a}_{\bullet })\leqslant \operatorname{e}(\mathfrak{a}_{1})$ . The assumption that $\mathfrak{a}_{m}\subseteq \mathfrak{m}_{x}^{\lfloor \unicode[STIX]{x1D6FF}m\rfloor }$ gives that $\operatorname{lct}(\mathfrak{a}_{\bullet })\leqslant \operatorname{lct}(\mathfrak{m}_{x})/\unicode[STIX]{x1D6FF}$ , the latter of which is ${<}+\!\infty$ .

The following proposition relates the normalized volume, an invariant of valuations, to the normalized multiplicity, an invariant of graded sequences of ideals.

Proposition 4.3 [Reference LiuLiu16, Theorem 27].

The following equality holds:

$$\begin{eqnarray}\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)=\inf _{\mathfrak{a}_{\bullet }\,\mathfrak{m}_{x}\text{-}\text{primary}}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet })=\inf _{\mathfrak{a}\,\mathfrak{m}_{x}\text{-}\text{primary}}\operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a}).\end{eqnarray}$$

The previous statement first appeared in [Reference LiuLiu16]. In the case when $x\in X$ is a smooth point, it was explained in [Reference LiLi15, Example 3.7]. We provide Liu’s proof, since the argument will be useful to us. The proposition is a consequence of the following lemma.

Lemma 4.4 [Reference LiuLiu16].

The following statements hold.

  1. (a) If $\mathfrak{a}_{\bullet }$ is a graded sequence of $\mathfrak{m}_{x}$ -primary ideals and $v\in \operatorname{Val}_{X,x}$ computes $\operatorname{lct}(\mathfrak{a}_{\bullet })$ (i.e. $A(v)/v(\mathfrak{a}_{\bullet })=\operatorname{lct}(\mathfrak{a}_{\bullet })$ ), then

    $$\begin{eqnarray}\widehat{\operatorname{vol}}(v)\leqslant \operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet }).\end{eqnarray}$$
  2. (b) If $v\in \operatorname{Val}_{X,x}$ , then

    $$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet }(v))^{n}\operatorname{e}(\mathfrak{a}_{\bullet }(v))\leqslant \widehat{\operatorname{vol}}(v).\end{eqnarray}$$

Proof. To prove statement (a), we first rescale $v$ so that $v(\mathfrak{a}_{\bullet })=1$ . Thus, $A_{X}(v)=A_{X}(v)/v(\mathfrak{a}_{\bullet })=\operatorname{lct}(\mathfrak{a}_{\bullet })$ . Since

$$\begin{eqnarray}1=v(\mathfrak{a}_{\bullet }):=\inf _{m\geqslant 0}\frac{v(\mathfrak{a}_{m})}{m},\end{eqnarray}$$

we see $v(\mathfrak{a}_{m})\geqslant m$ and, thus, $\mathfrak{a}_{m}\subseteq \mathfrak{a}_{m}(v)$ for all  $m$ . This implies $\operatorname{e}(\mathfrak{a}_{\bullet }(v))\leqslant \operatorname{e}(\mathfrak{a}_{\bullet })$ , and the desired inequality follows.

In order to show statement (b), we note

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet }(v)):=\min _{w}\frac{A_{X}(w)}{w(\mathfrak{a}_{\bullet }(v))}\leqslant \frac{A_{X}(v)}{v(\mathfrak{a}_{\bullet }(v))}=A_{X}(v),\end{eqnarray}$$

where the last equality follows from Lemma 3.5. Thus,

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet }(v))^{n}\operatorname{e}(\mathfrak{a}_{\bullet }(v))\leqslant A_{X}(v)^{n}\operatorname{e}(\mathfrak{a}_{\bullet }(v))=\widehat{\operatorname{vol}}(v).\Box\end{eqnarray}$$

Proof of Proposition 4.3.

The first equality follows immediately from the previous proposition and the fact that given a graded sequence $\mathfrak{a}_{\bullet }$ there exists a valuation $v^{\ast }\in \operatorname{Val}_{X}$ that computes $\operatorname{lct}(\mathfrak{a}_{\bullet })$ (see Theorem B.1). The last equality follows from Lemma 4.1.◻

Remark 4.5. Above, we provided a dictionary between the normalized volume of a valuation and the normalized multiplicity of a graded sequence of ideals. The normalized multiplicity also extends to a functional on the set of (formal) plurisubharmonic functions in the sense of [Reference Boucksom, Favre and JonssonBFJ08]. In a slightly different setting, similar functionals, arising from non-Archimedean analogues of functionals in Kähler geometry, were explored in [Reference Boucksom, Hisamoto and JonssonBHJ16].

4.1 Normalized volume minimizers

Proposition 4.6. If there exists a graded sequence of $\mathfrak{m}_{x}$ -primary ideals $\tilde{\mathfrak{a}}_{\bullet }$ such that

$$\begin{eqnarray}\operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })^{n}\operatorname{e}(\tilde{\mathfrak{a}}_{\bullet })=\inf _{\mathfrak{a}_{\bullet }\,\mathfrak{m}_{x}\text{-}\text{primary}}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet }),\end{eqnarray}$$

then there exists $v^{\ast }\in \operatorname{Val}_{X,x}$ that is a minimizer of $\widehat{\operatorname{vol}}_{X,x}$ . Furthermore, if there exists an $\mathfrak{m}_{x}$ -primary ideal $\tilde{\mathfrak{a}}$ such that

$$\begin{eqnarray}\operatorname{lct}(\tilde{\mathfrak{a}})^{n}\operatorname{e}(\tilde{\mathfrak{a}})=\inf _{\mathfrak{a}\,\mathfrak{m}_{x}\text{-}\text{primary}}\operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a}),\end{eqnarray}$$

then we may choose $v^{\ast }$ to be divisorial.

Proof. Assume there exists such a graded sequence $\tilde{\mathfrak{a}}_{\bullet }$ . By Theorem B.1, we may choose a valuation $v^{\ast }$ that computes $\operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })$ . By Lemma 4.4,

$$\begin{eqnarray}\widehat{\operatorname{vol}}(v^{\ast })\leqslant \operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })^{n}\operatorname{e}(\tilde{\mathfrak{a}}_{\bullet }).\end{eqnarray}$$

By Proposition 4.3, the result follows.

When there exits such an ideal $\tilde{\mathfrak{a}}$ , the same argument shows that if $v^{\ast }$ computes $\operatorname{lct}(\tilde{\mathfrak{a}})$ , then $v^{\ast }$ is our desired valuation. Furthermore, we may choose $v^{\ast }$ divisorial.◻

Lemma 4.7. If $v^{\ast }$ is a minimizer of $\widehat{\operatorname{vol}}_{X,x}$ , then

$$\begin{eqnarray}A_{X}(v^{\ast })\leqslant \frac{A_{X}(w)}{w(\mathfrak{a}_{\bullet }(v^{\ast }))}\end{eqnarray}$$

for all $w\in \operatorname{Val}_{X,x}$ . Furthermore, equality holds if and only if $w=\unicode[STIX]{x1D706}v^{\ast }$ for some $\unicode[STIX]{x1D706}\in \mathbf{R}_{{>}0}$ .

Remark 4.8. The above technical statement can be restated as follows. If $v^{\ast }$ is a normalized volume minimizer, then $v^{\ast }$ computes $\operatorname{lct}(\mathfrak{a}_{\bullet }(v^{\ast }))$ and $v^{\ast }$ is the only valuation (up to scaling) that computes $\operatorname{lct}(\mathfrak{a}_{\bullet }(v^{\ast }))$ .

A conjecture of Jonsson and Mustaţă states that valuations computing log canonical thresholds of graded sequences on smooth varieties are always quasimonomial [Reference Jonsson and MustaţǎJM12, Conjecture B]. Their conjecture in the klt case implies [Reference LiLi15, Conjecture 6.1.3], which says that normalized volume minimizers are quasimonomial.

Proof. We fix $w\in \operatorname{Val}_{X,x}$ and rescale $w$ so that $w(\mathfrak{a}_{\bullet }(v^{\ast }))=1$ . Thus, we are reduced to showing that $A_{X}(v^{\ast })\leqslant A_{X}(w)$ and equality holds if and only if $w=v^{\ast }$ .

By definition, we have

$$\begin{eqnarray}1=w(\mathfrak{a}_{\bullet }(v^{\ast })):=\inf _{m\geqslant 0}\frac{w(\mathfrak{a}_{m}(v^{\ast }))}{m},\end{eqnarray}$$

and, thus, $w(\mathfrak{a}_{m}(v^{\ast }))\geqslant m$ . The latter implies that $\mathfrak{a}_{m}(v^{\ast })\subseteq \mathfrak{a}_{m}(w)$ , so

$$\begin{eqnarray}\operatorname{vol}(w)\leqslant \operatorname{vol}(v^{\ast }).\end{eqnarray}$$

If $A_{X}(w)<A_{X}(v^{\ast })$ , then

$$\begin{eqnarray}A_{X}(w)^{n}\operatorname{vol}(w)<A_{X}(v^{\ast })^{n}\operatorname{vol}(v^{\ast })\end{eqnarray}$$

and this would contradict our assumption on $v^{\ast }$ . Furthermore, if $A(v^{\ast })=A(w)$ , then we must have that $\operatorname{vol}(v^{\ast })=\operatorname{vol}(w)$ . Since $v^{\ast }\leqslant w$ and $\operatorname{vol}(v^{\ast })=\operatorname{vol}(w)$ , then $v^{\ast }=w$ by [Reference Li and XuLX16, Proposition 2.12].◻

Proposition 4.9. Let $v^{\ast }\in \operatorname{Val}_{X,x}$ be a minimizer of $\widehat{\operatorname{vol}}_{X,x}$ . If $v^{\ast }=\operatorname{ord}_{E}$ , where $E$ is a prime divisor on a normal variety which is proper and birational over  $X$ , then we have the following.

  1. (a) The graded ${\mathcal{O}}_{X}$ -algebra ${\mathcal{O}}_{X}\oplus \mathfrak{a}_{1}(v^{\ast })\oplus \mathfrak{a}_{2}(v^{\ast })\oplus \cdots \,$ is finitely generated.

  2. (b) The valuation $v^{\ast }$ corresponds to a Kollar component (see [Reference Li and XuLX16]).

  3. (c) The number $\widehat{\operatorname{vol}}(v^{\ast })$ is rational.

The previous proposition was independently observed in [Reference Li and XuLX16, Theorem 1.5]. In fact, prior to our contribution, the original draft of [Reference Li and XuLX16] proved that if (a) holds then (b) holds. Our argument is different from that of [Reference Li and XuLX16].

Proof. By Lemma 4.7, it follows that $\operatorname{lct}(\mathfrak{a}_{\bullet }(v^{\ast }))=A(v^{\ast })$ . The finite generation of the desired ${\mathcal{O}}_{X}$ -algebra is a consequence of [Reference BlumBlu16, Theorem 1.4.1]. Additionally, the second sentence of Lemma 4.7 allows us to apply [Reference BlumBlu16, Proposition 4.4]. Thus, $v^{\ast }$ corresponds to a Kollar component.

