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

Fano threefolds in positive characteristic I

Hiromu Tanaka Graduate School of Mathematical Sciences, The University of Tokyo, 3-8-1 Komaba, Meguro-ku, Tokyo 153-8914, JAPAN [email protected]
Abstract.

Over an algebraically closed field of positive characteristic, we classify smooth Fano threefolds of Picard number one whose anti-canonical linear systems are not very ample. Furthermore, we also prove that an anti-canonically embedded Fano threefold of genus at least five is an intersection of quadrics.

Key words and phrases:
Fano threefolds, positive characteristic.
2020 Mathematics Subject Classification:
14J45, 14J30, 14G17

1. Introduction

One of important classes of algebraic varieties is the class of Fano varieties, that is, smooth projective varieties whose anti-canonical divisors are ample. The classification of Fano varieties has been an interesting topic in algebraic geometry. The projective line 1\mathbb{P}^{1} is the unique one-dimensional Fano variety. Two-dimensional Fano varieties are called del Pezzo surfaces. It is classically known that a del Pezzo surface is isomorphic to either 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} or the blowup of 2\mathbb{P}^{2} along at most general eight points. Study of three-dimensional Fano varieties, called Fano threefolds, was initiated by Fano himself. The classification of Fano threefolds was completed by Iskovskih, Shokurov, and Mori–Mukai [Isk77], [Isk78], [Sho79a], [Sho79b], [MM81], [MM83], [MM03] (cf. [IP99], [Tak89]).

It is natural to consider the classification of Fano threefolds also in positive characteristic. In this direction, Megyesi, Shepherd-Barron, and Saito studied the case when rX=2r_{X}=2 (of index two), ρ(X)=1\rho(X)=1, and ρ(X)=2\rho(X)=2, respectively [Meg98], [SB97], [Sai03]. The purpose of this series of papers is to complete the classification of Fano threefolds in positive characteristic. This article focus on the case when ρ(X)=1\rho(X)=1 and |KX||-K_{X}| is not very ample.

Theorem 1.1 (Theorem 5.3, Section 6).

Let kk be an algberaically closed field of characteristic p>0p>0 and let XX be a Fano threefold over kk, i.e., XX is a three-dimensional smooth projective variety over kk such that KX-K_{X} is ample. Let rXr_{X} be the index of XX (for its definition, see Definition 2.1). Assume that ρ(X)=1\rho(X)=1 and |KX||-K_{X}| is not very ample. Then |KX||-K_{X}| is base point free and the morphism

φ|KX|:Xh0(X,KX)1\varphi_{|-K_{X}|}:X\to\mathbb{P}^{h^{0}(X,-K_{X})-1}

induced by |KX||-K_{X}| is a double cover onto its image Y:=φ|KX|(X)Y:=\varphi_{|-K_{X}|}(X), i.e., the induced morphism XYX\to Y is a finite surjective morphism such that the induced field extension K(X)K(Y)K(X)\supset K(Y) is of degree two. Furthermore, one and only one of the following holds.

  1. (1)

    rX=1,(KX)3=2r_{X}=1,(-K_{X})^{3}=2, and XX is a double cover of 3\mathbb{P}^{3}.

  2. (2)

    rX=1,(KX)3=4r_{X}=1,(-K_{X})^{3}=4, and XX is a double cover of a smooth quadric hypersurface in 4\mathbb{P}^{4}.

  3. (3)

    rX=2,(KX)3=8,r_{X}=2,(-K_{X})^{3}=8, and XX is isomorphic to a weighted hypersurface of (1,1,1,2,3)\mathbb{P}(1,1,1,2,3) of degree 66.

For the case when |KX||-K_{X}| is very ample, we shall establish the following theorem.

Theorem 1.2 (Theorem 7.13).

Let kk be an algberaically closed field of characteristic p>0p>0 and let XX be a Fano threefold over kk such that |KX||-K_{X}| is very ample. Set g:=12(KX)3+1g:=\frac{1}{2}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}(-K_{X})^{3}}+1. If ρ(X)=1\rho(X)=1 and g5g\geq 5, then the image φ|KX|(X)g+1\varphi_{|-K_{X}|}(X)\subset\mathbb{P}^{g+1} is an intersection of quadrics.

Based on Theorem 1.2, we shall classify the case when |KX||-K_{X}| is very ample and ρ(X)=1\rho(X)=1 in the second part [Tan-Fano2]. Theorem 1.1 and Theorem 1.2 are claimed in [SB97]. However, [SB97] seems to contain several logical gaps, whilst [SB97] introduced many ingenious techniques. Some proofs of this paper are base on the ideas introduced in [SB97].

1.1. Overview of the proofs and contents

Given a Fano threefold XX, it is easy to see that one of (I)–(III) holds (Lemma 6.1).

  1. (I)

    |KX||-K_{X}| is not base point free.

  2. (II)

    |KX||-K_{X}| is base point free and φ|KX|\varphi_{|-K_{X}|} is birational onto its image.

  3. (III)

    |KX||-K_{X}| is base point free and φ|KX|\varphi_{|-K_{X}|} is a double cover onto its image.

Following the strategy by Mori–Mukai in characteristic zero, it is important to classify primitive Fano threefolds, e.g. the case when ρ(X)=1\rho(X)=1. In this paper, we focus on the case when ρ(X)=1\rho(X)=1 and |KX||-K_{X}| is not very ample. The main part is to prove that (I) and (II) do not happen under these conditions. As for the case (III), we may apply almost the same argument as in characteristic zero.

1.1.1. Generic elephants

In characteristic zero, if XX is a Fano threefold, then it is known that a general element of |KX||-K_{X}|, called a general elephant, is smooth. It seems to be hard to conclude the same conclusion in positive characteristic because of the failure of the Bertini theorem for base point free divisors. On the other hand, we shall establish the following weaker result.

Theorem 1.3 (Corollary 4.5).

Let kk be an algebraically closed field of characteristic p>0p>0. Let XX be a Fano threefold over kk with ρ(X)=1\rho(X)=1. Then the generic member SS of |KX||-K_{X}| is regular and geometrically integral.

In particular, the generic member SS of |KX||-K_{X}| is a regular prime divisor on X×kκX\times_{k}\kappa for the purely transcendental field extension kκ:=K((H0(X,KX)))k\subset\kappa:=K(\mathbb{P}(H^{0}(X,-K_{X}))).

Remark 1.4.

The author does not know whether there exists a Fano threefold in positive characteristic such that any member of |KX||-K_{X}| is not smooth.

1.1.2. K3-like surfaces

Let XX be a Fano threefold over kk with ρ(X)=1\rho(X)=1. By Theorem 1.3, the generic member SS of |KX||-K_{X}| is regular. It is easy to see, by adjunction formula, that KS0K_{S}\sim 0 and H1(S,𝒪S)=0H^{1}(S,\mathcal{O}_{S})=0, so that such a surface SS is called a K3-like surface. Although the base field κ=K((H0(X,KX)))\kappa=K(\mathbb{P}(H^{0}(X,-K_{X}))) of SS is no longer perfect, we shall see that geometrically integral K3-like surfaces enjoy similar properties to those of K3 surfaces, e.g. the following hold:

  • H1(S,L)=0H^{1}(S,-L)=0 for a nef and big Cartier divisor LL on SS (Theorem 3.4).

  • If LL is a Cartier divisor on SS, then the mobile part of |L||L| is base point free (Corollary 3.12).

Combining with some arguments from [Isk77], we shall settle the case when |KX||-K_{X}| is not base point free, i.e., if XX is a Fano threefold such that |KX||-K_{X}| is not base point free, then ρ(X)2\rho(X)\geq 2 (Theorem 5.3).

1.1.3. Intersection of quadrics

The proof of Theorem 1.2 is subtler than that of characteristic zero. As in characteristic zero, we first consider the intersection WW of all the quadrics containing XX, which can be shown to be a 44-dimensional variety of minimal degree. In characteristic zero, we can easily deduce that WW is smooth. In positive characteristic, it seems to be hard to obtain the same conclusion, whilst we get a weaker conclusion: dimSingW1\dim{\operatorname{Sing}}\,W\leq 1. We then apply case study depending on dimSingW\dim{\operatorname{Sing}}\,W. The main new ingredients are the following two results.

  1. (i)

    Lefschetz hyperplane section theorem for generic members.

  2. (ii)

    Non-existence of smooth prime divisors on a cone over the Veronese surface.

To explain how to use (i) and (ii), let us focus on the case when dimSingW=1\dim{\operatorname{Sing}}\,W=1 and dim(XSingW)=0\dim(X\cap{\operatorname{Sing}}\,W)=0 (cf. Proposition 7.12). Let SS be the generic member of |KX||-K_{X}|. By (i), SKX×kκS\sim-K_{X\times_{k}\kappa} and SS is a regular projective surface with ρ(S)=1\rho(S)=1. Note that SS is disjoint from the blowup centre SingW×kκ{\operatorname{Sing}}\,W\times_{k}\kappa. By taking the blowup WWW^{\prime}\to W along SingW{\operatorname{Sing}}\,W, we obtain a surjective morphsim SZ×kκS\to Z\times_{k}\kappa, where WW is a cone over a variety ZZ of minimal degree. As dimSingW=1\dim{\operatorname{Sing}}\,W=1 implies dimZ=2\dim Z=2, we see that 1=ρ(S)ρ(Z×kκ)ρ(Z)1=\rho(S)\geq\rho(Z\times_{k}\kappa)\geq\rho(Z), and hence ρ(Z)=1\rho(Z)=1. It is easy to see that such a case only occurs when ZZ is the Veronese surface. However, this contradicts (ii), because XX is a smooth prime divisor on a cone WW over the Veronese surface ZZ.

Acknowledgements: The author would like to thank Tatsuro Kawakami for constructive comments and answering questions. The author would like to thank Natsuo Saito for kindly sharing his private note on [SB97]. The author also thanks the referee for reading the manuscript carefully and for suggesting several improvements. The author was funded by JSPS KAKENHI Grant numbers JP22H01112 and JP23K03028.

2. Preliminaries

2.1. Notation

In this subsection, we summarise notation used in this paper.

  1. (1)

    We will freely use the notation and terminology in [Har77]. In particular, D1D2D_{1}\sim D_{2} means linear equivalence of Weil divisors.

  2. (2)

    Throughout this paper, we work over an algebraically closed field kk of characteristic p>0p>0 unless otherwise specified.

  3. (3)

    For an integral scheme XX, we define the function field K(X)K(X) of XX as the local ring 𝒪X,ξ\mathcal{O}_{X,\xi} at the generic point ξ\xi of XX. For an integral domain AA, K(A)K(A) denotes the function field of SpecA{\operatorname{Spec}}\,A.

  4. (4)

    For a scheme XX, its reduced structure XredX_{{\operatorname{red}}} is the reduced closed subscheme of XX such that the induced closed immersion XredXX_{{\operatorname{red}}}\to X is surjective.

  5. (5)

    For a field κ\kappa, we say that XX is a variety over 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. a surface, resp. a threefold) over κ\kappa if XX is a variety over κ\kappa of dimension one (resp. two, resp. three).

  6. (6)

    A variety XX over a field is regular (resp. normal) if the local ring 𝒪X,x\mathcal{O}_{X,x} at any point xXx\in X is regular (resp. an integrally closed domain).

  7. (7)

    For a normal variety XX over a field κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}, we define the canonical divisor KXK_{X} as a Weil divisor on XX such that 𝒪X(KX)ωX/κ\mathcal{O}_{X}(K_{X})\simeq\omega_{X/{\kappa}}, where ωX/κ\omega_{X/{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}} denotes the dualising sheaf (cf. [Tan18, Section 2.3]). Canonical divisors are unique up to linear equivalence. Note that ωX/κωX/κ\omega_{X/{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}}\simeq\omega_{X/{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa^{\prime}}} for any field extension κκ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}\subset{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa^{\prime}} that induces a factorisation XSpecκSpecκX\to{\operatorname{Spec}}\,{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa^{\prime}}\to{\operatorname{Spec}}\,{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa} ([Tan18, Lemma 2.7]).

  8. (8)

    Let XX be a variety over a field κ\kappa and let LL be an invertible sheaf on XX. Given a finite-dimensional nonzero κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}-vector subspace VH0(X,L)V\subset H^{0}(X,L), |V||V| denotes the associated linear system and φ|V|:XdimV1\varphi_{|V|}:X\dashrightarrow\mathbb{P}^{\dim V-1} denotes the induced rational map. In particular, we have |H0(X,L)|=|L||H^{0}(X,L)|=|L|. Note that the base locus or the base scheme, denoted by Bs|V|{\operatorname{Bs}}|V|, is the closed subscheme whose ideal sheaf IBs|V|I_{{\operatorname{Bs}}|V|} satisfies the following equation:

    Im(Vk𝒪XH0(X,L)k𝒪XL)=IBs|V|L.{{\operatorname{Im}}}(V\otimes_{k}\mathcal{O}_{X}\hookrightarrow H^{0}(X,L)\otimes_{k}\mathcal{O}_{X}\to L)=I_{{\operatorname{Bs}}|V|}\cdot L.

    We have the induced morphism:

    φ|V||XBs|V|:XBs|V|dimV1.\varphi_{|V|}|_{X\setminus{\operatorname{Bs}}\,|V|}:X\setminus{\operatorname{Bs}}\,|V|\to\mathbb{P}^{\dim V-1}.

    Note that V0V\neq 0 implies Bs|V|X{\operatorname{Bs}}|V|\neq X. Set

    Imφ|V|:=Im(φ|V||XBs|V|)¯,{\operatorname{Im}}\,\varphi_{|V|}:=\overline{{\operatorname{Im}}(\varphi_{|V|}|_{X\setminus{\operatorname{Bs}}\,|V|})},

    where the right hand side is the closure of Im(φ|V||XBs|V|){\operatorname{Im}}(\varphi_{|V|}|_{X\setminus{\operatorname{Bs}}\,|V|}) in dimV1\mathbb{P}^{\dim V-1}.

  9. (9)

    Given a projective variety XX and an invertible sheaf LL, we set R(X,L):=m=0H0(X,Lm)R(X,L):=\bigoplus_{m=0}^{\infty}H^{0}(X,L^{\otimes m}), which is a graded kk-algebra. For a Cartier divisor DD on XX, we set R(X,D):=m=0H0(X,𝒪X(D)m)=m=0H0(X,𝒪X(mD))R(X,D):=\bigoplus_{m=0}^{\infty}H^{0}(X,\mathcal{O}_{X}(D)^{\otimes m})=\bigoplus_{m=0}^{\infty}H^{0}(X,\mathcal{O}_{X}(mD)).

  10. (10)

    Let XX be a projective normal variety over a field. We say that a Cartier divisor DD is nef if DC0D\cdot C\geq 0 for any curve CC on XX. Given Cartier divisors D1D_{1} and D2D_{2}, the numerical equivalence D1D2D_{1}\equiv D_{2} means that D1C=D2CD_{1}\cdot C=D_{2}\cdot C for any curve CC on XX. For a nef Cartier divisor DD, its numerical dimension ν(X,D)\nu(X,D) is defined as the maximum ν{0,1,,dimX}\nu\in\{0,1,...,\dim X\} such that DνHdimXν0D^{\nu}\cdot H^{\dim X-\nu}\neq 0 for every (equivalently, for one) ample Cartier divisor HH.

  11. (11)

    Let XnX\subset\mathbb{P}^{n} be a projective variety which is a closed subscheme of n\mathbb{P}^{n}. For s>0s\in\mathbb{Z}_{>0}, the ss-th cone Yn+sY\subset\mathbb{P}^{n+s} (of XX) is defined by the same defining ideal of XX. Specifically, if

    X=Projk[x0,,xn](f1,,fr)Projk[x0,,xn]=nX={\operatorname{Proj}}\,\frac{k[x_{0},...,x_{n}]}{(f_{1},...,f_{r})}\subset{\operatorname{Proj}}\,k[x_{0},...,x_{n}]=\mathbb{P}^{n}

    for homogeneous polynomials f1,,frk[x0,,xn]f_{1},...,f_{r}\in k[x_{0},...,x_{n}], then

    Y=Projk[x0,,xn,y1,,ys](f1,,fr)Projk[x0,,xn,y1,,ys]=n+s.Y={\operatorname{Proj}}\,\frac{k[x_{0},...,x_{n},y_{1},...,y_{s}]}{(f_{1},...,f_{r})}\subset{\operatorname{Proj}}\,k[x_{0},...,x_{n},y_{1},...,y_{s}]=\mathbb{P}^{n+s}.

    A cone of XX is the ss-th cone for some s>0s\in\mathbb{Z}_{>0}.

  12. (12)

    Given a projective variety XX over a field κ\kappa, ρ(X)\rho(X) denotes its Picard number, i.e., ρ(X):=dim((PicX/))\rho(X):=\dim_{\mathbb{Q}}(({\operatorname{Pic}}\,X/\equiv)\otimes_{\mathbb{Z}}\mathbb{Q}), where \equiv is the numerical equivalence.

2.1.1. Fano threefolds

Definition 2.1.

We say that XX is a Fano threefold if XX is a three-dimensional smooth projective variety over kk such that KX-K_{X} is ample. The index rXr_{X} of XX is defined as the largest positive integer rr such that there exists a Cartier divisor HH on XX satisfying KXrH-K_{X}\sim rH.

Remark 2.2.

Let XX be a Fano threefold. We shall frequently use the following notations.

  1. (1)

    Set g:=(KX)32+1g:=\frac{(-K_{X})^{3}}{2}+1, which is called the genus of XX.

  2. (2)

    Set g:=h0(X,KX)2g^{\prime}:=h^{0}(X,-K_{X})-2.

If KX-K_{X} is very ample, then the genus of XX coincides with the genus of the smooth curve HHH\cap H^{\prime} for general members H,H|KX|H,H^{\prime}\in|-K_{X}|. We shall prove the following (Corollary 2.6):

g=g+h1(X,KX)g.g^{\prime}=g+h^{1}(X,-K_{X})\geq g.

2.1.2. Canonical surfaces

For a field κ\kappa and a normal surface SS over κ\kappa, we say that SS is canonical if (S,0)(S,0) is canonical in the sense of [Kol13, Definition 2.8]. For the minimal resolution μ:TS\mu:T\to S of SS, it is well known that the following are equivalent [Kol13, Theorem 2.29].

  1. (i)

    SS is canonical.

  2. (ii)

    KSK_{S} is Cartier and KTμKSK_{T}\sim\mu^{*}K_{S}.

In this paper, we only need the characterisation of canonical surfaces by (ii). Hence (ii) may be considered as the definition of canonical surfaces.

2.2. Cohomologies of Fano threefolds

Throughout this subsection, we work over an algebraically closed field kk of characteristic p>0p>0.

Lemma 2.3.

Let XX be a Fano threefold. Then there exist a sequence

X=:X0φ0X1φ1φ1XX=:X_{0}\xrightarrow{\varphi_{0}}X_{1}\xrightarrow{\varphi_{1}}\cdots\xrightarrow{\varphi_{\ell-1}}X_{\ell}

such that the following properties hold.

  1. (1)

    For every i{0,1,,}i\in\{0,1,...,\ell\}, XiX_{i} is a Fano threefold.

  2. (2)

    For every i{0,1,,1}i\in\{0,1,...,\ell-1\}, φi:XiXi+1\varphi_{i}:X_{i}\to X_{i+1} is a blowup of Xi+1X_{i+1} along a smooth curve on Xi+1X_{i+1}.

  3. (3)

    Either

    1. (a)

      ρ(X)=1\rho(X_{\ell})=1 or

    2. (b)

      there exists a contraction f:XYf:X_{\ell}\to Y of an extremal ray such that YY is a smooth projective rational surface and ff is generically smooth.

This lemma is a weaker version of [Kaw21, Lemma 3.2]. Since [Kaw21, Lemma 3.2] uses a reference [Sai03] which contains a logical gap (Remark 2.5(1)), we give a proof for the sake of completeness.

Proof.

We may assume that XX is primitive, i.e., there exists no blowup XXX\to X^{\prime} of a Fano threefold XX^{\prime} along a smooth curve. Assume ρ(X)2\rho(X)\geq 2. It is enough to show (b). By the same argument as in [MM83, Section 8, (8.1), (8.2)] (which is applicable by [Kol91, Main Theorem 1.1]), there exists a contraction f:XYf:X\to Y of an extremal ray such that YY is a smooth projective surface. If ff is not generically smooth, then XX has a 1\mathbb{P}^{1}-bundle structure f:XYf^{\prime}:X\to Y^{\prime} which is a contraction of another extremal ray [MS03, Corollary 8 and Remark 10]. Hence we may assume that f:XYf:X\to Y is generically smooth. By [Eji19, Corollary 4.10(2)], KY-K_{Y} is big, and hence YY is a smooth ruled surface. On the other hand, YY is rationally chain connected, because so is XX [Kol96, Ch. V, Theorem 2.13]. Therefore, YY is rational. ∎

Theorem 2.4.

Let XX be a Fano threefold. Let DD be a nef Cartier divisor on XX with ν(X,D)2\nu(X,D)\geq 2. Then the following hold.

  1. (1)

    Hj(X,D)=0H^{j}(X,-D)=0 for any j1j\leq 1.

  2. (2)

    Hi(X,KX+D)=0H^{i}(X,K_{X}+D)=0 for any i2i\geq 2.

  3. (3)

    Hi(X,𝒪X)=0H^{i}(X,\mathcal{O}_{X})=0 for any i>0i>0. In particular, χ(X,𝒪X)=1\chi(X,\mathcal{O}_{X})=1.

  4. (4)

    PicXρ(X){\operatorname{Pic}}\,X\simeq\mathbb{Z}^{\oplus\rho(X)}.

Proof.

The assertion follows from [Kaw21, Corollary 3.6 and Corollary 3.7]. Note that [Kaw21, Corollary 3.6 and Corollary 3.7] follows from [Kaw21, Theorem 3.5], which depends on [Kaw21, Lemma 3.2]. As mentioned above, [Kaw21, Lemma 3.2] relies on [Sai03], which contains a logical gap. Although Lemma 2.3 is weaker than [Kaw21, Lemma 3.2], Lemma 2.3 is enough to establish [Kaw21, Theorem 3.5].

Remark 2.5.

[Sai03, the proof of Lemma 2.4] claims that the same argument as in [MM83, Proposition 6.2] or [MM86, Corollary 4.6] works in positive characteristic. However, there actually exists an example for which the argument does not work [Tanb, Remark 7.6].

Given a smooth projective threefold XX and a Cartier divisor DD, the Riemann–Roch theorem is given as follows:

(2.5.1) χ(X,𝒪X(D))=112D(DKX)(2DKX)+112Dc2(X)+χ(X,𝒪X)\chi(X,\mathcal{O}_{X}(D))=\frac{1}{12}D\cdot(D-K_{X})\cdot(2D-K_{X})+\frac{1}{12}D\cdot c_{2}(X)+\chi(X,\mathcal{O}_{X})
(2.5.2) χ(X,𝒪X)=124(KX)c2(X).\chi(X,\mathcal{O}_{X})=\frac{1}{24}(-K_{X})\cdot c_{2}(X).
Corollary 2.6.

Let XX be a Fano threefold. Let HH be a Cartier divisor such that the numerical equivalence HqKXH\equiv-qK_{X} holds for some q0q\in\mathbb{Q}_{\geq 0} (i.e., HC=qKXCH\cdot C=-qK_{X}\cdot C holds for any curve CC on XX). Then the following holds

h0(X,H)h1(X,H)=χ(X,H)=112q(q+1)(2q+1)(KX)3+2q+1.h^{0}(X,H)-h^{1}(X,H)=\chi(X,H)=\frac{1}{12}q(q+1)(2q+1)(-K_{X})^{3}+2q+1.

In particular, if m0m\in\mathbb{Z}_{\geq 0}, then

h0(X,mKX)h1(X,mKX)=χ(X,mKX)=112m(m+1)(2m+1)(KX)3+2m+1,h^{0}(X,-mK_{X})-h^{1}(X,-mK_{X})=\chi(X,-mK_{X})=\frac{1}{12}m(m+1)(2m+1)(-K_{X})^{3}+2m+1,

and hence h0(X,KX)h1(X,KX)=(KX)32+3=g+2h^{0}(X,-K_{X})-h^{1}(X,-K_{X})=\frac{(-K_{X})^{3}}{2}+3=g+2.

Proof.

If q=0q=0, then there is nothing to show (Theorem 2.4). Assume q>0q>0. We then have ν(X,H)=ν(X,KX)=3\nu(X,H)=\nu(X,-K_{X})=3. By (2.5.2) and HqKXH\equiv-qK_{X}, we have

(2.6.1) Hc2(X)=q(KX)c2(X)=24qχ(X,𝒪X)=24q.H\cdot c_{2}(X)=q(-K_{X})\cdot c_{2}(X)=24q\chi(X,\mathcal{O}_{X})=24q.

Then the assertion holds by the following computation:

h0(X,H)h1(X,H)\displaystyle h^{0}(X,H)-h^{1}(X,H) =(1)χ(X,H)\displaystyle\overset{(1)}{=}\chi(X,H)
=(2)112H(HKX)(2HKX)+112Hc2(X)+χ(X,𝒪X)\displaystyle\overset{(2)}{=}\frac{1}{12}H\cdot(H-K_{X})\cdot(2H-K_{X})+\frac{1}{12}H\cdot c_{2}(X)+\chi(X,\mathcal{O}_{X})
=(3)112(qKX)(qKXKX)(2qKXKX)+2q+1\displaystyle\overset{(3)}{=}\frac{1}{12}(-qK_{X})\cdot(-qK_{X}-K_{X})\cdot(-2qK_{X}-K_{X})+2q+1
=112q(q+1)(2q+1)(KX)3+2q+1.\displaystyle=\frac{1}{12}q(q+1)(2q+1)(-K_{X})^{3}+2q+1.

Here (1) holds by Theorem 2.4, (2) follows from (2.5.1), and we obtain (3) by (2.6.1) and χ(X,𝒪X)=1\chi(X,\mathcal{O}_{X})=1 (Theorem 2.4). ∎

2.3. Bertini theorems

2.3.1. Generic members

Notation 2.7.

Let κ\kappa be a field. Let XX be a variety over κ\kappa, let LL be a Cartier divisor on XX, and let VH0(X,L)V\subset H^{0}(X,L) be a nonzero finite-dimensional κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}-vector subspace. We then have the universal family XL,VunivX^{{\operatorname{univ}}}_{L,V} that parameterises the effective Cartier divisors of |V||V| whose parameter space is (V)\mathbb{P}(V). Then its generic fibre XL,VgenX^{{\operatorname{gen}}}_{L,V} is called the generic member of |V||V|. To summarise, we have the following diagram in which all the squares are cartesian:

XL,VgenXL,VunivX×κK((V))X×κ(V)SpecK((V))(V).\begin{CD}X^{{\operatorname{gen}}}_{L,V}@>{}>{}>X^{{\operatorname{univ}}}_{L,V}\\ @V{}V{}V@V{}V{}V\\ X\times_{\kappa}K(\mathbb{P}(V))@>{}>{}>X\times_{\kappa}\mathbb{P}(V)\\ @V{}V{}V@V{}V{}V\\ {\operatorname{Spec}}\,K(\mathbb{P}(V))@>{}>{}>\mathbb{P}(V).\end{CD}

Note that if XX is normal, then we have XL,VgenLK((V))X^{{\operatorname{gen}}}_{L,V}\sim L_{K(\mathbb{P}(V))} for the pullback LK((V))L_{K(\mathbb{P}(V))} of LL to X×κK((V))X\times_{\kappa}K(\mathbb{P}(V)).

Remark 2.8.

We use Notation 2.7. Let XX^{\prime} be a non-empty open subset of XX and let i:XXi:X^{\prime}\hookrightarrow X be the induced open immersion. We then obtain

VH0(X,L)H0(X,L)forL:=L|X.V\subset H^{0}(X,L)\hookrightarrow H^{0}(X^{\prime},L^{\prime})\qquad\text{for}\qquad L^{\prime}:=L|_{X^{\prime}}.

Set VV^{\prime} to be the image of VV in H0(X,L)H^{0}(X^{\prime},L^{\prime}). We then obtain the following cartesian diagram [Tana, Proposition 5.10]:

XL,VgenXL,VgenXiX.\begin{CD}X^{\prime{\operatorname{gen}}}_{L^{\prime},V^{\prime}}@>{}>{}>X^{{\operatorname{gen}}}_{L,V}\\ @V{}V{}V@V{}V{}V\\ X^{\prime}@>{i}>{}>X.\end{CD}
Theorem 2.9.

Let κ\kappa be a field. Let XX be a variety over κ\kappa, let LL be a Cartier divisor on XX, and let VH0(X,L)V\subset H^{0}(X,L) be a nonzero finite-dimensional κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}-vector subspace. Set X:=XBs|V|X^{\prime}:=X\setminus{\operatorname{Bs}}\,|V| and let α:XL,VgenX\alpha:{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}X^{{\operatorname{gen}}}_{L,V}}\to X be the induced morphism. If XX^{\prime} is regular, then α1(X)\alpha^{-1}(X^{\prime}) is regular.

Proof.

After replacing XX by XX^{\prime}, we may assume that |V||V| is base point free (Remark 2.8). Then the assertion follows from [Tana, Theorem 4.9 and Remark 5.8]. ∎

2.3.2. General members

Proposition 2.10.

We work over an algebraically closed field kk. Let f:XNf:X\to\mathbb{P}^{N} be a morphism from a variety XX. If dimf(X)¯2\dim\overline{f(X)}\geq 2, then f1(H)f^{-1}(H) is irreducible for a general hyperplane HNH\subset\mathbb{P}^{N}.

Proof.

See [Jou83, 4) of Theoreme 6.3]. ∎

The following proposition is a known result [Fuj90, Theorem 2.7]. However, we here give a proof for the sake of compleness, because [Fuj90, Theorem 2.7] depends on [Wei62] which is written in a classical language of algebraic geometry.

Proposition 2.11.

We work over an algebraically closed field kk. Let XX be a projective normal variety and let LL be a Cartier divisor on XX such that dimBs|L|dimX2\dim{\operatorname{Bs}}\,|L|\leq\dim X-2 and dim(Imφ|L|)2\dim({\operatorname{Im}}\,\varphi_{|L|})\geq 2. Then general members of |L||L| are prime divisors. In particular, the generic member of |L||L| is geometrically integral.

Proof.

We first reduce the problem to the case when |L||L| is base point free. Let

μ:XX\mu:X^{\prime}\to X

be the normalisation of the resolution of the indeterminacies of φ|L|:Xh0(X,L)1\varphi_{|L|}:X\dashrightarrow\mathbb{P}^{h^{0}(X,L)-1}. We then have

μL=M+F,\mu^{*}L=M+F,

where |M||M| is base point free and FF is the fixed part of |μL||\mu^{*}L|. Then we obtain the induced morphisms:

φ|M|:X𝜇Xφ|L|h0(X,L)1.\varphi_{|M|}:X^{\prime}\xrightarrow{\mu}X\overset{\varphi_{|L|}}{\dashrightarrow}\mathbb{P}^{h^{0}(X,L)-1}.

By construction, we have

H0(X,M)H0(X,M+F)=H0(X,μL)H0(X,L).H^{0}(X^{\prime},M)\xrightarrow{\simeq}H^{0}(X^{\prime},M+F)=H^{0}(X^{\prime},\mu^{*}L)\simeq H^{0}(X,L).

Furthermore, this composite isomorphism induces the bijection: |M||L|,DμD|M|\to|L|,D\mapsto\mu_{*}D. Therefore, if a general member DD of |M||M| is a prime divisor, then also its push-forward μD\mu_{*}D is a prime divisor. Hence we may assume that |D||D| is base point free after replacing (X,L)(X,L) by (X,M)(X^{\prime},M).

We have the following induced morphisms:

φ|L|:X𝜑YN,Y:=φ|L|(X),N:=h0(X,L)1.\varphi_{|L|}:X\xrightarrow{\varphi}Y\hookrightarrow\mathbb{P}^{N},\qquad Y:=\varphi_{|L|}(X),\qquad N:=h^{0}(X,L)-1.

By construction, it holds that

H0(X,L)H0(Y,𝒪N(1)|Y)H0(N,𝒪N(1)).H^{0}(X,L)\simeq H^{0}(Y,\mathcal{O}_{\mathbb{P}^{N}}(1)|_{Y})\simeq H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(1)).

In particular, a general member HY|𝒪N(1)|Y|H_{Y}\in|\mathcal{O}_{\mathbb{P}^{N}}(1)|_{Y}| (a general hyperplane section) is an integral scheme.

For a suitable non-empty open subset YYY^{\circ}\subset Y, which is normal, and its inverse image X:=φ1(Y)XX^{\circ}:=\varphi^{-1}(Y^{\circ})\subset X, the Noether normalisation theorem, applied for the generic fibre X×YSpecK(Y)X\times_{Y}{\operatorname{Spec}}\,K(Y), induces following factorisation:

X𝛼Z:=Y×rpr1Y,X^{\circ}\xrightarrow{\alpha}Z:=Y^{\circ}\times\mathbb{P}^{r}\xrightarrow{{\rm pr}_{1}}Y^{\circ},

where α:XZ=Y×r\alpha:X^{\circ}\to Z=Y^{\circ}\times\mathbb{P}^{r} is a finite surjective morphism of normal varieties and the latter morphism pr1:Y×rY{\rm pr}_{1}:Y^{\circ}\times\mathbb{P}^{r}\to Y^{\circ} is the projection (cf. [Tan20, Lemma 2.14]). For a general hyperplane section HYYH_{Y}\subset Y, its inverse image HY|ZH_{Y}|_{Z} to ZZ is an integral scheme, because it can be written as (HYY)×r(H_{Y}\cap Y^{\circ})\times\mathbb{P}^{r}. Furthermore, we take the decomposition via the separable closure:

X𝛽W𝛾Z=Y×rpr1Y,X^{\circ}\xrightarrow{\beta}W\xrightarrow{\gamma}Z=Y^{\circ}\times\mathbb{P}^{r}\xrightarrow{{\rm pr}_{1}}Y^{\circ},

where

  • WW is a normal variety,

  • β:XW\beta:X^{\circ}\to W is a finite surjective purely inseaprable morphism, and

  • γ:WZ\gamma:W\to Z is a finite surjective separable morphism.

Then also the pullback HY|WH_{Y}|_{W} of HYH_{Y} to WW is integral (i.e., a prime divisor). Indeed, HY|XH_{Y}|_{X} is irreducible (Proposition 2.10), so it suffices to prove that HY|WH_{Y}|_{W} is reduced. It is S1S_{1}, and hence it suffices to prove that HY|WH_{Y}|_{W} is R0R_{0}. This holds because we may assume that HY|ZH_{Y}|_{Z} intersects with the étale locus of γ:WZ\gamma:W\to Z.

Since β:XW{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\beta:}X^{\circ}\to W is a finite purely inseparable morphism, we have the following factorisation for some e>0e\in\mathbb{Z}_{>0}:

Fe:WX𝛽W,F^{e}:W\to X^{\circ}\xrightarrow{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\beta}}W,

where Fe:WWF^{e}:W\to W denotes the ee-th iterated absolute Frobenius morphism. By (Fe)(HY|W)=peHY|W(F^{e})^{*}(H_{Y}|_{W})=p^{e}H_{Y}|_{W}, it holds that HY|X=pd(HY|X)redH_{Y}|_{X^{\circ}}=p^{d}(H_{Y}|_{X^{\circ}})_{{\operatorname{red}}} for some 0de0\leq d\leq e. A general member of |L||L| is nothing but the pullback HY|XH_{Y}|_{X} of a general member HYH_{Y} of |𝒪N(1)|Y||\mathcal{O}_{\mathbb{P}^{N}}(1)|_{Y}|. In particular,

(HY|X)|X=HY|X=pd(HY|X)red.(H_{Y}|_{X})|_{X^{\circ}}=H_{Y}|_{X^{\circ}}=p^{d}(H_{Y}|_{X^{\circ}})_{{\operatorname{red}}}.

Suppose that a general member of |L||L| is not ingeral. Since HY|XH_{Y}|_{X} is irreducible (Proposition 2.10), we have HY|X=pDH_{Y}|_{X}=pD for some effective Weil divisor DD.

Therefore, a general member of |L||L| is of the form pDpD. Replacing LL by pDpD, we may assume that L=pLL=pL^{\prime} for some Weil divisor LL^{\prime}. This implies that the image of

F:H0(X,L)H0(X,L),sspF:H^{0}(X,L^{\prime})\to H^{0}(X,L),\qquad s\mapsto s^{p}

is dense with respect to the Zariski topology of the affine space H0(X,L)H^{0}(X,L). As the image Im(F){\operatorname{Im}}(F) is a closed subset in H0(X,L)H^{0}(X,L), FF is surjective. Therefore, any member of |L||L| can be written as pDpD for some Weil divisor DD.

