Hostname: page-component-76fb5796d-22dnz Total loading time: 0 Render date: 2024-04-27T15:33:47.176Z Has data issue: false hasContentIssue false

Angehrn-Siu-Helmke’s method applied to abelian varieties

Published online by Cambridge University Press:  11 May 2023

Zhi Jiang*
Affiliation:
Shanghai Center for Mathematical Sciences, Fudan University, China; E-mail: zhijiang@fudan.edu.cn

Abstract

We apply Angehrn-Siu-Helmke’s method to estimate basepoint freeness thresholds of higher dimensional polarized abelian varieties. We showed that a conjecture of Caucci holds for very general polarized abelian varieties in the moduli spaces $\mathcal {A}_{g, l}$ with only finitely many possible exceptions of primitive polarization types l in each dimension g. We improved the bound of basepoint freeness thresholds of any polarized abelian $4$ -folds and simple abelian $5$ -folds.

Type
Algebraic and Complex Geometry
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2023. Published by Cambridge University Press

1 Introduction

We work over the field of complex numbers throughout this paper.

Syzygies of abelian varieties have attracted lots of attention in recent years. Recall the following question asked by Ito for $p\geq 0$ and by Lozovanu for $p=-1$ as well (see [Reference ItoIto1] and [Reference LozovanuLoz]).

Conjecture 1.1. Let $(A, L)$ be a polarized abelian variety of dimension g and $p\geq -1$ an integer. If $(L^g)>(g(p+2)))^g$ and $(L^d\cdot B)>(d(p+2))^d$ for any abelian subvariety B of A of dimension $0<d<g$ , then L satisfies Property $(N_p)$ .

We summarize recent progress towards this conjecture.

1.1 $\mathbb Q$ -twisted sheaves

Given a coherent sheaf $\mathcal {F}$ on A and a rational number $t\in \mathbb Q$ , following [Reference Jiang and PareschiJP], we formally define the $\mathbb Q$ -twisted sheaves $\mathcal {F}\langle tL \rangle $ . We say that $\mathcal {F}\langle tL\rangle $ is IT $^0$ if the i-th cohomological support loci $V^i(\mu _b^*\mathcal {F}\otimes L^{\otimes b^2t})$ is empty for each $i>0$ , where b is an integer such that $b^2t\in \mathbb Z$ . Note that this definition does not depend on the choice of b. Assume that D is an effective $\mathbb Q$ -divisor on A such that D is $\mathbb Q$ -equivalent to $tL$ . We will also write $\mathcal {F}\langle D\rangle =\mathcal {F}\langle tL \rangle $ .

Similarly, we say that $\mathcal {F}\langle tL \rangle $ is M-regular (resp. GV) if

$$ \begin{align*}\operatorname{\mathrm{codim}}_{\operatorname{\mathrm{Pic}}^0(A)}V^i(\mu_b^*\mathcal{F}\otimes L^{\otimes (b^2t)})>i\end{align*} $$

(resp. $\operatorname {\mathrm {codim}}_{\operatorname {\mathrm {Pic}}^0(A)}V^i(\mu _b^*\mathcal {F}\otimes L^{\otimes (b^2t)})\geq i$ ) for all $i>0$ . We can similarly define the cohomology ranks of $\mathcal {F}\langle tL\rangle $ :

$$ \begin{align*} h^i(A, \mathcal{F}\langle tL\rangle):=\frac{1}{b^{2\dim A}}h^i(A, \mu_b^*\mathcal{F}\otimes L^{\otimes b^2t}\otimes Q), \end{align*} $$

where $Q\in \operatorname {\mathrm {Pic}}^0(A)$ is general. The main result of [Reference Jiang and PareschiJP] says that the function $t\rightarrow h^i(A, \mathcal {F}\langle tL\rangle )$ is locally polynomial on a left or right neighborhood of a rational number and can be extended to a continuous function from $\mathbb R$ to $\mathbb R$ . We call this function the i-th cohomological rank function of $\mathcal {F}$ and denote it by $h^i_{\mathcal {F}, L}(t)$ .

Let $\mathcal {I}_o$ be the ideal sheaf of the neutral element o of A. The basepoint freeness threshold $\beta (L)$ is defined to be

$$ \begin{align*} \beta(L):=\inf\{t\in \mathbb Q\mid \mathcal{I}_o\;\langle tL\rangle \;\text{is}\; \mathrm{IT}^0\}. \end{align*} $$

By the main theorem of [Reference HaconHac], $\beta (L)$ is also equal to

$$ \begin{align*} \inf\{t\in \mathbb Q\mid \mathcal{I}_o\;\langle tL\rangle \;\text{is}\; \mathrm{GV}\}. \end{align*} $$

It was observed in [Reference Jiang and PareschiJP] that $\beta (L)\leq 1$ and equality holds iff $|L|$ has a basepoint. This shows that $\beta (L)$ may vary in families. More generally, if for some rational number $t=\frac {a}{b}$ , $V^i(\mu _b^*\mathcal {I}_o\otimes L^{\otimes (ab)})$ is a nonempty proper subset of $\operatorname {\mathrm {Pic}}^0(A)$ , $\beta (L)=t$ . However, there is, by far, no general way to determine $\beta (L)$ . It is also not clear whether or not $\beta (L)$ is always a rational number.

By [Reference Jiang and PareschiJP, Section 8] and [Reference CaucciC, Theorem 1.1], knowing the exact number of the basepoint freeness threshold $\beta (L)$ helps to understand the syzygies of L.

Theorem 1.2. For $p\geq -1$ , if $\beta (L)<\frac {1}{p+2}$ , L satisfies Property $(N_p)$ .

Remark 1.3. Ito refined this criteria in [Reference ItoIto4, Theorem 1.5] by showing that if $\mathcal {I}_o\langle \frac {1}{p+2}L\rangle $ is M-regular for some integer $p> 0$ , L satisfies Property $(N_p)$ .

Caucci then asked the following question, which would imply Conjecture 1.1.

Conjecture 1.4. Let $(A, L)$ be a polarized abelian variety of dimension g and $p\geq -1$ an integer. If $(L^g)>(g(p+2)))^g$ and $(L^d\cdot B)>(d(p+2))^d$ for any abelian subvariety B of A of dimension $0<d<g$ , then $\beta (L)<\frac {1}{p+2}$ .

1.2 Known results

By Theorem 1.2, in order to solve Conjecture 1.1 and Conjecture 1.4, it suffices to prove that

(1) $$ \begin{align} &\beta(L)\leq n(L):=\\ \nonumber \mathrm{inf}\{\frac{d}{\sqrt[d]{(L^d\cdot B)}}\mid\qquad & B \text{ is an abelian subvariety of dimension}\; 1\leq d \leq g\}. \end{align} $$

There are various ways to estimate $\beta (L)$ . In [Reference ItoIto2], Ito showed that the Angehrn-Siu method (see [Reference Angehrn and SiuAS, Reference HelmkeHel, Reference KawamataK]), which was initially applied to attack Fujita’s basepoint-freeness conjecture, can also be used to estimate $\beta (L)$ . To be more precise, let

$$ \begin{align*} &\quad r'(L):=\mathrm{inf}\{t\in \mathbb Q\mid \text{ there exists a } \mathbb{Q}\text{-divisor}\; D\sim_{\mathbb Q}tL \\ &\text{such that }\; o\; \text{is an isolated component of}\; \mathrm{Nklt}(A, D)\}, \end{align*} $$

where the non-klt locus $\mathrm {Nklt}(A, D)$ is the subscheme of A defined by the multiplier ideal $\mathcal {J}(A, D)$ (see, for instance, [Reference LazarsfeldLaz2, Section 9 and 10]). Ito proved in [Reference ItoIto2] that we always have $\beta (L)\leq r'(L)$ , $ r'(L)\leq n(L)$ when $(A, L)$ is a polarized simple abelian 3-fold, and Conjecture 1.1 and Conjecture 1.4 hold for any polarized abelian 3-fold.

In higher dimensions, we have proved in [Reference JiangJz] via generic vanishing that $\beta (L)\leq 2n(L).$ This implies that given a polarized abelian variety $(A, L)$ , if $(L^d\cdot B)>(2(p+2)d)^d$ for any abelian subvariety B of dimension $1\leq d\leq g$ , L satisfies Property $(N_p)$ .

Based on Bridgeland’s stability condition on surfaces, Lahoz-Rojas [Reference Lahoz and RojasLR] and Rojas [Reference RojasR] almost determined the cohomological rank functions $h^i_{\mathcal {I}_o, L}$ for any polarized abelian surface $(A, L)$ of Picard number $1$ . When L is of polarization type $(1, d)$ , Rojas proved that $\beta (L)=\frac {1}{\sqrt {d}}$ when d is a perfect square or $\beta (L)=\frac {2y}{x-1}$ where $(x, y)$ is the minimal or the second minimal positive solution of the Pell’s equation $X^2-4dY^2=1$ when d is not a perfect square. In either case, we have $\beta (L)< \frac {\sqrt {d}+1}{d}<\frac {\sqrt {2}}{\sqrt {d}}=n(L)$ when $d\geq 6$ .

Via a degeneration method, Ito studied in [Reference ItoIto3] the syzygies of general polarized abelian varieties of type $(1,\ldots ,1, d)$ . He proved that when $d\geq \frac {(p+2)^{g+1}-1}{p+1}$ , L satisfies Property $(N_p)$ .

Inspired by the results of Ito and Rojas, one may believe that $\beta (L)$ should be quite close to $\displaystyle\frac {1}{\sqrt[g]{h^0(A, L)}}=\frac {\sqrt[g]{g!}}{\sqrt[g]{(L^g)}}$ for general polarized abelian variety $(A, L)$ of dimension g.

1.3 Main results

In this paper, we follow Ito’s approach to apply the Angehrn-Siu method to estimate $\beta (L)$ for polarized abelian varieties. We improve the known upper bounds for $\beta (L)$ for higher dimensional polarized abelian varieties.

We denote, respectively, by g and l the dimension and the polarization type of a polarized abelian variety $(A, L)$ . There is a quasi-projective variety $\mathcal {A}_{g, l}$ parametrizing such polarized abelian varieties. We may assume that l is primitive.

