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

Vanishing theorems for Fano threefolds in positive characteristic

Tatsuro Kawakami Department of Mathematics, Graduate School of Science, Kyoto University, Kyoto 606-8502, Japan [email protected]  and  Hiromu Tanaka Graduate School of Mathematical Sciences, The University of Tokyo, 3-8-1 Komaba, Meguro-ku, Tokyo 153-8914, JAPAN [email protected]
Abstract.

We prove that Kodaira vanishing holds for an arbitrary smooth Fano threefold in positive characteristic. To this end, we show that it is quasi-FF-split when the Picard number or the Fano index is larger than one. We also establish Akizuki-Nakano vanishing for smooth Fano threefolds when the Picard number or the Fano index is larger than one, and therefore they lift to the ring of Witt vectors.

Key words and phrases:
Fano threefolds, (quasi-)FF-split, Kodaira vanishing
2020 Mathematics Subject Classification:
14J45, 13A35, 14F17

1. Introduction

In the context of the minimal model program, Fano varieties play a significant role in the classification of algebraic varieties. The classification of Fano varieties in characteristic zero has a long history. In the early nineteenth century, Gino Fano established a partial classification result for smooth Fano threefolds, in order to attack the rationality problem of cubic threefolds. In 1980s, the classification of smooth Fano threefolds was carried out by Mori–Mukai ([MM81], [MM83]), based on earlier works by Iskovskih and Shokurov. Recently, this classification result has been extended to the case of positive characteristic [SB97, Meg98, FanoI, FanoII, FanoIII, FanoIV].

In this paper, our focus lies in vanishing theorems for them. Kodaira vanishing for Fano varieties is particularly crucial since their anti-canonical divisors are ample. For example, Kodaira vanishing implies Hi(X,𝒪X(D))=0H^{i}(X,\mathcal{O}_{X}(D))=0 for every nef divisor DD on a smooth Fano variety XX. From now on, we work over an algebraically closed field kk of characteristic p>0p>0. Let XX be a smooth Fano threefold over kk. Then, for an ample Cartier divisor AA, the vanishing H1(X,𝒪X(A))=0H^{1}(X,\mathcal{O}_{X}(-A))=0 has been proven by Shepherd-Barron and the first author [SB97], [Kaw2, Theorem 1.1] (cf. [FanoI, Theorem 2.4]). If ρ(X)=rX=1\rho(X)=r_{X}=1 for the Picard number ρ(X)\rho(X) and the index rXr_{X}, then Kodaira vanishing is established in [FanoI, Corollary 4.5]. In this case, the vanishing is simple since AA can be written as AnKXA\sim-nK_{X}. On the other hand, as the Picard number increases, the description of ample divisors become more complicated, making it harder to establish Kodaira vanishing.

To address this issue, we turn to FF-splitting, as it implies Kodaira vanishing. Consequently, investigating FF-splitting of smooth Fano threefolds leads us to our goal: Kodaira vanishing for smooth Fano threefolds. However, it is known that smooth del Pezzo surfaces of characteristic p5p\leq 5 are not necessarily FF-split. Then, taking the product with 1\mathbb{P}^{1}, we can exhibit non-FF-split smooth Fano threefolds. To overcome this issue, we shall use quasi-FF-splitting, introduced by Yobuko [Yob19], which is a weaker condition than FF-splitting. Although quasi-FF-splitting is not as restrictive as FF-splitting, quasi-FF-split varieties satisfy various useful vanishing theorems including Kodaira vanishing.

Since every smooth del Pezzo surface is known to be quasi-FF-split [KTTWYY1, Corollary 4.7], it is tempting to extend this result to smooth Fano threefolds. However, the Fermat quartic hypersurface {x04+x14+x24+x34+x44=0}k4\{x_{0}^{4}+x_{1}^{4}+x_{2}^{4}+x_{3}^{4}+x_{4}^{4}=0\}\subset\mathbb{P}_{k}^{4} is not quasi-FF-split in characteristic three [KTY, Example 7.12]. Moreover, we find another non-quasi-FF-split smooth Fano threefold, which is a weighted hypersurface X{x06+x16+x26+x36+y2=0}(1,1,1,1,3)[x0:x1:x2:x3:y]X\coloneqq\{x_{0}^{6}+x_{1}^{6}+x_{2}^{6}+x_{3}^{6}+y^{2}=0\}\subset\mathbb{P}(1,1,1,1,3)_{[x_{0}:x_{1}:x_{2}:x_{3}:y]} in characteristic five. On the other hand, each of these examples satisfies ρ(X)=rX=1\rho(X)=r_{X}=1. Hence it is natural to expect quasi-FF-splitting for many smooth Fano threefolds such that ρ(X)>1\rho(X)>1 or rX>1r_{X}>1. In fact, we prove the following theorem.

Theorem A.

Let XX be a smooth Fano threefold over an algebraically closed field of positive characteristic. Then the following hold.

  1. (1)

    XX is quasi-FF-split if ρ(X)>1\rho(X)>1 or rX>1r_{X}>1, where ρ(X)\rho(X) is the Picard number and rXr_{X} denotes the index.

  2. (2)

    Hi(X,𝒪X(A))=0H^{i}(X,\mathcal{O}_{X}(-A))=0 for every i<3i<3 and every ample Cartier divisor AA.

Surprisingly, Theorem A asserts that Fano threefolds with wild conic bundle structures (i.e., conic bundles which are not generically smooth) are all quasi-FF-split. These varieties are known to be non-FF-split. Therefore, quasi-FF-splitting is a notion which can be applicable to such pathological varieties in positive characteristic, allowing it to establish many useful vanishing theorems. This is a significant advantage of quasi-FF-splitting.

1.1. Akizuki-Nakano vanishing

As another remarkable property of quasi-FF-splitting, it is known that every quasi-FF-split smooth variety lifts to W2(k)W_{2}(k) together with an arbitarary effective Cartier divisor [AZ21], [KTTWYY1, Section 7.2]. From this, we can deduce the logarithmic Akizuki-Nakano vanishing if p>2p>2. Moreover, we can also prove the non-logarithmic Akizuki-Nakano vanishing holds for every p>0p>0. To summarise, we obtain the following theorem.

Theorem B.

Let XX be a smooth Fano threefold over an algebraically closed field of characteristic p>0p>0 such that ρ(X)>1\rho(X)>1 or rX>1r_{X}>1. Then the following hold.

  1. (1)

    Hj(X,ΩXi𝒪X(A))=0H^{j}(X,\Omega^{i}_{X}\otimes\mathcal{O}_{X}(-A))=0 for every ample Cartier divisor AA and every pair (i,j)(i,j) of integers ii and jj satisfying i+j<3i+j<3.

  2. (2)

    Assume p2p\neq 2. Take a reduced divisor EE with simple normal crossing support and an ample \mathbb{Q}-divisor AA such that the support of the fractional part of AA is contained in EE. Then

    Hj(X,ΩXi(logE)𝒪X(A))=0H^{j}(X,\Omega^{i}_{X}(\log E)\otimes\mathcal{O}_{X}(-\lceil A\rceil))=0

    holds for every pair (i,j)(i,j) of integers ii and jj satisfying i+j<3i+j<3.

As an immediate consequence, we obtain the following theorem.

Theorem C.

Every smooth Fano threefold XX over an algebraically closed field kk of positive characteristic such that ρ(X)>1\rho(X)>1 or rX>1r_{X}>1 lifts to the ring W(k)W(k) of Witt vectors.

1.2. FF-splitting

It is well known that smooth del Pezzo surfaces are FF-split if p>5p>5. Then it is natural ask when a smooth Fano threefold is FF-split. In the process of the proof of Theorem A, we show the following theorem.

Theorem D.

Every smooth Fano threefold over an algebraically closed field of characteristic p>5p>5 such that ρ(X)>1\rho(X)>1 or rX>1r_{X}>1 is FF-split.

Remark 1.1.

FF-splitting of some smooth Fano threefolds has been proven by Totaro in a different way [Totaro(Fano), the proof of Lemma 1.5].

Remark 1.2.
  1. (1)

    As mentioned above, if SS is a non-FF-split del Pezzo surface, then X:=S×1X:=S\times\mathbb{P}^{1} is a smooth Fano threefold which is not FF-split. Since this construction can be applicable for p{2,3,5}p\in\{2,3,5\}, the assumption p>5p>5 in Theorem D is optimal.

  2. (2)

    If p=7p=7, then the Fermat quartic hypersurface {x04+x14+x24+x34+x44=0}k4\{x_{0}^{4}+x_{1}^{4}+x_{2}^{4}+x_{3}^{4}+x_{4}^{4}=0\}\subset\mathbb{P}_{k}^{4} is not FF-split by Fedder’s criterion.

  3. (3)

    If p=11p=11, then we shall prove that there exists a smooth Fano threefold which is not FF-split (Example 8.4).

  4. (4)

    The authors do not know whether there exists a non-FF-split smooth Fano threefold of characteristic p13p\geq 13.

1.3. Strategy of proofs

We now overview how to show that a smooth Fano threefold with ρ(X)2\rho(X)\geq 2 is quasi-FF-split (Theorem A(1)). In order to show that XX is quasi-FF-split, we shall use one of the following strategies.

  1. (A)

    Inversion of adjunction for FF-splitting.

  2. (B)

    Inversion of adjunction for quasi-FF-splitting.

  3. (C)

    Cartier operator criterion for quasi-FF-splitting.

(A) For many cases (e.g., 3ρ(X)53\leq\rho(X)\leq 5 except for No. 3-10), we can prove that XX is FF-split. For example, let us consider the case when XX is of No. 2-30, i.e., there is a blowup X3X\to\mathbb{P}^{3} along a smooth conic BB on 3\mathbb{P}^{3}. Take the plane D(2)D(\simeq\mathbb{P}^{2}) containing BB. For the the proper transform DXD_{X} of DD on XX, we have the following implications:

D is F-split(i)(3,D) is F-split(ii)(X,DX) is F-split(iii)X is F-split.D\text{ is $F$-split}\overset{{\rm(i)}}{\Rightarrow}(\mathbb{P}^{3},D)\text{ is $F$-split}\overset{{\rm(ii)}}{\Leftrightarrow}(X,D_{X})\text{ is $F$-split}\overset{{\rm(iii)}}{\Rightarrow}X\text{ is $F$-split}.

Of course, D(2)D(\simeq\mathbb{P}^{2}) is FF-split. The implication (i) follows from the inversion of adjunction for FF-splitting, e.g., if (Y,D)(Y,D) is FF-split and (KY+D)-(K_{Y}+D) is ample, then (Y,D)(Y,D) is FF-split. The equivalence (ii) is assured by KX+DX=f(K3+D)K_{X}+D_{X}=f^{*}(K_{\mathbb{P}^{3}}+D). Finally, (iii) holds by definition.

(B) Even if the strategy (A) does not work, we can apply a quasi-FF-split version of (A) for most of the remaining cases. The authors has proved that an inversion of adjunction for log Calabi-Yau pairs in [Kawakami-Tanaka(dPvar)]. This allows us to prove the following statement: a smooth Fano threefold XX is quasi-FF-split if there are smooth prime divisors SS and SS^{\prime} such that KX+S+S0K_{X}+S+S^{\prime}\sim 0 and SSS\cap S^{\prime} is a smooth curve which is quasi-FF-split (Corollary 2.18).

As other technical issues, we shall encounter the following obstructions:

  • SS is not necessarily smooth. For some cases, the generic member is enough as a replacement (cf. Section 2.3). To this end, we shall need to treat algebraic varieties defined over an imperfect field.

  • If the first prime divisor SS is a smooth weak del Pezzo surface, then it is often hard to find SS^{\prime} such that SSS\cap S^{\prime} is smooth. For example, if |KS||-K_{S}| has no smooth member, then SSS\cap S^{\prime} can not be smooth. In order to avoid such a pathological phenomenon, we shall establish some properties on weak del Pezzo surfaces, e.g., if VV is a surface over a C1C_{1}-field KK of characteristic two and its base change V×SpecKSpecK¯V\times_{\operatorname{Spec}K}\operatorname{Spec}\overline{K} is a Langer surface (defined as the base change of the blowup of 𝔽22\mathbb{P}^{2}_{\mathbb{F}_{2}} along all the 𝔽2\mathbb{F}_{2}-rational points), then ρ(V)=8\rho(V)=8 (Lemma 5.5).

(C) Except when XX is one of 2-2, 2-6, 2-8, and 3-10, we may apply one of (A) and (B). For these remaining cases, we shall apply a quasi-FF-splitting criterion via Cartier operator, which has been established in [KTTWYY1, Theorem F] (cf. Proposition 2.20). To this end, we need to prove Hj(X,ΩXi(pKX))=0H^{j}(X,\Omega^{i}_{X}(p^{\ell}K_{X}))=0 for suitable triples (i,j,)(i,j,\ell). Even if XX is explicitly given, it is often hard to compute such cohomologies directly. The main strategy is to embed XX into a (typically toric) fourfold PP, and apply Bott vanishing for PP. For example, if XX is of No. 2-2, then we can find such an embedding with P=1×2(𝒪1×2𝒪1×2(1,2))P=\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{2}}(-1,-2)). Although PP is not necessaliry toric for the other cases, we shall find a closed embedding XPX\hookrightarrow P to a fourfold PP which almost satisfies Bott vanishing. For more details on (C), see Section 6.

Acknowledgements. The authors express their gratitude to Burt Totaro for valuable comments. They also thank Teppei Takamatsu and Shou Yoshikawa for useful conversations. Kawakami was supported by JSPS KAKENHI Grant number JP22J00272. Tanaka was supported by JSPS KAKENHI Grant number JP22H01112 and JP23K03028.

2. Preliminaries

2.1. Notation

In this subsection, we summarise notation and basic definitions used in this article.

  1. (1)

    Throughout the paper, pp denotes a prime number and we set 𝔽p/p\mathbb{F}_{p}\coloneqq\mathbb{Z}/p\mathbb{Z}. Unless otherwise specified, we work over an algebraically closed field kk of characteristic p>0p>0. We denote by F:XXF\colon X\to X the absolute Frobenius morphism on an 𝔽p\mathbb{F}_{p}-scheme XX.

  2. (2)

    We say that XX is a variety (over a field κ\kappa) if XX is an integral scheme that is separated and of finite type over κ\kappa. We say that XX is a curve (resp. surface, resp. threefold) if XX is a variety of dimension one (resp. two, resp. three).

  3. (3)

    For a variety XX, we define the function field K(X)K(X) of XX as the stalk 𝒪X,ξ\mathcal{O}_{X,\xi} at the generic point ξ\xi of XX.

  4. (4)

    We say that an \mathbb{R}-divisor DD on a normal variety XX is simple normal crossing if for every point xSuppDx\in\operatorname{Supp}D, the local ring 𝒪X,x\mathcal{O}_{X,x} is regular and there exists a regular system of parameters x1,,xdx_{1},\ldots,x_{d} of the maximal ideal 𝔪\mathfrak{m} of 𝒪X,x\mathcal{O}_{X,x} and 1rd1\leq r\leq d such that Supp(D|Spec𝒪X,x)=Spec(𝒪X,x/(x1xr))\operatorname{Supp}(D|_{\operatorname{Spec}\mathcal{O}_{X,x}})=\operatorname{Spec}(\mathcal{O}_{X,x}/(x_{1}\cdots x_{r})).

  5. (5)

    Given an integral normal Noetherian scheme XX, a projective birational morphism π:YX\pi\colon Y\to X is called a log resolution (of singularities) of XX if YY is regular and Exc(f)\mathrm{Exc}(f) is a simple normal crossing divisor.

  6. (6)

    We say that an 𝔽p\mathbb{F}_{p}-scheme XX is FF-finite if the absolute Frobenius morphism F:XXF:X\to X is a finite morphism. We say that an 𝔽p\mathbb{F}_{p}-algebra RR is FF-finite if SpecR\operatorname{Spec}R is FF-finite. In particular, a field κ\kappa is FF-finite if and only if [κ:κp]<[\kappa:\kappa^{p}]<\infty. If XX is a variety over an FF-finite field, then XX is FF-finite.

  7. (7)

    Given a normal variety XX and an \mathbb{R}-divisor DD, we define the subsheaf 𝒪X(D)\mathcal{O}_{X}(D) of the constant sheaf K(X)K(X) on XX by the following formula

    Γ(U,𝒪X(D))={φK(X)(div(φ)+D)|U0}\Gamma(U,\mathcal{O}_{X}(D))=\{\varphi\in K(X)\mid\left({\rm div}(\varphi)+D\right)|_{U}\geq 0\}

    for every open subset UU of XX. In particular, 𝒪X(D)=𝒪X(D)\mathcal{O}_{X}(\lfloor{D}\rfloor)=\mathcal{O}_{X}({D}).

  8. (8)

    Given a field KK and KK-schemes XX and YY, we say that XX is KK-isomorphic to YY if there exists an isomorphism θ:XY\theta:X\to Y over KK.

  9. (9)

    Given a closed subscheme XX of n\mathbb{P}^{n}, we set 𝒪X(a):=𝒪n(a)|X\mathcal{O}_{X}(a):=\mathcal{O}_{\mathbb{P}^{n}}(a)|_{X} for aa\in\mathbb{Z} unless otherwise specified. Similarly, if YY is a closed subscheme of n×m\mathbb{P}^{n}\times\mathbb{P}^{m}, then we define 𝒪Y(a,b)𝒪n×m(a,b)\mathcal{O}_{Y}(a,b)\coloneqq\mathcal{O}_{\mathbb{P}^{n}\times\mathbb{P}^{m}}(a,b) for a,ba,b\in\mathbb{Z}.

  10. (10)

    Given a coherent sheaf \mathcal{F} and a Cartier divisor DD on a variety XX, we set (D)𝒪X(D)\mathcal{F}(D)\coloneqq\mathcal{F}\otimes\mathcal{O}_{X}(D) unless otherwise specified. Note that BnΩXi(pnD)B_{n}\Omega_{X}^{i}(p^{n}D) (resp. ZnΩXi(pnD)Z_{n}\Omega_{X}^{i}(p^{n}D)) does not mean BnΩXi𝒪X(pnD)B_{n}\Omega_{X}^{i}\otimes\mathcal{O}_{X}(p^{n}D) (resp. ZnΩXi𝒪X(pnD)Z_{n}\Omega_{X}^{i}\otimes\mathcal{O}_{X}(p^{n}D)) even if DD is Cartier (cf. Subsection 2.2).

  11. (11)

    Given two closed subschemes YY and ZZ on a scheme XX, we denote by YZY\cap Z the scheme-theoretic intersection, i.e., YZY×XZY\cap Z\coloneqq Y\times_{X}Z.

2.2. Cartier operators

In this section, we recall the fundamental facts on the higher Cartier operators ([Ill79], [KTTWYY1]).

Let XX be a smooth variety over a perfect field of characteristic p>0p>0 and DD a Cartier divisor on XX. The Frobenius pushforward of the de Rham complex

FΩX:F𝒪XFdFΩXFdF_{*}\Omega^{\bullet}_{X}\colon F_{*}\mathcal{O}_{X}\xrightarrow{F_{*}d}F_{*}\Omega_{X}\xrightarrow{F_{*}d}\cdots

is a complex of 𝒪X\mathcal{O}_{X}-modules. Tensoring with 𝒪X(D)\mathcal{O}_{X}(D), we obtain a complex

FΩX:F𝒪X(pD)Fd𝒪X(D)FΩX(pD)Fd𝒪X(D)F_{*}\Omega^{\bullet}_{X}\colon F_{*}\mathcal{O}_{X}(pD)\xrightarrow{F_{*}d\otimes\mathcal{O}_{X}(D)}F_{*}\Omega_{X}(pD)\xrightarrow{F_{*}d\otimes\mathcal{O}_{X}(D)}\cdots

We define locally free 𝒪X\mathcal{O}_{X}-modules as follows.

B1ΩXi(pD)Im(Fd:FΩXi1(pD)FΩXi(pD)),Z1ΩXi(pD)Ker(Fd:FΩXi(pD)FΩXi+1(pD)).\begin{array}[]{rl}&B^{1}\Omega^{i}_{X}(pD)\coloneqq\mathrm{Im}(F_{*}d:F_{*}\Omega^{i-1}_{X}(pD)\to F_{*}\Omega^{i}_{X}(pD)),\\ &Z_{1}\Omega^{i}_{X}(pD)\coloneqq\operatorname{Ker}(F_{*}d:F_{*}\Omega^{i}_{X}(pD)\to F_{*}\Omega^{i+1}_{X}(pD)).\\ \end{array}

We have an isomorphism

Z1ΩXi(pD)/B1ΩXi(pD)C(D)ΩXi(D).Z_{1}\Omega_{X}^{i}(pD)/B_{1}\Omega_{X}^{i}(pD)\overset{C(D)}{\simeq}\Omega_{X}^{i}(D).

resulting from the Cartier isomorphism. In fact, tensoring with 𝒪X(D)\mathcal{O}_{X}(D) with the usual Cartier isomorphism

Z1ΩXi/B1ΩXi𝐶ΩXi,Z_{1}\Omega_{X}^{i}/B_{1}\Omega_{X}^{i}\overset{C}{\simeq}\Omega_{X}^{i},

we obtain the above isomorphism.

Taking the Frobenius pushforward, we obtain

FZ1ΩXi(p2D)FZ1ΩXi(p2D)/FB1ΩXi(p2D)FC(D)FΩXi(pD).F_{*}Z_{1}\Omega_{X}^{i}(p^{2}D)\to F_{*}Z_{1}\Omega_{X}^{i}(p^{2}D)/F_{*}B_{1}\Omega_{X}^{i}(p^{2}D)\overset{F_{*}C(D)}{\simeq}F_{*}\Omega_{X}^{i}(pD).

We denote by B2ΩXi(p2D)B_{2}\Omega_{X}^{i}(p^{2}D) and Z2ΩXi(p2D)Z_{2}\Omega_{X}^{i}(p^{2}D) the preimages of B1ΩXi(pD)FΩXi(pD)B_{1}\Omega_{X}^{i}(pD)\subset F_{*}\Omega_{X}^{i}(pD) and Z1ΩXi(pD)FΩXi(pD)Z_{1}\Omega_{X}^{i}(pD)\subset F_{*}\Omega_{X}^{i}(pD) by the above map. Inductively, we define locally 𝒪X\mathcal{O}_{X}-module BnΩXi(pnD)B_{n}\Omega_{X}^{i}(p^{n}D) and ZnΩXi(pnD)Z_{n}\Omega_{X}^{i}(p^{n}D) for all n0n\geq 0. Moreover, we set B0ΩXi(D)=0B_{0}\Omega_{X}^{i}(D)=0 and Z0ΩXi(pD)=ΩXi(pD)Z_{0}\Omega_{X}^{i}(pD)=\Omega_{X}^{i}(pD).

Lemma 2.1.

Then we have the following exact sequences

(2.1.1) 0BnΩXi(pnD)ZnΩXi(pnD)ΩXi(D)0.\displaystyle 0\to B_{n}\Omega^{i}_{X}(p^{n}D)\to Z_{n}\Omega^{i}_{X}(p^{n}D)\to\Omega^{i}_{X}(D)\to 0.
(2.1.2) 0ZnΩXi(pnD)FZn1ΩXi(pnD)B1ΩXi+1(pD)0.\displaystyle 0\to Z_{n}\Omega_{X}^{i}(p^{n}D)\to F_{*}Z_{n-1}\Omega_{X}^{i}(p^{n}D)\to B_{1}\Omega_{X}^{i+1}(pD)\to 0.

for all i0i\geq 0 and all n1n\geq 1.

Proof.

The assertion follows from [KTTWYY1, (5.7.1) and Lemma 5.8]. ∎

Remark 2.2.

Taking n=1n=1, we have the following exact sequence:

(2.2.1) 0B1ΩXi(pD)Z1ΩXi(pD)C(D)ΩXi(D)0.\displaystyle 0\to B_{1}\Omega^{i}_{X}(pD)\to Z_{1}\Omega^{i}_{X}(pD)\xrightarrow{C(D)}\Omega^{i}_{X}(D)\to 0.
(2.2.2) 0Z1ΩXi(pD)FΩXi(pD)Fd𝒪X(D)B1ΩXi+1(pD)0.\displaystyle 0\to Z_{1}\Omega_{X}^{i}(pD)\to F_{*}\Omega_{X}^{i}(pD)\xrightarrow{F_{*}d\otimes\mathcal{O}_{X}(D)}B_{1}\Omega_{X}^{i+1}(pD)\to 0.

for all i0i\geq 0.

Remark 2.3.

Taking D=0D=0, we have short exact sequences

(2.3.1) 0BnΩXiZnΩXiCnΩXi0,\displaystyle 0\to B_{n}\Omega^{i}_{X}\to Z_{n}\Omega^{i}_{X}\xrightarrow{C^{n}}\Omega^{i}_{X}\to 0,
(2.3.2) 0ZnΩXiFZn1ΩXiFdFCn1B1ΩXi+10,\displaystyle 0\to Z_{n}\Omega_{X}^{i}\to F_{*}Z_{n-1}\Omega_{X}^{i}\xrightarrow{F_{*}d\circ F_{*}C^{n-1}}B_{1}\Omega_{X}^{i+1}\to 0,

which coincides with [KTTWYY1, (2.15.1) and Lemma 2.16] respectively. Then we can confirm that

(2.1.1)=(2.3.1)𝒪X(D)and(2.1.2)=(2.3.2)𝒪X(D)\eqref{exact:B}=\eqref{exact:B,D=0}\otimes\mathcal{O}_{X}(D)\,\,\text{and}\,\,\eqref{exact:Z}=\eqref{exact:Z,D=0}\otimes\mathcal{O}_{X}(D)

holds for all i0i\geq 0. In particular,

BnΩXi(pnD)=BnΩXi𝒪X(D)and\displaystyle B_{n}\Omega_{X}^{i}(p^{n}D)=B_{n}\Omega_{X}^{i}\otimes\mathcal{O}_{X}(D)\,\,\text{and}
ZnΩXi(pnD)=ZnΩXi𝒪X(D)\displaystyle Z_{n}\Omega_{X}^{i}(p^{n}D)=Z_{n}\Omega_{X}^{i}\otimes\mathcal{O}_{X}(D)

hold for all n0n\geq 0.

2.3. Generic members

Let XX be a regular projective variety XX over a field kk and let DD be a Cartier divisor on XX satisfying h0(X,𝒪X(D))2h^{0}(X,\mathcal{O}_{X}(D))\geq 2. For a base point free linear system Λ|D|\Lambda\subset|D| and the corresponding linear subspace VΛH0(X,𝒪X(D))V_{\Lambda}\subset H^{0}(X,\mathcal{O}_{X}(D)), the generic member XΛgenX^{\operatorname{gen}}_{\Lambda} of Λ\Lambda is defined by the following diagram:

XΛgenXΛunivX×kκX×k(kn)XSpecκ(kn)SpeckΘpr2pr1κ:=K((kn))\leavevmode\hbox to172.13pt{\vbox to91.28pt{\pgfpicture\makeatletter\hbox{\hskip 86.06342pt\lower-45.69011pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-86.06342pt}{-45.59027pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 13.04025pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-8.73471pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${X^{\operatorname{gen}}_{\Lambda}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 13.04025pt\hfil&\hfil\hskip 38.20691pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-9.9014pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${X^{\mathrm{univ}}_{\Lambda}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 14.20694pt\hfil\cr\vskip 18.00005pt\cr\hfil\hskip 18.26686pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-13.96132pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${X\times_{k}\kappa}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 18.26686pt\hfil&\hfil\hskip 49.75838pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-21.45287pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${X\times_{k}(\mathbb{P}^{n}_{k})^{*}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 25.7584pt\hfil&\hfil\hskip 32.84023pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.53471pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${X}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 8.84026pt\hfil\cr\vskip 18.00005pt\cr\hfil\hskip 17.32524pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-13.0197pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\operatorname{Spec}\kappa}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 17.32524pt\hfil&\hfil\hskip 38.67786pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-10.37234pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${(\mathbb{P}^{n}_{k})^{*}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 14.67789pt\hfil&\hfil\hskip 42.03815pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-13.73264pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${\operatorname{Spec}\,k}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 18.03818pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} { {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{{ {\pgfsys@beginscope \pgfsys@setdash{}{0.0pt}\pgfsys@roundcap\pgfsys@roundjoin{} {}{}{} {}{}{} \pgfsys@moveto{-2.07988pt}{2.39986pt}\pgfsys@curveto{-1.69989pt}{0.95992pt}{-0.85313pt}{0.27998pt}{0.0pt}{0.0pt}\pgfsys@curveto{-0.85313pt}{-0.27998pt}{-1.69989pt}{-0.95992pt}{-2.07988pt}{-2.39986pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}{}{}{{}}\pgfsys@moveto{15.10657pt}{-43.09027pt}\pgfsys@lineto{49.3871pt}{-43.09027pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{49.58708pt}{-43.09027pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-50.27132pt}{-43.09027pt}\pgfsys@lineto{-15.04916pt}{-43.09027pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.84918pt}{-43.09027pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-35.1825pt}{-40.7375pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\Theta}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{68.02524pt}{-12.82774pt}\pgfsys@lineto{68.02524pt}{-34.38615pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{68.02524pt}{-34.58614pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{0.22868pt}{-15.32774pt}\pgfsys@lineto{0.22868pt}{-32.52783pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{0.22868pt}{-32.72781pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{2.58145pt}{-24.95416pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{{\rm pr}_{2}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-67.79655pt}{-14.77217pt}\pgfsys@lineto{-67.79655pt}{-34.49728pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-67.79655pt}{-34.69727pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-49.3297pt}{-6.46803pt}\pgfsys@lineto{-26.12968pt}{-6.46803pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-25.9297pt}{-6.46803pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{{ {\pgfsys@beginscope \pgfsys@setdash{}{0.0pt}\pgfsys@miterjoin\pgfsys@roundcap{{{}} {{}} {} {{{{}{}{}{}}}{{}{}{}{}}} } \pgfsys@moveto{0.0pt}{1.95987pt}\pgfsys@curveto{0.6848pt}{1.95987pt}{1.23993pt}{1.52113pt}{1.23993pt}{0.97993pt}\pgfsys@curveto{1.23993pt}{0.43874pt}{0.6848pt}{0.0pt}{0.0pt}{0.0pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}\pgfsys@moveto{0.22868pt}{19.85458pt}\pgfsys@lineto{0.22868pt}{4.0944pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{1.0}{1.0}{0.0}{0.22868pt}{19.85458pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{0.22868pt}{3.89442pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{}{{}}\pgfsys@moveto{-67.79655pt}{19.85458pt}\pgfsys@lineto{-67.79655pt}{2.12495pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}{{{}}{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{1.0}{1.0}{0.0}{-67.79655pt}{19.85458pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-67.79655pt}{1.92497pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-54.5563pt}{29.56754pt}\pgfsys@lineto{-14.57822pt}{29.56754pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.37823pt}{29.56754pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{14.63562pt}{21.91188pt}\pgfsys@lineto{58.63184pt}{-1.47849pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.88297}{-0.46944}{0.46944}{0.88297}{58.80841pt}{-1.57234pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{26.18709pt}{-6.46803pt}\pgfsys@lineto{58.58502pt}{-6.46803pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{58.785pt}{-6.46803pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{38.27075pt}{-2.75418pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{{\rm pr}_{1}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}\qquad\kappa:=K((\mathbb{P}^{n}_{k})^{*})

where

  1. (1)

    XΛunivX^{\mathrm{univ}}_{\Lambda} denotes the universal family that parametrises all the members of Λ\Lambda,

  2. (2)

    κ=K((kn))\kappa=K((\mathbb{P}^{n}_{k})^{*}) is the function field of the projective space (kn)(\mathbb{P}^{n}_{k})^{*} and Θ:SpecK((kn))(kn)\Theta:\operatorname{Spec}K((\mathbb{P}^{n}_{k})^{*})\to(\mathbb{P}^{n}_{k})^{*} is the induced morphism.

Then the following hold.

  1. (3)

    κ/k\kappa/k is a purely transcendental extension of finite transcendence degree.

  2. (4)

    X×kκX\times_{k}\kappa is a regular projective variety.

  3. (5)

    XΛgenX^{\operatorname{gen}}_{\Lambda} is a regular prime divisor [Tan-Bertini, Theorem 4.9(4)(12)].

For more details, we refer to [Tan-Bertini]. By abuse of notation, also (XΛgen)×κκ(X^{\operatorname{gen}}_{\Lambda})\times_{\kappa}\kappa^{\prime} is called the generic member when κ/κ\kappa^{\prime}/\kappa is a purely transcendental extension, because we shall encounter the situation as in the following remark.

Remark 2.4.

We now consider the case when we have two base point free linear systems Λ1\Lambda_{1} and Λ2\Lambda_{2} on XX. As above, we obtain two generic members XΛ1genX^{\operatorname{gen}}_{\Lambda_{1}} on X×κκ1X\times_{\kappa}\kappa_{1} and XΛ2genX^{\operatorname{gen}}_{\Lambda_{2}} on X×κκ2X\times_{\kappa}\kappa_{2}. For κ1=k(s1,,sa)\kappa_{1}=k(s_{1},\ldots,s_{a}) and κ2=k(t1,,tb)\kappa_{2}=k(t_{1},\ldots,t_{b}), i.e., each of {si}\{s_{i}\} and {tj}\{t_{j}\} is a transcendental basis, we set

κ:=Frac(k(s1,,sa)kk(t1,,tb))=k(s1,,sa,t1,,tb).\kappa:=\mathrm{Frac}(k(s_{1},\ldots,s_{a})\otimes_{k}k(t_{1},\ldots,t_{b}))=k(s_{1},\ldots,s_{a},t_{1},\ldots,t_{b}).

For (XΛ1gen)κ:=XΛ1gen×κ1κ(X^{\operatorname{gen}}_{\Lambda_{1}})_{\kappa}:=X^{\operatorname{gen}}_{\Lambda_{1}}\times_{\kappa_{1}}\kappa and (XΛ2gen)κ:=XΛ2gen×κ2κ(X^{\operatorname{gen}}_{\Lambda_{2}})_{\kappa}:=X^{\operatorname{gen}}_{\Lambda_{2}}\times_{\kappa_{2}}\kappa,

  1. ()(\star)

    the sum (XΛ1gen)κ+(XΛ2gen)κ(X^{\operatorname{gen}}_{\Lambda_{1}})_{\kappa}+(X^{\operatorname{gen}}_{\Lambda_{2}})_{\kappa} is a simple normal crossing divisor.

The similar statement holds even if we start with finitely many base point free linear systems Λ1,,Λr\Lambda_{1},\ldots,\Lambda_{r} on XX, i.e., the sum

(XΛ1gen)κ++(XΛrgen)κ(X^{\operatorname{gen}}_{\Lambda_{1}})_{\kappa}+\cdots+(X^{\operatorname{gen}}_{\Lambda_{r}})_{\kappa}

of the generic members (XΛ1gen)κ,,(XΛrgen)κ(X^{\operatorname{gen}}_{\Lambda_{1}})_{\kappa},\ldots,(X^{\operatorname{gen}}_{\Lambda_{r}})_{\kappa} is simple normal crossing, where κ:=Frac(κ1kkκr)\kappa:=\mathrm{Frac}(\kappa_{1}\otimes_{k}\cdots\otimes_{k}\kappa_{r}) and ()κ(-)_{\kappa} denotes the base change to κ\kappa.

Proof of ()(\star).

Both (XΛ1gen)κ(X^{\operatorname{gen}}_{\Lambda_{1}})_{\kappa} and (XΛ2gen)κ(X^{\operatorname{gen}}_{\Lambda_{2}})_{\kappa} are clearly regular prime divisors. It suffices to show that the scheme-theoretic intersection (XΛ1gen)κ(XΛ2gen)κ(X^{\operatorname{gen}}_{\Lambda_{1}})_{\kappa}\cap(X^{\operatorname{gen}}_{\Lambda_{2}})_{\kappa} is regular. For each i{1,2}i\in\{1,2\}, let DiD_{i} be the Cartier divisor with Λi|Di|\Lambda_{i}\subset|D_{i}| and let ViH0(X,𝒪X(Di))V_{i}\subset H^{0}(X,\mathcal{O}_{X}(D_{i})) be the kk-vector subspace corresponding to Λi\Lambda_{i}. Consider the restriction map:

ρ:H0(X×kκ1,𝒪X×kκ1(D2×kκ1))H0(XΛ1gen,𝒪X×kκ1(D2×kκ1)|XΛ1gen).\rho:H^{0}(X\times_{k}\kappa_{1},\mathcal{O}_{X\times_{k}\kappa_{1}}(D_{2}\times_{k}\kappa_{1}))\to H^{0}(X^{\operatorname{gen}}_{\Lambda_{1}},\mathcal{O}_{X\times_{k}\kappa_{1}}(D_{2}\times_{k}\kappa_{1})|_{X^{\operatorname{gen}}_{\Lambda_{1}}}).

Then the generic member (XΛ2gen)κ(X^{\operatorname{gen}}_{\Lambda_{2}})_{\kappa} coincides with the base change of the generic member of V2×kκH0(X×kκ1,𝒪X×kκ1(D2×kκ1))V_{2}\times_{k}\kappa\subset H^{0}(X\times_{k}\kappa_{1},\mathcal{O}_{X\times_{k}\kappa_{1}}(D_{2}\times_{k}\kappa_{1})). Therefore, the intersection (XΛ1gen)κ(XΛ2gen)κ(X^{\operatorname{gen}}_{\Lambda_{1}})_{\kappa}\cap(X^{\operatorname{gen}}_{\Lambda_{2}})_{\kappa} coincides with the generic member of Image(ρ){\rm Image}(\rho) by [Tan-Bertini, Proposition 5.10(2)], which is regular [Tan-Bertini, Theorem 4.9(4)]. ∎

2.4. FF-splitting criteria

Definition 2.5.

Let XX be a normal variety and let Δ\Delta be an effective \mathbb{Q}-divisor on XX.

  1. (1)

    We say that (X,Δ)(X,\Delta) is FF-split if

    𝒪XFeFe𝒪XFe𝒪X((pe1)Δ)\mathcal{O}_{X}\xrightarrow{F^{e}}F_{*}^{e}\mathcal{O}_{X}\hookrightarrow F_{*}^{e}\mathcal{O}_{X}(\lfloor(p^{e}-1)\Delta\rfloor)

    splits as an 𝒪X\mathcal{O}_{X}-module homomorphism for every e>0e\in\mathbb{Z}_{>0}.

  2. (2)

    We say that (X,Δ)(X,\Delta) is sharply FF-split if

    𝒪XFeFe𝒪XFe𝒪X((pe1)Δ)\mathcal{O}_{X}\xrightarrow{F^{e}}F_{*}^{e}\mathcal{O}_{X}\hookrightarrow F_{*}^{e}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil)

    splits as an 𝒪X\mathcal{O}_{X}-module homomorphism for some e>0e\in\mathbb{Z}_{>0}.

