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

Kawamata–Miyaoka type inequality for \mathbb{Q}-Fano varieties with canonical singularities

Haidong Liu Haidong Liu, Sun Yat-Sen University, Department of Mathematics, Guangzhou, 510275, China [email protected],[email protected] https://sites.google.com/view/liuhaidong  and  Jie Liu Jie Liu, Institute of Mathematics, Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing, 100190, China [email protected] http://www.jliumath.com
Abstract.

Let XX be an nn-dimensional normal \mathbb{Q}-factorial projective variety with canonical singularities and Picard number one such that XX is smooth in codimension two, KX-K_{X} is ample and n2n\geq 2. We prove that XX satisfies the following Kawamata–Miyaoka type inequality:

c1(X)n<4c2(X)c1(X)n2.c_{1}(X)^{n}<4c_{2}(X)\cdot c_{1}(X)^{n-2}.

If additionally XX is a threefold with terminal singularities, then a stronger inequality is also obtained.

Key words and phrases:
Kawamata–Miyaoka type inequality, Fano varieties, Fano threefolds, second Chern class
2020 Mathematics Subject Classification:
Primary 14J45; Secondary 14J10, 14J30.

1. Introduction

Throughout this paper, we work over the complex number field \mathbb{C}. We denote the Picard number of a normal projective variety XX by ρ(X)\rho(X). A normal projective variety is called \mathbb{Q}-Fano (resp. weak \mathbb{Q}-Fano) if it is \mathbb{Q}-factorial and its anti-canonical divisor is ample (resp. nef and big). According to the minimal model program (MMP), \mathbb{Q}-Fano varieties with mild singularities (e.g. terminal, canonical, klt, etc.) are one of the building blocks of algebraic varieties.

According to [Bir21, Theorem 1.1], \mathbb{Q}-Fano varieties with canonical singularities and fixed dimension form a bounded family. Thus one may ask if we can find the full list of them. This goal is only achieved in the large index cases. The key ingredient is to produce special divisors on XX and then the problem is reduced to surfaces [Muk89, San96, Mel99]. One may expect to apply similar ideas in the general case to reduce the classification problem of higher dimensional \mathbb{Q}-Fano varieties to the classification of some special lower dimensional varieties. The very first step of this approach is to prove the existence of certain effective divisors using the Riemann–Roch formula. To this end, one needs to control the positivity of higher Chern classes. For instance, for any \mathbb{Q}-Fano threefold XX with terminal singularities and ρ(X)=1\rho(X)=1, Kawamata proved in [Kaw92, Proposition 1] that there exists a positive number b3b_{3} independent of XX such that

c1(X)3b3c2(X)c1(X).c_{1}(X)^{3}\leq b_{3}c_{2}(X)\cdot c_{1}(X).

This inequality plays a prominent role in the classification of \mathbb{Q}-Fano threefolds with terminal singularities and Picard number one. We refer the reader to [Suz04, Pro13, Pro22, BK22] and the references therein for more details.

In any dimension n2n\geq 2, according to [IJL23, Corollary 1.7], there exists a positive number bnb_{n} depending only on nn such that any nn-dimensional weak \mathbb{Q}-Fano variety XX with terminal singularities satisfies

c1(X)nbnc2(X)c1(X)n2.c_{1}(X)^{n}\leq b_{n}c_{2}(X)\cdot c_{1}(X)^{n-2}. (1.0.1)

Such an inequality is called a Kawamata–Miyaoka type inequality. In the viewpoint of the explicit classification of weak \mathbb{Q}-Fano varieties with terminal singularities, it is natural to ask if we can find an effective or even the smallest constant bnb_{n} satisfying (1.0.1). For smooth Fano varieties with Picard number one, an effective constant bnb_{n} has already been found in [Liu19, Theorem 1.1]. Our main result in this paper is the following effective version of (1.0.1) for \mathbb{Q}-Fano varieties with canonical singularities and Picard number one.

Theorem 1.1.

Let XX be an nn-dimensional \mathbb{Q}-Fano variety with canonical singularities and ρ(X)=1\rho(X)=1 such that n2n\geq 2 and XX is smooth in codimension two. Then we have

c1(X)n<4c2(X)c1(X)n2.c_{1}(X)^{n}<4c_{2}(X)\cdot c_{1}(X)^{n-2}.

For the proof of Theorem 1.1, we will follow the same strategy as that of [Liu19], which can be traced back to Miyaoka’s pioneering work [Miy87] on the pseudo-effectivity of the second Chern class of generically nef sheaves. However, the main tool used in [Liu19] is the theory of minimal rational curves which only holds on smooth varieties. In this paper, we will use the theory of Fano foliations developed in the last decade [AD13, AD14, Dru17] to overcome the difficulty caused by singularities and our main contribution is an optimal upper bound for the slopes of rank one subsheaves of tangent sheaves of \mathbb{Q}-Fano varieties with canonical singularities and Picard number one, i.e., Proposition 3.8.

In dimension three, using Reid’s orbifold Riemann–Roch formula and the known classification of \mathbb{Q}-Fano threefolds with terminal singularities and Picard number one, we can refine our general arguments to get the following stronger inequality.

Theorem 1.2.

Let XX be a \mathbb{Q}-Fano threefold with terminal singularities and ρ(X)=1\rho(X)=1. Then we have

c1(X)3258c2(X)c1(X).c_{1}(X)^{3}\leq\frac{25}{8}c_{2}(X)\cdot c_{1}(X).

Our statement is actually a bit more precise: Except very few cases, we can replace the constant 25/825/8 by 33, or even by 121/41121/41 if the Fano index of XX is at least 44 (see § 2.2 for the definition of Fano index). We refer the reader to Theorem 4.4 and Remark 4.5 for the precise statement.

Acknowledgements

We thank the referees for their detailed reports, which help us to significantly improve the exposition of this paper. We would like to thank Wenhao Ou and Qizheng Yin for very helpful discussions and Yuri Prokhorov for useful communications. J. Liu is supported by the National Key Research and Development Program of China (No. 2021YFA1002300), the NSFC grants (No. 12001521 and No. 12288201), the CAS Project for Young Scientists in Basic Research (No. YSBR-033) and the Youth Innovation Promotion Association CAS.

2. Preliminaries

Varieties and manifolds will always be supposed to be irreducible. We will freely use standard terminologies and results of the minimal model program (MMP) as explained in [KM98].

2.1. Slope and stability of torsion free shaves

Let XX be an nn-dimensional normal projective variety and let F\scr{F} be a torsion free coherent sheaf with rank r1r\geq 1 on XX. The reflexive hull of F\scr{F} is defined to be its double dual F\scr{F}^{**}. We say that F\scr{F} is reflexive if F=F\scr{F}=\scr{F}^{**}. For any positive integer mm, we denote by [m]F\wedge^{[m]}\scr{F} the reflexive hull of mF\wedge^{m}\scr{F} and the determinant det(F)\det(\scr{F}) of F\scr{F} is defined as [r]F\wedge^{[r]}\scr{F}. Let f:YXf\colon Y\rightarrow X be a surjective morphism. We denote by f[]Ff^{[*]}\scr{F} the reflexive pull-back of F\scr{F}; in other words, f[]Ff^{[*]}\scr{F} is the reflexive hull of fFf^{*}\scr{F}. If XX is smooth in codimension kk, then the Chern classes ci(F)c_{i}(\scr{F}) are defined as the elements jci(F|Xreg)j_{*}c_{i}(\scr{F}|_{X_{\operatorname{reg}}}) sitting in the Chow groups Ai(X)A^{i}(X) for iki\leq k, where XregX_{\operatorname{reg}} is the smooth locus of XX and j:Ai(Xreg)Ai(X)j\colon A^{i}(X_{\operatorname{reg}})\hookrightarrow A^{i}(X) is the natural inclusion by taking closure [BS58, § 6]. In particular, if XX is smooth in codimension two, then c1(F)2c_{1}(\scr{F})^{2} and c2(F)c_{2}(\scr{F}) are well-defined in A2(X)A^{2}(X) and the discriminant Δ(F)\Delta(\scr{F}) of F\scr{F} is defined as

Δ(F)=2rc2(F)(r1)c1(F)2.\Delta(\scr{F})=2rc_{2}(\scr{F})-(r-1)c_{1}(\scr{F})^{2}.

Let D1,,Dn1D_{1},\dots,D_{n-1} be a collection of nef \mathbb{Q}-Cartier \mathbb{Q}-divisors on XX. Denote by α\alpha the one cycle D1Dn1D_{1}\cdots D_{n-1}. The slope μα(F)\mu_{\alpha}(\scr{F}) of F\scr{F} (with respect to α\alpha) is defined as

μα(F)c1(F)αr.\mu_{\alpha}(\scr{F})\coloneqq\frac{c_{1}(\scr{F})\cdot\alpha}{r}.

We note that c1(F)αc_{1}(\scr{F})\cdot\alpha is well-defined because the DiD_{i}’s are \mathbb{Q}-Cartier. One can then define (semi-)stability of F\scr{F} (with respect to α\alpha) as usual. Moreover, there exists the so-called Harder–Narasimhan filtration of F\scr{F}, i.e., a filtration as follows

0=F0F1Fm1Fm=F0=\scr{F}_{0}\subsetneq\scr{F}_{1}\subsetneq\cdots\subsetneq\scr{F}_{m-1}\subsetneq\scr{F}_{m}=\scr{F}

such that the graded pieces, GiFi/Fi1\scr{G}_{i}\coloneqq\scr{F}_{i}/\scr{F}_{i-1}, are semi-stable torsion free sheaves with strictly decreasing slopes [Miy87, Theorem 2.1]. We call F1\scr{F}_{1} the maximal destabilising subsheaf of F\scr{F}. We can also define

μαmax(F)μα(F1)andμαmin(F)μα(Gm).\mu_{\alpha}^{\max}(\scr{F})\coloneqq\mu_{\alpha}(\scr{F}_{1})\quad\textup{and}\quad\mu_{\alpha}^{\min}(\scr{F})\coloneqq\mu_{\alpha}(\scr{G}_{m}).