We have an isomorphism:

F:H0(X,L)H0(X,L),ttp.F:H^{0}(X,L^{\prime})\xrightarrow{\simeq}H^{0}(X,L),\qquad t\mapsto t^{p}.

Fix a kk-linear basis of H0(X,L)H^{0}(X,L^{\prime}):

t0,,tNH0(X,L)t_{0},...,t_{N}\in H^{0}(X,L^{\prime})

and their images to H0(X,L)H^{0}(X,L):

t0p,,tNpH0(X,L),t_{0}^{p},...,t_{N}^{p}\in H^{0}(X,L),

which form a kk-linear basis of H0(X,L)H^{0}(X,L). We then obtain

φ|L|:XN,x[(t0(x))p::(tN(x))p].\varphi_{|L|}:X\to\mathbb{P}^{N},\qquad x\mapsto[(t_{0}(x))^{p}:\cdots:(t_{N}(x))^{p}].

Therefore, we get the following factorisation:

φ|L|:Xφ|L|N𝐹N,x[t0(x)::tN(x)][(t0(x))p::(tN(x))p].\varphi_{|L|}:X\xrightarrow{\varphi_{|L^{\prime}|}}\mathbb{P}^{N}\xrightarrow{F}\mathbb{P}^{N},\qquad x\mapsto[t_{0}(x):\cdots:t_{N}(x)]\mapsto[(t_{0}(x))^{p}:\cdots:(t_{N}(x))^{p}].

By Y=φ|L|(X)Y=\varphi_{|L|}(X), we obtain the factorisation

XY𝐹Y,X\to Y\xrightarrow{F}Y,

which induces the following isomorphisms

H0(Y,𝒪N(1)|Y)H0(Y,𝒪N(p)|Y)H0(X,L),H^{0}(Y,\mathcal{O}_{\mathbb{P}^{N}}(1)|_{Y})\xrightarrow{\simeq}H^{0}(Y,{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\mathcal{O}_{\mathbb{P}^{N}}(p)|_{Y}})\xrightarrow{\simeq}H^{0}(X,L),

because the composition is an isomorphism and each map is injective. This implies that any member of |𝒪N(p)|Y||\mathcal{O}_{\mathbb{P}^{N}}(p)|_{Y}| is non-reduced. However, this is a contradiction, because HY|𝒪N(p)|Y|H\cap Y\in|\mathcal{O}_{\mathbb{P}^{N}}(p)|_{Y}| is an integral scheme for a general member H|𝒪N(p)|H\in|\mathcal{O}_{\mathbb{P}^{N}}(p)|. ∎

Lemma 2.12.

We work over an algebraically closed field kk. Let XX be a projective CM variety with dimX2\dim X\geq 2 and let

ψ:XY\psi:X\to Y

be a birational morphism to a projective variety YY. Fix a closed embedding YNY\subset\mathbb{P}^{N}. If HH is a general hyperplane of N\mathbb{P}^{N}, then

ψ1(YH)YH\psi^{-1}(Y\cap H)\to Y\cap H

is a birational morphism from a projective CM variety ψ1(YH)\psi^{-1}(Y\cap H) to a projective variety YHY\cap H. In particular, if H1,,HrNH_{1},...,H_{r}\subset\mathbb{P}^{N} are general hyperplanes with 1rdimX11\leq r\leq\dim X-1, then

ψ1(YH1Hr)YH1Hr\psi^{-1}(Y\cap H_{1}\cap\cdots\cap H_{r})\to Y\cap H_{1}\cap\cdots\cap H_{r}

is a birational morphism from a projective CM variety ψ1(YH1Hr)\psi^{-1}(Y\cap H_{1}\cap\cdots\cap H_{r}) to a projective variety YH1HrY\cap H_{1}\cap\cdots\cap H_{r}.

Proof.

Fix a general hyperplane HH. By the classical Bertini theorem, YHY\cap H is an integral scheme. It follows from Proposition 2.10 that ψ1(YH)\psi^{-1}(Y\cap H) is irreducible. Since ψ1(YH)\psi^{-1}(Y\cap H) is an effective Cartier divisor on XX, also ψ1(YH)\psi^{-1}(Y\cap H) is CM. Since ψ1(YH)YH\psi^{-1}(Y\cap H)\to Y\cap H is isomorphic over the generic point of YHY\cap H, ψ1(YH)\psi^{-1}(Y\cap H) is generically reduced. Then ψ1(YH)\psi^{-1}(Y\cap H) is reduced, because ψ1(YH)\psi^{-1}(Y\cap H) is R0R_{0} and S1S_{1}. Therefore, ψ1(YH)\psi^{-1}(Y\cap H) is an integral scheme. ∎

2.4. Δ\Delta-genera

Throughout this subsection, we work over an algebraically closed field kk of characteristic p>0p>0.

Definition 2.13.

We say that (X,L)(X,L) is a polarised variety if XX is a projective variety and LL is an ample invertible sheaf or an ample Cartier divisor. Set

Δ(X,L):=dimX+LdimXh0(X,L).\Delta(X,L):=\dim X+L^{\dim X}-h^{0}(X,L).

We say that polarised varieties (X,L)(X,L) and (X,L)(X^{\prime},L^{\prime}) are isomorphic, denoted by (X,L)(X,L)(X,L)\simeq(X^{\prime},L^{\prime}), if there exists a kk-isomorphism θ:XX\theta:X\xrightarrow{\simeq}X^{\prime} such that LθLL\simeq\theta^{*}L^{\prime}. A polarised variety (X,L)(X,L) is called smooth if XX is smooth.

Theorem 2.14.

Let (X,L)(X,L) be a polarised variety. Then the following hold.

  1. (1)

    Δ(X,L)0\Delta(X,L)\geq 0, i.e., h0(X,L)dimX+LdimXh^{0}(X,L)\leq\dim X+L^{\dim X}.

  2. (2)

    If Δ(X,L)=0\Delta(X,L)=0, then |L||L| is very ample.

Proof.

See [Fuj82b, Theorem (2.1) and Theorem (4.2)]. ∎

Theorem 2.15.

Let (X,L)(X,L) be a smooth polarised variety such that Δ(X,L)=0\Delta(X,L)=0. Set n:=dimXn:=\dim X. Then one of the following holds.

  1. (1)

    Ln=1L^{n}=1 and (X,L)(n,𝒪n(1))(X,L)\simeq(\mathbb{P}^{n},\mathcal{O}_{\mathbb{P}^{n}}(1)).

  2. (2)

    Ln=2L^{n}=2 and (X,L)(Qn,𝒪n+1(1)|Qn)(X,L)\simeq(Q^{n},\mathcal{O}_{\mathbb{P}^{n+1}}(1)|_{Q^{n}}), where Qnn+1Q^{n}\subset\mathbb{P}^{n+1} is a smooth quadric hypersurface.

  3. (3)

    Ln3L^{n}\geq 3 and X1(E)X\simeq\mathbb{P}_{\mathbb{P}^{1}}(E), where EE is a locally free sheaf on 1\mathbb{P}^{1} of rank nn.

  4. (4)

    (X,L)(2,𝒪2(2))(X,L)\simeq(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(2)) and L2=4L^{2}=4.

Proof.

See [Fuj82b, Corollary (4.3) and Theorem (4.9)]. ∎

Remark 2.16.

Let (X,L)(X,L) be a polarised variety such that XX is not smooth and Δ(X,L)=0\Delta(X,L)=0. Then |L||L| is very ample (Theorem 2.14(2)). Set N:=h0(X,L)1N:=h^{0}(X,L)-1 and we fix a closed embedding XNX\subset\mathbb{P}^{N} induced by |L||L|. Then Δ(X,L)=0\Delta(X,L)=0 implies degX=1+codimX\deg X=1+{\rm codim}\,X. Such a variety is called a variety of minimal degree, since the inequality degX1+codimX\deg X\geq 1+{\rm codim}\,X holds in general. By [Fuj82b, Theorem (4.11)] or [EH87, Theorem 1], it is known that (X,L)(X,L) is a cone over a smooth polarised variety (Z,LZ)(Z,L_{Z}) with Δ(Z,LZ)=0\Delta(Z,L_{Z})=0. More specifically, it holds by [EH87, page 4] that

  • Z=Projk[x0,,xNr]/(f1,,fs)NrZ={\operatorname{Proj}}\,k[x_{0},...,x_{N-r}]/(f_{1},...,f_{s})\subset\mathbb{P}^{N-r} for r:=dimXdimZr:=\dim X-\dim Z and some f1,,fsk[x0,,xNr]f_{1},...,f_{s}\in k[x_{0},...,x_{N-r}],

  • X=Projk[x0,,xN]/(f1,,fs)NX={\operatorname{Proj}}\,k[x_{0},...,x_{N}]/(f_{1},...,f_{s})\subset\mathbb{P}^{N}, i.e., the same equations as those of ZZ define XNX\subset\mathbb{P}^{N}.

Remark 2.17.

Let κ\kappa be a (not necessarily algebraically closed) field. For a geometrically integral projective variety XX over κ\kappa and an ample divisor LL, we set

Δ(X,L):=dimX+LdimXh0(X,L).\Delta(X,L):=\dim X+L^{\dim X}-h^{0}(X,L).

By definition, we have

Δ(X,L)=Δ(X×κκ¯,αL)0,\Delta(X,L)=\Delta(X\times_{\kappa}\overline{\kappa},\alpha^{*}L)\geq 0,

where κ¯\overline{\kappa} denotes the algebraic closure of κ\kappa and α:X×κκ¯X\alpha:X\times_{\kappa}\overline{\kappa}\to X is the induced morphism.

2.5. The case of index 2\geq 2

Throughout this subsection, we work over an algebraically closed field kk of characteristic p>0p>0.

Theorem 2.18.

Let XX be a Fano threefold. Then the following hold.

  1. (1)

    1rX41\leq r_{X}\leq 4.

  2. (2)

    rX=4r_{X}=4 if and only if X3X\simeq\mathbb{P}^{3}.

  3. (3)

    rX=3r_{X}=3 if and only if XX is isomorphic to a smooth quadric hypersurface in 4\mathbb{P}^{4}.

Proof.

See [Meg98, Proposition 4]. ∎

Proposition 2.19.

Let (X,L)(X,L) be a polarised variety. Fix a>0a\in\mathbb{Z}_{>0} and δH0(X,aL)\delta\in H^{0}(X,aL). Let DD be the member of |aL||aL| corresponding to δ\delta. Let

ρ:m=0H0(X,mL)m=0H0(D,m(L|D))\rho:\bigoplus_{m=0}^{\infty}H^{0}(X,mL)\to\bigoplus_{m=0}^{\infty}H^{0}(D,m(L|_{D}))

be the restriction map. Let ξ1,,ξr\xi_{1},...,\xi_{r} be homogeneous elements of m=0H0(X,mL)\bigoplus_{m=0}^{\infty}H^{0}(X,mL). If m=0H0(D,m(L|D))\bigoplus_{m=0}^{\infty}H^{0}(D,m(L|_{D})) is generated by ρ(ξ1),,ρ(ξr)\rho(\xi_{1}),...,\rho(\xi_{r}) as a kk-algebra, then m=0H0(X,mL)\bigoplus_{m=0}^{\infty}H^{0}(X,mL) is generated by δ,ξ1,,ξr\delta,\xi_{1},...,\xi_{r}.

Proof.

See [Fuj90, Theorem (2.3)]. ∎

Definition 2.20 ((1.5) and (1.6) of [Fuj82b]).

Let (X,L)(X,L) be a polarised variety and set n:=dimXn:=\dim X.

  1. (1)

    We say that X=:XnXn1X1X=:X_{n}\supset X_{n-1}\supset\cdots\supset X_{1} is a ladder (of (X,L)(X,L)) if, for every i{n,n1,,2}i\in\{n,n-1,...,2\}, Xi1X_{i-1} is a member of |L|Xi||L|_{X_{i}}| which is an integral scheme. By definition, we have dimXi=i\dim X_{i}=i for each i{n,n1,,1}i\in\{n,n-1,...,1\}.

  2. (2)

    We say that X=:XnXn1X1X=:X_{n}\supset X_{n-1}\supset\cdots\supset X_{1} is a regular ladder of (X,L)(X,L) if this is a ladder and the restriction map

    H0(Xi,L|Xi)H0(Xi1,L|Xi1)H^{0}(X_{i},L|_{X_{i}})\to H^{0}(X_{i-1},L|_{X_{i-1}})

    is surjective for every i{n,n1,,2}i\in\{n,n-1,...,2\}. We say that (X,L)(X,L) has a regular ladder if there exists a regular ladder of (X,L)(X,L).

Theorem 2.21.

Let XX be a Fano threefold with rX=2r_{X}=2. Let HH be a Cartier divisor such that KX2H-K_{X}\sim 2H. Then the following hold.

  1. (1)

    Δ(X,H)=1\Delta(X,H)=1.

  2. (2)

    (X,H){\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}(X,H)} has a regular ladder.

  3. (3)

    If H32H^{3}\geq 2, then |H||H| is base point free.

  4. (4)

    If H33H^{3}\geq 3, then |H||H| is very ample.

Proof.

The assertion (1) follows from [Meg98, Lemma 5]. Then (2)–(4) hold by [Fuj82b, Corollary (5.5)]. ∎

Lemma 2.22.

Let CC be a projective Gorenstein curve such that ωC𝒪C\omega_{C}\simeq\mathcal{O}_{C}. Let LL be a Cartier divisor on CC. Then the following hold.

  1. (1)

    If degL3\deg L\geq 3, then R(X,L)R(X,L) is generated by H0(X,L)H^{0}(X,L) as a kk-algebra.

  2. (2)

    If degL=2\deg L=2, then R(X,L)R(X,L) is generated by η1,η2H0(X,L)\eta_{1},\eta_{2}\in H^{0}(X,L) and η3H0(X,2L)\eta_{3}\in H^{0}(X,2L) as a kk-algebra.

  3. (3)

    If degL=1\deg L=1, then R(X,L)R(X,L) is generated by η1H0(X,L),η2H0(X,2L),η3H0(X,3L)\eta_{1}\in H^{0}(X,L),\eta_{2}\in H^{0}(X,2L),\eta_{3}\in H^{0}(X,3L) as a kk-algebra.

Proof.

By [Tan21, Proposition 11.11], the following hold.

  1. (i)

    If degL3\deg L\geq 3, then R(X,L)R(X,L) is generated by H0(X,L)H^{0}(X,L).

  2. (ii)

    If degL=2\deg L=2, then R(X,L)R(X,L) is generated by H0(X,L)H0(X,2L)H^{0}(X,L)\oplus H^{0}(X,2L).

  3. (iii)

    If degL=1\deg L=1, then R(X,L)R(X,L) is generated by H0(X,L)H0(X,2L)H0(X,3L)H^{0}(X,L)\oplus H^{0}(X,2L)\oplus H^{0}(X,3L).

Then (i) implies (1).

We only prove (3), as the proof of (2) is similar. By the Riemann–Roch theorem, we have h0(X,mL)=mh^{0}(X,mL)=m for any m1m\geq 1. Fix a nonzero element η1H0(X,L)\eta_{1}\in H^{0}(X,L), so that we have H0(X,L)=kη1H^{0}(X,L)=k\eta_{1}. Then η12H0(X,2L)\eta_{1}^{2}\in H^{0}(X,2L) is nonzero, and hence we can find η2H0(X,2L)\eta_{2}\in H^{0}(X,2L) such that H0(X,2L)=kη12kη2H^{0}(X,2L)=k\eta_{1}^{2}\oplus k\eta_{2}. Since

×η1:H0(X,2L)H0(X,3L),sη1s\times\eta_{1}:H^{0}(X,2L)\to H^{0}(X,3L),\qquad s\mapsto\eta_{1}s

is injective, η13,η1η2H0(X,3L)\eta_{1}^{3},\eta_{1}\eta_{2}\in H^{0}(X,3L) are linearly independent over kk. Therefore, we can find η3H0(X,3L)\eta_{3}\in H^{0}(X,3L) such that

H0(X,3L)=kη13kη1η2kη3.H^{0}(X,3L)=k\eta_{1}^{3}\oplus k\eta_{1}\eta_{2}\oplus k\eta_{3}.

Hence η1,η2,η3\eta_{1},\eta_{2},\eta_{3} generate R(X,L)R(X,L) as a kk-algebra. Thus (3) holds. ∎

Theorem 2.23.

Let XX be a Fano threefold with rX=2r_{X}=2. Let HH be a Cartier divisor such that KX2H-K_{X}\sim 2H. Then 1H371\leq H^{3}\leq 7. Furthermore, the following hold.

  1. (1)

    If H3=1H^{3}=1, then XX is isomorphic to a weighted hypersurface in (1,1,1,2,3)\mathbb{P}(1,1,1,2,3) of degree 66.

  2. (2)

    If H3=2H^{3}=2, then XX is isomorphic to a weighted hypersurface in (1,1,1,1,2)\mathbb{P}(1,1,1,1,2) of degree 44.

  3. (3)

    If H3=3H^{3}=3, then XX is isomorphic to a cubic hypersurface in 4\mathbb{P}^{4}.

  4. (4)

    If H3=4H^{3}=4, then XX is isomorphic to a complete intersection of two quadrics in 5\mathbb{P}^{5}.

  5. (5)

    If H3=5H^{3}=5, then XX is isomorphic to Gr(2,5)H1H2H3{\rm Gr}(2,5)\cap H_{1}\cap H_{2}\cap H_{3} in 9\mathbb{P}^{9}, where the inclusion Gr(2,5)9{\rm Gr}(2,5)\subset\mathbb{P}^{9} is the Plücker embedding and H1,H2,H3H_{1},H_{2},H_{3} are general hyperplanes of 9\mathbb{P}^{9}.

  6. (6)

    If H3=6H^{3}=6, then XX is isomorphic to either 1×1×1\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} or a member of |𝒪2×2(1,1)||\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1)| on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2}.

  7. (7)

    If H3=7H^{3}=7, then XX is isomorphic to a blowup of 3\mathbb{P}^{3} at a point.

Proof.

It follows from [Meg98, Theorem 6] that (3)–(7) holds. Let us show (1) and (2). By Theorem 2.21, there exists a regular ladder of (X,H)(X,H):

XSC,X\supset S\supset C,

and hence the restriction map

H0(X,H)H0(C,HC),HC:=H|CH^{0}(X,H)\to H^{0}(C,H_{C}),\qquad H_{C}:=H|_{C}

is surjective.

We now prove (1). By Proposition 2.19, it is enough to find

η1H0(C,HC),η2H0(C,2HC),η3H0(C,3HC)\eta_{1}\in H^{0}(C,H_{C}),\quad\eta_{2}\in H^{0}(C,2H_{C}),\quad\eta_{3}\in H^{0}(C,3H_{C})

such that R(C,HC)R(C,H_{C}) is generated by η1,η2,η3\eta_{1},\eta_{2},\eta_{3}. This follows from Lemma 2.22(3). Thus (1) holds. By Proposition 2.19 and Lemma 2.22(2), we can apply a similar argument for (2) to that of (1). ∎

Corollary 2.24.

Let XX be a Fano threefold with rX2r_{X}\geq 2. Then the following hold.

  1. (1)

    If (rX,(KX/2)3)(2,1)(r_{X},(-K_{X}/2)^{3})\neq(2,1), then R(X,KX)R(X,-K_{X}) is generated by H0(X,KX)H^{0}(X,-K_{X}), and hence |KX||-K_{X}| is very ample.

  2. (2)

    If (rX,(KX/2)3)=(2,1)(r_{X},(-K_{X}/2)^{3})=(2,1), then |KX||-K_{X}| is base point free and φ|KX|\varphi_{|-K_{X}|} is a double cover onto its image (in particular, |KX||-K_{X}| is not very ample).

Proof.

Let us show (1). If rX3r_{X}\geq 3, then the assertion holds by Theorem 2.18. We may assume that rX=2r_{X}=2. Fix an ample Cartier divisor HH on XX with KX2H-K_{X}\sim 2H. If H33H^{3}\geq 3, then |H||H| is very ample, and hence also |KX|=|2H||-K_{X}|=|2H| is very ample. Thus we may assume that (rX,H3)=(2,2)(r_{X},H^{3})=(2,2). In this case, |H||H| is base point free. By the proof of Theorem 2.23, R(X,H)R(X,H) is generated by x0,x1,x2,x3H0(X,H)x_{0},x_{1},x_{2},x_{3}\in H^{0}(X,H) and yH0(X,2H)y\in H^{0}(X,2H) as a kk-algebra. This immediately deduces that R(X,2H)R(X,2H) is generated by H0(X,2H)H^{0}(X,2H), because if a monomial x0a0x1a1x2a2x3a3ybx_{0}^{a_{0}}x_{1}^{a_{1}}x_{2}^{a_{2}}x_{3}^{a_{3}}y^{b} is of even degree, i.e., a0+a1+a2+a3+2b2>0a_{0}+a_{1}+a_{2}+a_{3}+2b\in 2\mathbb{Z}_{>0}, then we get a0+a1+a2+a320a_{0}+a_{1}+a_{2}+a_{3}\in 2\mathbb{Z}_{\geq 0}, and hence x0a0x1a1x2a2x3a3ybx_{0}^{a_{0}}x_{1}^{a_{1}}x_{2}^{a_{2}}x_{3}^{a_{3}}y^{b} can be written as a multiple of elements of H0(X,2H)H^{0}(X,2H). Therefore, |KX|=|2H||-K_{X}|=|2H| is very ample.

The assertion (2) follows from [Meg98, Theorem 8]. ∎

3. K3-like surfaces

Throughout this section, we work over a field κ\kappa of characteristic p>0p>0. In our applications, we have κ=k(t1,,tN)\kappa=k(t_{1},...,t_{N}), which is a purely transcendental field extension over an algebraically closed field kk of characteristic p>0p>0. In particular, κ\kappa can not be assumed to be perfect.

Definition 3.1.

We work over a field κ\kappa of characteristic p>0p>0. We say that SS is a K3-like surface if SS is a projective normal surface such that H1(S,𝒪S)=0H^{1}(S,\mathcal{O}_{S})=0 and KS0K_{S}\sim 0.

We are mainly interested in the following two cases: geometrically integral regular K3-like surfaces and geometrically integral canonical K3-like surfaces, i.e., SS has at worst canonical singularities.

3.1. Vanishing theorems

The purpose of this subsection is to establish the Kodaira (or Ramnujam) vanishing theorem for geometrically integral K3-like surfaces (Theorem 3.4). To this end, we shall establish a vanishing of Mumford type for geometrically integral surfaces (Theorem 3.3). We start with the following auxiliary result.

Proposition 3.2.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a projective normal surface and let DD be a nef and big effective \mathbb{Q}-Cartier \mathbb{Z}-divisor. Then H0(D,𝒪D)H^{0}(D,\mathcal{O}_{D}) is a field and

H0(S,𝒪S)H0(D,𝒪D)H^{0}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}S},\mathcal{O}_{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}S})\hookrightarrow H^{0}(D,\mathcal{O}_{D})

is a finite purely inseparable extension.

Proof.

For κ:=H0(S,𝒪S)\kappa^{\prime}:=H^{0}(S,\mathcal{O}_{S}), we have the following induced morphisms:

SSpecκSpecκ.S\to{\operatorname{Spec}}\,\kappa^{\prime}\to{\operatorname{Spec}}\,\kappa.

Recall that κ\kappa^{\prime} is a field which is a finitely generated κ\kappa-module [Har77, Ch. III, Theorem 8.8(b)], i.e., κκ\kappa\subset\kappa^{\prime} is a field extension of finite degree. Replacing κ\kappa by κ=H0(S,𝒪S){\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa^{\prime}=}H^{0}(S,\mathcal{O}_{S}), we may assume that H0(S,𝒪S)=κH^{0}(S,\mathcal{O}_{S})=\kappa. Then the assertion follows from [Eno, Collorary 3.13 and Corollary 3.17]. ∎

Theorem 3.3.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral projective normal surface. Let DD be a nef and big effective \mathbb{Q}-Cartier \mathbb{Z}-divisor such that h0(S,𝒪S(D))2h^{0}(S,\mathcal{O}_{S}(D))\geq 2. Then the following hold.

  1. (1)

    The induced map H0(S,𝒪S)H0(D,𝒪D)H^{0}(S,\mathcal{O}_{S})\to H^{0}(D,\mathcal{O}_{D}) is an isomorphism.

  2. (2)

    The induced map H1(S,𝒪S(D))H1(S,𝒪S)H^{1}(S,\mathcal{O}_{S}(-D))\to H^{1}(S,\mathcal{O}_{S}) is injective.

Proof.

Since SS is geometrically integral, we have H0(S,𝒪S)=κH^{0}(S,\mathcal{O}_{S})=\kappa. As the assertion (2) immediately holds by (1), it suffices to show (1). Taking the base change to the separable closure of κ\kappa, we may assume that κ\kappa is separably closed. Since SS is geometrically integral, there exists a κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}-rational point PSP\in S around which SS is smooth. In particular, 𝒪S(D)\mathcal{O}_{S}(D) is invertible around PSP\in S. We then obtain a short exact sequence

0𝒪S(D)𝔪P𝒪S(D)𝒪S(D)𝒪P0,0\to\mathcal{O}_{S}(D)\otimes\mathfrak{m}_{P}\to\mathcal{O}_{S}(D)\to\mathcal{O}_{S}(D)\otimes\mathcal{O}_{P}\to 0,

which induces another exact sequence:

0H0(S,𝒪S(D)𝔪P)H0(S,𝒪S(D))H0(P,𝒪S(D)|P)κ.0\to H^{0}(S,\mathcal{O}_{S}(D)\otimes\mathfrak{m}_{P})\to H^{0}(S,\mathcal{O}_{S}(D))\to H^{0}(P,\mathcal{O}_{S}(D)|_{P})\simeq{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}.

By h0(S,𝒪S(D))2>1=h0(P,𝒪S(D)|P)h^{0}(S,\mathcal{O}_{S}(D))\geq 2>1=h^{0}(P,\mathcal{O}_{S}(D)|_{P}), we get H0(S,𝒪S(D)𝔪P)0H^{0}(S,\mathcal{O}_{S}(D)\otimes\mathfrak{m}_{P})\neq 0. Hence we may assume that PSuppDP\in{\operatorname{Supp}}\,D. By Proposition 3.2, we have field extensions

κ=H0(S,𝒪S)H0(D,𝒪D)H0(P,𝒪P)=κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}=H^{0}(S,\mathcal{O}_{S})\hookrightarrow H^{0}(D,\mathcal{O}_{D})\hookrightarrow H^{0}(P,\mathcal{O}_{P})={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}

whose composite is identity. Therefore, H0(S,𝒪S)H0(D,𝒪D)H^{0}(S,\mathcal{O}_{S})\to H^{0}(D,\mathcal{O}_{D}) is an isomorphism. Thus (1) holds. ∎

Theorem 3.4.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral regular K3-like surface. Let LL be a nef and big Cartier divisor. Then the following hold.

  1. (1)

    Hj(S,𝒪S(L))=0H^{j}(S,\mathcal{O}_{S}(-L))=0 for any j<2j<2.

  2. (2)

    Hi(S,𝒪S(L))=0H^{i}(S,\mathcal{O}_{S}(L))=0 for any i>0i>0.

  3. (3)

    h0(S,𝒪S(L))=2+12L2h^{0}(S,\mathcal{O}_{S}(L))=2+\frac{1}{2}L^{2}.

Proof.

By Serre duality, we have H2(S,L)=0H^{2}(S,L)=0. It follows from the Riemann–Roch theorem that

h0(S,L)h0(S,L)h1(S,L)=χ(S,L)=χ(S,𝒪S)+12L(LKS)=2+12L2>2.h^{0}(S,L)\geq h^{0}(S,L)-h^{1}(S,L)=\chi(S,L)=\chi(S,\mathcal{O}_{S})+\frac{1}{2}L\cdot(L-K_{S})=2+\frac{1}{2}L^{2}>2.

We then get H1(S,L)=0H^{1}(S,-L)=0 (Theorem 3.3), which implies H1(S,L)=0H^{1}(S,L)=0 by Serre duality. Thus (1) and (2) hold, which implies (3). ∎

3.2. Linear systems

In this subsection, we study base locus of divisors on geometrically integral regular K3-like surfaces (Theorem 3.16). The hardest part is to prove that nef and big divisors without fixed components is base point free (Proposition 3.10). Many arguments are based on the proofs of the corresponding results over algebraically closed fields [Huy16, Chapter 2].

Proposition 3.5.

We work over a field κ\kappa of characteristic p>0p>0. Let CC be a projective regular curve such that degKC0\deg K_{C}\geq 0. Then |KC||K_{C}| is base point free.

Proof.

Replacing κ\kappa by H0(C,𝒪C)H^{0}(C,\mathcal{O}_{C}), the problem is reduced to the case when H0(C,𝒪C)=κH^{0}(C,\mathcal{O}_{C})=\kappa. Set

g:=h1(C,𝒪C)=h0(C,KC).g:=h^{1}(C,\mathcal{O}_{C})=h^{0}(C,K_{C}).

By the Riemann–Roch theorem and Serre duality, we obtain

χ(C,𝒪C)=χ(C,KC)=degKC+χ(C,𝒪C),-\chi(C,\mathcal{O}_{C})=\chi(C,K_{C})=\deg K_{C}+\chi(C,\mathcal{O}_{C}),

which implies

degKC=2g2.\deg K_{C}=2g-2.

If degKC=0\deg K_{C}=0, i.e., g=h1(C,𝒪C)=1g=h^{1}(C,\mathcal{O}_{C})=1, then H0(C,KC)0H^{0}(C,K_{C})\neq 0, which implies KC0K_{C}\sim 0.

We may assume that 2g2=degKC>02g-2=\deg K_{C}>0, i.e., g2g\geq 2. Then KCK_{C} is ample. By h0(C,KC)=g>0h^{0}(C,K_{C})=g>0, there exists an effective Cartier divisor DD on CC with KCDK_{C}\sim D.

Fix a closed point PCP\in C. It suffices to show that P|KC|P\not\in|K_{C}|. We may assume that PSuppDP\in{\operatorname{Supp}}\,D, i.e., we can write

D=aP+DD=aP+D^{\prime}

for some a>0a\in\mathbb{Z}_{>0} and an effective Cartier divisor DD^{\prime} with PSuppDP\not\in{\operatorname{Supp}}\,D^{\prime}. We have the following exact sequence:

H0(C,𝒪C(KC))H0(P,𝒪C(KC)|P)H1(C,𝒪C(KCP))H1(C,𝒪C(KC))0.H^{0}(C,\mathcal{O}_{C}(K_{C}))\to H^{0}(P,\mathcal{O}_{C}(K_{C})|_{P})\to H^{1}(C,\mathcal{O}_{C}(K_{C}-P))\to H^{1}(C,\mathcal{O}_{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}C}(K_{C}))\to 0.

If dimκH1(C,KCP)=1\dim_{\kappa}H^{1}(C,K_{C}-P)=1, then H1(C,KCP)H1(C,KC)H^{1}(C,K_{C}-P)\xrightarrow{\simeq}H^{1}(C,K_{C}), and hence we are done. Therefore, we may assume that dimκH1(C,KCP)2\dim_{\kappa}H^{1}(C,K_{C}-P)\geq 2, i.e., dimκH0(C,𝒪C(P))2\dim_{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}}H^{0}(C,\mathcal{O}_{C}(P))\geq 2. Then PP is linearly equivalent to an effective Cartier divisor EE with EPE\neq P. Therefore, we obtain PEP\not\in E. This implies that |P||P| is base point free. Hence PP is not a base point of |KC||K_{C}|. ∎

Proposition 3.6.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a regular K3-like surface. Let LL be a Cartier divisor on SS. If LL is nef and L2=0L^{2}=0, then |L||L| is base point free.

Proof.

Replacing κ\kappa by H0(S,𝒪S)H^{0}(S,\mathcal{O}_{S}), we may assume that H0(S,𝒪S)=κH^{0}(S,\mathcal{O}_{S})=\kappa.

We first treat the case when L0L\equiv 0. It suffices to show that L0L\sim 0, i.e., H0(S,L)0H^{0}(S,L)\neq 0. Otherwise, we have that H0(S,L)=H0(S,L)=0H^{0}(S,L)=H^{0}(S,-L)=0, and hence H2(S,L)=0H^{2}(S,L)=0 by Serre duality. Therefore, the Riemann–Roch theorem implies that

0h1(S,L)=χ(S,L)=12L2+2=2>0,0\geq-h^{1}(S,L)=\chi(S,L)=\frac{1}{2}L^{2}+2=2>0,

which is absurd.

Let us handle the remaining case: L0L\not\equiv 0. By Serre duality, we obtain H2(S,L)=0H^{2}(S,L)=0. Then the Riemann–Roch theorem implies that

h0(S,L)h0(S,L)h1(S,L)=χ(S,L)=12L2+2=2.h^{0}(S,L)\geq h^{0}(S,L)-h^{1}(S,L)=\chi(S,L)=\frac{1}{2}L^{2}+2=2.

Take the decomposition L=M+FL=M+F, where MM is the mobile part of |L||L| and FF is the fixed part of |L||L|. Note that MM is nef. We have

0=L2=L(M+F)=LM+LF.0=L^{2}=L\cdot(M+F)=L\cdot M+L\cdot F.

It follows from LM0L\cdot M\geq 0 and LF0L\cdot F\geq 0 that

LM=LF=0.L\cdot M=L\cdot F=0.

By M20,FM0M^{2}\geq 0,F\cdot M\geq 0, and (M+F)M=LM=0(M+F)\cdot M=L\cdot M=0, we get

M2=MF=0.M^{2}=M\cdot F=0.

Therefore, we obtain

F2=(LM)F=0.F^{2}=(L-M)\cdot F=0.

If F0F\neq 0, then

h0(S,F)h0(S,F)h1(S,F)=χ(S,F)=12F2+2=2,h^{0}(S,F)\geq h^{0}(S,F)-h^{1}(S,F)=\chi(S,F)=\frac{1}{2}F^{2}+2=2,

which is a contradiction. Therefore, we have F=0F=0, i.e., the base locus Bs|L|{\rm Bs}\,|L| is zero-dimensional (if non-empty). However, if Bs|L|{\rm Bs}\,|L| is non-empty and zero-dimensional, then we again obtain a contradiction by L2=0L^{2}=0. Thus |L||L| is base point free. ∎

Lemma 3.7.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a regular K3-like surface. Let CC be a prime divisor on SS with C20C^{2}\geq 0. Then the following hold.

  1. (1)

    dimBs|C|0\dim{\operatorname{Bs}}\,|C|\leq 0.

  2. (2)

    If CC is regular, then |C||C| is base point free.

Proof.

If C2=0C^{2}=0, then |C||C| is base point free by Proposition 3.6. Hence we may assume that C2>0C^{2}>0. We then obtain h0(S,C)2h^{0}(S,C)\geq 2 by the Riemann–Roch theorem, which implies (1).

Let us show (2). Consider the exact sequence:

H0(S,𝒪S(C))H0(C,𝒪S(C)|C)H0(C,KC)H1(S,𝒪S)=0.H^{0}(S,\mathcal{O}_{S}(C))\to H^{0}(C,\mathcal{O}_{S}(C)|_{C})\simeq H^{0}(C,K_{C})\to H^{1}(S,\mathcal{O}_{S})=0.

As |KC||K_{C}| is base point free (Proposition 3.5), |C||C| is base point free. ∎

Lemma 3.8.

We work over a field κ\kappa of characteristic p>0p>0. Let CC be a projective Gorenstein curve such that ωC1\omega_{C}^{-1} is ample. Let LL be a nef line bundle on CC. Then |L||L| is base point free.

Proof.

Replacing κ\kappa by H0(C,𝒪C)H^{0}(C,\mathcal{O}_{C}), we may assume that H0(C,𝒪C)=κH^{0}(C,\mathcal{O}_{C})=\kappa. By [Kol13, Lemma 10.6], there exists an ample invertible sheaf HH on CC such that PicC[H]{\operatorname{Pic}}\,C\simeq\mathbb{Z}[H], where [H][H] denotes the isomorphism class of HH. Therefore, it suffices to show that |H||H| is base point free. Again by [Kol13, Lemma 10.6], one of the following holds.

  1. (1)

    Cκ1C\simeq\mathbb{P}^{1}_{\kappa}.

