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 III

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

We classify primitive Fano threefolds in positive characteristic whose Picard numbers are at least two. We also classify Fano theefolds of Picard rank two.

Key words and phrases:
Fano threefolds, positive characteristic, classification.
2010 Mathematics Subject Classification:
14G17, 14E30.

1. Introduction

This article is the third part of our series of papers. In the first and second parts [TanI], [TanII], we studied Fano threefolds XX in positive characteristic with ρ(X)=1\rho(X)=1. In this paper, we classify primitive Fano threefolds XX in positive characteristic with ρ(X)2\rho(X)\geq 2. More precisely, the main theorem is as follows.

Theorem 1.1 (Theorem 5.18, Theorem 5.23, Theorem 5.34, Theorem 6.1).

Let XX be a primitive Fano threefold with ρ(X)2\rho(X)\geq 2. Then XX is isomorphic to one of the following varieties.

No. ρ(X)\rho(X) (KX)3(-K_{X})^{3} Description
2-2 22 66 a split double cover of 2×1\mathbb{P}^{2}\times\mathbb{P}^{1} with 𝒪2×1(2,1)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(2,1)
2-6 22 1212 a split double cover of WW with 2ωW1\mathcal{L}^{\otimes 2}\simeq\omega_{W}^{-1}
2-6 22 1212 a smooth divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (2,2)(2,2)
2-8 22 1414 a split double cover of V7=(𝒪2𝒪2(1))V_{7}=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)) with 2ωV71\mathcal{L}^{\otimes 2}\simeq\omega_{V_{7}}^{-1}
2-18 22 2424 a split double cover of 2×1\mathbb{P}^{2}\times\mathbb{P}^{1} with 𝒪2×1(1,1)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(1,1)
2-24 22 3030 a smooth divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,2)(1,2)
2-32 22 4848 WW
2-34 22 5454 2×1\mathbb{P}^{2}\times\mathbb{P}^{1}
2-35 22 5656 V7=(𝒪2𝒪2(1))V_{7}=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1))
2-36 22 6262 (𝒪2𝒪2(2))\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))
3-1 33 1212 a split double cover of 1×1×1\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} with 𝒪1×1×1(1,1,1)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1,1)
3-2 33 1414 a smooth member of the complete linear system |𝒪P(2)π𝒪1×1(2,3)||\mathcal{O}_{P}(2)\otimes\pi^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,3)| on the 2\mathbb{P}^{2}-bundle π:P=(𝒪1×1𝒪1×1(1,1)2)1×1\pi\colon P=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2})\to\mathbb{P}^{1}\times\mathbb{P}^{1}
3-27 33 4848 1×1×1\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}
3-31 33 5252 (𝒪1×1𝒪1×1(1,1))\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))

In the above table, we use the following notation and terminologies.

  1. (1)

    We say that f:XYf:X\to Y is a split double cover if ff is a finite surjective morphism of projective normal varieites such that 𝒪Yf𝒪X\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X} splits as an 𝒪Y\mathcal{O}_{Y}-module homomorphism and the induced field extension K(X)K(Y)K(X)\supset K(Y) is of degree two.

  2. (2)

    For a split double cover f:XYf:X\to Y, we set :=(f𝒪X/𝒪Y)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1}, which is an invertible sheaf on YY (Remark 2.2).

  3. (3)

    WW is a smooth prime divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree 𝒪(1,1)\mathcal{O}(1,1). Note that such a threefold is unique up to isomorphisms (Lemma 5.16).

We also classify Fano threefolds with ρ(X)=2\rho(X)=2. Although this case has been treated already in [Sai03], their strategies are different.

Theorem 1.2 (Section 5).

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2. Then XX is isomorphic to one of the following varieties.