The following easy result is well-known to experts and we include a complete proof for the lack of explicit references.

Lemma 2.1.

Let f:YXf\colon Y\rightarrow X be a birational morphism between nn-dimensional normal projective varieties. Let D1,,Dn1D_{1},\dots,D_{n-1} be a collection of nef \mathbb{Q}-Cartier \mathbb{Q}-divisors on XX. Set αD1Dn1\alpha\coloneqq D_{1}\cdots D_{n-1} and αfα\alpha^{\prime}\coloneqq f^{*}\alpha. Let F\scr{F} and F\scr{F}^{\prime} be non-zero torsion free sheaves on XX and YY, respectively.

  1. (2.1.1)

    If fFf^{*}\scr{F} agrees with F\scr{F}^{\prime} away from the ff-exceptional locus, then we have

    μα(F)=μα(F).\mu_{\alpha}(\scr{F})=\mu_{\alpha^{\prime}}(\scr{F}^{\prime}).
  2. (2.1.2)

    If F=fF/tor\scr{F}^{\prime}=f^{*}\scr{F}/\operatorname{tor}, then we have

    μαmax(F)=μαmax(F)\mu_{\alpha}^{\max}(\scr{F})=\mu_{\alpha^{\prime}}^{\max}(\scr{F}^{\prime})  and  μαmin(F)=μαmin(F)\mu_{\alpha}^{\min}(\scr{F})=\mu_{\alpha^{\prime}}^{\min}(\scr{F}^{\prime}).

Proof.

The first statement follows from the equality fc1(F)=c1(F)f_{*}c_{1}(\scr{F}^{\prime})=c_{1}(\scr{F}) and the projection formula.

Next, we assume that F=fF/tor\scr{F}^{\prime}=f^{*}\scr{F}/\operatorname{tor} and let E\scr{E} be the maximal destabilising subsheaf of F\scr{F}. Then E=fE/tor\scr{E}^{\prime}=f^{*}\scr{E}/\operatorname{tor} is a subsheaf of F\scr{F}^{\prime}. So we have μαmax(F)μαmax(F)\mu_{\alpha^{\prime}}^{\max}(\scr{F}^{\prime})\geq\mu_{\alpha}^{\max}(\scr{F}) since μα(E)=μα(E)=μαmax(F)\mu_{\alpha^{\prime}}(\scr{E}^{\prime})=\mu_{\alpha}(\scr{E})=\mu_{\alpha}^{\max}(\scr{F}) by the first statement.

Let now M\scr{M}^{\prime} be the maximal destabilising subsheaf of F\scr{F}^{\prime}. Then M\scr{M}^{\prime} descends to a coherent subsheaf MU\scr{M}_{U} of F|U\scr{F}|_{U}, where UU is the maximal open subset of XX such that f1(U)Uf^{-1}(U)\rightarrow U is an isomorphism. Then there exists a torsion free coherent subsheaf M\scr{M} of F\scr{F} such that M|U=MU\scr{M}|_{U}=\scr{M}_{U}, which yields

μαmax(F)μα(M)=μα(M)=μαmax(F).\mu_{\alpha}^{\max}(\scr{F})\geq\mu_{\alpha}(\scr{M})=\mu_{\alpha^{\prime}}(\scr{M}^{\prime})=\mu_{\alpha^{\prime}}^{\max}(\scr{F}^{\prime}).

Hence we have μαmax(F)=μαmax(F)\mu_{\alpha}^{\max}(\scr{F})=\mu_{\alpha^{\prime}}^{\max}(\scr{F}^{\prime}). The proof of the last equality μαmin(F)=μαmin(F)\mu_{\alpha}^{\min}(\scr{F})=\mu_{\alpha^{\prime}}^{\min}(\scr{F}^{\prime}) is very similar, so we leave it to the reader. ∎

2.2. Fano indices

Let XX be a normal projective variety such that its canonical divisor KXK_{X} is \mathbb{Q}-Cartier. Then the Gorenstein index rXr_{X} of XX is defined as the smallest positive integer mm such that mKXmK_{X} is Cartier. For a \mathbb{Q}-Fano variety XX, we can define the Fano index of XX in two different ways:

ιXmax{qKXqA,ACl(X)},ι^Xsup{qKXqB,BCl(X)}.\begin{split}\iota_{X}&\coloneqq\max\{q\in\mathbb{Z}\mid-K_{X}\sim qA,\quad A\in\operatorname{Cl}(X)\},\\ \hat{\iota}_{X}&\coloneqq\sup\{q\in\mathbb{Q}\mid-K_{X}\sim_{\mathbb{Q}}qB,\quad B\in\operatorname{Cl}(X)\}.\end{split}

Let XX be a \mathbb{Q}-Fano variety with klt singularities. Then the Picard group Pic(X)\operatorname{Pic}(X) is a finitely generated torsion free \mathbb{Z}-module and the numerical equivalence coincides with the \mathbb{Q}-linear equivalence [IP99, Proposition 2.1.2]. In particular, we have

ρ(X)Cl(X)/=Cl(X)/,\mathbb{Z}^{\rho(X)}\cong\operatorname{Cl}(X)/_{\sim_{\mathbb{Q}}}=\operatorname{Cl}(X)/_{\equiv},

where “\sim_{\mathbb{Q}}” means \mathbb{Q}-linear equivalence and “\equiv” means numerical equivalence. This implies directly that the Fano indices ιX\iota_{X} and ι^X\hat{\iota}_{X} are well-defined positive integers such that ιX|ι^X\iota_{X}|\hat{\iota}_{X}.

2.3. Basic notions of foliations

The tangent sheaf TX\scr{T}_{X} of a normal projective variety XX is defined to be the dual sheaf (ΩX1)(\Omega^{1}_{X})^{*} and the Chern classes ci(X)c_{i}(X) are defined to be the Chern classes of TX\scr{T}_{X} whenever they are well-defined in the sense of the definition in § 2.1.

Definition 2.2.

A foliation on a normal projective variety XX is a non-zero coherent subsheaf FTX\scr{F}\subsetneq\scr{T}_{X} such that

  1. (2.2.1)

    F\scr{F} is saturated in TX\scr{T}_{X}, i.e. TX/F\scr{T}_{X}/\scr{F} is torsion free, and

  2. (2.2.2)

    F\scr{F} is closed under the Lie bracket.

The canonical divisor of F\scr{F} is any Weil divisor KFK_{\scr{F}} such that OX(KF)det(F)\scr{O}_{X}(-K_{\scr{F}})\cong\det(\scr{F}).

A foliation F\scr{F} on a normal projective variety XX is called algebraically integrable if the leaf of F\scr{F} through a general point of XX is an algebraic variety. For an algebraically integrable foliation F\scr{F} with rank r>0r>0 such that KFK_{\scr{F}} is \mathbb{Q}-Cartier, there exists a unique proper subvariety TT^{\prime} of the Chow variety of XX whose general point parametrises the closure of a general leaf of F\scr{F} (viewed as a reduced and irreducible cycle in XX). Let TT be the normalisation of TT^{\prime} and let UT×XU\rightarrow T^{\prime}\times X be the normalisation of the universal family, with induced morphisms:

U{U}X{X}T{T}ν\scriptstyle{\nu}π\scriptstyle{\pi} (2.2.1)

Then ν:UX\nu\colon U\rightarrow X is birational and, for a general point tTt\in T, the image ν(π1(t))X\nu(\pi^{-1}(t))\subsetneq X is the closure of a leaf of F\scr{F}. We shall call the diagram (2.2.1) the family of leaves of F\scr{F} [AD14, Lemma 3.9]. Thanks to [AD14, Lemma 3.7 and Remark 3.12], there exists a canonically defined effective \mathbb{Q}-divisor Δ\Delta on UU such that

KFU+ΔνKF,K_{\scr{F}_{U}}+\Delta\sim_{\mathbb{Q}}\nu^{*}K_{\scr{F}}, (2.2.2)

where FU\scr{F}_{U} is the algebraically integrable foliation on UU induced by π\pi. In particular, the divisor Δ\Delta is ν\nu-exceptional as νKFU=KF\nu_{*}K_{\scr{F}_{U}}=K_{\scr{F}}. For a general fibre FF of π\pi, the pair (F,ΔF)(F,\Delta_{F}) is called the general log leaf of F\scr{F}, where ΔF=Δ|F\Delta_{F}=\Delta|_{F} [AD14, Definition 3.11 and Remark 3.12]. Here the restriction Δ|F\Delta|_{F} is well-defined because UU is smooth at codimension one points of FF. The following proposition plays a key role in the proof of Proposition 3.8.

Proposition 2.3 ([Dru17, Proposition 4.6]).

Let F\scr{F} be an algebraically integrable foliation on a normal projective variety XX such that KFK_{\scr{F}} is \mathbb{Q}-Cartier. If KF-K_{\scr{F}} is nef and big, then the general log leaf (F,ΔF)(F,\Delta_{F}) is not klt.