  2. (2)

    CC is a conic on κ2\mathbb{P}^{2}_{\kappa} and HωC1H\simeq\omega_{C}^{-1}.

Hence we may assume that (2) holds. For the embedding Cκ2C\subset\mathbb{P}^{2}_{\kappa}, we obtain

𝒪κ2(1)|C(ωκ2𝒪2(C))|CωC.\mathcal{O}_{\mathbb{P}^{2}_{\kappa}}(-1)|_{C}\simeq(\omega_{\mathbb{P}^{2}_{\kappa}}\otimes\mathcal{O}_{\mathbb{P}^{2}}(C))|_{C}\simeq\omega_{C}.

Therefore, HωC1𝒪k2(1)|CH\simeq\omega_{C}^{-1}\simeq\mathcal{O}_{\mathbb{P}^{2}_{k}}(1)|_{C}, which implies that |H||H| is very ample. ∎

Lemma 3.9.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a regular projective surface. Let DD be an effective \mathbb{Z}-divisor. Then the following hold.

  1. (1)

    If DD is nef and big, then DD is 11-connected, i.e., C1C2>0C_{1}\cdot C_{2}>0 holds for all nonzero effective \mathbb{Z}-divisors C1,C2C_{1},C_{2} with D=C1+C2D=C_{1}+C_{2}.

  2. (2)

    If DD is 11-connected, then H0(D,𝒪D)H^{0}(D,\mathcal{O}_{D}) is a field.

Proof.

Let us show (1). By the same argument as in [Har77, Ch. V, Theorem 1.9], the Hodge index theorem holds even over an imperfect field, i.e., the signature of the intersection product is (1,ρ(S)1)(1,\rho(S)-1). Then the argument as in [Huy16, Remark 1.7] works without any changes. Thus (1) holds.

Let us show (2). Take a maximal effective \mathbb{Z}-divisor DD^{\prime} such that 0DD0\leq D^{\prime}\leq D and H0(D,𝒪D)H^{0}(D^{\prime},\mathcal{O}_{D^{\prime}}) is a field (such DD^{\prime} exists, since a prime divisor satisfies this condition). Assume DDD\neq D^{\prime}. Then we have D(DD)>0D^{\prime}\cdot(D-D^{\prime})>0. Then DC>0D^{\prime}\cdot C>0 for some prime divisor CSupp(DD)C\subset{\operatorname{Supp}}\,(D-D^{\prime}). We have an exact sequence

0𝒪S(D)|C𝒪D+C𝒪D0,0\to\mathcal{O}_{S}(-D^{\prime})|_{C}\to\mathcal{O}_{D^{\prime}+C}\to\mathcal{O}_{D^{\prime}}\to 0,

which induces the following injection

H0(D+C,𝒪D+C)H0(D,𝒪D).H^{0}(D^{\prime}+C,\mathcal{O}_{D^{\prime}+C})\hookrightarrow H^{0}(D^{\prime},\mathcal{O}_{D^{\prime}}).

Then H0(D+C,𝒪D+C)H^{0}(D^{\prime}+C,\mathcal{O}_{D^{\prime}+C}) is an intermediate ring between κ\kappa and H0(D,𝒪D)H^{0}(D^{\prime},\mathcal{O}_{D^{\prime}}). Since the field extension κH0(D,𝒪D)\kappa\subset H^{0}(D^{\prime},\mathcal{O}_{D^{\prime}}) is of finite degree, H0(D+C,𝒪D+C)H^{0}(D^{\prime}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+C},\mathcal{O}_{D^{\prime}+C}) is a field. This contradicts the maximality of DD^{\prime}. Thus (2) holds.

Proposition 3.10.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral regular K3-like surface. Let MM be a nef and big Cartier divisor on SS. If dimBs|M|0\dim{\rm Bs}|M|\leq 0, then |M||M| is base point free.

Proof.

Taking the base change to the separable closure of κ\kappa, we may assume that κ\kappa is separably closed. We prove the assertion by induction on M2M^{2}. We may start with the case when M2=1M^{2}=1 as the base of this induction. By M22M^{2}\in 2\mathbb{Z}, there is nothing to show. In what follows, we assume that |M~||{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{M}}| is base point free if M~{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{M}} is a nef and big Cartier divisor on SS such that M~2<M2{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{M}}^{2}<M^{2} and dimBs|M~|0\dim{\rm Bs}\,|{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\widetilde{M}}|\leq 0.

Suppose that |M||M| is not base point free. Let us derive a contradiction. Note that Bs|M|{\rm Bs}\,|M| (set-theoretically) consists of finitely many closed points Q:=Q1,Q2,,QrQ:=Q_{1},Q_{2},...,Q_{r}. Set d(Q):=[k(Q):κ]d(Q):=[k(Q):{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}].

Step 1.

There exists an index 1ir1\leq i\leq r such that the generic member MM^{\prime} of |M||M| is not regular at QiQ_{i}.

Proof of Step 1.

Suppose the contrary, i.e., suppose that MM^{\prime} is regular at each of Q1,,QrQ_{1},...,Q_{r}. By Theorem 2.9, MM^{\prime} is regular outside Q1,,QrQ_{1},...,Q_{r}. In particular, MM^{\prime} is a regular curve on a regular K3-like surface. Then |Mκκ|=|M||M\otimes_{\kappa}\kappa^{\prime}|=|M^{\prime}| is base point free (Lemma 3.7(2)), where κ:=K((H0(S,M)))\kappa^{\prime}:=K(\mathbb{P}(H^{0}(S,M))). Hence also |M||M| is base point free, which is absurd. This completes the proof of Step 1. ∎

After possibly permuting the indices, we may assume that i=1i=1. In particular, any member N|M|N\in|M| is not regular at QQ. Set

μ:=minD|M|{multQD}.\mu:={\operatorname{min}}_{D\in|M|}\{{\operatorname{mult}}_{Q}\,D\}.

We have μ2\mu\geq 2. Let f:TSf:T\to S be the blowup at QQ. Set E:=Ex(f)E:={\operatorname{Ex}}(f) and N:=fM2EN:=f^{*}M-2E.

Step 2.

It holds that H1(T,N)0H^{1}(T,-N)\neq 0.

Proof of Step 2.

By the exact sequence

H0(T,fM)0H0(E,fM)H1(T,fME)H^{0}(T,f^{*}M)\xrightarrow{0}H^{0}(E,f^{*}M)\to H^{1}(T,f^{*}M-E)

and H0(E,fM)0H^{0}(E,f^{*}M)\neq 0, we obtain H1(T,fME)0H^{1}(T,f^{*}M-E)\neq 0. By Serre duality and KTfKS+EEK_{T}\sim f^{*}K_{S}+E\sim E, we get H1(T,(fM2E))0H^{1}(T,-(f^{*}M-2E))\neq 0. This completes the proof of Step 2. ∎

Step 3.

fMμEf^{*}M-\mu^{\prime}E is nef for any 0μμ0\leq\mu^{\prime}\leq\mu. In particular, NN is nef.

Proof of Step 3.

Since fMf^{*}M is nef, it suffices to show that fMμEf^{*}M-\mu E is nef. Pick two general members M1,M2|M|M_{1},M_{2}\in|M|. We then obtain

multQM1=multQM2=μ,dim(M1M2)=0.{\operatorname{mult}}_{Q}M_{1}={\operatorname{mult}}_{Q}M_{2}=\mu,\qquad\dim(M_{1}\cap M_{2})=0.

Therefore, fM1μEf^{*}M_{1}-\mu E and fM2μEf^{*}M_{2}-\mu E share no irreducible components. In particular, fMμEf^{*}M-\mu E is nef. This completes the proof of Step 3. ∎

Step 4.

The following hold.

  1. (1)

    N2=0N^{2}=0.

  2. (2)

    μ=2\mu=2.

  3. (3)

    r=1r=1, i.e., the set-theoretic equation Bs|M|={Q}{\rm Bs}|M|=\{Q\} holds.

  4. (4)

    M2=4d(Q)M^{2}=4d(Q).

  5. (5)

    For any M1|M|M_{1}\in|M|, we have multQM1=2{\operatorname{mult}}_{Q}M_{1}=2.

  6. (6)

    |N||N| is base point free.

Proof of Step 4.

Note that NN is nef by Step 3.

Let us show (1). If N2>0N^{2}>0, then NN is nef and big. Then also fNf_{*}N is nef and big, and hence h0(T,N)h0(S,fN)2h^{0}(T,N)\geq h^{0}(S,f_{*}N)\geq 2. Since TT is geometrically integral and H1(T,𝒪T)=0H^{1}(T,\mathcal{O}_{T})=0, we get H1(T,N)=0H^{1}(T,-N)=0 (Theorem 3.3), which contradicts Step 2. Thus (1) holds.

Let us show (2). Suppose the contrary, i.e., μ3\mu\geq 3. Note that fMμEf^{*}M-\mu^{\prime}E is nef for any 0μ30\leq\mu^{\prime}\leq 3. By (fM)2>0(f^{*}M)^{2}>0 and (fM3E)20(f^{*}M-3E)^{2}\geq 0, we have that N2=(fM2E)2>0N^{2}=(f^{*}M-2E)^{2}>0, which contradicts (1). Thus (2) holds.

Let us show (3). Suppose r2r\geq 2. Pick two general members M1,M2|M|M_{1},M_{2}\in|M|. Then we have that

multQM1=multQM2=2,dim(M1M2)=0.{\operatorname{mult}}_{Q}M_{1}={\operatorname{mult}}_{Q}M_{2}=2,\qquad\dim(M_{1}\cap M_{2})=0.

Then

(fM12E)(fM22E)(f^{*}M_{1}-2E)\cap(f^{*}M_{2}-2E)

is zero-dimensional. By r2r\geq 2,

(fM12E)(fM22E),(f^{*}M_{1}-2E)\cap(f^{*}M_{2}-2E)\neq\emptyset,

i.e., this set contains the point f1(Q2)f^{-1}(Q_{2}). Hence, we obtain N2=(fM12E)(fM22E)>0N^{2}=(f^{*}M_{1}-2E)\cdot(f^{*}M_{2}-2E)>0, which contradicts (1). Thus (3) holds.

Let us show (4). We have that 0=N2=(fM2E)2=M2+4E2=M24d(Q)0=N^{2}=(f^{*}M-2E)^{2}=M^{2}+4E^{2}=M^{2}-4d(Q), and hence (4) holds.

Let us show (5). Suppose the contrary, i.e., multQM13{\operatorname{mult}}_{Q}M_{1}\geq 3 for some M1|M|M_{1}\in|M|. Pick a general member M2|M|M_{2}\in|M|. We have that multQM2=2{\operatorname{mult}}_{Q}M_{2}=2, and dim(M1M2)0\dim(M_{1}\cap M_{2})\leq 0. It holds that N2=(fM12E)(fM22E)>0N^{2}=(f^{*}M_{1}-2E)\cdot(f^{*}M_{2}-2E)>0, which contradicts (1). Thus, (5) holds.

The assertion (6) follows from the proof of (3). This completes the proof of Step 4. ∎

Step 5.

A general member of |N||N| is a disjoint union of irreducible effective \mathbb{Z}-divisors.

Proof of Step 5.

Let

π:=φ|N|:TBκh0(T,N)1.\pi:=\varphi_{|N|}:T\to B\subset\mathbb{P}_{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}^{h^{0}(T,N)-1}.

be the morphism induced by |N||N|, where BB denotes its image. Since |N||N| is base point free with N0N\not\equiv 0 (Step 2), BB is a (possibly non-regular) curve. Let

π:TπB𝜃B\pi:T\xrightarrow{\pi^{\prime}}B^{\prime}\xrightarrow{\theta}B

be the Stein factorisation. Note that each of BB and BB^{\prime} has infinitely many κ\kappa-rational points. Pick a general closed point bBb\in B. Then θ1(b)\theta^{-1}(b) consists of finitely many general closed points of BB^{\prime}. Since general fibres of TBT\to B^{\prime} are (geometrically) irreducible, π(b)\pi^{*}(b) is a disjoint union of irreducible divisors. This implies that a general member of |N||N| is a disjoint union of irreducible divisors. This completes the proof of Step 5. ∎

Step 6.

Fix a non-empty open subset TTT^{\prime}\subset T on which ff is an isomorphism. Then there exist a κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}-rational point RTR\in T^{\prime} and N|N|N^{\prime}\in|N| such that RNR\in N^{\prime} and NN^{\prime} is a disjoint union of irreducible divisors. Set M:=fNM^{\prime}:=f_{*}N^{\prime}. In particular, MM^{\prime} passes through the κ\kappa-rational point RS:=f(R)SR_{S}:=f(R)\in S.

Proof of Step 6.

Fix two general members N1,N2|N|N_{1},N_{2}\in|N|, so that each of N1N_{1} and N2N_{2} is a disjoint union of irreducible effective \mathbb{Z}-divisors (Step 6). Let VH0(T,N)V\subset H^{0}(T,N) be the two-dimensional κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}-vector subspace corresponding to the pencil generated by N1,N2N_{1},N_{2}. For the pencil |V||V| corresponding to VV, a general member N|V|N^{\prime}\in|V| is a disjoint union of irreducible divisors. Therefore, except for finitely many members N~1,,N~t\widetilde{N}_{1},...,\widetilde{N}_{t}, any member of |V||V| is a disjoint union of irreducible divisors. Pick a κ\kappa-rational point RTR\in T^{\prime} such that Rj=1tN~jR\not\in\bigcup_{j=1}^{t}\widetilde{N}_{j}. Consider an exact sequence:

0H0(T,N𝔪R)H0(T,N)H0(R,N|R).0\to H^{0}(T,N\otimes\mathfrak{m}_{R})\to H^{0}(T,N)\to H^{0}(R,N|_{R}).

By dimV=2\dim V=2 and dimH0(R,N|R)=1\dim H^{0}(R,N|_{R})=1, we have that H0(T,N𝔪R)V{0}H^{0}(T,N\otimes\mathfrak{m}_{R})\cap V\neq\{0\}. Then there is a member N|N|N^{\prime}\in|N| such that RNR\in N^{\prime}. By the choice of RR, we have that NN~1,,NN~tN^{\prime}\neq\widetilde{N}_{1},...,N^{\prime}\neq\widetilde{N}_{t}. Then NN^{\prime} is a disjoint union of irreducible divisors, as required. This completes the proof of Step 6. ∎

Step 7.

NN^{\prime} is not 11-connected, i.e., there exists a decomposition

N=C1,T+C2,TN^{\prime}=C_{1,T}+C_{2,T}

for some nonzero effective \mathbb{Z}-divisors C1,T,C2,TC_{1,T},C_{2,T} such that C1,TC2,T0C_{1,T}\cdot C_{2,T}\leq 0. Set Ci:=fCi,TC_{i}:=f_{*}C_{i,T}.

Proof of Step 7.

If NN^{\prime} is 11-connected, then H0(N,𝒪N)H^{0}(N^{\prime},\mathcal{O}_{N^{\prime}}) is a field (Lemma 3.9). Since NN^{\prime} contains a κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}-rational point RR (Step 6), it holds that H0(N,𝒪N)=κH^{0}(N^{\prime},\mathcal{O}_{N^{\prime}})=\kappa. Then we have an exact sequence

H0(T,𝒪T)H0(N,𝒪N)H1(T,N)H1(T,𝒪T)=0.H^{0}(T,\mathcal{O}_{T})\xrightarrow{\simeq}H^{0}(N^{\prime},\mathcal{O}_{N^{\prime}})\to H^{1}(T,-N^{\prime})\to H^{1}(T,\mathcal{O}_{T})=0.

Thereofore, we obtain

0=H1(T,N)H1(T,N),0=H^{1}(T,-N^{\prime})\simeq H^{1}(T,-N),

which contradicts Step 2. This completes the proof of Step 7. ∎

Step 8.

The following hold.

  1. (i)

    C10C_{1}\neq 0 and C20C_{2}\neq 0.

  2. (ii)

    For each i{1,2}i\in\{1,2\}, we have QCiQ\in C_{i}.

  3. (iii)

    For each i{1,2}i\in\{1,2\}, we have multQCi=1{\operatorname{mult}}_{Q}C_{i}=1 and Ci,T=fCiEC_{i,T}=f^{*}C_{i}-E.

  4. (iv)

    C1,TC2,T=C1C2d(Q)C_{1,T}\cdot C_{2,T}=C_{1}\cdot C_{2}-d(Q).

Proof of Step 8.

The assertion (i) follows from the fact that NN^{\prime} does not contain EE.

Let us show (ii). If one of C1C_{1} and C2C_{2}, say C1C_{1}, does not contain QQ, then C1,TE=0C_{1,T}\cdot E=0, which implies

C1,TC2,T=C1,TffC2,T=C1C2>0.C_{1,T}\cdot C_{2,T}=C_{1,T}\cdot f^{*}f_{*}C_{2,T}=C_{1}\cdot C_{2}>0.

Here C1C2>0C_{1}\cdot C_{2}>0 follows from MfN=C1+C2M\sim f_{*}N^{\prime}=C_{1}+C_{2} and MM is 11-connected (Lemma 3.9(1)). This contradicts Step 7. Thus (ii) holds.

Let us show (iii). It follows from (ii) that multQC11{\operatorname{mult}}_{Q}C_{1}\geq 1 and multQC21{\operatorname{mult}}_{Q}C_{2}\geq 1. By multQ(M)=2{\operatorname{mult}}_{Q}(M^{\prime})=2 and M=C1+C2M^{\prime}=C_{1}+C_{2}, we get multQC1=multQC2=1{\operatorname{mult}}_{Q}C_{1}={\operatorname{mult}}_{Q}C_{2}=1. Since Ci,TC_{i,T} does not contain EE, we have that Ci,T=fCiEC_{i,T}=f^{*}C_{i}-E. Thus (iii) holds.

The assertion (iv) follows from

C1,TC2,T=(fC1E)(fC2E)=C1C2d(Q).C_{1,T}\cdot C_{2,T}=(f^{*}C_{1}-E)\cdot(f^{*}C_{2}-E)=C_{1}\cdot C_{2}-d(Q).

This completes the proof of Step 8. ∎

Step 9.

Each of C1C_{1} and C2C_{2} is a prime divisor. Furthermore, one of the following holds.

  1. (I)

    MM^{\prime} and NN^{\prime} are irreducible. Furthermore, M=2CM^{\prime}=2C and N=2CTN^{\prime}=2C_{T} for C:=C1=C2C:=C_{1}=C_{2} and CT:=C1,T=C2,TC_{T}:=C_{1,T}=C_{2,T}.

  2. (II)

    C1C2C_{1}\neq C_{2}, C1,TC2,T=0C_{1,T}\cdot C_{2,T}=0, and C1C2=d(Q)C_{1}\cdot C_{2}=d(Q). Moreover, C1,TC2,T=C_{1,T}\cap C_{2,T}=\emptyset and M=C1+C2M^{\prime}=C_{1}+C_{2} is simple normal crossing at QQ.

Proof of Step 9.

Recall that NN^{\prime} is a disjoint union of irreducible divisors and M=fNM^{\prime}=f_{*}N^{\prime} is connected. We can write

N=N1++Ns,N^{\prime}=N^{\prime}_{1}+\cdots{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}+}N^{\prime}_{s},

where each NiN^{\prime}_{i} is an irreducible effective \mathbb{Z}-divisor and Ni1Ni2=N^{\prime}_{i_{1}}\cap N^{\prime}_{i_{2}}=\emptyset for any pair (i1,i2)(i_{1},i_{2}) with i1i2i_{1}\neq i_{2}. Note that ENE\not\subset{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}N^{\prime}}, and hence Mi:=fNi0M^{\prime}_{i}:=f_{*}N^{\prime}_{i}\neq 0 for any ii.

We now show that each MiM^{\prime}_{i} passes through QQ. Suppose the contrary, i.e., NiE=N^{\prime}_{i}\cap E=\emptyset. Then f:STf:S\to T is an isomorphism around NiN^{\prime}_{i}, and hence MiM^{\prime}_{i} is disjoint from jiMj\bigcup_{j\neq i}M^{\prime}_{j}. This is a contradiction.

By multQM=2{\operatorname{mult}}_{Q}M^{\prime}=2, we have s2s\leq 2. We first treat the case when s=1s=1. In this case, NN^{\prime} and MM^{\prime} are irreducible. Hence (I) holds.

We may assume that s=2s=2. By C1C2d(Q)C_{1}\cdot C_{2}\geq d(Q) and C1,TC2,T0C_{1,T}\cdot C_{2,T}\leq 0, we obtain C1C2=d(Q)C_{1}\cdot C_{2}=d(Q) and C1,TC2,T=0C_{1,T}\cdot C_{2,T}=0. In this case, we obtain (II). This completes the proof of Step 9. ∎

Step 10.

(I) does not hold.

Proof of Step 10.

Suppose that (I) holds. By M2>0M^{2}>0, we have that

M2=M2=(2C)2=4C2>C2>0.M^{2}=M^{\prime 2}=(2C)^{2}=4C^{2}>C^{2}>0.

It holds that Bs|C|0{\operatorname{Bs}}\,|C|\leq 0 (Lemma 3.7(1)), and hence the induction hypothesis can be applied for |C||C|, so that |C||C| is base point free. Then also |M|=|2C||M|=|2C| is base point free, which is a contradiction. This completes the proof of Step 10. ∎

Step 11.

(II) does not hold.

Proof of Step 11.

Suppose that (II) holds. By symmetry, we may assume that C12C22C_{1}^{2}\geq C_{2}^{2}. By C1C2=d(Q)>0C_{1}\cdot C_{2}=d(Q)>0, we get

M2=(C1+C2)2=C12+2C1C2+C22=C12+C22+2d(Q)>C12+C22.M^{2}=(C_{1}+C_{2})^{2}=C_{1}^{2}+2C_{1}\cdot C_{2}+C_{2}^{2}=C_{1}^{2}+C_{2}^{2}+2d(Q)>C_{1}^{2}+C_{2}^{2}.

By M2=4d(Q)M^{2}=4d(Q), it holds that

C12+C22=2d(Q)>0.C_{1}^{2}+C_{2}^{2}=2d(Q)>0.

In particular, we get C12>0C_{1}^{2}>0.

If C220C_{2}^{2}\geq 0, we obtain

M2>C12+C22Ci2M^{2}>C_{1}^{2}+C_{2}^{2}\geq C_{i}^{2}

for any i{1,2}i\in\{1,2\}. Hence |C1||C_{1}| and |C2||C_{2}| are base point free by the induction hypothesis, Proposition 3.6, and Lemma 3.7(1). Then also |M|=|C1+C2||M|=|C_{1}+C_{2}| is base point free, which is a contradiction.

We may assume that C22<0C_{2}^{2}<0. In this case, |𝒪S(M)|C2||\mathcal{O}_{S}(M)|_{C_{2}}| is base point free (Lemma 3.8). As C1C_{1} is nef and big, we have the following exact sequence (Theorem 3.4):

H0(S,𝒪S(M))H0(C2,𝒪S(M)|C2)H1(S,𝒪S(C1))=0.H^{0}(S,\mathcal{O}_{S}(M))\to H^{0}(C_{2},\mathcal{O}_{S}(M)|_{C_{2}})\to H^{1}(S,\mathcal{O}_{S}(C_{1}))=0.

Therefore, Bs|M|{\rm Bs}|M| is disjoint from C2C_{2}. This contradicts QBs|M|Q\in{\operatorname{Bs}}\,|M| and QC2Q\in C_{2}. This completes the proof of Step 11. ∎

Step 9, Step 10, and Step 11 complete the proof of Proposition 3.10. ∎

Corollary 3.11.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral regular K3-like surface. Let CC be a prime divisor on SS. If C20C^{2}\geq 0, then |C||C| is base point free.

Proof.

If C2=0C^{2}=0, then apply Proposition 3.6. Otherwise, CC is nef and big. In this case, |C||C| is base point free by Lemma 3.7(1) and Proposition 3.10. ∎

Corollary 3.12.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral regular K3-like surface. Let LL be a Cartier divisor on SS. Then Bs|L|=F{\rm Bs}\,|L|=F for the fixed part FF of LL.

Note that the scheme-theoretic equation Bs|L|=F{\operatorname{Bs}}\,|L|=F holds, i.e., the equation IBs|L|=IFI_{{\operatorname{Bs}}\,|L|}=I_{F} holds for the corresponding ideal sheaves IBs|L|,IFI_{{\operatorname{Bs}}\,|L|},I_{F} on SS.

Proof.

Let MM be the mobile part of |L||L|. If M=0M=0, then there is nothing to show. If M0M\neq 0, then MM is a nef Cartier divisor with dimBs|M|0\dim{\operatorname{Bs}}|M|\leq 0. Then MM is base point free by Proposition 3.6 and Proposition 3.10. ∎

3.3. Nef divisors with base points

Lemma 3.13.

We work over a field κ\kappa. Let XX be a projective variety. For a Cartier divisor LL such that |L||L| is base point free, let φ|L|:Xh0(X,L)1\varphi_{|L|}:X\to\mathbb{P}^{h^{0}(X,L)-1} be the induced morphism. Assume that we have morphisms whose composition is φ|L|\varphi_{|L|}:

φ|L|:X𝜓Y𝜃N\varphi_{|L|}:X\xrightarrow{\psi}Y\xrightarrow{\theta}\mathbb{P}^{N}

with N:=h0(X,L)1N:=h^{0}(X,L)-1 such that YY is a projective variety and ψ:XY\psi:X\to Y is surjective (e.g., YY is the image of φ|L|\varphi_{|L|} or the Stein factorisation of φ|L|\varphi_{|L|}). Then the following induced maps are isomorphisms:

φ|L|:H0(N,𝒪N(1))θ,H0(Y,θ𝒪N(1))ψ,H0(X,L).\varphi_{|L|}^{*}:H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(1))\xrightarrow{\theta^{*},\simeq}H^{0}(Y,\theta^{*}\mathcal{O}_{\mathbb{P}^{N}}(1))\xrightarrow{\psi^{*},\simeq}H^{0}(X,L).
Proof.

By construction, φ|L|=ψθ\varphi_{|L|}^{*}=\psi^{*}\circ\theta^{*} is an isomorphism. In particular, ψ\psi^{*} is surjective. Since ψ:XY\psi:X\to Y is surjective, 𝒪Yψ𝒪X\mathcal{O}_{Y}\to\psi_{*}\mathcal{O}_{X} is injective, and hence ψ\psi^{*} is injective. Therefore, ψ\psi^{*} is an isomorphism. Thus the remaining map θ\theta^{*} is an isomorphism. ∎

Lemma 3.14.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral regular K3-like surface. Let π:SB\pi:S\to B be a morphism to a regular projective curve BB with π𝒪S=𝒪B\pi_{*}\mathcal{O}_{S}=\mathcal{O}_{B}. Then the following hold.

  1. (1)

    BB is smooth and KB-K_{B} is ample. In particular, Bκ1B\simeq\mathbb{P}^{1}_{\kappa} if and only if BB has a κ\kappa-rational point.

  2. (2)

    For a nef Cartier divisor NBN_{B} on BB, it hold that

    dimκH0(S,πNB)=degκNB+1.\dim_{\kappa}H^{0}(S,\pi^{*}N_{B})=\deg_{\kappa}N_{B}+1.
Proof.

We have the injection H1(B,𝒪B)H1(S,𝒪S)=0H^{1}(B,\mathcal{O}_{B})\hookrightarrow H^{1}(S,\mathcal{O}_{S})=0 induced by the corresponding Leray spectral sequence. It follows from degκKB=2h1(S,𝒪S)2=2\deg_{\kappa}K_{B}=2h^{1}(S,\mathcal{O}_{S})-2=-2 that KB-K_{B} is ample. Then BB is a geometrically integral conic on κ2\mathbb{P}^{2}_{\kappa} [Kol13, Lemma 10.6], and hence BB is smooth. Since B×κκ¯κ¯1B\times_{\kappa}\overline{\kappa}\simeq\mathbb{P}^{1}_{\overline{\kappa}}, it holds by Châtelet’s theorem [GS17, Theorem 5.1.3] that Bκ1B\simeq\mathbb{P}^{1}_{\kappa} if and only if BB has a κ\kappa-rational point. Thus (1) holds. By π𝒪S=𝒪B\pi_{*}\mathcal{O}_{S}=\mathcal{O}_{B} and the Riemann–Roch theorem for regular projective curves, we get

dimκH0(S,πNB)=dimκH0(B,NB)=χ(B,NB)=degκNB+1,\dim_{\kappa}H^{0}(S,\pi^{*}N_{B})=\dim_{\kappa}H^{0}(B,N_{B})=\chi(B,N_{B})=\deg_{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}N_{B}+1,

where H1(B,NB)=0H^{1}(B,N_{B})=0 follows from Serre duality. Thus (2) holds. ∎

Lemma 3.15.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral regular K3-like surface. For an effective Cartier divisor FF and nef and big Cartier divisors LL and MM, assume that

L=M+Fandh0(S,L)=h0(S,M).L=M+F\qquad\text{and}\qquad h^{0}(S,L)=h^{0}(S,M).

Then F=0F=0.

Proof.

Suppose F0F\neq 0. Let us derive a contradiction. We can write F=i=1rFiF=\sum_{i=1}^{r}F_{i}, where r1r\geq 1 and each FiF_{i} is a prime divisor (Fi=FjF_{i}=F_{j} might hold even if iji\neq j). Note that h0(S,L)=h0(S,M)h^{0}(S,L)=h^{0}(S,M) implies that the following induced injection is an isomorphism:

H0(S,M)H0(S,M+F)=H0(S,L).H^{0}(S,M)\xrightarrow{\simeq}H^{0}(S,M+F)=H^{0}(S,L).

We have Hi(S,L)=Hi(S,M)=0H^{i}(S,L)=H^{i}(S,M)=0 for i>0i>0 (Theorem 3.4), and hence χ(S,L)=h0(S,L)=h0(S,M)=χ(S,M)\chi(S,L)=h^{0}(S,L)=h^{0}(S,M)=\chi(S,M). By the Riemann–Roch theorem, we conclude L2=M2L^{2}=M^{2}. We then get M2=L2=(M+F)2=M2+2MF+F2M^{2}=L^{2}=(M+F)^{2}=M^{2}+2M\cdot F+F^{2}, which implies

MF+LF=2MF+F2=0.M\cdot F+L\cdot F=2M\cdot F+F^{2}=0.

By MF0M\cdot F\geq 0 and LF0L\cdot F\geq 0, we have that MF=LF=0M\cdot F=L\cdot F=0. Since MM and LL are nef, we get MFi=LFi=0M\cdot F_{i}=L\cdot F_{i}=0 and hence FFi=0F\cdot F_{i}=0 for all 1ir1\leq i\leq r. Therefore, FF is nef with F2=i=1rFFi=0F^{2}=\sum_{i=1}^{r}F\cdot F_{i}=0. Then |F||F| is base point free by Proposition 3.6. This contradicts the fact that FF is contained in the fixed part of |L||L|. ∎

Theorem 3.16.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral regular K3-like surface. Let LL be a nef and big Cartier divisor on SS. Set g:=12L2+1g:=\frac{1}{2}L^{2}+1. Let

L=M+FL=M+F

be the decomposition into the mobile part MM of |L||L| and the fixed part FF of |L||L|. Assume that |L||L| is not base point free. Then the following hold.

  1. (1)

    h0(S,L)=12L2+2=g+1h^{0}(S,L)=\frac{1}{2}L^{2}+2=g+1 and g2g\geq 2.

  2. (2)

    Bs|L|=F{\operatorname{Bs}}\,|L|=F as closed subschemes of SS.

  3. (3)

    |M||M| is base point free and M2=0M^{2}=0.

For B:=φ|M|(S)B:=\varphi_{|M|}(S), we have the induced morphisms:

φ|M|:S𝜋Bg.\varphi_{|M|}:S\xrightarrow{\pi}B\hookrightarrow\mathbb{P}^{g}.
  1. (4)

    Δ(B,𝒪g(1)|B)=0\Delta(B,\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})=0, π𝒪S=𝒪B\pi_{*}\mathcal{O}_{S}=\mathcal{O}_{B}, deg(𝒪g(1)|B)=g\deg(\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})=g, and BB is a smooth projective curve with H1(B,𝒪B)=0H^{1}(B,\mathcal{O}_{B})=0.

  2. (5)

    FF is a prime divisor which is a section of π:SB\pi:S\to B, i.e., π|F:FB\pi|_{F}:F\xrightarrow{\simeq}B. Moreover, H0(F,𝒪F)=κH^{0}(F,\mathcal{O}_{F})=\kappa, F2=2F^{2}=-2, and LF=g2L\cdot F=g-2.

  3. (6)

    If LF=0L\cdot F=0, then L2=2L^{2}=2 and g=2g=2.

  4. (7)

    If LF>0L\cdot F>0, then LC>0L\cdot C>0 for any curve CSC{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\subset}S with CFC\cap F\neq\emptyset.

Furthermore, if SS has a κ\kappa-rational point, then FBκ1F\simeq B\simeq\mathbb{P}^{1}_{\kappa}.

Proof.

Taking the base change to the separable closure, we may assume that κ\kappa is separably closed. In particular, SS has a κ\kappa-rational point and it suffices to show (1)–(7).

The assertions (1) and (2) hold by Theorem 3.4 and Corollary 3.12, respectively. By F0F\neq 0, we get M2=0M^{2}=0 (Lemma 3.15). Since MM is nef, |M||M| is base point free (Proposition 3.6). Thus (3) holds.

Let us show (4). Let

π:SBB\pi:S\to B^{\prime}\to B

be the Stein factorisation of π:SB\pi:S\to B. By Remark 2.17, we have

0Δ(B,𝒪g(1)|B)=dimB+deg(𝒪g(1)|B)h0(B,𝒪g(1)|B).0\leq\Delta(B,\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})=\dim B+\deg(\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})-h^{0}(B,\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B}).

For :=deg(BB)\ell:=\deg(B^{\prime}\to B), it follows from Bκ1B^{\prime}\simeq\mathbb{P}^{1}_{\kappa} and

H0(S,L)H0(S,M)H0(B,𝒪g(1)|B)H0(B,𝒪g(1)|B)(Lemma 3.13)H^{0}(S,L)\simeq H^{0}(S,M)\simeq H^{0}(B^{\prime},\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B^{\prime}})\simeq H^{0}(B,\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})\qquad(\text{Lemma \ref{l-H^0-image}})

that

  • h0(B,𝒪g(1)|B)=g+1h^{0}(B,\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})=g+1 and

  • deg(𝒪g(1)|B)=1deg(𝒪g(1)|B)=1(h0(B,𝒪g(1)|B)1)=g\deg(\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})=\frac{1}{\ell}\deg(\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B^{\prime}})=\frac{1}{\ell}(h^{0}(B^{\prime},\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B^{\prime}})-1)=\frac{g}{\ell}.

Hence we obtain

0Δ(B,𝒪g(1)|B)=1+g(g+1)=g(1).0\leq\Delta(B,\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}1+\frac{g}{\ell}-(g+1)=\frac{g}{\ell}(1-\ell).}

Therefore, Δ(B,𝒪g(1)|B)=0\Delta(B,\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})=0 and =1\ell=1. Since BB has a κ\kappa-rational point, we get Bκ1B\simeq\mathbb{P}^{1}_{\kappa}. Then BBB^{\prime}\to B is a finite birational morphism to a normal variety, and hence an isomorphism. Hence deg(𝒪g(1)|B)=g\deg(\mathcal{O}_{\mathbb{P}^{g}}(1)|_{B})=g. Thus (4) holds.

Let us show (5). We first prove that

  1. (5a)

    FF is a prime divisor, and

  2. (5b)

    π(F)=B\pi(F)=B.

Since L=M+FL=M+F is big, there exists a prime divisor CSuppFC\subset{\operatorname{Supp}}\,F such that π(C)=B\pi(C)=B. We can write F=C+FF=C+F^{\prime} for some effective \mathbb{Z}-divisor FF^{\prime}. Suppose F0F^{\prime}\neq 0. It suffices to derive a contradiction. By Lemma 3.15, it is enough to prove that M+CM+C is nef and big. For a κ\kappa-rational point QB=κ1Q\in B=\mathbb{P}^{1}_{\kappa} and κC:=H0(C,𝒪C)\kappa_{C}:=H^{0}(C,\mathcal{O}_{C}), we have MgπQM\sim g\pi^{*}Q and

degκC(M+C)|CdegκCg(πQ|C)+(2)g20.\deg_{\kappa_{C}}(M+C)|_{C}\geq\deg_{\kappa_{C}}g(\pi^{*}Q|_{C})+(-2)\geq g-2\geq 0.