  3. (3)

    We say that (X,Δ)(X,\Delta) is globally FF-regular if, given an effective \mathbb{Z}-divisor EE, there exists e>0e\in\mathbb{Z}_{>0} such that

    𝒪XFeFe𝒪XFe𝒪X((pe1)Δ+E)\mathcal{O}_{X}\xrightarrow{F^{e}}F_{*}^{e}\mathcal{O}_{X}\hookrightarrow F_{*}^{e}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil+E)

    splits as an 𝒪X\mathcal{O}_{X}-module homomorphism.

We say that XX is FF-split (resp. globally FF-regular) if so is (X,0)(X,0).

Remark 2.6.

We have the following implications:

globally F-regularsharply F-splitF-split\text{globally $F$-regular}\Longrightarrow\text{sharply $F$-split}\Longrightarrow\text{$F$-split}

where the former implication is clear and the latter one holds by the same argument as in [Sch08g]*Proposition 3.3. Moreover, if the condition (\star) holds, then (X,Δ)(X,\Delta) is sharply F-split if and only if (X,Δ)(X,\Delta) is FF-split.

  1. (\star)

    (pe1)Δ(p^{e}-1)\Delta is a \mathbb{Z}-divisor for some e>0e\in\mathbb{Z}_{>0}

In particular, XX is FF-split if and only if F:𝒪XF𝒪XF\colon\mathcal{O}_{X}\to F_{*}\mathcal{O}_{X} splits as an 𝒪X\mathcal{O}_{X}-module homomorphism. For more foundational properties, we refer to [SS10].

In what follows, we summarise some FF-splitting criteria, which are well known to experts.

Proposition 2.7.

Let f:XYf\colon X\to Y be a birational morphism of normal projective varieties. Take an effective \mathbb{Q}-divisor ΔY\Delta_{Y} on YY such that (pe1)(KY+ΔY)(p^{e}-1)(K_{Y}+\Delta_{Y}) is Cartier for some e>0e\in\mathbb{Z}_{>0}. Assume that the \mathbb{Q}-divisor Δ\Delta defined by KX+Δ=f(KY+ΔY)K_{X}+\Delta=f^{*}(K_{Y}+\Delta_{Y}) is effective. Then (X,Δ)(X,\Delta) is FF-split if and only if (Y,ΔY)(Y,\Delta_{Y}) is FF-split.

Proof.

If (X,Δ)(X,\Delta) is FF-split, then so is (Y,ΔY)(Y,\Delta_{Y}) (take the pushforward). As for the opposite implication, the same argument as [HX15]*the first paragraph of the proof of Proposition 2.11 works. ∎

Proposition 2.8.

Let κ\kappa be an FF-finite field of characteristic p>0p>0. Let XX be a normal Gorenstein projective variety over κ\kappa. Take a normal prime Cartier divisor SS and an effective \mathbb{Q}-Cartier \mathbb{Q}-divisor BB on XX such that SSuppBS\not\subset\operatorname{Supp}\,B. Assume that

  1. (1)

    (S,B|S)(S,B|_{S}) is FF-split, and

  2. (2)

    there exists e>0e\in\mathbb{Z}_{>0} such that (pe1)(KX+S+B)(p^{e}-1)(K_{X}+S+B) is Cartier and

    H1(X,𝒪X(S(pe1)(KX+S+B)))=0.H^{1}(X,\mathcal{O}_{X}(-S-(p^{e}-1)(K_{X}+S+B)))=0.

Then (X,S+B)(X,S+B) is FF-split.

Proof.

The same argument as in [CTW17, Lemma 2.7] works. ∎

Corollary 2.9.

Let κ\kappa be an FF-finite field of characteristic p>0p>0. Let XX be a normal Gorenstein projective variety over κ\kappa. Take a normal prime Cartier divisor SS such that SS is FF-split and (KX+S)-(K_{X}+S) is ample. Then (X,S)(X,S) is FF-split.

Proof.

By applying Proposition 2.8 with B=0B=0, it suffices to show that H1(X,𝒪X(S(pe1)(KX+S))))=0H^{1}(X,\mathcal{O}_{X}(-S-(p^{e}-1)(K_{X}+S))))=0 for e0e\gg 0, which follows from Serre vanishing. ∎

Remark 2.10.

In the setting of Corollary 2.9, even if SS is quasi-FF-split, XX is not necessarily quasi-FF-split [KTY, Example 7.7].

Corollary 2.11.

Let XX be a smooth Fano threefold over kk. Assume that there exist a field extension kκk\subset\kappa and effective divisors D1,D2,D3D_{1},D_{2},D_{3} on Xκ:=X×kκX_{\kappa}:=X\times_{k}\kappa such that

  1. (1)

    κ\kappa is an FF-finite field,

  2. (2)

    KXκ+D1+D2+D30K_{X_{\kappa}}+D_{1}+D_{2}+D_{3}\sim 0,

  3. (3)

    D1D_{1} is a normal prime divisor,

  4. (4)

    D1D2D_{1}\cap D_{2} is one-dimensional, smooth, and

  5. (5)

    D1D2D3D_{1}\cap D_{2}\cap D_{3} is non-empty, zero-dimensional, and smooth over κ\kappa, and

  6. (6)

    H1(Xκ,𝒪Xκ(D1))=H1(Xκ,𝒪Xκ(D2))=H1(Xκ,𝒪Xκ(D3))=0H^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D_{1}))=H^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D_{2}))=H^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D_{3}))=0 (cf. Lemma 2.12).

Then (X×kκ,D1+D2+D3)(X\times_{k}\kappa,D_{1}+D_{2}+D_{3}) is FF-split. In particular, XX is FF-split.

Proof.

First of all, we show that

(2.11.1) H1(D1,𝒪D1(D2|D1))=0.H^{1}(D_{1},\mathcal{O}_{D_{1}}(-D_{2}|_{D_{1}}))=0.

To this end, it suffices to prove H1(Xκ,𝒪X(D2))=0H^{1}(X_{\kappa},\mathcal{O}_{X}(-D_{2}))=0 and H2(Xκ,𝒪Xκ(D2D1))=0H^{2}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D_{2}-D_{1}))=0. The former one follows from (6). The latter one holds by

h2(Xκ,𝒪Xκ(D2D1))=h1(Xκ,𝒪Xκ(KX+D1+D2))=(2)h1(Xκ,𝒪Xκ(D3))=(6)0.h^{2}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D_{2}-D_{1}))=h^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(K_{X}+D_{1}+D_{2}))\overset{{\rm(2)}}{=}h^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D_{3}))\overset{{\rm(6)}}{=}0.

This completes the proof of (2.11.1).

By H1(Xκ,𝒪Xκ(D1))=(6)0H^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D_{1}))\overset{{\rm(6)}}{=}0 and (2.11.1), we obtain

κ=H0(Xκ,𝒪Xκ)H0(D1,𝒪D1)H0(D1D2,𝒪D1D2).\kappa=H^{0}(X_{\kappa},\mathcal{O}_{X_{\kappa}})\xrightarrow{\simeq}H^{0}(D_{1},\mathcal{O}_{D_{1}})\xrightarrow{\simeq}H^{0}(D_{1}\cap D_{2},\mathcal{O}_{D_{1}\cap D_{2}}).

Therefore, C:=D1D2C:=D_{1}\cap D_{2} is a smooth projective curve. By (4) and (5), CC and CD3C\cap D_{3} are smooth over κ\kappa. In particular, (C×κκ¯,(CD3)×κκ¯)(κ¯1,P+Q)(C\times_{\kappa}\overline{\kappa},(C\cap D_{3})\times_{\kappa}\overline{\kappa})\simeq(\mathbb{P}^{1}_{\overline{\kappa}},P+Q) for the algebraic closure κ¯\overline{\kappa} of κ\kappa and some distinct points PP and QQ. Then (C,CD3)(C,C\cap D_{3}) is FF-split.

By Proposition 2.8, (D1,(D2+D3)|D1)(D_{1},(D_{2}+D_{3})|_{D_{1}}) is FF-split, where Proposition 2.8 is applicable by (2.11.1). Again by Proposition 2.8 and (6), (X×kκ,D1+D2+D3)(X\times_{k}\kappa,D_{1}+D_{2}+D_{3}) is FF-split. ∎

Lemma 2.12.

Let XX be a smooth Fano threefold over kk and let kκk\subset\kappa be a field extension. Take a divisor on XκX×kκX_{\kappa}\coloneqq X\times_{k}\kappa. Assume that one of the following conditions hold.

  1. (1)

    DD is nef and ν(Xκ,D)2\nu(X_{\kappa},D)\geq 2.

  2. (2)

    There exists a morphism π:Xκ1\pi\colon X\to\mathbb{P}^{1}_{\kappa} such that π𝒪X=𝒪κ1\pi_{*}\mathcal{O}_{X}=\mathcal{O}_{\mathbb{P}^{1}_{\kappa}} and 𝒪Xκ(D)π𝒪1(1)\mathcal{O}_{X_{\kappa}}(D)\simeq\pi^{*}\mathcal{O}_{\mathbb{P}^{1}}(1).

Then H1(Xκ,𝒪Xκ(D))=0H^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D))=0.

Proof.

Taking the base change to the algebraic closure κ¯\overline{\kappa} of κ\kappa, we may assume that κ\kappa is algebraically closed. Then the problem is reduced to the case when k=κk=\kappa. If (1) holds, then we may apply [Kaw1, Corollary 3.6]. If (2) holds, then we may assume that DD is a general fibre of π:X1\pi\colon X\to\mathbb{P}^{1}, which is a prime divisor, and hence H1(X,𝒪X)=0H^{1}(X,\mathcal{O}_{X})=0 implies H1(X,𝒪X(D))=0H^{1}(X,\mathcal{O}_{X}(-D))=0. ∎

Remark 2.13.

Let κ\kappa be an FF-finite field and let CC be a Gorenstein projective curve over κ\kappa. If KC-K_{C} is ample and there exists a Cartier divisor DD satisfying degD=1\deg D=1, then Cκ1C\simeq\mathbb{P}^{1}_{\kappa} [Kol13, Lemma 10.6], and hence CC is FF-split.

Example 2.14.

If XX is a normal toric variety and DD is a torus-invariant reduced divisor, then (X,D)(X,D) is FF-split. This follows essentially from [Fuj07, 2.6] (cf. [Tan-toric, Proposition 2.17]).

Proposition 2.15.

Let XX be a smooth Fano threefold. Suppose that the following condition.

  1. (1)

    H0(X,ΩX2(pKX))=0H^{0}(X,\Omega_{X}^{2}(pK_{X}))=0.

  2. (2)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega_{X}^{1}(K_{X}))=0.

  3. (3)

    H2(X,ΩX1(pKX))=0H^{2}(X,\Omega_{X}^{1}(-pK_{X}))=0.

  4. (4)

    H1(X,ΩX2(pKX))=0H^{1}(X,\Omega_{X}^{2}(-pK_{X}))=0.

Then XX is FF-split.

Proof.

Firstly, the global FF-splitting of XX (i.e., splitting of F:𝒪XF𝒪XF\colon\mathcal{O}_{X}\to F_{*}\mathcal{O}_{X}) is equivalent to the surjectivity of the evaluation map

Hom𝒪X(F𝒪X,𝒪X)FHom𝒪X(𝒪X,𝒪X)(H0(X,𝒪X)).\mathrm{Hom}_{\mathcal{O}_{X}}(F_{*}\mathcal{O}_{X},\mathcal{O}_{X})\xrightarrow{F^{*}}\mathrm{Hom}_{\mathcal{O}_{X}}(\mathcal{O}_{X},\mathcal{O}_{X})(\simeq H^{0}(X,\mathcal{O}_{X})).

By Grothendieck duality, they are equivalent to the injectivity of

F:H3(X,ωX)H3(X,F𝒪XωX).F:H^{3}(X,\omega_{X})\to H^{3}(X,F_{*}\mathcal{O}_{X}\otimes\omega_{X}).

Thus it suffices to show that H2(X,B1ΩX1(pKX))=0H^{2}(X,B_{1}\Omega_{X}^{1}(pK_{X}))=0. By (2.2.1) and (2), this vanishing can be reduced to H2(X,Z1ΩX1(pKX))H^{2}(X,Z_{1}\Omega^{1}_{X}(pK_{X})). By (2.2.2), it is enough to prove

H1(X,B1ΩX2(pKX))=0andH2(X,ΩX1(pKX))=0.H^{1}(X,B_{1}\Omega^{2}_{X}(pK_{X}))=0\,\,\,\text{and}\,\,\,H^{2}(X,\Omega^{1}_{X}(pK_{X}))=0.

The latter one follows from (4) and Serre duality. It suffices to show the first one. By (2.2.1) and the condition (1), this vanishing is reduced to

H1(X,Z1ΩX2(pKX))=0.H^{1}(X,Z_{1}\Omega_{X}^{2}(pK_{X}))=0.

By 2.2.2, it suffices to show

H0(X,B1ΩX3(pKX))=0andH1(X,ΩX2(pKX))=0.H^{0}(X,B_{1}\Omega^{3}_{X}(pK_{X}))=0\,\,\,\text{and}\,\,\,H^{1}(X,\Omega^{2}_{X}(pK_{X}))=0.

The first vanishing holds by B1ΩX3(pKX)FωXp+1B_{1}\Omega_{X}^{3}(pK_{X})\subset F_{*}\omega^{p+1}_{X} and the ampleness of ωX1\omega^{-1}_{X}. The latter one follows from (3) and Serre duality. ∎

2.5. Quasi-FF-splitting criteria

We recall that definition of FF-splitting and quasi-FF-splitting. We refer to [Kawakami-Tanaka(dPvar), Section 3] for details.

Definition 2.16.

Let XX be a normal variety. We define a Wn𝒪XW_{n}\mathcal{O}_{X}-module QX,neQ^{e}_{X,n} and a Wn𝒪XW_{n}\mathcal{O}_{X}-module homomorphism ΦX,Δ,nΔ,e\Phi^{\Delta,e}_{X,\Delta,n} by the following pushout diagram:

Wn𝒪X{W_{n}\mathcal{O}_{X}}FeWn𝒪X{F_{*}^{e}W_{n}\mathcal{O}_{X}}𝒪X{\mathcal{O}_{X}}QX,ne.{Q^{e}_{X,n}.}Fe\scriptstyle{F^{e}}Rn1\scriptstyle{R^{n-1}}ΦX,ne\scriptstyle{\Phi^{e}_{X,n}}

Applying ()omWn𝒪X(,WnωX(KX))(-)^{*}\coloneqq\mathcal{H}om_{W_{n}\mathcal{O}_{X}}(-,W_{n}\omega_{X}(-K_{X})) to ΦX,ne\Phi^{e}_{X,n}, we have a Wn𝒪XW_{n}\mathcal{O}_{X}-module homomorphism

(ΦX,ne):(QX,ne)𝒪X.(\Phi^{e}_{X,n})^{*}\colon(Q^{e}_{X,n})^{*}\to\mathcal{O}_{X}.

We say that XX is quasi-FeF^{e}-split if

H0(X,(ΦX,ne)):H0(X,(QX,ne))H0(X,𝒪X)H^{0}(X,(\Phi^{e}_{X,n})^{*})\colon H^{0}(X,(Q^{e}_{X,n})^{*})\to H^{0}(X,\mathcal{O}_{X})

is surjective.

Remark 2.17.

Let XX be a normal variety. Then XX is FF-split if and only if

H0(X,F):H0(X,(F𝒪X))H0(X,𝒪X)H^{0}(X,F^{*})\colon H^{0}(X,(F_{*}\mathcal{O}_{X})^{*})\to H^{0}(X,\mathcal{O}_{X})

is surjective for every e>0e>0, where ()om𝒪X(,W1ωX(KX))=om𝒪X(,𝒪X)(-)^{*}\coloneqq\mathcal{H}om_{\mathcal{O}_{X}}(-,W_{1}\omega_{X}(-K_{X}))=\mathcal{H}om_{\mathcal{O}_{X}}(-,\mathcal{O}_{X}). In particular, if XX is FF-split, then XX is quasi-FeF^{e}-split for every e>0e>0.

Corollary 2.18.

Let kk be an algebraically closed field of characteristic p>0p>0 and let kκk\subset\kappa be a field extension, where κ\kappa is FF-finite. Let XX be a smooth Fano threefold over kk. Assume that there exist a prime divisor DD and a reduced divisor DD^{\prime} on X×kκX\times_{k}\kappa such that

  1. (a)

    KX×kκ+D+D0K_{X\times_{k}\kappa}+D+D^{\prime}\sim 0,

  2. (b)

    DD is regular, DDD\cap D^{\prime} is smooth over κ\kappa, DSuppDD\not\subset\operatorname{Supp}D^{\prime},

  3. (c)

    H1(X×kκ,𝒪X×kκ(D))=0H^{1}(X\times_{k}\kappa,\mathcal{O}_{X\times_{k}\kappa}(-D))=0, and

  4. (d)

    H1(X×kκ,𝒪X×kκ(D))=0H^{1}(X\times_{k}\kappa,\mathcal{O}_{X\times_{k}\kappa}(-D^{\prime}))=0.

Then XX is 22-quasi-FeF^{e}-split for all e>0e>0.

Proof.

Set XκX×kκX_{\kappa}\coloneqq X\times_{k}\kappa. We show that (X×kκ,D1++Dn1)(X\times_{k}\kappa,D_{1}+\cdots+D_{n-1}) is weakly 22-quasi-FF-split (see [Kawakami-Tanaka(dPvar), Definition 5.13] for the definition of weak quasi-FF-splitting).

Since D+DD+D^{\prime} is ample, D+DD+D^{\prime} is connected, and hence D|D=DDD^{\prime}|_{D}=D\cap D^{\prime}\neq\emptyset. By [Kawakami-Tanaka(dPvar), Corollary 5.17], it is enough to show that (D,D|D)(D,D^{\prime}|_{D}) is weakly 22-quasi-FF-split. Again by [Kawakami-Tanaka(dPvar), Corollary 5.17], it suffices to prove that

  1. (i)

    D|DD^{\prime}|_{D} is a prime divisor,

  2. (ii)

    H1(D,𝒪D(D|D))=0H^{1}(D,\mathcal{O}_{D}(-D^{\prime}|_{D}))=0, and

  3. (iii)

    D|DD^{\prime}|_{D} is 22-quasi-FF-split.

Consider an exact sequence

0𝒪Xκ(DD)𝒪Xκ(D)𝒪D(D|D)0,0\to\mathcal{O}_{X_{\kappa}}(-D-D^{\prime})\to\mathcal{O}_{X_{\kappa}}(-D^{\prime})\to\mathcal{O}_{D}(-D^{\prime}|_{D})\to 0,

which induces the following one:

0=(d)H1(Xκ,𝒪Xκ(D))H1(D,𝒪D(D|D))H2(Xκ,𝒪Xκ(DD)).0\overset{{\rm(d)}}{=}H^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D^{\prime}))\to H^{1}(D,\mathcal{O}_{D}(-D^{\prime}|_{D}))\to H^{2}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D-D^{\prime})).

Then (ii) follows from

H2(Xκ,𝒪Xκ(DD))H2(Xκ,𝒪Xκ(KXκ))H2(X,𝒪X(KX))kκ=0.H^{2}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-D-D^{\prime}))\simeq H^{2}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(K_{X_{\kappa}}))\simeq H^{2}(X,\mathcal{O}_{X}(K_{X}))\otimes_{k}\kappa=0.

By the induced exact sequence

H0(D,𝒪D)H0(D|D,𝒪D|D)H1(D,𝒪D(D|D))=(ii)0,H^{0}(D,\mathcal{O}_{D})\to H^{0}(D^{\prime}|_{D},\mathcal{O}_{D^{\prime}|_{D}})\to H^{1}(D,\mathcal{O}_{D}(-D^{\prime}|_{D}))\overset{{\rm(ii)}}{=}0,

H0(D|D,𝒪D|D)H^{0}(D^{\prime}|_{D},\mathcal{O}_{D^{\prime}|_{D}}) is a field. Since D|D=DDD^{\prime}|_{D}=D\cap D^{\prime} is smooth, we obtain (i). Let us show (iii). Since D|DD^{\prime}|_{D} is a smooth projective curve over κ\kappa with KD|D0K_{D^{\prime}|_{D}}\sim 0, the base change (D|D)×κκ¯(D^{\prime}|_{D})\times_{\kappa}\overline{\kappa} to the algebraic closure κ¯\overline{\kappa} of κ\kappa is an elliptic curve, which is 2-quasi-FF-split ([KTTWYY1, Remark 2.11]). Thus so is D|DD^{\prime}|_{D} [KTY, Proposition 2.12], and (iii) holds.

Now, we show that XX is quasi-FF-split. Since (X×kκ,D1++Dn1)(X\times_{k}\kappa,D_{1}+\cdots+D_{n-1}) is weakly 22-quasi-FF-split it follows that X×kκX\times_{k}\kappa is 22-quasi-FF-split. By [KTY, Proposition 2.12], XX is 22-quasi-FF-split. ∎

Remark 2.19.

The assumptions (c) and (d) automatically hold when k=κk=\kappa and DD^{\prime} is connected. Indeed, we have an exact sequence H0(X,𝒪X)H0(E,𝒪E)H1(X,𝒪X(E))H1(X,𝒪X)=0H^{0}(X,\mathcal{O}_{X})\xrightarrow{\simeq}H^{0}(E,\mathcal{O}_{E})\to H^{1}(X,\mathcal{O}_{X}(-E))\to H^{1}(X,\mathcal{O}_{X})=0 for any reduced connected divisor EE on XX.

Although the following result is contained in [KTTWYY1, Theorem F], we include a proof for the reader’s convenience, as its proof is quite short.

Proposition 2.20.

Let XX be a smooth Fano threefold. Suppose that the following condition.

  1. (1)

    H0(X,ΩX2(piKX))=0H^{0}(X,\Omega_{X}^{2}(p^{i}K_{X}))=0 for all i>0i>0.

  2. (2)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega_{X}^{1}(K_{X}))=0.

  3. (3)

    H2(X,ΩX1(piKX))=0H^{2}(X,\Omega_{X}^{1}(-p^{i}K_{X}))=0 for all i>0i>0.

Then XX is quasi-FF-split.

Proof.

It is enough to show that H2(X,BnΩX1(pnKX))=0H^{2}(X,B_{n}\Omega_{X}^{1}(p^{n}K_{X}))=0 for some n>0n>0. By (2.1.1) and (2), it suffices to prove H2(X,ZnΩX1(pnKX))=0H^{2}(X,Z_{n}\Omega^{1}_{X}(p^{n}K_{X}))=0 for some n>0n>0. By (2.1.2), this vanishing is reduced to

H1(X,B1ΩX2(pKX))=0andH2(X,Zn1ΩX1(pnKX))=0.H^{1}(X,B_{1}\Omega^{2}_{X}(pK_{X}))=0\,\,\,\text{and}\,\,\,H^{2}(X,Z_{n-1}\Omega^{1}_{X}(p^{n}K_{X}))=0.

Repeating this procedure, the vanishing H2(X,ZnΩX1(pnKX))=0H^{2}(X,Z_{n}\Omega^{1}_{X}(p^{n}K_{X}))=0 is reduced to

H1(X,B1ΩX2(plKX))=0andH2(X,ΩX1(pnKX))=0.H^{1}(X,B_{1}\Omega^{2}_{X}(p^{l}K_{X}))=0\,\,\,\text{and}\,\,\,H^{2}(X,\Omega^{1}_{X}(p^{n}K_{X}))=0.

for every l{1,,n}l\in\{1,\ldots,n\}. Taking n0n\gg 0, we may assume H2(X,ΩX1(pnKX))=0H^{2}(X,\Omega^{1}_{X}(p^{n}K_{X}))=0 by Serre vanishing. By (2.1.1) and (1), it suffices to show that H1(X,Z1ΩX2(plKX))=0H^{1}(X,Z_{1}\Omega^{2}_{X}(p^{l}K_{X}))=0 for every l{1,,n}l\in\{1,\ldots,n\}. By Serre duality, we have

H1(X,ΩX2(plKX))H2(X,ΩX1(plKX))=(3)0H^{1}(X,\Omega_{X}^{2}(p^{l}K_{X}))\simeq H^{2}(X,\Omega_{X}^{1}(-p^{l}K_{X}))\overset{{\rm(3)}}{=}0

for every l>0l>0. By (2.1.2), the problem is finally reduced to H0(X,B1ΩX3(plKX))=0H^{0}(X,B_{1}\Omega_{X}^{3}(p^{l}K_{X}))=0 for all l{1,,n}l\in\{1,\ldots,n\}. This holds because B1ΩX3(plKX)FωXpl+1B_{1}\Omega_{X}^{3}(p^{l}K_{X})\subset F_{*}\omega^{p^{l}+1}_{X} and ωX1\omega^{-1}_{X} is ample. ∎

2.6. Weak del Pezzo surfaces

In this subsection, we recall when canonical weal del Pezzo surfaces are FF-split.

Definition 2.21.

Let SS be a normal Gorenstein projective surface.

  1. (1)

    We say that SS is weak del Pezzo if KS-K_{S} is nef and big.

  2. (2)

    We say that SS is del Pezzo if KS-K_{S} is ample.

Theorem 2.22.

Let SS be a canonical weak del Pezzo surface. Then SS is FF-split if p>5p>5 or KS25K_{S}^{2}\geq 5.

Proof.

See [KT, Theorem A]. ∎

3. Fano threefolds with ρ4\rho\geq 4

Proposition 3.1.

Let XX be a smooth Fano threefold with ρ(X)6\rho(X)\geq 6. Then XX is quasi-FF-split.

Proof.

In this case, XS×1X\simeq S\times\mathbb{P}^{1} for a smooth del Pezzo surface [FanoIV, Section 7.6]. Since SS is quasi-FF-split and 1\mathbb{P}^{1} is FF-split, S×1S\times\mathbb{P}^{1} is quasi-FF-split [KTY, Proposition 6.7]. ∎

Proposition 3.2.

Let XX be a smooth Fano threefold with ρ(X)=5\rho(X)=5. Then XX is FF-split.

Proof.

Recall that XX is of No. 5-1, 5-2, or 5-3 [FanoIV, Section 7.5]. If XX is 5-3, then XS×1X\simeq S\times\mathbb{P}^{1} for a smooth del Pezzo surface SS with ρ(S)=4\rho(S)=4, and hence FF-split (e.g. SS is toric, and hence so is XX).

Assume that XX is 5-1. Then X=BlB1B2B3YX=\operatorname{Bl}_{B_{1}\amalg B_{2}\amalg B_{3}}Y, where YBlCQY\coloneqq\operatorname{Bl}_{C}Q is a blowup of QQ along a conic CC and B1,B2,B3B_{1},B_{2},B_{3} are mutually distinct one-dimensional fibres of σ:YBlCQQ\sigma\colon Y\coloneqq\operatorname{Bl}_{C}Q\to Q [FanoIV, Section 7.5]. Since the smallest linear subvariety C\langle C\rangle of 4\mathbb{P}^{4} containing CC is a plane, we obtain CC=H¯1H¯2C\subset\langle C\rangle=\overline{H}_{1}\cap\overline{H}_{2} for suitable hyperplanes H¯1\overline{H}_{1} and H¯2\overline{H}_{2} on 4\mathbb{P}^{4}. Set HiQH¯iH_{i}\coloneqq Q\cap\overline{H}_{i} for each i{1,2}i\in\{1,2\}. We then get a scheme-theoretic equality C=H1H2C=H_{1}\cap H_{2}. Note that each HiH_{i} is a (possibly singular) quadric surface in H¯i=3\overline{H}_{i}=\mathbb{P}^{3}, which is smooth along C=H1H2C=H_{1}\cap H_{2}. It is easy to see that Δ\Delta is effective for the divisor Δ\Delta defined by h(KQ+H1+H2)=KX+Δh^{*}(K_{Q}+H_{1}+H_{2})=K_{X}+\Delta, where h:XQh\colon X\to Q denotes the induced birational morphism. It is enough to show that (Q,H1+H2)(Q,H_{1}+H_{2}) is FF-split (Proposition 2.7). Take a general hyperplane section H3H_{3}. Since KQ+H1+H2+H30K_{Q}+H_{1}+H_{2}+H_{3}\sim 0 and all H3,H3H2H_{3},H_{3}\cap H_{2}, and H3H2H1H_{3}\cap H_{2}\cap H_{1} are smooth, (Q,H1+H2+H3)(Q,H_{1}+H_{2}+H_{3}) is FF-split (Corollary 2.11), and hence (Q,H1+H2)(Q,H_{1}+H_{2}) is FF-split.

Assume that XX is No. 5-2. Then X=BlBBYX=\operatorname{Bl}_{B\amalg B^{\prime}}Y, where YBlL1L23Y\coloneqq\operatorname{Bl}_{L_{1}\amalg L_{2}}\mathbb{P}^{3} is a blowup of 3\mathbb{P}^{3} along a disjoint union of lines L1L_{1} and L2L_{2}, and BB and BB^{\prime} are mutually distinct one-dimensional fibres of σ:Y=BlL1L233\sigma\colon Y=\operatorname{Bl}_{L_{1}\amalg L_{2}}\mathbb{P}^{3}\to\mathbb{P}^{3} lying over L1L_{1} [FanoIV, Section 7.5]. Take two planes H1H_{1} and H2H_{2} on 3\mathbb{P}^{3} such that H1H2=L1H_{1}\cap H_{2}=L_{1}. Note that each HiL2H_{i}\cap L_{2} is a smooth point. Pick a plane H3H_{3} containing L2L_{2}. Then it is easy to see that Δ\Delta is effective for the divisor Δ\Delta defined by h(K3+H1+H2+H3)=KX+Δh^{*}(K_{\mathbb{P}^{3}}+H_{1}+H_{2}+H_{3})=K_{X}+\Delta. It is enough to show that (3,H1+H2+H3)(\mathbb{P}^{3},H_{1}+H_{2}+H_{3}) is FF-split (Proposition 2.7). Pick a general hyperplane H4H_{4}. Apply Corollary 2.11 by setting D1H1,D2H2D_{1}\coloneqq H_{1},D_{2}\coloneqq H_{2}, and D3H3+H4D_{3}\coloneqq H_{3}+H_{4}. Then (X,D1+D2+D3)(X,D_{1}+D_{2}+D_{3}) is FF-split, and hence (3,H1+H2+H3)(\mathbb{P}^{3},H_{1}+H_{2}+H_{3}) is FF-split. ∎

Proposition 3.3.

Let XX be a smooth Fano threefold with ρ(X)=4\rho(X)=4. Then XX is FF-split.

Proof.

We treat the following six cases separately:

  1. (1)

    4-4, 4-10, 4-12.

  2. (2)

    4-3, 4-6, 4-8, 4-13.

  3. (3)

    4-5, 4-7, 4-9, 4-11.

  4. (4)

    4-11.

  5. (5)

    4-2.

  6. (6)

    4-1.

(1) If XX is 4-4, then there is a smooth curve BB on XX such that the blowup X~\widetilde{X} of XX along BB is Fano [FanoIV, Proposition 5.31]. Since X~\widetilde{X} is FF-split (Proposition 3.2), so is XX. If XX is 4-10, then we can write XS×1X\simeq S\times\mathbb{P}^{1} for a smooth del Pezzo surface SS with KS2=7K_{S}^{2}=7 [FanoIV, Section 7.4]. In this case, XX is clearly FF-split. Assume that XX is 4-12. Then we have X=BlBBY2-33X=\operatorname{Bl}_{B\amalg B^{\prime}}Y_{\text{2-33}}, where Y2-33=BlL3Y_{\text{2-33}}=\operatorname{Bl}_{L}\mathbb{P}^{3} is the blowup of 3\mathbb{P}^{3} along a line LL, and BB and BB^{\prime} are mutually disjoint one-dimensional fibres of the induced blowup ρ:Y2-33=BlL33\rho:Y_{\text{2-33}}=\operatorname{Bl}_{L}\mathbb{P}^{3}\to\mathbb{P}^{3}. In this case, we can apply a similar argument to that of 5-2 in the proof of Proposition 3.2.

(2) Assume that XX is one of 4-3, 4-6, 4-8, 4-13. In this case, there is a blowup h:X11×21×31h:X\to\mathbb{P}^{1}_{1}\times\mathbb{P}^{1}_{2}\times\mathbb{P}^{1}_{3} along a curve BB of tridegree (1,1,c)(1,1,c) for some c0c\geq 0 [FanoIV, Section 7.4]. Let B11×21B^{\prime}\subset\mathbb{P}^{1}_{1}\times\mathbb{P}^{1}_{2} be the image of BB, which is a curve of bidegree (1,1)(1,1). Then the induced morphism B1B^{\prime}\to\mathbb{P}^{1} to the first direct product factor is an isomorphism. Hence B1B^{\prime}\simeq\mathbb{P}^{1}. Set D1×1×1D\subset\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} to be the inverse image of BB^{\prime}, which satisfies D|𝒪1×1×1(1,1,0)|D\in|\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1,0)| and D1×1D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}. Then it is easy to see that Δ\Delta is effective for the divisor Δ\Delta defined by h(K1×1×1+D)=KX+Δh^{*}(K_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}+D)=K_{X}+\Delta. It is enough to show that (1×1×1,D)(\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1},D) is FF-split (Proposition 2.7), which follows from the fact that DD is FF-split and (K1×1×1+D)-(K_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}+D) is ample (Corollary 2.9).

(3) Assume that XX is one of 4-5, 4-7, 4-9. Let τ:𝔽12\tau:\mathbb{F}_{1}\to\mathbb{P}^{2} be the blowdown of the (1)(-1)-curve Γ\Gamma and set P:=τ(Γ)P:=\tau(\Gamma). Then there is a blowup f:XY=𝔽1×1f:X\to Y=\mathbb{F}_{1}\times\mathbb{P}^{1} along a smooth curve BB lying over B𝔽1B^{\prime}\subset\mathbb{F}_{1}, where BB^{\prime} is disjoint from the (1)(-1)-curve Γ\Gamma and BB^{\prime} is the inverse image of a line L2L\subset\mathbb{P}^{2} [FanoIV, Section 7.4]. Then the induced composite morphism

h:X𝔽1×12×1h:X\to\mathbb{F}_{1}\times\mathbb{P}^{1}\to\mathbb{P}^{2}\times\mathbb{P}^{1}

is the blowup along (P×1)B¯(P\times\mathbb{P}^{1})\amalg\overline{B}, where B¯2×1\overline{B}\subset\mathbb{P}^{2}\times\mathbb{P}^{1} denotes the image of B𝔽1×1B\subset\mathbb{F}_{1}\times\mathbb{P}^{1}. In particular, we get (P×1)B¯L×1L×1(P\times\mathbb{P}^{1})\amalg\overline{B}\subset L^{\prime}\times\mathbb{P}^{1}\cup L\times\mathbb{P}^{1} for a line LL^{\prime} passing through PP. Then it is easy to see that Δ\Delta is effective for the divisor Δ\Delta defined by h(K2×1+L×1+L×1)=KX+Δh^{*}(K_{\mathbb{P}^{2}\times\mathbb{P}^{1}}+L\times\mathbb{P}^{1}+L^{\prime}\times\mathbb{P}^{1})=K_{X}+\Delta. It is enough to show that (2×1,L×1+L×1)(\mathbb{P}^{2}\times\mathbb{P}^{1},L\times\mathbb{P}^{1}+L^{\prime}\times\mathbb{P}^{1}) is FF-split (Proposition 2.7). This holds, because we may assume that L×1+L×1L\times\mathbb{P}^{1}+L^{\prime}\times\mathbb{P}^{1} is a torus-invariant reduced divisor on a toric variety 2×1\mathbb{P}^{2}\times\mathbb{P}^{1}.

(4) Assume that XX is 4-11. Then there is a blowup f:X𝔽1×1f\colon X\to\mathbb{F}_{1}\times\mathbb{P}^{1} along CC, where C=Γ×{t}C=\Gamma\times\{t\} for the (1)(-1)-curve Γ\Gamma on 𝔽1\mathbb{F}_{1} and a closed point tt of 1\mathbb{P}^{1} [FanoIV, Section 7.4]. For the blowup τ×id:𝔽1×12×1\tau\times{\rm id}\colon\mathbb{F}_{1}\times\mathbb{P}^{1}\to\mathbb{P}^{2}\times\mathbb{P}^{1}, consider the composite birational morphism:

h:X𝑓𝔽1×1τ×id2×1.h\colon X\xrightarrow{f}\mathbb{F}_{1}\times\mathbb{P}^{1}\xrightarrow{\tau\times{\rm id}}\mathbb{P}^{2}\times\mathbb{P}^{1}.

For the blowup centre Pτ(Γ)P\coloneqq\tau(\Gamma) of τ:𝔽12\tau\colon\mathbb{F}_{1}\to\mathbb{P}^{2}, pick two lines LL and LL^{\prime} on 2\mathbb{P}^{2} passing through PP. Then we have

KY+(Γ×1)+LY+LY=(τ×id)(K2×1+(L×1)+(L×1)),K_{Y}+(\Gamma\times\mathbb{P}^{1})+L_{Y}+L^{\prime}_{Y}=(\tau\times{\rm id})^{*}(K_{\mathbb{P}^{2}\times\mathbb{P}^{1}}+(L\times\mathbb{P}^{1})+(L^{\prime}\times\mathbb{P}^{1})),

where LYL_{Y} and LYL^{\prime}_{Y} denote the proper transforms of L×1L\times\mathbb{P}^{1} and L×1L^{\prime}\times\mathbb{P}^{1}, respectively. Since C=Γ×{t}C=\Gamma\times\{t\} is contained in Γ×1\Gamma\times\mathbb{P}^{1}, we see that the divisor Δ\Delta defined by KX+Δ=h(K2×1+(L×1)+(L×1))K_{X}+\Delta=h^{*}(K_{\mathbb{P}^{2}\times\mathbb{P}^{1}}+(L\times\mathbb{P}^{1})+(L^{\prime}\times\mathbb{P}^{1})) is effective. Then XX is FF-split, because (2×1,L×1+L×1)(\mathbb{P}^{2}\times\mathbb{P}^{1},L\times\mathbb{P}^{1}+L^{\prime}\times\mathbb{P}^{1}) is FF-split (Proposition 2.7).

(5) Assume that XX is 4-2. Then XX is a blowup of Y=1×1(𝒪𝒪(1,1))Y=\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}\oplus\mathcal{O}(1,1)) along an elliptic curve BB on a section SS of the 1\mathbb{P}^{1}-bundle π:Y1×1\pi:Y\to\mathbb{P}^{1}\times\mathbb{P}^{1} disjoint from the negative section SS^{\prime} of π\pi [FanoIV, Proposition 5.28]. Note that SS1×1S\simeq S^{\prime}\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}. We have that KY+S+SπK1×1K_{Y}+S+S^{\prime}\sim\pi^{*}K_{\mathbb{P}^{1}\times\mathbb{P}^{1}}. Indeed, since KY+S+SK_{Y}+S+S^{\prime} is π\pi-numerically trivial, we can write KY+S+SπDK_{Y}+S+S^{\prime}\sim\pi^{*}D for some Cartier divisor DD on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}. By restricting to SS, we obtain DK1×1D\sim K_{\mathbb{P}^{1}\times\mathbb{P}^{1}}.