No. (KX)3(-K_{X})^{3} Description Extremal rays
2-1 44 blowup of V1V_{1} along an elliptic curve of degree 11 D1+E1D_{1}+E_{1}
2-2 66 a split double cover of 2×1\mathbb{P}^{2}\times\mathbb{P}^{1} with 𝒪(2,1)\mathcal{L}\simeq\mathcal{O}(2,1) C1+D1C_{1}+D_{1}
2-3 88 blowup of V2V_{2} along an elliptic curve of degree 22 D1+E1D_{1}+E_{1}
2-4 1010 blowup of 3\mathbb{P}^{3} along a curve of genus 1010 degree 99 D1+E1D_{1}+E_{1}
2-5 1212 blowup of V3V_{3} along an elliptic curve of degree 33 D1+E1D_{1}+E_{1}
2-6 1212 a smooth divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (2,2)(2,2), or a split double cover of WW with 2ωW1\mathcal{L}^{\otimes 2}\simeq\omega_{W}^{-1} C1+C1C_{1}+C_{1}
2-7 1414 blowup of QQ along a curve of genus 55 degree 88 D1+E1D_{1}+E_{1}
2-8 1414 a split double cover of V7=(𝒪2𝒪2(1))V_{7}=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)) with 2ωV71\mathcal{L}^{\otimes 2}\simeq\omega_{V_{7}}^{-1} C1+E3orE4C_{1}+E_{3}\,{\rm or}\,E_{4}
2-9 1616 blowup of 3\mathbb{P}^{3} along a curve of genus 55 and degree 77 C1+E1C_{1}+E_{1}
2-10 1616 blowup of V4V_{4} along an elliptic curve of degree 44 D1+E1D_{1}+E_{1}
2-11 1818 blowup of V3V_{3} along a line C1+E1C_{1}+E_{1}
2-12 2020 blowup of 3\mathbb{P}^{3} along a curve of genus 33 and degree 66 E1+E1E_{1}+E_{1}
2-13 2020 blowup of QQ along a curve of genus 22 and degree 66 C1+E1C_{1}+E_{1}
2-14 2020 blowup of V5V_{5} along an elliptic curve of degree 55 D1+E1D_{1}+E_{1}
2-15 2222 blowup of 3\mathbb{P}^{3} along a curve of genus 33 and degree 66 E1+E3orE4E_{1}+E_{3}\,{\rm or}\,E_{4}
2-16 2222 blowup of V4V_{4} along a conic C1+E1C_{1}+E_{1}
2-17 2424 blowup of 3\mathbb{P}^{3} along an elliptic curve curve of degree 55 E1+E1E_{1}+E_{1}
2-18 2424 a split double cover of 2×1\mathbb{P}^{2}\times\mathbb{P}^{1} with 𝒪2×1(1,1)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(1,1) C1+D2C_{1}+D_{2}
2-19 2626 blowup of V4V_{4} along a line E1+E1E_{1}+E_{1}
2-20 2626 blowup of V5V_{5} along a cubic rational curve C1+E1C_{1}+E_{1}
2-21 2828 blowup of QQ along a rational curve of degree 44 E1+E1E_{1}+E_{1}
2-22 3030 blowup of 3\mathbb{P}^{3} along a rational curve of degree 44 E1+E1E_{1}+E_{1}
2-23 3030 blowup of QQ along an elliptic curve of degree 44 E1+E3orE4E_{1}+E_{3}\,{\rm or}\,E_{4}
2-24 3030 a smooth divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,2)(1,2) C1+C2C_{1}+C_{2}
2-25 3232 blowup of 3\mathbb{P}^{3} along an elliptic curve of degree 44 D2+E1D_{2}+E_{1}
2-26 3434 blowup of QQ along a cubic rational curve E1+E1E_{1}+E_{1}
2-27 3838 blowup of 3\mathbb{P}^{3} along a cubic rational curve C2+E1C_{2}+E_{1}
2-28 4040 blowup of 3\mathbb{P}^{3} along an elliptic curve of degree 33 E1+E5E_{1}+E_{5}
2-29 4040 blowup of QQ along a conic D2+E1D_{2}+E_{1}
2-30 4646 blowup of 3\mathbb{P}^{3} along a conic E1+E2E_{1}+E_{2}
2-31 4646 blowup of QQ along a line C2+E1C_{2}+E_{1}
2-32 4848 WW C2+C2C_{2}+C_{2}
2-33 5454 blowup of 3\mathbb{P}^{3} along a line D3+E1D_{3}+E_{1}
2-34 5454 2×1\mathbb{P}^{2}\times\mathbb{P}^{1} C2+D3C_{2}+D_{3}
2-35 5656 V7=(𝒪2𝒪2(1))V_{7}=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)) C2+E2C_{2}+E_{2}
2-36 6262 (𝒪2𝒪2(2))\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)) C2+E5C_{2}+E_{5}

In the above table, we use the following notation and terminologies in addition to (1)–(3) from Theorem 1.1.