Hence M+CM+C is nef. Since also MM is nef, M+aCM+aC is nef for every 0a10\leq a\leq 1. For some 0<ϵ10<\epsilon\ll 1, it follows from M2=0M^{2}=0 and MC>0M\cdot C>0 that

(M+ϵC)2=M2+2ϵMC+ϵ2C2=ϵ(2MC+ϵC2)>0.(M+\epsilon C)^{2}=M^{2}+2\epsilon M\cdot C+\epsilon^{2}C^{2}=\epsilon(2M\cdot C+\epsilon C^{2})>0.

Hence M+ϵCM+\epsilon C is big. Therefore, M+CM+C is nef and big. Thus (5a) and (5b) hold.

Set δ:=deg(π|F:FB)=[K(F):K(B)]\delta:=\deg(\pi|_{F}:F\to B)=[K(F):K(B)], κF:=H0(F,𝒪F)\kappa_{F}:=H^{0}(F,\mathcal{O}_{F}), and dF:=[κF:κ]d_{F}:=[\kappa_{F}:\kappa]. By the following commutative diagram

Fκ1=BSpecκFSpecκ,\begin{CD}F@>{}>{}>\mathbb{P}^{1}_{\kappa}=B\\ @V{}V{}V@V{}V{}V\\ {\operatorname{Spec}}\,\kappa_{F}@>{}>{}>{\operatorname{Spec}}\,\kappa,\end{CD}

we get the factorisation:

π|F:F𝛼κF1𝛽κ1=B.\pi|_{F}:F\xrightarrow{\alpha}\mathbb{P}^{1}_{\kappa_{F}}\xrightarrow{\beta}\mathbb{P}^{1}_{\kappa}=B.

We then obtain

δ=deg(π|F:Fκ1)=deg(α)deg(β)=deg(α)[κF:κ]dF>0.\delta=\deg(\pi|_{F}:F\to\mathbb{P}^{1}_{\kappa})=\deg(\alpha)\cdot\deg(\beta)=\deg(\alpha)[\kappa_{F}:\kappa]\in d_{F}\mathbb{Z}_{>0}.

It holds that

2g2=L2=(M+F)2=M2+2MF+F2=0+2δg+F22dFg2dF,2g-2=L^{2}=(M+F)^{2}=M^{2}+2M\cdot F+F^{2}=0+2\delta g+F^{2}\geq 2d_{F}g-2d_{F},

where the last inequality follows from δdF>0\delta\in d_{F}\mathbb{Z}_{>0} and

F2=(KS+F)F=degκωF=2dimκH1(F,𝒪F)2dimκH0(F,𝒪F)2dF.F^{2}=(K_{S}+F)\cdot F=\deg_{\kappa}\omega_{F}=2\dim_{\kappa}H^{1}(F,\mathcal{O}_{F})-2\dim_{\kappa}H^{0}(F,\mathcal{O}_{F})\geq-2d_{F}.

We then obtain dF=1d_{F}=1 and the above inequalities are equalities: δ=1\delta=1 and F2=2F^{2}=-2. Since π|F:FB=κ1\pi|_{F}:F\to B=\mathbb{P}^{1}_{\kappa} is birational, this is an isomorphism. Finally, we get LF=(M+F)F=g2L\cdot F=(M+F)\cdot F=g-2. Thus (5) holds.

Let us show (6). By g2=LF=0g-2=L\cdot F=0, we have g=2g=2. Hence L2=2g2=2L^{2}=2g-2=2. Thus (6) holds.

Let us show (7). Assume LF>0L\cdot F>0. Fix a curve CC on SS with CFC\cap F\neq\emptyset. It suffices to show LC>0L\cdot C>0. If C=FC=F, then LC=LF>0L\cdot C=L\cdot F>0 by our assumption. If CFC\neq F, then we get FC>0F\cdot C>0, and hence

LC=(M+F)C=MC+FCFC>0.L\cdot C=(M+F)\cdot C=M\cdot C+F\cdot C\geq F\cdot C>0.

Thus (7) holds. ∎

Theorem 3.17.

We work over a field κ\kappa of characteristic p>0p>0. Let SS be a geometrically integral canonical K3-like surface. Let LL be an ample Cartier divisor on SS such that |L||L| is not base point free. Then the following hold.

  1. (1)

    Bs|L|{\operatorname{Bs}}\,|L| is irreducible.

  2. (2)

    dimBs|L|=0\dim{\operatorname{Bs}}\,|L|=0 or dimBs|L|=1\dim{\operatorname{Bs}}\,|L|=1.

  3. (3)

    If dimBs|L|=1\dim{\operatorname{Bs}}\,|L|=1, then SS is smooth around Bs|L|{\operatorname{Bs}}\,|L|.

  4. (4)

    If dimBs|L|=0\dim{\operatorname{Bs}}\,|L|=0, then Bs|L|{\operatorname{Bs}}\,|L| is scheme-theoretically equal to a reduced point.

Proof.

Let μ:TS\mu:T\to S be the minimal resolution of SS. We then have the following equation of sets:

μ1(Bs|L|)=Bs|μL|=F,\mu^{-1}({\operatorname{Bs}}\,|L|)={\operatorname{Bs}}\,|\mu^{*}L|=F,

where FF is a prime divisor on TT (Theorem 3.16). By the following equation of sets

Bs|L|=μ(μ1(Bs|L|))=μ(F),{\operatorname{Bs}}\,|L|=\mu(\mu^{-1}({\operatorname{Bs}}\,|L|))=\mu(F),

Bs|L|{\operatorname{Bs}}\,|L| is an irreducible closed subset whose dimension is zero or one. Thus (1) and (2) hold.

Let us show (3). Assume dimBs|L|=1\dim{\operatorname{Bs}}\,|L|=1. In this case, we obtain μLF>0\mu^{*}L\cdot F>0. It follows from Theorem 3.16(7) that μLC>0\mu^{*}L\cdot C>0 for any curve CC on TT with CFC\cap F\neq\emptyset. In other words, Ex(μ)F={\operatorname{Ex}}(\mu)\cap F=\emptyset, i.e., μ:TS\mu:T\to S is isomorphic around FF. Since μ(F)\mu(F) is a smooth Cartier divisor on SS (Theorem 3.16), SS is smooth around μ(F)\mu(F). Thus (3) holds.

Let us show (4). Taking the base change to the separable closure of κ\kappa, we may assume that κ\kappa is separably closed. Assume dimBs|L|=0\dim{\operatorname{Bs}}\,|L|=0. In this case, P:=μ(F)=(Bs|L|)redP:=\mu(F)=({\operatorname{Bs}}\,|L|)_{{\operatorname{red}}} is one point by (1). Hence we get μLF=0\mu^{*}L\cdot F=0. By Theorem 3.16, we have L2=2L^{2}=2, g:=12L2+1=2g:=\frac{1}{2}L^{2}+1=2, μL=M+FE1+E2+F\mu^{*}L=M+F\sim E_{1}+E_{2}+F, and

φ|M|:T𝜋B=κ1κ2,\varphi_{|M|}:T\xrightarrow{\pi}B=\mathbb{P}^{1}_{\kappa}\hookrightarrow\mathbb{P}^{2}_{\kappa},

where FF is a section of π\pi and E1E_{1} and E2E_{2} are fibres of π\pi over general κ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\kappa}-rational points of BB. Set E1,S:=μE1E_{1,S}:=\mu_{*}E_{1} and E2,S:=μE2E_{2,S}:=\mu_{*}E_{2}, so that we get LE1,S+E2,SL\sim E_{1,S}+E_{2,S}. We have the induced morphisms μ|Ei:EiEi,S\mu|_{E_{i}}:E_{i}\to E_{i,S}. The remaining proof is divided into the following five steps.

  1. (i)

    For each i{1,2}i\in\{1,2\}, it holds that (μ|Ei)𝒪Ei=𝒪Ei,S(\mu|_{E_{i}})_{*}\mathcal{O}_{E_{i}}=\mathcal{O}_{E_{i,S}}, i.e., μ|Ei:EiEi,S\mu|_{E_{i}}:E_{i}\to E_{i,S} is an isomorphism.

  2. (ii)

    Each Ei,SE_{i,S} is a projective Gorenstein curve such that Ei,SE_{i,S} is regular around PP and dimκH0(Ei,S,𝒪Ei,S)=dimκH1(Ei,S,𝒪Ei,S)=1\dim_{\kappa}H^{0}(E_{i,S},\mathcal{O}_{E_{i,S}})=\dim_{\kappa}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}H^{1}(E_{i,S},\mathcal{O}_{E_{i,S}})}=1.

  3. (iii)

    The scheme-theoretic equation E1,SE2,S=PE_{1,S}\cap E_{2,S}=P holds.

  4. (iv)

    The scheme-theoretic equation Bs|L|E1,S+E2,S|Ei,S=Bs|L|Ei,S|=P{\operatorname{Bs}}\,|L|_{E_{1,S}+E_{2,S}}|\cap E_{i,S}={\operatorname{Bs}}\,|L|_{E_{i,S}}|=P holds for each i{1,2}i\in\{1,2\}.

  5. (v)

    The scheme-theoretic equation Bs|L|=P{\operatorname{Bs}}\,|L|=P holds.

Let us show (i). Fix i{1,2}i\in\{1,2\}. It suffices to show R1μ𝒪T(Ei)=0R^{1}\mu_{*}\mathcal{O}_{T}(-E_{i})=0. This follows from the relative Kawamata–Viehweg vanishing theorem [Kol13, Theorem 10.4], which can be applied by

(Ei(KT+23F))F=1+0+43>0.\left(-E_{i}-\left(K_{T}+\frac{2}{3}F\right)\right)\cdot F=-1+0+\frac{4}{3}>0.

Thus (i) holds.

Let us show (ii). Fix i{1,2}i\in\{1,2\}. Note that fibres over general κ\kappa-rational points of B=κ1B=\mathbb{P}^{1}_{\kappa} are irreducible. By EiF=1E_{i}\cdot F=1, EiE_{i} is a prime divisor and EiFE_{i}\cap F is a κ\kappa-rational point, which implies that H0(Ei,𝒪Ei)=κH^{0}(E_{i},\mathcal{O}_{E_{i}})=\kappa. By the adjunction formula ωEi𝒪T(KT+Ei)|Ei𝒪Ei\omega_{E_{i}}\simeq\mathcal{O}_{T}(K_{T}+E_{i})|_{E_{i}}\simeq\mathcal{O}_{E_{i}}, we obtain H1(Ei,𝒪Ei)κH^{1}(E_{i},\mathcal{O}_{E_{i}})\simeq\kappa by Serre duality. It follows from EiF=1E_{i}\cdot F=1 that EiE_{i} is regular around EiFE_{i}\cap F, and hence Ei,SE_{i,S} is regular around PP by (i). Thus (ii) holds.

Let us show (iii). By 2=L2=L(E1,S+E2,S)2=L^{2}=L\cdot(E_{1,S}+E_{2,S}) and the ampleness of LL, it holds that LE1,S=LE2,S=1L\cdot E_{1,S}=L\cdot E_{2,S}=1. We have L2E2,SL\sim 2E_{2,S} and E1,S(2E2,S)=1E_{1,S}\cdot(2E_{2,S})=1, which implies the scheme-theoretic equation E1,S(2E2,S)=PE_{1,S}\cap(2E_{2,S})=P. Therefore, we obtain the following scheme-theoretic inclusions:

{P}E1,SE2,SE1,S(2E2,S)={P},\{P\}\subset E_{1,S}\cap E_{2,S}\subset E_{1,S}\cap(2E_{2,S})=\{P\},

and hence E1,SE2,S=PE_{1,S}\cap E_{2,S}=P. Thus (iii) holds.

Let us show (iv). By (iii), we have the following exact sequence:

0H0(E1,S+E2,S,L|E1,S+E2,S)H0(E1,S,L|E1,S)H0(E2,S,L|E2,S)H0(P,L|P).0\to H^{0}(E_{1,S}+E_{2,S},L|_{E_{1,S}+E_{2,S}})\to H^{0}(E_{1,S},L|_{E_{1,S}})\oplus H^{0}(E_{2,S},L|_{E_{2,S}})\to H^{0}(P,L|_{P}).

Note that L|E1,S𝒪T(2E2,S)|E1,S𝒪E1,S(P)L|_{E_{1,S}}\simeq\mathcal{O}_{T}(2E_{2,S})|_{E_{1,S}}\simeq\mathcal{O}_{E_{1,S}}(P) by LE1,S=1L\cdot E_{1,S}=1. By the Riemann–Roch theorem, we obtain h0(E1,S,𝒪E1,S(P))=1h^{0}(E_{1,S},\mathcal{O}_{E_{1,S}}(P))=1 (note that E1,SE_{1,S} is regular around PP by (ii)). Hence we get the isomorphism:

H0(E1,S,𝒪E1,S)H0(E1,S,𝒪E1,S(P)),H^{0}(E_{1,S},\mathcal{O}_{E_{1,S}})\xrightarrow{\simeq}H^{0}(E_{1,S},\mathcal{O}_{E_{1,S}}(P)),

which implies that H0(E1,S,L|E1,S)H0(P,L|P)H^{0}(E_{1,S},L|_{E_{1,S}})\to H^{0}(P,L|_{P}) is zero. Similarly, H0(E2,S,L|E2,S)H0(P,L|P)H^{0}(E_{2,S},L|_{E_{2,S}})\to H^{0}(P,L|_{P}) is zero. Therefore, Bs|L|Ei,S|=P{\operatorname{Bs}}\,|L|_{E_{i,S}}|=P for each i{1,2}i\in\{1,2\} and we get the induced isomorphism:

H0(E1,S+E2,S,L)H0(E1,S,L)H0(E2,S,L).H^{0}(E_{1,S}+E_{2,S},L)\xrightarrow{\simeq}H^{0}(E_{1,S},L)\oplus H^{0}(E_{2,S},L).

Then each projection H0(E1,S+E2,S,L|E1,S+E2,S)H0(Ei,S,L|Ei,S)H^{0}(E_{1,S}+E_{2,S},L|_{E_{1,S}+E_{2,S}})\to H^{0}(E_{i,S},L|_{E_{i,S}}) is surjective. Therefore, (iv) holds.

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

H0(S,L)H0(E1,S+E2,S,L|E1,S+E2,S)H1(S,𝒪S)=0,H^{0}(S,L)\to H^{0}(E_{1,S}+E_{2,S},L|_{E_{1,S}+E_{2,S}})\to H^{1}(S,\mathcal{O}_{S})=0,

and hence

Bs|L|=Bs|L|E1,S+E2,S|.{\operatorname{Bs}}\,|L|={\operatorname{Bs}}\,|L|_{E_{1,S}+E_{2,S}}|.

For the ideals 𝔪P𝒪E1,S+E2,S\mathfrak{m}_{P}\subset\mathcal{O}_{E_{1,S}+E_{2,S}} and IBs|L|E1,S+E2,S|𝒪E1,S+E2,SI_{{\operatorname{Bs}}|L|_{E_{1,S}+E_{2,S}}|}\subset\mathcal{O}_{E_{1,S}+E_{2,S}} of PP and Bs|L|E1,S+E2,S|{\operatorname{Bs}}|L|_{E_{1,S}+E_{2,S}}| respectively, it suffices to show that 𝔪P=IBs|L|E1,S+E2,S|\mathfrak{m}_{P}=I_{{\operatorname{Bs}}|L|_{E_{1,S}+E_{2,S}}|}. By the induced injection

ρ:=ρ1ρ2:𝒪E1,S+E2,S𝒪E1,S𝒪E2,Swhereρi:𝒪E1,S+E2,S𝒪Ei,S,\rho:=\rho_{1}\oplus\rho_{2}:\mathcal{O}_{E_{1,S}+E_{2,S}}\hookrightarrow\mathcal{O}_{E_{1,S}}\oplus\mathcal{O}_{E_{2,S}}\quad\text{where}\quad\rho_{i}:\mathcal{O}_{E_{1,S}+E_{2,S}}\to\mathcal{O}_{E_{i,S}},

it is enough to check ρ(𝔪P)=ρ(IBs|L|E1,S+E2,S|)\rho(\mathfrak{m}_{P})=\rho(I_{{\operatorname{Bs}}|L|_{E_{1,S}+E_{2,S}}|}), which is equivalent to ρ1(𝔪P)=ρ1(IBs|L|E1,S+E2,S|)\rho_{1}(\mathfrak{m}_{P})=\rho_{1}(I_{{\operatorname{Bs}}|L|_{E_{1,S}+E_{2,S}}|}) and ρ2(𝔪P)=ρ2(IBs|L|E1,S+E2,S|)\rho_{2}(\mathfrak{m}_{P})=\rho_{2}(I_{{\operatorname{Bs}}|L|_{E_{1,S}+E_{2,S}}|}). These follow from (iv). Thus (v) holds. ∎

Corollary 3.18.

We work over an algebraically closed field kk of characteristic p>0p>0. Let XX be a Fano threefold. Assume that the generic member SS of |KX||-K_{X}| is a geometrically integral canonical K3-like surface. Then SS is regular.

Proof.

Set κ:=K((H0(X,KX)))\kappa:=K(\mathbb{P}(H^{0}(X,-K_{X}))), Xκ:=X×kκX_{\kappa}:=X\times_{k}\kappa, and L:=(KXκ)|SL:=(-K_{X_{\kappa}})|_{S}, which is an ample Cartier divisor on SS. Note that SS is a prime divisor on XκX_{\kappa} and SS is regular outside Bs|KXκ|{\operatorname{Bs}}\,|-K_{X_{\kappa}}| (Theorem 2.9). It holds that Bs|KXκ|=Bs|L|{\operatorname{Bs}}\,|-K_{X_{\kappa}}|={\operatorname{Bs}}\,|L|, because we have an exact sequence

H0(Xκ,𝒪Xκ(KXκ))H0(S,L)H1(Xκ,𝒪Xκ)H^{0}(X_{\kappa},\mathcal{O}_{X_{\kappa}}(-K_{X_{\kappa}}))\to H^{0}(S,L)\to H^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}})

and H1(Xκ,𝒪Xκ)H1(X,𝒪X)kκ=0H^{1}(X_{\kappa},\mathcal{O}_{X_{\kappa}})\simeq H^{1}(X,\mathcal{O}_{X})\otimes_{k}\kappa=0 (Theorem 2.4). By Theorem 3.17(2), there are the following two cases: dimBs|L|=1\dim{\operatorname{Bs}}\,|L|=1 and dimBs|L|=0\dim{\operatorname{Bs}}\,|L|=0. If dimBs|L|=1\dim{\operatorname{Bs}}\,|L|=1, then SS is smooth around Bs|L|{\operatorname{Bs}}\,|L| (Theorem 3.17(3)), and hence SS is regular. We may assume that dimBs|L|=0\dim{\operatorname{Bs}}\,|L|=0. Then Bs|KXκ|=Bs|L|{\operatorname{Bs}}\,|-K_{X_{\kappa}}|={\operatorname{Bs}}\,|L| is a reduced point PP (Theorem 3.17(4)). By Bs|KX|×kκBs|KXκ|{\operatorname{Bs}}\,|-K_{X}|\times_{k}\kappa\simeq{\operatorname{Bs}}\,|-K_{X_{\kappa}}|, also Bs|KX|{\operatorname{Bs}}\,|-K_{X}| is a reduced point QQ, which is nothing but the image of PP. Therefore, general members of |KX||-K_{X}| are smooth at QQ by Lemma 3.19 below. Hence SS is smooth around Bs|L|{\operatorname{Bs}}\,|L|, as required. ∎

Lemma 3.19.

We work over an algebraically closed field kk. Let XX be a smooth projective variety and fix a closed point QXQ\in X. Let LL be a Cartier divisor such that the scheme-theoretic equation

Bs|L|=QW{\operatorname{Bs}}\,|L|=Q\amalg W

holds for some closed subscheme WW of XX with QWQ\not\in W. Then general members of |L||L| are smooth at QQ.

Proof.

Let 𝔪Q𝒪X,Q\mathfrak{m}_{Q}\subset\mathcal{O}_{X,Q} be the maximal ideal. By Bs|L|=QW{\operatorname{Bs}}\,|L|=Q\amalg W, there exists an effective divisor D|L|D\in|L| such that fD𝔪Qf_{D}\in\mathfrak{m}_{Q} and fD𝔪Q2f_{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}D}\not\in\mathfrak{m}_{Q}^{2}, where fD𝒪X,Qf_{D}\in\mathcal{O}_{X,Q} is an defining element of DD which is uniquely determined up to 𝒪X,Q×\mathcal{O}_{X,Q}^{\times}. Hence DD is smooth at QQ. Therefore, also general members of |L||L| are smooth at QQ. ∎

4. Generic elephants

4.1. The case dimImφ|KX|=1\dim{\operatorname{Im}}\,\varphi_{|-K_{X}|}=1

Proposition 4.1.

We work over an algebraically closed field kk of characteristic p>0p>0. Let XX be a Fano threefold. Assume that dim(Imφ|KX|)=1\dim({\operatorname{Im}}\,\varphi_{|-K_{X}|})=1. Then the following hold.

  1. (1)

    The mobile part |D||D| of |KX||-K_{X}| is base point free.

  2. (2)

    For the image Y:=Imφ|D|Y:={\operatorname{Im}}\,\varphi_{|D|} and the morphism ψ:XY\psi:X\to Y induced by φ|D|\varphi_{|D|}, it holds that Y1Y\simeq\mathbb{P}^{1} and ψ𝒪X=𝒪Y\psi_{*}\mathcal{O}_{X}=\mathcal{O}_{Y}.

  3. (3)

    If |KX||-K_{X}| is not base point free, then KX2E=1-K_{X}^{2}\cdot E=1 for a fibre EE of ψ\psi.

Proof.

Take the decomposition

|KX|=D0+|D||-K_{X}|=D_{0}+|D|

into the fixed part D0D_{0} of |KX||-K_{X}| and the mobile part DD of |KX||-K_{X}|. Set

g:=12(KX)3+1=χ(X,KX)2andg:=h0(X,KX)2.g:=\frac{1}{2}(-K_{X})^{3}+1=\chi(X,-K_{X})-2\qquad\text{and}\qquad g^{\prime}:=h^{0}(X,-K_{X})-2.

By H2(X,KX)=H3(X,KX)=0H^{2}(X,-K_{X})=H^{3}(X,-K_{X})=0 (Theorem 2.4), we have

2g=χ(X,KX)2=h0(X,KX)h1(X,KX)2h0(X,KX)2=g.2\leq g=\chi(X,-K_{X})-2=h^{0}(X,-K_{X})-h^{1}(X,-K_{X})-2\leq h^{0}(X,-K_{X})-2=g^{\prime}.

We then have φ:=φ|KX|:Xg+1\varphi:=\varphi_{|-K_{X}|}:X\dashrightarrow\mathbb{P}^{g^{\prime}+1}. Let σ:X~X\sigma:\widetilde{X}\to X be a desingularisation of the resolution of the indeterminacies φ|KX|\varphi_{|-K_{X}|}, so that we have φ~:X~g+1\widetilde{\varphi}:\widetilde{X}\to\mathbb{P}^{g^{\prime}+1}. For the largest open subset XX^{\prime} of XX on which φ|KX|\varphi_{|-K_{X}|} is defined, it holds that dim(XX)1\dim(X\setminus X^{\prime})\leq 1 and X~:=σ1(X)X\widetilde{X}^{\prime}:=\sigma^{-1}(X^{\prime})\to X^{\prime} can be assumed to be an isomorphism. Take the decomposition

σD=D~+F\sigma^{*}D=\widetilde{D}+F

into the movile part D~\widetilde{D} and the fixed part FF, where |D~||\widetilde{D}| is base point free. We then get

(4.1.1) H0(X~,D~)H0(X~,σD)H0(X,D)H0(X,KX)H^{0}(\widetilde{X},\widetilde{D})\simeq H^{0}(\widetilde{X},\sigma^{*}D)\simeq H^{0}(X,D)\simeq H^{0}(X,-K_{X})

and φ|D~|:X~g+1\varphi_{|\widetilde{D}|}:\widetilde{X}\to\mathbb{P}^{g^{\prime}+1} with 𝒪X~(D~)φ|D~|𝒪g+1(1)\mathcal{O}_{\widetilde{X}}(\widetilde{D})\simeq\varphi_{|\widetilde{D}|}^{*}\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1). Set Y:=φ(X)¯=φ~(X~)Y:=\overline{\varphi(X^{\prime})}=\widetilde{\varphi}(\widetilde{X}). Let

φ~:X~𝜓Yg+1\widetilde{\varphi}:\widetilde{X}\xrightarrow{\psi}Y\hookrightarrow\mathbb{P}^{g^{\prime}+1}

be the induced morphisms.

We now show that

  1. (a)

    Δ(Y,𝒪g+1(1)|Y)=0\Delta(Y,\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Y})=0,

  2. (b)

    Y1Y\simeq\mathbb{P}^{1},

  3. (c)

    deg(𝒪g+1(1)|Y)=g+1\deg(\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Y})=g^{\prime}+1, and

  4. (d)

    ψ𝒪X~=𝒪Y\psi_{*}\mathcal{O}_{\widetilde{X}}=\mathcal{O}_{Y}.

Let ψ:X~ZY\psi:\widetilde{X}\to Z\to Y be the Stein factorisation of ψ\psi. By

H1(Z,𝒪Z)H1(X~,𝒪X~)H1(X,𝒪X)=0,H^{1}(Z,\mathcal{O}_{Z})\hookrightarrow H^{1}(\widetilde{X},\mathcal{O}_{\widetilde{X}})\simeq H^{1}(X,\mathcal{O}_{X})=0,

we obtain H1(Z,𝒪Z)=0H^{1}(Z,\mathcal{O}_{Z})=0, which in turn implies Z1Z\simeq\mathbb{P}^{1}. In particular, we get

(4.1.2) deg(𝒪g+1(1)|Z)=h0(Z,𝒪g+1(1)|Z)1=h0(X~,D~)1=g+1,\deg(\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Z})=h^{0}(Z,\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Z})-1=h^{0}(\widetilde{X},\widetilde{D})-1=g^{\prime}+1,

where the first equality holds by the Riemann–Roch theorem, the second one follows from Lemma 3.13, and the last one holds by (4.1.1). Set :=deg(ZY)>0\ell:=\deg(Z\to Y)\in\mathbb{Z}_{>0}. It holds that

0\displaystyle 0 Δ(Y,𝒪g+1(1)|Y)\displaystyle\leq\Delta(Y,\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Y})
=dimY+deg(𝒪g+1(1)|Y)h0(Y,𝒪g+1(1)|Y)\displaystyle=\dim Y+\deg(\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Y})-h^{0}(Y,\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Y})
=1+deg(𝒪g+1(1)|Z)(g+2)\displaystyle=1+\frac{\deg(\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Z})}{\ell}-(g^{\prime}+2)
=(g+1)(1),\displaystyle=\frac{(g^{\prime}+1)(1-\ell)}{\ell},

where the second equality follows from Lemma 3.13 and (4.1.1), and the third one holds by (4.1.2). Then =1\ell=1 and Δ(Y,𝒪g+1(1)|Y)=0\Delta(Y,\mathcal{O}_{\mathbb{P}^{g^{\prime}+1}}(1)|_{Y})=0, which deduces that ZYZ\to Y is birational and Y1Y\simeq\mathbb{P}^{1}. Thus (a) and (b) hold. Since YY is normal and ZYZ\to Y is a finite birational morphism, ZYZ\to Y is an isomorphism. Thus (d) holds. Finally, (c) follows from (4.1.2). This completes the proof of (a)–(d).

Fix a closed point PYP\in Y and set E~:=ψP\widetilde{E}:=\psi^{*}P and E:=σE~E:=\sigma_{*}\widetilde{E}. We then obtain D~(g+1)E~\widetilde{D}\sim(g^{\prime}+1)\widetilde{E} and

D=σσD=σ(D~+F)=σD~(g+1)E.D=\sigma_{*}\sigma^{*}D=\sigma_{*}(\widetilde{D}+F)=\sigma_{*}\widetilde{D}\sim(g^{\prime}+1)E.

By KXD0+DD0+(g+1)E-K_{X}\sim D_{0}+D\sim D_{0}+(g^{\prime}+1)E, we obtain

2g2\displaystyle 2g-2 =\displaystyle= (KX)3\displaystyle(-K_{X})^{3}
=\displaystyle= (D0+(g+1)E)(KX)2\displaystyle(D_{0}+(g^{\prime}+1)E)\cdot(-K_{X})^{2}
=\displaystyle= D0(KX)2+(g+1)E(KX)2\displaystyle D_{0}\cdot(-K_{X})^{2}+(g^{\prime}+1)E\cdot(-K_{X})^{2}
=\displaystyle= D0(KX)2+(g+1)E(D0+(g+1)E)(KX)\displaystyle D_{0}\cdot(-K_{X})^{2}+(g^{\prime}+1)E\cdot(D_{0}+(g^{\prime}+1)E)\cdot(-K_{X})
=\displaystyle= D0(KX)2+(g+1)ED0(KX)+(g+1)2E2(KX).\displaystyle D_{0}\cdot(-K_{X})^{2}+(g^{\prime}+1)E\cdot D_{0}\cdot(-K_{X})+(g^{\prime}+1)^{2}E^{2}\cdot(-K_{X}).

Since D0D_{0} is effective, KX-K_{X} is ample, and |E||E| is base point free outside a closed subset of codimension two, it holds that

D0(KX)20,ED0(KX)0,E2(KX)0.D_{0}\cdot(-K_{X})^{2}\geq 0,\qquad E\cdot D_{0}\cdot(-K_{X})\geq 0,\qquad E^{2}\cdot(-K_{X})\geq 0.

We now show that E2(KX)=0E^{2}\cdot(-K_{X})=0. Suppose the contrary, i.e., E2(KX)>0E^{2}\cdot(-K_{X})>0. The above equation deduces

2g>2g2(g+1)2E2(KX)(g+1)22g,2g>2g-2\geq(g^{\prime}+1)^{2}E^{2}\cdot(-K_{X})\geq(g^{\prime}+1)^{2}\geq 2g^{\prime},

which contradicts ggg^{\prime}\geq g. Hence, we have E2(KX)=0E^{2}\cdot(-K_{X})=0.

Pick two general members E~1,E~2|E~|\widetilde{E}_{1},\widetilde{E}_{2}\in|\widetilde{E}|, which are distinct prime divisors [Bad01, Corollary 7.3]. Set Ei:=σE~iE_{i}:=\sigma_{*}\widetilde{E}_{i}. Then E1E_{1} and E2E_{2} are distinct prime divisors. We get

DE1E2.D\sim E_{1}\sim E_{2}.

Therefore, we obtain

E1E2(KX)=E2(KX)=0,E_{1}\cdot E_{2}\cdot(-K_{X})=E^{2}\cdot(-K_{X})=0,

which implies that E1E2=E_{1}\cap E_{2}=\emptyset, i.e., |D||D| is base point free. In particular, σ:X~X\sigma:\widetilde{X}\xrightarrow{\simeq}X. Hence (1) and (2) holds.

Let us show (3). Assume that |KX||-K_{X}| is not base point free, which is equivalent to D00D_{0}\neq 0. Since KX-K_{X} is ample, we get D0(KX)2>0D_{0}\cdot(-K_{X})^{2}>0 and E(KX)2>0E\cdot(-K_{X})^{2}>0. The latter one implies

ED0(KX)=E(D0+(g+1)E)(KX)=E(KX)2>0.E\cdot D_{0}\cdot(-K_{X})=E\cdot(D_{0}+(g^{\prime}+1)E)\cdot(-K_{X})=E\cdot(-K_{X})^{2}>0.

We have

2g2=D0(KX)2+(g+1)ED0(KX).2g-2=D_{0}\cdot(-K_{X})^{2}+(g^{\prime}+1)E\cdot D_{0}\cdot(-K_{X}).

By ggg^{\prime}\geq g, we get ED0(KX)=1E\cdot D_{0}\cdot(-K_{X})=1, and hence E(KX)2=1E\cdot(-K_{X})^{2}=1. Thus (3) holds. ∎

Corollary 4.2.

We work over an algebraically closed field kk of characteristic p>0p>0. Let XX be a Fano threefold with ρ(X)=1\rho(X)=1. Then dim(Imφ|KX|)2\dim({\operatorname{Im}}\,\varphi_{|-K_{X}|})\geq 2.

Proof.

Suppose that dim(Imφ|KX|)1\dim({\operatorname{Im}}\,\varphi_{|-K_{X}|})\leq 1. By h0(X,KX)χ(X,KX)=(KX)32+32h^{0}(X,-K_{X})\geq\chi(X,-K_{X})=\frac{(-K_{X})^{3}}{2}+3\geq 2, we get dim(Imφ|KX|)=1\dim({\operatorname{Im}}\,\varphi_{|-K_{X}|})=1. It follows from Proposition 4.1 that there is a surjective morphism XYX\to{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}Y} to a curve Y{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}Y}. This contradicts ρ(X)=1\rho(X)=1. ∎

4.2. The case dimImφ|KX|2\dim{\operatorname{Im}}\,\varphi_{|-K_{X}|}\geq 2

4.3Mumford pullback.

We work over a field κ\kappa. Let

μ:TminT\mu:T^{{\operatorname{min}}}\to T

be the minimal resolution of a normal surface TT and let E1,,EnE_{1},...,E_{n} be all the μ\mu-exceptional prime divisors, i.e., Ex(μ)=E1En{\operatorname{Ex}}(\mu)=E_{1}\cup\cdots\cup E_{n}.

  1. (1)

    Given a \mathbb{Q}-divisor DD on TT, we define μD\mu^{*}D as a unique \mathbb{Q}-divisor on TminT^{{\operatorname{min}}} satisfying the following properties (a) and (b).

    1. (a)

      μμD=D\mu_{*}\mu^{*}D=D.

    2. (b)

      μDE1==μDEn=0\mu^{*}D\cdot E_{1}=\cdots=\mu^{*}D\cdot E_{n}=0.

    Equivalently, for the proper transform DD^{\prime} of DD on TminT^{{\operatorname{min}}}, μD\mu^{*}D is given by μD:=D+i=1neiEi\mu^{*}D:=D^{\prime}+\sum_{i=1}^{n}e_{i}E_{i}, where the rational numbers e1,,ene_{1},...,e_{n} are uniquely determined by (b), because the n×nn\times n matrix (EiEj)(E_{i}\cdot E_{j}) is invertible [Kol13, Theorem 10.1].

  2. (2)

    There exists an effective \mathbb{Q}-divisor Δ\Delta such that

    KTmin+Δ=μKT.K_{T^{{\operatorname{min}}}}+\Delta=\mu^{*}K_{T}.

    Furthermore, TT is canonical if and only if Δ=0\Delta=0 [Tan18, Theorem 4.13(1)].

Theorem 4.4.

We work over an algebraically closed field kk of characteristic p>0p>0. Let XX be a Fano threefold. Assume that (I) or (II) holds.

  1. (I)

    ρ(X)=1\rho(X)=1.

  2. (II)

    dim(Imφ|KX|)2\dim({\rm Im}\,\varphi_{|-K_{X}|})\geq 2.

Let

|KX|=|M|+F|-K_{X}|={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}|M|}+F

be the decomposition into the mobile part MM of |KX||-K_{X}| and the fixed part FF of |KX||-K_{X}|. Then F=0F=0 and the generic member SS of |KX||-K_{X}| is a geometrically integral regular K3-like surface.

Proof.

Since (I) implies (II) (Corollary 4.2), we may assume that (II) holds. Then general members of |M||M| are prime divisors (Proposition 2.11). Replacing MM by a general member of |M||M|, the problem is reduced to the case when MM is a prime divisor. By [CP08, Proposition 4.2], there exists a sequence of blowups

(4.4.1) σ:X~:=XσX1σ1σ2X1σ1X0:=X\sigma:\widetilde{X}:=X_{\ell}\xrightarrow{\sigma_{\ell}}X_{\ell-1}\xrightarrow{\sigma_{\ell-1}}\cdots\xrightarrow{\sigma_{2}}X_{1}\xrightarrow{\sigma_{1}}X_{0}:=X

such that

  • for each ii, σi:XiXi1\sigma_{i}:X_{i}\to X_{i-1} is a blowup along either a point or a smooth curve,

  • the centre σi(Ex(σi))\sigma_{i}({\operatorname{Ex}}(\sigma_{i})) is contained in the proper transform MiM_{i} of MM, and

  • the mobile part M~\widetilde{M} of |σM||\sigma^{*}M| is base point free.