Then, the \mathbb{Q}-divisor

(KY+S+(1ϵ)S)=ϵS+πK1×1-(K_{Y}+S+(1-\epsilon)S^{\prime})=\epsilon S^{\prime}+\pi^{*}K_{\mathbb{P}^{1}\times\mathbb{P}^{1}}

is ample for 0<ϵ10<\epsilon\ll 1. After perturbing ϵ\epsilon, the problem is reduced to the case when (pe1)(KY+S+(1ϵ)S)(p^{e}-1)(K_{Y}+S+(1-\epsilon)S^{\prime}) is Cartier for some e>0e>0. Replacing ee by some ee>0e^{\prime}\in e\mathbb{Z}_{>0}, we may assume that H1(X,𝒪X(S(pe1)(KY+S+(1ϵ)S)))=0H^{1}(X,\mathcal{O}_{X}(-S-(p^{e}-1)(K_{Y}+S+(1-\epsilon)S^{\prime})))=0. Since (S,(1ϵ)S|S)=(1×1,0)(S,(1-\epsilon)S^{\prime}|_{S})=(\mathbb{P}^{1}\times\mathbb{P}^{1},0) is sharply FF-split, (X,S+(1ϵ)S)(X,S+(1-\epsilon)S^{\prime}) is sharply FF-split (Proposition 2.8). Hence XX is FF-split.

(6) Assume that XX is 4-1. Then XX is a prime divisor on 11×21×31×41\mathbb{P}^{1}_{1}\times\mathbb{P}^{1}_{2}\times\mathbb{P}^{1}_{3}\times\mathbb{P}^{1}_{4} of multi-degree (1,1,1,1)(1,1,1,1) [FanoIV, Section 7.4]. For each i{1,2,3,4}i\in\{1,2,3,4\}, we set Hi:=πi𝒪i1(1)H_{i}:=\pi_{i}^{*}\mathcal{O}_{\mathbb{P}^{1}_{i}}(1) and Hi:=pri𝒪i1(1)H^{\prime}_{i}:=\operatorname{pr}_{i}^{*}\mathcal{O}_{\mathbb{P}^{1}_{i}}(1), where πi\pi_{i} and pri\operatorname{pr}_{i} denote the the induced morphisms:

πi:X11×21×31×41prii1.\pi_{i}:X\hookrightarrow\mathbb{P}^{1}_{1}\times\mathbb{P}^{1}_{2}\times\mathbb{P}^{1}_{3}\times\mathbb{P}^{1}_{4}\xrightarrow{\operatorname{pr}_{i}}\mathbb{P}^{1}_{i}.

It holds that KXH1+H2+H3+H4-K_{X}\sim H_{1}+H_{2}+H_{3}+H_{4}. Note that

H1H2H3=H1H2H3(H1+H2+H3+H4)=H1H2H3H4=1.H_{1}\cdot H_{2}\cdot H_{3}=H_{1}^{\prime}\cdot H^{\prime}_{2}\cdot H^{\prime}_{3}\cdot(H^{\prime}_{1}+H^{\prime}_{2}+H^{\prime}_{3}+H^{\prime}_{4})=H^{\prime}_{1}\cdot H^{\prime}_{2}\cdot H^{\prime}_{3}\cdot H^{\prime}_{4}=1.

Take the generic members H1gen,H2gen,H3gen,H4genH_{1}^{\operatorname{gen}},H_{2}^{\operatorname{gen}},H_{3}^{\operatorname{gen}},H_{4}^{\operatorname{gen}} of H1,H2,H3,H4H_{1},H_{2},H_{3},H_{4}, where each HigenH_{i}^{\operatorname{gen}} is an effective Cartier divisor on X×kκX\times_{k}\kappa for suitable purely transcendental field extension κ/k\kappa/k (Remark 2.4). Set D1:=H1gen,D2:=H2gen,D3:=H3gen+H4genD_{1}:=H_{1}^{\operatorname{gen}},D_{2}:=H_{2}^{\operatorname{gen}},D_{3}:=H_{3}^{\operatorname{gen}}+H_{4}^{\operatorname{gen}}. Then D1genD2genD_{1}^{\operatorname{gen}}\cap D_{2}^{\operatorname{gen}} is smooth, because deg(H3gen|D1D2)=1\deg(H_{3}^{\operatorname{gen}}|_{D_{1}\cap D_{2}})=1 (Remark 2.13). By Corollary 2.11, XX is FF-split. ∎

4. Fano threefolds with ρ=3\rho=3 (except for 3-10)

The purpose of this subsection is to prove that an arbitrary smooth Fano threefold XX with ρ(X)=3\rho(X)=3 is FF-split except for 3-10 (Proposition 4.4). We start with some complicated cases: 3-1, 3-3 and 3-4.

Lemma 4.1.

Let XX be a smooth Fano threefold of No. 3-1. Then the following hold.

  1. (1)

    Let φ:X1×1×1\varphi\colon X\to\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} be a finite double cover such that (φ𝒪X/𝒪Y)1𝒪1×1×1(1,1,1)(\varphi_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1,1). Then φ\varphi is separable, i.e., the induced field extension K(X)/K(Y)K(X)/K(Y) is separable.

  2. (2)

    XX is FF-split.

Proof.

Let us show (1). Set Y1×1×1Y\coloneqq\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} and 𝒪1×1×1(1,1,1)\mathcal{L}\coloneqq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1,1). By [CD89, Proposition 0.1.2] and [Ful98, Lemma 3.2], it suffices to show that

degc3(ΩY12)0.\deg c_{3}(\Omega_{Y}^{1}\otimes\mathcal{L}^{\otimes 2})\neq 0.

Set ipriΩ112\mathcal{M}_{i}\coloneqq\operatorname{pr}_{i}^{*}\Omega_{\mathbb{P}^{1}}^{1}\otimes\mathcal{L}^{\otimes 2} for every i{1,2,3}i\in\{1,2,3\}, i.e.,

1𝒪(0,2,2),2𝒪(2,0,2),3𝒪(2,2,0).\mathcal{M}_{1}\coloneqq\mathcal{O}(0,2,2),\qquad\mathcal{M}_{2}\coloneqq\mathcal{O}(2,0,2),\qquad\mathcal{M}_{3}\coloneqq\mathcal{O}(2,2,0).\qquad

We have ΩY12=123\Omega_{Y}^{1}\otimes\mathcal{L}^{\otimes 2}=\mathcal{M}_{1}\oplus\mathcal{M}_{2}\oplus\mathcal{M}_{3}. Then the following holds (cf. [Har77, Appendix A, Section 3]):

i=0ci(ΩY12)ti\displaystyle\sum_{i=0}^{\infty}c_{i}(\Omega_{Y}^{1}\otimes\mathcal{L}^{\otimes 2})t^{i} =\displaystyle= ct(ΩY12)\displaystyle c_{t}(\Omega_{Y}^{1}\otimes\mathcal{L}^{\otimes 2})
=\displaystyle= ct(1)ct(2)ct(3)\displaystyle c_{t}(\mathcal{M}_{1})c_{t}(\mathcal{M}_{2})c_{t}(\mathcal{M}_{3})
=\displaystyle= (1+c1(1)t)(1+c1(2)t)(1+c1(3)t).\displaystyle(1+c_{1}(\mathcal{M}_{1})t)(1+c_{1}(\mathcal{M}_{2})t)(1+c_{1}(\mathcal{M}_{3})t).

Therefore, we get

degc3(ΩY12)=deg(c1(1)c1(2)c1(3))=123>0,\deg c_{3}(\Omega_{Y}^{1}\otimes\mathcal{L}^{\otimes 2})=\deg(c_{1}(\mathcal{M}_{1})c_{1}(\mathcal{M}_{2})c_{1}(\mathcal{M}_{3}))=\mathcal{M}_{1}\cdot\mathcal{M}_{2}\cdot\mathcal{M}_{3}>0,

as required. Thus (1) holds.

Let us show (2). There is a finite separable double cover φ:XY=1×1×1\varphi\colon X\to Y=\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} as in (1) [FanoIII, Theorem 6.7]. Moreover, for each i{1,2,3}i\in\{1,2,3\}, the composition φi:X𝜑1×1×1pri1\varphi_{i}\colon X\xrightarrow{\varphi}\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}\xrightarrow{\operatorname{pr}_{i}}\mathbb{P}^{1} is the contraction of an extremal ray [FanoIII, Remark 6.8]. Since φ\varphi is separable, if H1,H2,H3H_{1},H_{2},H_{3} are general members, then the scheme-theoretic intersection H1H2H3H_{1}\cap H_{2}\cap H_{3} is reduced two points. Let H1gen,H2gen,H3genH_{1}^{\operatorname{gen}},H_{2}^{\operatorname{gen}},H_{3}^{\operatorname{gen}} be their generic members. Then the regular curve H1genH2genH_{1}^{\operatorname{gen}}\cap H_{2}^{\operatorname{gen}} is automatically smooth (Remark 2.13). Then XX is FF-split (Corollary 2.11, Lemma 2.12). ∎

Lemma 4.2.

Let YY be a smooth Fano threefold of No. 2-18. Let f1:Y1f_{1}\colon Y\to\mathbb{P}^{1} and f2:Y2f_{2}\colon Y\to\mathbb{P}^{2} be the contractions of the extremal rays. Take a general point P1P\in\mathbb{P}^{1} and a general line LL on 2\mathbb{P}^{2}. Then H0(Γ,𝒪Γ)=kH^{0}(\Gamma,\mathcal{O}_{\Gamma})=k for the scheme-theoretic intersection Γf11(P)f21(L)\Gamma\coloneqq f_{1}^{-1}(P)\cap f_{2}^{-1}(L).

Proof.

Set S1f11(P)S_{1}\coloneqq f_{1}^{-1}(P) and S2f21(L)S_{2}\coloneqq f_{2}^{-1}(L). By [FS20, Theorem 15.2] and [BT22, Theorem 3.3], S1S_{1} is a canonical del Pezzo surface. In particular, S1S_{1} is a rational surface. Since KYS1+2S2-K_{Y}\sim S_{1}+2S_{2} [FanoIII, Proposition 5.9 and Section 9.2], the divisor KY|YK=2S2|YK-K_{Y}|_{Y_{K}}=2S_{2}|_{Y_{K}} is ample for the generic fibre YKY_{K} of f1:Y1f_{1}\colon Y\to\mathbb{P}^{1}, where KK(1)K\coloneqq K(\mathbb{P}^{1}). Hence S2|S1(=Γ)S_{2}|_{S_{1}}(=\Gamma) is ample, as S1S_{1} is chosen to be a general fibre of f1f_{1}. Therefore, H1(S1,𝒪S1(Γ))=0H^{1}(S_{1},\mathcal{O}_{S_{1}}(-\Gamma))=0 by [CT19, Proposition 3.3] (or [Muk13, Theorem 3]), and hence k=H0(S1,𝒪S1)H0(Γ,𝒪Γ)k=H^{0}(S_{1},\mathcal{O}_{S_{1}})\xrightarrow{\simeq}H^{0}(\Gamma,\mathcal{O}_{\Gamma}). ∎

Lemma 4.3.

Let XX be a smooth Fano threefold of No. 3-3 or 3-4. Then XX is FF-split.

Proof.

For each case, XX has exactly three extremal rays and there is a conic bundle structure f:X1×1f\colon X\to\mathbb{P}^{1}\times\mathbb{P}^{1} [FanoIV, Section 7.3]. In what follows, we shall use their properties obtained in [FanoIV, Propositions 4.33 and 4.35]. For each i{1,2}i\in\{1,2\}, let

φi:X𝑓1×1pri1=:Zi\varphi_{i}\colon X\xrightarrow{f}\mathbb{P}^{1}\times\mathbb{P}^{1}\xrightarrow{\operatorname{pr}_{i}}\mathbb{P}^{1}=:Z_{i}

be the induced composite contraction. Note that each φi\varphi_{i} corresponds to a two-dimensional extremal face of NE(X)\operatorname{NE}(X). Let φ3:XZ3=2\varphi_{3}\colon X\to Z_{3}=\mathbb{P}^{2} be the contraction of the remaining two-dimensional extremal face of NE(X)\operatorname{NE}(X). For each i{1,2,3}i\in\{1,2,3\}, let HiH_{i} be the pullback of the ample generator on ZiZ_{i}. We recall KXH1+H2+H3-K_{X}\sim H_{1}+H_{2}+H_{3} ([FanoIV, Propositions 4.33 and 4.35]). In particular,

2=KXH1H2=(H1+H2+H3)H1H2=H1H2H3.2=-K_{X}\cdot H_{1}\cdot H_{2}=(H_{1}+H_{2}+H_{3})\cdot H_{1}\cdot H_{2}=H_{1}\cdot H_{2}\cdot H_{3}.

For each i{1,2,3}i\in\{1,2,3\}, HiH^{\prime}_{i} denotes the generic member of HiH_{i}, which is a regular prime divisor on XX×kkX^{\prime}\coloneqq X\times_{k}k^{\prime} for a suitable purely transcendental extension k/kk^{\prime}/k (Remark 2.4). Note that H1H2H^{\prime}_{1}\cap H^{\prime}_{2} is a smooth curve, because ff is not wild [MS03, Corollary 8 and Remark 9].

We now finish the proof by assuming that

  1. (i)

    H1(X,KX+H1+2H3)=H1(X,KX+H2+2H3)=0H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{1}+2H^{\prime}_{3})=H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{2}+2H^{\prime}_{3})=0, and

  2. (ii)

    H2(X,KX+2H3)=0H^{2}(X^{\prime},K_{X^{\prime}}+2H^{\prime}_{3})=0.

By Corollary 2.11 and Lemma 2.12, it suffices to find H3′′|H3|H^{\prime\prime}_{3}\in|H^{\prime}_{3}| such that H1H2H3′′H^{\prime}_{1}\cap H^{\prime}_{2}\cap H^{\prime\prime}_{3} is smooth and zero-dimensional. Since kk^{\prime} is an infinite field and H1H2H^{\prime}_{1}\cap H^{\prime}_{2} is isomorphic to a smooth conic on k2\mathbb{P}^{2}_{k^{\prime}}, the Bertini theorem enables us to find a smooth zero-dimensional effective Cartier divisor DD on H1H2H^{\prime}_{1}\cap H^{\prime}_{2} such that H3|H1H2DH^{\prime}_{3}|_{H^{\prime}_{1}\cap H^{\prime}_{2}}\sim D. Therefore, it is enough to show that the restriction maps

H0(X,𝒪X(H3))𝛼H0(H1,𝒪X(H3)|H1)𝛽H0(H1H2,𝒪X(H3)|H1H2)H^{0}(X^{\prime},\mathcal{O}_{X}(H^{\prime}_{3}))\xrightarrow{\alpha}H^{0}(H^{\prime}_{1},\mathcal{O}_{X}(H^{\prime}_{3})|_{H^{\prime}_{1}})\xrightarrow{\beta}H^{0}(H^{\prime}_{1}\cap H^{\prime}_{2},\mathcal{O}_{X}(H^{\prime}_{3})|_{H^{\prime}_{1}\cap H^{\prime}_{2}})

are surjective. The restriction map α\alpha is surjective, because

H1(X,H3H1)H1(X,KX+H2+2H3)=(i)0.H^{1}(X^{\prime},H^{\prime}_{3}-H^{\prime}_{1})\simeq H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{2}+2H^{\prime}_{3})\overset{{\rm(i)}}{=}0.

The problem is reduced to the surjectivity of β\beta. To this end, it suffices to prove H1(H1,𝒪X(H2+H3)|H1)=0H^{1}(H^{\prime}_{1},\mathcal{O}_{X^{\prime}}(-H^{\prime}_{2}+H^{\prime}_{3})|_{H^{\prime}_{1}})=0. We have an exact sequence

H1(X,H2+H3)H1(H1,𝒪X(H2+H3)|H1)H2(X,H1H2+H3).H^{1}(X^{\prime},-H^{\prime}_{2}+H^{\prime}_{3})\to H^{1}(H^{\prime}_{1},\mathcal{O}_{X^{\prime}}(-H^{\prime}_{2}+H^{\prime}_{3})|_{H^{\prime}_{1}})\to H^{2}(X^{\prime},-H^{\prime}_{1}-H^{\prime}_{2}+H^{\prime}_{3}).

By

H2+H3KX+H1+2H3andH1H2+H3KX+2H3,-H^{\prime}_{2}+H^{\prime}_{3}\sim K_{X^{\prime}}+H^{\prime}_{1}+2H^{\prime}_{3}\qquad{\rm and}\qquad-H^{\prime}_{1}-H^{\prime}_{2}+H^{\prime}_{3}\sim K_{X^{\prime}}+2H^{\prime}_{3},

(i) and (ii) imply H1(H1,(H2+H3)|H1)=0H^{1}(H^{\prime}_{1},(-H^{\prime}_{2}+H^{\prime}_{3})|_{H^{\prime}_{1}})=0. Therefore, it is enough to prove (i) and (ii).

Claim.

The following hold.

  1. (1)

    HiH^{\prime}_{i} is a regular weak del Pezzo surface for every i{1,2,3}i\in\{1,2,3\}.

  2. (2)

    H0(Hi,𝒪Hi)=kH^{0}(H^{\prime}_{i},\mathcal{O}_{H^{\prime}_{i}})=k^{\prime} for every i{1,2,3}i\in\{1,2,3\}.

  3. (3)

    H2(X,KX+Hi)=0H^{2}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{i})=0 for every i{1,2,3}i\in\{1,2,3\}.

  4. (4)

    H2(X,KX+Hi+Hj)=H2(X,KX+2H3)=0H^{2}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{i}+H^{\prime}_{j})=H^{2}(X^{\prime},K_{X^{\prime}}+2H^{\prime}_{3})=0 for 1i<j31\leq i<j\leq 3.

  5. (5)

    H1(Hi,𝒪Hi)=0H^{1}(H^{\prime}_{i},\mathcal{O}_{H^{\prime}_{i}})=0 and H1(Hi,KHi)=0H^{1}(H^{\prime}_{i},K_{H^{\prime}_{i}})=0 for every i{1,2,3}i\in\{1,2,3\}.

  6. (6)

    H0(H1H3,𝒪H1H3)=kH^{0}(H^{\prime}_{1}\cap H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{1}\cap H^{\prime}_{3}})=k^{\prime} and H0(H2H3,𝒪H2H3)=kH^{0}(H^{\prime}_{2}\cap H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{2}\cap H^{\prime}_{3}})=k^{\prime}.

  7. (7)

    H1(H3,KH3+(H1+H3)|H3)=0H^{1}(H^{\prime}_{3},K_{H^{\prime}_{3}}+(H^{\prime}_{1}+H^{\prime}_{3})|_{H^{\prime}_{3}})=0

Proof of Claim.

Since HiH^{\prime}_{i} is the generic member of a base point free linear system |Hi||H_{i}|, HiH^{\prime}_{i} is a regular prime divisor on XX^{\prime}. If i{1,2}i\in\{1,2\}, then KHi-K_{H^{\prime}_{i}} is ample, because HiH^{\prime}_{i} is the generic fibre of φi:X1\varphi_{i}\colon X\to\mathbb{P}^{1}. We see that KH3-K_{H^{\prime}_{3}} is nef and big by KH3(H1+H2)|H3-K_{H^{\prime}_{3}}\sim(H^{\prime}_{1}+H^{\prime}_{2})|_{H^{\prime}_{3}} and (H1+H2)2H3=2H1H2H3>0(H^{\prime}_{1}+H^{\prime}_{2})^{2}\cdot H^{\prime}_{3}=2H^{\prime}_{1}\cdot H^{\prime}_{2}\cdot H^{\prime}_{3}>0. Thus (1) holds. If i{1,2}i\in\{1,2\}, then (2) holds by the fact that HiH_{i} is (a base change of) the generic fibre of a contraction φi:X1\varphi_{i}\colon X\to\mathbb{P}^{1}. We have H0(H3,𝒪H3)=kH^{0}(H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{3}})=k^{\prime}, because general members of the complete linear system |H3||H_{3}| are geometrically integral [FanoI, Proposition 2.10]. Thus, (2) holds.

Let us show (3). Consider an exact sequence

0=H2(X,KX)H2(X,KX+Hi)H2(Hi,KHi)H3(X,KX)H3(X,KX+Hi)=0.0=H^{2}(X^{\prime},K_{X^{\prime}})\to H^{2}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{i})\to H^{2}(H^{\prime}_{i},K_{H^{\prime}_{i}})\\ \to H^{3}(X^{\prime},K_{X^{\prime}})\to H^{3}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{i})=0.

By (2) and Serre duality, we obtain h2(Hi,KHi)=1h^{2}(H^{\prime}_{i},K_{H^{\prime}_{i}})=1 and h3(X,KX)=1h^{3}(X^{\prime},K_{X^{\prime}})=1. Therefore, H2(X,KX+Hi)=0H^{2}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{i})=0. Thus (3) holds.

Let us show (4). If iji\neq j, then H2(X,KX+Hi+Hj)=0H^{2}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{i}+H^{\prime}_{j})=0 by an exact sequence

0=H2(X,KX+Hi)H2(X,KX+Hi+Hj)H2(Hj,KHj+Hi)=0,0=H^{2}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{i})\to H^{2}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{i}+H^{\prime}_{j})\to H^{2}(H^{\prime}_{j},K_{H^{\prime}_{j}}+H^{\prime}_{i})=0,

where H2(Hj,KHj+Hi)=0H^{2}(H^{\prime}_{j},K_{H^{\prime}_{j}}+H^{\prime}_{i})=0 follows from Serre duality and H0(Hj,𝒪X(Hi)|Hj)=0H^{0}(H^{\prime}_{j},\mathcal{O}_{X^{\prime}}(-H^{\prime}_{i})|_{H^{\prime}_{j}})=0. Similarly, H2(X,KX+2H3)=0H^{2}(X^{\prime},K_{X^{\prime}}+2H^{\prime}_{3})=0 by H0(H3,𝒪X(H3)|H3)=0H^{0}(H^{\prime}_{3},\mathcal{O}_{X^{\prime}}(-H^{\prime}_{3})|_{H^{\prime}_{3}})=0. Thus (4) holds.

Let us show (5). By an exact sequence

0=H1(X,𝒪X)H1(Hi,𝒪Hi)H2(X,𝒪X(Hi)),0=H^{1}(X^{\prime},\mathcal{O}_{X^{\prime}})\to H^{1}(H^{\prime}_{i},\mathcal{O}_{H^{\prime}_{i}})\to H^{2}(X^{\prime},\mathcal{O}_{X^{\prime}}(-H^{\prime}_{i})),

it suffices to prove H2(X,𝒪X(Hi))=0H^{2}(X^{\prime},\mathcal{O}_{X^{\prime}}(-H^{\prime}_{i}))=0, which follows from (4) by using HiKX+H1+H2+H3Hi-H^{\prime}_{i}\sim K_{X^{\prime}}+H^{\prime}_{1}+H^{\prime}_{2}+H^{\prime}_{3}-H^{\prime}_{i}. Thus (5) holds.

Let us show (6). Since XX has exactly three extremal rays, there is the extremal ray RR such that RR is the intersection of the extremal faces corresponding to X1X\to\mathbb{P}^{1} and X2X\to\mathbb{P}^{2}. Let f:XYf\colon X\to Y be the contraction of RR. If XX is 3-3, then Y=1×2Y=\mathbb{P}^{1}\times\mathbb{P}^{2} and f:XY1×2f\colon X\to Y\coloneqq\mathbb{P}^{1}\times\mathbb{P}^{2} and H1H3H^{\prime}_{1}\cap H^{\prime}_{3} is isomorphic to the corresponding intersection H1YH3YH_{1}^{\prime Y}\cap H_{3}^{\prime Y} on YY, because H1YH3YH_{1}^{\prime Y}\cap H_{3}^{\prime Y} is disjoint from f(Ex(f))f(\operatorname{Ex}(f)). Therefore, H1H3H^{\prime}_{1}\cap H^{\prime}_{3} is geometrically integral. If XX is 3-4, then Y=𝔽1Y=\mathbb{F}_{1} or YY is a smooth Fano threefold of No. 2-18. If Y=𝔽1Y=\mathbb{F}_{1}, then H1H3H^{\prime}_{1}\cap H^{\prime}_{3} is a smooth fibre of f:X𝔽1f\colon X\to\mathbb{F}_{1}. The other case follows from Lemma 4.2 by using the upper semi-continuity [Har77, Ch. III, Theorem 12.8]. Thus (6) holds.

Let us show (7). We have

H1H32=H2H32=1,H_{1}\cdot H^{2}_{3}=H_{2}\cdot H^{2}_{3}=1,

because 2=KXH32=(H1+H2)H322=-K_{X}\cdot H^{2}_{3}=(H_{1}+H_{2})\cdot H_{3}^{2} and a fibre ζH32\zeta\equiv H_{3}^{2} of φ3\varphi_{3} is not contracted by φi\varphi_{i} for each i{1,2}i\in\{1,2\}. Fix a general member H3′′H^{\prime\prime}_{3} of |H3||H^{\prime}_{3}|. By Serre duality, it is enough to show H1(H3,𝒪H3(D))=0H^{1}(H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{3}}(-D))=0 for an effective Cartier divisor D(H1+H3′′)|H3D\coloneqq(H^{\prime}_{1}+H^{\prime\prime}_{3})|_{H^{\prime}_{3}} on H3H^{\prime}_{3}. By (2), the problem is reduced to H0(D,𝒪D)=kH^{0}(D,\mathcal{O}_{D})=k^{\prime}. Clearly, DD is nef. It holds that D2=(H1+H3)2H3H1H32=1>0D^{2}=(H_{1}+H_{3})^{2}\cdot H_{3}\geq H_{1}\cdot H^{2}_{3}=1>0. Hence DD is nef and big. Then H0(D,𝒪D)H^{0}(D,\mathcal{O}_{D}) is a field [Eno, Corollary 3.17]. Since SuppD\operatorname{Supp}D contains H1H3H3′′H^{\prime}_{1}\cap H^{\prime}_{3}\cap H^{\prime\prime}_{3} which is a kk^{\prime}-rational point by H1H3H3′′=H1H32=1H^{\prime}_{1}\cdot H^{\prime}_{3}\cdot H^{\prime\prime}_{3}=H_{1}\cdot H_{3}^{2}=1, we obtain field extensions

k=H0(H3,𝒪H3)H0(D,𝒪D)H0(H1H3H3′′,𝒪H1H3H3′′)=k,k^{\prime}=H^{0}(H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{3}})\hookrightarrow H^{0}(D,\mathcal{O}_{D})\hookrightarrow H^{0}(H^{\prime}_{1}\cap H^{\prime}_{3}\cap H^{\prime\prime}_{3},\mathcal{O}_{H^{\prime}_{1}\cap H^{\prime}_{3}\cap H^{\prime\prime}_{3}})=k^{\prime},

which implies H0(D,𝒪D)=kH^{0}(D,\mathcal{O}_{D})=k^{\prime}, as required. This completes the proof of Claim. ∎

It is enough to show (i) and (ii). As (ii) has been settled by Claim(4), let us show (i). By H0(H3,𝒪H3)=kH^{0}(H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{3}})=k^{\prime} and H0(H1H3,𝒪H1H3)=kH^{0}(H^{\prime}_{1}\cap H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{1}\cap H^{\prime}_{3}})=k^{\prime} (Claim(2)(6)), we get H3(H3,𝒪H3(H1|H3))=0H^{3}(H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{3}}(-H^{\prime}_{1}|_{H^{\prime}_{3}}))=0 by the following exact sequence:

H0(H3,𝒪H3)H0(H1H3,𝒪H1H3)H1(H3,𝒪H3(H1|H3))H1(H3,𝒪H3)=0.H^{0}(H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{3}})\xrightarrow{\simeq}H^{0}(H^{\prime}_{1}\cap H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{1}\cap H^{\prime}_{3}})\to H^{1}(H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{3}}(-H^{\prime}_{1}|_{H^{\prime}_{3}}))\to H^{1}(H^{\prime}_{3},\mathcal{O}_{H^{\prime}_{3}})=0.

By Serre duality, we get

H1(H3,KH3+H1|H3)=0.H^{1}(H^{\prime}_{3},K_{H^{\prime}_{3}}+H^{\prime}_{1}|_{H^{\prime}_{3}})=0.

We have the following exact sequences:

0=H1(X,KX)H1(X,KX+H1)H1(H1,KH1)=0,0=H^{1}(X^{\prime},K_{X^{\prime}})\to H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{1})\to H^{1}(H^{\prime}_{1},K_{H^{\prime}_{1}})=0,
0=H1(X,KX+H1)H1(X,KX+H1+H3)H1(H3,KH3+H1|H3)=00=H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{1})\to H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{1}+H^{\prime}_{3})\to H^{1}(H^{\prime}_{3},K_{H^{\prime}_{3}}+H^{\prime}_{1}|_{H^{\prime}_{3}})=0
0=H1(X,KX+H1+H3)H1(X,KX+H1+2H3)H1(H3,KH3+(H1+H3)|H3)=0,0=H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{1}+H^{\prime}_{3})\to H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{1}+2H^{\prime}_{3})\\ \to H^{1}(H^{\prime}_{3},K_{H^{\prime}_{3}}+(H^{\prime}_{1}+H^{\prime}_{3})|_{H^{\prime}_{3}})=0,

where the last equality follows from Claim (7). Thus H1(X,KX+H1+2H3)=0H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{1}+2H^{\prime}_{3})=0. Similarly, we obtain H1(X,KX+H2+2H3)=0H^{1}(X^{\prime},K_{X^{\prime}}+H^{\prime}_{2}+2H^{\prime}_{3})=0. Thus (i) holds. ∎

Proposition 4.4.

Let XX be a smooth Fano threefold with ρ(X)=3\rho(X)=3.

  1. (1)

    If XX is not 3-10, then XX is FF-split.

  2. (2)

    Assume that XX is 3-10. Then XX is FF-split if and only if XX has no wild conic bunlde structure.

Proof.

We may assume that XX is none of 3-1, 3-3, and 3-4 (Lemma 4.1, Lemma 4.3). Note that if XX has a wild conic bundle structure, then XX is not FF-split [GLP15, Theorem 2.1 or Corollary 2.5]. In what follows, we assume that any conic bundle structure from XX is generically smooth. Under this additional assumption, it suffices to show that XX is FF-split. We divide the proof into the following four cases.

  1. (1)

    3-27, 3-28, 3-31.

  2. (2)

    3-5, 3-8, 3-12, 3-13, 3-15, 3-16, 3-17, 3-19, 3-20, 3-21, 3-22, 3-23, 3-24, 3-26, 3-29.

  3. (3)

    3-6, 3-10, 3-18, 3-25.

  4. (4)

    3-2, 3-7, 3-9, 3-11, 3-14, 3-30.

(1) In this case, XX is toric [FanoIV, Subsection 7.3], and hence FF-split.

(2) In this case, there exist a smooth Fano threefold YY with ρ(Y)=2\rho(Y)=2, a 1\mathbb{P}^{1}-bundle g:Y2g\colon Y\to\mathbb{P}^{2}, and a subsection BYB\subset Y of gg such that XBlBYX\simeq\operatorname{Bl}_{B}Y, B2g(B)1B_{\mathbb{P}^{2}}\coloneqq g(B)\simeq\mathbb{P}^{1}, and YY is one of 2-32, 2-34, 2-35 [FanoIV, Subsection 7.3, cf. Theorem 4.23]. Set Sg1(B)S\coloneqq g^{-1}(B), which is a 1\mathbb{P}^{1}-bundle over 1\mathbb{P}^{1}, and hence FF-split. Let TT be the pullback of the ample generator by the contraction of the other extremal ray RR. By [FanoIII, Proposition 5.9(3)], we can write KYaS+2T-K_{Y}\sim aS+2T, where aa is the length of RR. Then we can check that a>1a>1 in each case [FanoIV, Subsection 7.2]. Thus (KY+S)(a1)S+bT-(K_{Y}+S)\sim(a-1)S+bT is ample by Kleiman’s criterion. Therefore, (Y,S)(Y,S) is FF-split (Proposition 2.8), and hence XX is FF-split (Proposition 2.7).

(3) In this case, we can write XBlCC3X\simeq\operatorname{Bl}_{C\amalg C^{\prime}}\mathbb{P}^{3} or XBlCCQX\simeq\operatorname{Bl}_{C\amalg C^{\prime}}Q for a disjoint union of smooth curves CC and CC^{\prime} on 3\mathbb{P}^{3} or QQ [FanoIV, Subsection 7.3]. We only treat the case when XX is 3-6, as the other cases are similar. In this case, XBlCC3X\simeq\operatorname{Bl}_{C\amalg C^{\prime}}\mathbb{P}^{3}, CC is a line, and we can write C=S1S2C^{\prime}=S_{1}\cap S_{2} for some quadric surfaces S1S_{1} and S2S_{2}. Take a general plane HH containing the line CC and a general quadric surface SS containing CC^{\prime}. Let HXH_{X} and SXS_{X} be the proper transforms of HH and SS, respectively. Although HXHH_{X}\to H and SXSS_{X}\to S are not necessarily isomorphisms, these birational morphisms are isomorphic over HSH\cap S. Therefore, we obtain HXSXHSH_{X}\cap S_{X}\xrightarrow{\simeq}H\cap S. This is nothing but a general fibre of the contraction X1×1X\to\mathbb{P}^{1}\times\mathbb{P}^{1} [FanoIV, Proposition 4.37]. Since this is not a wild conic bundle, we get HSHXSX1H\cap S\simeq H_{X}\cap S_{X}\simeq\mathbb{P}^{1}. Then (3,H+S)(\mathbb{P}^{3},H+S) is FF-split by applying Proposition 2.8 twice. Hence XX is FF-split (Proposition 2.7).

(4) In what follows, we treat the remaining cases separately.

3-2: We use the same notation as in [FanoIV, Proposition 4.32]. We have KX2H1+H2+D-K_{X}\sim 2H_{1}+H_{2}+D and a conic bundle f:X1×1f:X\to\mathbb{P}^{1}\times\mathbb{P}^{1}, where D1×1D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}, and each HiH_{i} is the pullback of 𝒪1(1)\mathcal{O}_{\mathbb{P}^{1}}(1) by X𝑓1×1pri1X\xrightarrow{f}\mathbb{P}^{1}\times\mathbb{P}^{1}\xrightarrow{\operatorname{pr}_{i}}\mathbb{P}^{1}. Moreover f|D:D=1×11×1f|_{D}\colon D=\mathbb{P}^{1}\times\mathbb{P}^{1}\to\mathbb{P}^{1}\times\mathbb{P}^{1} is a finite double cover which can be written as id×ψ{\rm id}\times\psi for some double cover ψ:11\psi\colon\mathbb{P}^{1}\to\mathbb{P}^{1}. For the generic member H2genH^{\operatorname{gen}}_{2} of |H2||H_{2}|, the intersection CH2genDC\coloneqq H_{2}^{\operatorname{gen}}\cap D is a regular curve of genus zero with KC+2H1|C0K_{C}+2H_{1}|_{C}\sim 0, because H1C=H1H2D>0H_{1}\cdot C=H_{1}\cdot H_{2}\cdot D>0. By deg(H1|C)=1\deg(H_{1}|_{C})=1, we get Cκ1C\simeq\mathbb{P}^{1}_{\kappa}. Hence XX is FF-split (Corollary 2.11).

3-7: In this case, f:XWf\colon X\to W is a blowup along an elliptic curve B=S1S2B=S_{1}\cap S_{2} with S1,S2|(1/2)KW|S_{1},S_{2}\in|-(1/2)K_{W}| [FanoIV, Subsection 7.3]. Let SS be a general member of |(1/2)KW||-(1/2)K_{W}| containing BB. Since BB is an ample effective Cartier divisor on SS, it follows that SS is smooth along BB and SS is normal. Note that the proper transform SXS_{X} of SS on XX is a fibre of a contraction π:X1\pi\colon X\to\mathbb{P}^{1}. Therefore, the geometric generic fibre XK¯X_{\overline{K}} of π\pi is normal, where KK(1)K\coloneqq K(\mathbb{P}^{1}) and K¯\overline{K} denotes the algebraic closure of KK. Since XKX_{K} is a regular del Pezzo surface, XK¯X_{\overline{K}} has at worst canonical singularities [BT22, Theorem 3.3]. Therefore, a general fibre SXS_{X} of π:X1\pi\colon X\to\mathbb{P}^{1} is a canonical del Pezzo surface. By KS2=6K_{S}^{2}=6, Theorem 2.22 shows that S(SX)S(\simeq S_{X}) is FF-split. Since (KW+S)-(K_{W}+S) is ample, we have (W,S)(W,S) is FF-split (Proposition 2.8), and hence XX is FF-split (Proposition 2.7).

3-9: By [FanoIV, Proposition 4.42], there is a blowup f:XY=Y2-36=2(𝒪𝒪(2))f\colon X\to Y=Y_{\text{2-36}}=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}\oplus\mathcal{O}(2)) along a smooth curve BB such that

  • BB is contained in a section SS of the 1\mathbb{P}^{1}-bundle structure π:2(𝒪𝒪(2))2\pi\colon\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}\oplus\mathcal{O}(2))\to\mathbb{P}^{2}, and

  • SS is disjoint from another section TT of π:2(𝒪𝒪(2))2\pi\colon\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}\oplus\mathcal{O}(2))\to\mathbb{P}^{2}.

Since π\pi is 1\mathbb{P}^{1}-bundle, we have KY+S+T=πK2K_{Y}+S+T=\pi^{*}K_{\mathbb{P}^{2}}. Since (KY+S+(1ϵ)T)=πK2+ϵT-(K_{Y}+S+(1-\epsilon)T)=-\pi^{*}K_{\mathbb{P}^{2}}+\epsilon T is ample for some 0<ϵ10<\epsilon\ll 1 and (S,0)(S,0) is FF-split, (Y,S+(1ϵ)T)(Y,S+(1-\epsilon)T) is FF-split (Proposition 2.8). Hence XX is FF-split (Proposition 2.7).

3-11: By [FanoIV, Subsection 7.3], there exists a blowup f:XV7f\colon X\to V_{7} along an elliptic curve B=S1S2B=S_{1}\cap S_{2} with S1,S2|(1/2)KV7|S_{1},S_{2}\in|-(1/2)K_{V_{7}}|. By the same argument as in that of 3-7, a general member SS of |(1/2)KV7||(-1/2)K_{V_{7}}| is a canonical del Pezzo surface with KS2=7K_{S}^{2}=7. Then SS is FF-split (Theorem 2.22). Since (KV7+S)-(K_{V_{7}}+S) is ample, (V7,S)(V_{7},S) is FF-split (Proposition 2.8), and hence XX is FF-split (Proposition 2.7).

3-14: By [FanoIV, Subsection 7.3], we have X=BlPC3X=\operatorname{Bl}_{P\amalg C}\mathbb{P}^{3}, where CC is a smooth cubic curve contained in a plane HH and PP is a point satisfying PHP\not\in H. Take two distinct planes HH^{\prime} and H′′H^{\prime\prime} containing PP. Then (3,H+H+H′′)(\mathbb{P}^{3},H+H^{\prime}+H^{\prime\prime}) is FF-split, and hence so is XX (Proposition 2.7).