Theorem 1.5. Let $(A, L)$ be a very general polarized abelian variety in $\mathcal {A}_{g, l}$ . When $g=4$ or $5$ or $6$ , we have $\beta (L)\leq n(L)$ . When $g\geq 7$ , we also have $\beta (L)\leq n(L)$ , except for possibly finitely many primitive polarization types l in each dimension g.

Theorem 1.6. Let $(A, L)$ be a polarized abelian $4$ -fold. Assume that $(L^4)>((2+\frac {4}{\sqrt {3}})(p+2))^4$ and $(L^d\cdot B)>((p+2)d)^d$ for any abelian subvariety B of dimension $1\leq d\leq 3$ . Then, $\beta (L)<\frac {1}{p+2}$ .

Remark 1.7. Note that $2+\frac {4}{\sqrt {3}}\approx 4.31$ . Thus, Theorem 1.6 is quite close to Conjecture 1.4 for abelian 4-folds.

Theorem 1.8. Let $(A, L)$ be a polarized abelian $5$ -fold. Assume that A is simple and $(L^5)>(8(p+2))^5$ . Then, $\beta (L)< \frac {1}{p+2}$ .

In higher dimensions, we also have slight improvements of $\beta (L)$ . For $g\geq 6$ , we define $\alpha _{g, g-2}=\sqrt [g-2]{\frac {15(g-3)!}{g-1}}$ , $\alpha _{g, 2}=\sqrt {\frac {5(g-2)+1}{g-1}}$ . For $3\leq d\leq g-3$ ,

$$ \begin{align*}\alpha_{g,d}=\sqrt[d]{\frac{\frac{5}{2}d!(g-d)}{\binom{g-1}{g-d}}},\end{align*} $$

and let

$$ \begin{align*}\alpha_g:=\mathrm{min}\{\alpha_{g,d}\mid 2\leq d\leq g-2\}.\end{align*} $$

Theorem 1.9. Let $(A, L)$ be a polarized abelian variety of dimension $g\geq 6$ . Assume that A is simple and $(L^g)>((2g-\alpha _g)(p+2))^g$ , $\beta (L)< \frac {1}{p+2}$ .

Note that $\alpha _6=\alpha _{6,3}=\sqrt [3]{\frac {9}{2}}$ , $\alpha _7=\alpha _{7, 3}=\sqrt [3]{4}$ , $\alpha _8=\alpha _{8, 3}=\sqrt [3]{\frac {25}{7}}$ .

2 Preliminaries

2.1 Log canonical centers

Let X be a smooth projective variety and $\Delta $ be an effective $\mathbb Q$ -divisor on X. We take $\mu : Y\rightarrow X$ a log resolution of $(X, \Delta )$ , and we write

$$ \begin{align*}K_Y=\mu^*(K_X+\Delta)+\sum_Ea(E, X, \Delta)E, \end{align*} $$

where E runs through prime divisors of Y and $\mu _*(\sum _Ea(E, X, \Delta )E)=-\Delta $ .

We say that $(X, \Delta )$ is log canonical at x if $a(E, X, \Delta )\geq -1$ for all prime divisors E on Y such that $x\in \mu (E)$ . For a prime divisor E on Y, if $(X, \Delta )$ is log canonical at the generic point of $\mu (E)$ and the discrepancy $a(E, X, \Delta )=-1$ , we call $\mu (E)$ a log canonical center of $(X, \Delta )$ .

When $(X, \Delta )$ is log canonical at x, there are finitely many log canonical centers containing x, and the intersection of two such log canonical centers is the union of certain log canonical centers containing x (see, for instance, [Reference KawamataK, Proposition 1.5] or [Reference FujinoF, Theorem 9.1]). Thus, there exists a unique minimal log canonical center Z of $(X, \Delta )$ through x. Moreover, Z is normal and has rational singularities around x. More precisely, locally around x, there exists a $\mathbb Q$ -effective divisor $\Delta _Z$ such that $(Z, \Delta _Z)$ is klt (see, for instance, [Reference Fujino and YoshinoriFG, Theorem 7.1]). However, the singularities of Z away from x cannot be controlled in general. The following result due to Xiaodong Jiang from [Reference JiangJx, Proposition 5.1] will be applied later.

Proposition 2.1. Let X be a smooth projective variety and $\Delta $ an effective $\mathbb Q$ -divisor on X with Z a log canonical center of $(X, \Delta )$ Footnote 1 . Let $\nu : \overline {Z}\rightarrow Z$ be the normalization. Then, there exists an effective $\mathbb Q$ -divisor $\Delta _{\overline {Z}}$ on $\overline {Z}$ such that $\nu ^*(K_X+\Delta )\sim _{\mathbb Q} K_{\overline {Z}}+\Delta _{\overline {Z}}$ .

We then have the immediate corollary.

Corollary 2.2. Under the above assumption, we have

$$ \begin{align*}\operatorname{\mathrm{vol}}(Z, (K_X+\Delta)|_Z)=\operatorname{\mathrm{vol}}(\overline{Z}, K_{\overline{Z}}+\Delta_{\overline{Z}})\geq \operatorname{\mathrm{vol}}(Z', K_{Z'}), \end{align*} $$

where $Z'$ is any $\mathbb Q$ -Gorenstein partial resolution of $\overline {Z}$ .

Proof. Let $\rho : Z'\rightarrow \overline {Z}$ be a $\mathbb Q$ -Gorenstein partial resolution and let $\mu : \tilde {Z}\xrightarrow {\sigma }Z'\xrightarrow {\rho } \overline {Z}$ be a log resolution of $(\overline {Z}, \Delta _{\overline {Z}})$ . Let $\Delta _{\tilde {Z}}=\mu _*^{-1}(\Delta _{\overline {Z}})$ . We then have $K_{\tilde {Z}}+\Delta _{\tilde {Z}}+E_1=\mu ^*(K_{\overline {Z}}+\Delta _{\overline {Z}})+E_2$ and $K_{\tilde {Z}}+E_3=\sigma ^*K_{Z'}+E_4$ , where $E_i$ are $\mathbb Q$ -effective $\mu $ -exceptional divisors for $1\leq i\leq 4$ . Thus, $ \operatorname {\mathrm {vol}}(\overline {Z}, K_{\overline {Z}}+\Delta _{\overline {Z}})=\operatorname {\mathrm {vol}}(\tilde {Z}, \mu ^*(K_{\overline {Z}}+\Delta _{\overline {Z}})+E_2+E_3)=\operatorname {\mathrm {vol}}(\tilde {Z}, K_{\tilde {Z}}+\Delta _{\tilde {Z}}+E_1+E_3)\geq \operatorname {\mathrm {vol}}(Z', K_{Z'})$ .

The reason that we are interested to know the lower bound of the restricted volume is that we like to apply Helmke’s induction to cut down log canonical centers.

We briefly recall Helmke’s work [Reference HelmkeHel]. Let D be an ample effective $\mathbb Q$ -divisor on a smooth projective variety X of dimension n. We assume that $x\in D$ and $\mathrm {mult}_xD=m>0$ . Let $c:=\mathrm {lct}(D, x)$ be the log canonical threshold of D at x; namely, c is the maximal rational number such that $(X, cD)$ is log canonical at x. Let Z be the minimal log canonical center of $(X, cD)$ through x and denote $d=\dim Z$ . Helmke and Ein independently introduced the local discrepancyFootnote 2 $b_x(X, cD)$ of $(X, cD)$ at x, which is the rational number

$$ \begin{align*} \mathrm{max}\{\mathrm{mult}_x E\mid\qquad & E\;\text{ is an effective}\; \mathbb Q\text{-divisor}\; \\ & \text{such that}\; (X, cD+E) \;\text{is log canonical at}\; x\}. \end{align*} $$

It is known that $0\leq b_x(X, cD)\leq n-cm$ , $b_x(X, cD)\leq d$ and $b_x(X, cD)=0$ iff $Z=\{x\}$ (see [Reference HelmkeHel] and [Reference EinE]).

Helmke’s induction can be summarized as follows (see [Reference HelmkeHel, Proposition 3.2 and Theorem 4.3]).

Proposition 2.3. Under the above assumption, assume that $c<1$ .

  1. (1) If $(D^{\dim Z}\cdot Z)>(\frac {b_x(X, cD)}{1-c})^{\dim Z}\mathrm {mult}_xZ$ , there exists a rational number $0<c'<1-c$ , an effective $\mathbb Q$ -divisor $D'\sim _{\mathbb Q}c'D$ , such that $(X, cD+D')$ is log canonical at x and the minimal log canonical center $Z'$ of $(X, cD+D')$ through x is a proper subset of Z.

  2. (2) If $m\geq n$ , let $c_1=c+c'$ and $D_1=cD+D'$ . Then, $\frac {b_x(X, D_1)}{1-c_1}<\frac {b_x(X, cD)}{1-c}\leq n$ .

  3. (3) $\mathrm {mult}_xZ\leq \binom {n-\lceil b_x(X, cD) \rceil }{n-d}$ .

Remark 2.4. Let $(A, L)$ be a polarized abelian variety. Assume that an effective $\mathbb Q$ -divisor $D\sim _{\mathbb Q}tL$ such that $(X, D)$ is log canonical around the neutral element o, and Z is the minimal log canonical center through o. Choose an integer $m\geq 1$ such that $\mathcal {I}_Z\otimes L^{\otimes m}$ is globally generated. Then, by a standard argument (see, for instance, [Reference KawamataK, Proposition 2.3]), we know that for a small perturbation $D'=(1-\epsilon )D+\eta H$ of D, where $0<\epsilon , \eta \ll 1$ and $H\in |\mathcal {I}_Z\otimes L^{\otimes m}|$ general, Z is an irreducible component of $\mathrm {Nklt}(A, D')$ . In particular, if $Z=\{o\}$ , we have $\beta (L)\leq r'(L)\leq t$ . When Z is an abelian subvariety of A and $\beta (L|_Z)\leq t$ , we also have $\beta (L)\leq t$ by Ito’s work [Reference ItoIto2, Proposition 6.6].

2.2 A generic vanishing approach

We recall some results from [Reference JiangJz].