2.4. Orbifold Riemann–Roch formula

Let XX be a \mathbb{Q}-factorial threefold with terminal singularities. Then XX has only isolated singularities and it follows from [Mor85, Theorems 12, 23 and 25] that each singular point of XX can be deformed to a (unique) collection of a finite number of terminal quotient singularities {Qi}iI\{Q_{i}\}_{i\in I}. Write the type of the orbifold point QiQ_{i} as 1ri(1,1,bi)\frac{1}{r_{i}}\left(1,-1,b_{i}\right) with gcd(bi,ri)=1\gcd(b_{i},r_{i})=1 and 0<biri/20<b_{i}\leq r_{i}/2. We can then associate to QiQ_{i} the pair (bi,ri)(b_{i},r_{i}) and the basket BXB_{X} of XX is defined to be the collection of all such pairs (permitting weights) appeared in the deformations of singular points of XX [Rei87, § 6]. Denote by X\mathcal{R}_{X} the collection of rir_{i} (permitting weights) appearing in BXB_{X}. For simplicity, we write X\mathcal{R}_{X} as a set of positive integers whose weights appear in superscripts, say for example,

X={2,2,3,5}={22,3,5}.\mathcal{R}_{X}=\{2,2,3,5\}=\{2^{2},3,5\}.

Note that lcm(X)\operatorname{lcm}(\mathcal{R}_{X}) coincides with the Gorenstein index rXr_{X} of XX by [Suz04, Lemma 1.2]. Let DD be a Weil divisor on XX. According to [Rei87, Theorem 10.2], we have

χ(D)=χ(𝒪X)+112D(DKX)(2DKX)+112c2(X)D+QcQ(D),\chi(D)=\chi(\mathcal{O}_{X})+\frac{1}{12}D\cdot(D-K_{X})\cdot(2D-K_{X})+\frac{1}{12}c_{2}(X)\cdot D+\sum_{Q}c_{Q}(D), (2.3.1)

where the last sum runs over BXB_{X}. Recall that the local Weil class group of XX at a singular point PP is generated by KXK_{X} [Kaw88, Corollary 5.2]; so DlKXD\sim lK_{X} around PP for some ll\in\mathbb{Z}. At an orbifold point QQ of type 1rQ(1,1,bQ)\frac{1}{r_{Q}}(1,-1,b_{Q}) associated to PP, the local index of DD at QQ is defined to be the unique integer 0iQ<rQ0\leq i_{Q}<r_{Q} such that lKXiQKXlK_{X}\sim i_{Q}K_{X} around QQ by passing to a deformation if necessary and then the number cQ(D)c_{Q}(D) is defined as

cQ(D)iQ(rQ21)12rQ+j=0iQ1jbQ¯(rQjbQ¯)2rQ,c_{Q}(D)\coloneqq-\frac{i_{Q}(r^{2}_{Q}-1)}{12r_{Q}}+\sum_{j=0}^{i_{Q}-1}\frac{\overline{jb_{Q}}(r_{Q}-\overline{jb_{Q}})}{2r_{Q}},

where the symbol ¯\overline{\bullet} means the smallest residue mod rQr_{Q} and j=010\sum_{j=0}^{-1}\coloneqq 0.

Let now XX be a \mathbb{Q}-Fano threefold with terminal singularities. In the case D=KXD=K_{X}, the formula (2.3.1) and Serre’s duality imply

1=124c2(X)c1(X)+124Q(rQ1rQ).1=\frac{1}{24}c_{2}(X)\cdot c_{1}(X)+\frac{1}{24}\sum_{Q}\left(r_{Q}-\frac{1}{r_{Q}}\right). (2.3.2)

In the case D=nKXD=-nK_{X} for some positive integer nn, the formula (2.3.1) implies

χ(nKX)\displaystyle\chi(-nK_{X}) =112n(n+1)(2n+1)c1(X)3+2n+1l(n+1),\displaystyle=\frac{1}{12}n(n+1)(2n+1)c_{1}(X)^{3}+2n+1-l(n+1), (2.3.3)

where l(n+1)=Qj=0njbQ¯(rQjbQ¯)2rQl(n+1)=\sum_{Q}\sum_{j=0}^{n}\frac{\overline{jb_{Q}}\left(r_{Q}-\overline{jb_{Q}}\right)}{2r_{Q}} and the first sum runs over BXB_{X}. When n=1n=1, together with the Kawamata–Viehweg vanishing theorem, we immediately get the following easy but useful lemma.

Lemma 2.4.

Let XX be a \mathbb{Q}-Fano threefold with terminal singularities. Then we have

h0(X,KX)=12c1(X)3+3l(2)0.\displaystyle h^{0}(X,-K_{X})=\frac{1}{2}c_{1}(X)^{3}+3-l(2)\in\mathbb{Z}_{\geq 0}. (2.4.1)

In the case KXιXA-K_{X}\sim\iota_{X}A, where AA is an ample Weil divisor, by [Suz04, Lemma 1.2], the numbers A3A^{3} and rXr_{X} satisfy

gcd(rX,ιX)=1andrXA3>0.\gcd(r_{X},\iota_{X})=1\quad\textup{and}\quad r_{X}A^{3}\in\mathbb{Z}_{>0}. (2.4.2)

If ιX3\iota_{X}\geq 3, combining (2.3.1) with the Kawamata–Viehweg vanishing yields

0=χ(tA)=1+t(ιX+t)(ιX+2t)12A3+t12ιXc2(X)c1(X)+QcQ(tA)0=\chi(tA)=1+\frac{t(\iota_{X}+t)(\iota_{X}+2t)}{12}A^{3}+\frac{t}{12\iota_{X}}c_{2}(X)\cdot c_{1}(X)+\sum_{Q}c_{Q}(tA) (2.4.3)

for ιX<t<0-\iota_{X}<t<0. In particular, when t=1t=-1, we obtain

A3=12(ιX1)(ιX2)(1c2(X)c1(X)12ιX+QcQ(A)).A^{3}=\frac{12}{(\iota_{X}-1)(\iota_{X}-2)}\left(1-\frac{c_{2}(X)\cdot c_{1}(X)}{12\iota_{X}}+\sum_{Q}c_{Q}(-A)\right). (2.4.4)

3. Kawamata–Miyaoka type inequality

3.1. Langer’s inequality and its variants

We discuss an inequality proved by Langer in [Lan04] and its several variants in this subsection.

Theorem 3.1 ([Lan04, Theorem 5.1]).

Let XX be an nn-dimensional normal projective variety such that n2n\geq 2 and XX is smooth in codimension two. Let D1,,Dn1D_{1},\dots,D_{n-1} be a collection of nef \mathbb{Q}-Cartier \mathbb{Q}-divisors on XX and set αD1Dn1\alpha\coloneqq D_{1}\cdots D_{n-1}. Then for any torsion free sheaf F\scr{F} with rank r>0r>0 on XX, we have

(D1α)(Δ(F)D2Dn1)+r2(μαmax(F)μα(F))(μα(F)μαmin(F))0.(D_{1}\cdot\alpha)(\Delta(\scr{F})\cdot D_{2}\cdots D_{n-1})+r^{2}(\mu_{\alpha}^{\max}(\scr{F})-\mu_{\alpha}(\scr{F}))(\mu_{\alpha}(\scr{F})-\mu^{\min}_{\alpha}(\scr{F}))\geq 0.
Proof.

If α\alpha is numerically trivial, then the left-hand side is equal to zero. Thus we may assume that α\alpha is not numerically trivial. Let f:XXf\colon X^{\prime}\rightarrow X be a resolution such that the induced morphism f1(Xreg)Xregf^{-1}(X_{\operatorname{reg}})\rightarrow X_{\operatorname{reg}} is an isomorphism. Set αfα\alpha^{\prime}\coloneqq f^{*}\alpha. Then α\alpha^{\prime} is also not numerically trivial. Applying [Lan04, Theorem 5.1] to FfF/tor\scr{F}^{\prime}\coloneqq f^{*}\scr{F}/\operatorname{tor} and α\alpha^{\prime} yields

(D1α)(Δ(F)D2Dn1)+r2(μαmax(F)μα(F))(μα(F)μαmin(F))0,(D^{\prime}_{1}\cdot\alpha^{\prime})(\Delta(\scr{F}^{\prime})\cdot D^{\prime}_{2}\cdots D^{\prime}_{n-1})+r^{2}(\mu^{\max}_{\alpha^{\prime}}(\scr{F}^{\prime})-\mu_{\alpha^{\prime}}(\scr{F}^{\prime}))(\mu_{\alpha^{\prime}}(\scr{F}^{\prime})-\mu_{\alpha^{\prime}}^{\min}(\scr{F}^{\prime}))\geq 0,

where DifDiD^{\prime}_{i}\coloneqq f^{*}D_{i}. Moreover, as codim(XXreg)3\operatorname{codim}(X\setminus X_{\operatorname{reg}})\geq 3, we have Δ(F)=fΔ(F)\Delta(\scr{F})=f_{*}\Delta(\scr{F}^{\prime}). Now the result follows from Lemma 2.1 and the projection formula. ∎

Recall that a Mehta–Ramanathan-general curve CC on an nn-dimensional normal projective variety is the complete intersection of n1n-1 sufficiently ample divisors in general positions.

Definition 3.2.

Let F\scr{F} be a torsion free sheaf on a normal projective variety XX. Then F\scr{F} is called generically nef (resp. generically ample) if F|C\scr{F}|_{C} is nef (resp. ample) for any Mehta–Ramanathan-general curve CC.

Remark 3.3.

By the modified Mumford–Mehta–Ramanathan theorem [Miy87, Corollary 3.13], a sheaf F\scr{F} is generically nef if and only if for any non-zero torsion free quotient Q\scr{Q} of F\scr{F}, we have c1(Q)D1Dn10c_{1}(\scr{Q})\cdot D_{1}\cdots D_{n-1}\geq 0 for any nef divisors DiD_{i}.

With the notions above, one can derive the following result.

Proposition 3.4.