To show that $\widehat{\operatorname{vol}}(v^{\ast })$ is rational, we note that the finite generation statement of (a) implies there exists $N>0$ so that $\mathfrak{a}_{mN}(v^{\ast })=(\mathfrak{a}_{N}(v^{\ast }))^{m}$ for all $m\in \mathbf{N}$ [Reference GrothendieckGro61, Lemma II.2.1.6.v]. By Lemma 4.4,

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet }(v^{\ast }))^{n}\operatorname{e}(\mathfrak{a}_{\bullet }(v^{\ast }))\leqslant \widehat{\operatorname{vol}}(v^{\ast }).\end{eqnarray}$$

Lemma 4.1 implies

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet }(v^{\ast }))^{n}\operatorname{e}(\mathfrak{a}_{\bullet }(v^{\ast }))=\operatorname{lct}(\mathfrak{a}_{N})^{n}\operatorname{e}(\mathfrak{a}_{N}),\end{eqnarray}$$

and the latter is a rational number. ◻

5 Limit points of collections of graded sequences of ideals

In this section we construct a space that parameterizes graded sequences of ideals on a fixed variety  $X$ . We use this parameter space to find ‘limit points’ of a collection of graded sequences of ideals on  $X$ . The ideas behind this construction arise from the work of de Fernex and Mustaţă [Reference de Fernex and MustaţăFM09], Kollár [Reference KollárKol08], and de Fernex et al. [Reference de Fernex, Ein and MustaţăFEM10, Reference de Fernex, Ein and MustaţăFEM11].

Before explaining our construction, we set our notation. We fix an affine variety $X=\operatorname{Spec}A$ , where $A=k[x_{1},\ldots ,x_{r}]/\mathfrak{p}$ . Let $\unicode[STIX]{x1D711}$ denote the map

$$\begin{eqnarray}R=k[x_{1},\ldots ,x_{r}]\overset{\unicode[STIX]{x1D711}}{\longrightarrow }A=k[x_{1},\ldots ,x_{r}]/\mathfrak{p}.\end{eqnarray}$$

We set $\mathfrak{m}_{R}:=(x_{1},\ldots ,x_{r})$ and assume that $\mathfrak{p}\subset \mathfrak{m}_{R}$ . Thus, $\mathfrak{m}_{A}=\unicode[STIX]{x1D711}(\mathfrak{m}_{R})$ is a maximal ideal of  $A$ .

5.1 Parameterizing ideals

We fix an integer $d>0$ and seek to parameterize ideals $\mathfrak{a}\subset A$ containing $\mathfrak{m}_{A}^{d}$ and contained in $\mathfrak{m}_{A}$ . Since $\mathfrak{m}_{R}^{d}\subseteq \unicode[STIX]{x1D711}^{-1}(\mathfrak{m}_{A}^{d})$ , such an ideal $\mathfrak{a}\subseteq A$ can be generated by $\mathfrak{m}_{A}^{d}$ and images of polynomials from $R$ of $\deg <d$ . Since there are $n_{d}=\binom{r+d-1}{r}-1$ monomials of positive degree less than $d$ in  $R$ , any such ideal $\mathfrak{m}_{A}^{d}\subseteq \mathfrak{a}\subseteq \mathfrak{m}_{A}$ can be generated by $\mathfrak{m}_{A}^{d}$ and the image of $n_{d}$ linear combinations of monomials. After setting $N_{d}=n_{d}^{2}$ , we get a map

$$\begin{eqnarray}\{k\text{-}\text{valued points of}~\mathbb{A}^{N_{d}}\}\longrightarrow \{\text{ideals}~\mathfrak{a}\subseteq A~\text{s.t.}~\mathfrak{m}_{A}^{d}\subseteq \mathfrak{a}\subseteq \mathfrak{m}_{A}\},\end{eqnarray}$$

where $\mathbb{A}^{N_{d}}$ parameterizes coefficients and generators of such ideals. The above map is surjective, but not injective (generators of an ideal are not unique). Additionally, we have a universal ideal $\mathscr{A}\subset {\mathcal{O}}_{X\times \mathbb{A}^{N_{d}}}$ such that $\mathscr{A}$ restricted to the fibers of $p:X\times \mathbb{A}^{N_{d}}\rightarrow \mathbb{A}^{N_{d}}$ give us our $\mathfrak{m}_{A}$ -primary ideals.

The construction in the previous paragraph follows the exposition of [Reference de Fernex and MustaţăFM09, § 3].

5.2 Parameterizing graded sequences of ideals

We proceed to parameterize graded sequences of ideals $\mathfrak{a}_{\bullet }$ of $A$ satisfying

We set

$$\begin{eqnarray}H_{d}:=\mathbb{A}^{N_{1}}\times \cdots \times \mathbb{A}^{N_{d}},\end{eqnarray}$$

where $N_{i}$ is chosen as in the previous section. For $d>c$ , let $\unicode[STIX]{x1D70B}_{d,c}:H_{d}\rightarrow H_{c}$ denote the natural projection maps. Our desired object is the following projective limit

$$\begin{eqnarray}H=\varprojlim H_{d}.\end{eqnarray}$$

For $d>0$ , let $\unicode[STIX]{x1D70B}_{d}:H\rightarrow H_{d}$ denote the natural map.

Note that the above projective limit exists in the category of schemes, since the maps in our directed system are all affine morphisms. Indeed, $H$ is isomorphic to an infinite-dimensional affine space.

Since a $k$ -valued point of $H$ is simply a sequence of $k$ -valued points of $\mathbb{A}^{N_{d}}$ for all $d\in \mathbf{N}$ , we have a surjection

$$\begin{eqnarray}\{k\text{-}\text{valued points of}~H\}\longrightarrow \{\text{sequences of ideals}~\mathfrak{b}_{\bullet }~\text{of}~A~\text{satisfying}~(\dagger )\}.\end{eqnarray}$$

Note that the sequences of ideals on the right-hand side are not necessarily graded.

Given a sequence of ideals $\mathfrak{b}_{\bullet }$ , we can construct a graded sequence $\mathfrak{a}_{\bullet }$ inductively by setting $\mathfrak{a}_{1}:=\mathfrak{b}_{1}$ and

$$\begin{eqnarray}\mathfrak{a}_{q}:=\mathfrak{b}_{q}+\mathop{\sum }_{m+n=q}\mathfrak{a}_{m}\cdot \mathfrak{a}_{n}.\end{eqnarray}$$

If $\mathfrak{b}_{\bullet }$ was graded to begin with, then $\mathfrak{a}_{\bullet }=\mathfrak{b}_{\bullet }$ . By the construction, it is clear that $\mathfrak{a}_{m}\cdot \mathfrak{a}_{n}\subseteq \mathfrak{a}_{m+n}$ . Thus, we have our desired map

$$\begin{eqnarray}\{k\text{-}\text{valued points of}~H\}\longrightarrow \{\text{graded sequences of ideals}~\mathfrak{a}_{\bullet }~\text{of}~A~\text{satisfying}~(\dagger )\}.\end{eqnarray}$$

Additionally, we have a universal graded sequence of ideals $\mathscr{A}_{\bullet }=\{\mathscr{A}_{m}\}_{m\in \mathbf{N}}$ on $X\times H$ . We will often abuse notation and refer to similarly defined ideals $\mathscr{A}_{1},\ldots ,\mathscr{A}_{d}$ on $X\times H_{d}$ .

The following technical lemma will be useful in the next proposition. The proof of the lemma relies on the fact that every descending sequence of nonempty constructible subsets of a variety over an uncountable field has nonempty intersection.

Lemma 5.1. If $\{W_{d}\}_{d\in \mathbf{N}}$ is a collection of nonempty subsets of $H_{d}$ such that

  1. (a) $W_{d}\subset H_{d}$ is a constructible, and

  2. (b)

    $$\begin{eqnarray}W_{d+1}\subseteq \unicode[STIX]{x1D70B}_{d+1,d}^{-1}(W_{d})\end{eqnarray}$$
    for each $d\in \mathbf{Z}_{{>}0}$ ,

then there exists a $k$ -valued point in

$$\begin{eqnarray}\mathop{\bigcap }_{d\in \mathbf{N}}\unicode[STIX]{x1D70B}_{d}^{-1}(W_{d}).\end{eqnarray}$$

Proof. Note that a $k$ -valued point in the above intersection is equivalent to a sequence of closed points $\{x_{d}\in W_{d}\}_{d\in \mathbf{N}}$ such that $\unicode[STIX]{x1D70B}_{d+1,d}(x_{d+1})=x_{d}$ . We proceed to construct such a sequence.

We first look to find a candidate for $x_{1}$ . Assumption (b) implies

$$\begin{eqnarray}W_{1}\supseteq \unicode[STIX]{x1D70B}_{2,1}(W_{2})\supseteq \unicode[STIX]{x1D70B}_{3,1}(W_{3})\supseteq \cdots\end{eqnarray}$$

is a descending sequence of nonempty sets. Note that $W_{1}$ is constructible by assumption and so are $\unicode[STIX]{x1D70B}_{d,1}(W_{d})$ for all $d$ by Chevalley’s theorem [Reference HartshorneHar77, Exercise II.3.9]. Thus,

$$\begin{eqnarray}W_{1}\cap \unicode[STIX]{x1D70B}_{2,1}(W_{2})\cap \unicode[STIX]{x1D70B}_{3,1}(W_{3})\cap \cdots\end{eqnarray}$$

is nonempty and we may choose a point $x_{1}$ lying in the set.

Next, we look at

$$\begin{eqnarray}W_{2}\cap \unicode[STIX]{x1D70B}_{2,1}^{-1}(x_{1})\supseteq \unicode[STIX]{x1D70B}_{3,2}(W_{3})\cap \unicode[STIX]{x1D70B}_{2,1}^{-1}(x_{1})\supseteq \unicode[STIX]{x1D70B}_{4,2}(W_{4})\cap \unicode[STIX]{x1D70B}_{2,1}^{-1}(x_{1}),\end{eqnarray}$$

and note that for $d\geqslant 2$ each $\unicode[STIX]{x1D70B}_{d,2}(W_{d})\cap \unicode[STIX]{x1D70B}_{2,1}^{-1}(x_{1})$ is nonempty by our choice of $x_{1}$ . By the same argument as before, we see

$$\begin{eqnarray}\unicode[STIX]{x1D70B}_{2,1}^{-1}(x_{1})\cap W_{2}\cap \unicode[STIX]{x1D70B}_{3,2}(W_{3})\cap \unicode[STIX]{x1D70B}_{4,2}(W_{4})\cap \cdots\end{eqnarray}$$

is nonempty and contains a closed point $x_{2}$ . Continuing in this manner, we construct the desired sequence.◻

5.3 Finding limit points

The proof of the following proposition relies on the previous construction of a space that parameterizes graded sequences of ideals. The proof is inspired by arguments in [Reference KollárKol08, Reference de Fernex, Ein and MustaţăFEM10, Reference de Fernex, Ein and MustaţăFEM11].

Proposition 5.2. Let $X$ be a klt variety and $x\in X$ a closed point. Assume there exists a collection of graded sequences of $\mathfrak{m}_{x}$ -primary ideals $\{\mathfrak{a}_{\bullet }^{(i)}\}_{i\in \mathbf{N}}$ and $\unicode[STIX]{x1D706}\in \mathbf{R}$ such that the following hold.

  1. (a) (Convergence from above) For every $\unicode[STIX]{x1D716}>0$ , there exists positive constants $M,N$ so that

    $$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{m}^{(i)})^{n}\operatorname{e}(\mathfrak{a}_{m}^{(i)})\leqslant \unicode[STIX]{x1D706}+\unicode[STIX]{x1D716}\end{eqnarray}$$
    for all $m\geqslant M$ and $i\geqslant N$ .
  2. (b) (Boundedness from below) For each $m,i\in \mathbf{N}$ , we have

    $$\begin{eqnarray}\mathfrak{m}_{x}^{m}\subseteq \mathfrak{a}_{m}^{(i)}.\end{eqnarray}$$
  3. (c) (Boundedness from above) There exists $\unicode[STIX]{x1D6FF}>0$ such that

    $$\begin{eqnarray}\mathfrak{a}_{m}^{(i)}\subseteq \mathfrak{m}^{\lfloor m\unicode[STIX]{x1D6FF}\rfloor }\end{eqnarray}$$
    for all $m,i\in \mathbf{N}$ .