Lemma 2.5. Let $(A, L)$ be a polarized abelian variety. Assume that $D\sim _{\mathbb Q} tL$ is an effective $\mathbb Q$ -divisor such that there exists an effective divisor $H\preceq D$ . Then, $\mathcal {I}_{K}\langle tL\rangle $ is GV, where K is the neutral component of the kernel of the morphism

$$ \begin{align*} &\varphi_H: A\rightarrow \operatorname{\mathrm{Pic}}^0(A)\\ & x\rightarrow t_x^*\mathcal{O}_A(H)\otimes \mathcal{O}_A(-H). \end{align*} $$

In particular, when H is ample, $\mathcal {I}_o\langle tL\rangle $ is GV and thus, $\beta (L)\leq t$ .

Proof. This follows directly from the proof of (1) of [Reference JiangJz, Proposition 4.1].

The following theorem is essentially the main result of [Reference JiangJz].

Theorem 2.6. Let $(A, L)$ be a polarized abelian variety. Assume that

  • there exists an irreducible normal subvariety Z of A such that $\mathcal {I}_Z\langle t_0L\rangle $ is IT $^0$ for some positive rational number $t_0$ and a smooth model of Z is of general type;

  • and there exists an effective $\mathbb Q$ -Weil divisor $D_Z$ on Z and a big $\mathbb Q$ -Cartier divisor $V_Z$ such that $K_Z+D_Z$ is $\mathbb Q$ -Cartier and $t_0L|_Z\sim _{\mathbb Q}2(K_Z+D_Z)+V_Z$ .

We have $\beta (L)\leq t_0$ .

Since [Reference JiangJz, Theorem 1.4] is not stated in this way, let’s briefly recall the proof for the readers’ convenience.

Proof. It suffices to show that $\mathcal {I}_0\langle t_0L\rangle $ is GV. After a translation, we may assume that $o\in Z$ is a smooth point of Z. We have the short exact sequence

$$ \begin{align*}0\rightarrow \mathcal{I}_Z\rightarrow \mathcal{I}_o\rightarrow \mathcal{I}_{o, Z}\rightarrow 0.\end{align*} $$

By the first condition, it suffices to show that $\mathcal {I}_{o, Z}\langle t_0L\rangle $ is GV.

We may also assume that Z is smooth (one may check [Reference JiangJz, Subsection 5.1] for the full argument). Then, since Z is of general type, $\omega _Z\otimes \mathcal {I}_{o, Z}$ is GV (see [Reference JiangJz, Lemma 2.6]). Then, one can take an integer M sufficiently divisible such that $Mt_0\in \mathbb Z$ and consider the multiplication-by-M map $\pi _M: A\rightarrow A$ and denote by $Z^M$ the inverse image $\pi _M^{-1}(Z)$ . By the second condition, $\mathcal {O}_{Z^M}(M^2t_0L-K_{Z^M})$ has a nontrivial section s. Indeed, the second condition implies that $M^2t_0L|_{Z^M}-K_{Z^M}$ is $\mathbb Q$ -equivalent to the sum of $K_{Z^M}$ and a big divisor, and one can conclude by Nadel’s vanishing and generic vanishing that $M^2t_0L|_{Z^M}-K_{Z^M}$ has a global section.Footnote 3

Via this section, we have a short exact sequence

$$ \begin{align*} 0\rightarrow K_{Z^M}\xrightarrow{\cdot s} \mathcal{O}_{Z^M}(M^2t_0L)\rightarrow \mathcal{Q}\rightarrow 0. \end{align*} $$

We then check that all terms in this short exact sequence are GV. We may assume that the zero locus of s does not intersect with $o_M:=\pi _M^{-1}(o)$ . Then, we have another short exact sequence

$$ \begin{align*} 0\rightarrow \mathcal{I}_{o_M}\otimes K_{Z^M}\xrightarrow{\cdot s} \mathcal{I}_{o_M}\otimes\mathcal{O}_{Z^M}(M^2t_0L)\rightarrow \mathcal{Q}\rightarrow 0. \end{align*} $$

Note that $\mathcal {I}_{o_M}\otimes K_{Z^M}=\pi _M^{*}(\mathcal {I}_o\otimes K_Z)$ is GV. Hence, $\mathcal {I}_{o_M}\otimes \mathcal {O}_{Z^M}(M^2t_0L)$ is also GV. This implies that $\mathcal {I}_{o,Z}\langle t_0L\rangle $ is GV.

We shall apply Theorem 2.6 in the following way.

Proposition 2.7. Let $(A, L)$ be a polarized simple abelian variety. Assume that an effective $\mathbb Q$ -divisor $D\sim _{\mathbb Q}t_0L$ and $\mathrm {lct}(D, o)<\frac {1}{2}$ . Then, $\beta (L)\leq 2\mathrm {lct}(D, o)t_0<t_0$ .

Proof. We take $c=\mathrm {lct}(D)$ the global log canonical threshold of D (i.e., c is the maximal rational number such that $(A, cD)$ is a log canonical pair). Then, $c\leq \mathrm {lct}(D, o)<\frac {1}{2}$ . After a small perturbation of $cD$ and taking a translation, we may assume that the log canonical pair $(A, cD)$ has only one log canonical center Z, which is thus an irreducible normal subvariety of A and $o\in Z$ . Therefore, the multiplier ideal sheaf $\mathcal {J}(A, cD)=\mathcal {I}_Z$ , and by Nadel’s vanishing, we have $\mathcal {I}_Z\langle 2ct_0L\rangle $ is IT $^0$ . Since A is simple, a smooth model of Z is of general type.

However, by the main theorem of [Reference Fujino and YoshinoriFG], we know that there exists an effective $\mathbb Q$ -Weil divisor $D_Z$ on Z such that $K_Z+D_Z$ is $\mathbb Q$ -Cartier and $cD|_Z\sim _{\mathbb Q}K_Z+D_Z$ . Thus, for any rational number $0<\epsilon \ll 1$ , $(2ct_0+\epsilon )L|_Z\sim _{\mathbb Q} 2(K_Z+D_Z)+\epsilon L|_Z$ . Therefore, by Theorem 2.6, $ \beta (L)\leq 2ct_0+\epsilon $ for any rational number $0<\epsilon \ll 1$ . We then have $\beta (L)\leq 2\mathrm {lct}(D, o)t_0<t_0.$

2.3 Canonical volumes of subvarieties of abelian varieties

Barja, Pardini and Stoppino studied higher dimensional Severi type inequalities for varieties of maximal Albanese dimension in [Reference Barja, Pardini and StoppinoBPS]. One of their main results is the following (see [Reference Barja, Pardini and StoppinoBPS, Corollary 6.12]).

Theorem 2.8. Let $a: X\rightarrow A$ be a morphism from a smooth projective variety of general type of dimension $d\geq 2$ to an abelian variety A. Assume that $a: X\rightarrow a(X)$ is of degree $1$ . Then, $\operatorname {\mathrm {vol}}(K_X)\geq \frac {5}{2}d!\chi (\omega _X) .$

However, Pareschi and Popa generalized the Castelnuovo-De Franchis inequality in [Reference Pareschi Giuseppe and PopaPP]. The following result is a simple corollary of [Reference Pareschi Giuseppe and PopaPP, Theorem 3.3].

Theorem 2.9. Let $a: X\rightarrow A$ be a morphism from a smooth projective variety of general type of dimension $d\geq 2$ to an abelian variety A of dimension g. Assume that A is simple and a is generically finite from X onto its image. Then, $\chi (\omega _X)\geq g-d$ .

Proof. Since a is generically finite onto its image, we have

$$ \begin{align*}\chi(\omega_X)=\chi(a_*\omega_X),\end{align*} $$

and for each $i\geq 0$ ,

$$ \begin{align*}V^i(\omega_X, a):=\{P\in\operatorname{\mathrm{Pic}}^0(A)\mid H^i(X, \omega_X\otimes a^*P)\neq 0\}=V^i(a_*\omega_X)\end{align*} $$

by Grauert-Riemenschneider vanishing (see for instance [Reference LazarsfeldLaz1, Theorem 4.3.9]. In particular, each irreducible component of $V^i(a_*\omega _X)$ is a translation of an abelian subvariety of $\operatorname {\mathrm {Pic}}^0(A)$ by the main result of [Reference Green and LazarsfeldGL].

However, by the main result of [Reference HaconHac], $a_*\omega _X$ is GV. Thus, $V^i(a_*\omega _X)$ is a union of translations of proper abelian subvariety of $\operatorname {\mathrm {Pic}}^0(A)$ for each $i\geq 1$ . Since A is simple by assumption, so is $\operatorname {\mathrm {Pic}}^0(A)$ . Thus, $V^i(a_*\omega _X)$ consists of finitely many points for each $i\geq 1$ . In particular, the neutral element $\mathcal {O}_A$ is an isolated component of $V^i(a_*\omega _X)$ for each $i\geq 1$ . According to [Reference Pareschi Giuseppe and PopaPP, Definition 3.1], the generic vanishing index of $a_*\omega _X$ is $g-d$ . By [Reference Pareschi Giuseppe and PopaPP, Theorem 3.3], we have $\chi (\omega _X)=\chi (a_*\omega _X)\geq g-d $ .

Remark 2.10. In [Reference Lazarsfeld Robert and PopaLP], Lazarsfeld and Popa conjectured that under that assumption of Theorem 2.9, $\chi (\omega _X)> g-d$ when g is large compared to $\chi (\omega _X)$ . Coandă verified that $\chi (\omega _X)\geq 3$ when $g\geq 5$ and $g-d=2$ ([Reference Lazarsfeld Robert and PopaLP, Proposition 4.10])

Combining Theorem 2.8 and 2.9, we have an estimation of the canonical volume of irreducible subvarieties of simple abelian varieties.

Corollary 2.11. Let Z be an irreducible subvariety of dimension d of a simple abelian variety A of dimension g. Let $\rho : X\rightarrow Z$ be a desingularization. Then, $\operatorname {\mathrm {vol}}(K_X)\geq \frac {5}{2}d!(g-d)$ .

When X is a surface, we can slightly improve Theorem 2.8.

Proposition 2.12. Let $a: S\rightarrow A$ be a morphism from a smooth projective surface of general type to an abelian variety A of dimension $g\geq 3$ . Assume that $a: S\rightarrow a(S)$ is of degree $1$ . Then, $\operatorname {\mathrm {vol}}(K_S)\geq 5\chi (\omega _S)+1$ .