Set k:=K((H0(X,KX)))k^{\prime}:=K(\mathbb{P}(H^{0}(X,-K_{X}))), which is a purely transcendental extension over kk of finite degree. In what follows, we set Y:=Y×kkY^{\prime}:=Y\times_{k}k^{\prime} for any kk-scheme YY. Applying the base change ()×kk(-)\times_{k}k^{\prime} to (4.4.1), we get a sequence of blowups (note that blowups commute with flat base changes [Liu02, Section 8, Proposition 1.12(c)]):

σ:X~:=XσX1σ1σ2X1σ1X0=X.\sigma^{\prime}:\widetilde{X}^{\prime}:=X^{\prime}_{\ell}\xrightarrow{\sigma^{\prime}_{\ell}}X^{\prime}_{\ell-1}\xrightarrow{\sigma^{\prime}_{\ell-1}}\cdots\xrightarrow{\sigma^{\prime}_{2}}X^{\prime}_{1}\xrightarrow{\sigma^{\prime}_{1}}X^{\prime}_{0}=X^{\prime}.

Let SS and S~\widetilde{S} be the generic members of |M||M| and |M~||\widetilde{M}|, respectively. By

H0(X,KX)H0(X,M)H0(X~,σM)H0(X~,M~),H^{0}(X,-K_{X})\simeq H^{0}(X,M)\simeq H^{0}(\widetilde{X},\sigma^{*}M)\simeq H^{0}(\widetilde{X},\widetilde{M}),

we have SM:=M×kkS\sim M^{\prime}:=M\times_{k}k^{\prime}, S~M~:=M~×kk\widetilde{S}\sim\widetilde{M}^{\prime}:=\widetilde{M}\times_{k}k^{\prime}, and σS~=S\sigma^{\prime}_{*}\widetilde{S}=S. As general members of each of |M||M| and |M~||\widetilde{M}| are prime divisors (Proposition 2.11), both SS and SS^{\prime} are geometrically integral prime divisors (indeed, for V:=H0(X,M)V:=H^{0}(X,M) and the family α:XM,Vuniv(V)\alpha:X^{{\operatorname{univ}}}_{M,V}\to\mathbb{P}(V) parametrising all the members of |M||M|, general members of |M||M| coincide with fibres over closed points of α\alpha and the generic member SS of |M||M| is nothing but the generic fibre of α\alpha (cf. Notation 2.7). Since general fibres of α\alpha are geometrically integral, so is the generic fibre SS [EGAIV3, Théorème 12.2.1(x)]. The same argument implies that also SS^{\prime} is geometrically integral). Furthermore, S~\widetilde{S} is regular, since |M~||\widetilde{M}| is base point free (Theorem 2.9). Let EiE_{i} be the prime divisor on X~\widetilde{X} that arises as the ii-th blowup, i.e., EiX~E_{i}\subset\widetilde{X} is the proper transform of Ex(σi)Xi{\operatorname{Ex}}(\sigma_{i})\subset X_{i}. There exist ni>0n_{i}\in\mathbb{Z}_{>0} and ai>0a_{i}\in\mathbb{Z}_{>0} such that the following hold.

  1. (1)

    σM=M~+iniEi\sigma^{*}M=\widetilde{M}+\sum_{i}n_{i}E_{i}.

  2. (2)

    σS=S~+iniEi\sigma^{\prime*}S=\widetilde{S}+\sum_{i}n_{i}E^{\prime}_{i}.

  3. (3)

    KX~=σKX+iaiEiK_{\widetilde{X}}=\sigma^{*}K_{X}+\sum_{i}a_{i}E_{i}.

  4. (4)

    KX~=σKX+iaiEiK_{\widetilde{X}^{\prime}}=\sigma^{\prime*}K_{X^{\prime}}+\sum_{i}a_{i}E^{\prime}_{i}.

Step 1.

If ii is an index such that the image σ(Ei)\sigma(E_{i}) on XX is one-dimensional, then it holds that niain_{i}\geq a_{i}.

Proof of Step 1.

Fix such ii. In order to prove niain_{i}\geq a_{i}, we may work with an open neighbourhood of the generic point ξi\xi_{i} of σ(Ei)\sigma(E_{i}). Over a suitable open neighbourhood UU of ξi\xi_{i}, σ|σ1(U):σ1(U)U\sigma|_{\sigma^{-1}(U)}:\sigma^{-1}(U)\to U is obtained by a sequence of blowups along smooth curves. We then inductively obtain

KXi+Mj+(effective divisor)=σj(KXj1+Mj1),K_{X_{i}}+M_{j}+(\text{effective divisor})=\sigma_{j}^{*}(K_{X_{j-1}}+M_{j-1}),

where M0:=MM_{0}:=M and MjM_{j} denotes the proper transform of MM. This implies niain_{i}\geq a_{i}, which completes the proof of Step 1. ∎

Let ν:SNS\nu:S^{N}\to S be the normalisation of SS and let μ:SminSN\mu:S^{{\operatorname{min}}}\to S^{N} be the minimal resolution of SNS^{N}. Since S~\widetilde{S} is regular, σ|S~:S~S\sigma^{\prime}|_{\widetilde{S}}:\widetilde{S}\to S factors through SminS^{{\operatorname{min}}}:

σ|S~:S~𝜆Smin𝜇SN𝜈S.\sigma^{\prime}|_{\widetilde{S}}:\widetilde{S}\xrightarrow{\lambda}S^{{\operatorname{min}}}\xrightarrow{\mu}S^{N}\xrightarrow{\nu}S.
Step 2.

The following hold.

  1. (5)

    KSNDK_{S^{N}}\sim-D for some effective \mathbb{Z}-divisor DD on SNS^{N}.

  2. (6)

    κ(S~,KS~)=κ(Smin,KSmin)κ(SN,KSN)0\kappa(\widetilde{S},K_{\widetilde{S}})=\kappa(S^{{\operatorname{min}}},K_{S^{{\operatorname{min}}}})\leq\kappa(S^{N},K_{S^{N}})\leq 0.

Proof of Step 2.

By (2) and (4), we get

(4.4.2) KX~+S~=σ(KX+S)+i(aini)Ei=σF+i(aini)Ei.K_{\widetilde{X}^{\prime}}+\widetilde{S}={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma^{\prime}}^{*}(K_{X^{\prime}}+S)+\sum_{i}(a_{i}-n_{i})E^{\prime}_{i}=-\sigma^{\prime*}F^{\prime}+\sum_{i}(a_{i}-n_{i})E^{\prime}_{i}.

Taking the pushforward of (4.4.2)|S~(\ref{e2-generic-ele})|_{\widetilde{S}} to SNS^{N} by μλ:S~SN\mu\circ\lambda:\widetilde{S}\to S^{N}, we obtain

KSN(μλ)KS~(μλ)(σF|S~)+i(aini)(μλ)(Ei|S~)0,K_{S^{N}}\sim(\mu\circ\lambda)_{*}K_{\widetilde{S}}\sim-(\mu\circ\lambda)_{*}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma^{\prime}}^{*}F|_{\widetilde{S}})+\sum_{i}(a_{i}-n_{i})(\mu\circ\lambda)_{*}(E_{i}|_{\widetilde{S}})\leq 0,

where σF|M~{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma^{\prime}}^{*}F|_{\widetilde{M}} is clearly effective and i(niai)(μλ)(Ei|S~)\sum_{i}(n_{i}-a_{i})(\mu\circ\lambda)_{*}(E_{i}|_{\widetilde{S}}) is effective by Step 1. Thus (5) holds.

Let us show (6). It follows from (5) that κ(SN,KSN)0\kappa(S^{N},K_{S^{N}})\leq 0. We have

H0(Smin,mKSmin)H0(SN,mKSN),H^{0}(S^{{\operatorname{min}}},mK_{S^{{\operatorname{min}}}})\hookrightarrow H^{0}(S^{N},mK_{S^{N}}),

which implies κ(Smin,KSmin)κ(SN,KSN)\kappa(S^{{\operatorname{min}}},K_{S^{{\operatorname{min}}}})\leq\kappa(S^{N},K_{S^{N}}). It is well known that κ(S~,KS~)=κ(Smin,KSmin)\kappa(\widetilde{S},K_{\widetilde{S}})=\kappa(S^{{\operatorname{min}}},K_{S^{{\operatorname{min}}}}). Thus (6) holds. This completes the proof of Step 2. ∎

Set A:=σ(KX)|S~A:={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma^{\prime}}^{*}(-K_{X^{\prime}})|_{\widetilde{S}} and

L:=KS~+A=(KX~+S~+σ(KX))|S~=(S~+iaiEi)|S~.L:=K_{\widetilde{S}}+A=\left(K_{\widetilde{X}^{\prime}}+\widetilde{S}+\sigma^{\prime*}(-K_{X^{\prime}})\right)\Big{|}_{\widetilde{S}}=\left(\widetilde{S}+\sum_{i}a_{i}E^{\prime}_{i}\right)\Big{|}_{\widetilde{S}}.

In particular,

(4.4.3) h0(S~,L)h0(S~,𝒪X~(S~)|S~).h^{0}(\widetilde{S},L)\geq h^{0}(\widetilde{S},\mathcal{O}_{\widetilde{X}^{\prime}}(\widetilde{S})|_{\widetilde{S}}).
Step 3.

It holds that h0(S~,A)2h^{0}(\widetilde{S},A)\geq 2.

Proof of Step 3.

It is enough to check the following inequalities:

h0(S~,A)=h0(S~,σ𝒪X(KX)|S~)h0(S,𝒪X(KX)|S)h0(S,𝒪X(S)|S)2.h^{0}(\widetilde{S},A)=h^{0}(\widetilde{S},{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma^{\prime}}^{*}\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}})|_{\widetilde{S}})\geq h^{0}(S,\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}})|_{S})\geq h^{0}(S,\mathcal{O}_{X^{\prime}}(S)|_{S})\geq 2.

It suffices to show the last inequality, since the other ones are obvious. By H1(X,𝒪X)=0H^{1}(X^{\prime},\mathcal{O}_{X^{\prime}})=0 and the exact sequence

0𝒪X𝒪X(S)𝒪X(S)|S0,0\to\mathcal{O}_{X^{\prime}}\to\mathcal{O}_{X^{\prime}}(S)\to\mathcal{O}_{X^{\prime}}(S)|_{S}\to 0,

we get

h0(S,𝒪X(S)|S)=h0(X,S)h0(X,𝒪X)=h0(X,KX)1h^{0}(S,\mathcal{O}_{X^{\prime}}(S)|_{S})=h^{0}(X^{\prime},S)-h^{0}(X^{\prime},\mathcal{O}_{X^{\prime}})=h^{0}(X^{\prime},-K_{X^{\prime}})-1
χ(X,KX)1=(KX)32+22.\geq\chi(X,-K_{X})-1=\frac{(-K_{X})^{3}}{2}+2\geq 2.

This completes the proof of Step 3. ∎

Step 4.

For g:=(KX)32+1g:=\frac{(-K_{X})^{3}}{2}+1 and g:=h0(X,KX)2g^{\prime}:=h^{0}(X,-K_{X})-2, it holds that

g1h0(S~,L)1h2(S~,𝒪S~)gh2(S~,𝒪S~).g-1\geq h^{0}(\widetilde{S},L)-1-h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})\geq g^{\prime}-h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}}).

Furthermore, if g=gg=g^{\prime} and h2(S~,𝒪S~)=1h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})=1, then F=0F=0.

Proof of Step 4.

In order to show the required inequalities, it is enough to prove the following inequalities:

2g2\displaystyle 2g-2 (α)σ(KX)(S~+iaiEi)S~\displaystyle\overset{(\alpha)}{\geq}\sigma^{\prime*}(-K_{X^{\prime}})\cdot\left(\widetilde{S}+\sum_{i}a_{i}E^{\prime}_{i}\right)\cdot\widetilde{S}
(β)2h0(S~,L)22h2(S~,𝒪S~)\displaystyle\overset{(\beta)}{\geq}2h^{0}(\widetilde{S},L)-2-2h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})
(γ)2g2h2(S~,𝒪S~).\displaystyle\overset{(\gamma)}{\geq}2g^{\prime}-2h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}}).

Let us show (α)(\alpha). It holds that

2g2\displaystyle 2g-2 =\displaystyle= σ(KX)3\displaystyle\sigma^{\prime*}(-K_{X^{\prime}})^{3}
=\displaystyle= σ(KX)2(σF+S~+iniEi)\displaystyle\sigma^{\prime*}(-K_{X^{\prime}})^{2}\cdot\left(\sigma^{\prime*}F^{\prime}+\widetilde{S}+\sum_{i}n_{i}E^{\prime}_{i}\right)
\displaystyle\geq σ(KX)2S~\displaystyle\sigma^{\prime*}(-K_{X^{\prime}})^{2}\cdot\widetilde{S}
=\displaystyle= σ(KX)(σF+S~+iniEi)S~\displaystyle\sigma^{\prime*}(-K_{X^{\prime}})\cdot\left(\sigma^{\prime*}F^{\prime}+\widetilde{S}+\sum_{i}n_{i}E^{\prime}_{i}\right)\cdot\widetilde{S}
\displaystyle\geq σ(KX)(S~+iaiEi)S~,\displaystyle\sigma^{\prime*}(-K_{X^{\prime}})\cdot\left(\widetilde{S}+\sum_{i}a_{i}E^{\prime}_{i}\right)\cdot\widetilde{S},

where the last inequality holds by σ(KX)σFS~0\sigma^{\prime*}(-K_{X^{\prime}})\cdot\sigma^{\prime*}F^{\prime}\cdot\widetilde{S}\geq 0 and the following facts.

  • If dimσ(Ei)=0\dim\sigma^{\prime}(E^{\prime}_{i})=0, then σ(KX)EiS~=0\sigma^{\prime*}(-K_{X^{\prime}})\cdot E^{\prime}_{i}\cdot\widetilde{S}=0.

  • If dimσ(Ei)=1\dim\sigma^{\prime}(E^{\prime}_{i})=1, then niain_{i}\geq a_{i} (Step 1).

Thus (α)(\alpha) holds.

Let us show (β)(\beta). By the Riemann–Roch theorem, we get

χ(S~,L)\displaystyle\chi(\widetilde{S},L) =χ(S~,𝒪S~)+12L(LKS~)\displaystyle=\chi(\widetilde{S},\mathcal{O}_{\widetilde{S}})+\frac{1}{2}L\cdot(L-K_{\widetilde{S}})
=χ(S~,𝒪S~)+12A(KS~+A)\displaystyle=\chi(\widetilde{S},\mathcal{O}_{\widetilde{S}})+\frac{1}{2}A\cdot(K_{\widetilde{S}}+A)
=χ(S~,𝒪S~)+12σ(KX)|S~(KX~+S~+σ(KX))|S~\displaystyle=\chi(\widetilde{S},\mathcal{O}_{\widetilde{S}})+\frac{1}{2}\sigma^{\prime*}(-K_{X^{\prime}})|_{\widetilde{S}}\cdot(K_{\widetilde{X}^{\prime}}+\widetilde{S}+\sigma^{\prime*}(-K_{X^{\prime}}))|_{\widetilde{S}}
=χ(S~,𝒪S~)+12σ(KX)(S~+iaiEi)S~.\displaystyle=\chi(\widetilde{S},\mathcal{O}_{\widetilde{S}})+\frac{1}{2}\sigma^{\prime*}(-K_{X^{\prime}})\cdot\left(\widetilde{S}+\sum_{i}a_{i}E^{\prime}_{i}\right)\cdot\widetilde{S}.

As A=σ(KX)|S~A=\sigma^{\prime*}(-K_{X^{\prime}})|_{\widetilde{S}} is big, we get

h2(S~,L)=h2(S~,KS~+A)=h0(S~,A)=0.h^{2}(\widetilde{S},L)=h^{2}(\widetilde{S},K_{\widetilde{S}}+A)=h^{0}(\widetilde{S},-A)=0.

Since S~\widetilde{S} is geometrically integral and h0(S~,A)2h^{0}(\widetilde{S},A)\geq 2 (Step 3), it follows from Theorem 3.3 that h1(S~,L)=h1(S~,A)h1(S~,𝒪S~)h^{1}(\widetilde{S},L)=h^{1}(\widetilde{S},-A)\leq h^{1}(\widetilde{S},\mathcal{O}_{\widetilde{S}}). Hence

12σ(KX)(S~+iaiEi)S~\displaystyle\,\,\frac{1}{2}\sigma^{\prime*}(-K_{X^{\prime}})\cdot\left(\widetilde{S}+\sum_{i}a_{i}E^{\prime}_{i}\right)\cdot\widetilde{S}
=\displaystyle= χ(S~,L)χ(S~,𝒪S~)\displaystyle\,\,\chi(\widetilde{S},L)-\chi(\widetilde{S},\mathcal{O}_{\widetilde{S}})
=\displaystyle= h0(S~,L)h1(S~,L)h0(S~,𝒪S~)+h1(S~,𝒪S~)h2(S~,𝒪S~)\displaystyle\,\,h^{0}(\widetilde{S},L)-h^{1}(\widetilde{S},L)-h^{0}(\widetilde{S},\mathcal{O}_{\widetilde{S}})+h^{1}(\widetilde{S},\mathcal{O}_{\widetilde{S}})-h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})
\displaystyle\geq h0(S~,L)1h2(S~,𝒪S~).\displaystyle\,\,h^{0}(\widetilde{S},L)-1-h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}}).

Note that we have h0(S~,𝒪S~)=1h^{0}(\widetilde{S},\mathcal{O}_{\widetilde{S}})=1, since S~\widetilde{S} is geometrically integral. Thus (β)(\beta) holds.

Let us show (γ)(\gamma). By H1(X~,𝒪X~)=0H^{1}(\widetilde{X}^{\prime},\mathcal{O}_{\widetilde{X}^{\prime}})=0 and the following exact sequence

0𝒪X~𝒪X~(S~)𝒪X~(S~)|S~0,0\to\mathcal{O}_{\widetilde{X}^{\prime}}\to\mathcal{O}_{\widetilde{X}^{\prime}}(\widetilde{S})\to\mathcal{O}_{\widetilde{X}^{\prime}}(\widetilde{S})|_{\widetilde{S}}\to 0,

we obtain

h0(S~,L)h0(S~,𝒪X~(S~)|S~)=h0(X~,S~)1=h0(X,KX)1=g+1.h^{0}(\widetilde{S},L)\geq h^{0}(\widetilde{S},\mathcal{O}_{\widetilde{X}^{\prime}}(\widetilde{S})|_{\widetilde{S}})=h^{0}(\widetilde{X}^{\prime},\widetilde{S})-1=h^{0}(X^{\prime},-K_{X^{\prime}})-1=g^{\prime}+1.

where the first inequality is guaranteed by (4.4.3). Thus (γ)(\gamma) holds.

Finally, assuming that g=gg=g^{\prime} and h2(S~,𝒪S~)=1h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})=1, it is enough to show that F=0F=0. In this case, all the inequalities (α),(β),(γ)(\alpha),(\beta),(\gamma) are equalities. By the proof of (α)(\alpha), we get σ(KX)2σF=0\sigma^{\prime*}(-K_{X^{\prime}})^{2}\cdot\sigma^{\prime*}F^{\prime}=0, which implies F=0F^{\prime}=0, i.e., F=0F=0. This completes the proof of Step 4. ∎

Step 5.

The following hold.

  1. (7)

    g=gg=g^{\prime}.

  2. (8)

    h2(S~,𝒪S~)=1h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})=1.

  3. (9)

    SS is a geometrically integral canonical K3-like surface.

Proof of Step 5.

It follows from (6) that h2(S~,𝒪S~)=h0(S~,KS~)1h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})=h^{0}(\widetilde{S},K_{\widetilde{S}})\leq 1. By ggg^{\prime}\geq g and Step 4, we obtain

gh0(S~,L)h2(S~,𝒪S~)g+1h2(S~,𝒪S~)g+11=g.g\geq h^{0}(\widetilde{S},L)-h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})\geq g^{\prime}+1-h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})\geq g+1-1=g.

Then all the inequalities in this equation are equalities, and hence

g=gandh2(S~,𝒪S~)=h0(S~,KS~)=1.g=g^{\prime}\qquad\text{and}\qquad h^{2}(\widetilde{S},\mathcal{O}_{\widetilde{S}})=h^{0}(\widetilde{S},K_{\widetilde{S}})=1.

Thus (7) and (8) hold.

Let us show (9). By (5), we have KSNDK_{S^{N}}\sim-D for some effective \mathbb{Z}-divisor DD. On the other hand, we get h0(Smin,KSmin)=h0(S~,KS~)=1h^{0}(S^{{\operatorname{min}}},K_{S^{{\operatorname{min}}}})=h^{0}(\widetilde{S},K_{\widetilde{S}})=1, which implies KSminEK_{S^{{\operatorname{min}}}}\sim E for some effective \mathbb{Z}-divisor EE. There exists an effective \mathbb{Q}-divisor Δ\Delta on SminS^{{\operatorname{min}}} such that KSmin+Δ=μKSNK_{S^{{\operatorname{min}}}}+\Delta=\mu^{*}K_{S^{N}} (4.3). We then get

E+ΔKSmin+Δ=μKSNμ(D),E+\Delta\equiv K_{S^{{\operatorname{min}}}}+\Delta=\mu^{*}K_{S^{N}}\equiv\mu^{*}(-D),

and hence E+Δ+μD0E+\Delta+\mu^{*}D\equiv 0. Since all of E,Δ,μDE,\Delta,\mu^{*}D are effective \mathbb{Q}-divisors, we obtain E=Δ=μD=0E=\Delta=\mu^{*}D=0 by taking the intersection with an ample divisor on SminS^{{\operatorname{min}}}. In particular, D=0D=0, KSN0K_{S^{N}}\sim 0, and KSmin0K_{S^{{\operatorname{min}}}}\sim 0. Then KSminμKSNK_{S^{{\operatorname{min}}}}\sim\mu^{*}K_{S^{N}}, which implies that SNS^{N} is canonical (4.3).

We have

ωS𝒪X(KX+S)|S𝒪X(F)|S.\omega_{S}\simeq\mathcal{O}_{X^{\prime}}(K_{X^{\prime}}+S)|_{S}\simeq\mathcal{O}_{X^{\prime}}(-F^{\prime})|_{S}.

Since SS is Gorenstein, we obtain νωS𝒪SN(KSN+C)\nu^{*}\omega_{S}\simeq\mathcal{O}_{S^{N}}(K_{S^{N}}+C) for the conductor CC, which is an effective \mathbb{Z}-divisor on SNS^{N} such that SuppC=Ex(ν){\operatorname{Supp}}\,C={\operatorname{Ex}}(\nu). We get

ν(𝒪X(F)|S)νωS𝒪SN(KSN+C)𝒪SN(C).\nu^{*}(\mathcal{O}_{X^{\prime}}(-F^{\prime})|_{S})\simeq\nu^{*}\omega_{S}\simeq\mathcal{O}_{S^{N}}(K_{S^{N}}+C)\simeq\mathcal{O}_{S^{N}}(C).

As CC is effective, it holds that C=0C=0, i.e., SS is normal. Hence SS is normal and has at worst canonical singularities. By KS0K_{S}\sim 0 and H1(S,𝒪S)=0H^{1}(S,\mathcal{O}_{S})=0, SS is a geometrically integral canonical K3-like surface. This completes the proof of Step 5. ∎

By Step 4 and Step 5, we obtain F=0F=0. Then the generic member SS of |KX|=|M||-K_{X}|=|M| is a geometrically integral canonical K3-like surface by Step 5. It follows from Corollary 3.18 that SS is regular. This completes the proof of Theorem 4.4. ∎

Corollary 4.5.

We work over an algebraically closed field kk of characteristic p>0p>0. Let XX be a Fano threefold. Assume that (I) or (II) holds.

  1. (I)

    ρ(X)=1\rho(X)=1.

  2. (II)

    dim(Im(φ|KX|))2\dim({\rm Im}(\varphi_{|-K_{X}|}))\geq 2.

Then the following hold.

  1. (1)

    Hi(X,nKX)=0H^{i}(X,nK_{X})=0 for all i{1,2}i\in\{1,2\} and nn\in\mathbb{Z}.

  2. (2)

    For all i>0i>0 and m0m\in\mathbb{Z}_{\geq 0}, it holds that Hi(X,mKX)=0H^{i}(X,-mK_{X})=0 and

    h0(X,mKX)=112m(m+1)(2m+1)(KX)3+2m+1.h^{0}(X,-mK_{X})=\frac{1}{12}m(m+1)(2m+1)(-K_{X})^{3}+2m+1.

    In particular, h0(X,KX)=12(KX)3+3=g+2h^{0}(X,-K_{X})=\frac{1}{2}(-K_{X})^{3}+3=g+2 for g:=12(KX)3+1g:=\frac{1}{2}(-K_{X})^{3}+1.

Proof.

Since (2) follows from (1) and Corollary 2.6, let us show (1). Fix nn\in\mathbb{Z}. Set k:=K((H0(X,KX)))k^{\prime}:=K(\mathbb{P}(H^{0}(X,-K_{X}))) and X:=X×kkX^{\prime}:=X\times_{k}k^{\prime}. Let SS be the generic member of |KX||-K_{X}|, which is a geometrically integral regular K3-like surface (Theorem 4.4). For \ell\in\mathbb{Z}, we have the following exact sequence:

0𝒪X(nKXS)𝒪X(nKX(1)S)𝒪X(nKX(1)S)|S0.0\to\mathcal{O}_{X^{\prime}}(nK_{X^{\prime}}-{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\ell}S)\to\mathcal{O}_{X^{\prime}}(nK_{X^{\prime}}-(\ell-1)S)\to\mathcal{O}_{X^{\prime}}(nK_{X^{\prime}}-(\ell-1)S)|_{S}\to 0.

By H1(S,𝒪X(nKX(1)S)|S)H1(S,𝒪X((n+1)KX))=0H^{1}(S,\mathcal{O}_{X^{\prime}}(nK_{X^{\prime}}-(\ell-1)S)|_{S})\simeq H^{1}(S,\mathcal{O}_{X^{\prime}}((n+\ell-1)K_{X^{\prime}}))=0 (Theorem 3.4), we have a surjection for any >0\ell>0:

H1(X,𝒪X(nKXS))H1(X,𝒪X(nKX)).H^{1}(X^{\prime},\mathcal{O}_{X^{\prime}}(nK_{X^{\prime}}-\ell S))\to H^{1}(X^{\prime},\mathcal{O}_{X^{\prime}}(nK_{X^{\prime}})).

For 0\ell\gg 0, the Serre vanishing theorem implies H1(X,𝒪X(nKX))=0H^{1}(X^{\prime},\mathcal{O}_{X^{\prime}}(nK_{X^{\prime}}))=0. By Serre duality, we get h2(X,𝒪X(nKX))=h1(X,𝒪X((1n)KX))=0h^{2}(X^{\prime},\mathcal{O}_{X^{\prime}}(nK_{X^{\prime}}))=h^{1}(X^{\prime},\mathcal{O}_{X^{\prime}}((1-n)K_{X^{\prime}}))=0. Thus (1) holds. ∎

5. The case when |KX||-K_{X}| is not base point free

Proposition 5.1.

We work over an algebraically closed field kk of characteristic p>0p>0. Let XX be a Fano threefold. Assume that Bs|KX|{\operatorname{Bs}}\,|-K_{X}|\neq\emptyset and dim(Imφ|KX|)2\dim({\rm Im}\,\varphi_{|-K_{X}|})\geq 2. Then the following hold.

  1. (1)

    The base scheme Z:=Bs|KX|Z:={\operatorname{Bs}}\,|-K_{X}| is isomorphic to k1\mathbb{P}^{1}_{k}.

  2. (2)

    For a general member TT of |KX||-K_{X}|, TT is a prime divisor which is smooth around ZZ.

  3. (3)

    KXZ=2gK_{X}\cdot Z=2-g.

  4. (4)

    There is an exact sequence

    0𝒪Z(g2)NZ/X𝒪Z(2)0,0\to\mathcal{O}_{Z}(g-2)\to N_{Z/X}\to\mathcal{O}_{Z}(-2)\to 0,

    where 𝒪Z()\mathcal{O}_{Z}(\ell) denotes the invertible sheaf of degree \ell on Zk1Z\simeq\mathbb{P}^{1}_{k}. In particular, degNZ/X=g4\deg N_{Z/X}=g-4.

Proof.

Set k:=K((H0(X,KX)))k^{\prime}:=K(\mathbb{P}(H^{0}(X,-K_{X}))). Let SS be the generic member of |KX||-K_{X}|. By Theorem 4.4, SS is a geometrically integral regular K3-like surface on X:=X×kkX^{\prime}:=X\times_{k}k^{\prime} with SKXS\sim-K_{X^{\prime}}.

Let us show (1). We have an exact sequence

0𝒪X(KXS)𝒪X(KX)𝒪X(KX)|S0.0\to\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}}-S)\to\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}})\to\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}})|_{S}\to 0.

By S|KX|S\in|-K_{X^{\prime}}| and H1(X,𝒪X(KXS))H1(X,𝒪X)=0H^{1}(X^{\prime},\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}}-S))\simeq H^{1}(X^{\prime},\mathcal{O}_{X^{\prime}})=0, it holds that Bs|KX|=Bs|KX|S|{\operatorname{Bs}}\,|-K_{X^{\prime}}|={\operatorname{Bs}}\,|-K_{X^{\prime}}|_{S}|. We then obtain

Bs|KX|×kkBs|KX|=Bs|KX|S|k1,{\operatorname{Bs}}\,|-K_{X}|\times_{k}k^{\prime}\simeq{\operatorname{Bs}}\,|-K_{X^{\prime}}|={\operatorname{Bs}}\,|-K_{X^{\prime}}|_{S}|\simeq\mathbb{P}^{1}_{k^{\prime}},

where the last isomorphism follows from Theorem 3.16, which is applicable because SS has a kk^{\prime}-rational point [Tana, Proposition 5.15(3)]. Then Bs|KX|S|{\operatorname{Bs}}\,|-K_{X^{\prime}}|_{S}| is a smooth projective curve over kk^{\prime} with

H1(Bs|KX|S|,𝒪Bs|KX|S|)=0.H^{1}({\operatorname{Bs}}\,|-K_{X^{\prime}}|_{S}|,\mathcal{O}_{{\operatorname{Bs}}\,|-K_{X^{\prime}}|_{S}|})=0.

These properties descend via the base change ()×kk(-)\times_{k}k^{\prime}, and hence Bs|KX|{\operatorname{Bs}}\,|-K_{X}| is a smooth projective curve over kk with H1(Bs|KX|,𝒪Bs|KX|)=0H^{1}({\operatorname{Bs}}\,|-K_{X}|,\mathcal{O}_{{\operatorname{Bs}}\,|-K_{X}|})=0, which implies Bs|KX|k1{\operatorname{Bs}}\,|-K_{X}|\simeq\mathbb{P}^{1}_{k}. Thus (1) holds.

Let us show (2). Recall that we have the universal family

ρ:𝒮X×k(H0(X,KX))pr2(H0(X,KX))\rho:\mathcal{S}\hookrightarrow X\times_{k}\mathbb{P}(H^{0}(X,-K_{X}))\xrightarrow{{\rm pr}_{2}}\mathbb{P}(H^{0}(X,-K_{X}))

parameterising the members of |KX||-K_{X}|. Here SS is the generic fibre of ρ\rho and a general member TT is a fibre of ρ\rho over a general closed point of (H0(X,KX))\mathbb{P}(H^{0}(X,-K_{X})). We have the following two inclusions, each of which is a closed immersion:

Bs|KX|×k(H0(X,KX))𝒮X×k(H0(X,KX)).{\operatorname{Bs}}\,|-K_{X}|\times_{k}\mathbb{P}(H^{0}(X,-K_{X}))\subset\mathcal{S}\subset X\times_{k}\mathbb{P}(H^{0}(X,-K_{X})).

After taking the generic fibres, these inclusions become

Bs|KX|×kkSX×kk=X.{\operatorname{Bs}}\,|-K_{X}|\times_{k}k^{\prime}\subset S\subset X\times_{k}k^{\prime}=X^{\prime}.

Note that Bs|KX|×kk(k1){\operatorname{Bs}}\,|-K_{X}|\times_{k}k^{\prime}(\simeq\mathbb{P}^{1}_{k^{\prime}}) is an effective Cartier divisor on a regular surface SS. Therefore, Bs|KX|{\operatorname{Bs}}\,|-K_{X}| is an effective Cartier divisor on TT for a general member T|KX|T\in|-K_{X}|. Since Bs|KX|{\operatorname{Bs}}\,|-K_{X}| is smooth by (1), TT is smooth around Bs|KX|{\operatorname{Bs}}\,|-K_{X}|. Thus (2) holds.

Let us show (3) and (4). By applying Theorem 3.16 for L:=𝒪X(KX)|SL:=\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}})|_{S}, we get

(Z×kk in S)2=2and𝒪X(KX)|S(Z×kk)=g2.(Z\times_{k}k^{\prime}\text{ in }S)^{2}=-2\qquad\text{and}\qquad\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}})|_{S}\cdot(Z\times_{k}k^{\prime})=g-2.

It holds that

KXZ=𝒪X(KX)(Z×kk)=𝒪X(KX)|S(Z×kk)=g2.-K_{X}\cdot Z=\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}})\cdot(Z\times_{k}k^{\prime})=\mathcal{O}_{X^{\prime}}(-K_{X^{\prime}})|_{S}\cdot(Z\times_{k}k^{\prime})=g-2.

Thus (3) holds. Since TT is smooth around ZZ, the following sequence is exact (cf. [Har77, Ch. II, Theorem 8.17]):

0NZ/TNZ/XNT/X|Z0.0\to N_{Z/T}\to N_{Z/X}\to N_{T/X}|_{Z}\to 0.

We have (Z in T)2=(Z×kk in S)2=2(Z\text{ in }T)^{2}=(Z\times_{k}k^{\prime}\text{ in }S)^{2}=-2 and TZ=KXZ=g2T\cdot Z=-K_{X}\cdot Z=g-2. Therefore, NZ/T𝒪Z(2)N_{Z/T}\simeq\mathcal{O}_{Z}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}-2}) and NT/X|Z𝒪Z(g2)N_{T/X}|_{Z}\simeq\mathcal{O}_{Z}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}g-2}). We get

degNZ/X=deg(detNZ/X)=degNZ/X+deg(NT/X|Z)=g4.\deg N_{Z/X}=\deg(\det N_{Z/X})=\deg N_{Z/X}+\deg(N_{T/X}|_{Z})=g-4.

Thus (4) holds. ∎

Proposition 5.2.

We work over an algebraically closed field kk of characeteristic p>0p>0. Let XX be a Fano threefold. Assume that Bs|KX|{\operatorname{Bs}}\,|-K_{X}|\neq\emptyset and dim(Imφ|KX|)2\dim({\rm Im}\,\varphi_{|-K_{X}|})\geq 2. Let σ:XX\sigma:X^{\prime}\to X be the blowup along Z:=Bs|KX|(k1)Z:={\operatorname{Bs}}\,|-K_{X}|(\simeq\mathbb{P}^{1}_{k}) (cf. Proposition 5.1) and let

(5.2.1) φ|KX|:X𝜓Yg+1\varphi_{|-K_{X^{\prime}}|}:X^{\prime}\xrightarrow{\psi}Y\hookrightarrow\mathbb{P}^{g+1}

be the induced morphisms, where Y:=φ|KX|(X)Y:=\varphi_{|-K_{X^{\prime}}|}(X^{\prime}) and the latter morphism Yg+1Y\hookrightarrow\mathbb{P}^{g+1} is the induced closed immersion. Set Z:=Ex(σ)Z^{\prime}:={\operatorname{Ex}}(\sigma). Then the following hold.

  1. (1)

    dimY=2\dim Y=2.

  2. (2)

    There exists a non-empty open subset YY^{\prime} of YY such that χ(Xy,𝒪Xy)=0\chi(X^{\prime}_{y},\mathcal{O}_{X^{\prime}_{y}})=0 for any point yYy\in Y^{\prime}.

  3. (3)

    ψ|Z:ZY\psi|_{Z^{\prime}}:Z^{\prime}\to Y is birational. Furthermore, either

    1. (a)

      ψ|Z\psi|_{Z^{\prime}} is an isomorphism, or

    2. (b)

      ψ|Z\psi|_{Z^{\prime}} is the contraction of the curve Γ\Gamma on ZZ^{\prime} with Γ2<0\Gamma^{2}<0.

  4. (4)

    ψ𝒪X=𝒪Y\psi_{*}\mathcal{O}_{X^{\prime}}=\mathcal{O}_{Y}.