Then, there exists a graded sequence of $\mathfrak{m}_{x}$ -primary ideals $\tilde{\mathfrak{a}}_{\bullet }$ on $X$ such that

$$\begin{eqnarray}\operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })^{n}\operatorname{e}(\tilde{\mathfrak{a}}_{\bullet })\leqslant \unicode[STIX]{x1D706}.\end{eqnarray}$$

Proof. It is sufficient to consider the case when $X$ is affine. Thus, we may assume that $X=\operatorname{Spec}A$ and $A=k[x_{1},\ldots ,x_{r}]/\mathfrak{p}$ as in the beginning of this section. In addition, we may assume that $x\in X$ corresponds to the maximal ideal $\mathfrak{m}_{A}$ . We recall that § 5.2 constructs a variety $H$ parameterizing graded sequences of ideals on $X$ satisfying (†). In addition, we have finite-dimensional truncations $H_{d}$ that parameterize the first $d$ elements of such a sequence.

Each graded sequence $\mathfrak{a}_{\bullet }^{(i)}$ satisfies (†) by assumptions (b) and (c). Thus, we may choose a point $p_{i}\in H$ corresponding to $\mathfrak{a}_{\bullet }^{(i)}$ . Note that $\unicode[STIX]{x1D70B}_{d}(p_{i})\in H_{d}$ corresponds to the first $d$ -terms of $\mathfrak{a}_{\bullet }^{(i)}$ .

Claim 1. We may choose infinite subsets $\mathbf{N}\supset I_{1}\supset I_{2}\supset \cdots \,$ and set

$$\begin{eqnarray}Z_{d}:=\overline{\{\unicode[STIX]{x1D70B}_{d}(p_{i})\mid i\in I_{d}\}}\end{eqnarray}$$

such that (∗∗) holds:

To prove Claim 1, we construct such a sequence inductively. First, we set $I_{1}=\mathbf{N}$ . Since $H_{1}\simeq \mathbb{A}^{0}$ is a point, (∗∗) is trivially satisfied for $d=1$ . After having chosen $I_{d}$ , choose $I_{d+1}\subset I_{d}$ so that (∗∗) is satisfied for $Z_{d+1}$ . By the Noetherianity of $H_{d}$ , such a choice is possible.

Claim 2. We have the inclusion $Z_{d+1}\subseteq \unicode[STIX]{x1D70B}_{d+1,d}^{-1}(Z_{d})$ for all $d\geqslant 1$ .

The proof of Claim 2 follows from the definition of $Z_{d}$ . Since $\unicode[STIX]{x1D70B}_{d}(p_{i})\in Z_{d}$ for all $i\in I_{d}$ and $I_{d}\supseteq I_{d+1}$ , it follows that $\unicode[STIX]{x1D70B}_{d+1,d}^{-1}(Z_{d})$ is a closed set containing $\unicode[STIX]{x1D70B}_{d+1}(p_{i})$ for $i\in I_{d+1}$ . The closure of the latter set of points is precisely $Z_{d+1}$ .

Claim 3. If $p\in Z_{d}$ is a closed point, we have $\mathscr{A}_{d}|_{p}\subseteq \mathfrak{m}^{\lfloor d\unicode[STIX]{x1D6FF}\rfloor }$ .

We now prove Claim 3. The set $\{p\in H_{d}\mid {\mathcal{A}}_{d}|_{p}\subseteq \mathfrak{m}^{\lfloor d\unicode[STIX]{x1D6FF}\rfloor }\}$ is a closed in $H_{d}$ . By assumption (c), $\unicode[STIX]{x1D70B}_{d}(p_{i})$ lies in the above closed set for all $i\in I_{d}$ . Thus, $Z_{d}\subseteq \{p\in H_{d}\mid {\mathcal{A}}_{d}|_{p}\subseteq \mathfrak{m}^{\lfloor d\unicode[STIX]{x1D6FF}\rfloor }\}$ .

We now return to the proof of the proposition. We look at the normalized multiplicity of the ideals parameterized by  $Z_{d}$ . By Propositions A.1 and A.2, for each  $d$ , we may choose a nonempty open set $U_{d}\subseteq Z_{d}$ such that

$$\begin{eqnarray}\operatorname{lct}(\mathscr{A}_{d}|_{p})^{n}\operatorname{e}(\mathscr{A}_{d}|_{p})=\unicode[STIX]{x1D706}_{d}\end{eqnarray}$$

is constant for $p\in U_{d}$ . Set

$$\begin{eqnarray}I_{d}^{\circ }=\{i\in I_{d}\mid \unicode[STIX]{x1D70B}_{d}(p_{i})\in U_{d}\}\subseteq I_{d},\end{eqnarray}$$

and note that $I_{d}\setminus I_{d}^{\circ }$ is finite. If this was not the case, then (∗∗) would not hold.

The finiteness of $I_{d}\setminus I_{d}^{\circ }$ has two consequences. First,

$$\begin{eqnarray}\lim _{d\rightarrow \infty }\sup \unicode[STIX]{x1D706}_{d}\leqslant \unicode[STIX]{x1D706},\end{eqnarray}$$

since $\unicode[STIX]{x1D70B}_{d}(p_{i})\in U_{d}$ for all $i\in I_{d}^{\circ }$ and assumption (a) of this proposition. Second, since $I_{d+1}\subset I_{d}$ for $d\in \mathbf{N}$ , we have

$$\begin{eqnarray}I_{d}^{\circ }\cap I_{d-1}^{\circ }\cdots \cap I_{1}^{\circ }\neq \emptyset .\end{eqnarray}$$

Claim 4. There exists a $k$ -valued point $\tilde{p}\in H$ such that $\unicode[STIX]{x1D70B}_{d}(\tilde{p})\in U_{d}$ for all $d\in \mathbf{Z}_{{>}0}$ .

Proving this claim will complete the proof. Indeed, a point $\tilde{p}\in H$ corresponds to a graded sequence of $\mathfrak{m}_{x}$ -primary ideals $\tilde{\mathfrak{a}}_{\bullet }$ . Since $\unicode[STIX]{x1D70B}_{d}(\tilde{p})\in U_{d}$ , we will have $\operatorname{lct}(\tilde{\mathfrak{a}}_{d})^{n}\operatorname{e}(\tilde{\mathfrak{a}}_{d})=\unicode[STIX]{x1D706}_{d}$ . In addition, Claim 2 implies and $\tilde{\mathfrak{a}}_{d}\subset \mathfrak{m}_{x}^{\lfloor d\unicode[STIX]{x1D6FF}\rfloor }$ for all $d\in \mathbf{Z}_{{>}0}$ . Thus,

$$\begin{eqnarray}\operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })^{n}\operatorname{e}(\tilde{\mathfrak{a}}_{\bullet })=\lim _{d\rightarrow \infty }\operatorname{lct}(\tilde{\mathfrak{a}}_{d})\operatorname{e}(\tilde{\mathfrak{a}}_{d})\leqslant \lim _{d\rightarrow \infty }\sup \unicode[STIX]{x1D706}_{d}\leqslant \unicode[STIX]{x1D706},\end{eqnarray}$$

and the proof will be complete.

We are left to prove Claim 4. In order to do so, we will apply Lemma 5.1 to find such a point $\tilde{p}\in H$ . First, we define constructible sets $W_{d}\subseteq H_{d}$ inductively. Set $W_{1}=U_{1}$ . After having chosen $W_{d}$ , set $W_{d+1}=\unicode[STIX]{x1D70B}_{d+1,d}^{-1}(W_{d})\cap U_{d}$ . For each $d\in \mathbf{N}$ we have the following:

  • $W_{d}$ is open in $Z_{d}$ and, thus, constructible in $H_{d}$ ;

  • $W_{d}$ is nonempty, since $W_{d}$ contains $\unicode[STIX]{x1D70B}_{d}(p_{i})$ for all $i\in I_{d}^{\circ }\cap I_{d-1}^{\circ }\cap \cdots \cap I_{1}^{\circ }$ , which is nonempty.

By Lemma 5.1, there exists a $k$ -valued point $\tilde{p}\in H$ such that $\unicode[STIX]{x1D70B}_{d}(\tilde{p})\in W_{d}\subset U_{d}$ for all $d\in \mathbf{Z}_{{>}0}$ . This completes the proof.◻

Remark 5.3. In the previous proof, we construct a graded sequence of ideal $\tilde{\mathfrak{a}}_{\bullet }$ based on a collection of graded sequences $\{\mathfrak{a}_{\bullet }^{(i)}\}_{i\in \mathbf{N}}$ . While the construction of $\tilde{\mathfrak{a}}_{\bullet }$ is inspired by past constructions of generic limits, $\tilde{\mathfrak{a}}_{\bullet }$ is not a generic limit in the sense of [Reference KollárKol08, 28].

We construct the precise analog as follows. We set

$$\begin{eqnarray}Z:=\mathop{\bigcap }_{d}\unicode[STIX]{x1D70B}_{d}^{-1}(Z_{d})\subseteq H,\end{eqnarray}$$

with $Z_{d}$ defined in the previous proof. The generic point of $Z$ gives a map $\operatorname{Spec}(K(Z))\rightarrow H$ , where $K(Z)$ is the function field of $Z$ . Thus, we get a graded sequence of ideals $\widehat{\mathfrak{a}}_{\bullet }$ on $X_{K(Z)}$ , the base change of $X$ by  $K(Z)$ .

In the previous proof, we wanted to construct a graded sequence on  $X$ , not a base change of  $X$ . Thus, $\tilde{\mathfrak{a}}_{\bullet }$ was chosen to be a graded sequence corresponding to a very general point in  $Z$ .

6 Proof of the Main Theorem

In this section we prove the Main Theorem. To prove the theorem, we apply the construction from § 5.

Proof of the Main Theorem.

We fix a klt variety $X$ and a closed point $x\in X$ . Next, we choose a sequence of valuations $(v_{i})_{i\in \mathbf{N}}$ in $\operatorname{Val}_{X,x}$ such that

$$\begin{eqnarray}\lim _{i}\widehat{\operatorname{vol}}(v_{i})=\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v).\end{eqnarray}$$

Additionally, after scaling our valuations, we may assume that $v_{i}(\mathfrak{m}_{x})=1$ for all $i\in \mathbf{N}$ . Note that this implies that $\mathfrak{m}_{x}^{m}\subset \mathfrak{a}_{m}(v_{i})$ for all  $m$ .

We claim that $\{\mathfrak{a}_{\bullet }(v_{i})\}_{i\in \mathbf{N}}$ satisfy the hypotheses of Proposition 5.2 with $\unicode[STIX]{x1D706}=\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)$ . After showing that this is the case, we will have that there exists a graded sequence of $\mathfrak{m}_{x}$ -primary ideals $\tilde{\mathfrak{a}}_{\bullet }$ such that

$$\begin{eqnarray}\operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })^{n}\operatorname{e}(\tilde{\mathfrak{a}}_{\bullet })\leqslant \inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v).\end{eqnarray}$$

By Theorem B.1, there exists a valuation $v^{\ast }\in \operatorname{Val}_{X,x}$ that computes $\operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })$ . Thus,

$$\begin{eqnarray}\widehat{\operatorname{vol}}(v^{\ast })\leqslant \operatorname{lct}(\tilde{\mathfrak{a}}_{\bullet })^{n}\operatorname{e}(\tilde{\mathfrak{a}}_{\bullet })=\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v),\end{eqnarray}$$

where the first inequality follows from Lemma 4.4. Thus, $v^{\ast }$ will be our normalized volume minimizer.