Proof. We follow the proof of [Reference Barja, Pardini and StoppinoBPS, Theorem 5.5 and Corollary 5.6], where it was proved that $\operatorname {\mathrm {vol}}(K_S)\geq 5\chi (\omega _S)$ . It suffices to show that the equality can never hold.

We may assume that S is minimal and $a^*: \operatorname {\mathrm {Pic}}^0(A)\rightarrow \operatorname {\mathrm {Pic}}^0(S)$ is injective.

Fix an ample divisor H on A. Following [Reference Barja, Pardini and StoppinoBPS], we consider the continuous rank functions: $F(t):=h^0_a(A, a_*\omega _S\otimes H^t)$ and $G(t):=h^0_a(A, a_*(\omega _S^{\otimes 2})\otimes H^{2t})$ . By definition, continuous rank functions are the $0$ -th cohomological rank functions. Thus, in our terminology, $F(t)=h^0_{a_*\omega _S, H}(t)$ and $G(t)=h^0_{a_*(\omega _S^{\otimes 2}), H}(2t)$ .

Following the proof in [Reference Barja, Pardini and StoppinoBPS, Theorem 5.5], we know that $D^{-1}G(t)\geq 6 D^{-1}F(t)$ for $t\leq 0$ , where $D^{-1}G(t)$ and $D^{-1}F(t)$ are, respectively, the left derivative of the functions F and G at t. We also observe that $G(t)=F(t)=0$ when $t\ll 0$ . Hence, if $G(0)=6F(0)$ (i.e., $\operatorname {\mathrm {vol}}(K_S)= 5\chi (\omega _S)$ ), $G(t)=6F(t)$ for all $t\leq 0$ .

Since $a_*(\omega _S^{\otimes 2})$ is an IT $^0$ sheaf on A, for $-1\ll t<0$ , $a_*(\omega _S^{\otimes 2})\langle 2tH\rangle $ remains to be IT $^0$ (see [Reference Jiang and PareschiJP, Theorem 5.2]). Thus,

$$ \begin{align*}G(t)=\chi(S, 2K_S+2tH)=2(H^2)_St^2+3(K_S\cdot H)_St+K_S^2+\chi(\omega_S)\end{align*} $$

is a degree $2 $ polynomial function for $-1\ll t<0.$

However,

$$ \begin{align*} F(t)&=\chi(S, K_S+tH)+h^1_{a_*\omega_S, H}(t)-h^2_{a_*\omega_S, H}(t)\\ &=\frac{1}{2}(H^2)_St^2+\frac{1}{2}(K_S\cdot H)_St+\chi(\omega_S)+h^1_{a_*\omega_S, H}(t)\\ &\quad - h^2_{a_*\omega_S, H}(t). \end{align*} $$

We recall the formula in [Reference Jiang and PareschiJP, Proposition 2.3]:

$$ \begin{align*} h^i_{a_*\omega_S, H}(t)=\frac{(-t)^g}{\chi(H)}\chi({\varphi}_H^*R^i\Phi_{\mathcal{P}}(a_*\omega_S)\otimes H^{-\frac{1}{t}}), \end{align*} $$

for $-1\ll t<0.$

Since $a^*: \operatorname {\mathrm {Pic}}^0(A)\rightarrow \operatorname {\mathrm {Pic}}^0(S)$ is injective, we know that $R^2\Phi _{\mathcal {P}}(a_*\omega _S)$ is the skyscraper sheaf at $o_{\hat {A}}$ . Hence, $h^2_{a_*\omega _S, H}(t)=\chi (H)(-t)^g$ for $-1\ll t<0.$

Similarly, since S is of general type, $a_*\omega _S$ is a M-regular sheaf on A (see Subsection 2.3). Then, $\operatorname {\mathrm {codim}}_{\operatorname {\mathrm {Pic}}^0(A)}V^1(a_*\omega _S)\geq 2$ . We also know that the support of $R^1\Phi _{\mathcal {P}}(a_*\omega _S)$ is contained in $V^1(a_*\omega _S)$ . Consider the Chern characters of $R^1\Phi _{\mathcal {P}}(a_*\omega _S)$ . We have $\mathrm {ch}_i(\varphi _H^*R^1\Phi _{\mathcal {P}}(a_*\omega _S))=0\in H^{2i}(A, \mathbb Q)$ for $i=0, 1$ . If the codimension of the support of $R^1\Phi _{\mathcal {P}}(a_*\omega _S)$ is equal to $2$ , we may assume that $\mathcal {Z}_1,\ldots , \mathcal {Z}_s$ are the codimension- $2$ components of its support and $a_i>0$ is the rank of $R^1\Phi _{\mathcal {P}}(a_*\omega _S)$ at the generic point of $\mathcal {Z}_i$ . Then, $\mathrm {ch}_2(\varphi _H^*R^1\Phi _{\mathcal {P}}(a_*\omega _S))=\sum _ia_i[\varphi _H^{-1}\mathcal {Z}_i]$ . Hence,

$$ \begin{align*}h^1_{a_*\omega_S, H}(t)=\alpha_{2}t^2+\sum_{3\leq i\leq g}\alpha_it^i\end{align*} $$

for $-1\ll t<0,$ where $\alpha _2\geq 0$ .

We compare the coefficient of $t^2$ for $F(t)$ and $G(t)$ , which are, respectively, $\frac {1}{2}(H^2)_S+\alpha _2$ and $2(H^2)_S$ and conclude that $G(t)\neq 6F(t)$ .

The following corollary seems to be new, and it would be interesting to characterize the surfaces where the equality holds.

Corollary 2.13. Let Z be a smooth projective surface of general type. Assume that $q(Z)\geq 4$ and Z is of maximal Albanese dimension, and either Z is birational to a product of two smooth projective curves of genus $2$ or $\operatorname {\mathrm {vol}}(Z, K_Z)\geq 16$ .

Proof. We first assume that Z is a minimal surface of general type. Then, $\chi (\omega _Z)\geq 1$ and $\operatorname {\mathrm {vol}}(Z, K_Z)=K_Z^2>0$ .

When $\chi (\omega _Z)=1$ , by the main result of [Reference Hacon and PardiniHP] (see also [Reference Jiang, Lahoz and TirabassiJLT]) and the assumption that $q(Z)\geq 4$ , we know that $q(Z)=4$ and Z is birational to a product of two smooth projective curves of genus $2$ .

We now consider the case $\chi (\omega _{Z})= 2$ and $q(Z)\geq 4$ . When $q(Z)\geq 5$ , then $\chi (\omega _Z)<q(Z)-2$ . By the Castelnuovo-de Francis theorem (see, for instance, [Reference Pareschi Giuseppe and PopaPP, Theorem A]), there exists a fibration $f: Z\rightarrow C$ from Z to a smooth projective curve of genus $\geq 2$ . We denote by F a general fiber of f. Then, $g(C)+g(F)\geq q(Z)\geq 5$ (see, for instance, [Reference DebarreD2, the Lemme in the appendix]). We then have

$$ \begin{align*}\operatorname{\mathrm{vol}}(Z, K_Z)=K_Z^2\geq 8(g(C)-1)(g(F)-1)\geq 16\end{align*} $$

(see, for instance, [Reference DebarreD2, the Corollaire in the appendix]). When $q(Z)=4$ , then $p_g(Z)=5$ . By the main result of [Reference Barja, Naranjo and PirolaBNP], we have $16\leq \operatorname {\mathrm {vol}}(Z, K_Z)\leq 18$ .

When $\chi (\omega _Z)\geq 3$ , let $Z'$ be the smooth minimal model of the Albanese image $a_Z(Z)$ . After birational modifications of Z, we have the factorization of the Albanese morphism of Z

$$ \begin{align*}a_Z: Z\xrightarrow{\tau} Z'\xrightarrow{\rho} a_Z(Z)\hookrightarrow A_Z,\end{align*} $$

where $\rho $ is birational.

If $a_Z$ is birational onto its image, we apply Proposition 2.12 to conclude that

$$ \begin{align*}\operatorname{\mathrm{vol}}(Z, K_{Z})\geq 16.\end{align*} $$

We then assume that $\deg \tau>1$ . If $Z'$ is of general type, we have $\operatorname {\mathrm {vol}}(Z, K_Z)\geq (\deg \tau ) K_{Z'}^2\geq 2(K_{Z'}^2)$ by the ramification formula. By the previous discussions, we have already seen that either $Z'$ is birational to a product of two smooth projective curves of genus $2$ or $K_{Z'}^2\geq 16$ . In either case, $\operatorname {\mathrm {vol}}(Z, K_Z)\geq 16$ . Finally, we consider the case that $Z'$ is not of general type. Since $a_Z(Z)$ generates $A_Z$ , we see easily that $\rho $ is an isomorphism and $Z'$ is fibred by an elliptic curve E in $A_Z$ , and the quotient $C':=Z'/E$ is a smooth projective curve of genus $\geq q(Z)-1\geq 3$ . We then have a surjective morphism $Z\rightarrow C'$ from Z to $C'$ . We conclude again by [Reference DebarreD2, the Corollaire in the appendix] that $\operatorname {\mathrm {vol}}(Z, K_Z)\geq 16$ .

2.4 Intersections with abelian subvarieties

When we apply Ito’s approach to study basepoint freeness thresholds of polarized abelian varieties $(A, L)$ , it may happen that the log canonical center Z of a pair $(A, D)$ is an abelian subvariety of A or is a subvariety fibred by an abelian subvariety of A.

Ito realized that one can deal with these cases with Poincaré’s reducibility theorem for polarized abelian varieties. When Z is an abelian subvariety, we have the following result [Reference ItoIto2, Proposition 6.6].

Proposition 2.14. Let $(A, L) $ be a polarized abelian variety. Assume that there exists an effective $\mathbb Q$ -divisor D such that an abelian subvariety B is an irreducible component of $\mathrm {Nklt}(A, D)$ and $tL-D$ is an ample $\mathbb Q$ -divisor. Assume, furthermore, that $\beta (L|_B)<t$ . Then, $\beta (L)<t$ .

When Z is fibred by an abelian subvariety B, it is important to estimate the intersection number $(L^{\dim B}\cdot B)$ . The author learned the following result from a private communication with Atsushi Ito.