Notation and assumptions are as in Theorem 3.1. Then for any generically nef torsion free sheaf F\scr{F} on XX with c1(F)D1c_{1}(\scr{F})\equiv D_{1}, we have

2c2(F)D2Dn1c1(F)2D2Dn1μαmax(F),2c_{2}(\scr{F})\cdot D_{2}\cdots D_{n-1}\geq c_{1}(\scr{F})^{2}\cdot D_{2}\cdots D_{n-1}-\mu_{\alpha}^{\max}(\scr{F}), (3.4.1)

and the equality holds only if either μαmin(F)=0\mu_{\alpha}^{\min}(\scr{F})=0 or F\scr{F} is semi-stable.

If additionally F\scr{F} is generically ample and the DiD_{i}’s are ample, then the inequality (3.4.1) is strict unless F\scr{F} is semi-stable and Δ(F)D2Dn1=0\Delta(\scr{F})\cdot D_{2}\cdots D_{n-1}=0.

Proof.

Since F\scr{F} is generically nef and the DiD_{i}’s are nef, we have μαmin(F)0\mu_{\alpha}^{\min}(\scr{F})\geq 0 by Remark 3.3. In particular, if D1α=0D_{1}\cdot\alpha=0, then F\scr{F} is actually semi-stable and

μαmax(F)=μα(F)=μαmin(F)=0.\mu_{\alpha}^{\max}(\scr{F})=\mu_{\alpha}(\scr{F})=\mu_{\alpha}^{\min}(\scr{F})=0.

Then the right-hand side of (3.4.1) is zero and thus the result follows from [Miy87, Theorem 6.1]. So we may assume that D1α>0D_{1}\cdot\alpha>0. Then Theorem 3.1 implies

(D1α)(Δ(F)D2Dn1)+r2(μαmax(F)μα(F))μα(F)0,(D_{1}\cdot\alpha)(\Delta(\scr{F})\cdot D_{2}\cdots D_{n-1})+r^{2}(\mu_{\alpha}^{\max}(\scr{F})-\mu_{\alpha}(\scr{F}))\mu_{\alpha}(\scr{F})\geq 0,

and the equality holds only if either μαmin(F)=0\mu_{\alpha}^{\min}(\scr{F})=0 or F\scr{F} is semi-stable. In particular, as D1α=rμα(F)>0D_{1}\cdot\alpha=r\mu_{\alpha}(\scr{F})>0, we obtain

Δ(F)D2Dn1+rμαmax(F)c1(F)2D2Dn10,\Delta(\scr{F})\cdot D_{2}\cdots D_{n-1}+r\mu_{\alpha}^{\max}(\scr{F})-c_{1}(\scr{F})^{2}\cdot D_{2}\cdots D_{n-1}\geq 0,

and the equality holds only if either μαmin(F)=0\mu_{\alpha}^{\min}(\scr{F})=0 or F\scr{F} is semi-stable. Then an easy computation yields the first statement. If F\scr{F} is generically ample and the DiD_{i}’s are ample, then μαmin(F)>0\mu_{\alpha}^{\min}(\scr{F})>0 and therefore the inequality (3.4.1) is strict unless F\scr{F} is semi-stable and Δ(F)D2Dn1=0\Delta(\scr{F})\cdot D_{2}\cdots D_{n-1}=0. ∎

Corollary 3.5.

Let XX be an nn-dimensional normal projective variety such that n2n\geq 2 and XX is smooth in codimension two. Let F\scr{F} be a generically nef torsion free sheaf on XX such that D1c1(F)D_{1}\coloneqq c_{1}(\scr{F}) is nef. Let D2,,Dn1D_{2},\dots,D_{n-1} be a collection of nef \mathbb{Q}-Cartier \mathbb{Q}-divisors on XX and set αD1Dn1\alpha\coloneqq D_{1}\cdots D_{n-1}. Then we have

r12rc1(F)2D2Dn1c2(F)D2Dn1,\frac{r^{\prime}-1}{2r^{\prime}}c_{1}(\scr{F})^{2}\cdot D_{2}\cdots D_{n-1}\leq c_{2}(\scr{F})\cdot D_{2}\cdots D_{n-1}, (3.5.1)

where rr^{\prime} is the rank of the maximal destabilising subsheaf of F\scr{F} with respect to α\alpha.

If additionally F\scr{F} is generically ample and the DiD_{i}’s are ample, then the inequality (3.5.1) is strict unless F\scr{F} is semi-stable and Δ(F)D2Dn1=0\Delta(\scr{F})\cdot D_{2}\cdots D_{n-1}=0.

Proof.

Let E\scr{E} be the maximal destabilising subsheaf of F\scr{F}. As F\scr{F} is generically nef, we have

μαmax(F)=μα(E)c1(F)αr.\mu_{\alpha}^{\max}(\scr{F})=\mu_{\alpha}(\scr{E})\leq\frac{c_{1}(\scr{F})\cdot\alpha}{r^{\prime}}.

Then the statements follow immediately from Proposition 3.4. ∎

3.2. Proof of Theorem 1.1

We start with the following result, which is essentially proved in [Ou23]. See also [Pet12, Theorem 1.3] for the smooth case.

Proposition 3.6.

Let XX be a weak \mathbb{Q}-Fano variety with klt singularities. Then TX\scr{T}_{X} is generically ample.

Proof.

We assume to the contrary that TX\scr{T}_{X} is not generically ample. Since TX\scr{T}_{X} is generically nef by [Ou23, Theorem 1.3], there exists a non-trivial torsion free quotient TXQ\scr{T}_{X}\rightarrow\scr{Q} such that det(Q)0\det(\scr{Q})\equiv 0 as in the proof of [Ou23, Theorem 1.7]. As XX is a weak \mathbb{Q}-Fano variety with klt singularities, there exists a positive integer mm such that det(Q)[m]OX\det(\scr{Q})^{[\otimes m]}\cong\scr{O}_{X} (see the proof of [IP99, Proposition 2.1.2]). In particular, there exists a finite quasi-étale cover f:X~Xf\colon\widetilde{X}\rightarrow X such that f[]det(Q)OX~f^{[*]}\det(\scr{Q})\cong\scr{O}_{\tilde{X}}. On the other hand, as Q\scr{Q} is a quotient of TX\scr{T}_{X}, it induces an injection

det(Q)ΩX[p]\det(\scr{Q}^{*})\rightarrow\Omega^{[p]}_{X}

of coherent sheaves, where 1prank(Q)dimX1\leq p\coloneqq\operatorname{rank}(\scr{Q})\leq\dim X. Then taking reflexive pull-back yields an injection

OX~f[]det(Q)ΩX~[p]=f[]ΩX[p].\scr{O}_{\widetilde{X}}\cong f^{[*]}\det(\scr{Q}^{*})\rightarrow\Omega_{\widetilde{X}}^{[p]}=f^{[*]}\Omega^{[p]}_{X}.

This means H0(X~,ΩX~[p])0H^{0}(\widetilde{X},\Omega^{[p]}_{\widetilde{X}})\not=0. However, since ff is a finite quasi-étale morphism, it follows that KX~=fKX-K_{\widetilde{X}}=-f^{*}K_{X} is a nef and big \mathbb{Q}-Cartier divisor and X~\widetilde{X} has only klt singularities by [KM98, Proposition 5.20]. Thus X~\widetilde{X} is rationally connected by [Zha06] and consequently H0(X~,ΩX~[p])=0H^{0}(\widetilde{X},\Omega^{[p]}_{\widetilde{X}})=0 for any 1pdimX~1\leq p\leq\dim\widetilde{X} by [GKKP11, Theorem 5.1], which is a contradiction. ∎

The following result is a slight generalisation of [IJL23, Corollary 7.5].

Proposition 3.7.

Let XX be an nn-dimensional \mathbb{Q}-factorial variety with klt singularities such that n2n\geq 2, XX is smooth in codimension two and D1KXD_{1}\coloneqq-K_{X} is nef. Let D2,,Dn1D_{2},\dots,D_{n-1} be a collection of nef \mathbb{Q}-Cartier \mathbb{Q}-divisors and set αD1Dn1\alpha\coloneqq D_{1}\cdots D_{n-1}. Then one of the following statements holds.

  1. (3.7.1)

    c1(X)2D2Dn14c2(X)D2Dn1c_{1}(X)^{2}\cdot D_{2}\cdots D_{n-1}\leq 4c_{2}(X)\cdot D_{2}\cdots D_{n-1} and the inequality is strict if the DiD_{i}’s are ample.

  2. (3.7.2)

    There exists a rational map f:XTf\colon X\dashrightarrow T, whose general fibres are rational curves, such that the relative tangent sheaf Tf\scr{T}_{f} is the maximal destabilising subsheaf of TX\scr{T}_{X} with respect to α\alpha.

Proof.

By [Ou23, Theorem 1.3], the tangent sheaf TX\scr{T}_{X} is generically nef. Let E\scr{E} be the maximal destabilising subsheaf of F\scr{F} with respect to α\alpha. Denote by rr^{\prime} the rank of E\scr{E}. Suppose first r2r^{\prime}\geq 2. Then by Corollary 3.5, we have

c1(X)2D2Dn12rr1c2(X)D2Dn14c2(X)D2Dn1.c_{1}(X)^{2}\cdot D_{2}\cdots D_{n-1}\leq\frac{2r^{\prime}}{r^{\prime}-1}c_{2}(X)\cdot D_{2}\cdots D_{n-1}\leq 4c_{2}(X)\cdot D_{2}\cdots D_{n-1}. (3.7.1)

Assume in addition that the DiD_{i}’s are ample. Then we have

c1(X)2D2Dn1=D12D2Dn1>0.c_{1}(X)^{2}\cdot D_{2}\cdots D_{n-1}=D_{1}^{2}\cdot D_{2}\cdots D_{n-1}>0.