  1. (4)

    QQ is the smooth quadric hypersurace in 4\mathbb{P}^{4}.

  2. (5)

    For 1d71\leq d\leq 7 with d6d\neq 6, let VdV_{d} be a Fano threefold of index 22 satisfying (KVd/2)3=d(-K_{V_{d}}/2)^{3}=d. Note that VdV_{d} is not determined uniquely by dd.

  3. (6)

    The centres of all the blowups are smooth curves. The degree of a curve BB on a Fano threefold YY is defined as 1rYKYB-\frac{1}{r_{Y}}K_{Y}\cdot B, where rYr_{Y} denotes the index of YY. A line (resp. conic) is a smooth rational curve of degree one (resp. two). Note that if |1rYKY||-\frac{1}{r_{Y}}K_{Y}| is very ample, then a curve of degree one (resp. two) is automatically a line (resp. conic).

For more details, see Table LABEL:table-pic2 in Section 9. The above theorems are positive-characteristic analogues of Mori–Mukai’s results in characteristic zero [MM81], [MM83]. The resulting tables are identical to those of characteristic zero. Our main strategy is the same as in characteristic zero, although we need to overcome some obstructions. For the primitive case, we shall follow the thesis by Ott [Ott], which is extensively detailed. We here point out some differences.

(I) Conic bundles. In characteristic zero, the discriminant divisor Δf\Delta_{f} of a threefold conic bundle f:XYf:X\to Y is reduced and normal crossing. However, both the conclusions fail in characteristic two. Because of this, we shall need some minor modifications of the proofs. For example, if Y=1×1Y=\mathbb{P}^{1}\times\mathbb{P}^{1} and Δf𝒪(2,0)\Delta_{f}\sim\mathcal{O}(2,0), then we immediately get a contradiction in characteristic zero, because Δf\Delta_{f} does not contain any smooth rational curve as a connected component.

(II) Double covers. In characteristic two, double covers are more complicated than those in characteristic p2p\neq 2. Fortunately, all the double covers f:XYf:X\to Y appearing in this paper can be checked to be split, i.e., 𝒪Yf𝒪X\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X} splits. Although it is hard to control general double covers in characteristic two, split double covers are quite similar to the ones in characteristic p2p\neq 2, e.g., they are explicitly given, κ(X)κ(Y)\kappa(X)\geq\kappa(Y), etc (Section 8).

Acknowledgements: This paper is based on the master thesis [Asa] of the first author, in which the primitive case of odd characteristic is carried out. The second 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] and [KM98]. 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)

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

  6. (6)

    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. A Fano threefold XX is imprimitive if there exists a Fano threefold YY and a smooth curve BB on YY such that XX is isomorphic to the blowup BlBY{\operatorname{Bl}}_{B}Y of YY along BB. We say that a Fano threefold XX is primitive if XX is not imprimitive.

2.2. Split double covers

Definition 2.1.
  1. (1)

    We say that a morphism f:XYf:X\to Y is a double cover if ff is a finite surjective morphism of normal varieties such that the induced field extension K(Y)K(X)K(Y)\subset K(X) is of degree two.

  2. (2)

    We say that a double cover f:XYf:X\to Y is split if 𝒪Yf𝒪X\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X} splits as an 𝒪Y\mathcal{O}_{Y}-module homomorphism.

Remark 2.2.

Let f:XYf:X\to Y be a double cover, where XX and YY are smooth varieties.