Lemma 2.15. Let D be an ample $\mathbb Q$ -divisor on A and let B be a abelian subvariety of A of dimension d. Let $\varphi : A\rightarrow A/B$ be the quotient morphism. Then, there exists an ample $\mathbb Q$ -divisor $H_B$ on $A/B$ such that $D-\varphi ^*H_B$ is a nef $\mathbb Q$ -divisor and $D^g=\binom {g}{d}(D^d\cdot B)(H_B^{g-d})_{A/B}$ .

Proof. Let M be a positive sufficiently divisible integer such that $MD$ is an integral divisor and $L=\mathcal {O}_A(MD)$ is an ample line bundle on A. By Poincaré’s reducibility (see, for instance, [Reference Birkenhake and LangeBL, Corollary 5.3.6]), there exists an abelian subvariety K of A such that the natural morphism $K\rightarrow A/B$ is an isogeny and the addition morphism $\pi : K\times B\rightarrow A$ induces an isogeny of polarized abelian varieties $(K, L_K)\times (B, L_B)\rightarrow (A, L)$ , where $L_K$ and $L_B$ are, respectively, the restriction of L on K and B. Then, one can consider the natural isogeny $\mu _K: K\rightarrow A/B$ . Note that $L|_K$ cannot descend to $A/B$ in general, but there exists an effective $\mathbb Q$ -divisor $D_B$ on $A/B$ such that $L|_K$ is algebraically equivalent to $\mu _K^*D_B$ as $\mathbb Q$ -divisors. Let $H_B=\frac {1}{M}D_B$ , and it is easy to check that $H_B$ satisfies the desired properties.

Corollary 2.16. Let $(A, L)$ be a polarized abelian variety of dimension g with $n(L)<\frac {1}{p+2}$ . Assume that Conjecture 1.4 holds in dimension $\leq g-1$ . Let $D=\frac {1}{p+2}L$ . Then, either

(2) $$ \begin{align} (D^d\cdot B)\geq \frac{(D^g)}{\binom{g}{d}(g-d)^{g-d}} \end{align} $$

for all abelian subvariety B of dimension $1\leq d\leq g-1$ or $\beta (L)<\frac {1}{p+2}$ .

Remark 2.17. Let $0<d<g$ be positive integers. Then,

$$ \begin{align*}\frac{g^g}{d^d\binom{g}{d}}>(g-d)^{g-d}.\end{align*} $$

More precisely, by Stirling’s formula, for any positive integer n, we have

$$ \begin{align*}\sqrt{2\pi}n^{n+\frac{1}{2}}e^{-n+\frac{1}{12n+1}}<n!<\sqrt{2\pi}n^{n+\frac{1}{2}}e^{-n+\frac{1}{12n}}.\end{align*} $$

Thus,

$$ \begin{align*}\frac{g^g}{d^d\binom{g}{d}}>(g-d)^{g-d}\sqrt{\frac{2\pi d(g-d)}{g}}.\end{align*} $$

Proof. If the inequalities (2) fail for some abelian subvarieties, we pick B such that $\dim B=d$ is maximal and $(D^d\cdot B)< \frac {(D^g)}{\binom {g}{d}(g-d)^{g-d}}$ . By Lemma 2.15, there exists an ample $\mathbb Q$ -divisor $H_{B}$ such that $D-\varphi ^*H_{B}$ is a nef $\mathbb Q$ -divisor and

$$ \begin{align*}(H_{B}^{g-d})_{A/B}=\frac{(D^g)}{\binom{g}{d}(D^d\cdot B)}>(g-d)^{g-d}.\end{align*} $$

Moreover, by the argument of Lemma 2.15, for each abelian subvariety C of $A/B$ of dimension $r>0$ , $(H_{B}^r\cdot C)_{A/B}=\frac {(D^{d+r}\cdot \varphi ^{-1}(C))}{\binom {d+r}{d}(D^d\cdot B)}$ . By the maximality of B, we conclude that

$$ \begin{align*} (H_{B}^r\cdot C)_{A/B}=\frac{(D^{d+r}\cdot \varphi^{-1}(C))}{\binom{d+r}{d}(D^d\cdot B)}>\qquad &\frac{ \binom{g}{d}(g-d)^{g-d}}{\binom{d+r}{d}\binom{g}{d+r}(g-d-r)^{g-d-r}}\\ &=\frac{ (g-d)^{g-d}}{(g-d-r)^{g-d-r}\binom{g-d}{r}}\\ & >r^r. \end{align*} $$

Since Conjecture 1.4 holds in dimension $g-d$ , $\mathcal {I}_{o_{A/B}}\langle (1-\epsilon )H_B\rangle $ is GV for some $0<\epsilon \ll 1$ . Thus, $\mathcal {I}_B\langle (1-\epsilon )\varphi ^*H_B\rangle $ is also GV. Since $D-\varphi ^*H_B$ is a nef $\mathbb Q$ -divisor, $\mathcal {I}_B\langle (1-\epsilon )D\rangle $ is GV. We then consider the short exact sequence

$$ \begin{align*}0\rightarrow \mathcal{I}_B\rightarrow \mathcal{I}_o\rightarrow \mathcal{I}_{o, B}\rightarrow 0.\end{align*} $$

Since by the assumption that Conjecture 1.4 holds in dimension d, $\mathcal {I}_{o, B}\langle (1-\epsilon )D\rangle $ is also GV and hence, $\mathcal {I}_o\langle (1-\epsilon )D\rangle $ is GV. Thus, $\beta (L)<\frac {1}{p+2}$ .

Lemma 2.18. Let D be an effective ample $\mathbb Q$ -divisor on an abelian variety A. Assume that $(A, D)$ is log canonical at o, and the minimal log canonical center Z through o is a subvariety of dimension d fibred by an abelian subvariety B of dimension $d'$ such that the desingularization of $Z/B$ is of general type. Then,

$$ \begin{align*}(D^d\cdot Z)\geq \binom{d}{d'}(D^{d'}\cdot B)\operatorname{\mathrm{vol}}(K_{(Z/B)'}),\end{align*} $$

where $(Z/B)'$ is a smooth model of the quotient $Z/B$ .

Proof. We consider the quotient morphism $A\rightarrow A/B$ . We may assume that $MD$ is an integral divisor corresponding to a line bundle L for some integer $M>0$ . By Poincaré’s reducibility theorem (see, for instance, [Reference Birkenhake and LangeBL, Corollary 5.3.6]), there exists an abelian subvariety K of A such that the natural morphism $K\rightarrow A/B$ is an isogeny, and the addition morphism $\pi : K\times B\rightarrow A$ induces an isogeny of polarized abelian varieties $(K, L_K)\times (B, L_B)\rightarrow (A, L)$ , where $L_K$ and $L_B$ are, respectively, the restriction of L on K and B.

Note that $\tilde {Z}:=Z\times _A(K\times B)$ is isomorphic to the product $\tilde {Z/B}\times B$ , where $\tilde {Z/B}=(Z/B)\times _{A/B}K$ . Thus,

$$ \begin{align*} (D^d\cdot Z)&=\frac{1}{M^d}(L^d\cdot Z)=\frac{1}{M^d\deg\pi}((L_K\boxtimes L_B)^d\cdot\tilde{Z})\\ &=\frac{1}{M^d\deg\pi}\big((L_K\boxtimes L_B)^d\cdot(\tilde{Z/B}\times B)\big)\\ &=\frac{\binom{d}{d'}}{M^{d-d'}\deg\pi}(D^{d'}\cdot B)(L_K^{d-d'}\cdot \tilde{Z/B}). \end{align*} $$

Let $\nu : \overline {Z}\rightarrow Z$ be its normalization. By Proposition 2.1, $\nu ^*D\sim _{\mathbb Q}K_{\overline {Z}}+V_{\overline {Z}}$ for some effective $\mathbb Q$ -divisor $V_{\overline {Z}}$ . Consider the pullback of this $\mathbb Q$ -linear equivalence on the normalization $ \overline {\tilde {Z}}$ of $\tilde {Z}$ and then restricting it to a general fiber of $ \overline {\tilde {Z}}\rightarrow B$ . We see that $\varsigma ^*(\frac {1}{M}L_{K})\sim _{\mathbb Q}K_{\widehat {Z/B}}+V_{\widehat {Z/B}} $ , where $\varsigma :\widehat {Z/B}\rightarrow \tilde {Z/B}$ is the normalization and $V_{\widehat {Z/B}}$ is an effective $\mathbb Q$ -divisor on $\widehat {Z/B}$ . Thus, $(L_K^{d-d'}\cdot \tilde {Z/B})= M^{d-d'}\operatorname {\mathrm {vol}}(\widehat {Z/B}, K_{\widehat {Z/B}}+V_{\widehat {Z/B}})\geq (\deg \pi )M^{d-d'}\operatorname {\mathrm {vol}}(K_{(Z/B)'})$ .

3 Very general polarized abelian varieties

In [Reference JiangJz], we applied Helmke’s induction to confirm that Conjecture 1.4 holds for Hodge theoretically very general polarized abelian varieties with special polarizations. We now show that the calculation indeed implies that Conjecture 1.4 holds for almost all generic polarized abelian varieties in fixed dimensions.

Recall that we say a polarized abelian variety $(A, L)$ is Hodge theoretically very general if $\dim _{\mathbb Q}H^{k,k}(A, \mathbb Q)=1$ for all $1\leq k\leq g-1$ . We observe that by Hard Lefschetz and Poincaré duality, $(A, L)$ is Hodge theoretically very general iff $\dim _{\mathbb Q}H^{\lfloor \frac {g}{2}\rfloor ,\lfloor \frac {g}{2}\rfloor }(A, \mathbb Q)=1$ . We also observe that when $(A, L)$ is Hodge theoretically very general, A is a simple abelian variety and hence, $n(L)=\displaystyle\frac {g}{\sqrt [g]{(L^g)}}$ . Note that in order to compare $\beta (L)$ and $n(L)$ , we can assume that L is primitive.