3-30: By [FanoIV, Subsection 7.3], there exist blowups

XV73,X\to V_{7}\to\mathbb{P}^{3},

where V73V_{7}\to\mathbb{P}^{3} is a blowup at a point P3P\in\mathbb{P}^{3} and the blowup centre of XV7X\to V_{7} is the proper transform of a line LL passing through PP. Take two planes H,HH,H^{\prime} containing LL. Then (3,H+H)(\mathbb{P}^{3},H+H^{\prime}) is F-split, and hence XX is F-split (Proposition 2.7). ∎

5. Fano threefolds with ρ=2\rho=2 (except for 2-2, 2-6, 2-8)

5.1. Quasi-FF-splitting (imprimitive case)

5.1.1. D+ED+E (imprimitive)

Proposition 5.1.

Let XX be a smooth Fano threefold with ρ(X)=2\rho(X)=2. If the types of the extremal rays are D+ED+E, then XX is quasi-FF-split.

Proof.

In this case, XX is imprimitive [FanoIV, Subsection 7.2], i.e., the types of the extremal rays are D+E1D+E_{1}. Let f:XYf:X\to Y (resp. π:X1\pi:X\to\mathbb{P}^{1}) be the contraction of type E1E_{1} (resp. type DD). Let BB be the smooth curve on YY that is the blowup centre of f:XYf:X\to Y, i.e., X=BlBYX=\operatorname{Bl}_{B}Y. By [FanoIV, Subsection 7.2], we may assume that one of the following holds.

  1. (1)

    Y=3Y=\mathbb{P}^{3} (2-4, 2-25, 2-33).

  2. (2)

    Y=QY=Q, where QQ is a smooth quadric hypersurface on 4\mathbb{P}^{4} (2-7, 2-29).

  3. (3)

    Y=VdY=V_{d} with 1d51\leq d\leq 5, where VdV_{d} is a smooth Fano threefold of index two satisfying (KVd)3=8d(-K_{V_{d}})^{3}=8d (2-1, 2-3, 2-5, 2-10, 2-14).

Claim.

A general fibre DXD_{X} of π:X1\pi:X\to\mathbb{P}^{1} is a canonical del Pezzo surface.

Proof of Claim.

By [FS20, Theorem 15.2], the generic fibre XKX_{K} of π\pi is geometrically normal, where KK denotes the function field of 1\mathbb{P}^{1}. Then its base change XK¯:=XK×KK¯X_{\overline{K}}:=X_{K}\times_{K}\overline{K} to the algebraic closure K¯\overline{K} has at worst canonical singularities [BT22, Theorem 3.3]. Hence a general fibre DXD_{X} of π\pi is normal and has at worst canonical singularities. This completes the proof of Claim. ∎

(1) Assume that Y=3Y=\mathbb{P}^{3}. In this case, B=DDB=D\cap D^{\prime} for some surfaces DD of degree 1e31\leq e\leq 3, i.e., YY is 2-4 (e=3e=3), 2-25 (e=2e=2), or 2-33 (e=1e=1). Although DD might be singular, DD is a normal prime divisor on XX, because DD is smooth along an effective ample Cartier divisor B=D|DB=D^{\prime}|_{D}. After replacing DD by a general member of the pencil generated by DD and DD^{\prime}, we may assume that DD is a canonical del Pezzo surface (Claim). If e{1,2}e\in\{1,2\}, then DD is FF-split (Theorem 2.22). Then (Y,D)(Y,D) is FF-split (Corollary 2.9), which implies that XX is FF-split (Proposition 2.7). We may assume that DD is a cubic surface, i.e., XX is 2-4. Recall that KXDX+H-K_{X}\sim D_{X}+H, where DXD_{X} denotes the proper transform of DD and Hf𝒪3(1)H\coloneqq f^{*}\mathcal{O}_{\mathbb{P}^{3}}(1). Replacing DXD_{X} and HH by general members of |DX||D_{X}| and |H||H|, we obtain DXHDfHD_{X}\cap H\simeq D\cap{f_{*}}H. Since DHD\cap H is a general hyperplane section of a normal cubic surface DD, we have DfHDXHD\cap{f_{*}}H\simeq D_{X}\cap H is an elliptic curve. Hence XX is quasi-FF-split (Corollary 2.18, Remark 2.19).

(2) Assume that Y=QY=Q. In this case, B=DDB=D\cap D^{\prime} for some surfaces D|𝒪Q(e)|D\in|\mathcal{O}_{Q}(e)| with 1e21\leq e\leq 2, i.e., YY is 2-29 (e=1e=1), or 2-7 (e=2)(e=2). By Claim, DD is a canonical del Pezzo surface. We have

KD2=(KQ+D)2D=2e(3e)2.K_{D}^{2}=(K_{Q}+D)^{2}\cdot D=2e(3-e)^{2}.

If e=1e=1, then KD2=8K_{D}^{2}=8, and thus DD is FF-split (Theorem 2.22). Then (Y,D)(Y,D) is FF-split (Corollary 2.9), which implies that XX is FF-split (Proposition 2.7). We may assume that e=2e=2, i.e., XX is 2-7. Set Hf𝒪Q(1)H\coloneqq f^{*}\mathcal{O}_{Q}(1). Replacing DD and fHf_{*}H by general members of |DX||D_{X}| and |𝒪Q(1)||\mathcal{O}_{Q}(1)|, we obtain DXHDfHD_{X}\cap H\simeq D\cap f_{*}H. Since DfHD\cap{f_{*}}H is a general hyperplane section of a canonical del Pezzo surface DD of degree 44 with KDfH|D=DfH-K_{D}\sim{f_{*}}H|_{D}=D\cap{f_{*}}H, it follows that DfHDXHD\cap f_{*}H\simeq D_{X}\cap H is an elliptic curve. Hence XX is quasi-FF-split (Corollary 2.18, Remark 2.19).

(3) Assume that Y=VdY=V_{d}. In this case, BB is an elliptic curve which is a complete intersection of two prime divisors D|(1/2)KY|D\in|-(1/2)K_{Y}| and D|(1/2)KY|D^{\prime}\in|-(1/2)K_{Y}|, i.e., YY is 2-1, 2-3, 2-5, 2-10, or 2-14. We then get KX+DX+DX+E0K_{X}+D_{X}+D^{\prime}_{X}+E\sim 0, where E:=Ex(f)E:=\operatorname{Ex}(f) and DXD_{X} and DXD^{\prime}_{X} are the proper transforms of DD and DD^{\prime}, respectively. Fix general members DD and DD^{\prime} of |(1/2)KY||-(1/2)K_{Y}| containing BB. Let DgenD^{\operatorname{gen}} be the generic member of the pencil generated by DD and DD^{\prime}. Let κ\kappa be the function field κ\kappa of this pencil. For every kk-scheme ZZ, we set Zκ:=Z×kκZ_{\kappa}:=Z\times_{k}\kappa. Then DgenDκ=BκD^{\operatorname{gen}}\cap D^{\prime}_{\kappa}=B_{\kappa} and XκYκX_{\kappa}\to Y_{\kappa} is the blowup along BκB_{\kappa}. Note that the proper transform DXgenD^{\operatorname{gen}}_{X} is regular [Tan-Bertini, Theorem 4.9]. Since DXgen((DX)κ+Eκ)=DXgenEκBκD^{\operatorname{gen}}_{X}\cap((D^{\prime}_{X})_{\kappa}+E_{\kappa})=D^{\operatorname{gen}}_{X}\cap E_{\kappa}\simeq B_{\kappa} is smooth over κ\kappa, we conclude that XX is quasi-FF-split (Corollary 2.18, Remark 2.19). ∎

5.1.2. E+EE+E (imprimitive)

Proposition 5.2.

Let XX be a smooth Fano threefold with ρ(X)=2\rho(X)=2. Assume that the types of the extremal rays are E1E_{1} and EE. Then the following hold.

  1. (1)

    XX is quasi-FF-split.

  2. (2)

    If XX is not 2-12, then XX is FF-split.

Proof.

Let f:XYf\colon X\to Y be a contraction of type E1E_{1} and let f:XYf^{\prime}\colon X\to Y^{\prime} be the contraction of the other extremal ray, which is of type EE. Let HYH_{Y} (resp. HYH_{Y^{\prime}}) be the ample Cartier divisor that generates PicY\operatorname{Pic}\,Y (resp. PicY\operatorname{Pic}\,Y^{\prime}). Set HfHYH\coloneqq f^{*}H_{Y} and HfHYH^{\prime}\coloneqq f^{\prime*}H_{Y^{\prime}}. The list of such smooth Fano threefolds is as follows [FanoIV, Subseciton 7.2]:

2-12, 2-15, 2-17, 2-19, 2-21, 2-22, 2-23, 2-26, 2-28, 2-30.\text{2-12, 2-15, 2-17, 2-19, 2-21, 2-22, 2-23, 2-26, 2-28, 2-30}.

In particular, YY is 3\mathbb{P}^{3}, QQ, or VdV_{d} with 3d53\leq d\leq 5. Then |HY||H_{Y}| is very ample, and hence we may assume that HYH_{Y} is a smooth prime divisor on YY. Note that we have KXH+H-K_{X}\sim H+H^{\prime} except when XX is 2-30 [FanoIII, Remark 3.4 and Proposition 5.9].

Step 1.

If XX is 2-15, 2-28, or 2-30, then XX is FF-split.

Proof of Step 1.

In this case, there is a blowup f:XY=3f:X\to Y=\mathbb{P}^{3} along a smooth curve BB such that BB is contained in a prime divisor DD on 3\mathbb{P}^{3} of degree 2\leq 2 ([FanoIII, Proposition 9.3], [FanoIV, Subsection 7.2]). Since DD is FF-split and (K3+D)-(K_{\mathbb{P}^{3}}+D) is ample, it follows that (3,D)(\mathbb{P}^{3},D) is FF-split (Corollary 2.9), which implies that XX is FF-split (Proposition 2.7). This completes the proof of Step 1. ∎

Step 2.

If XX is 2-17 2-19, 2-21, 2-22, 2-23, or 2-26, then XX is FF-split.

Proof of Step 2.

Replace HH by a general member of |H||H|. We now prove that HH is a smooth prime divisor that is FF-split. We first treat the case when XX is 2-17, 2-19, or 2-22. In this case, there is a blowup f:XY=3f\colon X\to Y=\mathbb{P}^{3} along a smooth curve BB of degree 5\leq 5. For a general plane HYH_{Y}, its pullback H=fHYH=f^{*}H_{Y} is nothing but the blowup of HYH_{Y} along HYBH_{Y}\cap B. Since HYBH_{Y}\cap B is smooth and KH=(KX+H)|H=H|H-K_{H}=-(K_{X}+H)|_{H}=H^{\prime}|_{H} is nef and big, it follows that HH is a smooth weak del Pezzo surface. By KH2KHY25=4K_{H}^{2}\geq K_{H_{Y}}^{2}-5=4, we have HH is FF-split [KT, Proposition 3.6]. For the the remaining case (i.e., XX is 2-21, 2-23, or 2-26), we can apply the same argument, because there is a blowup f:XY=Qf\colon X\to Y=Q along a smooth curve BB of degree 4\leq 4, and hence we may apply [KT, Proposition 3.6]. Therefore, HH is a smooth prime divisor which is FF-split.

In order to prove that (X,H)(X,H) is FF-split, it is enough to show that

H1(X,H(pe1)(KX+H))=H1(X,KX+peH)=0H^{1}(X,-H-(p^{e}-1)(K_{X}+H))=H^{1}(X,K_{X}+p^{e}H^{\prime})=0

for some e>0e>0 by KXH+H-K_{X}\sim H+H^{\prime} and Proposition 2.8. Set EEx(f)E^{\prime}\coloneqq\operatorname{Ex}(f^{\prime}). By the Fujita vanishing theorem [Fuj17, Theorem 3.8.1], we can find e0>0e_{0}\in\mathbb{Z}_{>0} and s0>0s_{0}\in\mathbb{Z}_{>0} such that

H1(X,KX+peHs0E)=0H^{1}(X,K_{X}+p^{e}H^{\prime}-s_{0}E^{\prime})=0

for every integer ee0e\geq e_{0}. Indeed, we can find a0,t0>0a_{0},t_{0}\in\mathbb{Z}_{>0} such that a0Ht0Ea_{0}H^{\prime}-t_{0}E is ample. Then, by Fujita vanishing, there exists m0m\gg 0 such that

H1(X,KX+m(a0Ht0E)+N)=0H^{1}(X,K_{X}+m(a_{0}H^{\prime}-t_{0}E^{\prime})+N)=0

for any nef Cartier divisor NN on XX. Then we take s0=mt0s_{0}=mt_{0} and e0>0e_{0}\in\mathbb{Z}_{>0} satisfying pema0p^{e}\geq ma_{0}.

Recall that 𝒪X(mHE)|E\mathcal{O}_{X}(mH^{\prime}-E^{\prime})|_{E^{\prime}} is ample for some m0m\gg 0 and EE^{\prime} satisfies the Kodaira vanishing theorem (for Cartier divisors), because EE^{\prime} is a smooth projective surface with negative Kodaira dimension or a singular quadric surface [FanoIII, Definition 3.3]. Hence, for each 0ss00\leq s\leq s_{0}, we can find e(s)>0e(s)\in\mathbb{Z}_{>0} such that

H1(E,𝒪X(KX+peHsE)|E)=H1(E,𝒪E(KE+peH(s+1)E))=0H^{1}(E^{\prime},\mathcal{O}_{X}(K_{X}+p^{e}H^{\prime}-sE^{\prime})|_{E^{\prime}})=H^{1}(E^{\prime},\mathcal{O}_{E^{\prime}}(K_{E^{\prime}}+p^{e}H^{\prime}-(s+1)E^{\prime}))=0

for every ee(s)e\geq e(s). Set

emax{e0,e(0),e(1),,e(s0)}.e\coloneqq\max\{e_{0},e(0),e(1),...,e(s_{0})\}.

It suffices to prove

H1(X,KX+peHsE)=0H^{1}(X,K_{X}+p^{e}H^{\prime}-sE^{\prime})=0

for every 0ss00\leq s\leq s_{0}. By descending induction on ss, this follows from

H1(X,KX+peH(s+1)E)H1(X,KX+peHsE)H1(S,𝒪X(KX+peHsE))=0.H^{1}(X,K_{X}+p^{e}H^{\prime}-(s+1)E^{\prime})\to H^{1}(X,K_{X}+p^{e}H^{\prime}-sE^{\prime})\\ \to H^{1}(S,\mathcal{O}_{X}(K_{X}+p^{e}H^{\prime}-sE^{\prime}))=0.

This completes the proof of Step 2. ∎

Step 3.

If XX is 2-12, then XX is quasi-FF-split.

Proof of Step 3.

In this case, the contraction of each extrmeal ray is a blowup XX of 3\mathbb{P}^{3} along a smooth curve BB of degree 66 [FanoIV, Subsection 7.2]. As in the argument in Step 2, replacing HH and HH^{\prime} by general members of |H||H| and |H||H^{\prime}| respectively, we may assume that HH and HH^{\prime} are smooth weak del Pezzo surfaces with KH2=3K_{H}^{2}=3.

We now finish the proof by assuming that the restriction map

ρ:H0(X,𝒪X(KXH))H0(H,𝒪H(KH))\rho\colon H^{0}(X,\mathcal{O}_{X}(-K_{X}-H))\to H^{0}(H,\mathcal{O}_{H}(-K_{H}))

is surjective. Since a general member CC of |KH||-K_{H}| is an elliptic curve [KN, Theorem 1.4] and HH^{\prime} is a general member in |H|=|KXH||H^{\prime}|=|-K_{X}-H|, it follows that HHH\cap H^{\prime} is an elliptic curve. Therefore, XX is quasi-FF-split by Corollary 2.18.

It suffices to show that ρ\rho is surjective. By an exact sequence

0𝒪X(KX2H)𝒪X(KXH)𝒪H(KH)0,0\to\mathcal{O}_{X}(-K_{X}-2H)\to\mathcal{O}_{X}(-K_{X}-H)\to\mathcal{O}_{H}(-K_{H})\to 0,

it is enough to prove that H1(X,KX2H)=0H^{1}(X,-K_{X}-2H)=0. Note that

KX2HKX+2H.-K_{X}-2H\sim K_{X}+2H^{\prime}.

Since HH^{\prime} is a smooth rational surface and H|HH^{\prime}|_{H^{\prime}} is nef and big, we have H1(H,KH+sH)=0H^{1}(H^{\prime},K_{H^{\prime}}+sH^{\prime})=0 for s0s\geq 0 by [Muk13, Theorem 3]. Thus, we have a surjection

H1(X,KX+sH)H1(X,KX+(s+1)H)H1(H,KH+sH)=0H^{1}(X,K_{X}+sH^{\prime})\to H^{1}(X,K_{X}+(s+1)H^{\prime})\to H^{1}(H^{\prime},K_{H^{\prime}}+sH^{\prime})=0

for every integer s0s\geq 0. Since we have H1(X,KX)=0H^{1}(X,K_{X})=0, using the above surjectivity for s{0,1}s\in\{0,1\}, we obtain H1(X,KX+2H)=0H^{1}(X,K_{X}+2H^{\prime})=0. This completes the proof of Step 3. ∎

Step 1, Step 2, and Step 3 complete the proof of Proposition 5.2. ∎

5.1.3. Langer surface

Definition 5.3.
  1. (1)

    For all the 𝔽2\mathbb{F}_{2}-points P1,,P7𝔽22P_{1},\ldots,P_{7}\in\mathbb{P}^{2}_{\mathbb{F}_{2}}, we set

    VL,𝔽2BlP1P7𝔽22.V_{L,\mathbb{F}_{2}}\coloneqq\operatorname{Bl}_{P_{1}\amalg\cdots\amalg P_{7}}\mathbb{P}^{2}_{\mathbb{F}_{2}}.

    For a field KK of characteristic two, we set VL,K:=VL,𝔽2×𝔽2KV_{L,K}:=V_{L,\mathbb{F}_{2}}\times_{\mathbb{F}_{2}}K, which is called the Langer surface over KK.

  2. (2)

    For a field of characteristic two and a zero-dimensional closed subscheme ZZ of K2\mathbb{P}^{2}_{K}, we say that ZZ is a Langer configuration if BlZK2\operatorname{Bl}_{Z}\mathbb{P}^{2}_{K} is KK-isomorphic to the Langer surface VL,KV_{L,K} over KK.

Lemma 5.4.

Let KK be an algebraically closed field of characteristic two. Take a Langer configuration ZK2Z\subset\mathbb{P}^{2}_{K}. Then there exists a KK-automorphism σ:K2K2\sigma:\mathbb{P}^{2}_{K}\xrightarrow{\simeq}\mathbb{P}^{2}_{K} such that σ(Z)=Z0\sigma(Z)=Z_{0}, where

Z0{[1:0:0],[0:1:0],[0:0:1],[1:1:0],[1:0:1],[0:1:1],[1:1:1]}.Z_{0}\coloneqq\{[1:0:0],[0:1:0],[0:0:1],[1:1:0],[1:0:1],[0:1:1],[1:1:1]\}.
Proof.

Fix a KK-isomorphism θ:BlZK2BlZ0K2=VL,K\theta\colon\operatorname{Bl}_{Z}\mathbb{P}^{2}_{K}\xrightarrow{\simeq}\operatorname{Bl}_{Z_{0}}\mathbb{P}^{2}_{K}=V_{L,K}. We have two birational contractions

φ:BlZK2K2,φ0:BlZ0K2K2,\varphi:\operatorname{Bl}_{Z}\mathbb{P}^{2}_{K}\to\mathbb{P}^{2}_{K},\qquad\varphi_{0}:\operatorname{Bl}_{Z_{0}}\mathbb{P}^{2}_{K}\to\mathbb{P}^{2}_{K},

where φ\varphi (resp. φ0\varphi_{0}) is the blowup along ZZ (resp. Z0Z_{0}). Recall that the Langer surface VL,KV_{L,K} has exactly 77 (1)(-1)-curves ([CT18, Theorem 5.4] or [KN, Lemma 4.5(4)]). Then both φ\varphi and φ0\varphi_{0} contracts all the (1)(-1)-curves on BlZK2\operatorname{Bl}_{Z}\mathbb{P}^{2}_{K} and BlZ0K2\operatorname{Bl}_{Z_{0}}\mathbb{P}^{2}_{K}. Therefore, we obtain a KK-automorphism σ:K2K2\sigma:\mathbb{P}^{2}_{K}\to\mathbb{P}^{2}_{K} which completes the following commutative diagram:

BlZK2{\operatorname{Bl}_{Z}\mathbb{P}^{2}_{K}}BlZ0K2{\operatorname{Bl}_{Z_{0}}\mathbb{P}^{2}_{K}}K2{\mathbb{P}^{2}_{K}}K2.{\mathbb{P}^{2}_{K}.}θ,\scriptstyle{\theta,\simeq}φ\scriptstyle{\varphi}φ0\scriptstyle{\varphi_{0}}σ,\scriptstyle{\sigma,\simeq}

This diagram shows that σ(Z)=Z0\sigma(Z)=Z_{0}. ∎

Lemma 5.5.

Let KK be a C1C_{1}-field of characteristic two and take its algebraic closure K¯\overline{K}. Let VV be a smooth projective surface over KK whose base change VK¯:=V×SpecKSpecK¯V_{\overline{K}}:=V\times_{\operatorname{Spec}K}\operatorname{Spec}{\overline{K}} is K¯\overline{K}-isomorphic to the Langer surface over K¯\overline{K}. Then the following hold.

  1. (1)

    If KK is perfect, then VV is KK-isomorphic to the Langer surface over KK.

  2. (2)

    ρ(V)=8\rho(V)=8.

Proof.

We now show the implication (1)\Rightarrow (2). Set KK^{\prime} to be the purely inseparable closure of KK in K¯\overline{K}, i.e.,

Ke=0K1/2e,K1/2e:={aK¯|a2eK}.K^{\prime}\coloneqq\bigcup_{e=0}^{\infty}K^{1/2^{e}},\qquad K^{1/2^{e}}:=\{a\in\overline{K}\,|\,a^{2^{e}}\in K\}.

Note that KK^{\prime} is a C1C_{1}-field, because being C1C_{1} is stable under algebraic extensions [GS17, Definition 6.2.1 and Lemma 6.2.4]. Therefore, (1) is applicable to the perfect C1C_{1}-field KK^{\prime} and the base change VK:=V×KKV_{K^{\prime}}:=V\times_{K}K^{\prime}. Therefore, VKV_{K^{\prime}} is the Langer surface over KK^{\prime}, and hence ρ(VK)=8\rho(V_{K^{\prime}})=8. Since the field extension KKK\subset K^{\prime} is purely inseparable, it holds that ρ(V)=ρ(VK)=8\rho(V)=\rho(V_{K^{\prime}})=8 [Tan18b, Proposition 2.4]. This completes the proof of the implication (1) \Rightarrow (2).

It suffices to show (1). Assume that KK is perfect. Recall that VK¯V_{\overline{K}} contains the exactly 77 (1)(-1)-curves E1,,E7E_{1},...,E_{7}. Set Γ¯E1++E7\overline{\Gamma}\coloneqq E_{1}+\cdots+E_{7} and ¯:=𝒪VK¯(Γ¯)\overline{\mathcal{L}}:=\mathcal{O}_{V_{\overline{K}}}(\overline{\Gamma}). We now show that

  1. (i)

    there is an invertible sheaf \mathcal{L} on VV such that α¯\alpha^{*}\mathcal{L}\simeq\overline{\mathcal{L}}, where α:VK¯V\alpha\colon V_{\overline{K}}\to V is the natural morphism and

  2. (ii)

    there exists an effective Cartier divisor Γ\Gamma on VV such that 𝒪X(Γ)\mathcal{O}_{X}(\Gamma)\simeq\mathcal{L} and the equality αΓ=Γ¯\alpha^{*}\Gamma=\overline{\Gamma} of Weil divisors holds.

Let us show (i). By H0(V,𝒪V)=KH^{0}(V,\mathcal{O}_{V})=K and Br(K)=0\operatorname{Br}(K)=0 [GS17, Proposition 6.2.3], we obtain PicVPic(VK¯)Gal(K¯/K)\operatorname{Pic}\,V\xrightarrow{\simeq}\operatorname{Pic}\,(V_{\overline{K}})^{\operatorname{Gal}(\overline{K}/K)} [FanoII, Proposition 2.3] (essentially due to [CTS21, Proposition 5.4.2]). Then it holds that

σ¯𝒪VK¯(σΓ¯)𝒪VK¯(Γ¯)¯\sigma^{*}\overline{\mathcal{L}}\simeq\mathcal{O}_{V_{\overline{K}}}(\sigma^{*}\overline{\Gamma})\simeq\mathcal{O}_{V_{\overline{K}}}(\overline{\Gamma})\simeq\overline{\mathcal{L}}

for every σGal(K¯/K)\sigma\in\operatorname{Gal}(\overline{K}/K), and hence ¯α\overline{\mathcal{L}}\simeq\alpha^{*}\mathcal{L} for some PicV\mathcal{L}\in\operatorname{Pic}\,V. Thus (i) holds. Let us show (ii). By the flat base change theorem, we obtain

H0(V,)KK¯H0(VK¯,L¯),H^{0}(V,\mathcal{L})\otimes_{K}\overline{K}\simeq H^{0}(V_{\overline{K}},\overline{L}),

which implies dimKH0(V,)=dimK¯H0(VK¯,¯)=dimK¯H0(VK¯,𝒪VK¯(Γ¯))=1\dim_{K}H^{0}(V,\mathcal{L})=\dim_{\overline{K}}H^{0}(V_{\overline{K}},\overline{\mathcal{L}})=\dim_{\overline{K}}H^{0}(V_{\overline{K}},\mathcal{O}_{V_{\overline{K}}}(\overline{\Gamma}))=1. In particular, there exists an effective Cartier divisor Γ\Gamma on VV such that 𝒪V(Γ)\mathcal{O}_{V}(\Gamma)\simeq\mathcal{L} and αΓ=Γ¯\alpha^{*}\Gamma=\overline{\Gamma}. Thus (ii) holds.

Since Γ¯=i=17Ei\overline{\Gamma}=\sum_{i=1}^{7}E_{i} is smooth, so is Γ\Gamma (note that Γ\Gamma might be irreducible, although Γ¯\overline{\Gamma} is not). Let φ:VW\varphi:V\to W be the contraction of Γ\Gamma, where WW is a smooth projective surface over KK. Then its base change φK¯:VK¯WK¯{\varphi}_{\overline{K}}:V_{\overline{K}}\to W_{\overline{K}} to K¯\overline{K} is the birational contraction of Γ¯\overline{\Gamma}, i.e., WK¯K¯2W_{\overline{K}}\simeq\mathbb{P}^{2}_{\overline{K}}. By Br(K)=0\operatorname{Br}(K)=0, we get a KK-isomorphism WK2W\simeq\mathbb{P}^{2}_{K} [CTS21, Proposition 7.1.6]. Via this isomorphism, we identify WW and K2\mathbb{P}^{2}_{K} (resp. WK¯W_{\overline{K}} and K¯2\mathbb{P}^{2}_{\overline{K}}):

φ:VW=K2,φK¯:VK¯WK¯=K¯2.\varphi:V\to W=\mathbb{P}^{2}_{K},\qquad{\varphi}_{\overline{K}}:V_{\overline{K}}\to W_{\overline{K}}=\mathbb{P}^{2}_{\overline{K}}.

Recall that φ\varphi is the blowup along some closed subscheme ZZ on W=K2W=\mathbb{P}^{2}_{K} [Har77, Ch. II, Theorem 7.17]. Since blowups commutes with flat base changes [Liu02, Section 8, Proposition 1.12(c)], φK¯\varphi_{\overline{K}} is the blowup along the base change ZK¯Z_{\overline{K}}. Since ZK¯Z_{\overline{K}} is a zero-dimensional reduced scheme consisting of 77 points, ZZ is a smooth zero-dimensional closed subscheme of K2\mathbb{P}^{2}_{K} satisfying h0(Z,𝒪Z)=7h^{0}(Z,\mathcal{O}_{Z})=7.

Set H:=Hilb𝔽2¯/𝔽¯227H:=\operatorname{Hilb}^{7}_{\mathbb{P}^{2}_{\overline{\mathbb{F}_{2}}/\overline{\mathbb{F}}_{2}}}, which is the Hilbert scheme of 𝔽¯22Spec𝔽¯2\mathbb{P}^{2}_{\overline{\mathbb{F}}_{2}}\to\operatorname{Spec}\overline{\mathbb{F}}_{2} that parametrises the zero-dimensional closed subschemes WW satisfying h0(W,𝒪W)=7h^{0}(W,\mathcal{O}_{W})=7. Let U:=Univ𝔽¯22/𝔽¯27U:=\operatorname{Univ}^{7}_{\mathbb{P}^{2}_{\overline{\mathbb{F}}_{2}}/\overline{\mathbb{F}}_{2}} be its universal family (cf. [FGI05, Section 5]): U𝔽¯22×𝔽¯2HHU\subset\mathbb{P}^{2}_{\overline{\mathbb{F}}_{2}}\times_{\overline{\mathbb{F}}_{2}}H\to H. Let HL(𝔽¯2)H(𝔽¯2)H_{L}(\overline{\mathbb{F}}_{2})\subset H(\overline{\mathbb{F}}_{2}) be the subset consisting of the Langer configurations over 𝔽¯2\overline{\mathbb{F}}_{2} (Definition 5.3(2)). Set G:=PGL3,𝔽¯2G:=\operatorname{PGL}_{3,\overline{\mathbb{F}}_{2}}, which is an algebraic group over 𝔽¯2\overline{\mathbb{F}}_{2}. By Lemma 5.4, HL(𝔽¯2)H_{L}(\overline{\mathbb{F}}_{2}) is equal to the G(𝔽¯2)G(\overline{\mathbb{F}}_{2})-orbit of [Z0]H(𝔽¯2)[Z_{0}]\in H(\overline{\mathbb{F}}_{2}), where

Z0:={[1:0:0],[0:1:0],[0:0:1],[1:1:0],[1:0:1],[0:1:1],[1:1:1]}𝔽¯22.Z_{0}:=\{[1:0:0],[0:1:0],[0:0:1],[1:1:0],[1:0:1],[0:1:1],[1:1:1]\}\subset\mathbb{P}^{2}_{\overline{\mathbb{F}}_{2}}.

Recall that the G(𝔽¯2)G(\overline{\mathbb{F}}_{2})-orbit HL(𝔽¯2)H_{L}(\overline{\mathbb{F}}_{2}) is a locally closed subset of H(𝔽¯2)H(\overline{\mathbb{F}}_{2}) [Mil16, Proposition 1.65(b)]. Since G=PGL3(𝔽¯2)G=\operatorname{PGL}_{3}(\overline{\mathbb{F}}_{2}) is irreducible, so is HL(𝔽¯2)H_{L}(\overline{\mathbb{F}}_{2}). There exists an integral locally closed subscheme HLH_{L} of HH whose set of the 𝔽¯2\overline{\mathbb{F}}_{2}-valued points coincides with HL(𝔽¯2)H_{L}(\overline{\mathbb{F}}_{2}). Set ULU×HHLU_{L}\coloneqq U\times_{H}H_{L}. We have the Langer configuration as the fibre of π:UH\pi\colon U\to H over [Z0][Z_{0}]:

{P1U,,P7U}=π1([Z0])U.\{P_{1}^{U},\ldots,P_{7}^{U}\}=\pi^{-1}([Z_{0}])\subset U.

Since G=PGL3,𝔽¯2G=\operatorname{PGL}_{3,\overline{\mathbb{F}}_{2}} equivariantly acts on π:UH\pi:U\to H, we have the orbits

OG(P1U),,OG(P7U)O_{G}(P_{1}^{U}),\ldots,O_{G}(P_{7}^{U})

of the above 77 points P1U,,P7UP_{1}^{U},\ldots,P_{7}^{U}. Since each OG(PiU)O_{G}(P_{i}^{U}) is a locally closed subset, this is a subvariety (integral scheme). Since any fibre of ULHLU_{L}\to H_{L} is geometrically reduced, it follows that ULU_{L} is reduced. Therefore, we get a scheme-theoretic equality

UL=OG(P1U)OG(P7U),U_{L}=O_{G}(P_{1}^{U})\amalg\cdots\amalg O_{G}(P_{7}^{U}),

because we have the corresponding set-theoretic equality.

Note that ZK2Z\subset\mathbb{P}^{2}_{K} corresponds to a KK-rational point [Z]HK(K)[Z]\in H_{K}(K) such that the corresponding K¯\overline{K}-rational point [ZK¯]HK¯(K¯)[Z_{\overline{K}}]\in H_{\overline{K}}(\overline{K}) is contained in (HL×𝔽¯2K¯)(K¯)(H_{L}\times_{\overline{\mathbb{F}}_{2}}\overline{K})(\overline{K}). Then the image

[Z]H(K)H[Z]\hookrightarrow H(K)\to H

is contained in the Langer locus HLH_{L} (over 𝔽¯2\overline{\mathbb{F}}_{2}). Therefore, Z[Z]Z\to[Z] is obtained by a base change of

ULHL,U_{L}\to H_{L},

and hence ZZ must be split up into 77 distinct points. ∎

5.1.4. C+EC+E (imprimitive)

Proposition 5.6.

Let XX be a smooth Fano threefold such that ρ(X)=2\rho(X)=2 and the types of the extremal rays are C+E1C+E_{1}. Then the following hold.

  1. (1)

    XX is quasi-FF-split.

  2. (2)

    If XX is not 2-9, then XX is FF-split.

Proof.

Let π:X2\pi:X\to\mathbb{P}^{2} (resp. g:XYg:X\to Y) be the contraction of the extremal ray of type CC (resp. E1E_{1}). Let SS be a general member of |π𝒪2(1)||\pi^{*}\mathcal{O}_{\mathbb{P}^{2}}(1)| and let HH be a general member of |g𝒪Y(1)||g^{*}\mathcal{O}_{Y}(1)|, where 𝒪Y(1)\mathcal{O}_{Y}(1) denotes the ample generator of PicY\operatorname{Pic}Y. Since Y=3,QY=\mathbb{P}^{3},Q, or VdV_{d} with 3d53\leq d\leq 5 [FanoIV, Subsection 7.2], the complete linear system |𝒪Y(1)||\mathcal{O}_{Y}(1)| is very ample. Note that HH is the blowup of a smooth surface H¯|𝒪Y(1)|\overline{H}\in|\mathcal{O}_{Y}(1)| along the zero-dimensional smooth closed subscheme BH¯B\cap\overline{H}, and hence HH is a smooth prime divisor. We have that

KXS+μH-K_{X}\sim S+\mu H

for the length μ\mu of the extremal ray of type CC [FanoIII, Proposition 5.9], i.e., if π\pi is of type CiC_{i} with i{1,2}i\in\{1,2\}, then μ=i\mu=i. Note that KH=(KX+H)|H=(S+(μ1)H)|H-K_{H}=-(K_{X}+H)|_{H}=(S+(\mu-1)H)|_{H} is nef and S2H=2iS^{2}\cdot H=\frac{2}{i} [FanoIII, Lemma 5.3 and Proposition 5.9(2)].

Step 1.

If π\pi is of type C2C_{2}, then XX is FF-split.

Proof of Step 1.

Assume that π\pi is of type C2C_{2}. In this case, XX is 2-27 or 2-31 [FanoIV, Subsection 7.2]. If XX is 2-27 (resp. 2-31), then X=BlCYX=\operatorname{Bl}_{C}\,Y for Y=3Y=\mathbb{P}^{3} (resp. Y=QY=Q), where CC is a smooth curve of degree 33 (resp. 11). Recall that KXS+2H-K_{X}\sim S+2H. We then get SH2=(KX)H22H3=(KY)𝒪Y(1)22𝒪Y(1)3=2S\cdot H^{2}=(-K_{X})\cdot H^{2}-2H^{3}=(-K_{Y})\cdot\mathcal{O}_{Y}(1)^{2}-2\mathcal{O}_{Y}(1)^{3}=2 in both cases. Since we have

KH2=(S+H)2H=S2H+2(SH2)+H3=1+4+H3,K_{H}^{2}=(S+H)^{2}\cdot H=S^{2}\cdot H+2(S\cdot H^{2})+H^{3}=1+4+H^{3},

it follows that HH is a smooth del Pezzo surface with KH2=6K_{H}^{2}=6 (resp. KH2=7K_{H}^{2}=7). Since (KX+H)S+H-(K_{X}+H)\sim S+H is ample, so is KH-K_{H}. Thus HH is FF-split (Theorem 2.22). Therefore, XX is FF-split (Corollary 2.9). This completes the proof of Step 1. ∎

Step 2.

Assume that π\pi is of type C1C_{1}. Then the following hold.

  1. (i)

    KH2=2K_{H}^{2}=2 and HH is a smooth weak del Pezzo surface.

  2. (ii)

    The induced composite morphism

    πH:HX𝜋2\pi_{H}\colon H\hookrightarrow X\xrightarrow{\pi}\mathbb{P}^{2}

    coincides with the morphism induced by the complete linear system |KH||-K_{H}|. Moreover, πH\pi_{H} is a generically finite morphism of degree two.

  3. (iii)

    If there exists a smooth prime divisor CC on HH satisfying CKHC\sim-K_{H}, then XX is quasi-FF-split.

Proof of Step 2.

It holds that

KH2=(KX+H)2H=S2H=2.K_{H}^{2}=(K_{X}+H)^{2}\cdot H=S^{2}\cdot H=2.

Thus (i) holds.

Let us show (ii). We have that π𝒪2(1)|HS|H(KX+H)|HKH\pi^{*}\mathcal{O}_{\mathbb{P}^{2}}(1)|_{H}\sim S|_{H}\sim-(K_{X}+H)|_{H}\sim-K_{H}. The Riemann–Roch theorem, together with (i), implies h0(H,KH)=KH2+1=3h^{0}(H,-K_{H})=K_{H}^{2}+1=3. By h0(H,KH)=3=h0(2,𝒪2(1))h^{0}(H,-K_{H})=3=h^{0}(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(1)), the composition πH:HX𝜋2\pi_{H}\colon H\hookrightarrow X\xrightarrow{\pi}\mathbb{P}^{2} coincides with the morphism induced by the complete linear system |KH||-K_{H}|. In particular, πH\pi_{H} is a generically finite morphism of degree two. Thus (ii) holds.