  1. (1)

    It is known that Coker(𝒪Yf𝒪X)=f𝒪X/𝒪Y{\operatorname{Coker}}(\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X})=f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y} is an invertible sheaf on YY [Kaw21, Lemma A.1]. We shall frequently use its inverse :=(f𝒪X/𝒪Y)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1}. We have the following exact sequence:

    0𝒪Yf𝒪X10.0\to\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X}\to\mathcal{L}^{-1}\to 0.
  2. (2)

    If p2p\neq 2, then the branched divisor DD satisfies 𝒪Y(D)2\mathcal{O}_{Y}(D)\simeq\mathcal{L}^{\otimes 2}.

Lemma 2.3.

Let f:XYf:X\to Y be a double cover of smooth projective varieties. For the invertible sheaf :=(f𝒪X/𝒪Y)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1} (Remark 2.2), it holds that ωXf(ωY)\omega_{X}\simeq f^{*}(\omega_{Y}\otimes\mathcal{L}).

Proof.

See [CD89, Proposition 0.1.3]. ∎

The following lemma gives criteria for the splitting of ff.

Lemma 2.4.

Let f:XYf:X\to Y be a double cover of smooth projective varieties. Set :=(f𝒪X/𝒪Y)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1}, which is an invertible sheaf (Remark 2.2). Assume that one of the following (1)–(3) holds.

  1. (1)

    p2p\neq 2.

  2. (2)

    H1(X,)=0H^{1}(X,\mathcal{L})=0.

  3. (3)

    Both (a) and (b) hold.

    1. (a)

      YY is FF-split, i.e., 𝒪YF𝒪Y\mathcal{O}_{Y}\to F_{*}\mathcal{O}_{Y} splits as an 𝒪Y\mathcal{O}_{Y}-module homomorphism, where F:YYF:Y\to Y denotes the absolute Frobenius morphism.

    2. (b)

      H1(Y,𝒪Y(E))=0H^{1}(Y,\mathcal{O}_{Y}(E))=0 for any effective Cartier divisor EE on YY.

Then ff is a split double cover, i.e., 𝒪Yf𝒪X\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X} splits.

Proof.

It is well known that ff is split when (1) holds. Note that 𝒪Yf𝒪X\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X} splits if and only if

0𝒪Yf𝒪X100\to\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X}\to\mathcal{L}^{-1}\to 0

splits. Since the extension class corresponding to this exact sequence is contained in H1(Y,)H^{1}(Y,\mathcal{L}), (2) implies that ff is split.

Assume (3). If ff is inseparable, then the absolute Frobenius morphism F:YYF:Y\to Y of YY factors through f:XYf:X\to Y:

F:YX𝑓Y,F:Y\to X\xrightarrow{f}Y,

and hence the splitting of 𝒪YF𝒪Y\mathcal{O}_{Y}\to F_{*}\mathcal{O}_{Y} implies the splitting of 𝒪Yf𝒪X\mathcal{O}_{Y}\to f_{*}\mathcal{O}_{X}. Therefore, we may assume that f:XYf:X\to Y is separable. In this case, the trace map Tr:f𝒪X𝒪Y{\operatorname{Tr}}:f_{*}\mathcal{O}_{X}\to\mathcal{O}_{Y} is nonzero, because the field-theoretic trace map Tr:K(X)K(Y){\operatorname{Tr}}:K(X)\to K(Y) is given by Tr(b)=b+σ(b){\operatorname{Tr}}(b)=b+\sigma(b) for the Galois involution K(X)K(X)K(X)\to K(X), and hence Tr(b)0{\operatorname{Tr}}(b)\neq 0 for bK(X)K(Y)b\in K(X)\setminus K(Y). As the composite 𝒪Y\mathcal{O}_{Y}-module homomorphism 𝒪Yf𝒪XTr𝒪Y\mathcal{O}_{Y}\hookrightarrow f_{*}\mathcal{O}_{X}\xrightarrow{{\operatorname{Tr}}}\mathcal{O}_{Y} is zero, we get a nonzero 𝒪Y\mathcal{O}_{Y}-module homomorphism 1f𝒪X/𝒪Y𝒪Y\mathcal{L}^{-1}\simeq f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y}\to\mathcal{O}_{Y}. Therefore, H0(Y,)0H^{0}(Y,\mathcal{L})\neq 0. In this case, (b) implies H1(Y,)=0H^{1}(Y,\mathcal{L})=0, i.e., (2) holds. Therefore, (3) implies that ff is split. ∎

2.3. Brauer groups

Definition 2.5 ([CTS21]*Definition 3.2.1).