Theorem 3.1. Let $(A, L)$ be a Hodge theoretically very general polarized abelian variety of type $l=(1,\delta _2,\ldots ,\delta _g)$ . Assume that

$$ \begin{align*} \delta:=\delta_2\cdots\delta_g\geq \mathrm{ max}\big\{\frac{(k(k+1)\cdots(g-1))^{\frac{g(g-1)}{k}}}{(g!)^{\frac{(g-1)(g-k)}{k}}}\mid 2\leq k\leq g-2\big\}. \end{align*} $$

Then, $\beta (L) \leq n(L).$

Proof. We apply Ito’s strategy that it suffices to show $r'(L)\leq n(L)$ . Let $t\in (n(L), n(L)+\epsilon )$ be a rational number, where $0<\epsilon \ll 1$ and denote by $D\sim _{\mathbb Q} tL$ an effective rational number such that $\mathrm {mult}_0D>g$ (there exists such D since $((tL)^g)>g^g$ ). We then just need to apply Proposition 2.3 to get a divisor $D'\sim _{\mathbb Q}cD$ with $0<c<1$ , and the neutral element o of A is a minimal log canonical center of $(A, D')$ . By Proposition 2.3, it suffices to verify that $(D^{k}\cdot Z)>\binom {g-1}{g-k}g^k$ for any irreducible subvariety Z of dimension k for $1\leq k\leq g-1$ .

Recall that $\beta (L)\leq 1$ and equality holds if and only if $|L|$ has a basepoint. Thus, we may assume that $(L^g)>g^g$ (i.e., $n(L)<1$ ). Then, since the class $[L]$ is the generator of $\mathrm {NS}(A)=\operatorname {\mathrm {Pic}}(A)/\operatorname {\mathrm {Pic}}^0(A)$ , we can assume that Z is of codimension $\geq 2$ . Since A is simple, by the main result of [Reference DebarreD1], for any irreducible curve Z of A, $(L\cdot Z)>\sqrt [g]{(L^g)}>g$ . Thus, it suffices to verify that $(D^{k}\cdot Z)>\binom {g-1}{g-k}g^k$ for any irreducible subvariety Z of dimension k, where $2\leq k\leq g-2$ .

For any irreducible subvariety Z of dimension k, we denote by $[Z]\in H^{g-k, g-k}(A, \mathbb Z)\subset H^{2g-2k}(A, \mathbb C)$ its cohomology class. Then, $[Z]$ is a positive integral multiple of $\frac {L^{g-k}}{(g-k)!\delta _2\cdots \delta _{g-k}}$ by the assumption that A is Hodge theoretically very general. Thus,

$$ \begin{align*} (D^k\cdot Z)> \frac{g^k}{(L^g)^{\frac{k}{g}}}(L^k\cdot Z) \geq g^k\frac{(L^g)^{\frac{g-k}{g}}}{(g-k)!\delta_2\cdots\delta_{g-k}}. \end{align*} $$

We just need to verify that

(3) $$ \begin{align} \frac{(L^g)^{\frac{g-k}{g}}}{(g-k)!\delta_2\cdots\delta_{g-k}}\geq \binom{g-1}{g-k}. \end{align} $$

Note that $\delta _2\cdots \delta _{g-k}\leq \delta ^{\frac {g-k-1}{g-1}}$ . Thus,

$$ \begin{align*} \frac{ (L^g)^{\frac{g-k}{g}}}{(g-k)!\delta_2\cdots\delta_{g-k}} \geq &\delta^{\frac{g-k}{g}-\frac{g-k-1}{g-1}}\frac{ (g!)^{\frac{g-k}{g}}}{(g-k)!}\\ &\ \ =\delta^{\frac{k}{g(g-1)}}\frac{ (g!)^{\frac{g-k}{g}}}{(g-k)!}. \end{align*} $$

The assumption on $\delta $ makes sure that $ \frac {(L^g)^{\frac {g-k}{g}}}{(g-k)!\delta _2\cdots \delta _{g-k}}\geq \binom {g-1}{g-k}$ for $2\leq k\leq g-2$ .

By the above theorem, we see that in order to verify Conjecture 1.1 and Conjecture 1.4 for very general polarized abelian varieties, it suffices to check finitely many families in each dimension. By the same computation with some extra efforts, we finish the proof of Theorem 1.5.

Proof. When $g=4$ , the assumption in Theorem 3.1 is simply that $\delta =h^0(A, L)\geq \frac {6^6}{24^3}=\frac {27}{8}$ . However, for $\delta \leq 3$ , $n(L)=\frac {4}{\sqrt [4]{24\delta }}>1$ we know that $|L|$ has basepoints and $\beta (L)=1$ .

When $g=5$ , we repeat the argument in the proof of Theorem 3.1. We first check (3) for $g=5$ and $k=2$ or $3$ , which are $(L^5)^{\frac {3}{5}}\geq 24\delta _2\delta _3$ and $(L^5)^{\frac {2}{5}}\geq 12\delta _2$ . It is easy to verify that both inequalities hold when $\delta _5\geq 5$ or $\delta _2\geq 3$ . When $\delta _2=1$ , $\delta \geq 5$ implies that both inequalities hold and if $\delta <5$ , $\beta (L)=1$ . When $\delta _2=2$ and $\delta _5\leq 4$ , the above inequalities fail only when the polarization type is $(1,2,2,2,2)$ . But in this case, $n(L)=\displaystyle\frac {5}{\sqrt [5]{120\delta (L)}}>1$ . Thus, we still have $\beta (L)\leq 1<n(L)$ .

When $g=6$ , we first remark that we may assume that l is not of the form $(1,\ldots , 1, \delta _6)$ . If l is of the form $(1,\ldots , 1, \delta _6)$ , since $(A, L)$ is very general, by [Reference ItoIto3, Theorem 1.5], $\beta (L)\leq \displaystyle\frac {1}{\lfloor \sqrt [6]{\delta _6}\rfloor }<\frac {6}{\sqrt [6]{720\delta _6}}$ . Thus, we will assume that $\delta _5\geq 2$ . We then check (3) for $k=2$ , $3$ and $4$ , which are $(L^6)\geq (20\delta _2)^3$ , $(L^6)\geq (60\delta _2\delta _3)^2$ and $(L^6)^{\frac {2}{3}}\geq 120\delta _2\delta _3\delta _4$ . Since $\delta _5\geq 2$ , we see that these inequalities hold when $\delta _6\geq 5$ . When $\delta _6=4$ , one can verify that these inequalities hold, except the polarization types $(1,4,4,4,4,4)$ or $(1,1, 1,1,2,4)$ or $(1, 2, 2, 2, 2, 2, 4)$ . When $\delta _6=3$ , these inequalities hold, except the polarization types $(1,3,3,3,3,3)$ . Note that for the polarization types $(1, 1, 1,1, 2,4)$ , or $(1,2,2,2,2,4)$ or with $\delta _6\leq 2$ , $n(L)>1$ and thus, $\beta (L)<n(L)$ . We finally need to consider the polarization types $(1,4,4,4,4,4)$ and $(1,3,3,3,3,3)$ . These two cases can be dealt by a result of Ito. When $(A, L)$ is a polarized abelian sixfold of polarization type $(1,4,4,4,4,4)$ (resp. $(1,3,3,3,3,3)$ ), let $(S, L')$ be a very general polarized abelian surface of polarization type $(1, 4)$ (resp. $(1, 3)$ ). By [Reference ItoIto5, Proposition 5.1], we always have $\beta (L)\leq \beta (L')$ . It is known that $\beta (L')=\frac {1}{2}$ (resp. $\frac {2}{3}$ ) (see [Reference ItoIto3, Proposition 4.3 and Lemma A.4]). We then still have $\beta (L)<n(L)$ for the polarization types $(1,4,4,4,4,4)$ and $(1,3,3,3,3,3)$ .

Remark 3.2. The last paragraph of the above proof is due to the anonymous referee.

4 Abelian fourfolds

4.1 The proof of Theorem 1.6

Proof. It suffices to show that $\beta (L)<\frac {1}{p+2}$ or equivalently $\mathcal {I}_o\langle \frac {1}{p+2}L\rangle $ is IT $^0$ , where o is the neutral element of A.

Since $(L^4)>((2+\frac {4}{\sqrt {3}})(p+2))^4$ , there exists an effective $\mathbb Q$ -divisor $D\sim _{\mathbb Q}\frac {1}{p+2}L$ such that $\mathrm {mult}_o(D)=m>2+\frac {4}{\sqrt {3}}$ . Let $c=\mathrm {lct}(D, o)\leq \frac {4}{m}$ be the log canonical threshold of D at o and let Z be the minimal log canonical center of $(A, cD)$ through o. By Proposition 2.7, we may assume that $c\geq \frac {1}{2}$ .

Step 1. We first deal with the case that Z is a divisor. By Lemma 2.5, $\mathcal {I}_K\langle \frac {c}{p+2}L\rangle $ is GV, where K is the kernel of $\varphi _{Z}$ . If K is a point, we are done. Otherwise, by the assumption that $(L^d\cdot B)>(d(p+2))^d$ for any abelian subvariety B of dimension $1\leq d\leq 3$ and Ito’s results [Reference ItoIto1, Reference ItoIto2], $\mathcal {I}_{o,K}\langle \frac {1}{p+2}L\rangle $ is IT $^0$ . Thus, from the short exact sequence

$$ \begin{align*}0\rightarrow \mathcal{I}_K\rightarrow \mathcal{I}_o\rightarrow \mathcal{I}_{o, K}\rightarrow 0,\end{align*} $$

we conclude that $\mathcal {I}_o\langle \frac {1}{p+2}L\rangle $ is IT $^0$ .

Step 2. If Z is a curve, by Proposition 2.3, Z is smooth at o and as soon as $(D\cdot Z)>4\geq \frac {b_o(A, cD)}{1-c}$ , there exists $D_1\sim _{\mathbb Q}c_1D$ with $c<c_1<1$ such that $(A, D_1)$ is log canonical at o and $\{o\}$ is a minimal log canonical center. Then, we have by Remark 2.4 that $\beta (L)\leq r'(L)\leq \frac {c_1}{p+2}<\frac {1}{p+2}$ .

By Corollary 2.2, we know that

$$ \begin{align*}(cD\cdot Z)\geq 2g(\overline{Z})-2,\end{align*} $$

where $\nu : \overline {Z}\rightarrow Z$ is the normalization. Thus, we are done once $g(\overline {Z})\geq 3$ .