Let us show (iii). By (ii), we have the induced isomorphism:

H0(H,KH)πH,H0(2,𝒪2(1))H^{0}(H,-K_{H})\xleftarrow{\pi_{H}^{*},\simeq}H^{0}(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(1))

via πH:H2\pi_{H}\colon H\to\mathbb{P}^{2}, i.e., πH:|𝒪2(1)||KH|\pi_{H}^{*}:|\mathcal{O}_{\mathbb{P}^{2}}(1)|\to|-K_{H}| is bijective. By our assumption, there exists a line L|𝒪2(1)|L\in|\mathcal{O}_{\mathbb{P}^{2}}(1)| such that C:=πH(L)|KH|C:=\pi^{*}_{H}(L)\in|-K_{H}| is a smooth prime divisor. This property holds even after replacing LL by a general member of |𝒪2(1)||\mathcal{O}_{\mathbb{P}^{2}}(1)|, and hence we may assume that S=πLS=\pi^{*}L. We then obtain SH=πLH=πH(L)=CS\cap H=\pi^{*}L\cap H=\pi^{*}_{H}(L)=C, which is a smooth elliptic curve. By KX+S+H0K_{X}+S+H\sim 0, XX is quasi-FF-split (Corollary 2.18). Thus (iii) holds. This completes the proof of Step 2. ∎

Step 3.

Assume that π\pi is of type C1C_{1}. Then there exists a smooth prime divisor CC on HH satisfying CKHC\sim-K_{H}.

Proof of Step 3.

Recall that HH is a general member of |g𝒪Y(1)||g^{*}\mathcal{O}_{Y}(1)|. Suppose that any member of |KH||-K_{H}| is singular. By Step 2(iii), it is enough to derive a contradiction. Note that, for every line LL on 2\mathbb{P}^{2} and every smooth member H|g𝒪Y(1)|H\in|g^{*}\mathcal{O}_{Y}(1)|, its intersection HπLH\cap\pi^{*}L is not smooth (Step 2(ii)).

We now show that every smooth member of |H||H| is isomorphic to the Langer surface over kk (Definition 5.3). The Stein factorisation HH^{\prime} of the composition πH:HX𝜋2\pi_{H}:H\hookrightarrow X\xrightarrow{\pi}\mathbb{P}^{2} is the anti-canonical model of HH (Step 2(ii)). Note that each fibre of πH:HX𝜋2\pi_{H}:H\hookrightarrow X\xrightarrow{\pi}\mathbb{P}^{2} is contained in a fibre of π:X2\pi:X\to\mathbb{P}^{2}, which is a conic. In particular, any singularity on HH^{\prime} is either A1A_{1} or A2A_{2}. By KH2=2K_{H}^{2}=2 (Step 2(i)) and [KN, Theorem 1.4], it holds that p=2p=2 and HH is isomorphic to the Langer surface.

Fix two general members H1H_{1} and H2H_{2} of |H||H|. In particular, ΓH1H2\Gamma\coloneqq H_{1}\cap H_{2} is a smooth curve and each of H1H_{1} and H2H_{2} is isomorphic to the Langer surface. Let σ:YX\sigma:Y\to X be the blowup along Γ=H1H2\Gamma=H_{1}\cap H_{2}, and hence we get the morphism α:Y1\alpha\colon Y\to\mathbb{P}^{1} induced by the pencil generated by H1YH_{1}^{Y} and H2YH_{2}^{Y}, where HiYσHiEx(σ)H^{Y}_{i}\coloneqq\sigma^{*}H_{i}-\operatorname{Ex}(\sigma), which coincides with the proper transform of HiH_{i} on YY. By construction, every general fibre of α\alpha is isomorphic to the Langer surface (as otherwise we could find a smooth member of |H||H| which is not the Langer surface). Set VY×1SpecKV\coloneqq Y\times_{\mathbb{P}^{1}}\operatorname{Spec}K to be the generic fibre of α:Y1\alpha:Y\to\mathbb{P}^{1}, where KFrac1K\coloneqq\mathrm{Frac}\,\mathbb{P}^{1}.

For the algebraic closure K¯\overline{K} of KK, it is enough to show, by Lemma 5.5, that the base change VK¯V×SpecKSpecK¯V_{\overline{K}}\coloneqq V\times_{\operatorname{Spec}K}\operatorname{Spec}\overline{K} is isomorphic to the Langer surface over K¯\overline{K}. In fact, this implies 8=ρ(V)=ρ(Y×1SpecK)=ρ(Y)ρ(1)=18=\rho(V)=\rho(Y\times_{\mathbb{P}^{1}}\operatorname{Spec}K)=\rho(Y)-\rho(\mathbb{P}^{1})=1, which is a contradiction. Fix a general fibre VV^{\prime} of α:Y1\alpha:Y\to\mathbb{P}^{1}. Let F1,,FnF_{1},\ldots,F_{n} be all the (2)(-2)-curves on VK¯V_{\overline{K}} (FF is called a (2)(-2)-curve if F2=2F^{2}=-2 and F1F\simeq\mathbb{P}^{1}). Since F1,,FnF_{1},\ldots,F_{n} can be defined around the generic point of the base and VV^{\prime} is general, we obtain the corresponding (2)(-2)-curves F1,,FnF^{\prime}_{1},\ldots,F^{\prime}_{n} on VV^{\prime}. By the invariance of intersection numbers for flat families, we see that FiFj=FiFjF^{\prime}_{i}\cdot F^{\prime}_{j}=F_{i}\cdot F_{j} for every i,ji,j. Since VV^{\prime} is the Langer surface over kk, there are exactly 77 (2)(-2)-curves and they are mutually disjoint [CT18, Theorem 5.4]. Then we see that n7n\leq 7 and FiFj=0F^{\prime}_{i}\cdot F^{\prime}_{j}=0 (i.e., FiFj=F^{\prime}_{i}\cap F^{\prime}_{j}=\emptyset) for every 1i<jn1\leq i<j\leq n. As |KV||-K_{V^{\prime}}| has no smooth member, neither does |KVK¯||-K_{V_{\overline{K}}}|. By [KN, Theorem 1.4], we get n=7n=7, i.e., VK¯V_{\overline{K}} is isomorphic to the Langer surface over K¯\overline{K}. This completes the proof of Step 3. ∎

Step 1, Step 2, and Step 3 complete the proof of Proposition 5.6. ∎

5.2. Quasi-FF-splitting (primitive case)

In Section 5.1, the quasi-FF-splitting for smooth Fano threefolds with ρ=2\rho=2 has been settled for the imprimitive case. Hence the remaining cases are as follows [FanoIV, Subsection 7.2]:

2-2, 2-6, 2-8, 2-18, 2-24, 2-32, 2-34, 2-35, 2-36.\text{2-2, 2-6, 2-8, 2-18, 2-24, 2-32, 2-34, 2-35, 2-36}.

In what follows, we shall settle the cases except for 2-2, 2-6, and 2-8. These cases will be treated in Section 6.

Lemma 5.7.

If XX is a smooth Fano threefold of No. 2-32, 2-34, 2-35, or 2-36, then XX is FF-split.

Proof.

It is well known that XX is toric except when it is 2-32. Assume that XX is 2-32, i.e., XX is a smooth hypersurface on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,1)(1,1). Note that f1:X2×2pr12f_{1}\colon X\hookrightarrow\mathbb{P}^{2}\times\mathbb{P}^{2}\xrightarrow{\operatorname{pr}_{1}}\mathbb{P}^{2} is a 1\mathbb{P}^{1}-bundle [FanoIV, Subsection 7.2]. Take a general member D|f1𝒪2(1)|D\in|f_{1}^{*}\mathcal{O}_{\mathbb{P}^{2}}(1)|, which is a 1\mathbb{P}^{1}-bundle over 1\mathbb{P}^{1}. Hence DD is FF-split. As in the proof of Proposition 4.4(2), we can see (KX+D)-(K_{X}+D) is ample by [FanoIII, Proposition 5.9(3)] and [FanoIV, Section 7.2]. Therefore, XX is FF-split (Corollary 2.9). ∎

Lemma 5.8.

Let YY be a smooth Fano threefold of No. 2-18. Then the following hold.

  1. (1)

    YY is FF-split.

  2. (2)

    Let BB be a smooth fibre of the contraction g:Y2g\colon Y\to\mathbb{P}^{2} of type C1C_{1}. Then the blowup XBlBYX\coloneqq\operatorname{Bl}_{B}Y of YY along BB is a smooth Fano threefold of No. 3-4.

The following argument is almost identical to that of [FanoIV, Proposition 4.35].

Proof.

By Proposition 4.4, (2) implies (1). Let us show (2). It is enough to show that KX-K_{X} is ample [FanoIV, Subsection 7.2]. By construction, we have the following commutative diagram except for f1,g11,g12f_{1},g_{11},g_{12}. Since φ1×φ2:XZ1×Z2=1×1\varphi_{1}\times\varphi_{2}\colon X\to Z_{1}\times Z_{2}=\mathbb{P}^{1}\times\mathbb{P}^{1} is not a finite morphism, its Stein factorisation f1:XSf_{1}\colon X\to S of φ1×φ2\varphi_{1}\times\varphi_{2} is not an isomorphism. We get dimS=2\dim S=2, because we have dimSdim(1×1)=2\dim S\leq\dim(\mathbb{P}^{1}\times\mathbb{P}^{1})=2, and S1×1pr11S\to\mathbb{P}^{1}\times\mathbb{P}^{1}\xrightarrow{\operatorname{pr}_{1}}\mathbb{P}^{1} is not an isomorphism. Then f1:XSf_{1}\colon X\to S is a contraction of a (possibly non KXK_{X}-negative) extremal ray. Since this extramal ray is contained in the two-dimensional extremal faces corresponding to φ1\varphi_{1} and φ2\varphi_{2}, NE(X)\operatorname{NE}(X) is generated by three extremal rays, which are corresponding to f1,f2,f3f_{1},f_{2},f_{3}.

XXSS𝔽1\mathbb{F}_{1}YY1\mathbb{P}^{1}1\mathbb{P}^{1}2\mathbb{P}^{2}f1\scriptstyle f_{1}f2\scriptstyle f_{2}f3\scriptstyle f_{3}φ1\scriptstyle\varphi_{1}φ2\scriptstyle\varphi_{2}φ3\scriptstyle\varphi_{3}g11\scriptstyle g_{11}g12\scriptstyle g_{12}g22\scriptstyle g_{22}g23=τ\scriptstyle g_{23}=\taug33\scriptstyle g_{33}g31\scriptstyle g_{31}

By the same argument as in [FanoIV, Proposition 4.35], we obtain

KXH1+H2+H3,-K_{X}\sim H_{1}+H_{2}+H_{3},

where each HiH_{i} is the pullback of the ample generator by φi\varphi_{i}. Since φ1×φ2×φ3:X1×1×2\varphi_{1}\times\varphi_{2}\times\varphi_{3}\colon X\to\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{2} is a finite morphism, KXH1+H2+H3-K_{X}\sim H_{1}+H_{2}+H_{3} is ample, as required. ∎

Lemma 5.9.

Let YY be a smooth Fano threefold of No. 2-24. Assume that the contraction ψ:Y2\psi\colon Y\to\mathbb{P}^{2} of type C1C_{1} is generically smooth. Then the following hold.

  1. (1)

    YY is FF-split.

  2. (2)

    Let BB be a smooth fibre of ψ:Y2\psi\colon Y\to\mathbb{P}^{2}. Then the blowup X:=BlBYX:=\operatorname{Bl}_{B}Y of YY along BB is a Fano threefold of No. 3-4.

The following argument is almost identical to that of [FanoIV, Proposition 4.40].

Proof.

Since (2) implies (1), it is enough to show (2). Let us show (2). It is enough to show that KX-K_{X} is ample [FanoIV, Subsection 7.2]. By construction, we get the following commutative diagram except for f3,g31,g33f_{3},g_{31},g_{33}. By the same argument as in [FanoIV, Proposition 4.40], we see that

  • φ3×φ1:X2×1\varphi_{3}\times\varphi_{1}:X\to\mathbb{P}^{2}\times\mathbb{P}^{1} is not a finite morphism, and

  • KXH1+H2+H3-K_{X}\sim H_{1}+H_{2}+H_{3},

where each HiH_{i} is the pullback of the ample generator by φi\varphi_{i}. Then XX has exactly three extremal rays and KX-K_{X} is ample.

XX𝔽1\mathbb{F}_{1}YY2×1\mathbb{P}^{2}\times\mathbb{P}^{1}1\mathbb{P}^{1}2\mathbb{P}^{2}2\mathbb{P}^{2}f1\scriptstyle f_{1}f2\scriptstyle f_{2}f3\scriptstyle f_{3}φ1\scriptstyle\varphi_{1}φ2\scriptstyle\varphi_{2}φ3\scriptstyle\varphi_{3}g11\scriptstyle g_{11}g12=τ\scriptstyle g_{12}=\taug22\scriptstyle g_{22}g23\scriptstyle g_{23}g33\scriptstyle g_{33}g31\scriptstyle g_{31}

Lemma 5.10.

Let YY be a smooth Fano threefold of No. 2-24. Assume that the contraction ψ:Y2\psi\colon Y\to\mathbb{P}^{2} of type C1C_{1} is not generically smooth. Then the following hold.

  1. (1)

    XX is isomorphic to {x0y02+x1y12+x2y22=0}x2×y2\{x_{0}y_{0}^{2}+x_{1}y_{1}^{2}+x_{2}y_{2}^{2}=0\}\subset\mathbb{P}^{2}_{x}\times\mathbb{P}^{2}_{y}, where x2Projk[x0,x1,x2](2)\mathbb{P}^{2}_{x}\coloneqq\mathrm{Proj}\,k[x_{0},x_{1},x_{2}](\simeq\mathbb{P}^{2}) and y2Projk[y0,y1,y2](2)\mathbb{P}^{2}_{y}\coloneqq\mathrm{Proj}\,k[y_{0},y_{1},y_{2}](\simeq\mathbb{P}^{2}).

  2. (2)

    XX is quasi-FF-split.

Proof.

Since (1) implies (2) [KTY, Example 7.13], it is enough to show (1). Recall that ψ\psi can be written as follows:

ψ:Xx2×y2pr1x2.\psi:X\hookrightarrow\mathbb{P}^{2}_{x}\times\mathbb{P}^{2}_{y}\xrightarrow{\operatorname{pr}_{1}}\mathbb{P}^{2}_{x}.

Since every fibre of ψ\psi is a non-reduced conic, we can write

X={f0(x)y02+f1(x)y12+f2(x)y22=0},X=\{f_{0}(x)y_{0}^{2}+f_{1}(x)y_{1}^{2}+f_{2}(x)y_{2}^{2}=0\},

where each fi(x)k[x0,x1,x2]f_{i}(x)\in k[x_{0},x_{1},x_{2}] is a homogeneous polynomial of degree 11. We see that

  1. ()(*)

    none of f0(x),f1(x),f2(x)f_{0}(x),f_{1}(x),f_{2}(x) is zero.

Indeed, if f2(x)=0f_{2}(x)=0, then the affine open subset of XX defined by {x00}{y20}\{x_{0}\neq 0\}\cap\{y_{2}\neq 0\} contains a singular point. After applying a suitable coordinate change, we may assume that f0(x)=x0f_{0}(x)=x_{0}. Therefore, we can write

X={x0y02+f1(x)y12+f2(x)y22=0}.X=\{x_{0}y_{0}^{2}+f_{1}(x)y_{1}^{2}+f_{2}(x)y_{2}^{2}=0\}.

We can write f1(x)=ax0+bx1+cx2f_{1}(x)=ax_{0}+bx_{1}+cx_{2}. By applying y0y0+ay1y_{0}\mapsto y_{0}+\sqrt{a}y_{1}, we may assume that a=0a=0. Note that ()(*) implies f1(x)0f_{1}(x)\neq 0. Replacing f1(x)=bx1+cx2(0)f_{1}(x)=bx_{1}+cx_{2}(\neq 0) by x1x_{1}, we may assume that f1(x)=x1f_{1}(x)=x_{1}, i.e.,

X={x0y02+x1y12+f2(x)y22=0}.X=\{x_{0}y_{0}^{2}+x_{1}y_{1}^{2}+f_{2}(x)y_{2}^{2}=0\}.

Similarly, by applying a coordinate change y0y0+dy2,y1y1+ed3y_{0}\mapsto y_{0}+dy_{2},y_{1}\mapsto y_{1}+ed_{3} for some d,ekd,e\in k, the problem is reduced to the case when f2(x)=αx2f_{2}(x)=\alpha x_{2} for some αk\alpha\in k. By ()(*), we may assume that α=1\alpha=1. Thus (1) holds. ∎

Remark 5.11.

Let XX be a smooth Fano threefold of No. 2-24. Combining Lemma 5.9, Lemma 5.10, and [KTY, Example 7.13], XX is 22-quasi-FF-split. Moreover, the following hold.

  1. (1)

    The following are equivalent.

    1. (a)

      The quasi-FF-split height is 11, i.e., XX is FF-split.

    2. (b)

      The contraction ψ:X2\psi:X\to\mathbb{P}^{2} of type C1C_{1} is generically smooth.

  2. (2)

    The following are equivalent.

    1. (a)

      The quasi-FF-split height is 22.

    2. (b)

      The contraction ψ:X2\psi:X\to\mathbb{P}^{2} of type C1C_{1} is not generically smooth (called wild).

    3. (c)

      X{x0y02+x1y12+x2y22=0}x2×y2X\simeq\{x_{0}y_{0}^{2}+x_{1}y_{1}^{2}+x_{2}y_{2}^{2}=0\}\subset\mathbb{P}^{2}_{x}\times\mathbb{P}^{2}_{y}.

Proposition 5.12.

Let XX be a smooth Fano threefold with ρ(X)=2\rho(X)=2. Then XX is quasi-FF-split if XX is none of 2-2, 2-6, and 2-8.

Proof.

If XX is imprimitive (resp. primitive), then the assertion follows from Proposition 5.1, Proposition 5.2, and Proposition 5.6 (resp. Section 5.2). ∎

5.3. FF-splitting

Theorem 5.13.

Assume p>5p>5. Let XX be a smooth Fano threefold with ρ(X)2\rho(X)\geq 2. If XX is neither 2-2 nor 2-6, then XX is FF-split.

Proof.

By Proposition 3.1, Proposition 3.2, Proposition 3.3, and Proposition 4.4, we may assume that ρ(X)=2\rho(X)=2. The types R1+R2R_{1}+R_{2} of the extremal rays R1R_{1} and R2R_{2} are as follows, because the case D+DD+D does not occur [FanoIV, Subsection 7.2]:

  1. (I)

    C+EC+E or D+ED+E.

  2. (II)

    E+EE+E.

  3. (III)

    C+DC+D.

  4. (IV)

    C+CC+C.

For each i{1,2}i\in\{1,2\}, let fi:XYif_{i}:X\to Y_{i} be the contraction of RiR_{i}, μi\mu_{i} denotes the length of RiR_{i}, and set HiH_{i} to be the pullback of the ample generator of PicYi()\operatorname{Pic}\,Y_{i}(\simeq\mathbb{Z}). Recall that we can write KXμ2H1+μ1H2-K_{X}\sim\mu_{2}H_{1}+\mu_{1}H_{2} [FanoIII, Proposition 5.9]. In what follows, we treat the above four cases separately.

(I) Since R1+R2R_{1}+R_{2} is C+EC+E or D+ED+E, we have Y1=1Y_{1}=\mathbb{P}^{1} or Y1=2Y_{1}=\mathbb{P}^{2}. Pick a general member SS of |H1||H_{1}|.

Claim.

SS is a canonical weak del Pezzo surface.

Proof of Claim.

We first treat the case when Y1=1Y_{1}=\mathbb{P}^{1}. Let SS^{\prime} be the geometric generic fibre of f1:XY1=1f_{1}\colon X\to Y_{1}=\mathbb{P}^{1}. Then SS^{\prime} is normal and KS-K_{S^{\prime}} is ample [FS20, Theorem 15.2], which implies that SS^{\prime} is canonical [BT22, Theorem 3.3]. Hence SS is a canonical del Pezzo surface for the case when Y1=1Y_{1}=\mathbb{P}^{1}.

It is enough to settle the case when Y=2Y=\mathbb{P}^{2}. In this case, SS is the inverse image of a general line LL on 2\mathbb{P}^{2}. Since the discriminant scheme Δf1\Delta_{f_{1}} of f1:XS=2f_{1}\colon X\to S=\mathbb{P}^{2} is a reduced divisor [Tan-conic, Proposition 7.2], the scheme theoretic intersection LΔf1L\cap\Delta_{f_{1}} is a zero-dimensional smooth scheme. For the resulting conic bundle g:S=X×Y1LLg\colon S=X\times_{Y_{1}}L\to L, we have Δg=LΔf1\Delta_{g}=L\cap\Delta_{f_{1}} [Tan-conic, Remark 3.4]. Since LL and Δg\Delta_{g} are smooth, also SS is smooth [Tan-conic, Theorem 4.4]. We have that KXS+H-K_{X}\sim S+H for Hμ1H2+(μ21)H1H\coloneqq\mu_{1}H_{2}+(\mu_{2}-1)H_{1}. In particular, HH and H|SH|_{S} are nef and big. By the adjunction formula: KS(KX+S)|SH|SK_{S}\sim(K_{X}+S)|_{S}\sim-H|_{S}, we have SS is a smooth weak del Pezzo surface. This completes the proof of Claim. ∎

By Claim, SS is FF-split (Theorem 2.22). We have KXS+H-K_{X}\sim S+H for Hμ1H2+(μ21)H1H\coloneqq\mu_{1}H_{2}+(\mu_{2}-1)H_{1}. By H1(X,S+(pe1)(KX+S))=H1(X,KX+peH)H^{1}(X,-S+(p^{e}-1)(K_{X}+S))=H^{1}(X,K_{X}+p^{e}H) and Proposition 2.8, it is enough to find e>0e\in\mathbb{Z}_{>0} such that

H1(X,KX+peH)=0H^{1}(X,K_{X}+p^{e}H)=0

by. Fix e1>0e_{1}\in\mathbb{Z}_{>0} such that pe1HDp^{e_{1}}H-D is ample for D:=Ex(f2)D:=\operatorname{Ex}(f_{2}). By the Serre vanishing theorem, there is e2>0e_{2}\in\mathbb{Z}_{>0} such that H1(X,KX+pe2(pe1HD))=0H^{1}(X,K_{X}+p^{e_{2}}(p^{e_{1}}H-D))=0. Set e:=e1+e2e:=e_{1}+e_{2}. It suffices to prove

H1(X,KX+peHsD)=0H^{1}(X,K_{X}+p^{e}H-sD)=0

for every 0spe20\leq s\leq p^{e_{2}} by descending induction on ss. The base case s=pe2s=p^{e_{2}} has been checked already. Fix an integer ss satisfying 0s<pe20\leq s<p^{e_{2}}. By the induction hypothesis, we have the following exact sequence

0=H1(X,KX+peH(s+1)D)H1(X,KX+peHsD)H1(D,KD+(peH(s+1)D)|D).0=H^{1}(X,K_{X}+p^{e}H-(s+1)D)\to H^{1}(X,K_{X}+p^{e}H-sD)\\ \to H^{1}(D,K_{D}+(p^{e}H-(s+1)D)|_{D}).

It suffices to show H1(D,KD+(peH(s+1)D)|D)=0H^{1}(D,K_{D}+(p^{e}H-(s+1)D)|_{D})=0. Note that peH(s+1)Dp^{e}H-(s+1)D is ample, because so are HH and peHpe2D(=pe2(pe1HD)p^{e}H-p^{e_{2}}D(=p^{e_{2}}(p^{e_{1}}H-D)). Then we get H1(D,KD+(peH(s+1)D)|D)=0H^{1}(D,K_{D}+(p^{e}H-(s+1)D)|_{D})=0 by the fact that DD is toric or a smooth ruled surface [Muk13, Theorem 3].

(II) Assume that R1+R2R_{1}+R_{2} is E+EE+E. The list of such Fano threefolds is as follows [FanoIV, Subsection 7.2]: 2-12, 2-15 2-17, 2-19, 2-21, 2-22, 2-23, 2-26, 2-28, 2-30. We treat the following two cases separately:

  1. (1)

    2-12, 2-15, 2-17, 2-19, 2-22, 2-28, 2-30.

  2. (2)

    2-21, 2-23, 2-26.

If (1) (resp. (2)) holds, then there is a blowup f1:XY1=3f_{1}:X\to Y_{1}=\mathbb{P}^{3} (resp. f1:XY1=Qf_{1}:X\to Y_{1}=Q) along a smooth curve BB on Y1Y_{1}. Note that H1H_{1} is the inverse image of the corresponding member H¯1\overline{H}_{1} on Y1Y_{1}. After replacing H1H_{1} by a general member of |H1||H_{1}|, we may assume that H1H_{1} is the blowup along a smooth zero-dimensional scheme BH¯1B\cap\overline{H}_{1} of a smooth surface H¯1\overline{H}_{1}, and hence H1H_{1} is a smooth projective surface. We have that KXH1+H2-K_{X}\sim H_{1}+H^{\prime}_{2} for some H2H2+NH^{\prime}_{2}\sim H_{2}+N with NN nef. Then it holds that KH1-K_{H_{1}} is nef and big. Hence H1H_{1} is a smooth weak del Pezzo surface. The same argument as in (I) deduces that XX is FF-split.

(III) Assume that the types of the extremal rays are C+DC+D. Since we are assuming that XX is not 2-2, XX is 2-18 or 2-34 [FanoIV, Subsection 7.2]. If XX is 2-34, i.e., X1×2X\simeq\mathbb{P}^{1}\times\mathbb{P}^{2}), then XX is clearly FF-split. The case when XX is 2-18 has been settled in Lemma 5.8.

(IV) Assume that R1+R2R_{1}+R_{2} is C+CC+C. Since we are assuming that XX is not 2-6, XX is 2-24 or 2-32. If XX is 2-24 (resp. 2-32), then XX is FF-split by Lemma 5.9 (resp. Lemma 5.7). Note that an arbitrary conic bundle XSX\to S is generically smooth by p>2p>2. ∎

6. FF-splitting and Quasi-FF-splitting via Cartier operators

6.1. Quasi-FF-splitting for 2-2, 2-6, 2-8, and 3-10

6.1.1. Preparation

Lemma 6.1.

Let XX be a smooth Fano threefold. Assume that XX is SRC. Then H0(X,ΩXi(D))=0H^{0}(X,\Omega_{X}^{i}(-D))=0 for every i>0i>0 and every pseudo-effective Cartier divisor DD on XX.

For the definition of SRC (separable rational connectedness), we refer to [Kol96].

Proof.

The assertion follows from the essentially same proof as in [Kaw1, Proposition 3.4] by using the fact that the restriction of a pseudo-effective Cartier divisor DD to a general curve is pseudo-effective. ∎

We repeatedly use the following basic lemma.

Lemma 6.2.

Let XX be a smooth Fano threefold. Assume that there exists a conic bundle f:XSf:X\to S, i.e., ff is a flat morphism to a smooth projective surface SS such that every fibre f1(s)f^{-1}(s) is isomorphic to a conic (cf. [Tan-conic, Definition 2.3]). Then the following hold.

  1. (1)

    XX is SRC.

  2. (2)

    H0(X,ΩXi(D))=0H^{0}(X,\Omega_{X}^{i}(-D))=0 for every i>0i>0 and every pseudo-effective Cartier divisor DD on XX.

In particular, if XX is one of No. 2-2, 2-6, 2-8, and 3-10, then XX satisfies the condition (1) in Proposition 2.20.

Proof.

Since (1) implies (2) (Lemma 6.1), it is enough to show (1). If ff is generically smooth, then SS is a smooth rational surface [FanoIV, Proposition 3.13], and hence XX is SRC by [GLP15, Theorem 0.5]. We may assume that ff is wild, i.e., not generically smooth. Then XX is 2-24 or 3-10 by [MS03, Corollary 8]. Assume that XX is 2-24. Then the contraction of the other extremal ray is of type CC and its contraction gives a generically smooth conic bundle structure. We are done by the generically smooth case. Assume that XX is 3-10. Then XX is obtained as a blowup of QQ (cf. [MS03, Corollary 8] or [FanoIV, Section 7]). Since QQ is rational (and hence SRC), so is XX. Thus (1) holds. ∎

6.3Double covers.

Given smooth projective varieties XX and YY, we say that f:XYf\colon X\to Y is a double cover if ff is a finite surjective morphism such that the induced field extension K(X)/K(Y)K(X)/K(Y) is of degree 22. Recall that we have an exact sequence

0𝒪Yf𝒪X𝒪Y(L)00\to\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X}\to\mathcal{O}_{Y}(-L)\to 0

for some Cartier divisor LL on YY [Kaw2, Lemma A.1]. In particular, 𝒪Y(L)(f𝒪X/𝒪Y)1\mathcal{O}_{Y}(L)\simeq(f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1}. Moreover, KXf(KY+L)K_{X}\sim f^{*}(K_{Y}+L) [CD89, Proposition 0.1.3]. We say a double cover f:XYf\colon X\to Y is split if the induced homomorphism 𝒪Yf𝒪X\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X} splits as an 𝒪Y\mathcal{O}_{Y}-module homomorphism (i.e., the above exact sequence splits). In this case, we obtain f𝒪X𝒪Y𝒪Y(L)f_{*}\mathcal{O}_{X}\simeq\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L).

Lemma 6.4.

Let f:XYf:X\to Y be a split double cover of smooth projective varieties. Let LL be a Cartier divisor LL on YY satisfying 𝒪Y(L)f𝒪X/𝒪Y\mathcal{O}_{Y}(-L)\simeq f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y} (cf. (6.3)). Then there exists a closed immersion j:XP:=Y(𝒪Y𝒪Y(L))j\colon X\hookrightarrow P:=\mathbb{P}_{Y}(\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L)) which satisfies the following properties.

  1. (1)

    KP+Xg(KY+L)K_{P}+X\sim g^{*}(K_{Y}+L), where g:P=Y(𝒪Y𝒪Y(L))Yg:P=\mathbb{P}_{Y}(\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L))\to Y denotes the induced 1\mathbb{P}^{1}-bundle.

  2. (2)

    For the section S:=Y(𝒪Y(L))S:=\mathbb{P}_{Y}(\mathcal{O}_{Y}(-L)) of g:P=Y(𝒪Y𝒪Y(L))Yg\colon P=\mathbb{P}_{Y}(\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L))\to Y corresponding to the second projection 𝒪Y𝒪Y(L)𝒪Y(L)\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L)\to\mathcal{O}_{Y}(-L), it holds that SX=S\cap X=\emptyset, S|SgL|SS|_{S}\sim-g^{*}L|_{S}, X2S2gLX-2S\sim 2g^{*}L, and 𝒪P(1)𝒪P(S)\mathcal{O}_{P}(1)\simeq\mathcal{O}_{P}(S),

  3. (3)

    For the section TY(𝒪Y)T\coloneqq\mathbb{P}_{Y}(\mathcal{O}_{Y}) of g:P=Y(𝒪Y𝒪Y(L))Yg:P=\mathbb{P}_{Y}(\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L))\to Y corresponding to the first projection 𝒪Y𝒪Y(L)𝒪Y\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L)\to\mathcal{O}_{Y}, it holds that ST=S\cap T=\emptyset, T|TgL|TT|_{T}\sim g^{*}L|_{T}, X2TX\sim 2T, TSgLT-S\sim g^{*}L, and 𝒪P(1)𝒪P(TgL)\mathcal{O}_{P}(1)\simeq\mathcal{O}_{P}(T-g^{*}L).

  4. (4)

    ΩP/Y1𝒪P(gL2S)\Omega_{P/Y}^{1}\simeq\mathcal{O}_{P}(-g^{*}L-2S).

Proof.

In what follows, we only treat the case when p=2p=2, as otherwise the problem is easier. By [CD89, the proof of Proposition 0.1.3] (note that ff splits if and only if ff corresponds to a splittable admissible triple), there is a closed immersion j:XPj^{\circ}:X\hookrightarrow P^{\circ} to the 𝔸1\mathbb{A}^{1}-bundle

P:=SpecY(d=0𝒪Y(dL)).P^{\circ}:=\operatorname{Spec}_{Y}(\bigoplus_{d=0}^{\infty}\mathcal{O}_{Y}(-dL)).

Since PP^{\circ} is an open subscheme of PP, we obtain a closed immersion j:XPj:X\to P over YY. By definition, we get ST=S\cap T=\emptyset, 𝒪P(1)|S𝒪S(L)\mathcal{O}_{P}(1)|_{S}\simeq\mathcal{O}_{S}(-L), and 𝒪P(1)|T𝒪T\mathcal{O}_{P}(1)|_{T}\simeq\mathcal{O}_{T}, where (g|S)L(g|_{S})^{*}L denotes LL for g|S:SYg|_{S}:S\xrightarrow{\simeq}Y by abuse of notation. We can write 𝒪P(1)𝒪P(S+gDS)𝒪P(T+gDT)\mathcal{O}_{P}(1)\simeq\mathcal{O}_{P}(S+g^{*}D_{S})\simeq\mathcal{O}_{P}(T+g^{*}D_{T}) for some Cartier divisors DSD_{S} and DTD_{T} on YY. By ST=S\cap T=\emptyset and 𝒪P(1)|T𝒪T\mathcal{O}_{P}(1)|_{T}\simeq\mathcal{O}_{T}, we obtain 𝒪T𝒪P(1)|T𝒪P(S+gDS)|T(g𝒪Y(DS))|T\mathcal{O}_{T}\simeq\mathcal{O}_{P}(1)|_{T}\simeq\mathcal{O}_{P}(S+g^{*}D_{S})|_{T}\simeq(g^{*}\mathcal{O}_{Y}(D_{S}))|_{T}, which implies DS0D_{S}\sim 0. Similarly, we obtain DTLD_{T}\sim-L by 𝒪S(L)𝒪P(1)|S𝒪P(T+gDT)|S(g𝒪Y(DT))|S\mathcal{O}_{S}(-L)\simeq\mathcal{O}_{P}(1)|_{S}\simeq\mathcal{O}_{P}(T+g^{*}D_{T})|_{S}\simeq(g^{*}\mathcal{O}_{Y}(D_{T}))|_{S}. Hence we get

𝒪P(1)STgL,\mathcal{O}_{P}(1)\sim S\sim T-g^{*}L,

which implies

S|SgL|S,T|TgL|T.S|_{S}\simeq-g^{*}L|_{S},\qquad T|_{T}\sim g^{*}L|_{T}.
Claim.

The following hold.

  1. (a)

    𝒪P(KP)𝒪P(2)g𝒪Y(2L)g𝒪Y(KY+L)\mathcal{O}_{P}(K_{P})\otimes\mathcal{O}_{P}(2)\otimes g^{*}\mathcal{O}_{Y}(2L)\simeq g^{*}\mathcal{O}_{Y}(K_{Y}+L).

  2. (b)

    XS=X\cap S=\emptyset.

  3. (c)

    KP+Xg(KY+L)K_{P}+X\sim g^{*}(K_{Y}+L).

Proof of Claim.

Let us show (c)’ below, which is weaker than (c):

  1. (c)’

    KP+Xg(KY+L)K_{P}+X\equiv g^{*}(K_{Y}+L), where \equiv denotes the numerical equivalence.

Since f:XYf:X\to Y is a double cover, we can find a Cartier divisor EE on YY such that KP+XgEK_{P}+X\sim g^{*}E. Then

fE=(gE)|X(KP+X)|XKXf(KY+L),f^{*}E=(g^{*}E)|_{X}\sim(K_{P}+X)|_{X}\sim K_{X}\sim f^{*}(K_{Y}+L),

which implies EKY+LE\equiv K_{Y}+L [Kle66, Corollary 1(ii) in page 304]. Thus (c)’ holds.

The assertion (a) holds by the following (cf. [FanoIII, Proposition 7.1(2)]):

𝒪P(KP)𝒪P(2)g(ωYdet(𝒪Y𝒪Y(L)))𝒪P(2)g𝒪Y(KYL).\mathcal{O}_{P}(K_{P})\simeq\mathcal{O}_{P}(-2)\otimes g^{*}(\omega_{Y}\otimes\det(\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L)))\simeq\mathcal{O}_{P}(-2)\otimes g^{*}\mathcal{O}_{Y}(K_{Y}-L).

Let us show (b). By (a) and (c)’, we obtain X𝒪P(2)+2gL2(S+gL)X\equiv\mathcal{O}_{P}(2)+2g^{*}L\sim 2(S+g^{*}L). This, together with S|SgL|SS|_{S}\sim-g^{*}L|_{S}, implies X|S0X|_{S}\equiv 0. By XSX\neq S, we obtain XS=X\cap S=\emptyset, as otherwise we could find a curve CC on SS which properly intersects XX. Thus (b) holds. Then it holds that

gE|S(KP+X)|S(b)KP|SKSS|Sg(KY+L)|S,g^{*}E|_{S}\sim(K_{P}+X)|_{S}\overset{{\rm(b)}}{\sim}K_{P}|_{S}\sim K_{S}-S|_{S}\sim g^{*}(K_{Y}+L)|_{S},

which implies EKY+LE\sim K_{Y}+L, i.e., (c) holds. This completes the proof of Claim. ∎

We can write X2SgFX-2S\sim g^{*}F for some Cartier divisor FF on YY. By 2gL|S(X2S)|LgF|S-2g^{*}L|_{S}\sim(X-2S)|_{L}\sim g^{*}F|_{S}, we obtain F2LF\sim 2L. Then X2S+2gL2TX\sim 2S+2g^{*}L\sim 2T. This completes the proofs of (2) and (3).

Let us show (4). We have an exact sequence

0gΩY1𝛼ΩP1ΩP/Y10,0\to g^{*}\Omega_{Y}^{1}\xrightarrow{\alpha}\Omega^{1}_{P}\to\Omega_{P/Y}^{1}\to 0,

where the injectivity of α\alpha can be checked by taking the corresponding sequence of the stalks at the generic point. Taking the wedge products, we get ωPgωYΩP/Y1\omega_{P}\simeq g^{*}\omega_{Y}\otimes\Omega_{P/Y}^{1}. By KP+Xg(KY+L)K_{P}+X\sim g^{*}(K_{Y}+L) and X2S+2gLX\sim 2S+2g^{*}L, we obtain

ΩP/Y1𝒪P(KPgKY)𝒪P(X+gL)𝒪P(gL2S).\Omega_{P/Y}^{1}\simeq\mathcal{O}_{P}(K_{P}-g^{*}K_{Y})\simeq\mathcal{O}_{P}(-X+g^{*}L)\simeq\mathcal{O}_{P}(-g^{*}L-2S).

Thus (4) holds. ∎

Lemma 6.5.

We use the same notation as Lemma 6.4. Fix qq\in\mathbb{Z} and take a Cartier divisor DD on YY. Assume that

  1. (1)

    Hq1(Y,D)=Hq1(Y,DL)=0H^{q-1}(Y,D)=H^{q-1}(Y,D-L)=0, and

  2. (2)

    Hq(Y,ΩY1(D))=Hq(Y,ΩY1(DL))=0H^{q}(Y,\Omega_{Y}^{1}(D))=H^{q}(Y,\Omega_{Y}^{1}(D-L))=0.

Then Hq(P,ΩP1(gD))=0H^{q}(P,\Omega_{P}^{1}(g^{*}D))=0.

Proof.

We have an exact sequence

0g(ΩY1(D))ΩP1(gD)ΩP/Y1(gD)0.0\to g^{*}(\Omega_{Y}^{1}(D))\to\Omega_{P}^{1}(g^{*}D)\to\Omega^{1}_{P/Y}(g^{*}D)\to 0.