In particular, the second inequality in (3.7.1) is strict unless r=2r^{\prime}=2. On the other hand, since D1KXD_{1}\coloneqq-K_{X} is ample, the sheaf TX\scr{T}_{X} is generically ample by Proposition 3.6. Then it follows from Corollary 3.5 that the first inequality in (3.7.1) is strict unless TX\scr{T}_{X} is semi-stable and Δ(TX)D2Dn1=0\Delta(\scr{T}_{X})\cdot D_{2}\cdots D_{n-1}=0. So we have

c1(X)2D2Dn1<4c2(X)D2Dn1c_{1}(X)^{2}\cdot D_{2}\cdots D_{n-1}<4c_{2}(X)\cdot D_{2}\cdots D_{n-1}

unless n=2n=2, TX\scr{T}_{X} is semi-stable and Δ(TX)=0\Delta(\scr{T}_{X})=0. In the latter case, it follows from [GKP21, Theorem 1.2] that XX is a quasi-Abelian surface, i.e., a finite quasi-étale quotient of an Abelian surface, which is absurd.

Suppose next r=1r^{\prime}=1. Then E\scr{E} is closed under the Lie bracket and thus it defines a rank one foliation on XX. Moreover, we have

nc1(E)α>c1(X)α0.nc_{1}(\scr{E})\cdot\alpha>c_{1}(X)\cdot\alpha\geq 0.

So KEK_{\scr{E}} is not pseudo-effective and [CP19, Theorem 1.1] says that E\scr{E} is algebraically integrable such that its general leaves are rational curves. Then we define f:XTf\colon X\dashrightarrow T to be the rational map induced by the universal family of leaves of E\scr{E}. ∎

Given two \mathbb{Q}-Cartier \mathbb{Q}-divisor classes δ\delta and δ\delta^{\prime} on a projective variety XX, we will denote by δδ\delta\leq\delta^{\prime} (resp. δ<δ\delta<\delta^{\prime}) if δδ\delta^{\prime}-\delta is effective (resp. effective and non-zero). The following result is the key ingredient of the proof of Theorem 1.1.

Proposition 3.8.

Let XX be a \mathbb{Q}-Fano variety with canonical singularities and ρ(X)=1\rho(X)=1 such that dim(X)2\dim(X)\geq 2. For any rank one subsheaf L\scr{L} of TX\scr{T}_{X}, we have

2c1(L)c1(X).2c_{1}(\scr{L})\leq c_{1}(X). (3.8.1)

If additionally XX has only terminal singularities, then we have

2rX+1rXc1(L)c1(X),\frac{2r_{X}+1}{r_{X}}c_{1}(\scr{L})\leq c_{1}(X), (3.8.2)

where rXr_{X} is the Gorenstein index of XX.

Proof.

Let L^\widehat{\scr{L}} be the saturation of L\scr{L} in TX\scr{T}_{X}. Then c1(L)c1(L^)c_{1}(\scr{L})\leq c_{1}(\widehat{\scr{L}}) and thus we may assume that L\scr{L} itself is saturated in TX\scr{T}_{X} such that c1(L)>0c_{1}(\scr{L})>0. Then L\scr{L} defines an algebraically integral rank one foliation with general leaves being rational curves by [CP19, Theorem 1.1]. Let π:UT\pi\colon U\rightarrow T be the family of leaves of L\scr{L} as in diagram (2.2.1)

U{U}X{X}T{T}ν\scriptstyle{\nu}π\scriptstyle{\pi}

and let Δ\Delta be the effective Weil \mathbb{Q}-divisor on UU as in (2.2.2) such that

KF+ΔνKL,K_{\scr{F}}+\Delta\sim_{\mathbb{Q}}\nu^{*}K_{\scr{L}},

where F\scr{F} is the foliation on UU induced by π\pi. Then the general fibre FF of π\pi is isomorphic to 1\mathbb{P}^{1}.

Since UU is smooth in a neighborhood of FF, the irreducible components of Δ\Delta are Cartier in a neighborhood of FF. Since KL=c1(L)-K_{\scr{L}}=c_{1}(\scr{L}) is ample, the general log leaf (F,ΔF)(F,\Delta_{F}) is not klt by Proposition 2.3, where ΔF=Δ|F\Delta_{F}=\Delta|_{F}. So deg(ΔF)1\deg(\Delta_{F})\geq 1 and we obtain

νKLF=(ΔKF)F=ΔF+2=deg(ΔF)+21.-\nu^{*}K_{\scr{L}}\cdot F=(-\Delta-K_{\scr{F}})\cdot F=-\Delta\cdot F+2=-\deg(\Delta_{F})+2\leq 1.

Denote by α\alpha the curve class of ν(F)\nu(F). Then by projection formula, we get

c1(L)α=KLα=νKLF1.c_{1}(\scr{L})\cdot\alpha=-K_{\scr{L}}\cdot\alpha=-\nu^{*}K_{\scr{L}}\cdot F\leq 1.

On the other hand, since XX has only canonical singularities, there exists an effective \mathbb{Q}-divisor ΔU\Delta_{U} on UU such that KU=νKX+ΔUK_{U}=\nu^{*}K_{X}+\Delta_{U}. Then applying the projection formula again yields

KXα=νKXF=(KU+ΔU)FKUF=2.-K_{X}\cdot\alpha=-\nu^{*}K_{X}\cdot F=(-K_{U}+\Delta_{U})\cdot F\geq-K_{U}\cdot F=2.

Hence 2c1(L)α2c1(X)α2c_{1}(\scr{L})\cdot\alpha\leq 2\leq c_{1}(X)\cdot\alpha and we are done as ρ(X)=1\rho(X)=1.

Finally we assume in addition that XX has only terminal singularities. Then we have Supp(ΔU)=Exc(ν)\operatorname{Supp}(\Delta_{U})=\operatorname{Exc}(\nu) since XX is \mathbb{Q}-factorial. Moreover, since rXKXr_{X}K_{X} is Cartier, the effective divisor rXΔUr_{X}\Delta_{U} is integral and as ρ(X)=1\rho(X)=1, there exists at least one irreducible component of Exc(ν)\operatorname{Exc}(\nu) which dominates TT. In particular, since rXΔUr_{X}\Delta_{U} is Cartier in a neighbourhood of FF, we have rXΔUF1r_{X}\Delta_{U}\cdot F\geq 1. Then the same argument as before applies to show

c1(L)α1andc1(X)α=KUF+ΔUF2+1rX.c_{1}(\scr{L})\cdot\alpha\leq 1\quad\textup{and}\quad c_{1}(X)\cdot\alpha=-K_{U}\cdot F+\Delta_{U}\cdot F\geq 2+\frac{1}{r_{X}}.

This finishes the proof as ρ(X)=1\rho(X)=1. ∎

The inequalities (3.8.1) and (3.8.2) in Proposition 3.8 above are both sharp as shown by the following example of the so-called normal generalised cones.

Example 3.9 (Normal generalised cones).

Let XX be a Fano manifold with ρ(X)=1\rho(X)=1 and let OX(1)\scr{O}_{X}(1) be the ample generator of Pic(X)\operatorname{Pic}(X). Denote by ii the Fano index of XX, i.e., OX(KX)OX(i)\scr{O}_{X}(-K_{X})\cong\scr{O}_{X}(i). Given a positive integer mm, we denote by Em\scr{E}_{m} the rank two vector bundle OX(m)OX\scr{O}_{X}(m)\oplus\scr{O}_{X}. Let YmY_{m} be the projectivised bundle (Em)\mathbb{P}(\scr{E}_{m}) with the natural projection πm:YmX\pi_{m}\colon Y_{m}\rightarrow X. Since the tautological line bundle OYm(1)O(Em)(1)\scr{O}_{Y_{m}}(1)\coloneqq\scr{O}_{\mathbb{P}(\scr{E}_{m})}(1) is big and semi-ample, it defines a birational contraction fm:YmZmf_{m}\colon Y_{m}\rightarrow Z_{m} to a normal projective variety ZmZ_{m}. The exceptional divisor EmE_{m} of fmf_{m} corresponds to the quotient EmOX\scr{E}_{m}\rightarrow\scr{O}_{X} which is contracted to a point under fmf_{m} such that

Em=c1(OYm(1))mπmc1(OX(1)).E_{m}=c_{1}(\scr{O}_{Y_{m}}(1))-m\pi_{m}^{*}c_{1}(\scr{O}_{X}(1)).

Note that Cl(Zm)\operatorname{Cl}(Z_{m}) is generated by the class Amfmπc1(OX(1))A_{m}\coloneqq f_{m*}\pi^{*}c_{1}(\scr{O}_{X}(1)), whose Gorenstein index is mm. So the variety ZmZ_{m} is \mathbb{Q}-factorial and ρ(Zm)=1\rho(Z_{m})=1. Moreover, an easy computation shows that ZmZ_{m} has canonical singularities if and only if 0<mi0<m\leq i and it has terminal singularities if and only if 0<m<i0<m<i. Let LmTZm\scr{L}_{m}\subset\scr{T}_{Z_{m}} be the rank one foliation induced by YmXY_{m}\rightarrow X; that is, the foliation defined by the induced rational map ZmXZ_{m}\dashrightarrow X. Then we have

c1(Lm)=mAmc_{1}(\scr{L}_{m})=mA_{m} and c1(Zm)=(m+i)Amc_{1}(Z_{m})=(m+i)A_{m}.

As a consequence, if m=im=i, then 2c1(Lm)=c1(Zm)2c_{1}(\scr{L}_{m})=c_{1}(Z_{m}); so the inequality (3.8.1) is optimal. On the other hand, if m=i1m=i-1, then the Gorenstein index rZmr_{Z_{m}} of ZmZ_{m} is equal to mm as gcd(m,i)=1\gcd(m,i)=1 and we have

c1(Lm)=mAm=m2m+1(2m+1)Am=m2m+1c1(Zm)=rZm2rZm+1c1(Zm).c_{1}(\scr{L}_{m})=mA_{m}=\frac{m}{2m+1}(2m+1)A_{m}=\frac{m}{2m+1}c_{1}(Z_{m})=\frac{r_{Z_{m}}}{2r_{Z_{m}}+1}c_{1}(Z_{m}).