For a scheme XX, the Brauer group of XX is defined by

Br(X):=He´t2(X,𝔾m),{\operatorname{Br}}(X):=H^{2}_{{\operatorname{\acute{e}t}}}(X,\mathbb{G}_{m}),

where He´t2()H^{2}_{{\operatorname{\acute{e}t}}}(-) denotes the étale cohomology. For a ring RR, we set

Br(R):=Br(SpecR)=He´t2(SpecR,𝔾m).{\operatorname{Br}}(R):={\operatorname{Br}}({\operatorname{Spec}}\,R)=H^{2}_{{\operatorname{\acute{e}t}}}({\operatorname{Spec}}\,R,\mathbb{G}_{m}).
Proposition 2.6.

The following hold.

  1. (1)

    If YY is a regular noetherian integral scheme, then we have an injective group homorphism Br(Y)Br(K(Y)){\operatorname{Br}}(Y)\hookrightarrow{\operatorname{Br}}(K(Y)).

  2. (2)

    If CC is a smooth curve over kk, then Br(C)=Br(K(C))=0{\operatorname{Br}}(C)={\operatorname{Br}}(K(C))=0.

  3. (3)

    If YY is a smooth projective rational variety over kk, then Br(Y)=0{\operatorname{Br}}(Y)=0.

Proof.

The assertion (1) follows from [Gro95, Corollaire 1.10]. The assertion (2) holds by (1) and [CTS21, Theorem 1.2.14]. The assertion (3) holds by [CTS21, Corollary 6.2.11]. ∎

Proposition 2.7.

Let f:XYf:X\to Y be a projective morphism of smooth varieties. Fix r>0r\in\mathbb{Z}_{>0}. Assume that Br(Y)=0{\operatorname{Br}}(Y)=0 and any fibre of ff is isomorphic to r\mathbb{P}^{r}. Then there exists a vector bundle EE of rank r+1r+1 such that XX is isomorphic to (E)\mathbb{P}(E) over YY.

For the reader’s convenience, we include a sketch of a proof. For more details, we refer to websites [Mur] and [Mus], which the following proof is based on.

Proof.

Fix a closed point yYy\in Y and set A:=𝒪Y,yA:=\mathcal{O}_{Y,y}. For the completion A^=𝒪^Y,y\widehat{A}=\widehat{\mathcal{O}}_{Y,y}, we have an isomorphism X×YSpecA^A^rX\times_{Y}{\operatorname{Spec}}\,\widehat{A}\simeq\mathbb{P}^{r}_{\widehat{A}} over A^\widehat{A} [Ser06, Corollary 1.2.15]. By Artin’s approxximation theorem, We obtain an AhA^{h}-isomorphism X×YSpecAhAhrX\times_{Y}{\operatorname{Spec}}\,A^{h}\simeq\mathbb{P}^{r}_{A^{h}} over the henselisation Ah=𝒪Y,yhA^{h}=\mathcal{O}_{Y,y}^{h}. Therefore, there exists an étale surjective morphism YYY^{\prime}\to Y such that X×YYYrX\times_{Y}Y^{\prime}\simeq\mathbb{P}^{r}_{Y^{\prime}}.

By the exact sequence

1𝔾mGLr+1PGLr+111\to\mathbb{G}_{m}\to{\rm GL}_{r+1}\to{\rm PGL}_{r+1}\to 1

on étale topology, we obtain the following exact sequence:

He´t1(Y,GLr+1)𝛼He´t1(Y,PGLr+1)He´t2(Y,𝔾m)=Br(Y)=0.H^{1}_{{\operatorname{\acute{e}t}}}(Y,{\rm GL}_{r+1})\xrightarrow{\alpha}H^{1}_{{\operatorname{\acute{e}t}}}(Y,{\rm PGL}_{r+1})\to H^{2}_{{\operatorname{\acute{e}t}}}(Y,\mathbb{G}_{m})={\operatorname{Br}}(Y)=0.