It is left to show that $\{\mathfrak{a}_{\bullet }(v_{i})\}_{i\in \mathbf{N}}$ satisfies the hypotheses of Proposition 5.2. Hypothesis (a) follows from Proposition 6.4, (b) from the assumption that $v_{i}(\mathfrak{m})=1$ for all $i\in \mathbf{N}$ , and (c) from Proposition 6.2.◻

We proceed to prove the two propositions mentioned in the previous paragraph. We emphasize that estimates from [Reference LiLi15] are essential in the proof of the following lemma and propositions.

Lemma 6.1. With the notation above, there exist positive constants $E,B$ such that: (a)  $A_{X}(v_{i})\leqslant E$ ; and (b)  $\operatorname{vol}(v_{i})\leqslant B$ for all $i\in \mathbf{N}$ .

Proof. By [Reference LiLi15, Theorem 3.3], there exists a constant $C$ such that

$$\begin{eqnarray}A_{X}(v)\leqslant C\cdot v(\mathfrak{m})\widehat{\operatorname{vol}}(v)\end{eqnarray}$$

for all $v\in \operatorname{Val}_{X,x}$ . Thus, we may set $E:=C\cdot \sup _{i}\widehat{\operatorname{vol}}(v_{i})<+\infty$ .

The bound on the volume follows from the inclusion $\mathfrak{m}_{x}^{m}\subset \mathfrak{a}_{m}(v_{i})$ for all $m\in \mathbf{N}$ . The inclusion gives that

$$\begin{eqnarray}\operatorname{vol}(v_{i})=\lim _{m\rightarrow \infty }\frac{\operatorname{e}(\mathfrak{a}_{m}(v_{i}))}{m^{n}}\leqslant \lim _{m\rightarrow \infty }\frac{\operatorname{e}(\mathfrak{m}_{x}^{n})}{m^{n}}=\operatorname{e}(\mathfrak{m}_{x}).\Box\end{eqnarray}$$

Proposition 6.2. With the notation above, there exists $\unicode[STIX]{x1D6FF}>0$ such that

$$\begin{eqnarray}\mathfrak{a}_{m}(v_{i})\subseteq \mathfrak{m}_{x}^{\lfloor \unicode[STIX]{x1D6FF}m\rfloor }\end{eqnarray}$$

for all $m,i\in \mathbf{N}$ .

Remark 6.3. Note that for an ideal $\mathfrak{a}$ , the order of vanishing of $\mathfrak{a}$ along $x$ is defined to be

$$\begin{eqnarray}\operatorname{ord}_{x}(\mathfrak{a}):=\max \{n\mid \mathfrak{a}\subseteq \mathfrak{m}_{x}^{n}\}.\end{eqnarray}$$

To prove the above proposition, it is sufficient to find $\unicode[STIX]{x1D6FF}^{\prime }>0$ such that

$$\begin{eqnarray}\unicode[STIX]{x1D6FF}^{\prime }m\leqslant \operatorname{ord}_{x}(\mathfrak{a}_{m}(v_{i}))\end{eqnarray}$$

for all $m,i\in \mathbf{N}$ .

Proof. By [Reference LiLi15, Proposition 2.3], there exists a constant $C$ such that for all $v\in \operatorname{Val}_{X,x}$ and $f\in {\mathcal{O}}_{X,x}$ ,

$$\begin{eqnarray}v(f)\leqslant C\cdot A_{X}(v)\operatorname{ord}_{x}(f).\end{eqnarray}$$

Thus,

$$\begin{eqnarray}m\leqslant v_{i}(\mathfrak{a}_{m}(v_{i}))\leqslant C\cdot A_{X}(v_{i})\operatorname{ord}_{x}(\mathfrak{a}_{m}(v_{i})),\end{eqnarray}$$

and

$$\begin{eqnarray}\frac{m}{CA_{X}(v_{i})}\leqslant \operatorname{ord}_{x}(\mathfrak{a}_{m}(v_{i})).\end{eqnarray}$$

By Lemma 6.1, there exists a positive constant $E$ such that $A_{X}(v_{i})\leqslant E$ for all  $i$ . We conclude

$$\begin{eqnarray}\frac{m}{C\cdot E}\leqslant \operatorname{ord}_{x}(\mathfrak{a}_{m}(v_{i})).\Box\end{eqnarray}$$

Proposition 6.4. With the notation above, for $\unicode[STIX]{x1D716}>0$ , there exist positive constants $M,N$ such that

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{m}(v_{i}))^{n}\operatorname{e}(\mathfrak{a}_{m}(v_{i}))\leqslant \inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)+\unicode[STIX]{x1D716}.\end{eqnarray}$$

for all $m\geqslant M$ and $i\geqslant N$ .

Proof. Since $\widehat{\operatorname{vol}}(v_{i})$ converges to $\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)$ as $i\rightarrow \infty$ , we may choose $N$ so that

$$\begin{eqnarray}\widehat{\operatorname{vol}}(v_{i})\leqslant \inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)+\unicode[STIX]{x1D716}/2\end{eqnarray}$$

for all $i\geqslant N$ . By Lemma 6.1, we have $E:=\sup A_{X}(v_{i})<\infty$ . Additionally, Lemma 6.1 allows us to apply Proposition 3.7 to find a constant $M$ so that

$$\begin{eqnarray}\operatorname{e}(\mathfrak{a}_{m}(v_{i}))\leqslant \operatorname{vol}(v_{i})+\unicode[STIX]{x1D716}/(2E^{n})\end{eqnarray}$$

for all integers $m\geqslant M$ . Thus,

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{m}(v_{i}))^{n}\operatorname{e}(\mathfrak{a}_{m}(v_{i}))\leqslant A_{X}(v_{i})^{n}(\operatorname{vol}(v_{i})+\unicode[STIX]{x1D716}/(2E^{n}))\leqslant \widehat{\operatorname{vol}}(v_{i})+\unicode[STIX]{x1D716}/2.\end{eqnarray}$$

for all $m\geqslant M$ and $i\in \mathbf{N}$ . We conclude that

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{m}(v_{i}))^{n}\operatorname{e}(\mathfrak{a}_{m}(v_{i}))\leqslant \widehat{\operatorname{vol}}(v_{i})+\unicode[STIX]{x1D716}/2\leqslant \inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)+\unicode[STIX]{x1D716}\end{eqnarray}$$

for all $m\geqslant M$ and $i\geqslant N$ .◻

7 The normalized volume over a log pair

The normalized volume function has been studied in the setting of log pairs [Reference Li and XuLX16, Reference Li and LiuLL16]. We explain that the arguments in this paper extend to the setting where $(X,\unicode[STIX]{x1D6E5})$ is a klt pair.

Recall that $(X,\unicode[STIX]{x1D6E5})$ is a log pair if $X$ is a normal variety and $\unicode[STIX]{x1D6E5}$ is an effective $\mathbf{Q}$ -divisor on $X$ such that $K_{X}+\unicode[STIX]{x1D6E5}$ is $\mathbf{Q}$ -Cartier.

7.1 Log discrepancies

If $(X,\unicode[STIX]{x1D6E5})$ is a log pair, the log discrepancy function $A_{(X,\unicode[STIX]{x1D6E5})}:\operatorname{Val}_{X}\rightarrow \mathbf{R}\cup \{+\infty \}$ is defined as follows. If $Y\rightarrow X$ is a proper birational morphism with $Y$ normal, and $E\subset Y$ a prime divisor, then the log discrepancy of $\operatorname{ord}_{E}$ over $(X,\unicode[STIX]{x1D6E5})$ is

$$\begin{eqnarray}A_{(X,\unicode[STIX]{x1D6E5})}(\operatorname{ord}_{E}):=1+(\text{coefficient of}~E~\text{in}~K_{Y}-f^{\ast }(K_{X}+\unicode[STIX]{x1D6E5})).\end{eqnarray}$$

As in [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15], $A_{(X,\unicode[STIX]{x1D6E5})}$ can then be extended to $\operatorname{Val}_{X}$ . If $(X,\unicode[STIX]{x1D6E5})$ is a log pair, we say $(X,\unicode[STIX]{x1D6E5})$ is klt if $A_{X}(v)>0$ for all divisorial valuations $v\in \operatorname{Val}_{X}$ .

7.2 Normalized volume minimizers

If $(X,\unicode[STIX]{x1D6E5})$ is a klt log pair and $x\in X$ is a closed point, the normalized volume of a valuation $v\in \operatorname{Val}_{X,x}$ over the pair $(X,\unicode[STIX]{x1D6E5})$ is defined to be

$$\begin{eqnarray}\widehat{\operatorname{vol}}_{(X,\unicode[STIX]{x1D6E5}),x}(v):=A_{(X,\unicode[STIX]{x1D6E5})}(v)^{n}\operatorname{vol}(v).\end{eqnarray}$$

We claim that if $(X,\unicode[STIX]{x1D6E5})$ is a klt pair and $x\in X$ is a closed point, then there exists a minimizer of $\widehat{\operatorname{vol}}_{(X,\unicode[STIX]{x1D6E5}),x}$ .

The main subtlety in extending our arguments to the log setting is in extending Theorem 3.6, which is a consequence of the subadditivity theorem. Takagi proved the following subadditivity theorem for log pairs.

Theorem 7.1 (Takagi [Reference TakagiTak13]).

Let $(X,\unicode[STIX]{x1D6E5})$ be a log pair, $\mathfrak{a},\mathfrak{b}$ ideals on $X$ , and $s,t\in \mathbf{Q}_{{\geqslant}0}$ . For $r\in \mathbf{Z}_{{>}0}$ so that $r(K_{X}+\unicode[STIX]{x1D6E5})$ , we have

$$\begin{eqnarray}\operatorname{Jac}_{X}\cdot {\mathcal{J}}((X,\unicode[STIX]{x1D6E5}),\mathfrak{a}^{s}\mathfrak{b}^{s}{\mathcal{O}}_{X}(-r\unicode[STIX]{x1D6E5})^{1/r})\subseteq {\mathcal{J}}((X,\unicode[STIX]{x1D6E5}),\mathfrak{a}^{s}){\mathcal{J}}((X,\unicode[STIX]{x1D6E5}),\mathfrak{b}^{t}).\end{eqnarray}$$

Takagi’s result implies the following generalization of Theorem 3.6 for log pairs. The remaining arguments in the paper extend to this setting.

Theorem 7.2. Let $(X,\unicode[STIX]{x1D6E5})$ be a log pair and $r\in \mathbf{Z}_{{>}0}$ such that $r(K_{X}+\unicode[STIX]{x1D6E5})$ is Carter. If $v\in \operatorname{Val}_{X}$ satisfies $A_{X}(v)<+\infty$ , then

$$\begin{eqnarray}(\operatorname{Jac}_{X}\cdot {\mathcal{O}}_{X}(-r\unicode[STIX]{x1D6E5}))^{\ell -1}\cdot \mathfrak{a}_{m}^{\ell }\subseteq (\operatorname{Jac}_{X}{\mathcal{O}}_{X}(-r\unicode[STIX]{x1D6E5}))^{\ell -1}\cdot \mathfrak{a}_{m\ell }\subseteq \mathfrak{a}_{m-e}^{\ell }\end{eqnarray}$$

for every $m\geqslant e$ and $\ell \in \mathbf{Z}_{{>}0}$ , where $\mathfrak{a}_{\bullet }:=\mathfrak{a}_{\bullet }(v)$ and $e:=\lceil A_{(X,\unicode[STIX]{x1D6E5})}(v)\rceil$ .