By (2), it is enough to show Hq(P,ΩP/Y1(gD))=0H^{q}(P,\Omega^{1}_{P/Y}(g^{*}D))=0, i.e., Hq(P,gDgL2S)=0H^{q}(P,g^{*}D-g^{*}L-2S)=0. Using an exact sequence 0𝒪P(S)𝒪P𝒪S00\to\mathcal{O}_{P}(-S)\to\mathcal{O}_{P}\to\mathcal{O}_{S}\to 0 twice, it suffices to prove that Hq(P,gDgL))=0H^{q}(P,g^{*}D-g^{*}L))=0 and Hq1(S,gDgLnS)=0H^{q-1}(S,g^{*}D-g^{*}L-nS)=0 with n{0,1}n\in\{0,1\}. By S|SgL|S-S|_{S}\sim g^{*}L|_{S}, these equalities follow from (2) and (1), respectively. ∎

Proposition 6.6.

We use the same notation of Lemma 6.4. Assume that dimX=dimY=3\dim X=\dim Y=3, and both LL and HKYLH\coloneqq-K_{Y}-L are ample. Moreover, suppose that the following hold.

  1. (0)
    1. (0a)

      Hj(Y,A))=0H^{j}(Y,-A))=0 for every j<3j<3 and every ample Cartier divisor AA on YY.

    2. (0b)

      Hj(Y,𝒪Y(piH))=Hj(Y,𝒪Y(piHL))=Hj(Y,𝒪Y(piH2L))=0H^{j}(Y,\mathcal{O}_{Y}(p^{i}H))=H^{j}(Y,\mathcal{O}_{Y}(p^{i}H-L))=H^{j}(Y,\mathcal{O}_{Y}(p^{i}H-2L))=0 for every j{1,2}j\in\{1,2\}.

    3. (0c)

      H3(Y,𝒪Y(piH2L))=H3(Y,𝒪Y(piH3L))=0H^{3}(Y,\mathcal{O}_{Y}(p^{i}H-2L))=H^{3}(Y,\mathcal{O}_{Y}(p^{i}H-3L))=0 for every i>0i>0.

  2. (1)
    1. (1a)

      H1(Y,ΩY1(HnL))=0H^{1}(Y,\Omega_{Y}^{1}(-H-nL))=0 for every n0n\geq 0.

    2. (1b)

      H2(Y,ΩY1(H2L))=H2(Y,ΩY1(H3L))=0H^{2}(Y,\Omega_{Y}^{1}(-H-2L))=H^{2}(Y,\Omega_{Y}^{1}(-H-3L))=0.

    3. (1c)

      Hj(Y,ΩY1(piH))=Hj(Y,ΩY1(piHL))=0H^{j}(Y,\Omega^{1}_{Y}(p^{i}H))=H^{j}(Y,\Omega^{1}_{Y}(p^{i}H-L))=0 for every j{2,3}j\in\{2,3\} and every i>0i>0.

  3. (2)

    H0(X,ΩX2(piKX))=0H^{0}(X,\Omega_{X}^{2}(p^{i}K_{X}))=0 for every i>0i>0.

Then the following hold.

  1. (I)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega^{1}_{X}(K_{X}))=0.

  2. (II)

    H2(X,ΩX1(piKX))=0H^{2}(X,\Omega^{1}_{X}(-p^{i}K_{X}))=0 for every i>0i>0.

  3. (III)

    XX is quasi-FF-split.

Proof.

By (2) and Proposition 2.20, (I) and (II) imply (III). In what follows, we shall prove (I) and (II). Note that we can write KXfH-K_{X}\sim f^{*}H for H:=KYLH:=-K_{Y}-L, and hence KX-K_{X} is ample.

Step 1: Proof of (I).   By the conormal exact sequence, we have an exact sequence

0𝒪X(KXX)ΩP1|X(KX)ΩX1(KX)0,0\to\mathcal{O}_{X}(K_{X}-X)\to\Omega^{1}_{P}|_{X}(K_{X})\to\Omega^{1}_{X}(K_{X})\to 0,

where ΩP1|X(KX):=(ΩP1|X)𝒪X(KX)\Omega^{1}_{P}|_{X}(K_{X}):=(\Omega^{1}_{P}|_{X})\otimes\mathcal{O}_{X}(K_{X}). It follows from KX=gH|X=fHK_{X}=-g^{*}H|_{X}=-f^{*}H, X|X=(X2S)|X=2gL|X=2fLX|_{X}=(X-2S)|_{X}=2g^{*}L|_{X}=2f^{*}L, and f𝒪X=𝒪Y𝒪Y(L)f_{*}\mathcal{O}_{X}=\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L) that

H2(X,𝒪X(KXX))\displaystyle H^{2}(X,\mathcal{O}_{X}(K_{X}-X)) =H2(X,f𝒪Y((H+2L)))\displaystyle=H^{2}(X,f^{*}\mathcal{O}_{Y}(-(H+2L)))
=H2(Y,𝒪Y(H2L))H2(Y,𝒪Y(H3L))\displaystyle=H^{2}(Y,\mathcal{O}_{Y}(-H-2L))\oplus H^{2}(Y,\mathcal{O}_{Y}(-H-3L))
=(0a)0.\displaystyle\overset{{\rm(0a)}}{=}0.

Thus it suffices to show H1(X,ΩP1|X(KX))=0H^{1}(X,\Omega^{1}_{P}|_{X}(K_{X}))=0.

By KX=gH|XK_{X}=-g^{*}H|_{X}, S|X=0S|_{X}=0, and X=2S2gL-X=-2S-2g^{*}L, we have the following exact sequence:

0ΩP1(gH2gL)ΩP1(gH+2S)ΩP1|X(KX)0.0\to\Omega^{1}_{P}(-g^{*}H-2g^{*}L)\to\Omega^{1}_{P}(-g^{*}H+2S)\to\Omega^{1}_{P}|_{X}(K_{X})\to 0.

Thus, in order to prove (I), it is enough to show

  1. (i)

    H1(P,ΩP1(gH+2S))=0H^{1}(P,\Omega^{1}_{P}(-g^{*}H+2S))=0 and

  2. (ii)

    H2(P,ΩP1(gH2gL))=0H^{2}(P,\Omega^{1}_{P}(-g^{*}H-2g^{*}L))=0.

Step 1-1: Proof of (i).   We have the following exact sequence:

0g(ΩY1(H))(2S)ΩP1(gH+2S)ΩP/Y1(gH+2S)0.0\to g^{*}(\Omega^{1}_{Y}(-H))(2S)\to\Omega^{1}_{P}(-g^{*}H+2S)\to\Omega^{1}_{P/Y}(-g^{*}H+2S)\to 0.

It holds that

H1(P,ΩP/Y1(gH+2S))H1(P,𝒪P(gHgL))H1(Y,𝒪Y(HL))=(0a)0.H^{1}(P,\Omega^{1}_{P/Y}(-g^{*}H+2S))\simeq H^{1}(P,\mathcal{O}_{P}(-g^{*}H-g^{*}L))\simeq H^{1}(Y,\mathcal{O}_{Y}(-H-L))\overset{{\rm(0a)}}{=}0.

Then it suffices to show H1(P,g(ΩY1(H))(2S))=0H^{1}(P,g^{*}(\Omega^{1}_{Y}(-H))(2S))=0. Using an exact sequence 0𝒪P(S)𝒪P𝒪S00\to\mathcal{O}_{P}(-S)\to\mathcal{O}_{P}\to\mathcal{O}_{S}\to 0 twice, the vanishing H1(P,g(ΩY(H))(2S))=0H^{1}(P,g^{*}(\Omega_{Y}(-H))(2S))=0 can be reduced, by S|S=gL|SS|_{S}=-g^{*}L|_{S}, to those of

  • H1(S,g(ΩY1(H))(nS))g|S:isomH1(Y,ΩY1(HnL))H^{1}(S,g^{*}(\Omega^{1}_{Y}(-H))(nS))\overset{g|_{S}:\text{isom}}{\simeq}H^{1}(Y,\Omega^{1}_{Y}(-H-nL)) for n{1,2}n\in\{1,2\} and

  • H1(P,gΩY1(H))H1(Y,ΩY1(H))H^{1}(P,g^{*}\Omega_{Y}^{1}(-H))\simeq H^{1}(Y,\Omega_{Y}^{1}(-H)).

Both of them follow from (1a). Thus (i) holds.

Step 1-2: Proof of (ii).   In order to show (ii), it is enough to verify the assumptions of Lemma 6.5 for the case when q=2q=2 and D=H2LD=-H-2L. The conditions Lemma 6.5(1) and Lemma 6.5(2) hold by (0a) and (1b), respectively. This completes the proofs of (ii) and (I).

Step 2: Proof of (II).   Fix i>0i\in\mathbb{Z}_{>0}. By the conormal exact sequence, we have the following exact sequence:

0𝒪X(piKXX)ΩP1|X(piKX)ΩX1(piKX)0.0\to\mathcal{O}_{X}(-p^{i}K_{X}-X)\to\Omega^{1}_{P}|_{X}(-p^{i}K_{X})\to\Omega^{1}_{X}(-p^{i}K_{X})\to 0.

It follows from KX=fHK_{X}=-f^{*}H, X|X=2fLX|_{X}=2f^{*}L, and f𝒪X=𝒪Y𝒪Y(L)f_{*}\mathcal{O}_{X}=\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L) that

H3(X,piKXX)\displaystyle H^{3}(X,-p^{i}K_{X}-X) =H3(X,pifH2fL)\displaystyle=H^{3}(X,p^{i}f^{*}H-2f^{*}L)
=H3(Y,𝒪Y(piH2L))H3(Y,𝒪Y(piH3L))\displaystyle=H^{3}(Y,\mathcal{O}_{Y}(p^{i}H-2L))\oplus H^{3}(Y,\mathcal{O}_{Y}(p^{i}H-3L))
=(0c)0.\displaystyle\overset{{\rm(0c)}}{=}0.

Thus it suffices to show H2(X,ΩP1|X(piKX))=0H^{2}(X,\Omega^{1}_{P}|_{X}(-p^{i}K_{X}))=0.

We get

Hq(P,ΩP1(pigH))=0foreveryq{2,3},H^{q}(P,\Omega^{1}_{P}(p^{i}g^{*}H))=0\qquad\text{for}\qquad\text{every}\qquad q\in\{2,3\},

because Lemma 6.5, for the case when q{2,3}q\in\{2,3\} and D=piHD=p^{i}H, is applicable by (0b) and (1c). Since KX=gH|XK_{X}=-g^{*}H|_{X} and X=2T-X=-2T, we have the following exact sequence:

0ΩP1(pigH2T)ΩP1(pigH)ΩP1|X(piKX)0.0\to\Omega^{1}_{P}(p^{i}g^{*}H-2T)\to\Omega^{1}_{P}(p^{i}g^{*}H)\to\Omega^{1}_{P}|_{X}(-p^{i}K_{X})\to 0.

Thus it suffices to show that

H3(P,ΩP1(pigH2T))=0.H^{3}(P,\Omega^{1}_{P}(p^{i}g^{*}H-2T))=0.

By T|T=gL|TT|_{T}=g^{*}L|_{T} and an exact sequence 0𝒪P(T)𝒪P𝒪T00\to\mathcal{O}_{P}(-T)\to\mathcal{O}_{P}\to\mathcal{O}_{T}\to 0, the problem is reduced to

  • H2(T,ΩP1|T(pigHnT))=0H^{2}(T,\Omega^{1}_{P}|_{T}(p^{i}g^{*}H-nT))=0 for n{0,1}n\in\{0,1\} and

  • H3(P,ΩP1(pigH))=0H^{3}(P,\Omega^{1}_{P}(p^{i}g^{*}H))=0.

The second vanishing has been settled already. Thus it suffices to show the first one. Fix n{0,1}n\in\{0,1\}. By the conormal exact sequence, we have the following exact sequence

0𝒪T(pigH(n+1)T)ΩP1|T(pigHnT)ΩT1(pigHnT)0.0\to\mathcal{O}_{T}(p^{i}g^{*}H-(n+1)T)\to\Omega^{1}_{P}|_{T}(p^{i}g^{*}H-nT)\to\Omega^{1}_{T}(p^{i}g^{*}H-nT)\to 0.

It follows from 𝒪T(pigH(n+1)T)TY𝒪Y(piH(n+1)L)\mathcal{O}_{T}(p^{i}g^{*}H-(n+1)T)\overset{T\simeq Y}{\simeq}\mathcal{O}_{Y}(p^{i}H-(n+1)L) that

H2(T,𝒪T(pigH(n+1)T)H2(Y,𝒪Y(piH(n+1)L))=(0b)0.\displaystyle H^{2}(T,\mathcal{O}_{T}(p^{i}g^{*}H-(n+1)T)\simeq H^{2}(Y,\mathcal{O}_{Y}(p^{i}H-(n+1)L))\overset{{\rm(0b)}}{=}0.

Then we are done by H2(T,ΩT1(pigHnT)H2(Y,ΩY1(piHnL))=(1c)0H^{2}(T,\Omega^{1}_{T}(p^{i}g^{*}H-nT)\simeq H^{2}(Y,\Omega^{1}_{Y}(p^{i}H-nL))\overset{{\rm(1c)}}{=}0. ∎

6.1.2. 2-2

Lemma 6.7.

A smooth Fano threefold XX of No. 2-2 satisfies the following properties:

  1. (1)

    There is a split double cover f:XY:=1×2f\colon X\to Y:=\mathbb{P}^{1}\times\mathbb{P}^{2}.

  2. (2)

    f𝒪X𝒪Y𝒪Y(L)f_{*}\mathcal{O}_{X}\simeq\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L) for a Cartier divisor LL satisfying 𝒪Y(L)𝒪Y(1,2)\mathcal{O}_{Y}(L)\simeq\mathcal{O}_{Y}(1,2).

  3. (3)

    XX is (isomorphic to) a divisor on PY(𝒪Y𝒪Y(L))P\coloneqq\mathbb{P}_{Y}(\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L)).

  4. (4)

    KP+X=gHK_{P}+X=-g^{*}H and KX=g(KY+L)=gH|XK_{X}=g^{*}(K_{Y}+L)=-g^{*}H|_{X}, where H𝒪Y(1,1)H\coloneqq\mathcal{O}_{Y}(1,1) and g:P=Y(𝒪Y𝒪Y(L))Yg:P=\mathbb{P}_{Y}(\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L))\to Y denotes the projection.

  5. (5)

    There exists a section SS of gg such that SX=S\cap X=\emptyset, S|SgL|SS|_{S}\sim-g^{*}L|_{S}, X2S2gLX-2S\sim 2g^{*}L, 𝒪P(1)𝒪P(S)\mathcal{O}_{P}(1)\simeq\mathcal{O}_{P}(S), and ΩP/Y1𝒪P(gL2S)\Omega_{P/Y}^{1}\simeq\mathcal{O}_{P}(-g^{*}L-2S).

  6. (6)

    There exists a section TT of gg such that ST=S\cap T=\emptyset, T|TgL|TT|_{T}\sim g^{*}L|_{T}, X2TX\sim 2T, TSgLT-S\sim g^{*}L, and 𝒪P(1)𝒪P(TgL)\mathcal{O}_{P}(1)\simeq\mathcal{O}_{P}(T-g^{*}L).

Proof.

The assertions (1) and (2) follow from [FanoIII, Subsection 9.2]. Then the remaining ones hold by Lemma 6.4. ∎

Lemma 6.8.

Let XX be a smooth Fano threefold of No. 2-2. Then the following hold.

  1. (1)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega^{1}_{X}(K_{X}))=0.

  2. (2)

    H2(X,ΩX1(piKX))=0H^{2}(X,\Omega^{1}_{X}(-p^{i}K_{X}))=0 for every i>0i>0.

  3. (3)

    XX is quasi-FF-split.

Proof.

We use the same notation as Lemma 6.7. It is enough to verify the conditions of Proposition 6.6 for Y=1×2,L=𝒪Y(1,1)Y=\mathbb{P}^{1}\times\mathbb{P}^{2},L=\mathcal{O}_{Y}(1,1), and H=𝒪Y(1,2)H=\mathcal{O}_{Y}(1,2). Note that mHLmH-L is ample when m2m\geq 2. By Lemma 6.2, Proposition 6.6(2) holds. Since YY satisfies Kodaira vanishing, it is easy to see that Proposition 6.6(0) holds. As YY satisfies Bott vanishing, it is obvious that Proposition 6.6(1) holds. ∎

6.1.3. 2-6-a

Definition 6.9.

Let XX be a smooth Fano threefold of No. 2-6. By [FanoIV, Section 7.2], one of the following holds up to isomorphisms.

  1. (2-6-a)

    XX is a hypersurface of P:=2×2P:=\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (2,2)(2,2). In this case, 𝒪X(KX)𝒪X(1,1)\mathcal{O}_{X}(-K_{X})\simeq\mathcal{O}_{X}(1,1), where 𝒪X(1,1):=𝒪P(1,1)\mathcal{O}_{X}(1,1):=\mathcal{O}_{P}(1,1). In this case, we say that XX is (a Fano threefold) of No. 2-6-a.

  2. (2-6-b)

    There is a split double cover f:XWf:X\to W satisfying f𝒪X𝒪W𝒪W(L)f_{*}\mathcal{O}_{X}\simeq\mathcal{O}_{W}\oplus\mathcal{O}_{W}(-L), where LL is a Cartier divisor on WW with 𝒪W(2L)ωW1\mathcal{O}_{W}(2L)\simeq\omega_{W}^{-1}. In this case, we say that XX is (a Fano threefold) of No. 2-6-b.

Lemma 6.10.

Let XX be a smooth Fano threefold of No. 2-6-a. Then the following hold.

  1. (1)

    Hi(X,𝒪X(n,n))=0H^{i}(X,\mathcal{O}_{X}(n,n))=0 for i{1,2}i\in\{1,2\} and nn\in\mathbb{Z}.

  2. (2)

    H3(X,𝒪X(n,n))=0H^{3}(X,\mathcal{O}_{X}(n,n))=0 for n0n\in\mathbb{Z}_{\geq 0}.

Proof.

Use an exact sequence 0𝒪2×2(2,2)𝒪2×2𝒪X00\to\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(-2,-2)\to\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}\to\mathcal{O}_{X}\to 0. ∎

Lemma 6.11.

Let XX be a smooth Fano threefold of No. 2-6-a. Then the following hold.

  1. (1)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega^{1}_{X}(K_{X}))=0.

  2. (2)

    H2(X,ΩX1(piKX))=0H^{2}(X,\Omega^{1}_{X}(-p^{i}K_{X}))=0 for every i>0i>0.

  3. (3)

    XX is quasi-FF-split.

Proof.

We use the notation of Definition 6.9. By Proposition 2.20 and Lemma 6.2, it suffices to show (1) and (2).

Step 1: Proof of (1).   By the conormal exact sequence, we have the following exact sequence:

0𝒪X(KXX)ΩP1|X(KX)ΩX1(KX)0.0\to\mathcal{O}_{X}(K_{X}-X)\to\Omega^{1}_{P}|_{X}(K_{X})\to\Omega^{1}_{X}(K_{X})\to 0.

Since we have

H2(X,KXX)=H2(X,𝒪X(3,3))=0H^{2}(X,K_{X}-X)=H^{2}(X,\mathcal{O}_{X}(-3,-3))=0

by Lemma 6.10, it suffices to show H1(X,ΩP1|X(KX))=0H^{1}(X,\Omega^{1}_{P}|_{X}(K_{X}))=0.

We have the following exact sequence:

0ΩP1(KP)ΩP1(KP+X)ΩP1|X(KX)00\to\Omega^{1}_{P}(K_{P})\to\Omega^{1}_{P}(K_{P}+X)\to\Omega^{1}_{P}|_{X}(K_{X})\to 0

By Bott vanishing, we have

  • H1(P,ΩP1(KP+X))=H1(P,ΩP1(1,1))=0H^{1}(P,\Omega^{1}_{P}(K_{P}+X))=H^{1}(P,\Omega^{1}_{P}(-1,-1))=0 and

  • H2(P,ΩP1(KP))=H2(P,ΩP1(3,3))=0H^{2}(P,\Omega^{1}_{P}(K_{P}))=H^{2}(P,\Omega^{1}_{P}(-3,-3))=0.

Therefore, (1) holds.

Step 2: Proof of (2).   Fix i>0i\in\mathbb{Z}_{>0}. By the conormal exact sequence, we have the following exact sequence

0𝒪X(piKXX)ΩP1|X(piKX)ΩX1(piKX)0.0\to\mathcal{O}_{X}(-p^{i}K_{X}-X)\to\Omega^{1}_{P}|_{X}(-p^{i}K_{X})\to\Omega^{1}_{X}(-p^{i}K_{X})\to 0.

We have

H3(X,𝒪X(piKXX))=H3(X,𝒪X(pi2,pi2))=0H^{3}(X,\mathcal{O}_{X}(-p^{i}K_{X}-X))=H^{3}(X,\mathcal{O}_{X}(p^{i}-2,p^{i}-2))=0

by Lemma 6.10. Thus, it suffices to show H2(X,ΩP1|X(piKX))=0H^{2}(X,\Omega^{1}_{P}|_{X}(-p^{i}K_{X}))=0.

We have the following exact sequence:

0ΩP1(pi(KP+X)X)ΩP1(pi(KP+X))ΩP1|X(piKX)0.0\to\Omega^{1}_{P}(-p^{i}(K_{P}+X)-X)\to\Omega^{1}_{P}(-p^{i}(K_{P}+X))\to\Omega^{1}_{P}|_{X}(-p^{i}K_{X})\to 0.

Then

  • H2(P,ΩP1(pi(KP+X)))=H2(P,ΩP1(pi,pi))=0H^{2}(P,\Omega^{1}_{P}(-p^{i}(K_{P}+X)))=H^{2}(P,\Omega^{1}_{P}(p^{i},p^{i}))=0 by Bott vanishing and

  • H3(P,ΩP1(pi(KP+X)X))=H3(P,ΩP1(pi2,pi2))=0H^{3}(P,\Omega^{1}_{P}(-p^{i}(K_{P}+X)-X))=H^{3}(P,\Omega^{1}_{P}(p^{i}-2,p^{i}-2))=0 by [Totaro(Fano), Proposition 1.3].

Therefore, the assertion holds. ∎

6.1.4. 2-6-b

Lemma 6.12.

A smooth Fano threefold XX of No. 2-6-b satisfies the following properties:

  1. (1)

    There is a split double cover f:XWf\colon X\to W, where WW is a smooth hypersurface of 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,1)(1,1).

  2. (2)

    f𝒪X𝒪W𝒪W(L)f_{*}\mathcal{O}_{X}\simeq\mathcal{O}_{W}\oplus\mathcal{O}_{W}(-L) for a Cartier divisor L=𝒪W(1,1)L=\mathcal{O}_{W}(1,1).

  3. (3)

    XX is a divisor of PW(𝒪W𝒪W(L))P\coloneqq\mathbb{P}_{W}(\mathcal{O}_{W}\oplus\mathcal{O}_{W}(-L)).

  4. (4)

    KP+X=gLK_{P}+X=-g^{*}L, KX=g(KW+L)=gL|XK_{X}=g^{*}(K_{W}+L)=-g^{*}L|_{X}.

  5. (5)

    There exists a section SS of gg such that SX=S\cap X=\emptyset, S|SgL|SS|_{S}\sim-g^{*}L|_{S}, X2S2gLX-2S\sim 2g^{*}L, 𝒪P(1)𝒪P(S)\mathcal{O}_{P}(1)\simeq\mathcal{O}_{P}(S), and ΩP/W1𝒪P(gL2S)\Omega_{P/W}^{1}\simeq\mathcal{O}_{P}(-g^{*}L-2S).

Proof.

The assertions (1) and (2) follow from [FanoIII, Subsection 9.2]. Then the remaining ones hold by Lemma 6.4. ∎

Lemma 6.13.

Let Y=3Y=\mathbb{P}^{3} (resp. Y=QY=Q, resp. Y=WY=W), where QQ is a smooth quadric hypersurface of 4\mathbb{P}^{4} and WW is a smooth hypersurface of 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,1)(1,1). Let L=𝒪3(3)L=\mathcal{O}_{\mathbb{P}^{3}}(3) (resp. 𝒪Q(2)\mathcal{O}_{Q}(2), resp. 𝒪W(1,1)\mathcal{O}_{W}(1,1)). Then the following hold.

  1. (1)

    H1(Y,𝒪Y(nL))=0H^{1}(Y,\mathcal{O}_{Y}(nL))=0 for nn\in\mathbb{Z}.

  2. (2)

    H2(Y,𝒪Y(nL))=0H^{2}(Y,\mathcal{O}_{Y}(nL))=0 for nn\in\mathbb{Z}.

  3. (3)

    H3(Y,𝒪Y(nL))=0H^{3}(Y,\mathcal{O}_{Y}(nL))=0 for n1n\in\mathbb{Z}_{\geq-1}.

  4. (4)

    H0(Y,ΩY1(nL))=0H^{0}(Y,\Omega^{1}_{Y}(nL))=0 for n0n\in\mathbb{Z}_{\leq 0}.

  5. (5)

    H1(Y,ΩY1(nL))=0H^{1}(Y,\Omega^{1}_{Y}(nL))=0 for n{0}n\in\mathbb{Z}\setminus\{0\}.

  6. (6)

    H2(Y,ΩY1(nL))=0H^{2}(Y,\Omega^{1}_{Y}(nL))=0 for n{1}n\in\mathbb{Z}\setminus\{-1\}.

  7. (7)

    H3(Y,ΩY1(nL))=0H^{3}(Y,\Omega^{1}_{Y}(nL))=0 for n0n\in\mathbb{Z}_{\geq 0}.

Proof.

Since YY is FF-split (Lemma 5.7), (1)-(3) hold. The assertions (4) and (7) follow from the fact that XX is SRC (Lemma 6.2). Let us prove (5) and (6). If Y=3Y=\mathbb{P}^{3}, then these follow from the Bott vanishing theorem and [Totaro(Fano), Proposition 1.3]. In what follows, we assume Y{Q,W}Y\in\{Q,W\}. If Y=QY=Q (resp. Y=WY=W), then

  • we have an embedding YPY\subset P for P:=4P:=\mathbb{P}^{4} (resp. P=2×2P=\mathbb{P}^{2}\times\mathbb{P}^{2}),

  • we set H:=𝒪4(1)H:=\mathcal{O}_{\mathbb{P}^{4}}(1) (resp. H:=𝒪2×2(1,1)H:=\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1)), and

  • we get LsH|YL\sim sH|_{Y} for s:=2s:=2 (resp. s:=1s:=1). It holds that YsHY\sim sH.

We have the following exact sequence:

0ΩP1((ns)H)ΩP1(nH)ΩP1(nH)|Y0.0\to\Omega_{P}^{1}((n-s)H)\to\Omega_{P}^{1}(nH)\to\Omega^{1}_{P}(nH)|_{Y}\to 0.

By Bott vanishing and [Totaro(Fano), Proposition 1.3], we have

  • H1(P,ΩP1(nH))=0H^{1}(P,\Omega^{1}_{P}(nH))=0 for n{0}n\in\mathbb{Z}\setminus\{0\},

  • H2(P,ΩP1(nH))=0H^{2}(P,\Omega^{1}_{P}(nH))=0 for nn\in\mathbb{Z}, and

  • H3(P,ΩP1(nH))=0H^{3}(P,\Omega^{1}_{P}(nH))=0 for nn\in\mathbb{Z}.

We then get

  • H1(Y,ΩP1(nH)|Y)=0H^{1}(Y,\Omega^{1}_{P}(nH)|_{Y})=0 for n{0}n\in\mathbb{Z}\setminus\{0\} and

  • H2(Y,ΩP1(nH)|Y)=0H^{2}(Y,\Omega^{1}_{P}(nH)|_{Y})=0 for nn\in\mathbb{Z}.

By YsHY\sim sH and the conormal exact sequence, we have the following exact sequence:

0𝒪Y(nHsH)ΩP1(nH)|YΩY1(nH)0.0\to\mathcal{O}_{Y}(nH-sH)\to\Omega_{P}^{1}(nH)|_{Y}\to\Omega^{1}_{Y}(nH)\to 0.

Recall that

  • H2(Y,𝒪Y(nH))=0H^{2}(Y,\mathcal{O}_{Y}(nH))=0 for nn\in\mathbb{Z} and

  • H3(Y,𝒪Y(nH))=0H^{3}(Y,\mathcal{O}_{Y}(nH))=0 for nsn\geq-s.

Hence we get

  1. (5)’

    H1(Y,ΩY1(nH))=0H^{1}(Y,\Omega^{1}_{Y}(nH))=0 for n0n\neq 0 and

  2. (6-a)

    H2(Y,ΩY1(nH))=0H^{2}(Y,\Omega^{1}_{Y}(nH))=0 for n0n\geq 0.

In particular, (5) holds. Comparing (6) with (6-a), it suffices to show (6-b) below by Serre duality.

  1. (6-b)

    H1(Y,ΩY2(nH))=0H^{1}(Y,\Omega^{2}_{Y}(nH))=0 for n>sn>s.

Taking the wedge product 2\bigwedge^{2} to the conormal exact sequence, we get the following exact sequence:

0ΩY1(nHsH)ΩP2(nH)|YΩY2(nH)0.0\to\Omega^{1}_{Y}(nH-sH)\to\Omega^{2}_{P}(nH)|_{Y}\to\Omega^{2}_{Y}(nH)\to 0.

Since we have H2(Y,ΩY1(nHsH))=0H^{2}(Y,\Omega_{Y}^{1}(nH-sH))=0 for n>sn>s (6-a), it is enough to prove H1(Y,ΩP2(nH)|Y)=0H^{1}(Y,\Omega^{2}_{P}(nH)|_{Y})=0 for n>sn>s. This holds by an exact sequence

0ΩP2(nHsH)ΩP2(nH)ΩP2(nH)|Y0,0\to\Omega^{2}_{P}(nH-sH)\to\Omega^{2}_{P}(nH)\to\Omega^{2}_{P}(nH)|_{Y}\to 0,

because Bott vanishing implies Hi(P,ΩP2(mH))=0H^{i}(P,\Omega_{P}^{2}(mH))=0 for i>0i>0 and m>0m>0. ∎

Lemma 6.14.

Let XX be a smooth Fano threefold of No. 2-6-b. Then the following hold.

  1. (1)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega^{1}_{X}(K_{X}))=0.

  2. (2)

    H2(X,ΩX1(piKX))=0H^{2}(X,\Omega^{1}_{X}(-p^{i}K_{X}))=0 for every i>0i>0.

  3. (3)

    XX is quasi-FF-split.

Proof.

It is enough to verify the conditions in Proposition 6.6. Note that we have H=KYL=LH=-K_{Y}-L=L. Then the conditions in Proposition 6.6 follow from Lemma 6.2 and Lemma 6.13. ∎

The above argument can be applied for some hyperelliptic Fano threefolds. Let us start by recalling the definition.

Definition 6.15.

We say that a smooth Fano threefold XX is hyperelliptic if XX is of index one, |KX||-K_{X}| is base point free, and the induced morphism f:XYφ|KX|(X)f\colon X\to Y\coloneqq\varphi_{|-K_{X}|}(X) is a double cover.

It is known that if ρ(X)=1\rho(X)=1, then YY is isomorphic to 3\mathbb{P}^{3} or QQ in the above notation [FanoI, Theorem 6.5]. The assumption p>5p>5 in Proposition 6.16(i) is sharp as we shall see later (Example 8.7).

Proposition 6.16.

Let XX be a hyperelliptic smooth Fano threefold such that ρ(X)=1\rho(X)=1. Let f:XYφ|KX|(X)f\colon X\to Y\coloneqq\varphi_{|-K_{X}|}(X) be the double cover induced by φ|KX|\varphi_{|-K_{X}|}. Assume the following.

  1. (i)

    If Y3Y\simeq\mathbb{P}^{3}, then p>5p>5.

  2. (ii)

    If YQY\simeq Q, then p>3p>3.

Then the following hold.

  1. (1)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega^{1}_{X}(K_{X}))=0.

  2. (2)

    H2(X,ΩX1(piKX))=0H^{2}(X,\Omega^{1}_{X}(-p^{i}K_{X}))=0 for every i>0i>0.

  3. (3)

    XX is quasi-FF-split.

Proof.

We have f𝒪X𝒪Y𝒪Y(L)f_{*}\mathcal{O}_{X}\simeq\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L) for a Cartier divisor L=𝒪Y(r1)L=\mathcal{O}_{Y}(r-1), where rr denotes the index of YY (i.e., if Y=3Y=\mathbb{P}^{3} (resp. Y=QY=Q), then r=4r=4 (resp. r=3r=3)). It is enough to verify the conditions in Proposition 6.6. Note that we have L=(r1)HL=(r-1)H and 𝒪Y(1)=𝒪Y(H)\mathcal{O}_{Y}(1)=\mathcal{O}_{Y}(H). Then Lemma 6.2 and Lemma 6.13 imply all the conditions in Proposition 6.6 except for Proposition 6.6(0c). By our assumptions (i) and (ii), Proposition 6.6(0c) directly follows from Serre duality, e.g.,

h3(Y,𝒪Y(piH3L))=h0(Y,𝒪Y(KY+3LpiH))=h0(Y,𝒪Y((2r3pi)H))=0h^{3}(Y,\mathcal{O}_{Y}(p^{i}H-3L))=h^{0}(Y,\mathcal{O}_{Y}(K_{Y}+3L-p^{i}H))=h^{0}(Y,\mathcal{O}_{Y}((2r-3-p^{i})H))=0

for every i>0i>0. ∎

6.1.5. 2-8

Lemma 6.17.

A smooth Fano threefold XX of No. 2-8 satisfies the following properties:

  1. (1)

    XX is (isomorphic to) a divisor on P2(𝒪2𝒪2(1)𝒪2(2))P\coloneqq\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)).

  2. (2)

    𝒪P(KP)=𝒪P(3)\mathcal{O}_{P}(K_{P})=\mathcal{O}_{P}(-3), 𝒪P(X)=𝒪P(2)\mathcal{O}_{P}(X)=\mathcal{O}_{P}(2), KX=𝒪X(1),K_{X}=\mathcal{O}_{X}(-1), and 𝒪X(X)=𝒪X(2)=𝒪X(2KX)\mathcal{O}_{X}(X)=\mathcal{O}_{X}(2)=\mathcal{O}_{X}(-2K_{X}).

  3. (3)

    |𝒪P(1)||\mathcal{O}_{P}(1)| is base point free and 𝒪P(1)\mathcal{O}_{P}(1) is big.

  4. (4)

    Let φ:PP\varphi:P\to P^{\prime} be the birational morphism to a normal projective variety PP^{\prime} such that φ𝒪P=𝒪P\varphi_{*}\mathcal{O}_{P}=\mathcal{O}_{P^{\prime}} and 𝒪P(1)φ𝒪P(1)\mathcal{O}_{P}(1)\simeq\varphi^{*}\mathcal{O}_{P^{\prime}}(1) for some ample invertible sheaf 𝒪P(1)\mathcal{O}_{P^{\prime}}(1) on PP^{\prime}. Then φ:PP\varphi:P\to P^{\prime} is a small birational morphism which is an isomorphism around XX.

We say that a birational morphism φ:PP\varphi:P\to P^{\prime} is small if dimEx(φ)dimP2\dim\operatorname{Ex}(\varphi)\leq\dim P-2.

Proof.

Note that a Fano threefold of No. 2-8 is characterised by the following properties (i) and (ii) [FanoIII, Theorem 5.34]:

  1. (i)

    XX is a Fano threefold with ρ(X)=2\rho(X)=2.

  2. (ii)

    One of the extremal rays is of type C1C_{1}, and the the other extremal ray is of type E3E_{3} or E4E_{4}.

Then we may apply [FanoIII, Proposition 5.29 and Lemma 5.30]. By [FanoIII, Proposition 5.29(3), Lemma 5.30], XX is a divisor on P2(𝒪2𝒪2(1)𝒪2(2))P\coloneqq\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)) satisfying X𝒪P(2)X\sim\mathcal{O}_{P}(2). Thus (1) holds. Moreover, [FanoIII, the proof of Lemma 5.30] implies that KP𝒪P(3)-K_{P}\sim\mathcal{O}_{P}(3). Then the adunction formula implies KX(KP+X)|X𝒪P(3+2)|X=𝒪X(1)K_{X}\sim(K_{P}+X)|_{X}\sim\mathcal{O}_{P}(-3+2)|_{X}=\mathcal{O}_{X}(-1). Thus (2) holds.

Let us show (3) and (4). We have three sections Γ0,Γ1,Γ2\Gamma_{0},\Gamma_{1},\Gamma_{2} of the induced 2\mathbb{P}^{2}-bundle g:P=2(E)2g:P=\mathbb{P}_{\mathbb{P}^{2}}(E)\to\mathbb{P}^{2}, where E:=𝒪2𝒪2(1)𝒪2(2)E:=\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2). corresponding to the projections of EE to the factors 𝒪2,𝒪2(1),𝒪2(2)\mathcal{O}_{\mathbb{P}^{2}},\mathcal{O}_{\mathbb{P}^{2}}(1),\mathcal{O}_{\mathbb{P}^{2}}(2), respectively. Similarly, we have the following three prime divisors which are 1\mathbb{P}^{1}-bundles over 2\mathbb{P}^{2}:

  • D0:=2(𝒪2(1)𝒪2(2))D_{0}:=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)), corresponding to E𝒪2(1)𝒪2(2)E\to\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2).

  • D1:=2(𝒪2𝒪2(2))D_{1}:=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)), corresponding to E𝒪2𝒪2(2)E\to\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2).

  • D2:=2(𝒪2𝒪2(1))D_{2}:=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)), corresponding to E𝒪2𝒪2(1)E\to\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1).

By construction, we have DiΓi=D_{i}\cap\Gamma_{i}=\emptyset for every i{0,1,2}i\in\{0,1,2\}. Fix a line LL on 2\mathbb{P}^{2} and set F:=gLF:=g^{*}L, which is a prime divisor on PP. For each i{0,1,2}i\in\{0,1,2\}, we can write 𝒪P(1)Di+bF\mathcal{O}_{P}(1)\sim D_{i}+bF for some bb\in\mathbb{Z}. It holds that

𝒪2(i)=𝒪P(1)|Γi=(Di+bF)|Γi=bF|Γi𝒪2(b),\mathcal{O}_{\mathbb{P}^{2}}(i)=\mathcal{O}_{P}(1)|_{\Gamma_{i}}=(D_{i}+bF)|_{\Gamma_{i}}=bF|_{\Gamma_{i}}\simeq\mathcal{O}_{\mathbb{P}^{2}}(b),

i.e., b=ib=i. Then we have that

𝒪P(1)D0D1+FD2+2F.\mathcal{O}_{P}(1)\sim D_{0}\sim D_{1}+F\sim D_{2}+2F.