So the inequality (3.8.2) is also optimal.

Corollary 3.10.

Let XX be a del Pezzo surface with du Val singularities and ρ(X)=1\rho(X)=1. Then TX\scr{T}_{X} is semi-stable.

Proof.

Since du Val singularities are quotient singularities, the surface XX is \mathbb{Q}-factorial and thus it is a \mathbb{Q}-Fano surface with canonical singularities. Then it follows from Proposition 3.8 that TX\scr{T}_{X} is semi-stable. ∎

Now we are ready to finish the proof of Theorem 1.1.

Proof of Theorem 1.1.

Set αc1(X)n1\alpha\coloneqq c_{1}(X)^{n-1}. By Proposition 3.7, we may assume that the maximal destabilising subsheaf E\scr{E} of TX\scr{T}_{X} has rank one. Then Proposition 3.8 implies that μα(E)c1(X)n/2\mu_{\alpha}(\scr{E})\leq c_{1}(X)^{n}/2. In particular, applying Proposition 3.4 to TX\scr{T}_{X} and c1(X)c_{1}(X) yields

2c2(X)c1(X)n2c1(X)nμα(E)c1(X)n22c_{2}(X)\cdot c_{1}(X)^{n-2}\geq c_{1}(X)^{n}-\mu_{\alpha}(\scr{E})\geq\frac{c_{1}(X)^{n}}{2}

and the equality holds only if TX\scr{T}_{X} is semi-stable, which is impossible as n2n\geq 2 and E\scr{E} has rank one. ∎

4. \mathbb{Q}-Fano threefolds with terminal singularities

Besides Reid’s orbifold Riemann–Roch formula, Kawamata’s result [Kaw92, Proposition 1] is another key result used in the studying of \mathbb{Q}-Fano threefolds with terminal singularities and ρ(X)=1\rho(X)=1 ([Suz04, Pro10, Pro22, BK22]). Though the inequality in Theorem 1.1 is already stronger than Kawamata’s one in dimension three, we can still ask if it can be improved or even be made optimal in the viewpoint of the explicit classification of \mathbb{Q}-Fano threefolds.

If XX is a smooth Fano threefold with ρ(X)=1\rho(X)=1, then TX\scr{T}_{X} is known to be stable ([PW95, Proposition 2.2 and Theorem 2.3]). In particular, the Bogomolov–Gieseker inequality implies

c1(X)33c2(X)c1(X).c_{1}(X)^{3}\leq 3c_{2}(X)\cdot c_{1}(X).

Thus one may ask if the tangent sheaves of \mathbb{Q}-Fano threefolds with terminal singularities and Picard number one are still (semi-)stable. Unfortunately, the answer to this question is negative as shown by the following easy example.

Example 4.1.

Let X(1,2,3,5)X\simeq\mathbb{P}(1,2,3,5) be one of the examples in [Pro10, Theorem 1.4]. The projection of XX onto the first two coordinates gives a rank two foliation FTX\scr{F}\subsetneq\scr{T}_{X}. It is easy to see that c1(F)=OX(8)c_{1}(\scr{F})=\scr{O}_{X}(8) and c1(TX)=OX(11)c_{1}(\scr{T}_{X})=\scr{O}_{X}(11) and hence

μ(F)=81122>1133=μ(TX),\mu(\scr{F})=\frac{8\cdot 11^{2}}{2}>\frac{11^{3}}{3}=\mu(\scr{T}_{X}),

which means that TX\scr{T}_{X} is not semi-stable.

Thus we cannot expect to improve Theorem 1.1 in dimension three by using the semi-stability of tangent sheaves. However, combining the known explicit classification in special cases with a refined argument of the proof of Theorem 1.1 will give us such an improvement. Indeed, if XX is a \mathbb{Q}-Fano threefold with terminal singularities and ρ(X)=1\rho(X)=1 such that ι^X9\hat{\iota}_{X}\geq 9, then by [Pro10, Proposition 3.6] and (2.3.2), an easy computation shows that either

c1(X)312141c2(X)c1(X),c_{1}(X)^{3}\leq\frac{121}{41}c_{2}(X)\cdot c_{1}(X),

or

ι^X=10andc1(X)3>4c2(X)c1(X).\hat{\iota}_{X}=10\quad\textup{and}\quad c_{1}(X)^{3}>4c_{2}(X)\cdot c_{1}(X).

Moreover, by [Pro10, Proposition 3.6 and Theorem 1.4 (iv)], the equality in the former case is attained by X(1,2,3,5)X\simeq\mathbb{P}(1,2,3,5); while the latter case contradicts Theorem 1.1 (cf. [Pro10, Section 5]). Thus we may assume that ι^X8\hat{\iota}_{X}\leq 8 in the sequel.

Theorem 4.2.

Let XX be a \mathbb{Q}-Fano threefold with terminal singularities and ρ(X)=1\rho(X)=1 such that ι^X8\hat{\iota}_{X}\leq 8 and TX\scr{T}_{X} is not semi-stable. Then ι^X4\hat{\iota}_{X}\geq 4 and the length of the Harder–Narasimhan filtration of TX\scr{T}_{X} is two.

Denote by F1\scr{F}_{1} the maximal destabilising subsheaf of TX\scr{T}_{X} with rank r1r_{1}. Let AA be an ample Weil divisor generating Cl(X)/\operatorname{Cl}(X)/_{\sim_{\mathbb{Q}}} and let ι^1\hat{\iota}_{1} be the positive integer such that c1(F1)ι^1Ac_{1}(\scr{F}_{1})\sim_{\mathbb{Q}}\hat{\iota}_{1}A. Then the possibilities for the pair (ι^1,r1)(\hat{\iota}_{1},r_{1}) are listed in the following table.

Table 1. Maximal destabilising subsheaves
ι^X\hat{\iota}_{X} 4 5 6 7 8
(ι^1,r1)(\hat{\iota}_{1},r_{1}) (3,2)(3,2) (2,1)(2,1) (4,2)(4,2) (5,2)(5,2) (3,1)(3,1) (5,2)(5,2) (6,2)(6,2) (3,1)(3,1) (6,2)(6,2) (7,2)(7,2)
Proof.

Let 0=F0F1Fl=TX0=\scr{F}_{0}\subsetneq\scr{F}_{1}\subsetneq\cdots\subsetneq\scr{F}_{l}=\scr{T}_{X} be the Harder–Narasimhan filtration of TX\scr{T}_{X} and let GiFi/Fi1\scr{G}_{i}\coloneqq\scr{F}_{i}/\scr{F}_{i-1} be the graded pieces. Denote by ι^i\hat{\iota}_{i} the integer such that c1(Gi)ι^iAc_{1}(\scr{G}_{i})\sim_{\mathbb{Q}}\hat{\iota}_{i}A and by rir_{i} the rank of Gi\scr{G}_{i}. Then clearly we have

ι^1r1>ι^X3,i=1lι^i=ι^Xandi=1lri=3.\frac{\hat{\iota}_{1}}{r_{1}}>\frac{\hat{\iota}_{X}}{3},\quad\sum_{i=1}^{l}\hat{\iota}_{i}=\hat{\iota}_{X}\quad\textup{and}\quad\sum_{i=1}^{l}r_{i}=3.

By Proposition 3.6, we have ι^i1\hat{\iota}_{i}\geq 1. Moreover, the sequence {ι^i/ri}\{\hat{\iota}_{i}/r_{i}\} is strictly decreasing and if r1=1r_{1}=1, then we have 2ι^1<ι^X2\hat{\iota}_{1}<\hat{\iota}_{X} by Proposition 3.8. Then a straightforward computation shows that ι^X4\hat{\iota}_{X}\geq 4 and if l=3l=3, then we have

r1=r2=r3=1andι^X=ι^1+ι^2+ι^33ι^133ι^X23.r_{1}=r_{2}=r_{3}=1\quad\textup{and}\quad\hat{\iota}_{X}=\hat{\iota}_{1}+\hat{\iota}_{2}+\hat{\iota}_{3}\leq 3\hat{\iota}_{1}-3\leq 3\lfloor\frac{\hat{\iota}_{X}}{2}\rfloor-3.

As ι^X8\hat{\iota}_{X}\leq 8, the inequality above implies that if l=3l=3, then (ι^X,ι^1)=(6,3)(\hat{\iota}_{X},\hat{\iota}_{1})=(6,3) or (8,4)(8,4), which are both impossible as 2ι^1<ι^X2\hat{\iota}_{1}<\hat{\iota}_{X}. Hence, we must have l=2l=2.

Finally, the possibilities of (ι^1,r1)(\hat{\iota}_{1},r_{1}) in each case are derived from an easy computation using the facts that 3ι^1>ι^Xr13\hat{\iota}_{1}>\hat{\iota}_{X}r_{1} and if r1=1r_{1}=1, then 2ι^1<ι^X2\hat{\iota}_{1}<\hat{\iota}_{X}. ∎

Some similar results have also been independently obtained by Sukuzi in [Suz24]. Combining Theorem 4.2 with Theorem 3.1, we get the following effective bounds.

Corollary 4.3.

Let XX be a \mathbb{Q}-Fano threefold with terminal singularities and ρ(X)=1\rho(X)=1 such that ι^X8\hat{\iota}_{X}\leq 8 and TX\scr{T}_{X} is not semi-stable. Denote by r1r_{1} the rank of the maximal destabilizing subsheaf F1\scr{F}_{1} of TX\scr{T}_{X}. Then we have

c1(X)3bc2(X)c1(X),c_{1}(X)^{3}\leq bc_{2}(X)\cdot c_{1}(X),

where bb can be chosen in each case as in the following table:

Table 2. Effective bound b
ι^X\hat{\iota}_{X} 4 5 6 7 8
r1=1r_{1}=1 // 10033\frac{100}{33} // 4916\frac{49}{16} 25685\frac{256}{85}
r1=2r_{1}=2 6421\frac{64}{21} 258\frac{25}{8} 165\frac{16}{5} 4915\frac{49}{15} 25677\frac{256}{77}

where the symbol “//” means that the case does not happen.