8 The toric setting

We use the notation of [Reference FultonFul93] for toric varieties. Let $N$ be a free abelian group of rank $n\geqslant 1$ and $M=N^{\ast }$ its dual. We write $N_{\mathbf{R}}:=N\otimes \mathbf{R}$ and $M_{\mathbf{R}}:=M\otimes \mathbf{R}$ . There is a canonical pairing

$$\begin{eqnarray}\langle \,,\,\rangle :N_{\mathbf{R}}\times M_{\mathbf{R}}\rightarrow \mathbf{R}.\end{eqnarray}$$

We say that an element $u\in N$ is primitive if $u$ cannot be written as $u=au^{\prime }$ for $a\in \mathbf{Z}_{{>}1}$ and $u\in N$ .

Fix a maximal dimension, strongly convex, rational, polyhedral cone $\unicode[STIX]{x1D70E}\subset N_{\mathbf{R}}$ . From the cone $\unicode[STIX]{x1D70E}$ , we get a toric variety $X_{\unicode[STIX]{x1D70E}}=\operatorname{Spec}R_{\unicode[STIX]{x1D70E}}$ , where $R_{\unicode[STIX]{x1D70E}}=k[\unicode[STIX]{x1D70E}^{\vee }\cap M]$ . Let $x\in X_{\unicode[STIX]{x1D70E}}$ denote the unique torus invariant point of $X_{\unicode[STIX]{x1D70E}}$ . We write $u_{1},\ldots ,u_{r}\in N$ for the primitive lattice points of $N$ that generate the one-dimensional faces of $\unicode[STIX]{x1D70E}$ . Each $u_{i}$ corresponds to a toric invariant divisor $D_{i}$ on $X_{\unicode[STIX]{x1D70E}}$ . Since the canonical divisor is given by $K_{X_{\unicode[STIX]{x1D70E}}}=-\!\sum D_{i}$ , the divisor $K_{X}$ is $\mathbf{Q}$ -Cartier if and only if there exists $w\in M\otimes \mathbf{Q}\subset M_{\mathbf{R}}$ such that $\langle u_{i},w\rangle =1$ for $i=1,\ldots ,r$ .

Given $u\in \unicode[STIX]{x1D70E}$ , we obtain a toric valuation $v_{u}\in \operatorname{Val}_{X}$ defined by

$$\begin{eqnarray}v_{u}\biggl(\mathop{\sum }_{m\in M\cap \unicode[STIX]{x1D70E}^{\vee }}\unicode[STIX]{x1D6FC}_{v}\unicode[STIX]{x1D712}^{v}\biggr)=\min \{\langle u,v\rangle \mid \unicode[STIX]{x1D6FC}_{v}\neq 0\}.\end{eqnarray}$$

If $u\in \unicode[STIX]{x1D70E}^{\vee }\cap N$ is primitive, the valuation $v_{u}$ corresponds to vanishing along a prime divisor on a toric variety proper and birational over $X_{\unicode[STIX]{x1D70E}}$ . For $u\in \unicode[STIX]{x1D70E}$ , $v_{u}$ has center equal to $x$ if and only if $u\in \operatorname{Int}(\unicode[STIX]{x1D70E})$ .

Let $\operatorname{Val}_{X_{\unicode[STIX]{x1D70E}},x}^{\text{toric}}\subset \operatorname{Val}_{X_{\unicode[STIX]{x1D70E}},x}$ denote the valuations on $X_{\unicode[STIX]{x1D70E}}$ of the form $v_{u}$ for $u\in \operatorname{Int}(\unicode[STIX]{x1D70E})$ . We refer to these valuations as the toric valuations at  $x$ . It is straightforward to compute the normalized volume of such a valuation. Assume $K_{X}$ is $\mathbf{Q}$ -Cartier and $w$ is the unique vector such that $\langle u_{i},w\rangle =1$ for $i=1,\ldots ,s$ . For $u\in \unicode[STIX]{x1D70E}$ , we have

$$\begin{eqnarray}A_{X_{\unicode[STIX]{x1D70E}}}(v_{u})=\langle u,w\rangle .\end{eqnarray}$$

For $u\in \unicode[STIX]{x1D70E}$ and $m\in \mathbf{N}$ , we set $H_{u}(m)=\{v\in M_{\mathbf{R}}\mid \langle u,v\rangle \geqslant m\}$ . Note that

$$\begin{eqnarray}\mathfrak{a}_{m}(v_{u})=(\unicode[STIX]{x1D712}^{v}\mid v\in H_{u}(m)\cap \unicode[STIX]{x1D70E}^{\vee }\cap M).\end{eqnarray}$$

In the case when $u\in \operatorname{Int}(\unicode[STIX]{x1D70E})$ ,

$$\begin{eqnarray}\operatorname{vol}(v_{u})=n!\cdot \operatorname{Vol}(\unicode[STIX]{x1D70E}^{\vee }\setminus H_{u}(1)),\end{eqnarray}$$

where $\operatorname{Vol}$ denotes the Euclidean volume.

8.1 Deformation to the initial ideal

As explained in [Reference EisenbudEis95], when $X_{\unicode[STIX]{x1D70E}}\simeq \mathbb{A}^{n}$ and $I\subset R_{\unicode[STIX]{x1D70E}}$ , there exists a deformation of $I$ to a monomial ideal. A similar argument works in our setting.

Following the approach of [Reference Kaveh and KhovanskiiKK14, § 6], we put a $\mathbf{Z}_{{\geqslant}0}^{n}$ order on the monomials of $R_{\unicode[STIX]{x1D70E}}$ . Fix $y_{1},\ldots ,y_{n}\in N\,\cap \,\unicode[STIX]{x1D70E}$ that are linearly independent in  $M_{\mathbf{R}}$ . Thus, we get an injective map $\unicode[STIX]{x1D70C}:M\rightarrow \mathbf{Z}^{n}$ by sending

$$\begin{eqnarray}v\longmapsto (\langle y_{1},v\rangle ,\ldots ,\langle y_{n},v\rangle ).\end{eqnarray}$$

Since each $y_{i}\in \unicode[STIX]{x1D70E}$ , we have $\unicode[STIX]{x1D70C}(M\cap \unicode[STIX]{x1D70E}^{\vee })\subseteq \mathbf{Z}_{{\geqslant}0}^{n}$ . After putting the lexigraphic order on $\mathbf{Z}_{{\geqslant}0}^{n}$ , we get an order ${>}$ on the monomials of  $R_{\unicode[STIX]{x1D70E}}$ .

An element $f\in R_{\unicode[STIX]{x1D70E}}$ may be written as a sum of scalar multiples of distinct monomials. The initial term of  $f$ , denoted $\operatorname{in}_{{>}}f$ , is the greatest term of $f$ with respect to the order  ${>}$ . For an ideal $I\subset R_{\unicode[STIX]{x1D70E}}$ , the initial ideal of $I$ is

$$\begin{eqnarray}\operatorname{in}_{{>}}I=(\operatorname{in}_{{>}}f\mid f\in I).\end{eqnarray}$$

Note that if $I$ is $\mathfrak{m}_{x}$ -primary, then so is $\operatorname{in}_{{>}}I$ . Also, if $\{I_{m}\}_{m\in \mathbf{N}}$ is a graded sequence of ideals of $R_{\unicode[STIX]{x1D70E}}$ , then so is $\{\operatorname{in}_{{>}}I_{m}\}_{m\in \mathbf{N}}$ . This follows from the fact that $\operatorname{in}_{{>}}f\cdot \operatorname{in}_{{>}}g=\operatorname{in}_{{>}}fg$ .

Lemma 8.1. If $I\subset R_{\unicode[STIX]{x1D70E}}$ is an $\mathfrak{m}_{x}$ -primary ideal, then

$$\begin{eqnarray}\operatorname{length}(R_{\unicode[STIX]{x1D70E}}/I)=\operatorname{length}(R_{\unicode[STIX]{x1D70E}}/\text{in}_{{>}}I).\end{eqnarray}$$

Proof. The proof is similar to the proof of [Reference EisenbudEis95, Theorem 15.3]. ◻

Similar to the argument in [Reference EisenbudEis95], we construct a deformation of $I$ to $\operatorname{in}_{{>}}I$ . Since $R_{\unicode[STIX]{x1D70E}}$ is Noetherian, we may choose elements $g_{1},\ldots ,g_{s}\in I$ such that

$$\begin{eqnarray}I=(g_{1},\ldots ,g_{s})\quad \text{and}\quad \operatorname{in}_{{>}}I=(\operatorname{in}_{{>}}g_{1},\ldots ,\operatorname{in}_{{>}}g_{s}).\end{eqnarray}$$

Fix an integral weight $\unicode[STIX]{x1D706}:M\cap \unicode[STIX]{x1D70E}^{\vee }\rightarrow \mathbf{Z}_{{\geqslant}0}$ such that

$$\begin{eqnarray}\operatorname{in}_{{>}_{\unicode[STIX]{x1D706}}}(g_{i})=\operatorname{in}_{{>}}(g_{i})\end{eqnarray}$$

for all $i$ . Note that ${>}_{\unicode[STIX]{x1D706}}$ denotes the order on the monomials induced by the weight function  $\unicode[STIX]{x1D706}$ .

Let $R_{\unicode[STIX]{x1D70E}}[t]$ denote the polynomial ring in one variable over $R_{\unicode[STIX]{x1D70E}}$ . For $g=\sum \unicode[STIX]{x1D6FC}_{m}\unicode[STIX]{x1D712}^{m}$ , we write $b:=\max \{\unicode[STIX]{x1D706}(m)\mid \unicode[STIX]{x1D6FC}_{m}\neq 0\}$ and set

$$\begin{eqnarray}\tilde{g}:=t^{b}\sum \unicode[STIX]{x1D6FC}_{m}t^{-\unicode[STIX]{x1D706}(m)}\unicode[STIX]{x1D712}^{m}.\end{eqnarray}$$

Next, let

$$\begin{eqnarray}\tilde{I} =(\tilde{g}_{1}\ldots \tilde{g}_{s})\subset R_{\unicode[STIX]{x1D70E}}[t].\end{eqnarray}$$

For $c\in k$ , we write $I_{c}$ for the image of $\tilde{I}$ under the map $R_{\unicode[STIX]{x1D70E}}[t]\rightarrow R_{\unicode[STIX]{x1D70E}}$ defined by $t\mapsto c$ . It is clear that $I_{1}=I$ and $I_{0}=\operatorname{in}_{{>}}I$ .

Proposition 8.2. If $I\subset R_{\unicode[STIX]{x1D70E}}$ is an $\mathfrak{m}_{x}$ primary ideal, then $\operatorname{lct}(\operatorname{in}_{{<}}(I))\leqslant \operatorname{lct}(I)$ .