  5. (5)

    There exists a non-empty open subset Y′′Y^{\prime\prime} of YY such that the scheme-theoretic fibre XyX^{\prime}_{y} is geometrically integral for every point yY′′y\in Y^{\prime\prime}.

Proof.

By

KX=σKX+Zandσ(KX)=KX+Z,K_{X^{\prime}}=\sigma^{*}K_{X}+Z^{\prime}\qquad\text{and}\qquad\sigma^{*}(-K_{X})=-K_{X^{\prime}}+Z^{\prime},

KX-K_{X^{\prime}} coincides with the mobile part of |σ(KX)||\sigma^{*}(-K_{X})|, which is base point free by construction. Hence we obtain the induced morphisms (5.2.1).

Let us show (1). By Y=Imφ|KX|Y={\rm Im}\,\varphi_{|-K_{X}|} and the assumption dim(Imφ|KX|)2\dim({\rm Im}\,\varphi_{|-K_{X}|})\geq 2, we have dimY2\dim Y\geq 2. In order to show dimY=2\dim Y=2, it suffices to show (KX)3=0(-K_{X^{\prime}})^{3}=0. For a fibre FF^{\prime} of ZZZ^{\prime}\to Z over a closed point of ZZ, we have σKXZ=(KXZ)F\sigma^{*}K_{X}\cdot Z^{\prime}=(K_{X}\cdot Z)F^{\prime}, ZF=1Z^{\prime}\cdot F^{\prime}=-1, and Z3=degZ(NZ/X)Z^{\prime 3}=-\deg_{Z}(N_{Z/X}) [Isk77, Lemma 2.11], which implies the following.

  • KX3=22gK_{X}^{3}=2-2g.

  • (σKX)2Z=0(\sigma^{*}K_{X})^{2}\cdot Z^{\prime}=0.

  • σKXZ2=(KXZ)FZ=KXZ=g2\sigma^{*}K_{X}\cdot Z^{\prime 2}=(K_{X}\cdot Z)F^{\prime}\cdot Z^{\prime}=-K_{X}\cdot Z=g-2 (Proposition 5.1).

  • Z3=degZ(NZ/X)=4gZ^{\prime 3}=-\deg_{Z}(N_{Z/X})=4-g (Proposition 5.1).

It holds that

KX3=(σKX+Z)3=KX3+3(σKX)2Z+3σKXZ2+Z3K_{X^{\prime}}^{3}=(\sigma^{*}K_{X}+Z^{\prime})^{3}=K_{X}^{3}+3(\sigma^{*}K_{X})^{2}\cdot Z^{\prime}+3\sigma^{*}K_{X}\cdot Z^{\prime 2}+Z^{\prime 3}
=(22g)+0+3(g2)+(4g)=0.=(2-2g)+0+3(g-2)+(4-g)=0.

Thus (1) holds.

Let us show (2). By the generic flatness, it suffices to show that χ(XK,𝒪XK)=0\chi(X^{\prime}_{K},\mathcal{O}_{X^{\prime}_{K}})=0, where K:=K(Y)K:=K(Y) and XKX^{\prime}_{K} denotes the generic fibre of ψ:XY\psi:X^{\prime}\to Y. Note that XKX^{\prime}_{K} is a regular projective curve over KK. Thus it is enough to check ωXK/K𝒪XK\omega_{X^{\prime}_{K}/K}\simeq\mathcal{O}_{X^{\prime}_{K}}. Note that ωX\omega_{X^{\prime}} is a pullback of an invertible sheaf on YY. For some non-empty smooth open subset Y1Y_{1} of YY and its inverse image X1:=ψ1(Y1)X^{\prime}_{1}:=\psi^{-1}(Y_{1}), we get

𝒪XKωX1/Y1|XKωXK/K,\mathcal{O}_{X^{\prime}_{K}}\simeq\omega_{X^{\prime}_{1}/Y_{1}}|_{X^{\prime}_{K}}\simeq\omega_{X^{\prime}_{K}/K},

where the latter isomorphism follows from [Con00, Theorem 3.6.1]. Thus (2) holds.

Let us show (3). The induced moprhism ψ|Z:ZY\psi|_{Z^{\prime}}:Z^{\prime}\to Y is surjective, because

(KX)2Z\displaystyle(-K_{X^{\prime}})^{2}\cdot Z^{\prime} =\displaystyle= (σKX+Z)2Z\displaystyle(\sigma^{*}K_{X}+Z^{\prime})^{2}\cdot Z^{\prime}
=\displaystyle= (σKX)2Z+2σKXZ2+Z3\displaystyle(\sigma^{*}K_{X})^{2}\cdot Z^{\prime}+2\sigma^{*}K_{X}\cdot Z^{\prime 2}+Z^{\prime 3}
=\displaystyle= 0+2(g2)+(4g)\displaystyle 0+2(g-2)+(4-g)
=\displaystyle= g=(KX)32+10.\displaystyle g=\frac{(-K_{X})^{3}}{2}+1\neq 0.

For :=deg(ψ|Z:ZY)\ell:=\deg(\psi|_{Z^{\prime}}:Z^{\prime}\to Y), the following holds:

0\displaystyle 0 Δ(Y,𝒪g+1(1)|Y)\displaystyle\leq\Delta(Y,\mathcal{O}_{\mathbb{P}^{g+1}(1)}|_{Y})
=dimY+(𝒪g+1(1)|Y)2h0(Y,𝒪g+1(1)|Y)\displaystyle=\dim Y+(\mathcal{O}_{\mathbb{P}^{g+1}}(1)|_{Y})^{2}-h^{0}(Y,\mathcal{O}_{\mathbb{P}^{g+1}}(1)|_{Y})
=2+(KX)2Zh0(X,KX)\displaystyle=2+\frac{(-K_{X^{\prime}})^{2}\cdot Z^{\prime}}{\ell}-h^{0}(X^{\prime},-K_{X^{\prime}})
=2+g(g+2)\displaystyle=2+\frac{g}{\ell}-(g+2)
=g(1),\displaystyle=\frac{g(1-\ell)}{\ell},

where the second equality follows from Lemma 3.13. Therefore, =1\ell=1 and Δ(Y,𝒪g+1(1)|Y)=0\Delta(Y,\mathcal{O}_{\mathbb{P}^{g+1}(1)}|_{Y})=0. In particular, ψ|Z:ZY\psi|_{Z^{\prime}}:Z^{\prime}\to Y is birational and YY is normal. Thus (3) holds.

Let us show (4). Let ψ:Xψ~Y~𝜃Y\psi:X^{\prime}\xrightarrow{\widetilde{\psi}}\widetilde{Y}\xrightarrow{\theta}Y be the Stein factorisation of ψ\psi. Since the composite morphism ψ|Z:Zψ~|ZY~𝜃Y\psi|_{Z^{\prime}}:Z^{\prime}\xrightarrow{\widetilde{\psi}|_{Z^{\prime}}}\widetilde{Y}\xrightarrow{\theta}Y is birational by (3), both ψ~|Z\widetilde{\psi}|_{Z^{\prime}} and θ\theta are birational. Then θ:Y~Y\theta:\widetilde{Y}\to Y is a finite birational morphism of normal projective surfaces by (3), and hence θ\theta is an isomorphism. Thus (4) holds.

Let us show (5). Fix a general closed point yYy\in Y. It is enough to show that the fibre XyX^{\prime}_{y} of ψ:XY\psi:X^{\prime}\to Y over yy is an integral scheme. By (4), the generic fibre of ψ:XY\psi:X^{\prime}\to Y is geometrically irreducible [Tan18-b, Lemma 2.2], and hence so is a general fibre XyX^{\prime}_{y} [EGAIV3, Proposition 9.7.8]. It is clear that dimXy=1\dim X^{\prime}_{y}=1. Since ψ|Z:ZY\psi|_{Z^{\prime}}:Z^{\prime}\to Y is birational, we get ZXy=1Z^{\prime}\cdot X^{\prime}_{y}=1. Hence XyX^{\prime}_{y} generically reduced [Bad01, Lemma 1.18], i.e., 𝒪Xy,γ\mathcal{O}_{X^{\prime}_{y},\gamma} is a field for the generic point γ\gamma of XyX^{\prime}_{y}. Since XyX^{\prime}_{y} is CM, XyX^{\prime}_{y} is S1S_{1} and R0R_{0}, i.e., XyX^{\prime}_{y} is reduced. Hence XyX^{\prime}_{y} is integral. Thus (5) holds.

Theorem 5.3.

We work over an algebraically closed field kk of characeteristic p>0p>0. Let XX be a Fano threefold with ρ(X)=1\rho(X)=1. Then |KX||-K_{X}| is base point free.

Proof.

Suppose that Bs|KX|{\operatorname{Bs}}\,|-K_{X}|\neq\emptyset. Let us derive a contradiction. By ρ(X)=1\rho(X)=1, we have dim(Imφ|KX|)2\dim({\rm Im}\,\varphi_{|-K_{X}|})\geq 2 (Corollary 4.2). Then we may apply Proposition 5.2 and we shall use the same notation as in the statement of Proposition 5.2. We have Z𝔽m=1(𝒪1𝒪1(m))Z^{\prime}\simeq\mathbb{F}_{m}=\mathbb{P}_{\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}}(m)) for some m0m\geq 0. By ρ(X)=1\rho(X)=1, we get ρ(X)=2\rho(X^{\prime})=2 and ρ(Y)=1\rho(Y)=1. It follows from Proposition 5.2 that Y(1,1,m)Y\simeq\mathbb{P}(1,1,m), which is nothing but the birational contraction of the curve ΓZ\Gamma\subset{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}Z^{\prime}} with Γ2<0\Gamma^{2}<0, where m:=Γ2>0m:=-\Gamma^{2}>0. Set y0:=ψ(Γ)Yy_{0}:=\psi(\Gamma)\in Y.

For a general fibre BZB_{Z^{\prime}} of σ|Z:Z=𝔽m1=Z\sigma|_{Z^{\prime}}:Z^{\prime}=\mathbb{F}_{m}\to\mathbb{P}^{1}=Z, we set B:=ψ(BZ)k1B:=\psi(B_{Z^{\prime}})\simeq\mathbb{P}^{1}_{k}. Then a general fibre of ψ1(B)B\psi^{-1}(B)\to B is one-dimensional and integral (Proposition 5.2). Furthermore, ψ1(B)B\psi^{-1}(B^{\circ})\to B^{\circ} is flat for some non-empty open subset BBB^{\circ}\subset B. Hence ψ1(B)\psi^{-1}(B^{\circ}) is integral and two-dimensional. Then its closure V:=ψ1(B)¯V:=\overline{\psi^{-1}(B^{\circ})}, equipped with the reduced scheme structure, is a prime divisor on XX^{\prime}. We have VXy=0V\cdot X^{\prime}_{y}=0 for a general closed point yYy\in Y.

Claim.

V=ψ1(B)redV=\psi^{-1}(B)_{{\operatorname{red}}}.

Proof of Claim.

The inclusion Vψ1(B)redV\subset\psi^{-1}(B)_{{\operatorname{red}}} is clear. Suppose that Vψ1(B)redV\not\supset\psi^{-1}(B)_{{\operatorname{red}}}. Then it holds that ψ1(b)redV\psi^{-1}(b)_{{\operatorname{red}}}\not\subset V for some bBb\in B. Since ψ1(b)\psi^{-1}(b) is connected (Proposition 5.2) and ψ1(b)V\psi^{-1}(b)\cap V\neq\emptyset, we can find a curve Cψ1(b)C\subset\psi^{-1}(b) such that VCV\cap C\neq\emptyset and CVC\not\subset V. In particular, VC>0V\cdot C>0. In order to derive a contradiction, let us prove that ρ(X)3\rho(X^{\prime})\geq 3. It is enough to show that HX,ψHY,VH_{X^{\prime}},\psi^{*}H_{Y},V are \mathbb{Z}-linearly independent in PicX/{\operatorname{Pic}}\,X^{\prime}/{\equiv}, where HXH_{X^{\prime}} and HYH_{Y} are ample Cartier divisors on XX^{\prime} and YY, respectively. For α,β,γ\alpha,\beta,\gamma\in\mathbb{Z}, assume that

αHX+βψHY+γV0.\alpha H_{X^{\prime}}+\beta\psi^{*}H_{Y}+\gamma V\equiv 0.

By taking the intersection with a general fibre of ψ\psi, we obtain α=0\alpha=0. Then consider the intersection with the above curve CC, which deduces γ=0\gamma=0 and hence β=0\beta=0. This completes the proof of Claim. ∎

Let ψV:VB\psi_{V}:V\to B be the morphism induced by ψ:XY\psi:X^{\prime}\to Y. Note that ψV\psi_{V} is a flat morphism from a projective surface VV to B1B\simeq\mathbb{P}^{1}. Since ψ1(B)\psi^{-1}(B^{\circ}) is an integral scheme, ψV:VB\psi_{V}:V\to B is a compatification of ψ1(B)B\psi^{-1}(B^{\circ})\to B^{\circ}. Therefore, we obtain

(5.3.1) χ(Vy,𝒪Vy)=χ(Xy,𝒪Xy)=0,\chi(V_{y},\mathcal{O}_{V_{y}})=\chi(X^{\prime}_{y},\mathcal{O}_{X^{\prime}_{y}})=0,
(5.3.2) χ(Vy0,𝒪Vy0)=χ(Vy,𝒪Vy),and𝒪X(Z)|VVy0=𝒪X(Z)|VVy\chi(V_{y_{0}},\mathcal{O}_{V_{y_{0}}})=\chi(V_{y},\mathcal{O}_{V_{y}}),\quad\text{and}\quad\mathcal{O}_{X^{\prime}}(Z^{\prime})|_{V}\cdot V_{y_{0}}=\mathcal{O}_{X^{\prime}}(Z^{\prime})|_{V}\cdot V_{y}

for a general closed point yBy\in B (Proposition 5.2). It holds that

(5.3.3) 𝒪X(Z)|VVy=ZXy=1.\mathcal{O}_{X^{\prime}}(Z^{\prime})|_{V}\cdot V_{y}=Z^{\prime}\cdot X^{\prime}_{y}=1.

On the other hand, Vy0=ψV1(y0)V_{y_{0}}=\psi_{V}^{-1}(y_{0}) is an effective Cartier divisor on VV, and hence pure one-dimensional. Let (Vy0)red=i=1sΓi(V_{y_{0}})_{{\operatorname{red}}}=\bigcup_{i=1}^{s}\Gamma_{i} be the irreducible decomposition with Γ1:=Γ\Gamma_{1}:=\Gamma. We then obtain

(5.3.4) 𝒪X(Z)|VVy0=𝒪X(Z)|V(i=1siΓi)=i=1si𝒪X(Z)Γi,\mathcal{O}_{X^{\prime}}(Z^{\prime})|_{V}\cdot V_{y_{0}}=\mathcal{O}_{X^{\prime}}(Z^{\prime})|_{V}\cdot\left(\sum_{i=1}^{s}\ell_{i}\Gamma_{i}\right)=\sum_{i=1}^{s}\ell_{i}\mathcal{O}_{X^{\prime}}(Z^{\prime})\cdot\Gamma_{i},

where i:=length𝒪Vy0,γi𝒪Vy0>0\ell_{i}:={\rm length}_{\mathcal{O}_{V_{y_{0}},\gamma_{i}}}\mathcal{O}_{V_{y_{0}}}>0 for the generic point γi\gamma_{i} of Γi\Gamma_{i} (cf. [Bad01, Lemma 1.18]). It follows from (5.3.2)–(5.3.4) that

i=1si𝒪X(Z)Γi=1.\sum_{i=1}^{s}\ell_{i}\mathcal{O}_{X^{\prime}}(Z^{\prime})\cdot\Gamma_{i}=1.

We now show that ZΓi>0Z^{\prime}\cdot\Gamma_{i}>0 for any ii. Otherwise, there exists ii such that ZΓi0Z^{\prime}\cdot\Gamma_{i}\leq 0. Fix ample Cartier divisors HXH_{X^{\prime}} and HYH_{Y} on XX^{\prime} and YY, respectively. It suffices to show that HX,ψHY,ZH_{X^{\prime}},\psi^{*}H_{Y},Z^{\prime} are \mathbb{Z}-linearly independent in PicX/{\operatorname{Pic}}\,X^{\prime}/\equiv, as this implies ρ(X)3\rho(X^{\prime})\geq 3. Assume

αHX+βψHY+γZ0\alpha H_{X^{\prime}}+\beta\psi^{*}H_{Y}+\gamma Z^{\prime}\equiv 0

for some α,β,γ\alpha,\beta,\gamma\in\mathbb{Z}. Taking the intersections with Γi\Gamma_{i} and a general fibre of ψ\psi, we obtain α=γ=0\alpha=\gamma=0 and hence β=0\beta=0. Therefore, ZΓi>0Z^{\prime}\cdot\Gamma_{i}>0 for any ii.

We then get s=1s=1 and 1=1\ell_{1}=1. In other words, (Vy0)red=Γ(V_{y_{0}})_{{\operatorname{red}}}=\Gamma and Vy0V_{y_{0}} is generically reduced, i.e., 𝒪Vy0,γ1\mathcal{O}_{V_{y_{0}},\gamma_{1}} is a field. Since VV is a prime divisor on a smooth threefold XX^{\prime}, VV is CM, and hence so is Vy0V_{y_{0}}. Therefore, Vy0V_{y_{0}} is reduced, i.e., Vy0=ΓV_{y_{0}}=\Gamma. Hence

0=χ(Vy,𝒪Vy)=χ(Vy0,𝒪Vy0)=h0(Γ,𝒪Γ)h1(Γ,𝒪Γ)=h0(Γ,𝒪Γ)=1,0=\chi(V_{y},\mathcal{O}_{V_{y}})=\chi(V_{y_{0}},\mathcal{O}_{V_{y_{0}}})={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}h^{0}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Gamma},\mathcal{O}_{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Gamma}})-h^{1}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Gamma},\mathcal{O}_{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Gamma}})=h^{0}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Gamma},\mathcal{O}_{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\Gamma}})=1,}

where the first equality follows from (5.3.1) and the second one holds by (5.3.2). This is a contradiction. ∎

6. The case when |KX||-K_{X}| is base point free

Throughout this section, we work over an algebraically closed field kk of characteristic p>0p>0.

6.1. Birational case

Lemma 6.1.

Let XX be a Fano threefold such that |KX||-K_{X}| is base point free. For Y:=Im(φ|KX|:Xh0(X,KX)1)Y:={\operatorname{Im}}(\varphi_{|-K_{X}|}:X\to\mathbb{P}^{h^{0}(X,-K_{X})-1}), let ψ:XY\psi:X\to Y be the morphism induced by φ|KX|\varphi_{|-K_{X}|}. Then degψ=1\deg\psi=1 or degψ=2\deg\psi=2.

Proof.

By construction, we can find an ample Cartier divisor HYH_{Y} on YY such that ψHYKX\psi^{*}H_{Y}\sim-K_{X} and ψ:H0(Y,HY)H0(X,KX)\psi^{*}:H^{0}(Y,H_{Y})\xrightarrow{\simeq}H^{0}(X,-K_{X}) (Lemma 3.13). It holds that

0Δ(Y,HY)=dimY+HY3h0(Y,HY)=3+(KX)3degψh0(X,KX).0\leq\Delta(Y,H_{Y})=\dim Y+H_{Y}^{3}-h^{0}(Y,H_{Y})=3+\frac{(-K_{X})^{3}}{\deg\psi}-h^{0}(X,-K_{X}).

By (KX)3=2g2(-K_{X})^{3}=2g-2 and h0(X,KX)=g+2h^{0}(X,-K_{X})=g+2 (Corollary 4.5), we obtain

g+2=h0(X,KX)3+(KX)3degψ=3+2g2degψ,g+2=h^{0}(X,-K_{X})\leq 3+\frac{(-K_{X})^{3}}{\deg\psi}=3+\frac{2g-2}{\deg\psi},

which implies 0(g1)(2degψ)0\leq(g-1)(2-\deg\psi). Hence we obtain degψ=1\deg\psi=1 or degψ=2\deg\psi=2. ∎

Theorem 6.2.

Let XX be a Fano threefold such that |KX||-K_{X}| is base point free and φ|KX|:Xh0(X,KX)1\varphi_{|-K_{X}|}:X\to\mathbb{P}^{h^{0}(X,-K_{X})-1} is birational onto its image. Then

m=0H0(X,𝒪X(mKX))\bigoplus_{m=0}^{\infty}H^{0}(X,\mathcal{O}_{X}(-mK_{X}))

is generated by H0(X,𝒪X(KX))H^{0}(X,\mathcal{O}_{X}(-K_{X})) as a kk-algebra. In particular, |KX||-K_{X}| is very ample.

Proof.

Set Y:=φ|KX|(X)Y:=\varphi_{|-K_{X}|}(X). We have the following induced morphisms:

φ|KX|:X𝜓Yh0(X,KX)1.\varphi_{|-K_{X}|}:X\xrightarrow{\psi}Y\hookrightarrow\mathbb{P}^{h^{0}(X,-K_{X})-1}.

Fix general hyperplane secionts HY,HYYH_{Y},H^{\prime}_{Y}\subset Y. Set

HX:=ψ1(HY),HX:=ψ1(HY),CY:=HYHY,CX:=HXHX.H_{X}:=\psi^{-1}(H_{Y}),\quad H^{\prime}_{X}:=\psi^{-1}(H^{\prime}_{Y}),\quad C_{Y}:=H_{Y}\cap H^{\prime}_{Y},\quad C_{X}:=H_{X}\cap H^{\prime}_{X}.

Note that all HX,HX,CX,HY,HY,CYH_{X},H^{\prime}_{X},C_{X},H_{Y},H^{\prime}_{Y},C_{Y} are integral schemes (Proposition 2.12). By applying Proposition 2.19 twice, it suffices to show that

R(CX,𝒪X(KX)|CX)R(C_{X},\mathcal{O}_{X}(-K_{X})|_{C_{X}})

is generated by H0(CX,𝒪X(KX)|CX)H^{0}(C_{X},\mathcal{O}_{X}(-K_{X})|_{C_{X}}). It follows from the adujction formula that CXC_{X} is a projective Gorenstein curve with

𝒪X(KX)|CX𝒪X(KX2KX)|CX(ωX𝒪X(HX)𝒪X(HX))|CXωCX.\mathcal{O}_{X}(-K_{X})|_{C_{X}}\simeq\mathcal{O}_{X}(K_{X}-2K_{X})|_{C_{X}}\simeq(\omega_{X}\otimes\mathcal{O}_{X}(H_{X})\otimes\mathcal{O}_{X}(H^{\prime}_{X}))|_{C_{X}}\simeq\omega_{C_{X}}.

Since |KX||-K_{X}| is base point free, also |KX|CX|=|ωCX||-K_{X}|_{C_{X}}|=|\omega_{C_{X}}| is base point free. By construction, φ|ωCX|\varphi_{|\omega_{C_{X}}|} is birational onto its image CYC_{Y}. Therefore, R(C,ωC)R(C,\omega_{C}) is generated by H0(C,ωC)H^{0}(C,\omega_{C}) [Fuj83, Theorem (A1) in page 39]. ∎

6.2. Hyperelliptic case

Definition 6.3.

We say that a Fano threefold XX is hyperelliptic if XX is of index one, |KX||-K_{X}| is base point free, and the morphism φ:Xφ|KX|(X)\varphi:X\to\varphi_{|-K_{X}|}(X), induced by φ|KX|:Xh0(X,KX)1\varphi_{|-K_{X}|}:X\to\mathbb{P}^{h^{0}(X,-K_{X})-1}, is of degree two.

Notation 6.4.

Let XX be a hyperelliptic Fano threefold and let φ:XY:=φ|KX|(X)g+1\varphi:X\to Y:=\varphi_{|-K_{X}|}(X)\subset\mathbb{P}^{g+1} be the morphism induced by φ|KX|\varphi_{|-K_{X}|}, where g:=h0(X,KX)2g:=h^{0}(X,-K_{X})-2. Since KX-K_{X} is ample, φ\varphi is a finite surjective morphism of projective threefolds such that [K(X):K(Y)]=2[K(X):K(Y)]=2. Set 𝒪Y():=𝒪g+1()|Y\mathcal{O}_{Y}(\ell):=\mathcal{O}_{\mathbb{P}^{g+1}}(\ell)|_{Y} for any \ell\in\mathbb{Z}.

Theorem 6.5.

We use Notation 6.4. Then the following hold.

  1. (1)

    Δ(Y,𝒪Y(1))=0\Delta(Y,\mathcal{O}_{Y}(1))=0.

  2. (2)

    YY is smooth.

  3. (3)

    One of the following holds.

    1. (i)

      Y3Y\simeq\mathbb{P}^{3}.

    2. (ii)

      YY is a smooth quadric hypersurface in 4\mathbb{P}^{4}.

    3. (iii)

      ρ(X)2\rho(X)\geq 2.

Proof.

Let us show (1). For the double cover φ:XY\varphi:X\to Y, it holds that

2g2=(KX)3=(degφ)(degY)=2degY,2g-2=(-K_{X})^{3}=(\deg\varphi)\cdot(\deg Y)=2\cdot\deg Y,

which implies degY=g1\deg Y=g-1. Therefore, we get

Δ(Y,𝒪Y(1))=dimY+𝒪Y(1)3h0(Y,𝒪Y(1))=3+degY(g+2)=0,\Delta(Y,\mathcal{O}_{Y}(1))=\dim Y+\mathcal{O}_{Y}(1)^{3}-h^{0}(Y,\mathcal{O}_{Y}(1))=3+\deg Y-(g+2)=0,

where the last equality follows from h0(Y,𝒪Y(1))=h0(X,KX)=g+2h^{0}(Y,\mathcal{O}_{Y}(1))=h^{0}(X,-K_{X})=g+2 (Lemma 3.13, Corollary 4.5) Thus (1) holds.

We now show that (1) and (2) imply (3). Assume that none of (i) and (ii) holds. By (1), (2), and Theorem 2.15, we have Y=1(E)Y=\mathbb{P}_{\mathbb{P}^{1}}(E) for some vector bundle EE on 1\mathbb{P}^{1} of rank 33. We then obtain ρ(X)ρ(Y)=2\rho(X)\geq\rho(Y)=2, and hence (iii) holds. Thus (1) and (2) imply (3).

It is enough to prove (2). Suppose that YY is singular. Let us derive a contradiction. By Remark 2.16, there are the following two cases.

  1. (a)

    YY is a cone over a smooth polarised surface (Z,𝒪Z(1))(Z,\mathcal{O}_{Z}(1)) with Δ(Z,𝒪Z(1))=0\Delta(Z,\mathcal{O}_{Z}(1))=0.

  2. (b)

    YY is a cone over a smooth polarised curve (Z,𝒪Z(1))(Z,\mathcal{O}_{Z}(1)) with Δ(Z,𝒪Z(1))=0\Delta(Z,\mathcal{O}_{Z}(1))=0.

Fix Cartier divisors HYH_{Y} and HZH_{Z} on YY and ZZ respectively such that 𝒪Y(HY)𝒪Y(1)\mathcal{O}_{Y}(H_{Y})\simeq\mathcal{O}_{Y}(1) and 𝒪Z(HZ)𝒪Z(1)\mathcal{O}_{Z}(H_{Z})\simeq\mathcal{O}_{Z}(1).

Assume (a). By Remark 2.16, we have

W=Z(𝒪Z𝒪Z(1))μYπZ,\begin{CD}W=\mathbb{P}_{Z}(\mathcal{O}_{Z}\oplus\mathcal{O}_{Z}(1))@>{\mu}>{}>Y\\ @V{}V{\pi}V\\ Z,\end{CD}

where π\pi is the induced projection and μ\mu is the birational morphism such that v:=μ(Z~)v:=\mu(\widetilde{Z}) is the vertex of YY for the section Z~W\widetilde{Z}\subset W of π\pi corresponding to the surjection: 𝒪Z𝒪Z(1)𝒪Z,(a,b)a\mathcal{O}_{Z}\oplus\mathcal{O}_{Z}(1)\to\mathcal{O}_{Z},\,\,(a,b)\mapsto a. By Theorem 2.15, (Z,𝒪Z(1))=(2,𝒪2(2))(Z,\mathcal{O}_{Z}(1))=(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(2)) or Z=𝔽nZ=\mathbb{F}_{n} for some n0n\geq 0. For the latter case: Z=𝔽nZ=\mathbb{F}_{n}, YY is not \mathbb{Q}-factorial (Proposition 8.5). However, this contradicts the fact that YY is the image of a finite morphsim φ:XY\varphi:X\to Y from a smooth variety XX [KM98, Lemma 5.16].

We may assume that (Z,𝒪Z(1))=(2,𝒪2(2))(Z,\mathcal{O}_{Z}(1))=(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(2)). We then have HZ2HZH_{Z}\sim 2H^{\prime}_{Z} for some Cartier divisor HZH^{\prime}_{Z} on ZZ, which implies

HYμπHZ2μπHZ=2HY,H_{Y}\sim\mu_{*}\pi^{*}H_{Z}\sim 2\mu_{*}\pi^{*}H^{\prime}_{Z}=2H^{\prime}_{Y},

where HY:=μπHZH^{\prime}_{Y}:=\mu_{*}\pi^{*}H^{\prime}_{Z} is a Weil divisor, which is not necessarily Cartier. Therefore, we get

KX=φHY2φHY,-K_{X}=\varphi^{*}H_{Y}\sim 2\varphi^{*}H^{\prime}_{Y},

where the linear equivalence can be checked after removing vv and φ1(v)\varphi^{-1}(v). This contradicts the assumption that XX is of index 11. This completes the case (a).

Assume (b). By Remark 2.16, we have (Z,𝒪Z(1))=(1,𝒪1(m))(Z,\mathcal{O}_{Z}(1))=(\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}}(m)) with m>0m>0 and Ym+2Y\subset\mathbb{P}^{m+2} is obtained from ZmZ\subset\mathbb{P}^{m} by applying cone construction:

W=Z(𝒪Z𝒪Z𝒪Z(1))μYπZ.\begin{CD}W=\mathbb{P}_{Z}(\mathcal{O}_{Z}\oplus\mathcal{O}_{Z}\oplus\mathcal{O}_{Z}(1))@>{\mu}>{}>Y\\ @V{}V{\pi}V\\ Z.\end{CD}

Furthermore, we get m2m\geq 2, as otherwise the cone YY would be smooth, i.e., Y3Y\simeq\mathbb{P}^{3}. Hence it holds that HYμπ(mP)H_{Y}\sim\mu_{*}\pi^{*}(mP) for a closed point P1P\in\mathbb{P}^{1}. This contradicts the assumption that XX is of index 11. ∎

Proposition 6.6.

We use Notation 6.4. Assume Y=3Y=\mathbb{P}^{3}. Then the following hold.

  1. (1)

    𝒪X(KX)φ𝒪3(1)\mathcal{O}_{X}(-K_{X})\simeq\varphi^{*}\mathcal{O}_{\mathbb{P}^{3}}(1).

  2. (2)

    (KX)3=2(-K_{X})^{3}=2.

  3. (3)

    If p2p\neq 2, then φ:XY=3\varphi:X\to Y=\mathbb{P}^{3} is a double cover ramified along a smooth prime divisor S3S\subset\mathbb{P}^{3} of degree 66.

Proof.

Since XX is of index one, we get (1). It is clear that (1) implies (2). For the branch divisor S3S\subset\mathbb{P}^{3}, we have

KX=ψ(KY+12S),K_{X}=\psi^{*}\left(K_{Y}+\frac{1}{2}S\right),

which deduces (3). ∎

Proposition 6.7.

We use Notation 6.4. Assume that YY is a smooth quadric hypersurface in 4\mathbb{P}^{4}. Then the following hold.

  1. (1)

    𝒪X(KX)φ(𝒪4(1)|Y)\mathcal{O}_{X}(-K_{X})\simeq\varphi^{*}(\mathcal{O}_{\mathbb{P}^{4}}(1)|_{Y}).

  2. (2)

    (KX)3=4(-K_{X})^{3}=4.

  3. (3)

    If p2p\neq 2, then φ:XY\varphi:X\to Y is a double cover ramified along a smooth prime divisor SYS\subset Y which is a complete intersection of YY and a hypersurface of degree 44.

Proof.

The same argument as in Proposition 6.6 works. ∎

Although the following proposition will not be used in the rest of this paper, we shall later need it in order to classify Fano threefolds with ρ=2\rho=2.

Proposition 6.8.

Let XX be a Fano threefold such that (KX)3=2(-K_{X})^{3}=2. Assume that dim(Imφ|KX|)2\dim({\operatorname{Im}}\,\varphi_{|-K_{X}|})\geq 2. Then the following hold.

  1. (1)

    XX is isomorphic to a weighted hypersurface in (1,1,1,1,3)\mathbb{P}(1,1,1,1,3) of degree 66.

  2. (2)

    ρ(X)=rX=1\rho(X)=r_{X}=1, where rXr_{X} denotes the index of XX.

  3. (3)

    h0(X,KX)=4h^{0}(X,-K_{X})=4 and the induced morphism φ|KX|:X3\varphi_{|-K_{X}|}:X\to\mathbb{P}^{3} is a double cover (cf. Proposition 6.6).

Proof.

First of all, we show that |KX||-K_{X}| is base point free. Suppose that |KX||-K_{X}| is not base point free. Let us derive a contradiction. By dim(Imφ|KX|)2\dim({\operatorname{Im}}\,\varphi_{|-K_{X}|})\geq 2, the generic member SS of |KX||-K_{X}| is a geometrically integral regular K3-like surface (Theorem 4.4). Set k:=K((H0(X,KX)))k^{\prime}:=K(\mathbb{P}(H^{0}(X,-K_{X}))) and X:=X×kkX^{\prime}:=X\times_{k}k^{\prime}. Then the restriction map

H0(X,KX)H0(S,KX|S)H^{0}(X^{\prime},-K_{X^{\prime}})\to H^{0}(S,-K_{X^{\prime}}|_{S})

is surjective, because H1(X,KXS)H1(X,𝒪X)H1(X,𝒪X)kk=0H^{1}(X,-K_{X^{\prime}}-S)\simeq H^{1}(X^{\prime},\mathcal{O}_{X^{\prime}})\simeq H^{1}(X,\mathcal{O}_{X})\otimes_{k}k^{\prime}=0. Hence |KX|S||-K_{X^{\prime}}|_{S}| is not base point free. It follows from Theorem 3.16(5) that the base locus FF of |KX|S||-K_{X^{\prime}}|_{S}| is a prime divisor on SS satisfying (KX|S)F=g2(-K_{X^{\prime}}|_{S})\cdot F=g-2 for g:=(KX|S)22+1g:=\frac{(-K_{X^{\prime}}|_{S})^{2}}{2}+1. We have g=(KX|S)22+1=(KX)32+1=2g=\frac{(-K_{X^{\prime}}|_{S})^{2}}{2}+1=\frac{(-K_{X})^{3}}{2}+1=2, which implies (KX|S)F=g2=0(-K_{X^{\prime}}|_{S})\cdot F=g-2=0. This is absurd, because (KX|S)(-K_{X^{\prime}}|_{S}) is ample. This completes the proof of the base point freeness of |KX||-K_{X}|.

By assuming (1), we now finish the proof. It is known that (1) implies ρ(X)=1\rho(X)=1 [Mor75, Theorem 3.7]. We have rX=1r_{X}=1, because rX2r_{X}\geq 2 implies (KX)323=8(-K_{X})^{3}\geq 2^{3}=8. By h0(X,KX)=12(KX)3+3=4h^{0}(X,-K_{X})=\frac{1}{2}(-K_{X})^{3}+3=4 (Corollary 4.5), we obtain a finite surjective morphism φ|KX|:X3\varphi_{|-K_{X}|}:X\to\mathbb{P}^{3}. By rX=14r_{X}=1\neq 4, φ|KX|\varphi_{|-K_{X}|} is a double cover (Lemma 6.1), and hence XX is hyperelliptic (Definition 6.3).

It is enough to show (1). Set 𝒪X():=𝒪X(KX)\mathcal{O}_{X}(\ell):=\mathcal{O}_{X}(-\ell K_{X}) for \ell\in\mathbb{Z}. Note that 𝒪X(KX)\mathcal{O}_{X}(-K_{X}) is 33-regular with respect to an ample globally generated invertible sheaf 𝒪X(KX)\mathcal{O}_{X}(-K_{X}) [Laz04, Section 1.8, especially Example 1.8.24], [FGI05, Subsection 5.2], i.e., the following holds (Corollary 4.5):

Hi(X,𝒪X((3i)KX))=0fori>0.H^{i}(X,\mathcal{O}_{X}(-(3-i)K_{X}))=0\qquad\text{for}\qquad i>0.