When $g(\overline {Z})=1$ , $Z=\overline {Z}$ is an elliptic curve. We conclude again by Proposition 2.14, Remark 2.4 and the assumption that $(L\cdot Z)>p+2$ and thus, $\beta (L|_Z)<\frac {1}{p+2}$ .

If $g(\overline {Z})=2$ , Z generates an abelian surface B of A. By Corollary 2.16, we may assume that $(D^2\cdot B)\geq \frac {4^4}{\binom {4}{2}2^2}= \frac {32}{3}$ . Then, by Hodge index theorem,

$$ \begin{align*}(D\cdot Z)_B\geq \sqrt{(Z^2)_B(D^2\cdot B)} \geq 8/\sqrt{3}>4.\end{align*} $$

Step 3. When Z is a surface, we need to apply Helmke’s induction. We may assume that Z is not an abelian surface. Otherwise, we conclude directly by Proposition 2.14 and Ito’s work in [Reference ItoIto1]. We know that $\mathrm {mult}_oZ\leq 3$ by Proposition 2.3. Let $\mu : \tilde {Z}\rightarrow \overline {Z}\rightarrow Z$ be the minimal resolution of the normalization $\overline {Z}$ of Z. Since Z is not an abelian variety, $\tilde {Z}$ is a surface of maximal Albanese dimension of Kodaira dimension $\geq 1$ . Since there exist no rational curves on $\overline {Z}$ , $\tilde {Z}$ is minimal.

Claim. $((cD)^2\cdot Z)\geq 16$ when $\mathrm {mult}_oZ=3$ ; $((cD)^2\cdot Z)\geq 16$ or $(D^2\cdot Z)\geq 16\sqrt [3]{6}$ when $\mathrm {mult}_oZ=2$ ; $(D^2\cdot Z)>16$ when Z is smooth at o.

When $\tilde {Z}$ is not of general type, Z is fibred by an elliptic curve E, and since Z has rational singularities around o, Z is indeed smooth at o. We need to show that $(D^2\cdot Z)>16$ . Let $C=Z/E$ be the quotient and $\tilde {C}$ be the normalization of C. By Corollary 2.16, we may assume $(D\cdot E)\geq \frac {64}{27}$ and by Lemma 2.18, we have

$$ \begin{align*}(D^2\cdot Z)\geq 2(2g(\tilde{C})-2)(D\cdot E).\end{align*} $$

Thus, we are done when $g(\tilde {C})\geq 3$ . When $g(\tilde {C})=2$ , Z generates an abelian $3$ -fold B of A and $C\hookrightarrow B/E$ is an ample divisor. Thus, $(C^2)_{B/E}=a\geq 2$ and hence, $Z^2$ is algebraically equivalent to $aE$ as $1$ -cycles of B. By Corollary 2.16, we may assume that $(D^3\cdot B)\geq 4^3$ . Then, by Hodge index,

$$ \begin{align*} (D^2\cdot Z)=((D|_B)^2\cdot Z)_B\quad &\geq \sqrt{(D^3)_B(D|_B\cdot Z^2)_B}\\ & \geq \sqrt{4^3\cdot a\cdot \frac{64}{27}}\geq 16\sqrt{\frac{32}{27}}>16. \end{align*} $$

We now assume that $\tilde {Z}$ is of general type.

If $q(\tilde {Z})\geq 4$ , by Corollary 2.13, $\operatorname {\mathrm {vol}}(K_{\tilde {Z}})\geq 16$ or $\tilde {Z}\simeq C_1\times C_2$ , where $C_i$ is a smooth projective curve of genus $2$ . In the latter case $\mu $ is the normalization of Z, and since Z is normal at o, it is smooth at o. We apply Proposition 2.1 and conclude that

$$ \begin{align*} (D^2\cdot Z)=(\mu^*(D)^2)_{\tilde{Z}}\quad & \geq (K_{\tilde{Z}}\cdot \mu^*D)_{\tilde{Z}} \\ &=2(C_1\cdot \mu^*D)_{\tilde{Z}} +2(C_2\cdot \mu^*D)_{\tilde{Z}}. \end{align*} $$

Note that the image of $C_i$ generates an abelian surface $B_i$ of A. By Corollary 2.16, we may assume that $(D^2\cdot B_i)\geq \frac {32}{3}$ . Thus, by Hodge index, $(C_i\cdot \mu ^*D)_{\tilde {Z}}\geq 8/\sqrt {3}$ . Therefore, we have $(D^2\cdot Z)\geq 32/\sqrt {3}>16$ . In the former case, $ ((cD)^2\cdot Z)=\operatorname {\mathrm {vol}}(Z, cD)\geq \operatorname {\mathrm {vol}}(K_{\tilde {Z}})\geq 16$ by Corollary 2.2.

If $q(\tilde {Z})=3$ , Z generates an abelian 3-fold $B\subset A$ . Then, Z is an ample divisor of B. Moreover, in this case, the embedded dimension of Z at o is at most $3$ . Thus, $\mathrm {mult}_o(Z)\leq 2$ by the well-known facts about isolated rational surface singularities (see [Reference ArtinA, Corollary 6]). As before, by Corollary 2.16, we may assume that $(D^3\cdot B)\geq 4^3$ . Since Z is an ample divisor of B, $(Z^3)_B\geq 6$ . Thus,

$$ \begin{align*}(D^2\cdot Z)\geq \sqrt[3]{(D^3\cdot B)^2(Z^3)_B}\geq 16\sqrt[3]{6}.\end{align*} $$

Step 4.

If Z is smooth at o, we have shown that $(D^2\cdot Z)>16$ . Thus, by Proposition 2.3, there exists an effective $\mathbb Q$ -divisor $D_1\sim _{\mathbb Q} c_1D$ with $c_1<1$ such that $(A, D_1)$ is log canonical at o whose minimal log canonical center $Z_1$ through o is a proper subset of Z and $\frac {b_o(A, D_1)}{1-c_1}\leq \frac {b_o(A, cD)}{1-c}\leq 4$ . We then finish the proof by going back to Step 2.

If $\mathrm {mult}_oZ= 3$ , we have already seen that $((cD)^2\cdot Z)\geq 16$ . In order to apply Helmke’s induction, we need to verify that

$$ \begin{align*}(D^2\cdot Z)>3\big(\frac{b_o(A, cD)}{1-c}\big)^2.\end{align*} $$

Note that $b_o(A, cD)\leq 4-cm<4-(2+\frac {4}{\sqrt {3}})c$ . It is elementary to verify that

$$ \begin{align*}\frac{16}{c^2}\geq 3\big(\frac{4-(2+\frac{4}{\sqrt{3}})c}{1-c}\big)^2\end{align*} $$

always holds. We then finish the proof as before.

If $\mathrm {mult}_oZ=2$ and $((cD)^2\cdot Z)\geq 16$ , we conclude as the multiplicity $3$ case. If $\mathrm {mult}_oZ=2$ and $(D^2\cdot Z)\geq 16\sqrt [3]{6}$ , we need to verify that

$$ \begin{align*}(D^2\cdot Z)>2\big(\frac{b_o(A, cD)}{1-c}\big)^2.\end{align*} $$

Since $c\geq \frac {1}{2}$ , we have $\frac {b_o(A, cD)}{1-c}<4-\frac {(\frac {4}{\sqrt {3}}-2)c}{1-c}\leq 6-\frac {4}{\sqrt {3}}$ . We then check that $16\sqrt [3]{6}>2(6-\frac {4}{\sqrt {3}})^2$ .

4.2 The proof of Theorem 1.8

We apply the same strategy as in the proof of Theorem 1.6.

Fix an effective $\mathbb Q$ -divisor such that $D\sim _{\mathbb Q}\frac {1}{p+2}L$ , such that $\mathrm {mult}_oD>8$ . Let $c_1=\mathrm {lct}(D, o)<\frac {5}{8}$ be the log canonical threshold of D at o and let $Z_1$ be the minimal log canonical center of $(A, c_1D)$ at o. By Proposition 2.7, we may assume that $c_1\geq \frac {1}{2}$ . Thus,

$$ \begin{align*}\frac{b_o(A, c_1D)}{1-c_1}\leq \frac{5-c_1\mathrm{mult}_oD}{1-c_1}<5-3\frac{c_1}{1-c_1}\leq 2.\end{align*} $$

If $Z_1$ is a divisor, we conclude by Lemma 2.5.

If $Z_1$ is a threefold, we apply Helmke’s induction. Let $\rho : \tilde {Z_1}\rightarrow Z_1$ be a desingularization. Then, $\operatorname {\mathrm {vol}}(K_{\tilde {Z_1}})\geq \frac {5}{2}\times 3!\times 3=45$ by Theorem 2.8, Theorem 2.9 and Remark 2.10. We also note that $\mathrm {mult}_oZ_1\leq 6$ . Note that

$$ \begin{align*}(D^3\cdot Z_1)\geq \operatorname{\mathrm{vol}}(K_{\tilde{Z_1}})/c_1^3>\frac{45}{(\frac{5}{8})^3}\geq 48>\big(\frac{b_o(A, c_1D)}{1-c_1}\big)^3\mathrm{mult}_{o}(Z_1).\end{align*} $$

Thus, there exists an effective $\mathbb Q$ -divisor $ D_2\sim _{\mathbb Q}c_2D$ such that $c_1<c_2<1$ , such that $(A, D_2)$ is log canonical at o, whose minimal lc center through o is a proper subvariety $Z_2$ contained in $Z_1$ , and $\frac {b_o(A, D_2)}{1-c_2}<\frac {b_o(A, c_1D)}{1-c_1}<2$ .

When $Z_2$ is a surface, we have $\mathrm {mult}_oZ_2\leq 4$ . Let $\tilde {Z}_2$ be its smooth model. Since A is simple, $\tilde {Z}_2$ is of general type and hence, $\operatorname {\mathrm {vol}}(K_{\tilde {Z_2}})\geq 16$ by Corollary 2.13. We then have

$$ \begin{align*}(D^2\cdot Z_2)>\operatorname{\mathrm{vol}}(K_{\tilde{Z_2}})\geq 16>\big(\frac{b_o(A, D_2)}{1-c_2}\big)^2\mathrm{mult}_oZ_2.\end{align*} $$

Thus, by Helmke’s induction, we may assume that $Z_2$ is a curve. In this case, we verify easily that $(D\cdot Z_2)>\frac {b_o(A, D_2)}{1-c_2}$ . Therefore, there exists an effective $\mathbb Q$ -divisor $D_3\sim _{\mathbb Q}c_3L$ with $c_2<c_3<1$ such that $(A, D_3)$ is log canonical at o and o is a minimal log canonical center of $(A, D_3)$ . We then finish the proof of Theorem 1.8.