This implies that the PGLr+1{\rm PGL}_{r+1}-torsor f:XYf:X\to Y comes from GLr+1{\rm GL}_{r+1}-torsor. Since any GLr+1{\rm GL}_{r+1}-torsor is nothing but a projective space bundle Y(E)\mathbb{P}_{Y}(E), the assertion holds by the following facts.

  • He´t1(X,GLr+1)H^{1}_{{\operatorname{\acute{e}t}}}(X,{\rm GL}_{r+1}) can be considered as a set of (Zariski) locally free sheaf of rank r+1r+1 [SGA1II, Exposé XI, Corollaire 5.3].

  • He´t1(Y,PGLr+1)H^{1}_{{\operatorname{\acute{e}t}}}(Y,{\rm PGL}_{r+1}) consists of the isomorphism classes of π:WY\pi:W\to Y such that WW is a flat proper morphism such that W×YYPYr+1W\times_{Y}Y^{\prime}\simeq P^{r+1}_{Y^{\prime}} for some étale surjective morphism YYY^{\prime}\to Y/. r+1\mathbb{P}^{r+1}.

  • α(E)=Y(E)\alpha(E)=\mathbb{P}_{Y}(E).

3. Extremal rays and Mori fibre spaces

3.1. Types of extremal rays

Definition 3.1.

Let XX be a smooth projective threefold. Let RR be a KXK_{X}-negative extremal ray of NE¯(X)\overline{{\operatorname{NE}}}(X). By [Kol91, (1.1.1)], there exists a unique morphism f:XYf\colon X\to Y, called the contraction (morphism) of RR, to a projective normal variety YY such that the following hold.

  1. (1)

    f𝒪X=𝒪Yf_{*}\mathcal{O}_{X}=\mathcal{O}_{Y}.

  2. (2)

    For any curve CC on XX, [C]R[C]\in R if and only if f(C)f(C) is a point.

Definition 3.2.

Let XX be a smooth projective threefold. Let RR be an extremal ray of NE¯(X)\overline{{\operatorname{NE}}}(X) and let f:XYf:X\to Y be the contraction of RR. We set

μR:=min{(KXC)XCisarationalcurveonXwith[C]R},\mu_{R}:=\min\left\{(-K_{X}\cdot C)_{X}\mid C\mathrm{~{}is~{}a~{}rational~{}curve~{}on~{}}X\mathrm{~{}with~{}}[C]\in R\right\},

which is called the length of an extremal ray RR. We say that \ell is an extremal rational curve if \ell is a rational curve on XX such that []R[\ell]\in R and KX=μR-K_{X}\cdot\ell=\mu_{R}.

Definition 3.3.

Let XX be a smooth projective threefold. Let RR be a KXK_{X}-negative extremal ray of NE¯(X)\overline{{\operatorname{NE}}}(X) and let f:XYf:X\to Y be the contraction of RR.

  1. (1)

    RR and ff are called of type CC if dimY=2\dim Y=2.

    • RR and ff are called of type C1C_{1} if ff is not smooth.

    • RR and ff are called of type C2C_{2} if ff is smooth.

  2. (2)

    RR and ff are called of type DD if dimY=1\dim Y=1. Let XKX_{K} be the generic fibre of ff, which is a regular projective surface over K:=K(Y)K:=K(Y).

    • RR and ff are called of type D1D_{1} if 1KXK271\leq K_{X_{K}}^{2}\leq 7.

    • RR and ff are called of type D2D_{2} if KXK2=8K_{X_{K}}^{2}=8.

    • RR and ff are called of type D3D_{3} if KXK2=9K_{X_{K}}^{2}=9.

  3. (3)

    RR and ff are called of type EE if dimY=3\dim Y=3. Set D:=Ex(f)D:={\operatorname{Ex}}(f).

    • RR and ff are called of type E1E_{1} if f(D)f(D) is a curve.