Then the induced kk-linear map

H0(X,KX)kH0(X,rKX)H0(X,(r+1)KX)H^{0}(X,-K_{X})\otimes_{k}H^{0}(X,-rK_{X})\to H^{0}(X,-(r+1)K_{X})

is surjective for every r3r\geq 3 [FGI05, Lemma 5.1(a)]. Then, as a kk-algebra, m=0H0(X,mKX)\bigoplus_{m=0}^{\infty}H^{0}(X,-mK_{X}) is generated by

H0(X,KX)H0(X,2KX)H0(X,3KX).H^{0}(X,-K_{X})\oplus H^{0}(X,-2K_{X})\oplus H^{0}(X,-3K_{X}).

It follows from Corollary 4.5 that

h0(X,mKX)=112m(m+1)(2m+1)(KX)3+2m+1h^{0}(X,-mK_{X})=\frac{1}{12}m(m+1)(2m+1)(-K_{X})^{3}+2m+1
=16m(m+1)(2m+1)+2m+1,=\frac{1}{6}m(m+1)(2m+1)+2m+1,

which implies the following:

  • h0(X,KX)=16123+3=4h^{0}(X,-K_{X})=\frac{1}{6}\cdot 1\cdot 2\cdot 3+3=4.

  • h0(X,2KX)=16235+5=10h^{0}(X,-2K_{X})=\frac{1}{6}\cdot 2\cdot 3\cdot 5+5=10.

  • h0(X,3KX)=16347+7=21h^{0}(X,-3K_{X})=\frac{1}{6}\cdot 3\cdot 4\cdot 7+7=21.

  • h0(X,4KX)=16459+9=39h^{0}(X,-4K_{X})=\frac{1}{6}\cdot 4\cdot 5\cdot 9+9=39.

  • h0(X,5KX)=165611+11=66h^{0}(X,-5K_{X})=\frac{1}{6}\cdot 5\cdot 6\cdot 11+11=66.

  • h0(X,6KX)=166713+13=104h^{0}(X,-6K_{X})=\frac{1}{6}\cdot 6\cdot 7\cdot 13+13=104.

Fix a kk-linear basis: H0(X,KX)=0i3kxiH^{0}(X,-K_{X})=\bigoplus_{0\leq i\leq 3}kx_{i}. Note that the subset {xixj| 0ij3}\{x_{i}x_{j}\,|\,0\leq i\leq j\leq 3\} (resp. {xixjxk| 0ijk3}\{x_{i}x_{j}x_{k}\,|\,0\leq i\leq j\leq k\leq 3\}) of H0(X,2KX)H^{0}(X,-2K_{X}) (resp. H0(X,3KX)H^{0}(X,-3K_{X})) is linearly independent over kk, as otherwise the induced morphism φ|KX|:X3\varphi_{|-K_{X}|}:X\to\mathbb{P}^{3} would factor through a hypersurface Z3Z\subset\mathbb{P}^{3} of degree 22 (resp. 33) which is defined by the corresponding linear dependence equation. Since the subspace generated by these elements is of dimension 1010 (resp. 2020), we can find an element x4H0(X,3KX)x_{4}\in H^{0}(X,-3K_{X}) such that H0(X,3KX)=S3(H0(X,KX))kx4H^{0}(X,-{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}3}K_{X})=S^{3}(H^{0}(X,-K_{X}))\oplus kx_{4}. Let k[y0,y1,y2,y3,y4]k[y_{0},y_{1},y_{2},y_{3},y_{4}] be the polynomial ring with degy0=degy1=degy2=degy3=1\deg y_{0}=\deg y_{1}=\deg y_{2}=\deg y_{3}=1 and degy4=3\deg y_{4}=3. We obtain a surjective graded kk-algebra homomorphism:

k[y0,y1,y2,y3,y4]m=0H0(X,mKX),yixi,k[y_{0},y_{1},y_{2},y_{3},y_{4}]\to\bigoplus_{m=0}^{\infty}H^{0}(X,-mK_{X}),\qquad y_{i}\mapsto x_{i},

which induces a closed immersion X(1,1,1,1,3)=Projk[y0,y1,y2,y3,y4]X\hookrightarrow\mathbb{P}(1,1,1,1,3)={\operatorname{Proj}}\,k[y_{0},y_{1},y_{2},y_{3},y_{4}]. Hence XX is a weighted hypersurface in (1,1,1,1,3)\mathbb{P}(1,1,1,1,3).

For d>0d\in\mathbb{Z}_{>0}, k[y0,y1,y2,y3,y4]dk[y_{0},y_{1},y_{2},y_{3},y_{4}]_{d} denotes the kk-linear supspace of k[y0,y1,y2,y3,y4]k[y_{0},y_{1},y_{2},y_{3},y_{4}] consisting of all the homogeneous elements of degree dd. We now see how to compute dimkk[y0,y1,y2,y3,y4]6\dim_{k}k[y_{0},y_{1},y_{2},y_{3},y_{4}]_{6}. Pick a monomial y0d0y1d1y2d2y3d3y4d4y_{0}^{d_{0}}y_{1}^{d_{1}}y_{2}^{d_{2}}y_{3}^{d_{3}}y_{4}^{d_{4}} with d0+d1+d2+d3+3d4=6d_{0}+d_{1}+d_{2}+d_{3}+{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}3}d_{4}=6. If d4=2d_{4}=2, then there is one solution. If d4=1d_{4}=1, then the number of the solutions of i=03di=3\sum_{i=0}^{3}d_{i}=3 is (63)=20\binom{6}{3}=20. If d4=0d_{4}=0, then the number of the solutions of i=03di=6\sum_{i=0}^{3}d_{i}=6 is (93)=84\binom{9}{3}=84. Hence dimkk[y0,,y4]6=1+20+84=105\dim_{k}k[y_{0},...,y_{4}]_{6}=1+20+84=105. Similarly, dimkk[y0,,y4]4=(43)+(73)=39\dim_{k}k[y_{0},...,y_{4}]_{4}=\binom{4}{3}+\binom{7}{3}=39 and dimkk[y0,,y4]5=(53)+(83)=66\dim_{k}k[y_{0},...,y_{4}]_{5}=\binom{5}{3}+\binom{8}{3}=66. Therefore, we obtain degX=6\deg X=6. Thus (1) holds. ∎

7. Intersection of quadrics

Throughout this section, we work over an algebraically closed field kk of characteristic p>0p>0.

7.1. Anti-canonically embedded Fano threefolds

Definition 7.1.

We say that Xg+1X\subset\mathbb{P}^{g+1} is an anti-canonically embedded Fano threefold if XX is a smooth projective threefold such that KX-K_{X} is very ample, XX is a closed subscheme of g+1\mathbb{P}^{g+1}, and the induced closed immersion Xg+1X\hookrightarrow\mathbb{P}^{g+1} is given by the complete linear system |KX||-K_{X}|. In this case, we have that h0(X,KX)=g+2h^{0}(X,-K_{X})=g+2 and degX=(KX)3=2g2\deg X=(-K_{X})^{3}=2g-2 (Corollary 4.5).

Remark 7.2.

Let Xg+1X\subset\mathbb{P}^{g+1} be an anti-canonically embedded Fano threefold. Then a general hyperplane section XgX\cap\mathbb{P}^{g} is a smooth K3 surface and the intersection Xg1X\cap\mathbb{P}^{g-1} with a general (g1)(g-1)-dimensional linear subvariety g1\mathbb{P}^{g-1} is a canonical curve of genes gg (Definition 7.3). In particular, g3g\geq 3.

Definition 7.3.

We say that Cg1C\subset\mathbb{P}^{g-1} is a canonical curve if CC is a smooth projective curve of genus gg such that |KC||K_{C}| is very ample, CC is a closed subscheme of g1\mathbb{P}^{g-1}, and the induced closed immersion Cg1C\hookrightarrow\mathbb{P}^{g-1} is given by the complete linear system |KC||K_{C}|.

7.2. Trigonal case

Notation 7.4.

Let Xg+1X\subset\mathbb{P}^{g+1} be an anti-canonically embedded Fano threefold with g5g\geq 5. Assume that XX is not an intersection of quadrics. Set

W:=XQ,Q:quadricQ,W:=\bigcap_{\begin{subarray}{c}X\subset Q,\\ Q:\text{quadric}\end{subarray}}Q,

where the right hand side denotes the scheme-theoretic intersection of all the quadric hypersurfaces of g+1\mathbb{P}^{g+1} containing XX.

Theorem 7.5 (The Noether–Enriques–Petri theorem).

Let Cg1C\subset\mathbb{P}^{g-1} be a canonical curve. Assume g5g\geq 5 and CC is not an intersection of quadrics. Set

WC:=CQ,Q:quadricQW_{C}:=\bigcap_{\begin{subarray}{c}C\subset Q,\\ Q:\text{quadric}\end{subarray}}Q

to be the scheme-theoretic intersection of all the quadric hypersurfaces in g1\mathbb{P}^{g-1} containing CC. Then the following hold.

  1. (1)

    WCW_{C} is a smooth projective surface with degWC=g2\deg W_{C}=g-2.

  2. (2)

    One of the following holds.

    1. (a)

      WCW_{C} is a rational scroll.

    2. (b)

      g=6g=6 and WC5W_{C}\subset\mathbb{P}^{5} is a Veronese surface.

Proof.

By [SD73, Theorem 4.7 and Lemma 4.8], (WC)red(W_{C})_{{\operatorname{red}}} is irreducible, two-dimensional, and deg(WC)red=g2\deg(W_{C})_{{\operatorname{red}}}=g-2. Then (WC)red(W_{C})_{{\operatorname{red}}} is a variety of minimal degree, and hence (WC)red(W_{C})_{{\operatorname{red}}} is an intersection of quadrics [SD74, Section 1 and Proposition 1.5(ii)] (cf. [Isk77, Lemma 2.5 and Remark 2.6]). Therefore, we obtain

WC=CQ,Q:quadricQ(WC)redQ,Q:quadricQ=(WC)red,W_{C}=\bigcap_{\begin{subarray}{c}C\subset Q,\\ Q:\text{quadric}\end{subarray}}Q\subset\bigcap_{\begin{subarray}{c}(W_{C})_{{\operatorname{red}}}\subset Q,\\ Q:\text{quadric}\end{subarray}}Q=(W_{C})_{{\operatorname{red}}},

which implies WC=(WC)redW_{C}=(W_{C})_{{\operatorname{red}}}, as required. Furthermore, WCW_{C} is smooth by [SD73, Lemma 4.10]. ∎

Lemma 7.6.

Let XNX\subset\mathbb{P}^{N} be a projective variety with dimX2\dim X\geq 2. Fix a hyperplane N1N\mathbb{P}^{N-1}\subset\mathbb{P}^{N} and set Y:=XN1Y:=X\cap\mathbb{P}^{N-1}. Assume that

  1. (i)

    H0(N,𝒪N(1))H0(X,𝒪N(1)|X)H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(1))\xrightarrow{\simeq}H^{0}(X,\mathcal{O}_{\mathbb{P}^{N}}(1)|_{X}) and

  2. (ii)

    m0(X,𝒪N(m)|X)\bigoplus_{m\geq 0}(X,\mathcal{O}_{\mathbb{P}^{N}}(m)|_{X}) is generated by H0(X,𝒪N(1)|X)H^{0}(X,\mathcal{O}_{\mathbb{P}^{N}}(1)|_{X}).

Then the following hold.

  1. (1)

    For

    • A:=A:= the set of the quadric hypersurfaces of N\mathbb{P}^{N} containing XX, and

    • B:=B:= the set of the quadric hypersurfaces of N1\mathbb{P}^{N-1} containing YY,

    the following map is bijective:

    AB,QQN1.A\to B,\qquad Q\mapsto Q\cap\mathbb{P}^{N-1}.
  2. (2)

    XNX\subset\mathbb{P}^{N} is an intersection of quadrics if and only if YN1Y\subset\mathbb{P}^{N-1} is an intersection of quadrics.

Proof.

See [Isk77, Lemma 2.10]. ∎

Proposition 7.7.

We use Notation 7.4. Then the following hold.

  1. (1)

    WW is a 44-dimensional projective normal variety such that Δ(W,𝒪g+1(1)|W)=0\Delta(W,\mathcal{O}_{\mathbb{P}^{g+1}}(1)|_{W})=0 and degW=g2\deg W=g-2.

  2. (2)

    dimSingW1\dim{\operatorname{Sing}}\,W\leq 1.

Proof.

Let us show (1). Let gg+1\mathbb{P}^{g}\subset\mathbb{P}^{g+1} be a hyperplane such that S:=XgS:=X\cap\mathbb{P}^{g} is smooth, and hence a K3 surface. Let g1g\mathbb{P}^{g-1}\subset\mathbb{P}^{g} be a hyperplane of g\mathbb{P}^{g} such that C:=Sg1C:=S\cap\mathbb{P}^{g-1} is smooth, and hence a canonical curve. Set 𝒪S(1):=𝒪g(1)|S\mathcal{O}_{S}(1):=\mathcal{O}_{\mathbb{P}^{g}}(1)|_{S} and 𝒪C(1):=𝒪g1(1)\mathcal{O}_{C}(1):=\mathcal{O}_{\mathbb{P}^{g-1}}(1). Note that R(S,𝒪S(1))R(S,\mathcal{O}_{S}(1)) and R(C,𝒪C(1))R(C,\mathcal{O}_{C}(1)) are generated by H0(S,𝒪S(1))H^{0}(S,\mathcal{O}_{S}(1)) and H0(C,𝒪C(1))H^{0}(C,\mathcal{O}_{C}(1)), respectively. In particular, the following 3 sets are bijectively corresponding via restriction (Lemma 7.6(1)):

  • The set of the quadric hypersurfaces of g+1\mathbb{P}^{g+1} containing XX.

  • The set of the quadric hypersurfaces of g\mathbb{P}^{g} containing SS.

  • The set of the quadric hypersurfaces of g1\mathbb{P}^{g-1} containing CC.

Recall that

W:=XQ,Q:quadricQW:=\bigcap_{\begin{subarray}{c}X\subset Q,\\ Q:\text{quadric}\end{subarray}}Q

is the scheme-theoretic intersection of the quadric hypersurfaces containing XX. Similarly, we set

WS:=SQ,Q:quadricQ,andWC:=CQ,Q:quadricQ.W_{S}:=\bigcap_{\begin{subarray}{c}S\subset Q,\\ Q:\text{quadric}\end{subarray}}Q,\qquad\text{and}\qquad W_{C}:=\bigcap_{\begin{subarray}{c}C\subset Q,\\ Q:\text{quadric}\end{subarray}}Q.

The above bijective correspondence deduces the following equality of closed subschemes of g+1\mathbb{P}^{g+1}:

WC=Wg1.W_{C}=W\cap\mathbb{P}^{g-1}.

By Theorem 7.5, WCg1W_{C}\subset\mathbb{P}^{g-1} is a surface of minimal degree, and hence WCW_{C} is an intersection of quadrics. Therefore, we obtain a decomposition

Wred=W1W2,W_{{\operatorname{red}}}=W_{1}\cup W_{2},

into reduced closed subschemes, where W1W_{1} is a 44-dimensional projective variety and dimW21\dim W_{2}\leq 1 (W2W_{2} is possibly reducible). We have XW1X\subset W_{1} and

WC=Wg1=W1g1.W_{C}=W\cap\mathbb{P}^{g-1}=W_{1}\cap\mathbb{P}^{g-1}.

Then degW1=degWC=g2\deg W_{1}=\deg W_{C}=g-2, and hence W1g+1W_{1}\subset\mathbb{P}^{g+1} is of minimal degree. In particular, W1W_{1} is an intersection of quadrics. Therefore,

W1WredW=XQ,Q:quadricQW1Q,Q:quadricQ=W1,W_{1}\subset W_{{\operatorname{red}}}\subset W=\bigcap_{\begin{subarray}{c}X\subset Q,\\ Q:\text{quadric}\end{subarray}}Q\subset\bigcap_{\begin{subarray}{c}W_{1}\subset Q,\\ Q:\text{quadric}\end{subarray}}Q=W_{1},

which implies W=W1W=W_{1}. Thus (1) holds.

Let us show (2). Suppose dimSingW2\dim{\operatorname{Sing}}\,W\geq 2. As WW is normal by (1), we obtain dimSingW=2\dim{\operatorname{Sing}}\,W=2. We take two general hyperplane sections and its intersection: g1=gg\mathbb{P}^{g-1}=\mathbb{P}^{g}\cap\mathbb{P}^{g}. Then C:=Xg1C:=X\cap\mathbb{P}^{g-1} and WC=Wg1W_{C}=W\cap\mathbb{P}^{g-1} are smooth (Theorem 7.5), which is a contradiction. Hence dimSingW1\dim{\rm Sing}W\leq 1. Thus (2) holds. ∎

Proposition 7.8.

We use Notation 7.4. If WW is smooth, then ρ(X)2\rho(X)\geq 2.

Proof.

We have the induced closed embeddings: XWg+1X\subset W\subset\mathbb{P}^{g+1}. Recall that degW=g252=3\deg W=g-2\geq 5-2=3. Since WW is smooth, we have (W,𝒪g+1(1)|W)=(4,𝒪4(1)),(W,𝒪g+1(1)|W)=(Q4,𝒪5(1)|Q4)(W,\mathcal{O}_{\mathbb{P}^{g+1}}(1)|_{W})=(\mathbb{P}^{4},\mathcal{O}_{\mathbb{P}^{4}}(1)),(W,\mathcal{O}_{\mathbb{P}^{g+1}}(1)|_{W})=(Q^{4},\mathcal{O}_{\mathbb{P}^{5}}(1)|_{Q^{4}}), or WW is a 3\mathbb{P}^{3}-bundle over 1\mathbb{P}^{1} (Theorem 2.15). If (W,𝒪g+1(1)|W)=(4,𝒪4(1))(W,\mathcal{O}_{\mathbb{P}^{g+1}}(1)|_{W})=(\mathbb{P}^{4},\mathcal{O}_{\mathbb{P}^{4}}(1)) or (W,𝒪g+1(1)|W)=(Q4,𝒪5(1)|Q4)(W,\mathcal{O}_{\mathbb{P}^{g+1}}(1)|_{W})=(Q^{4},\mathcal{O}_{\mathbb{P}^{5}}(1)|_{Q^{4}}), then we have degW2\deg W\leq 2, which is a contradiction. Hence WW is a 3\mathbb{P}^{3}-bundle over 1\mathbb{P}^{1}. Let π:W1\pi:W\to\mathbb{P}^{1} be the projection. If π(X)=1\pi(X)=\mathbb{P}^{1}, then ρ(X)2\rho(X)\geq 2.

Therefore, we may assume that π(X)\pi(X) is a point. In this case, XX is contained in a fibre of a 3\mathbb{P}^{3}-bundle π:W1\pi:W\to\mathbb{P}^{1}, and hence X3X\simeq\mathbb{P}^{3}. Then Xg+1X\subset\mathbb{P}^{g+1} is a Veronese variety, which is an intersection of quadrics (Lemma 7.9). Hence this case is excluded, because we assume, in Notation 7.4, that XX is not an intersection of quadrics. ∎

Lemma 7.9.

For positive integers nn and dd, set VV to be the image of the closed immersion φ|𝒪n(d)|:nN\varphi_{|\mathcal{O}_{\mathbb{P}^{n}}(d)|}:\mathbb{P}^{n}\hookrightarrow\mathbb{P}^{N}, where N:=(n+dn)1N:=\binom{n+d}{n}-1. Then VNV\subset\mathbb{P}^{N} is an intersection of quadrics.

The variety VV as above is called a Veronese variety.

Proof.

See [Sha13, Chapter 1, Subsection 4.4, Example 1.28] or [Her06, Theorem IV.25 and Appendix D]. ∎

Note that the singular locus SingW{\operatorname{Sing}}\,W of WW is a linear subvariety of g+1\mathbb{P}^{g+1}. In particular, the following hold.

  • dimSingW=0\dim{\operatorname{Sing}}\,W=0 if and only if SingW{\operatorname{Sing}}\,W is a point.

  • dimSingW=1\dim{\operatorname{Sing}}\,W=1 if and only if SingW{\operatorname{Sing}}\,W is a line.

Proposition 7.10.

We use Notation 7.4. Assume ρ(X)=1\rho(X)=1. Then dimSingW0\dim{\operatorname{Sing}}\,W\neq 0.

Proof.

Suppose that dimSingW=0\dim{\operatorname{Sing}}\,W=0, i.e., P:=SingWP:={\operatorname{Sing}}\,W is a point. Let us derive a contradiction. Note that Wg+1W\subset\mathbb{P}^{g+1} is a cone over a smooth threefold ZgZ\subset\mathbb{P}^{g} of minimal degree. By degZ=degW=g23\deg Z=\deg W=g-2\geq 3, ZZ is a 2\mathbb{P}^{2}-bundle over 1\mathbb{P}^{1} (Theorem 2.15). In particular, ρ(Z)=2\rho(Z)=2.

For the blowup μ~:WW\widetilde{\mu}:W^{\prime}\to W of WW at PP, we get the induced 1\mathbb{P}^{1}-bundle π~:WZ\widetilde{\pi}:W^{\prime}\to Z. For the proper transform X:=μ1(X)X^{\prime}:=\mu_{*}^{-1}(X) of XX, we have the induced morphisms μ:XX\mu:X^{\prime}\to X and π:=π~|X:XZ\pi:=\widetilde{\pi}|_{X^{\prime}}:X^{\prime}\to Z.

W{W^{\prime}}Z{Z}W{W}π~\scriptstyle{\widetilde{\pi}}μ~\scriptstyle{\widetilde{\mu}}            X{X^{\prime}}Z{Z}X.{X.}π\scriptstyle{\pi}μ\scriptstyle{\mu}

We treat the following two cases (I) and (II) separately:

(I)PX(II)PX.{\rm(I)}\,P\not\in X\hskip 56.9055pt{\rm(II)}\,P\in X.

(I) Assume PXP\not\in X. In this case, we get μ:XX\mu:X^{\prime}\xrightarrow{\simeq}X, and hence ρ(X)=ρ(X)=1\rho(X^{\prime})=\rho(X)=1. Since any fibre of π~\widetilde{\pi} is one-dimensional, it follows from ρ(X)=1\rho(X^{\prime})=1 that π:XZ\pi:X^{\prime}\to Z is surjective. Then the inequality 1=ρ(X)ρ(Z)1=\rho(X)\geq\rho(Z) implies ρ(Z)=1\rho(Z)=1. However, this contradicts ρ(Z)=2\rho(Z)=2.

(II) Assume PXP\in X. In this case, μ:XX\mu:X^{\prime}\to X is the blowup of XX at PP. In particular, XX^{\prime} is a smooth projective threefold with ρ(X)=2\rho(X^{\prime})=2. Since any fibre of π~\widetilde{\pi} is one-dimensional, it holds that dimπ(X)=2\dim\pi(X^{\prime})=2 or dimπ(X)=3\dim\pi(X^{\prime})=3.

Let us show dimπ(X)2\dim\pi(X^{\prime})\neq 2. Suppose dimπ(X)=2\dim\pi(X^{\prime})=2. In order to apply Lemma 7.11, let us confirm its assumptions, i.e., every fibre of π|X:Xπ(X)\pi|_{X^{\prime}}:X^{\prime}\to\pi(X^{\prime}) is 1\mathbb{P}^{1} and π(X)=2\pi(X^{\prime})=\mathbb{P}^{2}. For a closed point zπ(X)z\in\pi(X^{\prime}), we obtain a scheme-theoretic inclusion XzWz=1X^{\prime}_{z}\subset W^{\prime}_{z}=\mathbb{P}^{1}. Hence any fibre XzX^{\prime}_{z} of π|X:Xπ(X)\pi|_{X^{\prime}}:X^{\prime}\to\pi(X^{\prime}) is isomorphic to 1\mathbb{P}^{1}. It holds that π(X)\pi(X^{\prime}) is a fibre of the 2\mathbb{P}^{2}-bundle Z1Z\to\mathbb{P}^{1}, as otherwise the composite morphism XZ1X^{\prime}\to Z\to\mathbb{P}^{1} is surjective, which leads to the following contradiction for the normalisation of π(X)N\pi(X^{\prime})^{N}:

2=ρ(X)>ρ(π(X)N)>ρ(1)=1.2=\rho(X^{\prime})>\rho(\pi(X^{\prime})^{N})>\rho(\mathbb{P}^{1})=1.

Therefore, π(X)2\pi(X^{\prime})\simeq\mathbb{P}^{2}, and hence we may apply Lemma 7.11, so that X3X\simeq\mathbb{P}^{3}. This contradicts Lemma 7.9, which completes the proof of dimπ(X)2\dim\pi(X^{\prime})\neq 2.

Hence we obtain dimπ(X)=3\dim\pi(X^{\prime})=3, and hence π:XZ\pi:X^{\prime}\to Z is surjective. We get a surjection to 1\mathbb{P}^{1}: g:X𝜋Z𝜌1g:X^{\prime}\xrightarrow{\pi}Z\xrightarrow{\rho}\mathbb{P}^{1}, where ρ\rho denotes the projection (recall that ZZ is a 2\mathbb{P}^{2}-bundle over 1\mathbb{P}^{1}). Taking the Stein factorisation of gg, we obtain a morphism

h:XBh:X^{\prime}\to B

to a smooth projective curve BB with h𝒪X=𝒪Bh_{*}\mathcal{O}_{X^{\prime}}=\mathcal{O}_{B}. Set E:=Ex(μ)=2E:={\operatorname{Ex}}(\mu)=\mathbb{P}^{2}. Then the composite morphism

2=EXB\mathbb{P}^{2}=E\hookrightarrow X\to B

is trivial. We then obtain the following factorisation:

h:X𝜇X𝑞B.h:X^{\prime}\xrightarrow{\mu}X\xrightarrow{q}B.

However, this is a contradiction, because qq is surjective and ρ(X)=1\rho(X)=1. ∎

The following lemma has been already used in the above proof. We shall establish a more general result of this lemma in [AT].

Lemma 7.11.

Let XX be a Fano threefold with ρ(X)=1\rho(X)=1. Let μ:YX\mu:Y\to X be a blowup at a point PXP\in X. Assume that there exists a morphism π:Y2\pi:Y\to\mathbb{P}^{2} such that every fibre of π\pi is isomorphic to 1\mathbb{P}^{1}. Then X3X\simeq\mathbb{P}^{3}.

Proof.

We now show that rX=2r_{X}=2 or rX=4r_{X}=4. Suppose rX=1r_{X}=1 or rX=3r_{X}=3. Let us derive a contradiction. We can write KXrXHX-K_{X}\sim r_{X}H_{X} for some ample Cartier divisor HXH_{X} on XX satisfying PicX=HX{\operatorname{Pic}}\,X=\mathbb{Z}H_{X}. In particular, we obtain PicY=(μHX)E{\operatorname{Pic}}\,Y=\mathbb{Z}(\mu^{*}H_{X})\oplus\mathbb{Z}E for E:=Ex(μ)E:={\operatorname{Ex}}(\mu). Let LL be a line on 2\mathbb{P}^{2} and set D:=πLD:=\pi^{*}L. In particular, DD is a 1\mathbb{P}^{1}-bundle over 1\mathbb{P}^{1}. By 𝒪Y(E)|E𝒪E(1)\mathcal{O}_{Y}(E)|_{E}\simeq\mathcal{O}_{E}(-1) and KY=μKX+2EK_{Y}=\mu^{*}K_{X}+2E, we obtain

KY2E=4,KYE2=2,E3=1.K_{Y}^{2}\cdot E=4,\qquad K_{Y}\cdot E^{2}=2,\qquad E^{3}=1.

We also have

μKXKY2=μKX(μKX+2E)2=KX3\mu^{*}K_{X}\cdot K_{Y}^{2}=\mu^{*}K_{X}\cdot(\mu^{*}K_{X}+2E)^{2}=K_{X}^{3}
(μKX)2KY=(μKX)2(μKX+2E)=KX3.(\mu^{*}K_{X})^{2}\cdot K_{Y}=(\mu^{*}K_{X})^{2}\cdot(\mu^{*}K_{X}+2E)=K_{X}^{3}.

We can write

D=aμKXbE,D=-a\mu^{*}K_{X}-bE,

where a13a\in\frac{1}{3}\mathbb{Z} and bb\in\mathbb{Z}. Since κ(Y,D)=2\kappa(Y,D)=2 and κ(Y,μ(KX))=3\kappa(Y,\mu^{*}(-K_{X}))=3, it holds that a>0a>0 and b>0b>0. Since DD is a 1\mathbb{P}^{1}-bundle over 1\mathbb{P}^{1}, it holds that

DKY2=D(KY+DD)2=KD22KD(a fibre of DL)=82(2)=12.D\cdot K_{Y}^{2}=D\cdot(K_{Y}+D-D)^{2}=K_{D}^{2}-2K_{D}\cdot(\text{a fibre of }D\to L)=8-2(-2)=12.

We obtain

12=(aμKXbE)KY2=aKX34b.12=(-a\mu^{*}K_{X}-bE)\cdot K_{Y}^{2}=-aK_{X}^{3}-4b.

By D2KY=deg(KY|a fibre of π)=2D^{2}\cdot K_{Y}=\deg(K_{Y}|_{\text{a fibre of }\pi})=-2 and μKXE=0\mu^{*}K_{X}\cdot E=0, we obtain

2=D2KY=(aμKXbE)2KY=a2KX3+2b2.-2=D^{2}\cdot K_{Y}=(-a\mu^{*}K_{X}-bE)^{2}\cdot K_{Y}=a^{2}K_{X}^{3}+2b^{2}.

Hence we get a(4b+12)=a2KX3=2b2+2a(4b+12)=-a^{2}K_{X}^{3}=2b^{2}+2, which implies 2a(b+3)=b2+12a(b+3)=b^{2}+1. We can write a=a/3a=a^{\prime}/3 for some aa^{\prime}\in\mathbb{Z}, so that

2a(b+3)=3(b2+1).2a^{\prime}(b+3)=3(b^{2}+1).

Then bb is odd, and hence we have b=2b+1b=2b^{\prime}+1 for some b0b^{\prime}\in\mathbb{Z}_{\geq 0}, which implies

2a(b+2)=3(2b2+2b+1).2a^{\prime}(b^{\prime}+2)=3(2b^{\prime 2}+2b^{\prime}+1).

This is a contradiction, because the left (resp. right) hand side is even (resp. odd). Therefore, rX=2r_{X}=2 or rX=4r_{X}=4.

Note that KY-K_{Y} is ample by Kleiman’s criterion. Then it follows from KY=μKX+2EK_{Y}=\mu^{*}K_{X}+2E that rY=2r_{Y}=2. By ρ(Y)=2\rho(Y)=2, there are only two possibilities (Theorem 2.23): Y|𝒪2×2(1,1)|Y\in|\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1)| or YY is a blowup of 3\mathbb{P}^{3} at a point. For the former case, the two projections π1,π2:Y2\pi_{1},\pi_{2}:Y\to\mathbb{P}^{2} correspond to the distinct extremal rays. Therefore, we have that YY is a blowup of 3\mathbb{P}^{3} at a point. In this case, we obtain X3X\simeq\mathbb{P}^{3}. ∎

Proposition 7.12.

We use Notation 7.4. Assume ρ(X)=1\rho(X)=1. Then dimSingW1\dim{\operatorname{Sing}}\,W\neq 1.

Proof.

Suppose dimSingW=1\dim{\operatorname{Sing}}\,W=1, i.e., :=SingW\ell:={\operatorname{Sing}}\,W is a line on g+1\mathbb{P}^{g+1}. Let us derive a contradiction. In this case, Wg+1W\subset\mathbb{P}^{g+1} is a cone over a smooth surface Zg1Z\subset\mathbb{P}^{g-1} of minimal degree. By degZ=degW=g23\deg Z=\deg W=g-2\geq 3, either ZZ is a Hirzebrugh surface or Zg1Z\subset\mathbb{P}^{g-1} is a Veronese surface with g=6g=6 Theorem 2.15). By Theorem 8.10, a cone WW over the Veronese surface ZZ does not contain a smooth prime divisor. Hence ZZ is a Hirzebruch surface. In particular, ρ(Z)=2\rho(Z)=2.

For the blowup μ~:WW\widetilde{\mu}:W^{\prime}\to W of WW along \ell, we get the induced 2\mathbb{P}^{2}-bundle π~:WZ\widetilde{\pi}:W^{\prime}\to Z. For the proper transform X:=μ1(X)X^{\prime}:=\mu_{*}^{-1}(X) of XX, we have the induced morphisms μ:XX\mu:X^{\prime}\to X and π:=π~|X:XZ\pi:=\widetilde{\pi}|_{X^{\prime}}:X^{\prime}\to Z.

W{W^{\prime}}Z{Z}W{W}π~\scriptstyle{\widetilde{\pi}}μ~\scriptstyle{\widetilde{\mu}}            X{X^{\prime}}Z{Z}X.{X.}π\scriptstyle{\pi}μ\scriptstyle{\mu}

In what follows, we treat the following two cases separately:

(I)X(II)X.{\rm(I)}\,\,\ell\not\subset X\hskip 56.9055pt{\rm(II)}\,\,\ell\subset X.

(I) Assume X\ell\not\subset X. In this case, XX\cap\ell is either empty or zero-dimensional. Let SX×kκS\subset X\times_{k}\kappa be the generic hyperplane section of XX. Note that we have ρ(S)=ρ(X)=1\rho(S)=\rho(X)=1 [Tana, Proposition 5.17]. Since blowups commute with flat base changes, μ~×kκ:W×kκW×kκ\widetilde{\mu}\times_{k}\kappa:W^{\prime}\times_{k}\kappa\to W\times_{k}\kappa is the blowup along ×kκ\ell\times_{k}\kappa. Then SS is disjoint from the blowup centre ×kκ\ell\times_{k}\kappa. By ρ(S)=1\rho(S)=1, either SS is contained in a fibre of π~×kκ:W×kκZ×kκ\widetilde{\pi}\times_{k}\kappa:W^{\prime}\times_{k}\kappa\to Z\times_{k}\kappa or the induced morphism πS:SZ×kκ\pi_{S}:S\to Z\times_{k}\kappa is surjective. The former case is impossible, because κ(S)=0=κ(2)\kappa(S)=0\neq-\infty=\kappa(\mathbb{P}^{2}) (note that any fibre of π~×kκ\widetilde{\pi}\times_{k}\kappa is 2\mathbb{P}^{2}). Then πS:SZ×kκ\pi_{S}:S\to Z\times_{k}\kappa is surjective. Since ZZ is a Hirzebruch surface, we obtain the following contradiction:

1=ρ(S)ρ(Z×kκ)ρ(Z)=2.1=\rho(S)\geq\rho(Z\times_{k}\kappa)\geq\rho(Z)=2.

(II) Assume X\ell\subset X. Set E:=Ex(μ)E:={\operatorname{Ex}}(\mu), which is a 1\mathbb{P}^{1}-bundle over =1\ell=\mathbb{P}^{1}.

Let us show that E≄1×1E\not\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}. Recall that \ell is a line on g+1\mathbb{P}^{g+1}. It follows from [IP99, Proposition 2.2.14], that degN/X=2g()2KX=1\deg N_{\ell/X}=2g(\ell)-2-K_{X}\cdot\ell=-1. We have E(N/X)E\simeq\mathbb{P}_{\ell}(N_{\ell/X}). Since degN/X\deg N_{\ell/X} is an odd number, we have that E≄1×1E\not\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}.

We now show that π|E:EZ\pi|_{E}:E\to Z is surjective. Otherwise, its image is either a point or a curve. For the former case, we get the factorisation: π:X𝜇XZ\pi:X^{\prime}\xrightarrow{\mu}X\to Z, which contradicts ρ(X)=1\rho(X)=1. Suppose the latter case, i.e., the image of π|E:EZ\pi|_{E}:E\to Z is a curve. Since EE is a Hirzebruch surface with E≄1×1E\not\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}, the induced morphism μ|E:E\mu|_{E}:E\to\ell is the unique morphism to a curve satisfying (μ|E)𝒪E=𝒪(\mu|_{E})_{*}\mathcal{O}_{E}=\mathcal{O}_{\ell}. Hence we again get the factorisation: π:X𝜇XZ\pi:X^{\prime}\xrightarrow{\mu}X\to Z, which contradicts ρ(X)=1\rho(X)=1. This completes the proof of the surjectivity of π|E:EZ\pi|_{E}:E\to Z.