4.3 The proof of Theorem 1.9

The proof of Theorem 1.9 is identical to that of Theorem 1.8. We first observe that $\alpha _g\leq \alpha _{g,2}=\sqrt {\frac {5(g-2)+1}{g-1}}<\sqrt {5}$ . We then fix an effective $\mathbb Q$ -divisor $D\sim \frac {1}{p+2}L$ such that $\mathrm {mult}_oD>2g-\alpha _g$ and thus let $c<\frac {g}{2g-\sqrt {5}}$ be the log canonical threshold of $(A, D)$ at o. We may assume that $c\geq \frac {1}{2}$ by Proposition 2.7. Then

$$ \begin{align*}\frac{b_o(A, cD)}{1-c}\leq \frac{g-c (\mathrm{mult}_oD)}{1-c}\leq g-(\mathrm{mult}_oD-g)<\alpha_g.\end{align*} $$

Let Z be the minimal log canonical threshold of $(A, cD)$ through o. Let $d=\dim Z$ . Then, $\mathrm {mult}_oD\leq \binom {g-1}{g-d}$ .

If $d=g-1$ , we conclude by Lemma 2.5. Thus, we may assume that $1\leq d\leq g-2$ .

For a smooth model $\tilde {Z}$ of Z, by Theorem 2.8, Theorem 2.9, Remark 2.10 and Proposition 2.12, we have $\operatorname {\mathrm {vol}}(K_{\tilde {Z}})\geq \frac {15}{2}(g-2)!$ when $d=g-2$ , $\operatorname {\mathrm {vol}}(K_{\tilde {Z}})\geq \frac {5}{2}d!(g-d)$ when $3\leq d\leq g-3$ , $\operatorname {\mathrm {vol}}(K_{\tilde {Z}})\geq 5(g-2)+1$ when $d=2$ and $\operatorname {\mathrm {vol}}(K_{\tilde {Z}})\geq 2(g-1)$ when $d=1$ . By Corollary 2.2, $(D^d\cdot Z)\geq \frac {1}{c^d}\operatorname {\mathrm {vol}}(K_{\tilde {Z}})>\operatorname {\mathrm {vol}}(K_{\tilde {Z}})$ . We then have

$$ \begin{align*}(D^d\cdot Z)> (\alpha_g)^d\mathrm{mult}_oD.\end{align*} $$

We then repeatedly apply Helmke’s induction and the above calculation to cut down the log canonical centers and finish the proof.

Acknowledgements

The author thanks Atsushi Ito for helpful communications and, in particular, Lemma 2.15 is due to him. The author is grateful to the anonymous referee who provides an argument to improve Theorem 1.5 and many helpful comments. The author also thanks Chen Jiang for helpful conversations and Federico Caucci for helpful comments.

Competing interest

The authors have no competing interest to declare.

Financial support

The author is supported by NSFC for Innovative Research Groups (No. 12121001), by the National Key Research and Development Program of China (No. 2020YFA0713200), by the Natural Science Foundation of China (No. 11871155 and No. 11731004) and by the Natural Science Foundation of Shanghai (No. 21ZR1404500).

Footnotes

1 Note that the notion of a pure log canonical center in [Reference JiangJx] coincides with the notion of a log canonical center here.

2 This is called the deficit of $(X, cD)$ at x in [Reference EinE] and [Reference Ye and ZhuYZ].

3 Without the coefficient $2$ in the second condition, we only know that $M^2t_0L|_{Z^M}-K_{Z^M}$ is $\mathbb Q$ -equivalent to an effective divisor, and this does not imply that the line bundle $M^2t_0L|_{Z^M}-K_{Z^M}$ has a global section.

References

Artin, M., ‘On isolated rational singularities of surfaces’, Amer. J. Math. 88 (1966), 129136.CrossRefGoogle Scholar
Angehrn, U. and Siu, Y.-T., ‘Effective freeness and point separation for adjoint bundles’, Invent. Math. 122(2) (1995), 291308.CrossRefGoogle Scholar
Birkenhake, C. and Lange, H., Complex Abelian Varieties (Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]) vol. 302, second edn., (Springer-Verlag, Berlin, 2004).Google Scholar
Barja, M. Á., Naranjo, J. C. and Pirola, G. P., ‘On the topological index of irregular surfaces’, J. Algebraic Geom. 16(3) (2007), 435458.CrossRefGoogle Scholar
Barja, M. Á., Pardini, R. and Stoppino, L., ‘Linear systems on irregular varieties’, J. Inst. Math. Jussieu 19(6) (2020), 20872125.CrossRefGoogle Scholar
Caucci, F., ‘The basepoint-freeness threshold and syzygies of abelian varieties’, Algebra Number Theory 14(4) (2020), 947960.CrossRefGoogle Scholar
Debarre, O., ‘Degree of curves in abelian varieties’, Bull. Soc. Math. Fr. 122 (1994), 101119.Google Scholar
Debarre, O., ‘Inéegalités numériques pour les surfaces de type général [Numerical inequalities for surfaces of general type]’, Bull. Soc. Math. France 110(3) (1982), 319346. With an appendix by A. Beauville.CrossRefGoogle Scholar
Ein, L., ‘Multiplier ideals, vanishing theorems and applications’, in Algebraic Geometry Santa Cruz 1995 (Proc. Sympos. Pure Math.) vol. 62, part 1, (Amer. Math. Soc., Providence, RI, 1997), 203219.CrossRefGoogle Scholar
Fujino, O., ‘Fundamental theorems for the log minimal model program’, Publ. Res. Inst. Math. Sci. 47(3) (2011), 727789.CrossRefGoogle Scholar
Fujino, O. and Yoshinori, G., ‘On canonical bundle formulas and subadjunctions’, Michigan Math. J. 61(2) (2012), 255264.CrossRefGoogle Scholar
Green, M. and Lazarsfeld, R., ‘Higher obstructions to deforming cohomology groups of line bundles’, J. Amer. Math. Soc. 4(1) (1991), 87103.CrossRefGoogle Scholar
Hacon, C. D., ‘A derived category approach to generic vanishing’, J. Reine Angew. Math. 575 (2004), 173187.Google Scholar
Hacon, C. D. and Pardini, R., ‘Birational characterization of products of curves of genus 2’, Math. Res. Lett. 12(1) (2005), 129140.CrossRefGoogle Scholar
Helmke, S., ‘On Fujita’s conjecture’, Duke Math. J. 88(2) (1997), 201216.CrossRefGoogle Scholar
Ito, A., ‘A remark on higher syzygies of abelian surfaces’, Comm. Algebra 46(12) (2018), 53425347.CrossRefGoogle Scholar
Ito, A., ‘Basepoint-freeness thresholds and higher syzygies on abelian threefolds’, Algebr. Geom. 9(6) (2022), 762787.CrossRefGoogle Scholar
Ito, A., ‘Higher syzygies on general polarized abelian varieties of type (1, … , 1, d)’, arXiv:2011.09687.Google Scholar
Ito, A., ‘M-regularity of Q-twisted sheaves and its application to linear systems on abelian varieties’, Trans. Amer. Math. Soc. 375(9) (2022), 66536673.Google Scholar
Ito, A., ‘Projective normality and basepoint-freeness thresholds of general polarized abelian varieties’, arXiv:2204.10035.Google Scholar
Jiang, X., ‘On the pluricanonical maps of varieties of intermediate Kodaira dimension’, Math. Ann. 356(3) (2013), 9791004.CrossRefGoogle Scholar
Jiang, Z., Lahoz, M. and Tirabassi, S., ‘Characterization of products of theta divisors’, Compos. Math. 150(8) (2014), 13841412.CrossRefGoogle Scholar
Jiang, Z., and Pareschi, G., ‘Cohomological rank functions on abelian varieties’, Ann. Sci. Éc. Norm. Supér. 53(4) (2020), 815846.CrossRefGoogle Scholar
Jiang, Z., ‘Cohomological rank functions and Syzygies of abelian varieties’, Math. Z. 300(4) (2022), 33413355.CrossRefGoogle Scholar
Kawamata, Y., ‘On Fujita’s freeness conjecture for 3-folds and 4-folds’, Math. Ann. 308(3) (1997), 491505.CrossRefGoogle Scholar
Lazarsfeld, R., ‘Positivity’, in Algebraic Geometry I (Classical setting: line bundles and linear series. Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]) vol. 48 (Springer-Verlag, Berlin, 2004).Google Scholar
Lazarsfeld, R., ‘Positivity’ in Algebraic Geometry II (Positivity for vector bundles, and multiplier ideals. Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]) vol. 49 (Springer-Verlag, Berlin, 2004).Google Scholar
Lozovanu, V., ‘Singular divisors and syzygies of polarized abelian threefolds’, arXiv:1803.08780v2.Google Scholar
Lazarsfeld Robert, R., and Popa, M., ‘Derivative complex, BGG correspondence, and numerical inequalities for compact K¨ahler manifolds’, Invent. Math. 182(3) (2010), 605633.CrossRefGoogle Scholar
Lahoz, M. and Rojas, A., ‘Chern degree functions’, Commun. Contemp. Math., arXiv:2105.03263.Google Scholar
Pareschi Giuseppe, G., and Popa, M., ‘Strong generic vanishing and a higher-dimensional Castelnuovo-de Franchis inequality’, Duke Math. J. 150(2) (2009), 269285.Google Scholar
Rojas, A., ‘The basepoint-freeness threshold of a very general abelian surface’, Sel. Math. New Ser. 28(34) (2022). URL: https://doi.org/10.1007/s00029-021-00757-9.CrossRefGoogle Scholar
Ye, F. and Zhu, Z., ‘On Fujita’s freeness conjecture in dimension 5’, Adv. Math. 371 (2020), 107210. With an appendix by Jun Lu.CrossRefGoogle Scholar