By D0D1D2=D_{0}\cap D_{1}\cap D_{2}=\emptyset, |𝒪P(1)||\mathcal{O}_{P}(1)| is base point free. Since D1D_{1} is relatively ample over 2\mathbb{P}^{2} and FF is the pullback of an ample divisor, the divisor 𝒪P(1)D1+F\mathcal{O}_{P}(1)\sim D_{1}+F is big. Thus (3) holds. Let φ:PP\varphi:P\to P^{\prime} be as in the statement of (4).

We now show that Ex(φ)=Γ0\operatorname{Ex}(\varphi)=\Gamma_{0}. By 𝒪P(1)|Γ0𝒪Γ0\mathcal{O}_{P}(1)|_{\Gamma_{0}}\simeq\mathcal{O}_{\Gamma_{0}}, φ(Γ0)\varphi(\Gamma_{0}) is a point. In particular, Ex(φ)Γ0\operatorname{Ex}(\varphi)\supset\Gamma_{0}. Pick a curve CC such that φ(C)\varphi(C) is a point. It suffices to show CΓ0C\subset\Gamma_{0}. We have 𝒪P(1)C=0\mathcal{O}_{P}(1)\cdot C=0. Note that g(C)g(C) is not a point, because 𝒪P(1)\mathcal{O}_{P}(1) is gg-ample. In particular, FC>0F\cdot C>0. Therefore,

0=𝒪P(1)C=(D1+F)C=(D2+2F)C.0=\mathcal{O}_{P}(1)\cdot C=(D_{1}+F)\cdot C=(D_{2}+2F)\cdot C.

By FC>0F\cdot C>0, we obtain D1C<0D_{1}\cdot C<0 and D2C<0D_{2}\cdot C<0. Hence CD1D2=Γ0C\subset D_{1}\cap D_{2}=\Gamma_{0}, as required.

It is enough to prove XΓ0=X\cap\Gamma_{0}=\emptyset. Suppose XΓ0X\cap\Gamma_{0}\neq\emptyset. By 𝒪P(X)|Γ0𝒪P(2)|Γ0𝒪Γ0\mathcal{O}_{P}(X)|_{\Gamma_{0}}\simeq\mathcal{O}_{P}(2)|_{\Gamma_{0}}\simeq\mathcal{O}_{\Gamma_{0}}, we obtain 2Γ0X\mathbb{P}^{2}\simeq\Gamma_{0}\subset X. Then the Stein factorisation ψ:XX\psi:X\to X^{\prime} of the composite morphism φ|X:XPP\varphi|_{X}:X\hookrightarrow P\to P^{\prime} is a birational morphism which contracts Γ02\Gamma_{0}\simeq\mathbb{P}^{2} to a point. This is absurd, because XX has no extremal ray of E2E_{2} or E5E_{5}. Thus (4) holds. ∎

Lemma 6.18.

A smooth Fano threefold of No. 2-8 is quasi-F-split.

Proof.

We use the notation of Lemma 6.17. By Proposition 2.20 and Lemma 6.2, it suffices to show that

  1. (1)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega^{1}_{X}(K_{X}))=0.

  2. (2)

    H2(X,ΩX1(piKX))=0H^{2}(X,\Omega^{1}_{X}(-p^{i}K_{X}))=0 for every i>0i>0.

Step 1: Proof of (1).   By the conormal exact sequence, we have the following exact sequence

0𝒪X(KXX)ΩP1|X(KX)ΩX1(KX)0.0\to\mathcal{O}_{X}(K_{X}-X)\to\Omega^{1}_{P}|_{X}(K_{X})\to\Omega^{1}_{X}(K_{X})\to 0.

Since 𝒪X(KX)=𝒪X(1)\mathcal{O}_{X}(K_{X})=\mathcal{O}_{X}(-1) and 𝒪X(X)=𝒪X(2)\mathcal{O}_{X}(X)=\mathcal{O}_{X}(2), we have an exact sequence

H2(P,𝒪P(3))H2(X,𝒪X(3))(=H2(X,KXX))H3(P,𝒪P(5)).H^{2}(P,\mathcal{O}_{P}(-3))\to H^{2}(X,\mathcal{O}_{X}(-3))(=H^{2}(X,K_{X}-X))\to H^{3}(P,\mathcal{O}_{P}(-5)).

Since PP is toric and 𝒪P(1)\mathcal{O}_{P}(1) is nef, it follows from KP𝒪P(3)K_{P}\sim\mathcal{O}_{P}(-3) and [Totaro(Fano), Proposition 1.3] that

  • H2(P,𝒪P(3))H2(P,𝒪P)=0H^{2}(P,\mathcal{O}_{P}(-3))\simeq H^{2}(P,\mathcal{O}_{P})=0 and

  • H3(P,𝒪P(5))H1(P,𝒪P(2))=0H^{3}(P,\mathcal{O}_{P}(-5))\simeq H^{1}(P,\mathcal{O}_{P}(2))=0.

Thus H2(X,KXX)=0H^{2}(X,K_{X}-X)=0, and it suffices to show H1(X,ΩP1|X(KX))=0H^{1}(X,\Omega^{1}_{P}|_{X}(K_{X}))=0.

Since we have a closed embedding XPX\subset P^{\prime} around which PP^{\prime} is smooth (Proposition 6.17(4)), we have the following exact sequence:

0ΩP[1](KP)ΩP[1](KP+X)ΩP[1]|X(KX)ΩP1|X(KX)0,0\to\Omega^{[1]}_{P^{\prime}}(K_{P^{\prime}})\to\Omega^{[1]}_{P^{\prime}}(K_{P^{\prime}}+X)\to\Omega^{[1]}_{P^{\prime}}|_{X}(K_{X})\simeq\Omega^{1}_{P}|_{X}(K_{X})\to 0,

By Bott vanishing [Fuj07, Theorem 1.1 or Corollary 1.3], we have

  • H1(P,ΩP[1](KP+X))=H1(P,ΩP[1](1))=0H^{1}(P^{\prime},\Omega^{[1]}_{P^{\prime}}(K_{P^{\prime}}+X))=H^{1}(P^{\prime},\Omega^{[1]}_{P^{\prime}}(-1))=0 and

  • H2(P,ΩP[1](KP))=H2(P,ΩP[1](3))=0H^{2}(P^{\prime},\Omega^{[1]}_{P^{\prime}}(K_{P^{\prime}}))=H^{2}(P^{\prime},\Omega^{[1]}_{P^{\prime}}(-3))=0.

This completes the proof of (1).

Step 2: Proof of (2).   By the conormal exact sequence, we have the following exact sequence

0𝒪X(piKXX)ΩP1|X(piKX)ΩX1(piKX)0.0\to\mathcal{O}_{X}(-p^{i}K_{X}-X)\to\Omega^{1}_{P}|_{X}(-p^{i}K_{X})\to\Omega^{1}_{X}(-p^{i}K_{X})\to 0.

We have

H3(X,𝒪X(piKXX))=H3(X,𝒪X(pi2))=0.H^{3}(X,\mathcal{O}_{X}(-p^{i}K_{X}-X))=H^{3}(X,\mathcal{O}_{X}(p^{i}-2))=0.

Thus it suffices to show that H2(X,ΩP1|X(piKX))=0H^{2}(X,\Omega^{1}_{P}|_{X}(-p^{i}K_{X}))=0.

We have the following exact sequence:

0ΩP1(pi(KP+X)X)ΩP1(pi(KP+X))ΩP1|X(piKX)0.0\to\Omega^{1}_{P}(-p^{i}(K_{P}+X)-X)\to\Omega^{1}_{P}(-p^{i}(K_{P}+X))\to\Omega^{1}_{P}|_{X}(-p^{i}K_{X})\to 0.

Since 𝒪P(1)\mathcal{O}_{P}(1) is nef and PP is a smooth toric variety, we have

  1. (1)

    H2(P,ΩP1(pi(KP+X)))=H2(P,ΩP1(pi))=0H^{2}(P,\Omega^{1}_{P}(-p^{i}(K_{P}+X)))=H^{2}(P,\Omega^{1}_{P}(p^{i}))=0 and

  2. (2)

    H3(P,ΩP1(pi(KP+X)X))=H3(P,ΩP1(pi2))=0H^{3}(P,\Omega^{1}_{P}(-p^{i}(K_{P}+X)-X))=H^{3}(P,\Omega^{1}_{P}(p^{i}-2))=0.

by [Totaro(Fano), Proposition 1.3]. Thus (2) holds. ∎

6.1.6. 3-10

Lemma 6.19.

Let XX be a smooth Fano threefold XX of No. 3-10 such that there is a wild conic bundle structure f:X1×1f:X\to\mathbb{P}^{1}\times\mathbb{P}^{1}. Then the following properties hold:

  1. (1)

    XX is (isomorphic to) a divisor on P1×1(𝒪𝒪(1,0)𝒪(0,1))P\coloneqq\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}\oplus\mathcal{O}(1,0)\oplus\mathcal{O}(0,1)) satisfying 𝒪P(X)𝒪P(2)\mathcal{O}_{P}(X)\sim\mathcal{O}_{P}(2).

  2. (2)

    Each of XX and 𝒪P(1)\mathcal{O}_{P}(1) is nef and big.

  3. (3)

    KP-K_{P} and (KP+X)-(K_{P}+X) are ample.

  4. (4)

    𝒪X(KX)𝒪X(1)f𝒪1×1(1,1)\mathcal{O}_{X}(K_{X})\simeq\mathcal{O}_{X}(-1)\otimes f^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1), where 𝒪X(1)𝒪P(1)|X\mathcal{O}_{X}(-1)\coloneqq\mathcal{O}_{P}(-1)|_{X}.

  5. (5)

    2i(KP+X)X-2^{i}(K_{P}+X)-X is nef for every i>0i>0.

Proof.

The assertion (1) follows from [MS03, Corollary 8]. Let us show (2). By X𝒪P(2)X\sim\mathcal{O}_{P}(2), it suffices to show that |𝒪P(1)||\mathcal{O}_{P}(1)| is base point free and 𝒪P(1)\mathcal{O}_{P}(1) is big. Set L0:=𝒪1×1,L1:=𝒪1×1(1,0),L2:=𝒪1×1(0,1)L_{0}:=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}},L_{1}:=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0),L_{2}:=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(0,1), and

EL0L1L2=𝒪𝒪(1,0)𝒪(0,1).E\coloneqq L_{0}\oplus L_{1}\oplus L_{2}=\mathcal{O}\oplus\mathcal{O}(1,0)\oplus\mathcal{O}(0,1).

We have three sections of π:P=1×1(𝒪𝒪(1,0)𝒪(0,1))1×1\pi\colon P=\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}\oplus\mathcal{O}(1,0)\oplus\mathcal{O}(0,1))\to\mathbb{P}^{1}\times\mathbb{P}^{1}:

  • Set Γ0:=(𝒪)\Gamma_{0}:=\mathbb{P}(\mathcal{O}), which is corresponding to the projection E=𝒪𝒪(1,0)𝒪(0,1)𝒪E=\mathcal{O}\oplus\mathcal{O}(1,0)\oplus\mathcal{O}(0,1)\to\mathcal{O}. We get 𝒪P(1)|Γ0=𝒪=L0\mathcal{O}_{P}(1)|_{\Gamma_{0}}=\mathcal{O}=L_{0}.

  • Set Γ1:=(𝒪(1,0))\Gamma_{1}:=\mathbb{P}(\mathcal{O}(1,0)), which is corresponding to the projection E=𝒪𝒪(1,0)𝒪(0,1)𝒪(1,0)E=\mathcal{O}\oplus\mathcal{O}(1,0)\oplus\mathcal{O}(0,1)\to\mathcal{O}(1,0). We get 𝒪P(1)|Γ1=𝒪(1,0)=L1\mathcal{O}_{P}(1)|_{\Gamma_{1}}=\mathcal{O}(1,0)=L_{1}.

  • Set Γ2:=(𝒪(0,1))\Gamma_{2}:=\mathbb{P}(\mathcal{O}(0,1)), which is corresponding to the projection E=𝒪𝒪(1,0)𝒪(0,1)𝒪(0,2)E=\mathcal{O}\oplus\mathcal{O}(1,0)\oplus\mathcal{O}(0,1)\to\mathcal{O}(0,2). We get 𝒪P(1)|Γ2=𝒪(0,1)=L2\mathcal{O}_{P}(1)|_{\Gamma_{2}}=\mathcal{O}(0,1)=L_{2}.

Similarly, we have three prime divisors on PP which are 1\mathbb{P}^{1}-bundles over 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}:

  • Set D0:=(L1L2)D_{0}:=\mathbb{P}(L_{1}\oplus L_{2}), which is corresponding to the projection EL1L2E\to L_{1}\oplus L_{2}.

  • Set D1:=(L0L2)D_{1}:=\mathbb{P}(L_{0}\oplus L_{2}), which is corresponding to the projection EL0L2E\to L_{0}\oplus L_{2}.

  • Set D2:=(L0L1)D_{2}:=\mathbb{P}(L_{0}\oplus L_{1}), which is corresponding to the projection EL0L1E\to L_{0}\oplus L_{1}.

By construction, we get ΓiDi=\Gamma_{i}\cap D_{i}=\emptyset for every i{0,1,2}i\in\{0,1,2\}.

We now show that

𝒪P(1)D0D1+π𝒪(1,0)D2+π𝒪(0,1).\mathcal{O}_{P}(1)\sim D_{0}\sim D_{1}+\pi^{*}\mathcal{O}(1,0)\sim D_{2}+\pi^{*}\mathcal{O}(0,1).

Fix i{0,1,2}i\in\{0,1,2\}. Note that we have 𝒪P(1)Di+πMi\mathcal{O}_{P}(1)\sim D_{i}+\pi^{*}M_{i} for some MiM_{i}. By restricting this to Γi\Gamma_{i}, we obtain

Li=𝒪P(1)|Γi=(Di+πMi)|Γi=Mi.L_{i}=\mathcal{O}_{P}(1)|_{\Gamma_{i}}=(D_{i}+\pi^{*}M_{i})|_{\Gamma_{i}}=M_{i}.

Hence we get 𝒪P(1)Di+Li\mathcal{O}_{P}(1)\sim D_{i}+L_{i}, as required.

It follows from D0D1D2=D_{0}\cap D_{1}\cap D_{2}=\emptyset that |𝒪P(1)||\mathcal{O}_{P}(1)| is base point free. We have

𝒪P(2)D1+D2+π𝒪(1,1).\mathcal{O}_{P}(2)\sim D_{1}+D_{2}+\pi^{*}\mathcal{O}(1,1).

Since 𝒪1×1(1,1)\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1) is ample and D1+D2D_{1}+D_{2} is an effective π\pi-ample divisor, 𝒪P(2)\mathcal{O}_{P}(2) is big. Thus (2) holds.

Let us show (3). The following holds (cf. [FanoIII, Proposition 7.1(2)]):

KP𝒪P(3)π(K1×1detE)𝒪P(3)π𝒪(1,1).K_{P}\simeq\mathcal{O}_{P}(-3)\otimes\pi^{*}(K_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\otimes\det E)\simeq\mathcal{O}_{P}(-3)\otimes\pi^{*}\mathcal{O}(-1,-1).

Since 𝒪P(1)\mathcal{O}_{P}(1) is nef and π\pi-ample, KP-K_{P} is ample. Similarly, (KP+X)-(K_{P}+X) is ample by

KP+X𝒪P(1)π𝒪(1,1).K_{P}+X\sim\mathcal{O}_{P}(-1)\otimes\pi^{*}\mathcal{O}(-1,-1).

Thus (3) holds. This linear equivalence implies (4). Finally, (5) follows from

2i(KP+X)X𝒪P(2i2)π𝒪(2i,2i).-2^{i}(K_{P}+X)-X\sim\mathcal{O}_{P}(2^{i}-2)\otimes\pi^{*}\mathcal{O}(2^{i},2^{i}).

Lemma 6.20.

A smooth Fano threefold XX of No. 3-10 is quasi-F-split.

Proof.

By Proposition 4.4, we may assume that p=2p=2 and XX has a wild conic bundle structure. In what follows, we use the notation of Lemma 6.19. By Proposition 2.20 and Lemma 6.2, it suffices to show that

  1. (1)

    H1(X,ΩX1(KX))=0H^{1}(X,\Omega^{1}_{X}(K_{X}))=0.

  2. (2)

    H2(X,ΩX1(2iKX))=0H^{2}(X,\Omega^{1}_{X}(-2^{i}K_{X}))=0 for every i>0i>0.

Step 1: Proof of (1).   By the conormal exact sequence, we have the following exact sequence:

0𝒪X(KXX)ΩP1|X(KX)ΩX1(KX)0.0\to\mathcal{O}_{X}(K_{X}-X)\to\Omega^{1}_{P}|_{X}(K_{X})\to\Omega^{1}_{X}(K_{X})\to 0.

Considering the restriction 𝒪P𝒪X\mathcal{O}_{P}\to\mathcal{O}_{X}, we have an exact sequence

H2(P,KP)H2(X,KXX)H3(P,KPX).H^{2}(P,K_{P})\to H^{2}(X,K_{X}-X)\to H^{3}(P,K_{P}-X).

Since XX is nef, we have H2(P,KP)H2(P,𝒪P)=0H^{2}(P,K_{P})\simeq H^{2}(P,\mathcal{O}_{P})=0 and H3(P,KPX)H1(P,X)=0H^{3}(P,K_{P}-X)\simeq H^{1}(P,X)=0 by [Totaro(Fano), Proposition 1.3]. Thus, we have H2(X,KXX)=0H^{2}(X,K_{X}-X)=0. Then it suffices to show H1(X,ΩP1|X(KX))=0H^{1}(X,\Omega^{1}_{P}|_{X}(K_{X}))=0.

We have the following exact sequence:

0ΩP1(KP)ΩP1(KP+X)ΩP1|X(KX)00\to\Omega^{1}_{P}(K_{P})\to\Omega^{1}_{P}(K_{P}+X)\to\Omega^{1}_{P}|_{X}(K_{X})\to 0

Since (KP+X)-(K_{P}+X) and KP-K_{P} are ample, we get

H1(P,ΩP1(KP+X))=0andH2(P,ΩP1(KP))=0H^{1}(P,\Omega^{1}_{P}(K_{P}+X))=0\,\,\,\text{and}\,\,\,H^{2}(P,\Omega^{1}_{P}(K_{P}))=0

by Bott vanishing. Thus (1) holds.

Step 2: Proof of (2).   By the conormal exact sequence, we have an exact sequence

0𝒪X(2iKXX)ΩP1|X(2iKX)ΩX1(2iKX)0.0\to\mathcal{O}_{X}(-2^{i}K_{X}-X)\to\Omega^{1}_{P}|_{X}(-2^{i}K_{X})\to\Omega^{1}_{X}(-2^{i}K_{X})\to 0.

Since KX=𝒪X(1)π𝒪1×1(1,1)K_{X}=\mathcal{O}_{X}(-1)\otimes\pi^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1) and X=𝒪P(2)X=\mathcal{O}_{P}(2), we have

H3(X,2iKXX)\displaystyle H^{3}(X,-2^{i}K_{X}-X) H0(X,(2i+1)KX+X)\displaystyle\simeq H^{0}(X,(2^{i}+1)K_{X}+X)
=H0(X,𝒪P(2i+1)π𝒪1×1(2i1,2i1))\displaystyle=H^{0}(X,\mathcal{O}_{P}(-2^{i}+1)\otimes\pi^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2^{i}-1,-2^{i}-1))
=0.\displaystyle=0.

Thus it suffices to show H2(X,ΩP1|X(2iKX))=0H^{2}(X,\Omega^{1}_{P}|_{X}(-2^{i}K_{X}))=0.

We have the following exact sequence:

0ΩP1(2i(KP+X)X)ΩP1(2i(KP+X))ΩP1|X(2iKX)00\to\Omega^{1}_{P}(-2^{i}(K_{P}+X)-X)\to\Omega^{1}_{P}(-2^{i}(K_{P}+X))\to\Omega^{1}_{P}|_{X}(-2^{i}K_{X})\to 0

Since 2i(KP+X)-2^{i}(K_{P}+X) and 2i(KP+X)X-2^{i}(K_{P}+X)-X is nef, we have

H2(P,ΩP1(2i(KP+X))=0andH3(P,ΩP1(2i(KP+X)X))=0H^{2}(P,\Omega^{1}_{P}(-2^{i}(K_{P}+X))=0\,\,\,\text{and}\,\,\,H^{3}(P,\Omega^{1}_{P}(-2^{i}(K_{P}+X)-X))=0

by [Totaro(Fano), Proposition 1.3]. Thus (2) holds. ∎

6.2. FF-splitting for 2-2 and 2-6

Proposition 6.21.

We use the same notation of Lemma 6.4. Assume that dimX=dimY=3\dim X=\dim Y=3. Moreover, suppose that the following hold.

  1. (0)
    1. (0a)

      H1(Y,pHmL)=0H^{1}(Y,pH-mL)=0 for m{2,3}m\in\{2,3\}.

    2. (0b)

      H2(Y,pHmL)=0H^{2}(Y,pH-mL)=0 for m{1,2,3,4}m\in\{1,2,3,4\}.

    3. (0c)

      H3(Y,pHmL)=0H^{3}(Y,pH-mL)=0 for m{1,2,3,4,5}m\in\{1,2,3,4,5\}.

  2. (1)
    1. (1a)

      H1(Y,ΩY1(pHmL))=0H^{1}(Y,\Omega_{Y}^{1}(pH-mL))=0 for m{1,2,3}m\in\{1,2,3\}.

    2. (1b)

      H2(Y,ΩY1(pHmL))=0H^{2}(Y,\Omega_{Y}^{1}(pH-mL))=0 for m{2,3,4}m\in\{2,3,4\}.

    3. (1c)

      H3(Y,ΩY1(pH4L))=0H^{3}(Y,\Omega_{Y}^{1}(pH-4L))=0.

  3. (2)
    1. (2a)

      H1(Y,ΩY2(pHnL))=0H^{1}(Y,\Omega_{Y}^{2}(pH-nL))=0 for n{0,1,2}n\in\{0,1,2\}.

    2. (2b)

      H2(Y,ΩY2(pH2L))=0H^{2}(Y,\Omega_{Y}^{2}(pH-2L))=0.

Then the following hold.

  1. (A)

    H2(X,ΩX1(pKXX))=0H^{2}(X,\Omega_{X}^{1}(-pK_{X}-X))=0.

  2. (B)

    H1(X,ΩP2|X(pKX))=0H^{1}(X,\Omega^{2}_{P}|_{X}(-pK_{X}))=0.

  3. (C)

    H1(X,Ω2(pKX))=0H^{1}(X,\Omega^{2}(-pK_{X}))=0.

Proof.

By taking the wedge product 2\bigwedge^{2} of the conormal exact sequence, we have an exact sequence

0ΩX1(pKXX)ΩP2|X(pKX)ΩX2(pKX)0.0\to\Omega^{1}_{X}(-pK_{X}-X)\to\Omega^{2}_{P}|_{X}(-pK_{X})\to\Omega^{2}_{X}(-pK_{X})\to 0.

Therefore, we get the following implication:

(A)+(B)(C).{\rm(A)}+{\rm(B)}\Rightarrow{\rm(C)}.

In what follows, we shall prove (A) and (B).

Step 1: Proof of (A).   By the conormal exact sequence, we have an exact sequence

0𝒪X(pKX2X)ΩP1|X(pKXX)ΩX1(pKXX)0.0\to\mathcal{O}_{X}(-pK_{X}-2X)\to\Omega^{1}_{P}|_{X}(-pK_{X}-X)\to\Omega^{1}_{X}(-pK_{X}-X)\to 0.

We recall that KX=gH|X=fHK_{X}=-g^{*}H|_{X}=-f^{*}H, X2S2gL=2fLX-2S\sim 2g^{*}L=2f^{*}L, S|X=0S|_{X}=0, and f𝒪X=𝒪Y𝒪Y(L)f_{*}\mathcal{O}_{X}=\mathcal{O}_{Y}\oplus\mathcal{O}_{Y}(-L). We then get

H3(X,pKX2X)\displaystyle H^{3}(X,-pK_{X}-2X) =H3(X,pfH4fL)\displaystyle=H^{3}(X,pf^{*}H-4f^{*}L)
=H3(Y,𝒪Y(pH4L))H3(Y,𝒪Y(pH5L))\displaystyle=H^{3}(Y,\mathcal{O}_{Y}(pH-4L))\oplus H^{3}(Y,\mathcal{O}_{Y}(pH-5L))
=(0c)0.\displaystyle\overset{{\rm(0c)}}{=}0.

Thus it suffices to show H2(X,ΩP1|X(pKXX))=0H^{2}(X,\Omega^{1}_{P}|_{X}(-pK_{X}-X))=0.

Since KX=gH|XK_{X}=-g^{*}H|_{X}, X=2S2gL-X=-2S-2g^{*}L, and S|X=0S|_{X}=0, we have the following exact sequence:

0ΩP1(pgL4gL2S)ΩP1(pgH2gL)ΩP1|X(pKXX)0.0\to\Omega^{1}_{P}(pg^{*}L-4g^{*}L-2S)\to\Omega^{1}_{P}(pg^{*}H-2g^{*}L)\to\Omega^{1}_{P}|_{X}(-pK_{X}-X)\to 0.

By applying Lemma 6.5 for q=2q=2 and D=pH2LD=pH-2L, (0a) and (1b) imply H2(P,ΩP1(pgH2gL))=0H^{2}(P,\Omega^{1}_{P}(pg^{*}H-2g^{*}L))=0. Then the problem is reduced to

H3(P,ΩP1(pgL4gL2S))=0.H^{3}(P,\Omega^{1}_{P}(pg^{*}L-4g^{*}L-2S))=0.

We have the following exact sequence:

0g(ΩY1(pH4L))(2S)ΩP1(pgH4gL2S)ΩP/Y1(pgH4gL2S)𝒪P(pgH5gL4S)00\to g^{*}(\Omega^{1}_{Y}(pH-4L))(-2S)\to\Omega^{1}_{P}(pg^{*}H-4g^{*}L-2S)\\ \to\Omega^{1}_{P/Y}(pg^{*}H-4g^{*}L-2S)\simeq\mathcal{O}_{P}(pg^{*}H-5g^{*}L-4S)\to 0

Thus it is enough to prove that

  1. (I)

    H3(P,g(ΩY1(pH4L))(2S))=0H^{3}(P,g^{*}(\Omega^{1}_{Y}(pH-4L))(-2S))=0 and

  2. (II)

    H3(P,𝒪P(pgH5gL4S))=0H^{3}(P,\mathcal{O}_{P}(pg^{*}H-5g^{*}L-4S))=0.

Step 1-1: Proof of (I).   By using an exact sequence 0𝒪P(S)𝒪P𝒪S00\to\mathcal{O}_{P}(-S)\to\mathcal{O}_{P}\to\mathcal{O}_{S}\to 0 twice, the problem is reduced to

  1. (Ia)

    H2(S,g(ΩY1(pH+(n4)L)))=0H^{2}(S,g^{*}(\Omega^{1}_{Y}(pH+(n-4)L)))=0 for n{0,1}n\in\{0,1\} and

  2. (Ib)

    H3(P,g(ΩY1(pH4L)))=0H^{3}(P,g^{*}(\Omega^{1}_{Y}(pH-4L)))=0.

By

H2(P,g(ΩY1(pH+(n4)L)))=H2(Y,ΩY1(pH+(n4)L)),H^{2}(P,g^{*}(\Omega^{1}_{Y}(pH+(n-4)L)))=H^{2}(Y,\Omega_{Y}^{1}(pH+(n-4)L)),

(Ia) follows from (1b). We have

H3(P,g(ΩY1(pH4L)))H3(Y,ΩY1(pH4L)),H^{3}(P,g^{*}(\Omega^{1}_{Y}(pH-4L)))\simeq H^{3}(Y,\Omega^{1}_{Y}(pH-4L)),

and hence (1c) implies (Ib). This completes thep proof of (I).

Step 1-2: Proof of (II).   By TS+gLT\sim S+g^{*}L, We have H3(P,𝒪P(pgH5gL4S))H3(P,𝒪P(pgHgL4T))H^{3}(P,\mathcal{O}_{P}(pg^{*}H-5g^{*}L-4S))\simeq H^{3}(P,\mathcal{O}_{P}(pg^{*}H-g^{*}L-4T)). By using an exact sequence 0𝒪P(T)𝒪P𝒪T00\to\mathcal{O}_{P}(-T)\to\mathcal{O}_{P}\to\mathcal{O}_{T}\to 0 four times, it is enough to prove (IIa) and (IIb) below.

  1. (IIa)

    H2(T,pgHgLnT)=0H^{2}(T,pg^{*}H-g^{*}L-nT)=0 for n{0,1,2,3}n\in\{0,1,2,3\}.

  2. (IIb)

    H3(P,pgHgL)=0H^{3}(P,pg^{*}H-g^{*}L)=0.

By

H2(T,pgHgLnT)H2(Y,pH(n+1)L),H^{2}(T,pg^{*}H-g^{*}L-nT)\simeq H^{2}(Y,pH-(n+1)L),

(IIa) follows from (0b). We have

H3(P,pgHgL)H3(Y,pHgL),H^{3}(P,pg^{*}H-g^{*}L)\simeq H^{3}(Y,pH-gL),

and hence (0c) implies (IIb). Thus (II) holds.

Step 2: Proof of (B).   Since KX=gH|XK_{X}=-g^{*}H|_{X}, S|X=0S|_{X}=0, and X=2S2gL-X=-2S-2g^{*}L, we have the following exact sequence:

0ΩP2(pgH2gL)ΩP2(pgH+2S)ΩP2|X(pKX)00\to\Omega^{2}_{P}(pg^{*}H-2g^{*}L)\to\Omega^{2}_{P}(pg^{*}H+2S)\to\Omega^{2}_{P}|_{X}(-pK_{X})\to 0

Thus it suffices to show that

  1. (III)

    H1(P,ΩP2(pgH+2S))=0H^{1}(P,\Omega^{2}_{P}(pg^{*}H+2S))=0 and

  2. (IV)

    H2(P,ΩP2(pgH2gL))=0H^{2}(P,\Omega^{2}_{P}(pg^{*}H-2g^{*}L))=0.

Step 2-1: Proof of (III).   Taking the wedge product 2\bigwedge^{2} of the relative exact sequence 0gΩY1ΩP1ΩP/Y100\to g^{*}\Omega^{1}_{Y}\to\Omega^{1}_{P}\to\Omega^{1}_{P/Y}\to 0, we get

0gΩY2ΩP2gΩY1ΩP/Y1gΩY1(gL2S)0.0\to g^{*}\Omega^{2}_{Y}\to\Omega^{2}_{P}\to g^{*}\Omega^{1}_{Y}\otimes\Omega^{1}_{P/Y}\simeq g^{*}\Omega^{1}_{Y}(-g^{*}L-2S)\to 0.

Thus we have

0g(ΩY2(pH))(2S)ΩP2(pgH+2S)g(ΩY1(pHL))0.0\to g^{*}(\Omega^{2}_{Y}(pH))(2S)\to\Omega^{2}_{P}(pg^{*}H+2S)\to g^{*}(\Omega^{1}_{Y}(pH-L))\to 0.

It holds that

H1(P,g(ΩY1(pHL)))=H1(Y,ΩY1(pHL))=(1a)0.H^{1}(P,g^{*}(\Omega^{1}_{Y}(pH-L)))=H^{1}(Y,\Omega^{1}_{Y}(pH-L))\overset{{\rm(1a)}}{=}0.

Then it suffices to show H1(P,g(ΩY2(pH))(2S))=0H^{1}(P,g^{*}(\Omega^{2}_{Y}(pH))(2S))=0. By S|S=gL|SS|_{S}=-g^{*}L|_{S} and an exact sequence 0𝒪P(S)𝒪P𝒪S00\to\mathcal{O}_{P}(-S)\to\mathcal{O}_{P}\to\mathcal{O}_{S}\to 0, the problem is reduced to the vanishings of the following:

  • H1(S,g(ΩY2(pHnL))H1(Y,ΩY2(pHnL))H^{1}(S,g^{*}(\Omega^{2}_{Y}(pH-nL))\simeq H^{1}(Y,\Omega^{2}_{Y}(pH-nL)) for n{1,2}n\in\{1,2\}.

  • H1(P,g(Ω2Y(pH)))H1(Y,Ω2Y(pH))H^{1}(P,g^{*}(\Omega^{2}_{Y}(pH)))\simeq H^{1}(Y,\Omega^{2}_{Y}(pH)).

Both of them follow from (2a). Thus (III) holds.

Step 2-2: Proof of (IV).   By the relative exact sequence 0gΩ1YΩ1PΩ1P/Y𝒪P(gL2S)00\to g^{*}\Omega^{1}_{Y}\to\Omega^{1}_{P}\to\Omega^{1}_{P/Y}\simeq\mathcal{O}_{P}(-g^{*}L-2S)\to 0, we get the following exact sequence:

0g(Ω2Y(pH2L))Ω2P(pgH2gL)g(ΩY1(pH3L))(2S)0.0\to g^{*}(\Omega^{2}_{Y}(pH-2L))\to\Omega^{2}_{P}(pg^{*}H-2g^{*}L)\to g^{*}(\Omega_{Y}^{1}(pH-3L))(-2S)\to 0.

We have H2(P,g(Ω1Y(pH2L)))=H2(Y,Ω2Y(pH2L))=(2b)0H^{2}(P,g^{*}(\Omega^{1}_{Y}(pH-2L)))=H^{2}(Y,\Omega^{2}_{Y}(pH-2L))\overset{{\rm(2b)}}{=}0. By S|S=gL|SS|_{S}=-g^{*}L|_{S} and an exact sequence 0𝒪P(S)𝒪P𝒪S00\to\mathcal{O}_{P}(-S)\to\mathcal{O}_{P}\to\mathcal{O}_{S}\to 0, the vanishing of H2(P,g(ΩY(pgH3gL)(2S))H^{2}(P,g^{*}(\Omega_{Y}(pg^{*}H-3g^{*}L)(-2S)) can be reduced to those of

  • H1(S,g(Ω1Y(pH+(3+n)L)H1(Y,ΩY1(pH+(3+n)L))H^{1}(S,g^{*}(\Omega^{1}_{Y}(pH+(-3+n)L)\simeq H^{1}(Y,\Omega_{Y}^{1}(pH+(-3+n)L)) for n{0,1}n\in\{0,1\} and

  • H2(P,g(Ω1Y(pH3L)))H2(Y,Ω1Y(pH3L))H^{2}(P,g^{*}(\Omega^{1}_{Y}(pH-3L)))\simeq H^{2}(Y,\Omega^{1}_{Y}(pH-3L)).

These follow from (1a) and (1b). Thus (IV) holds. ∎

6.2.1. 2-2

Lemma 6.22.

A smooth Fano threefold XX of No. 2-2 is FF-split if p7p\geq 7.

Proof.

We follow the notation of Lemma 6.7. It is enough to verify the conditions (1)-(4) in Proposition 2.15. Proposition 2.15(1) holds by Lemma 6.2. Lemma 6.8 implies Proposition 2.15(2) and Proposition 2.15(3).

It suffices to show Proposition 2.15(4). It is enough to verify the conditions of Proposition 6.21. Recall that Y=1×2Y=\mathbb{P}^{1}\times\mathbb{P}^{2}, H=𝒪Y(1,1)H=\mathcal{O}_{Y}(1,1), and L=𝒪Y(1,2)L=\mathcal{O}_{Y}(1,2). Since YY is toric, YY satisfies Bott vanishing. Then it is enough to check the following (concerning Proposition 6.21(0), use the fact that pH4LKYpH-4L-K_{Y} is ample):

  1. (0)

    H3(Y,pH5L)=0H^{3}(Y,pH-5L)=0.

  2. (1)

    H2(Y,Ω1Y(pH4L))=H3(Y,Ω1Y(pH4L))=0H^{2}(Y,\Omega^{1}_{Y}(pH-4L))=H^{3}(Y,\Omega^{1}_{Y}(pH-4L))=0.

The assertion (0) follows from

H3(Y,pH5L)=H3(Y,𝒪Y(p5,p10))H1(1,𝒪1(p5))H2(2,𝒪2(p10))=0.H^{3}(Y,pH-5L)=H^{3}(Y,\mathcal{O}_{Y}(p-5,p-10))\simeq H^{1}(\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}}(p-5))\otimes H^{2}(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(p-10))=0.

Let us show (1). We have

Ω1Y(pH4L)\displaystyle\Omega^{1}_{Y}(pH-4L) =Ω1Y(p4,p8)\displaystyle=\Omega^{1}_{Y}(p-4,p-8)
pr1Ω11(p4)pr2𝒪2(p8)\displaystyle\simeq\mathrm{pr}_{1}^{*}\Omega^{1}_{\mathbb{P}^{1}}(p-4)\otimes\mathrm{pr}_{2}^{*}\mathcal{O}_{\mathbb{P}^{2}}(p-8)
pr1𝒪1(p4)pr2Ω12(p8)\displaystyle\oplus\mathrm{pr}_{1}^{*}\mathcal{O}_{\mathbb{P}^{1}}(p-4)\otimes\mathrm{pr}_{2}^{*}\Omega^{1}_{\mathbb{P}^{2}}(p-8)

Then it holds that

H3(Y,Ω1Y(pH4L)\displaystyle H^{3}(Y,\Omega^{1}_{Y}(pH-4L)
=\displaystyle= H1(Ω11(p4))H2(𝒪2(p8))H1(𝒪1(p4))H2(Ω12(p8))\displaystyle H^{1}(\Omega^{1}_{\mathbb{P}^{1}}(p-4))\otimes H^{2}(\mathcal{O}_{\mathbb{P}^{2}}(p-8))\oplus H^{1}(\mathcal{O}_{\mathbb{P}^{1}}(p-4))\otimes H^{2}(\Omega^{1}_{\mathbb{P}^{2}}(p-8))
=\displaystyle= 0\displaystyle 0

and

H2(Y,Ω1Y(pH4L)\displaystyle H^{2}(Y,\Omega^{1}_{Y}(pH-4L)
=\displaystyle= H0(Ω11(p4))H2(𝒪2(p8))H1(Ω11(p4))H1(𝒪2(p8))\displaystyle H^{0}(\Omega^{1}_{\mathbb{P}^{1}}(p-4))\otimes H^{2}(\mathcal{O}_{\mathbb{P}^{2}}(p-8))\oplus H^{1}(\Omega^{1}_{\mathbb{P}^{1}}(p-4))\otimes H^{1}(\mathcal{O}_{\mathbb{P}^{2}}(p-8))
\displaystyle\oplus H0(𝒪1(p4)H2(Ω12(p8))H1(𝒪1(p4))H1(Ω12(p8))\displaystyle H^{0}(\mathcal{O}_{\mathbb{P}^{1}}(p-4)\otimes H^{2}(\Omega^{1}_{\mathbb{P}^{2}}(p-8))\oplus H^{1}(\mathcal{O}_{\mathbb{P}^{1}}(p-4))\otimes H^{1}(\Omega^{1}_{\mathbb{P}^{2}}(p-8))
=\displaystyle= 0,\displaystyle 0,

where we have H2(Ω12(p8))H0(Ω12(8p))=0H^{2}(\Omega^{1}_{\mathbb{P}^{2}}(p-8))\simeq H^{0}(\Omega^{1}_{\mathbb{P}^{2}}(8-p))^{*}=0 by p7p\geq 7 and the Euler exact sequence [Har77, Ch. II, Example 8.20.1], where ():=Homk(,k)(-)^{*}:=\operatorname{Hom}_{k}(-,k). Indeed, if p>7p>7, then this immediately follows from Bott vanishing. When p=7p=7, use the Euler exact sequence 0Ω21(1)𝒪23𝒪2(1)00\to\Omega_{\mathbb{P}^{2}}^{1}(1)\to\mathcal{O}_{\mathbb{P}^{2}}^{\oplus 3}\to\mathcal{O}_{\mathbb{P}^{2}}(1)\to 0, which is still exact even after applying H0(2,)H^{0}(\mathbb{P}^{2},-) by Bott vanishing, and hence h0(Ω21(1))=h0(𝒪23)h0(𝒪2(1))=0h^{0}(\Omega_{\mathbb{P}^{2}}^{1}(1))=h^{0}(\mathcal{O}_{\mathbb{P}^{2}}^{\oplus 3})-h^{0}(\mathcal{O}_{\mathbb{P}^{2}}(1))=0. ∎

6.2.2. 2-6-a

Lemma 6.23.