Proof.

Set αc1(X)2\alpha\coloneqq c_{1}(X)^{2}. Let AA be an ample Weil divisor generating Cl(X)/\operatorname{Cl}(X)/_{\sim_{\mathbb{Q}}}. Denote by ι^1\hat{\iota}_{1} the positive integer such that c1(F1)ι^1Ac_{1}(\scr{F}_{1})\sim_{\mathbb{Q}}\hat{\iota}_{1}A. Then we have

μαmax(TX)=μα(F1)=ι^1r1Ac1(X)2=ι^1ι^Xr1c1(X)3.\mu_{\alpha}^{\max}(\scr{T}_{X})=\mu_{\alpha}(\scr{F}_{1})=\frac{\hat{\iota}_{1}}{r_{1}}A\cdot c_{1}(X)^{2}=\frac{\hat{\iota}_{1}}{\hat{\iota}_{X}r_{1}}c_{1}(X)^{3}.

On the other hand, by Theorem 4.2, we also have

μαmin(TX)=μα(TX/F1)=ι^Xι^13r1Ac1(X)2=ι^Xι^1(3r1)ι^Xc1(X)3.\mu_{\alpha}^{\min}(\scr{T}_{X})=\mu_{\alpha}(\scr{T}_{X}/\scr{F}_{1})=\frac{\hat{\iota}_{X}-\hat{\iota}_{1}}{3-r_{1}}A\cdot c_{1}(X)^{2}=\frac{\hat{\iota}_{X}-\hat{\iota}_{1}}{(3-r_{1})\hat{\iota}_{X}}c_{1}(X)^{3}.

Then applying Theorem 3.1 to F=TX\scr{F}=\scr{T}_{X} yields

6c2(X)c1(X)\displaystyle 6c_{2}(X)\cdot c_{1}(X) 2c1(X)3(3ι^1ι^Xr11)(13(ι^Xι^1)(3r1)ι^X)c1(X)3\displaystyle\geq 2c_{1}(X)^{3}-\left(\frac{3\hat{\iota}_{1}}{\hat{\iota}_{X}r_{1}}-1\right)\left(1-\frac{3(\hat{\iota}_{X}-\hat{\iota}_{1})}{(3-r_{1})\hat{\iota}_{X}}\right)c_{1}(X)^{3}
2c1(X)3(3ι^1ι^Xr1)2r1(3r1)ι^X2c1(X)3.\displaystyle\geq 2c_{1}(X)^{3}-\frac{(3\hat{\iota}_{1}-\hat{\iota}_{X}r_{1})^{2}}{r_{1}(3-r_{1})\hat{\iota}_{X}^{2}}c_{1}(X)^{3}.

The result then follows from Table LABEL:tab1 in Theorem 4.2 by an easy computation. ∎

Now we are ready to prove the main result in this section and Theorem 1.2 is a direct consequence of it. Its proof is a combination of Corollary 4.3 and Reid’s orbifold Riemann–Roch formula.

Theorem 4.4.

Let XX be a \mathbb{Q}-Fano threefold with terminal singularities and ρ(X)=1\rho(X)=1. Then we have

c1(X)3258c2(X)c1(X),c_{1}(X)^{3}\leq\frac{25}{8}c_{2}(X)\cdot c_{1}(X), (4.4.1)

where the equality holds only if ι^X=ιX=5\hat{\iota}_{X}=\iota_{X}=5 and X={3,72}\mathcal{R}_{X}=\{3,7^{2}\}. If we assume in addition that ι^X4\hat{\iota}_{X}\not=4 and 55, then we have

c1(X)3<3c2(X)c1(X).c_{1}(X)^{3}<3c_{2}(X)\cdot c_{1}(X). (4.4.2)
Proof.

If TX\scr{T}_{X} is semi-stable, then by the Bogomolov–Gieseker inequality and [GKP21, Theorem 1.2], we have

c1(X)3<3c2(X)c1(X).c_{1}(X)^{3}<3c_{2}(X)\cdot c_{1}(X).

Thus we may assume that TX\scr{T}_{X} is not semi-stable and 5ι^X85\leq\hat{\iota}_{X}\leq 8 by Corollary 4.3. In particular, it follows from [Pro22, Proposition 3.3] that ι^X=ιX\hat{\iota}_{X}=\iota_{X}. Let AA be an ample Weil divisor such that KXι^XA-K_{X}\sim\hat{\iota}_{X}A. Then we use the orbifold Riemann–Roch formula (§ 2.4) and a computer program written in Python, whose algorithm is similar to that in [Pro10, Lemma 3.5] and is sketched as follows, to find out the possible baskets and numerical invariants of XX.

Step 1. As c2(X)c1(X)>0c_{2}(X)\cdot c_{1}(X)>0 by [Kaw92, Proposition 1], we can list huge but finitely many possibilities of X\mathcal{R}_{X} and c2(X)c1(X)c_{2}(X)\cdot c_{1}(X) satisfying (2.3.2).

Step 2. For each ιX=ι^X5\iota_{X}=\hat{\iota}_{X}\geq 5, we calculate the number A3A^{3} by (2.4.4) and pick up those X\mathcal{R}_{X} satisfying both (2.4.2) and (2.4.3).

Step 3. Recall from the beginning of this section that if ι^X9\hat{\iota}_{X}\geq 9, then the following (sharp) inequality always holds

c1(X)312141c2(X)c1(X).c_{1}(X)^{3}\leq\frac{121}{41}c_{2}(X)\cdot c_{1}(X). (4.4.3)

Thus, for each 5ι^X85\leq\hat{\iota}_{X}\leq 8, we take bι^Xb_{\hat{\iota}_{X}} as the maximum in the corresponding column in Table LABEL:tab2. Then we search for all candidates X\mathcal{R}_{X} satisfying the following

12141c2(X)c1(X)<c1(X)3bι^Xc2(X)c1(X)\frac{121}{41}c_{2}(X)\cdot c_{1}(X)<c_{1}(X)^{3}\leq b_{\hat{\iota}_{X}}c_{2}(X)\cdot c_{1}(X) (4.4.4)

to see if the inequality (4.4.3) also holds in these cases.

Finally we obtain only three possibilities for the numerical type of XX satisfying the inequality (4.4.4) for 5ι^X85\leq\hat{\iota}_{X}\leq 8, which are listed in the following table:

ι^X\hat{\iota}_{X} X\mathcal{R}_{X} c2(X)c1(X)c_{2}(X)\cdot c_{1}(X) c1(X)3c_{1}(X)^{3}
55 {4,7}\{4,7\} 37528\frac{375}{28} 112528\frac{1125}{28}
55 {3,72}\{3,7^{2}\} 16021\frac{160}{21} 50021\frac{500}{21}
77 {22,8}\{2^{2},8\} 1058\frac{105}{8} 3438\frac{343}{8}

The first case and the third case are ruled out by [Pro13, 7.5] and [Pro13, Claim 6.5], respectively. An easy computation shows 8c1(X)3=25c2(X)c1(X)8c_{1}(X)^{3}=25c_{2}(X)\cdot c_{1}(X) in the second case, which finishes the proof. ∎

Remark 4.5.

In the setting of Theorem 4.4, if ιX=ι^X=4\iota_{X}=\hat{\iota}_{X}=4 and XX satisfies the following inequality

12141c2(X)c1(X)<c1(X)36421c2(X)c1(X),\frac{121}{41}c_{2}(X)\cdot c_{1}(X)<c_{1}(X)^{3}\leq\frac{64}{21}c_{2}(X)\cdot c_{1}(X),

then applying the same computer program as in the proof of Theorem 4.4 shows that there is only one possible numerical type for such XX:

X={7,13},c2(X)c1(X)=38491andc1(X)3=115291.\mathcal{R}_{X}=\{7,13\},\quad c_{2}(X)\cdot c_{1}(X)=\frac{384}{91}\quad\textup{and}\quad c_{1}(X)^{3}=\frac{1152}{91}.

In particular, in the setting of Theorem 4.4, if c1(X)33c2(X)c1(X)c_{1}(X)^{3}\geq 3c_{2}(X)\cdot c_{1}(X), then one of the following statements holds:

  1. (4.5.1)

    ιXι^X=4\iota_{X}\not=\hat{\iota}_{X}=4;

  2. (4.5.2)

    ιX=ι^X=4\iota_{X}=\hat{\iota}_{X}=4 and X={7,13}\mathcal{R}_{X}=\{7,13\};

  3. (4.5.3)

    ιX=ι^X=5\iota_{X}=\hat{\iota}_{X}=5 and X={3,72}\mathcal{R}_{X}=\{3,7^{2}\}.

Moreover, if we assume in addition that ι^X4\hat{\iota}_{X}\geq 4, then ``3"``\geq 3" in the inequality above can be replaced by ``>121/41"``>121/41".

By [IJL23, Theorem 1.2], we have c2(X)c1(X)1/252c_{2}(X)\cdot c_{1}(X)\geq 1/252 for any weak \mathbb{Q}-Fano threefold XX with terminal singularities. Now we can improve further this lower bound in the Picard number one case.

Corollary 4.6.