Since π|E:EZ\pi|_{E}:E\to Z is surjective, also π:XZ\pi:X^{\prime}\to Z is surjective. This leads to the following contradiction: 2=ρ(X)>ρ(Z)=22=\rho(X^{\prime})>\rho(Z)=2. ∎

We are ready to prove the main result of this section.

Theorem 7.13.

Let Xg+1X\subset\mathbb{P}^{g+1} be an anti-canonically embedded Fano threefold with g5g\geq 5 and ρ(X)=1\rho(X)=1. Then XX is an intersection of quadrics.

Proof.

Suppose that XX is not an intersection of quadrics. We use Notation 7.4. By Proposition 7.7, WW is a 44-dimensional normal variety with dimSingW1\dim{\operatorname{Sing}}\,W\leq 1. Hence the assertion follows from Proposition 7.8, Proposition 7.10, and Proposition 7.12. ∎

8. Appendix: Singular varieties of minimal degree

8.1. Rational normal scrolls

We summarise some basic properties on rational normal scrolls. In order to treat rational normal scrolls and cones over the Veronese surface simultaneously, we consider a slightly generalised setting, i.e., we consider a projective space bundle over n\mathbb{P}^{n} instead of over 1\mathbb{P}^{1}. Most of the arguments in this subsection are based on [EH87].

Notation 8.1.

Fix n>0n\in\mathbb{Z}_{>0}. Let 0a0a1ad0\leq a_{0}\leq a_{1}\leq\cdots\leq a_{d} be integers with ad>0a_{d}>0. For

E:=𝒪n(a0)𝒪n(ad),E:=\mathcal{O}_{\mathbb{P}^{n}}(a_{0})\oplus\cdots\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{d}),

we set

n(a0,,ad):=(E):=n(E)=n(𝒪n(a0)𝒪n(ad)),\mathbb{P}_{\mathbb{P}^{n}}(a_{0},...,a_{d}):=\mathbb{P}(E):=\mathbb{P}_{\mathbb{P}^{n}}(E)=\mathbb{P}_{\mathbb{P}^{n}}(\mathcal{O}_{\mathbb{P}^{n}}(a_{0})\oplus\cdots\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{d})),

which is a d\mathbb{P}^{d}-bundle over n\mathbb{P}^{n}. Let Sn(a0,,ad)S_{\mathbb{P}^{n}}(a_{0},...,a_{d}) be its image by 𝒪(E)(1)\mathcal{O}_{\mathbb{P}(E)}(1). Let

φ|𝒪(E)(1)|:n(a0,,ad)Sn(a0,,ad)1+ih0(n,𝒪n(ai))\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|}:\mathbb{P}_{\mathbb{P}^{n}}(a_{0},...,a_{d})\to S_{\mathbb{P}^{n}}(a_{0},...,a_{d})\subset\mathbb{P}^{-1+\sum_{i}h^{0}(\mathbb{P}^{n},\mathcal{O}_{\mathbb{P}^{n}}(a_{i}))}

be the morphism induced by the complete linear system |𝒪(E)(1)||\mathcal{O}_{\mathbb{P}(E)}(1)|, which is base point free (cf. Theorem 8.4(1)). Fix a hyperplane HH on n\mathbb{P}^{n} and set F:=πHF:=\pi^{*}H, where π:n(E)n\pi:\mathbb{P}_{\mathbb{P}^{n}}(E)\to\mathbb{P}^{n} denotes the projection. For each 0id0\leq i\leq d, we have

  • the section of π\pi

    Γi:=n(𝒪n(ai))n(E)\Gamma_{i}:=\mathbb{P}_{\mathbb{P}^{n}}(\mathcal{O}_{\mathbb{P}^{n}}(a_{i}))\subset\mathbb{P}_{\mathbb{P}^{n}}(E)

    corresponding to the projection E𝒪n(ai)E\to\mathcal{O}_{\mathbb{P}^{n}}(a_{i}), and

  • the d1\mathbb{P}^{d-1}-bundle over n\mathbb{P}^{n}

    Di:=n(𝒪n(a0)𝒪n(ai1)𝒪n(ai+1)𝒪n(ad))n(E)D_{i}:=\mathbb{P}_{\mathbb{P}^{n}}(\mathcal{O}_{\mathbb{P}^{n}}(a_{0})\oplus\cdots\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{i-1})\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{i+1})\oplus\cdots\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{d}))\subset\mathbb{P}_{\mathbb{P}^{n}}(E)

    corresponding to the projection

    E𝒪n(a0)𝒪n(ai1)𝒪n(ai+1)𝒪n(ad).E\to\mathcal{O}_{\mathbb{P}^{n}}(a_{0})\oplus\cdots\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{i-1})\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{i+1})\oplus\cdots\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{d}).
Remark 8.2.

We use Notation 8.1. By construction, the following hold for each 0id0\leq i\leq d.

  1. (1)

    We have DiΓi=D_{i}\cap\Gamma_{i}=\emptyset.

  2. (2)

    𝒪(E)(1)|Γi=𝒪(E)(1)|(𝒪n(ai))=𝒪(𝒪n(ai))(1)=𝒪n(ai)\mathcal{O}_{\mathbb{P}(E)}(1)|_{\Gamma_{i}}=\mathcal{O}_{\mathbb{P}(E)}(1)|_{\mathbb{P}(\mathcal{O}_{\mathbb{P}^{n}}(a_{i}))}=\mathcal{O}_{\mathbb{P}(\mathcal{O}_{\mathbb{P}^{n}}(a_{i}))}(1)=\mathcal{O}_{\mathbb{P}^{n}}(a_{i}).

  3. (3)

    By (2), φ|𝒪(E)(1)|(Γi)\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|}(\Gamma_{i}) is a point if and only if ai=0a_{i}=0.

Lemma 8.3.

We use Notation 8.1. For each 0id0\leq i\leq d, it holds that

𝒪(E)(1)Di+aiF.\mathcal{O}_{\mathbb{P}(E)}(1)\sim D_{i}+a_{i}F.
Proof.

We can write Di𝒪(E)(1)+rFD_{i}\sim\mathcal{O}_{\mathbb{P}(E)}(1)+rF for some rr\in\mathbb{Z}. Via π|Γi:Γin\pi|_{\Gamma_{i}}:\Gamma_{i}\xrightarrow{\simeq}\mathbb{P}^{n}, it holds that

𝒪n𝒪(E)(Di)|Γi(𝒪(E)(1)𝒪(E)(rF))|Γi𝒪n(ai+r).\mathcal{O}_{\mathbb{P}^{n}}\simeq\mathcal{O}_{\mathbb{P}(E)}(D_{i})|_{\Gamma_{i}}\simeq(\mathcal{O}_{\mathbb{P}(E)}(1)\otimes\mathcal{O}_{\mathbb{P}(E)}(rF))|_{\Gamma_{i}}\simeq\mathcal{O}_{\mathbb{P}^{n}}(a_{i}+r).

Hence ai+r=0a_{i}+r=0. We are done. ∎

Theorem 8.4.

We use Notation 8.1.

  1. (1)

    |𝒪(E)(1)||\mathcal{O}_{\mathbb{P}(E)}(1)| is base point free.

  2. (2)

    |𝒪(E)(1)||\mathcal{O}_{\mathbb{P}(E)}(1)| is very ample if and only if 0<a00<a_{0}, i.e., all of a0,,ada_{0},...,a_{d} are positive.

  3. (3)

    K(E)+i=0dDiπKnK_{\mathbb{P}(E)}+\sum_{i=0}^{d}D_{i}\sim\pi^{*}K_{\mathbb{P}^{n}}.

  4. (4)

    S(0,,0,a0,,ad)s1+ih0(n,𝒪n(ai))S(0,...,0,a_{0},...,a_{d})\subset\mathbb{P}^{s{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}-1}+\sum_{i}h^{0}(\mathbb{P}^{n},\mathcal{O}_{\mathbb{P}^{n}}(a_{i}))} is the ss-th cone over S(a0,,ad)1+ih0(n,𝒪n(ai))S(a_{0},...,a_{d})\subset\mathbb{P}^{{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}-1+}\sum_{i}h^{0}(\mathbb{P}^{n},\mathcal{O}_{\mathbb{P}^{n}}(a_{i}))}, where ss is the number of 0 appearing in S(0,,0,a0,,ad)S(0,...,0,a_{0},...,a_{d}).

Proof.

The assertion (1) follows from Lemma 8.3, D0Dd=D_{0}\cap\cdots\cap D_{d}=\emptyset, and the fact that |F||F| is base point free.

Let us show (2). Assume that 𝒪(E)(1)\mathcal{O}_{\mathbb{P}(E)}(1) is very ample. Then all of a0,,ada_{0},...,a_{d} are positive by Remark 8.3(3). Let us prove the opposite implication. Assume that all of a0,,ada_{0},...,a_{d} are positive. As n(E)\mathbb{P}_{\mathbb{P}^{n}}(E) is a smooth projective toric variety, it is enough to show that 𝒪(E)(1)\mathcal{O}_{\mathbb{P}(E)}(1) is ample [CLS11, Theorem 6.1.15]. Fix a curve CC on n(E)\mathbb{P}_{\mathbb{P}^{n}}(E). Since |𝒪(E)(1)||\mathcal{O}_{\mathbb{P}(E)}(1)| is base point free by (1), it suffices to show that 𝒪(E)(1)C>0\mathcal{O}_{\mathbb{P}(E)}(1)\cdot C>0. If π(C)\pi(C) is a point, then this follows from the fact that 𝒪(E)(1)\mathcal{O}_{\mathbb{P}(E)}(1) is π\pi-ample. Hence we may assume that π(C)\pi(C) is a curve. By D0Dd=D_{0}\cap\cdots\cap D_{d}=\emptyset, there exists 0id0\leq i\leq d such that CDiC\not\subset D_{i}. Then

𝒪(E)(1)C=DiC+aiFCaiFC>0.\mathcal{O}_{\mathbb{P}(E)}(1)\cdot C=D_{i}\cdot C+a_{i}F\cdot C\geq a_{i}F\cdot C>0.

Thus (2) holds.

Let us show (3). Fix a fibre d\mathbb{P}^{d} of π:(E)n\pi:\mathbb{P}(E)\to\mathbb{P}^{n}. Then we obtain

(K(E)+i=0dDi)|d=Kd+i=0d(Di|d)0,(K_{\mathbb{P}(E)}+\sum_{i=0}^{d}D_{i})|_{\mathbb{P}^{d}}=K_{\mathbb{P}^{d}}+\sum_{i=0}^{d}(D_{i}|_{\mathbb{P}^{d}})\sim 0,

because D0|d,,Dd|dD_{0}|_{\mathbb{P}^{d}},...,D_{d}|_{\mathbb{P}^{d}} are hyperplanes. Hence we get K(E)+i=0dDiπLK_{\mathbb{P}(E)}+\sum_{i=0}^{d}D_{i}\sim\pi^{*}L for some Cartier divisor LL on n\mathbb{P}^{n}. By taking the restriction to the section Γ0=n\Gamma_{0}=\mathbb{P}^{n} of π\pi, we get

L(K(E)+i=0dDi)|Γ0=(K(E)+i=1dDi)|Γ0KΓ0,L\sim(K_{\mathbb{P}(E)}+\sum_{i=0}^{d}D_{i})|_{\Gamma_{0}}=(K_{\mathbb{P}(E)}+\sum_{i=1}^{d}D_{i})|_{\Gamma_{0}}\sim K_{\Gamma_{0}},

where the latter linear equivalence follows from adjunction formula and D1Dn=Γ0D_{1}\cap\cdots\cap D_{n}=\Gamma_{0}. Thus (3) holds.

Concerning (4), we may apply the same argument as in [EH87, page 6]. ∎

Proposition 8.5.

We use Notation 8.1. Then the following are equivalent.

  1. (1)

    Sn(a0,,ad)S_{\mathbb{P}^{n}}(a_{0},...,a_{d}) is \mathbb{Q}-factorial.

  2. (2)

    One of the following holds.

    1. (a)

      a0>0a_{0}>0, i.e., all of a0,,ada_{0},...,a_{d} are positive.

    2. (b)

      a0==ad1=0a_{0}=\cdots=a_{d-1}=0, i.e., only ada_{d} is positive.

Furthermore, if (b) holds, then φ|𝒪(E)(1)|\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|} is a birational morphism with Ex(φ|𝒪(E)(1)|)=Dd{\operatorname{Ex}}(\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|})=D_{d}.

Proof.

Let us show (1) \Rightarrow (2). Assume that (2) does not hold. Then a0=0a_{0}=0 and ad1>0a_{d-1}>0. In particular, ad>0a_{d}>0. Let CC be a curve on n(E)\mathbb{P}_{\mathbb{P}^{n}}(E) such that φ|𝒪(E)(1)|(C)\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|}(C) is a point. In order to prove that S(a0,,ad)S(a_{0},...,a_{d}) is not \mathbb{Q}-factorial, it is enough to show that CDd1DdC\subset D_{d-1}\cap D_{d}. It follows that π(C)\pi(C) is a curve, because π\pi and φ|𝒪(E)(1)|\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|} correspond to distinct extremal rays. For i{d1,d}i\in\{d-1,d\}, we have

0=𝒪(E)(1)C=(Di+aiF)C=DiC+aiFC>DiC,0=\mathcal{O}_{\mathbb{P}(E)}(1)\cdot C=(D_{i}+a_{i}F)\cdot C=D_{i}\cdot C+a_{i}F\cdot C>D_{i}\cdot C,

which implies CDiC\subset D_{i}. This completes the proof of (1) \Rightarrow (2).

Let us show (2) \Rightarrow (1). Assume (2). It suffices to prove that S(a0,,ad)S(a_{0},...,a_{d}) is \mathbb{Q}-factorial. If (a) holds, then φ|𝒪(E)(1)|\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|} is a closed immersion (Theorem 8.4), so that we get (a0,,ad)S(a0,,ad)\mathbb{P}(a_{0},...,a_{d})\simeq S(a_{0},...,a_{d}), and hence S(a0,,ad)S(a_{0},...,a_{d}) is \mathbb{Q}-factorial. We may assume that a0==ad1=0a_{0}=\cdots=a_{d-1}=0. We then have E=𝒪nd𝒪n(ad)E=\mathcal{O}_{\mathbb{P}^{n}}^{\oplus d}\oplus\mathcal{O}_{\mathbb{P}^{n}}(a_{d}) with ad>0a_{d}>0. By

Dd=n(𝒪nd)n×d1,D_{d}=\mathbb{P}_{\mathbb{P}^{n}}(\mathcal{O}_{\mathbb{P}^{n}}^{\oplus d})\simeq\mathbb{P}^{n}\times\mathbb{P}^{d-1},

we see that dimDd>dimφ(Dd)\dim D_{d}>\dim\varphi(D_{d}). Hence φ|𝒪(E)(1)|:n(a0,,ad)Sn(a0,,ad)\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|}:\mathbb{P}_{\mathbb{P}^{n}}(a_{0},...,a_{d})\to S_{\mathbb{P}^{n}}(a_{0},...,a_{d}) is a divisorial contraction, i.e., Ex(φ|𝒪(E)(1)|){\operatorname{Ex}}(\varphi_{|\mathcal{O}_{\mathbb{P}(E)}(1)|}) is a prime divisor. Since (a0,,ad)\mathbb{P}(a_{0},...,a_{d}) is toric, S(a0,,ad)S(a_{0},...,a_{d}) is \mathbb{Q}-factorial [CLS11, Proposition 15.4.5]. ∎

We shall need the following lemma in the next subsection. Although this result is well known, we give a proof for the sake of completeness.

Lemma 8.6.

Let Xn=Projk[x0,,xn]X\subset\mathbb{P}^{n}={\operatorname{Proj}}\,k[x_{0},...,x_{n}] be a projective variety. Take s>0s\in\mathbb{Z}_{>0} and let Yn+s=Projk[x0,,xn,y1,,ys]Y\subset\mathbb{P}^{n+s}={\operatorname{Proj}}\,k[x_{0},...,x_{n},y_{1},...,y_{s}] be the ss-th cone of XX. Set LL to be the vertex, i.e., LL is the reduced closed subscheme of n+s\mathbb{P}^{n+s} whose closed points are {[x0::xn:y1::ys]n+s(k)|x0==xn=0,y1,,ysk}\{[x_{0}:\cdots:x_{n}:y_{1}:...:y_{s}]\in\mathbb{P}^{n+s}(k)\,|\,x_{0}=...=x_{n}=0,y_{1},\cdots,y_{s}\in k\}. Then, for 𝒪X(1):=𝒪n(1)|X\mathcal{O}_{X}(1):=\mathcal{O}_{\mathbb{P}^{n}}(1)|_{X} and Z:=X(𝒪Xs𝒪X(1))Z:=\mathbb{P}_{X}(\mathcal{O}^{\oplus s}_{X}\oplus\mathcal{O}_{X}(1)), there exist the following morphisms

Z=X(𝒪Xs𝒪X(1)){Z=\mathbb{P}_{X}(\mathcal{O}_{X}^{\oplus s}\oplus\mathcal{O}_{X}(1))}X{X}Y,{Y,}π\scriptstyle{\pi}σ\scriptstyle{\sigma}

where π:Z=X(𝒪Xs𝒪X(1))X\pi:Z=\mathbb{P}_{X}(\mathcal{O}_{X}^{\oplus s}\oplus\mathcal{O}_{X}(1))\to X denotes the induced s\mathbb{P}^{s}-bundle and σ:ZY\sigma:Z\to Y is the blowup along the vertex LL.

Proof.

The proof consists of the following two steps:

  1. (I)

    The case when X=nX=\mathbb{P}^{n}.

  2. (II)

    The general case.

(I) We first treat the case when X=nX=\mathbb{P}^{n}. In this case, we have Y=n+sY=\mathbb{P}^{n+s} and Z=n(𝒪ns𝒪n(1))Z=\mathbb{P}_{\mathbb{P}^{n}}(\mathcal{O}_{\mathbb{P}^{n}}^{\oplus s}\oplus\mathcal{O}_{\mathbb{P}^{n}}(1)). By Theorem 8.4(4), we get the above diagram, where π\pi is the induced s\mathbb{P}^{s}-bundle and σ\sigma is a birational morphism. It is enough to show that σ:ZY\sigma:Z\to Y is the blowup along the vertex LL. Note that LL is scheme-theoretically equal to the base scheme of the linear system that induces the dominant rational map

Y=n+sn=X,[x0::xn:y1::ys][x0::xn].Y=\mathbb{P}^{n+s}\dashrightarrow\mathbb{P}^{n}=X,\qquad[x_{0}:\cdots:x_{n}:y_{1}:\cdots:y_{s}]\mapsto[x_{0}:\cdots:x_{n}].

Then the blowup σ:ZY\sigma^{\prime}:Z^{\prime}\to Y along LL coincides with the resolution of its indeterminacies. By the universal property of blowups [Har77, Ch. II, Proposition 7.14], we obtain a factorisation σ:ZσZ𝜃Y\sigma:Z\xrightarrow{\sigma^{\prime}}Z^{\prime}\xrightarrow{\theta}Y. Then θ:ZZ\theta:Z\to Z^{\prime} is an isomorphism, because θ:ZZ\theta:Z\to Z^{\prime} is a birational morphism of normal projective varieties satisfying ρ(Z)=ρ(Z)\rho(Z)=\rho(Z^{\prime}).

(II) Let us treat the general case. Set X~:=n,Y~:=n+s,\widetilde{X}:=\mathbb{P}^{n},\widetilde{Y}:=\mathbb{P}^{n+s}, and Z~:=n(𝒪ns𝒪n(1))\widetilde{Z}:=\mathbb{P}_{\mathbb{P}^{n}}(\mathcal{O}_{\mathbb{P}^{n}}^{\oplus s}\oplus\mathcal{O}_{\mathbb{P}^{n}}(1)). By the case (I), we have induced morphisms

Z~=n(𝒪ns𝒪n(1)){\widetilde{Z}=\mathbb{P}_{\mathbb{P}^{n}}(\mathcal{O}_{\mathbb{P}^{n}}^{\oplus s}\oplus\mathcal{O}_{\mathbb{P}^{n}}(1))}X~=n{\widetilde{X}=\mathbb{P}^{n}}Y~=n+s,{\widetilde{Y}=\mathbb{P}^{n+s},}π~\scriptstyle{\widetilde{\pi}}σ~\scriptstyle{\widetilde{\sigma}}

where π~\widetilde{\pi} is the induced s\mathbb{P}^{s}-bundle and σ~\widetilde{\sigma} is the blowup along LL. Note that we have induced closed embeddings XX~X\subset\widetilde{X} and YY~Y\subset\widetilde{Y}. We also have a natural closed embedding ZZ~Z\subset\widetilde{Z}, because there is the following cartesian diagram

Z{Z}Z~{\widetilde{Z}}X{X}X~,{\widetilde{X},}π\scriptstyle{\pi}π~\scriptstyle{\widetilde{\pi}}

where the vertical arrows are the induced s\mathbb{P}^{s}-bundles and the horizontal ones are closed immersions. In particular, Z=π~1(X)Z=\widetilde{\pi}^{-1}(X). For the blowup ZZ^{\prime} of YY along the vertex LL (note that LYL\subset Y), we have a closed embeddding ZZ~Z^{\prime}\subset\widetilde{Z} [Har77, Ch. II, Corollary 7.15]. By construction, we have a dominant rational map YXY\dashrightarrow X compatible with Y~X~\widetilde{Y}\dashrightarrow\widetilde{X}, i.e., the induced morphisms

YLXandY~LX~Y\setminus L\to X\quad\text{and}\quad\widetilde{Y}\setminus L\to\widetilde{X}
which are given by[x0::xn:y1::ys][x0::xn].\text{which are given by}\quad[x_{0}:\cdots:x_{n}:y_{1}:\cdots:y_{s}]\mapsto[x_{0}:\cdots:x_{n}].

Hence we get the morphism π:ZX\pi^{\prime}:Z^{\prime}\to X induced by ZZ~X~Z^{\prime}\hookrightarrow\widetilde{Z}\to\widetilde{X}. For every closed point xXx\in X, it holds that π1(x)=π~1(x)\pi^{\prime-1}(x)=\widetilde{\pi}^{-1}(x), because we have dimπ1(x)dimZdimX=dimYdimX=s\dim\pi^{\prime-1}(x)\geq\dim Z^{\prime}-\dim X=\dim Y-\dim X=s and a scheme-theoretic inclusion π1(x)π~1(x)s\pi^{\prime-1}(x)\subset\widetilde{\pi}^{-1}(x)\simeq\mathbb{P}^{s}. Therefore, we get the following set-theoretic equation:

Z(k)=xX(k)π~1(x)=Z(k),Z(k)=\bigcup_{x\in X(k)}\widetilde{\pi}^{-1}(x)=Z^{\prime}(k),

where Z(k)Z(k) (resp. Z(k)Z^{\prime}(k)) denotes the subset of ZZ (resp. ZZ^{\prime}) consisting of all the closed points. Since both ZZ and ZZ^{\prime} are reduced closed subschemes of Z~\widetilde{Z}, we get a scheme-theoretic equation Z=ZZ=Z^{\prime}. ∎

8.2. Cones over the Veronese surface

The purpose of this subsection is to prove that the dd-th cone over the Veronese surface does not have a smooth prime divisor when d2d\geq 2 (Theorem 8.10). Throughout this subsection, we shall use the following notation.

Notation 8.7.

We work over an algebraically closed field kk. Let T5T\subset\mathbb{P}^{5} be the Veronese surface, i.e., TT is the image of the Veronese embedding:

ν:25,[x:y:z][x2:y2:z2:xy:yz:zx].\nu:\mathbb{P}^{2}\hookrightarrow\mathbb{P}^{5},\qquad[x:y:z]\mapsto[x^{2}:y^{2}:z^{2}:xy:yz:zx].

Fix an integer d1d\geq 1. Let Vd+5V\subset\mathbb{P}^{d+5} be the dd-th cone over TT. We then obtain the following diagram

W:=2(𝒪2d𝒪2(2)){W:=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}^{\oplus{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}d}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))}T2{T\simeq\mathbb{P}^{2}}V,{V,}π\scriptstyle{\pi}σ\scriptstyle{\sigma}

where π\pi denotes the projection and σ{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma} is the blowup along the vertex (Lemma 8.6). We set

a0:=0,a1:=0,,ad1=0,ad:=2,andE:=𝒪2d𝒪2(2)=i=0d𝒪2(ai).a_{0}:=0,\,\,a_{1}:=0,\,\,...,\,\,a_{d-1}=0,\,\,a_{d}:=2,\quad\text{and}\quad E:=\mathcal{O}_{\mathbb{P}^{2}}^{\oplus d}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)=\bigoplus_{i=0}^{d}\mathcal{O}_{\mathbb{P}^{2}}(a_{i}).

We may use Notation 8.1 for n:=2n:=2. In particular,

  1. (1)

    HH is a line on T=2T=\mathbb{P}^{2} and F:=πHF:=\pi^{*}H.

  2. (2)

    σ:WV\sigma:W\to V is a birational morphism with Ex(σ)=Dd{\operatorname{Ex}}({\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma})=D_{d} (Proposition 8.5).

  3. (3)

    For each 0id0\leq i\leq d, we have a section Γi\Gamma_{i} of π\pi and a d1\mathbb{P}^{d-1}-bundle DiWD_{i}\subset W over T=2T=\mathbb{P}^{2}, which are mutually disjoint.

For every \ell\in\mathbb{Z}, let 𝒪W():=𝒪(E)()\mathcal{O}_{W}(\ell):=\mathcal{O}_{\mathbb{P}(E)}(\ell). We set FV:=σFF_{V}:=\sigma_{*}F. Note that Dd2×d1D_{d}\simeq\mathbb{P}^{2}\times\mathbb{P}^{d-1} and the induced morphism π|Dd:DdT=2\pi|_{D_{d}}:D_{d}\to T=\mathbb{P}^{2} coincides with the first projection. Fix a closed point Qd1Q\in\mathbb{P}^{d-1} and a line ζ\zeta on 2×{Q}\mathbb{P}^{2}\times\{Q\}. Hence

ζ2×{Q}2×d1=Dd.\zeta\subset\mathbb{P}^{2}\times\{Q\}\subset\mathbb{P}^{2}\times\mathbb{P}^{d-1}=D_{d}.

We now summarise some basic properties in Proposition 8.8 and Proposition 8.9.

Proposition 8.8.

We use Notation 8.7. Then the following hold.

  1. (1)

    ClV=FV{\operatorname{Cl}}\,V=\mathbb{Z}F_{V}.

  2. (2)

    Fζ=1F\cdot\zeta=1 and Ddζ=2D_{d}\cdot\zeta=-2.

  3. (3)

    σFV=F+12Dd\sigma^{*}F_{V}=F+\frac{1}{2}D_{d}. Furthermore, FVF_{V} is not Cartier.

  4. (4)

    For every Weil divisor BB on VV, 2B2B is Cartier.

In particular, it holds that

CaClV=(2FV){\rm CaCl}\,V=\mathbb{Z}(2F_{V})

for the subgroup CaClV{\rm CaCl}\,V of ClV{\operatorname{Cl}}\,V consisting of the linear equivalence classes of the Cartier divisors.

Proof.

The assertion (1) follows from ClW=Dd+F{\operatorname{Cl}}\,W=\mathbb{Z}D_{d}+\mathbb{Z}F and

ClV=σ(ClW)=σDd+σF=FV.{\operatorname{Cl}}\,V={\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma}_{*}({\operatorname{Cl}}\,W)=\mathbb{Z}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma}_{*}D_{d}+\mathbb{Z}{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\sigma}_{*}F=\mathbb{Z}F_{V}.

Let us show (2). We have

Fζ=F|2×{Q}ζ=(line)(line)=1.F\cdot\zeta=F|_{\mathbb{P}^{2}\times\{Q\}}\cdot\zeta=(\text{line})\cdot(\text{line})=1.

Since σ(ζ)\sigma(\zeta) is a point, it follows from Lemma 8.3 that

0=𝒪W(1)ζ=(Dd+2F)ζ.0=\mathcal{O}_{W}(1)\cdot\zeta=(D_{d}+2F)\cdot\zeta.

In particular, Ddζ=2D_{d}\cdot\zeta=-2. Thus (2) holds.

Let us show (3). Since we can uniquely write F+cDd=σFVF+cD_{d}=\sigma^{*}F_{V} for some c0c\in\mathbb{Q}_{\geq 0}, the equalities (Dd+2F)ζ=0(D_{d}+2F)\cdot\zeta=0 and σFVζ=0\sigma^{*}F_{V}\cdot\zeta=0, together with Ddζ0{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}D_{d}}\cdot\zeta\neq 0, imply σFV=F+12Dd\sigma^{*}F_{V}=F+\frac{1}{2}D_{d}. In particular, FVF_{V} is not Cartier. Thus (3) holds.

Let us show (4). Since σ:WV\sigma:W\to V is given by the complete linear system |𝒪W(1)||\mathcal{O}_{W}(1)|, there exists an ample Cartier divisor HVH_{V} on VV such that 𝒪W(1)σ𝒪V(HV)𝒪V(σHV)\mathcal{O}_{W}(1)\simeq\sigma^{*}\mathcal{O}_{V}(H_{V})\simeq\mathcal{O}_{V}(\sigma^{*}H_{V}). By

σHVDd+2F=σ(2FV),\sigma^{*}H_{V}\sim D_{d}+2F=\sigma^{*}(2F_{V}),

we obtain HV=σσHVσσ(2FV)=2FVH_{V}=\sigma_{*}\sigma^{*}H_{V}\sim\sigma_{*}\sigma^{*}(2F_{V})=2F_{V}. Hence 2FV2F_{V} is Cartier. Thus (4) holds by (1). ∎

Proposition 8.9.

We use Notation 8.7. Then the following hold.

  1. (1)

    KW+i=0dDiπK2K_{W}+\sum_{i=0}^{d}D_{i}\sim\pi^{*}K_{\mathbb{P}^{2}}.

  2. (2)

    KW(d+1)Dd+(2d+3)F-K_{W}\sim(d+1)D_{d}+(2d+3)F.

  3. (3)

    KV(2d+3)FV-K_{V}\sim(2d+3)F_{V}.

  4. (4)

    KW=μKV+12DdK_{W}=\mu^{*}K_{V}+\frac{1}{2}D_{d}.

  5. (5)

    VV is \mathbb{Q}-facotiral and terminal.

Proof.

The assertion (1) follows from Theorem 8.4. By 𝒪W(1)D0Dd1Dd+2F\mathcal{O}_{W}(1)\sim D_{0}\sim\cdots\sim D_{d-1}\sim D_{d}+2F (Lemma 8.3), we obtain

KW+(d+1)Dd+2dFKW+i=0dDiπK23F,K_{W}+(d+1)D_{d}+2dF\sim K_{W}+\sum_{i=0}^{d}D_{i}\sim\pi^{*}K_{\mathbb{P}^{2}}\sim-3F,

which implies

KW(d+1)Dd+(2d+3)FandKVσ((d+1)Dd+(2d+3)F)=(2d+3)FV.-K_{W}\sim(d+1)D_{d}+(2d+3)F\quad\text{and}\quad-K_{V}\sim\sigma_{*}((d+1)D_{d}+(2d+3)F)=(2d+3)F_{V}.

Thus (2) and (3) hold.

Let us show (4) and (5). There exists aa\in\mathbb{Q} such that

KW=σKV+aDd.K_{W}=\sigma^{*}K_{V}+aD_{d}.

Recall that we have Fζ=1F\cdot\zeta=1 and Ddζ=2D_{d}\cdot\zeta=-2 (Proposition 8.8). These, together with (2), imply

KWζ=((d+1)Dd+(2d+3)F)ζ=2(d+1)+(2d+3)=1.-K_{W}\cdot\zeta=((d+1)D_{d}+(2d+3)F)\cdot\zeta=-2(d+1)+(2d+3)=1.

By σKVζ=0\sigma^{*}K_{V}\cdot\zeta=0, we get a=1/2a=1/2, and hence (4) holds. In particular, VV is terminal. Note that VV is \mathbb{Q}-factorial by Proposition 8.5. Thus (5) holds. ∎

Theorem 8.10.

Let Vd+5V\subset\mathbb{P}^{d+5} be the dd-th cone over the Veronese surface T5T\subset\mathbb{P}^{5}. If d2d\geq 2, then there exists no smooth prime divisor on VV.

Proof.

In what follows, we use Notation 8.7. Suppose that there is a smooth prime divisor XX on VV. Let us derive a contradiction. We have XbFVX\sim bF_{V} for some b>0b\in\mathbb{Z}_{>0} (Proposition 8.8). Recall that dimSingV=d11\dim{\operatorname{Sing}}\,V=d-1\geq 1.

Claim 8.11.

The following hold.

  1. (1)

    bb is odd.

  2. (2)

    SingVX{\operatorname{Sing}}\,V\subset X.

Proof of Claim 8.11.

Let us show (1). Suppose that bb is even. By XbFVX\sim bF_{V}, XX is an effective Cartier divisor (Proposition 8.8). Since XX is smooth, VV is smooth around XX. As 2X2X is an ample effective Cartier divisor, SingV{\operatorname{Sing}}\,V must be (at most) zero-dimensional, which contradicts dimSingV=d11\dim{\operatorname{Sing}}\,V=d-1\geq 1. Hence (1) holds.

Let us show (2). Suppose SingVX{\operatorname{Sing}}\,V\not\subset X. Fix vSingVv\in{\operatorname{Sing}}\,V with vXv\not\in X. By construction, we have SingVFV{\operatorname{Sing}}\,V\subset F_{V}. For an open neighbourhood UvU_{v} of vVv\in V satisfying UvX=U_{v}\cap X=\emptyset and 2FV|Uv02F_{V}|_{U_{v}}\sim 0 (cf. Proposition 8.8), it follows from XbFVX\sim bF_{V} that

0=X|UvbFV|UvFV|Uv,0=X|_{U_{v}}\sim bF_{V}|_{U_{v}}\sim F_{V}|_{U_{v}},

where the latter linear equivalence holds by (1) and 2FV|Uv02F_{V}|_{U_{v}}\sim 0. Taking the pullback by the restriction σ:σ1(Uv)Uv\sigma^{\prime}:\sigma^{-1}(U_{v})\to U_{v} of σ\sigma, we obtain

0σ(FV|Uv)=σ(FV)|σ1(Uv)=(F+12Dd)|σ1(Uv).0\sim\sigma^{\prime*}(F_{V}|_{U_{v}})=\sigma^{*}(F_{V})|_{\sigma^{-1}(U_{v})}=\left(F+\frac{1}{2}D_{d}\right)\Big{|}_{\sigma^{-1}(U_{v})}.

This contradicts Dd|σ1(Uv)D_{d}|_{\sigma^{-1}(U_{v})}\neq\emptyset. Thus (2) holds. This complete the proof of Claim 8.11. ∎

Recall that σ:WV\sigma:W\to V is the blowup along the vertex SingV{\operatorname{Sing}}\,V (Lemma 8.6). Therefore, for the proper transform Y:=σ1XY:=\sigma_{*}^{-1}X, the induced birational morphism τ:YX\tau:Y\to X is the blowup along SingV{\operatorname{Sing}}\,V [Har77, Ch. II, Corollary 7.15]. Hence also YY is a smooth projective variety. By Claim 8.11(2), we obtain

Y+αDd=σXY+\alpha D_{d}=\sigma^{*}X

for some α>0\alpha\in\mathbb{Q}_{>0}. We have α1/2\alpha\geq 1/2 by Proposition 8.8. By codimVSingV=3{\rm codim}_{V}\,{\operatorname{Sing}}\,V=3, we obtain (KV+X)|X=KX(K_{V}+X)|_{X}=K_{X}, i.e., DiffX(0)=0{\operatorname{Diff}}_{X}(0)=0 [Kol13, Proposition 4.5(1)]. This, together with KW=σKV+12DdK_{W}=\sigma^{*}K_{V}+\frac{1}{2}D_{d} (Proposition 8.9), implies

KY=σKX+(12α)(Dd|Y).K_{Y}=\sigma^{*}K_{X}+\left(\frac{1}{2}-\alpha\right)(D_{d}|_{Y}).

By 12α0\frac{1}{2}-\alpha\leq 0, this is a contradiction, because σ:YX\sigma:Y\to X is the blowup along SingV{\operatorname{Sing}}\,V, so that the coefficient of Ds|YD_{s}|_{Y} must be positive (note that the coefficient of KYσKXK_{Y}-\sigma^{*}K_{X} is uniquely determined by the negativity lemma [KM98, Lemma 3.39]). ∎

References