Proof. We consider the automorphism of $\unicode[STIX]{x1D711}:R_{\unicode[STIX]{x1D70E}}[t,t^{-1}]\rightarrow R_{\unicode[STIX]{x1D70E}}[t,t^{-1}]$ that sends $\unicode[STIX]{x1D712}^{m}$ to $t^{\unicode[STIX]{x1D706}(m)}\unicode[STIX]{x1D712}^{m}$ . Note that $\unicode[STIX]{x1D711}$ sends $\tilde{I} R_{\unicode[STIX]{x1D70E}}[t,t^{-1}]$ to $IR_{\unicode[STIX]{x1D70E}}[t,t^{-1}]$ . Therefore, for each $c\in k^{\ast }$ , we get an automorphism $\unicode[STIX]{x1D711}_{c}:R_{\unicode[STIX]{x1D70E}}\rightarrow R_{\unicode[STIX]{x1D70E}}$ such that $\unicode[STIX]{x1D711}_{c}(I_{c})=I$ . Thus, $\operatorname{lct}(I_{c})=\operatorname{lct}(I)$ for all $c\in k^{\ast }$ . Since $I_{c}$ is $\mathfrak{m}_{x}$ -primary for all $c\in k$ , we may apply Proposition A.3 to see $\operatorname{lct}(I_{0})\leqslant \operatorname{lct}(I)$ . Since $\operatorname{in}_{{>}}(I)=I_{0}$ , we are done.◻

8.2 Proof of Theorem 1.4

Theorem 1.4 is a direct consequence of Proposition 8.3. Our proof of Proposition 8.3 is inspired by the main argument in [Reference MustaţǎMus02].

Proposition 8.3. Let $\mathfrak{a}_{\bullet }$ be a graded sequence of $\mathfrak{m}_{x}$ -primary ideals on $X_{\unicode[STIX]{x1D70E}}$ . We have that

$$\begin{eqnarray}\operatorname{lct}(\operatorname{in}_{{>}}(\mathfrak{a}_{\bullet }))^{n}\operatorname{e}(\operatorname{in}(\mathfrak{a}_{\bullet }))\leqslant \operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet }).\end{eqnarray}$$

Proof. We first note that

$$\begin{eqnarray}\operatorname{e}(\operatorname{in}_{{>}}(\mathfrak{a}_{\bullet })):=\limsup _{m\rightarrow \infty }\frac{\operatorname{length}({\mathcal{O}}_{X_{\unicode[STIX]{x1D70E}},x}/\text{in}_{{>}}(\mathfrak{a}_{m}))}{m^{n}/n!}=\limsup _{m\rightarrow \infty }\frac{\operatorname{length}({\mathcal{O}}_{X_{\unicode[STIX]{x1D70E}},x}/\mathfrak{a}_{m})}{m^{n}/n!}=:\operatorname{e}(\mathfrak{a}_{\bullet }),\end{eqnarray}$$

where the second equality follows from Proposition 8.1. By Proposition 8.2,

$$\begin{eqnarray}\operatorname{lct}(\operatorname{in}_{{>}}(\mathfrak{a}_{\bullet }))\leqslant \operatorname{lct}(\mathfrak{a}_{\bullet }).\end{eqnarray}$$

The result follows. ◻

Proof of Theorem 1.4.

Since $\operatorname{Val}_{X,x}^{\text{toric}}\subset \operatorname{Val}_{X,x}$ , we have

$$\begin{eqnarray}\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)\leqslant \inf _{v\in \operatorname{Val}_{x,x}^{\text{toric}}}\widehat{\operatorname{vol}}(v).\end{eqnarray}$$

We proceed to show the reveres inequality. Note that

(8.1) $$\begin{eqnarray}\inf _{v\in \operatorname{Val}_{X,x}}\widehat{\operatorname{vol}}(v)=\inf _{\mathfrak{a}_{\bullet }\,\mathfrak{m}_{x}\text{-}\text{primary}}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet })=\inf _{\substack{ \mathfrak{a}_{\bullet }\,\mathfrak{m}_{x}\text{-}\text{primary} \\ \text{monomial}}}\operatorname{lct}(\mathfrak{a}_{\bullet })^{n}\operatorname{e}(\mathfrak{a}_{\bullet }),\end{eqnarray}$$

where the first equality follows from Proposition 4.3 and the second from Proposition 8.3. The last infimum in (8.1) is equal to

$$\begin{eqnarray}\inf _{\substack{ \mathfrak{a}\,\mathfrak{m}_{x}\text{-}\text{primary} \\ \text{monomial}}}\operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a})\end{eqnarray}$$

by Lemma 4.1. Thus, it is sufficient to show that

$$\begin{eqnarray}\inf _{v\in \operatorname{Val}_{X,x}^{\text{toric}}}\widehat{\operatorname{vol}}(v)\leqslant \inf _{\substack{ \mathfrak{a}\,\mathfrak{m}_{x}\text{-}\text{primary} \\ \text{monomial}}}\operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a}).\end{eqnarray}$$

Let $\mathfrak{a}$ be an $\mathfrak{m}_{x}$ -primary monomial ideal. Since $\mathfrak{a}$ is a monomial ideal, there exists a toric valuation $v^{\ast }\in \operatorname{Val}_{X,x}^{\text{toric}}$ such that $v^{\ast }$ computes $\operatorname{lct}(\mathfrak{a})$ . (This follows from the fact that there exists a toric log resolution of  $\mathfrak{a}$ .) By Proposition 4.4,

$$\begin{eqnarray}\widehat{\operatorname{vol}}(v^{\ast })\leqslant \operatorname{lct}(\mathfrak{a})^{n}\operatorname{e}(\mathfrak{a}),\end{eqnarray}$$

and the proof is complete. ◻

Figure 1. The cone $\unicode[STIX]{x1D70E}$ is drawn. The toric variety $X_{\unicode[STIX]{x1D70E}}$ is isomorphic to the cone over $\mathbb{P}^{2}$ blown up at a point.

8.3 An example of nondivisorial volume minimizer

Let $V$ denote $\mathbb{P}^{2}$ blown up at a point. Note that $V$ is a Fano variety. The affine cone $C(V,-K_{V})=\operatorname{Spec}(\bigoplus _{m\geqslant 0}H^{0}(V,-K_{V}))$ is isomorphic to the toric variety $X_{\unicode[STIX]{x1D70E}}$ , where $\unicode[STIX]{x1D70E}\subseteq \mathbf{R}^{3}$ is the cone in Figure 1. Let $x$ denote the torus invariant point of $X_{\unicode[STIX]{x1D70E}}$ .

We seek to find a minimizer of the function $\operatorname{Val}_{X_{\unicode[STIX]{x1D70E}},x}^{\text{toric}}\rightarrow \mathbf{R}_{{>}0}$ defined by $v_{u}\mapsto \widehat{\operatorname{vol}}(v_{u})$ . Since the normalized volume is invariant under scaling, it is sufficient to consider elements $u\in \operatorname{Int}(\unicode[STIX]{x1D70E})$ of the form $u=(a,b,1)\in \operatorname{Int}(\unicode[STIX]{x1D70E})$ . We have

$$\begin{eqnarray}A_{X_{\unicode[STIX]{x1D70E}}}(v_{(a,b,1)})=\langle (a,b,1),(0,0,1)\rangle =1.\end{eqnarray}$$

The normalized volume of $v_{(a,b,c)}$ is

$$\begin{eqnarray}\widehat{\operatorname{vol}}(v_{(a,b,1)}):=A_{X}(v_{(a,b,1)})^{3}\operatorname{vol}(v_{(a,b,1)})=3!\operatorname{Vol}(\unicode[STIX]{x1D70E}^{\vee }\setminus H_{(a,b,1)}(1)).\end{eqnarray}$$

After computing the previous volume, we see that the function is minimized at

$$\begin{eqnarray}(a^{\ast },b^{\ast },1)=(4/3-\sqrt{13}/3,4/3-\sqrt{13}/3,1),\end{eqnarray}$$

with $\widehat{\operatorname{vol}}(v_{(a^{\ast },b^{\ast },1)})=\frac{1}{12}(46+13\sqrt{13})$ . By Theorem 1.4, the toric volume minimizer $v^{\ast }=v_{(a^{\ast },b^{\ast },1)}$ is also a minimizer of $\widehat{\operatorname{vol}}_{X_{\unicode[STIX]{x1D70E}},x}$ . Since $\widehat{\operatorname{vol}}(v_{(a^{\ast },b^{\ast },1)})$ is irrational, Proposition 4.9 implies there cannot be a divisorial minimizer of $\widehat{\operatorname{vol}}_{X_{\unicode[STIX]{x1D70E}},x}$ .

Acknowledgements

I would like to thank my advisor Mircea Mustaţă for guiding my research and sharing his ideas with me. I also thank Mattias Jonsson, Chi Li, Yuchen Liu, and Chenyang Xu for comments on a previous draft of this paper. Lastly, I am grateful to David Stapleton for producing the graphic in § 8.3.

Appendix A Multiplicities and log canonical thresholds in families

In this section we provide information on the behavior of the Hilbert–Samuel multiplicity and log canonical threshold in a family. The content of this section is well known to experts, but does not necessarily appear in the literature in the desired form. The following statements will be useful in the proof of Proposition 5.2.

A.1 Multiplicities

We recall an interpretation of the Hilbert–Samuel multiplicity described in [Reference RamanujamRam73]. Let $X$ be a proper variety of dimension $n$ and $x\in X$ a closed point. If $\mathfrak{a}\subseteq {\mathcal{O}}_{X}$ is an $\mathfrak{m}_{x}$ -primary ideal and $Y\rightarrow X$ a proper birational morphism such that $Y$ is nonsingular and $\mathfrak{a}\cdot {\mathcal{O}}_{Y}$ is an invertible sheaf, then $\operatorname{e}(\mathfrak{a})=(-1)^{n-1}D^{n}$ , where $D$ is the effective divisor on $Y$ defined by $\mathfrak{a}\cdot {\mathcal{O}}_{Y}$ .

The following proposition is well known. Related statements appear in [Reference BennettBen70] and [Reference LipmanLip82].

Proposition A.1. Let $X$ and $T$ be varieties and $x\in X$ a closed point. If $\mathfrak{a}\subseteq {\mathcal{O}}_{X\times T}$ is an ideal such that $\mathfrak{a}_{t}:=\mathfrak{a}\cdot {\mathcal{O}}_{X\times \{t\}}$ is $\mathfrak{m}_{x}$ -primary for all closed points $t\in T$ , then there exists a nonempty open set $U\subseteq T$ such that the function $U\ni t\mapsto \operatorname{e}(\mathfrak{a}_{t})$ is constant.

Proof. It is sufficient to consider the case when $X$ is proper. Indeed, let $V\subseteq X$ be an open affine subset of $X$ containing  $x$ . Replace $X$ with a projective closure of  $V$ .

Now, let $f:Y\rightarrow X\times T$ be a log resolution of $\mathfrak{a}$ . Set ${\mathcal{L}}=\mathfrak{a}\cdot {\mathcal{O}}_{Y}$ and write $g:Y\rightarrow T$ for the map $f$ composed with the projection $X\times T\rightarrow T$ .

Choose a nonempty open set $U\subseteq T$ such that $g^{-1}(U)$ is flat over $U$ and $Y_{t}\rightarrow X\times \{t\}$ is birational for all $t\in U$ . By [Reference KollárKol96, Proposition 2.9], $U\ni t\mapsto ({\mathcal{L}}_{t})^{n}$ is constant. Thus, we are done.◻

A.2 Log canonical thresholds

We prove two statements on the behavior of the log canonical threshold along a family of ideals. For similar statements, see [Reference LazarsfeldLaz04, Example 9.3.17] and [Reference KollárKol97, Lemma 8.6].