    • RR and ff are called of type E2E_{2} if f(D)f(D) is a point, D2D\simeq\mathbb{P}^{2}, and 𝒪X(D)|D𝒪D(1)\mathcal{O}_{X}(D)|_{D}\simeq\mathcal{O}_{D}(-1).

    • RR and ff are called of type E3E_{3} if f(D)f(D) is a point, DD is isomorphic to a smooth quadric surface QQ on 3\mathbb{P}^{3}, and 𝒪X(D)|D𝒪3(1)|Q\mathcal{O}_{X}(D)|_{D}\simeq\mathcal{O}_{\mathbb{P}^{3}}(-1)|_{Q}.

    • RR and ff are called of type E4E_{4} if f(D)f(D) is a point, DD is isomorphic to a singular quadric surface QQ on 3\mathbb{P}^{3}, and 𝒪X(D)|D𝒪3(1)|Q\mathcal{O}_{X}(D)|_{D}\simeq\mathcal{O}_{\mathbb{P}^{3}}(-1)|_{Q}.

    • RR and ff are called of type E5E_{5} if f(D)f(D) is a point, D2D\simeq\mathbb{P}^{2}, and 𝒪X(D)|D𝒪D(2)\mathcal{O}_{X}(D)|_{D}\simeq\mathcal{O}_{D}(-2).

  4. (4)

    RR and ff are called of type FF if dimY=0\dim Y=0.

Remark 3.4.

Let XX be a smooth projective threefold. Let RR be a KXK_{X}-negative extremal ray of NE¯(X)\overline{{\operatorname{NE}}}(X) and let f:XYf:X\to Y be the contraction of RR.

  1. (1)

    Assume that RR is of type CC. Then the following hold.

    • RR is of type C1C_{1} if and only if μR=1\mu_{R}=1.

    • RR is of type C2C_{2} if and only if μR=2\mu_{R}=2.

  2. (2)

    Assume that RR is of type DD. Then the following hold [TanII, Proposition 3.17].

    • RR is of type D1D_{1} if and only if μR=1\mu_{R}=1.

    • RR is of type D2D_{2} if and only if μR=2\mu_{R}=2.

    • RR is of type D3D_{3} if and only if μR=3\mu_{R}=3.

  3. (3)

    Assume that RR is of type EE. Then the following hold [TanII, Proposition 3.22].

    • If RR of type E1E_{1}, then μR=1\mu_{R}=1.

    • If RR of type E2E_{2}, then μR=2\mu_{R}=2.

    • If RR of type E3E_{3}, then μR=1\mu_{R}=1.

    • If RR of type E4E_{4}, then μR=1\mu_{R}=1.

    • If RR of type E5E_{5}, then μR=1\mu_{R}=1.

3.2. Del Pezzo fibrations

Proposition 3.5.

Let XX be a smooth projective threefold and let f:XYf:X\to Y be a contraction of a KXK_{X}-negative extremal ray of type DD. For K:=K(Y)K:=K(Y) and its algebraic closure K¯:=K(Y)¯\overline{K}:=\overline{K(Y)}, let XKX_{K} and XK¯X_{\overline{K}} be the generic fibre and the geometric generic fibre of ff, respectively. Then the following hold.

  1. (1)

    For a closed point yYy\in Y, the effective Cartier divisor fyf^{*}y is a prime divisor, i.e., any scheme-theoretic fibre of ff is geometrically integral.

  2. (2)

    XK¯X_{\overline{K}} is a projective canonical del Pezzo surface, i.e., XK¯X_{\overline{K}} is a projective normal surface such that VK¯V_{\overline{K}} has at worst canonical singularities and KVK¯-K_{V_{\overline{K}}} is ample.

  3. (3)

    1KXK2=KXK¯291\leq K_{X_{K}}^{2}=K_{X_{\overline{K}}}^{2}\leq 9.

  4. (4)

    PicXK¯{\operatorname{Pic}}X_{\overline{K}} is a finitely generated free abelian groups. Furthermore, PicXK{\operatorname{Pic}}X_{K}\simeq\mathbb{Z}.