Let XX be a \mathbb{Q}-Fano threefold with terminal singularities and ρ(X)=1\rho(X)=1. Then we have

c2(X)c1(X)29546,c_{2}(X)\cdot c_{1}(X)\geq\frac{29}{546},

where the equality holds only if X={2,3,7,13}\mathcal{R}_{X}=\{2,3,7,13\} and ιX=ι^X=1\iota_{X}=\hat{\iota}_{X}=1.

Proof.

Firstly we use a computer program to find out all possible baskets X\mathcal{R}_{X} satisfying (2.3.2) and

0<c2(X)c1(X)<110.0<c_{2}(X)\cdot c_{1}(X)<\frac{1}{10}.

Then we pick up all possible values of c1(X)3c_{1}(X)^{3} satisfying both (2.4.1) and Theorem 4.4. Note that (2.4.4) is not used and so this algorithm works for arbitrary ι^X\hat{\iota}_{X}. Finally we obtain only three possibilities for the numerical type of such XX as listed in the following table:

X\mathcal{R}_{X} c2(X)c1(X)c_{2}(X)\cdot c_{1}(X) c1(X)3c_{1}(X)^{3}
{22,33,13}\{2^{2},3^{3},13\} 113\frac{1}{13} 113\frac{1}{13}
{22,33,13}\{2^{2},3^{3},13\} 113\frac{1}{13} 313\frac{3}{13}
{2,3,7,13}\{2,3,7,13\} 29546\frac{29}{546} 61546\frac{61}{546}

Now the desired inequality follows immediately and the equality holds only in the last case with ιX=1\iota_{X}=1, which can be derived from (2.4.2) as follows:

rXc1(X)3ιX3=546×61546ιX3=61ιX30.r_{X}\cdot\frac{c_{1}(X)^{3}}{\iota_{X}^{3}}=546\times\frac{61}{546\iota_{X}^{3}}=\frac{61}{\iota_{X}^{3}}\in\mathbb{Z}_{\geq 0}.

The same argument also works for ι^X\hat{\iota}_{X} by the proof of [Suz04, Lemma 1.2 (2)]. ∎

Remark 4.7.
  1. (4.7.1)

    By Theorem 4.4, we can further rule out some of the possible Hilbert series PX(t)=mh0(X,mKX)tmP_{X}(t)=\sum_{m\in\mathbb{N}}h^{0}(X,-mK_{X})t^{m} of \mathbb{Q}-Fano threefolds with terminal singularities and Picard number one (for a complete list of possible Hilbert series, see [BK22] and the references therein).

  2. (4.7.2)

    Similar to the geography problem of surfaces, for any \mathbb{Q}-Fano variety XX with terminal singularities and ρ(X)=1\rho(X)=1, we can ask what the sharp bounds ana_{n} and bnb_{n} depending only on n=dimXn=\dim X are such that

    anc2(X)c1(X)n2c1(X)nbnc2(X)c1(X)n2.a_{n}c_{2}(X)\cdot c_{1}(X)^{n-2}\leq c_{1}(X)^{n}\leq b_{n}c_{2}(X)\cdot c_{1}(X)^{n-2}.

    The existence of ana_{n} and bnb_{n} follows from the boundedness of such varieties [Bir21] and bn<4b_{n}<4 by Theorem 1.1. We refer the reader to [DS22, LX24] for other related works on this kind of problem and [LL24] for our future work to prove b3<3b_{3}<3.

References

  • [AD13] Carolina Araujo and Stéphane Druel. On Fano foliations. Adv. Math., 238:70–118, 2013.
  • [AD14] Carolina Araujo and Stéphane Druel. On codimension 1 del Pezzo foliations on varieties with mild singularities. Math. Ann., 360(3-4):769–798, 2014.
  • [Bir21] Caucher Birkar. Singularities of linear systems and boundedness of Fano varieties. Ann. of Math. (2), 193(2):347–405, 2021.
  • [BK22] Gavin Brown and Alexander Kasprzyk. Kawamata boundedness for Fano threefolds and the Graded Ring Database. arXiv preprint arXiv:2201.07178, 2022.
  • [BS58] Armand Borel and Jean-Pierre Serre. Le théorème de Riemann-Roch. Bull. Soc. Math. France, 86:97–136, 1958.
  • [CP19] Frédéric Campana and Mihai Păun. Foliations with positive slopes and birational stability of orbifold cotangent bundles. Publ. Math. Inst. Hautes Études Sci., 129:1–49, 2019.
  • [Dru17] Stéphane Druel. On foliations with nef anti-canonical bundle. Trans. Amer. Math. Soc., 369(11):7765–7787, 2017.
  • [DS22] Rong Du and Hao Sun. Inequalities of Chern classes on nonsingular projective nn-folds with ample canonical or anti-canonical line bundles. J. Differential Geom., 122(3):377–398, 2022.
  • [GKKP11] Daniel Greb, Stefan Kebekus, Sándor J. Kovács, and Thomas Peternell. Differential forms on log canonical spaces. Publ. Math. Inst. Hautes Études Sci., (114):87–169, 2011.
  • [GKP21] Daniel Greb, Stefan Kebekus, and Thomas Peternell. Projectively flat klt varieties. J. Éc. polytech. Math., 8:1005–1036, 2021.
  • [IJL23] Masataka Iwai, Chen Jiang, and Haidong Liu. Miyaoka type inequality for terminal threefolds with nef anti-canonical divisors. Sci. China Math., to appear, 2023.
  • [IP99] Vasily Alexeevich Iskovskikh and Yuri G. Prokhorov. Algebraic geometry. V, volume 47 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, 1999. Fano varieties, A translation of ıt Algebraic geometry. 5 (Russian), Ross. Akad. Nauk, Vseross. Inst. Nauchn. i Tekhn. Inform., Moscow, Translation edited by A. N. Parshin and I. R. Shafarevich.
  • [Kaw88] Yujiro Kawamata. Crepant blowing-up of 3-dimensional canonical singularities and its application to degenerations of surfaces. Ann.of Math. (2), 127(1):93–163, 1988.
  • [Kaw92] Yujiro Kawamata. Boundedness of \mathbb{Q}–Fano threefolds. In Proceedings of the International Conference on Algebra, Part 3 (Novosibirsk, 1989), volume 131 of Contemp. Math., pages 439–445. Amer. Math. Soc., Providence, RI, 1992.
  • [KM98] János Kollár and Shigefumi Mori. Birational geometry of algebraic varieties, volume 134 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 1998.
  • [Lan04] Adrian Langer. Semistable sheaves in positive characteristic. Ann. of Math. (2), 159(1):251–276, 2004.
  • [Liu19] Jie Liu. Second Chern class of Fano manifolds and anti-canonical geometry. Math. Ann., 375(1-2):655–669, 2019.
  • [LL24] Haidong Liu and Jie Liu. Kawamata–Miyaoka type inequality for \mathbb{Q}-Fano varieties with canonical singularities II: terminal \mathbb{Q}-Fano threefolds. Épijournal Géom. Algébrique, to appear, 2024.
  • [LX24] Xing Lu and Jian Xiao. The inequalities of Chern classes and Riemann–Roch type inequalities. Adv. Math., 458:109982, 2024.
  • [Mel99] Massimiliano Mella. Existence of good divisors on Mukai varieties. J. Algebraic Geom., 8(2):197–206, 1999.
  • [Miy87] Yoichi Miyaoka. The Chern classes and Kodaira dimension of a minimal variety. In Algebraic geometry, Sendai, 1985, volume 10 of Adv. Stud. Pure Math., pages 449–476. North-Holland, Amsterdam, 1987.
  • [Mor85] Shigefumi Mori. On 33-dimensional terminal singularities. Nagoya Math. J., 98:43–66, 1985.
  • [Muk89] Shigeru Mukai. Biregular classification of Fano 33-folds and Fano manifolds of coindex 33. Proc. Nat. Acad. Sci. U.S.A., 86(9):3000–3002, 1989.
  • [Ou23] Wenhao Ou. On generic nefness of tangent sheaves. Math. Z., 304(4):58, 2023.
  • [Pet12] Thomas Peternell. Varieties with generically nef tangent bundles. J. Eur. Math. Soc. (JEMS), 14(2):571–603, 2012.
  • [Pro10] Yu. G. Prokhorov. \mathbb{Q}-Fano threefolds of large Fano index, I. Doc. Math., 15:843–872, 2010.
  • [Pro13] Yu. G. Prokhorov. On Fano threefolds of large Fano index and large degree. Sb. Math., 204(4):347–382, 2013.
  • [Pro22] Yu. G. Prokhorov. Rationality of \mathbb{Q}-Fano threefolds of large Fano index. In London Mathematical Society Lecture Note Series, volume 478, pages 253–274. Cambridge University Press, 2022.
  • [PW95] Thomas Peternell and Jarosław A. Wiśniewski. On stability of tangent bundles of Fano manifolds with b2=1b_{2}=1. J. Algebraic Geom., 4(2):363–384, 1995.
  • [Rei87] Miles Reid. Young person’s guide to canonical singularities. In Algebraic geometry, Bowdoin, 1985 (Brunswick, Maine, 1985), volume 46 of Proc. Sympos. Pure Math., pages 345–414. Amer. Math. Soc., Providence, RI, 1987.
  • [San96] Takeshi Sano. Classification of non-Gorenstein {\mathbb{Q}}-Fano dd-folds of Fano index greater than d2d-2. Nagoya Math. J., 142:133–143, 1996.
  • [Suz04] Kaori Suzuki. On Fano indices of \mathbb{Q}-Fano 3-folds. Manuscripta Math., 114(2):229–246, 2004.
  • [Suz24] Kaori Suzuki. The graded ring database for Fano 3-folds and the Bogomolov stability bound. Ann Univ Ferrara, to appear, 2024.
  • [Zha06] Qi Zhang. Rational connectedness of log 𝐐{\bf Q}-Fano varieties. J. Reine Angew. Math., 590:131–142, 2006.