Proposition A.2. Let $X$ and $T$ be varieties and assume that $X$ is klt. Fix an ideal $\mathfrak{a}\subseteq {\mathcal{O}}_{X\times T}$ , and set $\mathfrak{a}_{t}:=\mathfrak{a}\cdot {\mathcal{O}}_{X\times \{t\}}$ for $t\in T$ . Then, there exists a nonempty open set $U\subseteq T$ such that the function $U\ni t\mapsto \operatorname{lct}(\mathfrak{a}_{t})$ is constant.

Proof. Let $\unicode[STIX]{x1D707}:Y\rightarrow X\times T$ be a log resolution of $\mathfrak{a}$ , and set $p^{\prime }=p\circ \unicode[STIX]{x1D707}$ .

Let $D$ be the divisor on $Y$ such that $\mathfrak{a}\cdot {\mathcal{O}}_{X^{\prime }}={\mathcal{O}}_{X^{\prime }}(-D)$ and $E_{1},\ldots ,E_{r}$ be the prime components of $\operatorname{Exc}(\unicode[STIX]{x1D707})+D_{\text{red}}$ . After shrinking $T$ , we may assume that each $E_{i}$ surjects onto $T$ and $T$ is smooth.

We claim that on a nonempty open set $U\subset T$ , $\unicode[STIX]{x1D707}_{t}:Y_{t}\rightarrow X_{t}$ is a log resolution of $\mathfrak{a}_{t}$ for all $t\in U$ , where $Y_{t}:=Y_{p^{-1}(t)}$ and $X_{t}:=X\times \{t\}$ . Indeed, by generic smoothness [Reference HartshorneHar77, Corollary III.10.7] applied to $X^{\prime }$ , each $E_{i}$ , and all the intersections of the $E_{i}$ , we may find such a locus $U\subset T$ .

Now, we have $(K_{Y/X\times T})|_{X_{t}}=K_{Y_{t}/X_{t}}$ and $\mathfrak{a}_{t}\cdot {\mathcal{O}}_{Y_{t}}={\mathcal{O}}_{Y_{t}}(-D|_{Y_{t}})$ for $t\in U$ . Therefore, $\operatorname{lct}(\mathfrak{a}_{t})=\min _{i=1,\ldots ,r}\operatorname{ord}_{E_{i}}(K_{Y/X\times T})/\text{ord}_{E_{i}}(D)$ for all $t\in U$ , and we are done.◻

Proposition A.3. Let $X$ be a klt variety, $T$ a smooth curve, and $t_{0}\in T$ a closed point. Fix an ideal $\mathfrak{a}\subseteq {\mathcal{O}}_{X\times T}$ , and set $\mathfrak{a}_{t}:=\mathfrak{a}\cdot {\mathcal{O}}_{X\times \{t\}}$ . If $V(\mathfrak{a})\subset X$ is proper over $T$ , then there exists an open neighborhood $t_{0}\in U\subseteq T$ such that

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{t_{0}})\leqslant \operatorname{lct}(\mathfrak{a}_{t})\end{eqnarray}$$

for all closed points $t\in U$ .

Remark A.4. The condition that $V(\mathfrak{a})$ is proper over $T$ holds if there exists $x\in X$ such that each ideal $\mathfrak{a}_{t}$ is $\mathfrak{m}_{x}$ -primary for all closed points $t\in T$ . Indeed, in this case $V(\mathfrak{a})_{\text{red}}=\{x\}\times T$ .

Proof. By Proposition A.2, we may choose a nonempty open set $W\subseteq T$ such that $\operatorname{lct}(\mathfrak{a}_{t})$ takes the constant value $\unicode[STIX]{x1D706}$ for all $t\in W$ . We will show $\operatorname{lct}(\mathfrak{a}_{t_{0}})\leqslant \unicode[STIX]{x1D706}$ . Then, $U:=W\cup \{t_{0}\}$ will satisfy the conclusion of our proposition.

We first show that $p(V({\mathcal{J}}(X,\times T,\unicode[STIX]{x1D706}\cdot \mathfrak{a})))=T$ . By [Reference LazarsfeldLaz04, Example 9.5.34], we may shrink $W$ so that ${\mathcal{J}}(X\times \{t\},\unicode[STIX]{x1D706}\cdot \mathfrak{a}_{t})={\mathcal{J}}(X,\unicode[STIX]{x1D706}\cdot \mathfrak{a})\cdot {\mathcal{O}}_{X\times \{t\}}$ for all $t\in W$ . Since ${\mathcal{J}}(X\times \{t\},\mathfrak{a}_{t}^{\unicode[STIX]{x1D706}})\neq {\mathcal{O}}_{X\times \{t\}}$ for all $t\in W$ , we see $W\subseteq (p(V({\mathcal{J}}(X,\times T,\unicode[STIX]{x1D706}\cdot \mathfrak{a}))))$ . Note that $p(V({\mathcal{J}}(X\times T,\unicode[STIX]{x1D706}\cdot \mathfrak{a})))$ is closed in $T$ , since ${\mathcal{J}}(X\times T,\unicode[STIX]{x1D706}\cdot \mathfrak{a})$ is closed in $V(\mathfrak{a})$ and $V(\mathfrak{a})$ is proper over  $T$ . Therefore, $p(V({\mathcal{J}}(X\times T,\unicode[STIX]{x1D706}\cdot \mathfrak{a})))=T$ .

Since $p({\mathcal{J}}(X\times T,\unicode[STIX]{x1D706}\cdot \mathfrak{a})\cdot {\mathcal{O}}_{X\times \{t_{0}\}})=T$ and ${\mathcal{J}}(X\times \{t_{0}\},\unicode[STIX]{x1D706}\cdot \mathfrak{a}_{t_{0}})\subseteq {\mathcal{J}}(X\times T,\unicode[STIX]{x1D706}\cdot \mathfrak{a})\cdot {\mathcal{O}}_{X\times \{t_{0}\}}$ by [Reference LazarsfeldLaz04, Theorem 9.5.16], ${\mathcal{J}}(X\times \{t_{0}\},\unicode[STIX]{x1D706}\cdot \mathfrak{a}_{t_{0}})$ is nontrivial. Thus, $\operatorname{lct}(\mathfrak{a}_{t_{0}})\leqslant \unicode[STIX]{x1D706}$ , and the proof is complete.◻

Appendix B Valuations computing log canonical thresholds of graded sequences

In [Reference Jonsson and MustaţǎJM12], the authors prove the existence of valuations computing log canonical thresholds of graded sequences of ideals on smooth varieties. We extend the result to klt varieties. While the statement is likely known to experts, it does not appear in the literature.

Theorem B.1. If $X$ is a klt variety and $\mathfrak{a}_{\bullet }$ a graded sequence of ideals on $X$ , then there exists $v^{\ast }\in \operatorname{Val}_{X}$ computing $\operatorname{lct}(\mathfrak{a}_{\bullet })$ .

The proof we give is similar in spirit to the proof of [Reference Jonsson and MustaţǎJM12, Theorem 7.3], but also relies on results in [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15].

Proposition B.2. If $X$ is a normal $\mathbf{Q}$ -Gorenstein variety and $\mathfrak{a}_{\bullet }$ a graded sequence of ideals on $X$ , then $v\mapsto v(\mathfrak{a}_{\bullet })$ is a continuous function on $\operatorname{Val}_{X}\cap \,\{A_{X}(v)<+\infty \}$ .

Proof. We reduce the result to the smooth case. Take a resolution of singularities $Y\rightarrow X$ and write $\mathfrak{a}_{\bullet }^{Y}$ for the graded sequence of ideals on $Y$ defined by $\mathfrak{a}_{m}^{Y}=\mathfrak{a}_{m}\cdot {\mathcal{O}}_{Y}$ . Thanks to [Reference Jonsson and MustaţǎJM12, Corollary 6.4], the function $v\mapsto v(\mathfrak{a}_{\bullet }^{Y})$ is continuous on $\operatorname{Val}_{Y}\cap \,\{A_{Y}(v)<+\infty \}$ .

Now, note that the natural map $\operatorname{Val}_{Y}\rightarrow \operatorname{Val}_{X}$ is a homeomorphism of topological spaces and $v(\mathfrak{a}_{\bullet })=v(\mathfrak{a}_{\bullet }^{Y})$ . Since $A_{X}(v)=A_{Y}(v)+v(K_{Y/X})$ , $A_{X}(v)<+\infty$ if and only if $A_{Y}(v)<+\infty$ . Thus, the proof is complete.◻

Now, we recall some formalism from [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15]. A normalizing subscheme on $X$ is a (nontrivial) closed subscheme of $X$ containing $\operatorname{Sing}(X)$ . If $N$ is a normalizing subscheme of $X$ , we set

$$\begin{eqnarray}\operatorname{Val}_{X}^{N}:=\{v\in \operatorname{Val}_{X}\mid v({\mathcal{I}}_{N})=1\}.\end{eqnarray}$$

Proposition B.3. Let $X$ be a normal $\mathbf{Q}$ -Gorenstein variety, $\mathfrak{a}_{\bullet }$ a graded sequence of ideals on $X$ , and $N$ a normalizing subscheme of $X$ such that $N$ contains the zero locus of $\mathfrak{a}_{1}$ .

  1. (a) The function $v\mapsto v(\mathfrak{a}_{\bullet })$ is bounded on $\operatorname{Val}_{X}^{N}$ .

  2. (b) For each $M\in \mathbf{R}$ , the set $\{A_{X}(v)\leqslant M\}\cap \operatorname{Val}_{X}^{N}$ is compact.

Proof. Statements (a) and (b) appear in [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15, Proposition 2.5] and [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15, Theorem 3.1], respectively. ◻

Proposition B.4. If $X$ is a klt variety and $N$ a normalized subscheme of $X$ , then there exists $\unicode[STIX]{x1D716}>0$ such that $A_{X}(v)>\unicode[STIX]{x1D716}$ for all $v\in \operatorname{Val}_{X}^{N}$ .

Proof. Let $\unicode[STIX]{x1D70B}:Y\rightarrow X$ be a good resolution of $N$ , and consider the continuous retraction map

$$\begin{eqnarray}r_{\unicode[STIX]{x1D70B}}^{N}:\operatorname{Val}_{X}^{N}\rightarrow \unicode[STIX]{x1D6E5}_{\unicode[STIX]{x1D70B}}^{N}\end{eqnarray}$$

described in [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15, § 2.2]. Since $X$ is assumed to be klt, there exists $\unicode[STIX]{x1D716}>0$ such that $A_{X}(v)>\unicode[STIX]{x1D716}$ for all $v\in \unicode[STIX]{x1D6E5}_{\unicode[STIX]{x1D70B}}^{N}$ . Since $A_{X}(v)\geqslant A_{X}(r_{\unicode[STIX]{x1D70B}}^{N}(v))$ for all $v\in \operatorname{Val}_{X}^{N}$ by [Reference Boucksom, de Fernex, Favre and UrbinatiBFFU15, Theorem 3.1], the proof is complete.◻

Proof of Theorem B.1.

If $\operatorname{lct}(\mathfrak{a}_{\bullet })=+\infty$ , then any $v\in \operatorname{Val}_{X}$ with $A_{X}(v)<+\infty$ must compute $\operatorname{lct}(\mathfrak{a}_{\bullet })$ . We now move on to the case when $\operatorname{lct}(\mathfrak{a}_{\bullet })<+\infty$ .