A smooth Fano threefold of No. 2-6-a is FF-split if p5p\geq 5.

Proof.

We follow the notation of Definition 6.9. It is enough to verify the conditions (1)-(4) in Proposition 2.15. Proposition 2.15(1) holds by Lemma 6.2. Lemma 6.14 implies Proposition 2.15(2) and Proposition 2.15(3).

It suffices to show Proposition 2.15(4). By the first paragraph of the proof of Proposition 6.21, it is enough to prove that

  1. (1)

    H2(X,ΩX1(pKXX))=0H^{2}(X,\Omega_{X}^{1}(-pK_{X}-X))=0.

  2. (2)

    H1(X,ΩP2|X(pKX))=0H^{1}(X,\Omega_{P}^{2}|_{X}(-pK_{X}))=0.

Step 1: Proof of (1).   By the conormal exact sequence, we have an exact sequence

0𝒪X(pKX2X)Ω1P|X(pKXX)Ω1X(pKXX)0.0\to\mathcal{O}_{X}(-pK_{X}-2X)\to\Omega^{1}_{P}|_{X}(-pK_{X}-X)\to\Omega^{1}_{X}(-pK_{X}-X)\to 0.

We have

H3(X,𝒪X(pKX2X))=H3(X,𝒪X(p4,p4))=0H^{3}(X,\mathcal{O}_{X}(-pK_{X}-2X))=H^{3}(X,\mathcal{O}_{X}(p-4,p-4))=0

by p5p\geq 5 and Lemma 6.10. Thus it suffices to show H2(X,Ω1P|X(pKXX)=0H^{2}(X,\Omega^{1}_{P}|_{X}(-pK_{X}-X)=0.

We have the following exact sequence:

0Ω1P(p(KP+X)2X)Ω1P(p(KP+X)X)Ω1P|X(pKXX)0.0\to\Omega^{1}_{P}(-p(K_{P}+X)-2X)\to\Omega^{1}_{P}(-p(K_{P}+X)-X)\to\Omega^{1}_{P}|_{X}(-pK_{X}-X)\to 0.

Then the required vanishing H2(X,Ω1P|X(pKXX)=0H^{2}(X,\Omega^{1}_{P}|_{X}(-pK_{X}-X)=0 follows from

H2(P,Ω1P(p(KP+X)X))=H2(P,Ω1P(p2,p2))=0H^{2}(P,\Omega^{1}_{P}(-p(K_{P}+X)-X))=H^{2}(P,\Omega^{1}_{P}(p-2,p-2))=0

and

H3(P,Ω1P(p(KP+X)2X))=H3(P,Ω1P(p4,p4))=0,H^{3}(P,\Omega^{1}_{P}(-p(K_{P}+X)-2X))=H^{3}(P,\Omega^{1}_{P}(p-4,p-4))=0,

where each vanishing follows from Bott vanishing. Thus (1) holds.

Step 2: Proof of (2).   We have the following exact sequence:

0Ω2P(p(KP+X)X)Ω2P(p(KP+X))Ω2P|X(pKX)0.0\to\Omega^{2}_{P}(-p(K_{P}+X)-X)\to\Omega^{2}_{P}(-p(K_{P}+X))\to\Omega^{2}_{P}|_{X}(-pK_{X})\to 0.

By Bott vanishing, we get

H1(P,Ω2P(p(KP+X)))=H1(P,Ω2P(p,p))=0H^{1}(P,\Omega^{2}_{P}(-p(K_{P}+X)))=H^{1}(P,\Omega^{2}_{P}(p,p))=0

and

H2(P,Ω2P(p(KP+X)X))=H2(P,Ω2P(p2,p2))=0.H^{2}(P,\Omega^{2}_{P}(-p(K_{P}+X)-X))=H^{2}(P,\Omega^{2}_{P}(p-2,p-2))=0.

Therefore, (2) holds. ∎

6.2.3. 2-6-b

Lemma 6.24.

A smooth Fano threefold XX of No. 2-6-b is FF-split if p5p\geq 5.

Proof.

We follow the notation of Lemma 6.12. It is enough to verify the conditions (1)-(4) in Proposition 2.15. Proposition 2.15(1) holds by Lemma 6.2. Lemma 6.14 implies Proposition 2.15(2) and Proposition 2.15(3).

It suffices to show Proposition 2.15(4). It is enough to verify the conditions of Proposition 6.21. Recall that W𝒪2×2(1,1)W\in\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1) and L=H=𝒪Y(1,1)L=H=\mathcal{O}_{Y}(1,1). Then all the conditions of Proposition 6.21 hold by Lemma 6.13 and Serre duality. ∎

By a similar argument, we obtain an analogous result for the hyperelliptic case. We shall later prove that the assumption on pp in (1) is optimal (Example 8.4).

Proposition 6.25.

Let XX be a smooth Fano threefold such that ρ(X)=rX=1\rho(X)=r_{X}=1 and |KX||-K_{X}| is not very ample, where rXr_{X} denotes the index of XX. Let f:XYf:X\to Y be the double cover induced by |KX||-K_{X}|, where Y{3,Q}Y\in\{\mathbb{P}^{3},Q\} (cf. [FanoI, Theorem 6.5]). Then the following hold.

  1. (1)

    If p13p\geq 13 and Y=3Y=\mathbb{P}^{3}, then XX is FF-split.

  2. (2)

    If p11p\geq 11 and Y=QY=Q, then XX is FF-split.

Proof.

We use the same notation of the proof of Proposition 6.16. By the same argument as in 2-6-b (cf. the proof of Lemma 6.24), it is enough to verify Proposition 2.15(4), which follows from Lemma 6.13. ∎

7. Proofs of the main theorems

In this section, we prove Theorems A, B, C, and D.

Lemma 7.1.

Let XX be a smooth projective variety. Take an ample Cartier divisor AA on XX. Suppose that XX is quasi-FF-split. Then there exists n0>0n_{0}>0 such that Hi(X,BnΩ1X(pnA))=0H^{i}(X,B_{n}\Omega^{1}_{X}(-p^{n}A))=0 for every nn0n\geq n_{0} and every i<dimXi<\dim X.

Proof.

Fix n1n_{1} such that XX is n1n_{1}-quasi-FF-split. Pick integers nn and \ell satisfying nn1n\geq\ell\geq n_{1}. By definition, XX is \ell-quasi-FF-split. By [KTTWYY1, Lemma 3.8 and Lemma 5.9], we get an isomorphism

QX,pnA,𝒪X(pnA)BΩX1(pnA).Q_{X,-p^{n-\ell}A,\ell}\simeq\mathcal{O}_{X}(-p^{n-\ell}A)\oplus B_{\ell}\Omega_{X}^{1}(-p^{n}A).

and an exact sequence

0FB1ΩX1(pnA)QX,pnA,F𝒪X(pn+1A)0.0\to F_{*}B_{\ell-1}\Omega_{X}^{1}(-p^{n}A)\to Q_{X,-p^{n-\ell}A,\ell}\to F_{*}\mathcal{O}_{X}(-p^{n-\ell+1}A)\to 0.

Since XX satisfies Kodaira vanishing [KTTWYY1, Theorem 3.15], the following hold:

Hi(X,BΩX1(pnA))Hi(X,QX,pnA,n)Hi(X,B1ΩX1(pnA)).H^{i}(X,B_{\ell}\Omega_{X}^{1}(-p^{n}A))\simeq H^{i}(X,Q_{X,-p^{n-\ell}A,n})\simeq H^{i}(X,B_{\ell-1}\Omega_{X}^{1}(-p^{n}A)).

By the Serre vanishing theorem, we can find n(i)n(i) such that the following holds for nn(i)n\geq n(i):

Hi(X,BnΩX1(pnA))\displaystyle H^{i}(X,B_{n}\Omega_{X}^{1}(-p^{n}A)) \displaystyle\simeq Hi(X,Bn1ΩX1(pnA))\displaystyle H^{i}(X,B_{n-1}\Omega_{X}^{1}(-p^{n}A))
\displaystyle\simeq Hi(X,Bn2ΩX1(pnA))\displaystyle H^{i}(X,B_{n-2}\Omega_{X}^{1}(-p^{n}A))
\displaystyle\simeq \displaystyle\cdots
\displaystyle\simeq Hi(X,Bn1ΩX1(pnA))=0.\displaystyle H^{i}(X,B_{n_{1}}\Omega_{X}^{1}(-p^{n}A))=0.

Hence we are done by setting n0max{n(0),,n(dimX1)}n_{0}\coloneqq\max\{n(0),\ldots,n(\dim X-1)\}. ∎

Lemma 7.2.

Let XX be a smooth projective threefold. Take an ample Cartier divisor AA. Suppose that

  1. (1)

    H0(X,ΩX2(pmA))=0H^{0}(X,\Omega_{X}^{2}(-p^{m}A))=0 for every integer m>0m>0, and

  2. (2)

    H2(X,BnΩX1(pnA))=0H^{2}(X,B_{n}\Omega_{X}^{1}(-p^{n}A))=0 for every sufficiently large integer n0n\gg 0.

Then

H1(X,ΩX1(A))=0.H^{1}(X,\Omega_{X}^{1}(-A))=0.
Proof.

Fix n0n\gg 0. By (2.1.1), we have an exact sequence

0BnΩX1(pnA)ZnΩX1(pnA)ΩX1(A)0.0\to B_{n}\Omega_{X}^{1}(-p^{n}A)\to Z_{n}\Omega_{X}^{1}(-p^{n}A)\to\Omega_{X}^{1}(-A)\to 0.

By (2), the required vanishing is reduced to H1(X,ZnΩX1(pnA))=0H^{1}(X,Z_{n}\Omega_{X}^{1}(-p^{n}A))=0. By (2.1.2), we have the following exact sequence

0ZnΩX1(pnA)FZn1ΩX1(pnA)B1ΩX2(pA)0.0\to Z_{n}\Omega_{X}^{1}(-p^{n}A)\to F_{*}Z_{n-1}\Omega_{X}^{1}(-p^{n}A)\to B_{1}\Omega_{X}^{2}(-pA)\to 0.

Since H0(X,B1ΩX2(pA))H0(X,ΩX2(pA))=0H^{0}(X,B_{1}\Omega_{X}^{2}(-pA))\subset H^{0}(X,\Omega_{X}^{2}(-pA))=0 by (1), it suffices to show that H1(X,Zn1ΩX1(pnA))=0H^{1}(X,Z_{n-1}\Omega_{X}^{1}(-p^{n}A))=0. By repeating this procedure, it is enough to prove

H1(X,Ω1X(pnA))=0,H^{1}(X,\Omega^{1}_{X}(-p^{n}A))=0,

which follows from n0n\gg 0 and the Serre vanishing theorem. ∎

Proposition 7.3.

Let XX be a smooth Fano threefold with ρ(X)2\rho(X)\geq 2. Then the following hold.

  1. (1)

    XX is SRC if XX is none of 2-1, 2-3, 2-5, 2-10, and 2-14.

  2. (2)

    H0(X,ΩX2(D))=0H^{0}(X,\Omega_{X}^{2}(-D))=0 for every pseudo-effective Cartier divisor DD on XX.

Proof.

Let us show (1). Note that XX is SRC if XX is rational or XX has a conic bundle structure (Lemma 6.2). We first treat the case when ρ(X)3\rho(X)\geq 3. In this case, XX has a conic bundle structure except when XX is 3-18 [FanoIV, Theorem 4.10, Theorem 5.2]. If XX is 3-18, then XX is rational by [FanoIV, Section 7.3]. Hence we may assume that ρ(X)=2\rho(X)=2. Since 3\mathbb{P}^{3} and QQ are rational, XX is rational or has an extremal ray of type CC except when XX is one of 2-1, 2-3, 2-5, 2-10, and 2-14 [FanoIV, Section 7.2]. Thus (1) holds.

Let us show (2). By (1) and Lemma 6.1, we may assume that XX is one of 2-1, 2-3, 2-5, 2-10, and 2-14. It follows from [FanoIV, Section 7.2] that

  • there is a blowup f:XVdf\colon X\to V_{d} to a smooth Fano threefold VdV_{d} of index 22 with (KVd)3=8d(-K_{V_{d}})^{3}=8d for some 1d51\leq d\leq 5 and

  • we have a del Pezzo fibration g:X1g\colon X\to\mathbb{P}^{1}.

Take a general fibre FF of gg. Then FF is a canonical del Pezzo surface by [FS20, Theorem 15.2] and [BT22, Theorem 3.3]. By construction, we see that FF has a smooth anti-canonical member. In fact, f:XYVdf\colon X\to Y\coloneqq V_{d} is a blowup along an elliptic curve which is a complete intersection of two members of |12KY||-\frac{1}{2}K_{Y}|, and this blowup centre is isomorphic to some member of |KF||-K_{F}|.

Since fΩ2X(D)f_{*}\Omega^{2}_{X}(-D) is torsion-free, we get

fΩ2X(D)(fΩ2X(D))=Ω2Y(DY),f_{*}\Omega^{2}_{X}(-D)\hookrightarrow(f_{*}\Omega^{2}_{X}(-D))^{**}=\Omega^{2}_{Y}(-D_{Y}),

where DYfDD_{Y}\coloneqq f_{*}D. Thus we have

H0(X,Ω2X(D))H0(Y,Ω2Y(DY)).H^{0}(X,\Omega^{2}_{X}(-D))\hookrightarrow H^{0}(Y,\Omega^{2}_{Y}(-D_{Y})).

Then it is enough to prove H0(Y,Ω2Y(DY))=0H^{0}(Y,\Omega^{2}_{Y}(-D_{Y}))=0. By Pic(Y)=\operatorname{Pic}(Y)=\mathbb{Z}, the divisor DYD_{Y} is nef. Set SfF|12KY|S\coloneqq f_{*}F\in|-\frac{1}{2}K_{Y}|. Since FF is general and the base locus of |12KY||-\frac{1}{2}K_{Y}| is either empty or zero-dimensional, it suffices to show H0(S,ΩY2(DY)|S)=0H^{0}(S,\Omega_{Y}^{2}(-D_{Y})|_{S})=0, because it implies H0(Y,ΩY2(DY))=0H^{0}(Y,\Omega_{Y}^{2}(-D_{Y}))=0 (indeed, if there is 0sH0(Y,ΩY2(DY))0\neq s\in H^{0}(Y,\Omega_{Y}^{2}(-D_{Y})), then we could find a member SSupp(s)S\not\subset\operatorname{Supp}(s), which implies 0s|FH0(S,ΩY2(DY)|S)0\neq s|_{F}\in H^{0}(S,\Omega_{Y}^{2}(-D_{Y})|_{S})). Since SS is smooth along the blowup centre, we get f|F:FSf|_{F}\colon F\xrightarrow{\simeq}S, and hence SS is a canonical del Pezzo surface. Let SsmS_{\mathrm{sm}} be the smooth locus of SS. By taking the exterior power 2\wedge^{2} of the conormal exact sequence and applying the tensor product with 𝒪Y(DY)\mathcal{O}_{Y}(-D_{Y}), we get an exact sequence

0ΩSsm1(SDY)Ω2Y(DY)|SsmΩSsm2(DY)0.0\to\Omega_{S_{\mathrm{sm}}}^{1}(-S-D_{Y})\to\Omega^{2}_{Y}(-D_{Y})|_{S_{\mathrm{sm}}}\to\Omega_{S_{\mathrm{sm}}}^{2}(-D_{Y})\to 0.

Applying the pushforward ii_{*} by the inclusion i:SsmSi\colon S_{\mathrm{sm}}\hookrightarrow S (which is left exact), we obtain an exact sequence

0ΩS[1](SDY)Ω2Y(DY)|SΩS[2](DY)0\to\Omega_{S}^{[1]}(-S-D_{Y})\to\Omega^{2}_{Y}(-D_{Y})|_{S}\to\Omega_{S}^{[2]}(-D_{Y})

By H0(S,ΩS[2](DY))=H0(S,𝒪S(KSDY)))=0H^{0}(S,\Omega_{S}^{[2]}(-D_{Y}))=H^{0}(S,\mathcal{O}_{S}(K_{S}-D_{Y})))=0, it is enough to prove H0(S,ΩS[1](SDY))=0H^{0}(S,\Omega_{S}^{[1]}(-S-D_{Y}))=0.

Pick a general member CC of |KS||-K_{S}|, which is an elliptic curve. Since SS is smooth along CC, we have ΩS[1](SDY)|C=ΩS1|C𝒪C(𝒪Y(SDY)|C)\Omega_{S}^{[1]}(-S-D_{Y})|_{C}=\Omega_{S}^{1}|_{C}\otimes_{\mathcal{O}_{C}}(\mathcal{O}_{Y}(-S-D_{Y})|_{C}). Then it suffices to show

H0(C,ΩS1|C𝒪C(𝒪Y(SDY)|C))=0.H^{0}(C,\Omega_{S}^{1}|_{C}\otimes_{\mathcal{O}_{C}}(\mathcal{O}_{Y}(-S-D_{Y})|_{C}))=0.

By the conormal exact sequence, we have an exact sequence

0(𝒪S(C)|C)(𝒪Y(SDY)|C)Ω1S|C(𝒪Y(SDY)|C)ΩC1(𝒪Y(SDY)|C)0.0\to(\mathcal{O}_{S}(-C)|_{C})\otimes(\mathcal{O}_{Y}(-S-D_{Y})|_{C})\to\Omega^{1}_{S}|_{C}\otimes(\mathcal{O}_{Y}(-S-D_{Y})|_{C})\\ \to\Omega_{C}^{1}\otimes(\mathcal{O}_{Y}(-S-D_{Y})|_{C})\to 0.

Then the required vanishing H0(C,Ω1S|C(𝒪Y(SDY)|C))=0H^{0}(C,\Omega^{1}_{S}|_{C}\otimes(\mathcal{O}_{Y}(-S-D_{Y})|_{C}))=0 follows from

  • H0(C,(𝒪S(C)|C)(𝒪Y(SDY)|C))=H0(C,𝒪Y(KY+S)|C𝒪Y(SDY)|C)=H0(C,𝒪Y(KYDY)|C))=0H^{0}(C,(\mathcal{O}_{S}(-C)|_{C})\otimes(\mathcal{O}_{Y}(-S-D_{Y})|_{C}))=H^{0}(C,\mathcal{O}_{Y}(K_{Y}+S)|_{C}\otimes\mathcal{O}_{Y}(-S-D_{Y})|_{C})=H^{0}(C,\mathcal{O}_{Y}(K_{Y}-D_{Y})|_{C}))=0, and

  • H0(C,Ω1C(𝒪Y(SDY)|C))=H0(C,𝒪Y(SDY)|C)=0H^{0}(C,\Omega^{1}_{C}\otimes(\mathcal{O}_{Y}(-S-D_{Y})|_{C}))=H^{0}(C,\mathcal{O}_{Y}(-S-D_{Y})|_{C})=0.

Proof of Theorem A.

Let us show (1). Assume rX>1r_{X}>1. If rX3r_{X}\geq 3, then X3X\simeq\mathbb{P}^{3} or XX is isomorphic to a smooth quadric threefold, and hence XX is FF-split. If rX=2r_{X}=2, then XX is quasi-FF-split by [Kawakami-Tanaka(dPvar), Theorem A and Remark 2.8]. Therefore, we may assume that ρ(X)2\rho(X)\geq 2. In this case, the assertion holds by former parts as follows:

  • ρ(X)6\rho(X)\geq 6: Proposition 3.1.

  • ρ(X)=5\rho(X)=5: Proposition 3.2.

  • ρ(X)=4\rho(X)=4: Proposition 3.3.

  • ρ(X)=3\rho(X)=3: Proposition 4.4and Lemma 6.20.

  • ρ(X)=2\rho(X)=2: Proposition 5.12, Lemma 6.8 (2-2), Lemma 6.11, Lemma 6.14, and Lemma 6.18.

We now show (2), i.e., Kodaira vanishing for XX. When ρ(X)2\rho(X)\geq 2 or rX2r_{X}\geq 2, this follows from (1) and Kodaira vanishing for quasi-FF-split varieties [KTTWYY1, Theorem 3.15]. When ρ(X)=rX=1\rho(X)=r_{X}=1, then Kodaira vanishing holds by [FanoI, Corollary 4.5]. ∎

Proof of Theorem B.

The assertion (2) follows from Theorem A(1) and [KTTWYY1, Corollary 7.6]. Let us show (1), i.e., Hj(X,ΩiX(A))=0H^{j}(X,\Omega^{i}_{X}(-A))=0 for i+j<3i+j<3. If i=0i=0 (resp. (i,j)=(1,0)(i,j)=(1,0), resp. (i,j)=(2,0)(i,j)=(2,0)), then this follows from Theorem A(2) (resp. [Kaw1, Theorem 3.5], resp. Proposition 7.3). The remaining case when (i,j)=(1,2)(i,j)=(1,2) is settled by Lemma 7.1 and Lemma 7.2. ∎

Proof of Theorem C.

By Theorem B(1), we have H2(X,TX)H1(X,ΩX1(KX))=0H^{2}(X,T_{X})\simeq H^{1}(X,\Omega_{X}^{1}(K_{X}))=0. Combining with H2(X,𝒪X)=0H^{2}(X,\mathcal{O}_{X})=0, we conclude from [FAG, Theorem 8.5.19] that XX lifts to W(k)W(k). ∎

Corollary 7.4.

Let f:XYf\colon X\to Y be a blowup along a smooth curve, where XX and YY are smooth Fano threefolds. Then CYC\subset Y admits a W(k)W(k)-lifting 𝒞𝒴\mathcal{C}\subset\mathcal{Y}, and the blowup Bl𝒞(𝒴)𝒴\mathrm{Bl}_{\mathcal{C}}(\mathcal{Y})\to\mathcal{Y} along 𝒞\mathcal{C} is a W(k)W(k)-lifting of f:XYf\colon X\to Y.

Proof.

By ρ(X)2\rho(X)\geq 2, Theorem B(1) implies

H2(X,TX)H1(X,Ω1X(KX))=0,H^{2}(X,T_{X})\simeq H^{1}(X,\Omega^{1}_{X}(K_{X}))=0,

and thus XX formally lifts to W(k)W(k) [FAG, Theorem 8.5.9(b)]. Moreover, we have H1(E,NE/X)=H1(E,𝒪X(E)|X)=H1(E,𝒪E(KEKX|E))=0H^{1}(E,N_{E/X})=H^{1}(E,\mathcal{O}_{X}(E)|_{X})=H^{1}(E,\mathcal{O}_{E}(K_{E}-K_{X}|_{E}))=0 by Kodaira vanishing [Muk13, Theorem 3]. Thus EE formally lifts to W(k)W(k) as a closed subscheme of XX by [Har2, Theorem 6.2]. Since Rf𝒪X=𝒪YRf_{*}\mathcal{O}_{X}=\mathcal{O}_{Y} and Rf𝒪E=𝒪CRf_{*}\mathcal{O}_{E}=\mathcal{O}_{C}, it follows that YY formally lifts to W(k)W(k) and CC formally lifts to W(k)W(k) as a closed subscheme by [AZ-nonlift, Proposition 2.3(1)]. Since H2(Y,𝒪Y)=0H^{2}(Y,\mathcal{O}_{Y})=0, they are algebraisable by [FAG, Corollary 8.5.6 and Corollary 8.4.5]. By [MR2791606, Theorem 2.5.8], for every closed point y𝒞y\in\mathcal{C}, there exists an affine open neighbourhood 𝒰𝒴\mathcal{U}\subset\mathcal{Y} of yy and an étale morphism ϕ~:𝒰SpecW(k)[t1,t2,t3]\tilde{\phi}\colon\mathcal{U}\to\operatorname{Spec}\,W(k)[t_{1},t_{2},t_{3}] such that 𝒞𝒰=V(ϕ~t1,ϕ~t2)\mathcal{C}\cap\mathcal{U}=V(\tilde{\phi}^{*}t_{1},\tilde{\phi}^{*}t_{2}). Set U𝒰W(k)kU\coloneqq\mathcal{U}\otimes_{W(k)}k and ϕϕ~W(k)k:USpeck[t1,t2,t3]\phi\coloneqq\tilde{\phi}\otimes_{W(k)}k\colon U\to\operatorname{Spec}\,k[t_{1},t_{2},t_{3}]. Then f|U=BlV(ϕt1,ϕt2)(U)f|_{U}=\mathrm{Bl}_{V(\phi^{*}t_{1},\phi^{*}t_{2})}(U), and we conclude that Bl𝒞(𝒰)𝒰\mathrm{Bl}_{\mathcal{C}}(\mathcal{U})\to\mathcal{U} is a W(k)W(k)-lifting of f|Uf|_{U}. ∎

Corollary 7.5.

Let XX be a smooth Fano threefold and let f:XYf\colon X\to Y be a morphism to a normal projective variety YY such that f𝒪X=𝒪Yf_{*}\mathcal{O}_{X}=\mathcal{O}_{Y} and ff is not isomorphism. Then ff lifts to W(k)W(k).

Proof.

Since ff is not isomorphism, we have ρ(X)2\rho(X)\geq 2. Then the Fano threefold XX admits a W(k)W(k)-lifting 𝒳\mathcal{X}. Let LYL_{Y} be an ample invertible sheaf on YY and set LfLYL\coloneqq f^{*}L_{Y}. By H2(X,𝒪X)=0H^{2}(X,\mathcal{O}_{X})=0, there exists a W(k)W(k)-lifting Pic(𝒳)\mathcal{L}\in\operatorname{Pic}(\mathcal{X}) of LL. We take m>0m\in\mathbb{Z}_{>0} such that f=φ|mL|f=\varphi_{|mL|}, where ϕ|mL|\phi_{|mL|} is the contraction associated to the complete linear system |mL||mL|. By Kodaira vanishing (Theorem A), we have Hi(X,𝒪X(mL))=0H^{i}(X,\mathcal{O}_{X}(mL))=0. This, together with upper-semicontinuity theorem, implies Hi(𝒳,𝒪X(m))=0H^{i}(\mathcal{X},\mathcal{O}_{X}(m\mathcal{L}))=0. It follows from the Grauert theorem that

H0(X,𝒪X(mL))H0(𝒳,𝒪X(m))W(k)k.H^{0}(X,\mathcal{O}_{X}(mL))\simeq H^{0}(\mathcal{X},\mathcal{O}_{X}(m\mathcal{L}))\otimes_{W(k)}k.

Let 𝒵\mathcal{Z} be the base locus of |m||m\mathcal{L}|. We show that 𝒵=\mathcal{Z}=\emptyset. Since 𝒳\mathcal{X} is projective over W(k)W(k), the image of 𝒵\mathcal{Z} in SpecW(k)\operatorname{Spec}\,W(k) is closed. Since |mL||mL| is free, the image does not contain the closed point of SpecW(k)\operatorname{Spec}\,W(k). Thus 𝒵=\mathcal{Z}=\emptyset. Then we can see that the contraction φm\varphi_{m\mathcal{L}} gives a lift of ff. ∎

Proof of Theorem D.

The assertion follows from Theorem 5.13 and Section 6.2. ∎

8. Examples

In this section, we gather examples of non-FF-split or non-quasi-FF-split smooth Fano threefolds.

Example 8.1 (X=S×1X=S\times\mathbb{P}^{1}).

Let SS be a smooth del Pezzo surface which is not FF-split. Then XS×1X\coloneqq S\times\mathbb{P}^{1} is a Fano threefold which is not FF-split. Therefore, if the characteristic pp of the base field kk and ρ\rho satisfies one of (1)-(3) below, then there exists a non-FF-split smooth Fano threefold XX over kk satisfying ρ(X)=ρ\rho(X)=\rho.

  1. (1)

    p=2p=2 and ρ=7\rho=7

  2. (2)

    p{2,3}p\in\{2,3\} and ρ=8\rho=8.

  3. (3)

    p{2,3,5}p\in\{2,3,5\} and ρ=9\rho=9.

Example 8.2 (Wild conic bundles).

Wild conic bundles are not FF-split. Indeed, if XX is FF-split and f:XSf:X\to S is a conic bundle, then ff is generically reduced [GLP15, Lemma 2.4] and hence not wild. Therefore, if p=2p=2 and XX is a smooth Fano threefold which is 2-24 or 3-10, then XX is not necessarily FF-split.

Example 8.3 (p=7p=7, non-FF-split).

Assume p=7p=7. Then

X:={x04+x14+x24+x34+x44=0}4X:=\{x_{0}^{4}+x_{1}^{4}+x_{2}^{4}+x_{3}^{4}+x_{4}^{4}=0\}\subset\mathbb{P}^{4}

is not FF-split. Indeed, we have

(x04+x14+x24+x34+x44)6(x07,x17,x27,x37,x47),(x_{0}^{4}+x_{1}^{4}+x_{2}^{4}+x_{3}^{4}+x_{4}^{4})^{6}\not\in(x_{0}^{7},x_{1}^{7},x_{2}^{7},x_{3}^{7},x_{4}^{7}),

and hence Fedder’s criterion [Fed83, Proposition 2.1] implies that XX is not FF-split (cf. [KTY, Proposition A.8]).

Example 8.4 (p=11p=11, non-FF-split).

Assume p=11p=11. Take

X:={x06+x16+x26+x36+y2=0}(1,1,1,1,3),X:=\{x_{0}^{6}+x_{1}^{6}+x_{2}^{6}+x_{3}^{6}+y^{2}=0\}\subset\mathbb{P}(1,1,1,1,3),

i.e., (1,1,1,1,3)=Projk[x0,x1,x2,x3,y]\mathbb{P}(1,1,1,1,3)=\mathrm{Proj}\,k[x_{0},x_{1},x_{2},x_{3},y] with degxi=1\deg x_{i}=1 and degy=3\deg y=3 for every 0i30\leq i\leq 3, and

X:=Projk[x0,x1,x2,x3,y](x06+x16+x26+x36+y2).X:=\mathrm{Proj}\,\frac{k[x_{0},x_{1},x_{2},x_{3},y]}{(x_{0}^{6}+x_{1}^{6}+x_{2}^{6}+x_{3}^{6}+y^{2})}.

Let us prove that

  1. (1)

    XX is not FF-split, and

  2. (2)

    XX is a smooth Fano threefold.

The assertion (1) follows from Fedder’s criterion [Fed83, Proposition 2.1] (cf. [KTY, Proposition A.8]).

Let us show (2). It is enough to prove that XX is smooth, as the other assertions in (2) follow from the adjunction formula and the fact that XX is an ample \mathbb{Q}-Cartier effective Weil divisor on (1,1,1,1,3)\mathbb{P}(1,1,1,1,3) (which implies the connectedness of XX). Suppose that [a0:a1:a2:a3:b][a_{0}:a_{1}:a_{2}:a_{3}:b] is a singular point of XX, where a0,a1,a2,a3,bka_{0},a_{1},a_{2},a_{3},b\in k. Recall that we have [a0:a1:a2:a3:b]=[λa0:λa1:λa2:λa3:λ3b][a_{0}:a_{1}:a_{2}:a_{3}:b]=[\lambda a_{0}:\lambda a_{1}:\lambda a_{2}:\lambda a_{3}:\lambda^{3}b] for every λk{0}\lambda\in k\setminus\{0\}. For X1:=x1/x0,X2:=x2/x0,X3:=x3/x0,Y:=y/x03X_{1}:=x_{1}/x_{0},X_{2}:=x_{2}/x_{0},X_{3}:=x_{3}/x_{0},Y:=y/x_{0}^{3}, we have

D+(x0)=Speck[X1,X2,X3,Y](𝔸4)(1,1,1,1,3).D_{+}(x_{0})=\operatorname{Spec}k\left[X_{1},X_{2},X_{3},Y\right](\simeq\mathbb{A}^{4})\subset\mathbb{P}(1,1,1,1,3).

and

XD+(x0)={1+X16+X26+X36+Y2=0},X\cap D_{+}(x_{0})=\{1+X_{1}^{6}+X_{2}^{6}+X_{3}^{6}+Y^{2}=0\},

which is smooth. Then it holds that a0=0a_{0}=0. By symmetry, we get a0=a1=a2=a3=0a_{0}=a_{1}=a_{2}=a_{3}=0, which implies [a0:a1:a2:a3:b]=[0:0:0:0:b]=[0:0:0:0:1][a_{0}:a_{1}:a_{2}:a_{3}:b]=[0:0:0:0:b]=[0:0:0:0:1]. However, XX does not pass through [0:0:0:0:1][0:0:0:0:1], which is absurd. Thus (2) holds.

Example 8.5 (No. 2-3, p=3p=3, non-FF-split).

Assume p=3p=3. We construct a Fano threefold XX which is 2-3 and not FF-split. Let V2V_{2} be a Fano threefold of index 22 such that (KV2)3=16(-K_{V_{2}})^{3}=16 and V2V_{2} is not FF-split (e.g., V2:={x04+x14+x24+x34+y2=0}V_{2}:=\{x_{0}^{4}+x_{1}^{4}+x_{2}^{4}+x_{3}^{4}+y^{2}=0\} in (1,1,1,1,2)=Projk[x0,x1,x2,x3,y]\mathbb{P}(1,1,1,1,2)=\mathrm{Proj}\,k[x_{0},x_{1},x_{2},x_{3},y] for degx0=degx1=degx2=degx3\deg x_{0}=\deg x_{1}=\deg x_{2}=\deg x_{3} and degy=2\deg y=2, cf. the proof of Example 8.4). Take a Cartier divisor HH on V2V_{2} such that 2HKV22H\sim-K_{V_{2}}. Then |H||H| is base point free and it induces a finite double cover. By a Bertini theorem [Spr98, Corollary 4.3], we may assume that HH is a smooth prime divisor on V2V_{2}, which is a smooth del Pezzo surface with KH2=2K_{H}^{2}=2. Pick a general member CC of |KH||-K_{H}|, which is a smooth elliptic curve [KN, Theorem 1.4]. By an exact sequence

H0(V2,𝒪V2(H))H0(H,𝒪V2(H)|H)H1(V2,𝒪V2(HH))=H1(V2,𝒪V2)=0,H^{0}(V_{2},\mathcal{O}_{V_{2}}(H))\to H^{0}(H,\mathcal{O}_{V_{2}}(H)|_{H})\to H^{1}(V_{2},\mathcal{O}_{V_{2}}(H-H))=H^{1}(V_{2},\mathcal{O}_{V_{2}})=0,

there exists a member H|H|H^{\prime}\in|H| such that HH=H|H=CH\cap H^{\prime}=H^{\prime}|_{H}=C. Let σ:XV2\sigma:X\to V_{2} be the blowup along CC. Since σ\sigma coincides with the resolution of the indeterminacies of the pencil generated by HH and HH^{\prime}, there is a contraction π:X1\pi:X\to\mathbb{P}^{1} of type DD such that the proper transforms of HH and HH^{\prime} are fibres of π\pi. By Kleimann’s criterion, XX is a smooth Fano threefold, which is of No. 2-3.

Example 8.6 (No. 2-1, p=5p=5, non-FF-split).

Assume p=5p=5. We construct a Fano threefold XX which is 2-1 and XX is not FF-split. Let V1V_{1} be a Fano threefold of index 22 such that (KV1)3=8(-K_{V_{1}})^{3}=8 and V1V_{1} is not FF-split. We can find such an example by setting

V1:={x06+x16+x26+y3+z2=0}(1,1,1,2,3)=Projk[x0,x1,x2,y,z],V_{1}:=\{x_{0}^{6}+x_{1}^{6}+x_{2}^{6}+y^{3}+z^{2}=0\}\subset\mathbb{P}(1,1,1,2,3)=\mathrm{Proj}\,k[x_{0},x_{1},x_{2},y,z],

where degx0=degx1=degx2=1,degy=2,degz=3\deg x_{0}=\deg x_{1}=\deg x_{2}=1,\deg y=2,\deg z=3 (cf. [Oka21, Section 3.1]). Take a Cartier divisor HH on V1V_{1} such that 2HKV12H\sim-K_{V_{1}}. Then we have a scheme-theoretic equality Bs|H|=P{\rm Bs}\,|H|=P for some closed point PP on XX. We take generic members Hgen1H^{\operatorname{gen}}_{1} and Hgen2H^{\operatorname{gen}}_{2} of |H||H| twice. Then C=Hgen1Hgen2C=H^{\operatorname{gen}}_{1}\cap H^{\operatorname{gen}}_{2} is a regular curve of genus one. By p=5>3p=5>3, CC is a smooth elliptic curve [PW22, Corollary 1.8]. Therefore, the intersection C:=H1H2C:=H_{1}\cap H_{2} of two general members H1H_{1} and H2H_{2} of |H||H| is a smooth elliptic curve. Take the blowup XV1X\to V_{1} along CC. Then we can apply the same argument as in Example 8.6.

Example 8.7 (p=5p=5, non-quasi-FF-split).

Assume p=5p=5. Take

X:={x06+x16+x26+x36+y2=0}(1,1,1,1,3),X:=\{x_{0}^{6}+x_{1}^{6}+x_{2}^{6}+x_{3}^{6}+y^{2}=0\}\subset\mathbb{P}(1,1,1,1,3),

i.e., (1,1,1,1,3)=Projk[x0,x1,x2,x3,y]\mathbb{P}(1,1,1,1,3)=\mathrm{Proj}\,k[x_{0},x_{1},x_{2},x_{3},y] with degxi=1\deg x_{i}=1 and degy=3\deg y=3 for every 0i30\leq i\leq 3, and

X:=Projk[x0,x1,x2,x3,y](x06+x16+x26+x36+y2).X:=\mathrm{Proj}\,\frac{k[x_{0},x_{1},x_{2},x_{3},y]}{(x_{0}^{6}+x_{1}^{6}+x_{2}^{6}+x_{3}^{6}+y^{2})}.

Then XX is a smooth Fano threefold by the same proof as in Example 8.4. Moreover, XX is not quasi-FF-split by [KTY, Corollary 4.19(i), Proposition A.8].

References