Let $N$ be the subscheme of $X$ defined by the ideal ${\mathcal{I}}_{\operatorname{Sing}(X)}\cdot \mathfrak{a}_{1}$ . Since $N$ is a normalizing subscheme of $X$ that contains the zero locus of $\mathfrak{a}_{1}$ , we may apply the previous two propositions to choose $B\in \mathbf{R}$ and $\unicode[STIX]{x1D716}>0$ so that $v(\mathfrak{a}_{\bullet })<B$ and $A_{X}(v)>\unicode[STIX]{x1D716}$ for all $v\in \operatorname{Val}_{X}^{N}$ .

Note that

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet })=\inf _{v\in \operatorname{Val}_{X}^{N}}\frac{A_{X}(v)}{v(\mathfrak{a}_{\bullet })}.\end{eqnarray}$$

Indeed, consider $v\in \operatorname{Val}_{X}$ such that $A_{X}(v)/v(\mathfrak{a}_{\bullet })<+\infty$ . Since $v(\mathfrak{a}_{\bullet })>0$ , then $v(\mathfrak{a}_{1})>0$ and, thus, $v({\mathcal{I}}_{N})>0$ . We see $w=(1/v({\mathcal{I}}_{N}))v\in \operatorname{Val}_{X}^{N}$ and $A_{X}(w)/w(\mathfrak{a}_{\bullet })=A_{X}(v)/v(\mathfrak{a}_{\bullet })$ .

Now, fix $L>\operatorname{lct}(\mathfrak{a}_{\bullet })$ . If $v\in \operatorname{Val}_{X}^{N}$ and $A_{X}(v)/v(\mathfrak{a}_{\bullet })\leqslant L$ , then

$$\begin{eqnarray}\unicode[STIX]{x1D716}<A_{X}(v)\leqslant Lv(\mathfrak{a}_{\bullet })\leqslant L\cdot B.\end{eqnarray}$$

Therefore,

$$\begin{eqnarray}\operatorname{lct}(\mathfrak{a}_{\bullet })=\inf _{v\in W}\frac{A_{X}(v)}{v(\mathfrak{a}_{\bullet })},\end{eqnarray}$$

where

$$\begin{eqnarray}W=\operatorname{Val}_{X}^{N}\cap \,\{A_{X}(v)\leqslant L\cdot B\}\cap \{Lv(\mathfrak{a}_{\bullet })\geqslant \unicode[STIX]{x1D716}\}.\end{eqnarray}$$

We claim that $W$ is compact. Indeed, $\operatorname{Val}_{X}^{N}\cap \,\{A_{X}(v)\leqslant L\cdot B\}$ is compact by the previous proposition. Since $v\mapsto v(\mathfrak{a}_{\bullet })$ is continuous on $\operatorname{Val}_{X}^{N}\cap \,\{A_{X}(v)\leqslant L\cdot B\}$ , $W$ is closed in $\operatorname{Val}_{X}^{N}\cap \,\{A_{X}(v)\leqslant L\cdot B\}$ , and, thus, compact as well. Since $v\mapsto A_{X}(v)/v(\mathfrak{a}_{\bullet })$ is lower semicontinuous on the compact set $W$ , there exists $v^{\ast }\in W$ such that $A_{X}(v^{\ast })/v^{\ast }(\mathfrak{a}_{\bullet })=\operatorname{lct}(\mathfrak{a}_{\bullet })$ .◻

Footnotes

1 Li showed that if $\widehat{\operatorname{vol}}_{X,x}$ is lower semicontinuous on $\operatorname{Val}_{X,x}$ , then there is a minimizer of $\widehat{\operatorname{vol}}_{X,x}$ [Reference LiLi15, Corollary 3.5]. Note that $\widehat{\operatorname{vol}}_{X,x}(v):=A_{X}(v)^{n}\operatorname{vol}(v)$ is a product of two functions. While $A_{X}$ is lower semicontinuous on $\operatorname{Val}_{X,x}$ , $\operatorname{vol}$ fails to be lower semicontinuous in general [Reference Favre and JonssonFJ04, Proposition 3.31].

2 The cosupport of an ideal $\mathfrak{a}\subseteq {\mathcal{O}}_{X}$ is defined as $\operatorname{Cosupp}(\mathfrak{a}):=\operatorname{Supp}({\mathcal{O}}_{X}/\mathfrak{a})$ .

3 This is equivalent to saying that each ideal $\mathfrak{a}_{m}(v)$ vanishes only at  $x$ .

References

Bennett, B., On the characteristic function of a local ring , Ann. of Math. (2) 91 (1970), 2587.CrossRefGoogle Scholar
Blum, H., On divisors computing mlds and lcts, Preprint (2016), arXiv:1605.09662v2.Google Scholar
Boucksom, S., Favre, C. and Jonsson, M., Valuations and plurisubharmonic singularities , Publ. Res. Inst. Math. Sci. 44 (2008), 449494.Google Scholar
Boucksom, S., de Fernex, T., Favre, C. and Urbinati, S., Valuation spaces and multiplier ideals on singular varieties , in Recent advances in algebraic geometry. Volume in honor of Rob Lazarsfeld’s 60th birthday, London Mathematical Society Lecture Note Series (Cambridge University Press, Cambridge, 2015), 2951.CrossRefGoogle Scholar
Boucksom, S., Hisamoto, T. and Jonsson, M., Uniform K-stability and asymptotics of energy functionals in Kähler geometry, Ann. Inst Fourier, to appear. Preprint (2016),arXiv:1603.01026.Google Scholar
Cutkosky, D., Multiplicities associated to graded families of ideals , Algebra Number Theory 7 (2013), 20592083.Google Scholar
Demailly, J.-P., Ein, L. and Lazarsfeld, R., A subadditivity property of multiplier ideals , Michigan Math. J. 48 (2000), 137156.Google Scholar
Ein, L., Lazarsfeld, R. and Smith, K. E., Uniform approximation of Abhyankar valuation ideals in smooth function fields , Amer. J. Math. 125 (2003), 409440.Google Scholar
Eisenbud, D., Commutative algebra, Graduate Texts in Mathematics, vol. 150 (Springer, New York, NY, 1995).Google Scholar
Eisenstein, E., Inversion of adjunction in high codimension, PhD thesis, University of Michigan (2011).Google Scholar
Favre, C. and Jonsson, M., The valuative tree, Lecture Notes in Mathematics, vol. 1853 (Springer, Berlin, 2004).Google Scholar
de Fernex, T., Birationally rigid hypersurfaces , Invent. Math. 192 (2013), 533566.CrossRefGoogle Scholar
de Fernex, T., Ein, L. and Mustaţă, M., Bounds for log canonical thresholds with applications to birational rigidity , Math. Res. Lett. 10 (2003), 219236.Google Scholar
de Fernex, T., Ein, L. and Mustaţă, M., Multiplicities and log canonical threshold , J. Algebraic Geom. 13 (2004), 603615.Google Scholar
de Fernex, T., Ein, L. and Mustaţă, M., Shokurovs ACC conjecture for log canonical thresholds on smooth varieties , Duke Math. J. 152 (2010), 93114.Google Scholar
de Fernex, T., Ein, L. and Mustaţă, M., Log canonical thresholds on varieties with bounded singularities , in Classification of algebraic varieties, EMS Series of Congress Reports (European Mathematical Society, Zürich, 2011), 221257.Google Scholar
de Fernex, T. and Mustaţă, M., Limits of log canonical thresholds , Ann. Sci. Éc. Norm. Supér. (4) 42 (2009), 491515.Google Scholar
Fulton, W., Introduction to toric varieties, Annals of Mathematics Studies, vol. 131 (Princeton University Press, Princeton, NJ, 1993).Google Scholar
Grothendieck, A., Éléments de géométrie algébrique. II. Étude globale élémentaire de quelques classes de morphismes , Inst. Hautes Études Sci. Publ. Math. 8 (1961), 5222.Google Scholar
Hartshorne, R., Algebraic geometry, Graduate Texts in Mathematics, vol. 52 (Springer, Heidelberg, 1977).Google Scholar
Jonsson, M. and Mustaţǎ, M., Valuations and asymptotic invariants for sequences of ideals , Ann. Inst. Fourier (Grenoble) 62 (2012), 21452209.Google Scholar
Kaveh, K. and Khovanskii, A. G., Convex bodies and multiplicities of ideals , Proc. Steklov Inst. Math. 286 (2014), 268284.CrossRefGoogle Scholar
Kollár, J., Rational curves on algebraic varieties, Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 32 (Springer, Berlin, 1996).Google Scholar
Kollár, J., Singularities of pairs , in Algebraic geometry – Santa Cruz 1995, Proceedings of Symposia in Pure Mathematics, vol. 62, Part 1 (American Mathematical Society, Providence, RI, 1997), 221287.Google Scholar
Kollár, J., Which powers of holomorphic functions are integrable? Preprint (2008), arXiv:0805.0756.Google Scholar
Kollár, J. and Mori, S., Birational geometry of algebraic varieties, Cambridge Tracts in Mathematics, vol. 134 (Cambridge University Press, Cambridge, 1998).CrossRefGoogle Scholar
Lazarsfeld, R., Positivity in algebraic geometry, I and II , Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 48–49 (Springer, Berlin, 2004).Google Scholar
Lazarsfeld, R. and Mustaţǎ, M., Convex bodies associated to linear series , Ann. Sci. Éc. Norm. Supér. (4) 42 (2009), 783835.CrossRefGoogle Scholar
Li, C., Minimizing normalized volumes of valuations, Math. Z., to appear. Preprint (2015),arXiv:1511.08164v3.Google Scholar
Li, C., K-semistability is equivariant volume minimization , Duke Math. J. 166 (2017), 31473218.CrossRefGoogle Scholar
Li, C. and Liu, Y., Kähler–Einstein metrics and volume minimization, Adv. Math., to appear. Preprint (2016), arXiv:1602.05094.Google Scholar
Li, C. and Xu, C., Stability of valuations and Kollár components, Preprint (2016), arXiv:1604.05398.Google Scholar
Lipman, J., Equimultiplicity, reduction, and blowing up , in Commutative algebra (Fairfax, Va. 1979), Lecture Notes in Pure and Applied Mathematics, vol. 68 (Dekker, New York, NY, 1982).Google Scholar
Liu, Y., The volume of singular Kähler–Einstein Fano varieties, Compositio Math., to appear. Preprint (2016), arXiv:1605.01034v2.Google Scholar
Martelli, D. and Sparks, J., Toric geometry, Sasaki–Einstein manifolds and a new infinite class of AdS/CFT duals , Comm. Math. Phys. 262 (2006), 5189.CrossRefGoogle Scholar
Mustaţǎ, M., On multiplicities of graded sequences of ideals , J. Algebra 256 (2002), 229249.Google Scholar
Ramanujam, C., On a geometric interpretation of multiplicity , Invent. Math. 22 (1973), 6367.Google Scholar
Takagi, S., Formulas for multiplier ideals on singular varieties , Amer. J. Math. 128 (2006), 13451362.Google Scholar
Takagi, S., A subadditivity formula for multiplier ideals associated to log pairs , Proc. Amer. Math. Soc. 141 (2013), 93102.Google Scholar
Teissier, B., Sur une inégalité à la Minkowski pour les multiplicités, appendix to D. Eisenbud and H. I. Levine, The degree of a C map germ , Ann. of Math. (2) 106 (1977), 1944.Google Scholar
Figure 0

Figure 1. The cone $\unicode[STIX]{x1D70E}$ is drawn. The toric variety $X_{\unicode[STIX]{x1D70E}}$ is isomorphic to the cone over $\mathbb{P}^{2}$ blown up at a point.