  5. (5)

    Assume that KXK2=9K_{X_{K}}^{2}=9. Then the following hold.

    1. (a)

      XKK2X_{K}\simeq\mathbb{P}^{2}_{K}. In particular, general fibres are 2\mathbb{P}^{2}.

    2. (b)

      There exist Cartier divisors DD on XX and EE on YY such that KX3D+fE-K_{X}\sim 3D+f^{*}E.

  6. (6)

    Assume that KXK2=8K_{X_{K}}^{2}=8. Then the following hold.

    1. (a)

      If XKX_{K} is smooth over KK and yy is a general closed point, then both XK¯X_{\overline{K}} and XyX_{y} are isomorphic to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}.

    2. (b)

      If XKX_{K} is not smooth over KK and yy is a general closed point, then p=2p=2 and both XK¯X_{\overline{K}} and XyX_{y} are isomorphic to (1,1,2)\mathbb{P}(1,1,2).

    3. (c)

      There exist Cartier divisors DD on XX and EE on YY such that KX2D+fE-K_{X}\sim 2D+f^{*}E.

Proof.

The assertions (1)–(4) follow from [TanII, Proposition 3.5]. As for (5) and (6), see [TanII, Lemma 3.14 and Lemma 3.15]. ∎

Remark 3.6.

By [FS20, Proposition 14.7], there actually exists a del Pezzo fibration f:XYf:X\to Y such that XX is smooth and the generic fibre is not smooth.

Proposition 3.7.

Let XX be a smooth projective threefold with an extremal ray RR of type DD. Let f:XYf:X\to Y be the contraction of RR. For K:=K(Y)K:=K(Y), XKX_{K} denotes the generic fibre of ff.

  1. (1)

    If KXK2=9K_{X_{K}}^{2}=9, then ff is a 2\mathbb{P}^{2}-bundle, i.e., there exists a locally free FF on YY of rank 33 such that XX is isomorphic to Y(F)\mathbb{P}_{Y}(F) over YY. In particular, any fibre is 2\mathbb{P}^{2}.

  2. (2)

    If KXK2=8K_{X_{K}}^{2}=8, then there exists a 3\mathbb{P}^{3}-bundle (E)\mathbb{P}(E) over YY and a closed immersion j:X(E)j:X\hookrightarrow\mathbb{P}(E) such that jy(Xy)2j_{y}(X_{y})\subset\mathbb{P}^{2} is a quadric surface on k2\mathbb{P}^{2}_{k} for every point yYy\in Y.

Proof.

Let us show (1). By Proposition 3.5(5), we have KX3D+fE-K_{X}\sim 3D+f^{*}E for some Cartier divisors DD on XX and EE on YY, respectively. We use Fujita’s Δ\Delta-genus: Δ(V,L)=dimV+LdimVh0(V,L)\Delta(V,L)=\dim V+L^{\dim V}-h^{0}(V,L). It holds that

Δ(XK¯,𝒪X(D)|XK¯)=2+𝒪X(D)2h0(XK¯,𝒪X(D)|XK¯)=0.\Delta(X_{\overline{K}},\mathcal{O}_{X}(D)|_{X_{\overline{K}}})=2+\mathcal{O}_{X}(D)^{2}-h^{0}(X_{\overline{K}},\mathcal{O}_{X}(D)|_{X_{\overline{K}}})=0.

Then the assertion (1) follows from the same argument as in [Fuj75, Corollary 5.4]. Similarly, the assertion (2) holds by using Proposition 3.5(6) (cf. [Fuj75, Corollary 5.5]). ∎

3.3. Contraction morphisms

Lemma 3.8.

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

  1. (1)

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

  2. (2)

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

  3. (3)

    KXc2(X)=24-K_{X}\cdot c_{2}(X)=24.

  4. (4)

    For any effective divisor DD on XX,

    Dc2(X)=6χ(D,𝒪D)+6χ(D,𝒪X(D)|D)2D3(KX)2DD\cdot c_{2}(X)=6\chi(D,\mathcal{O}_{D})+6\chi(D,\mathcal{O}_{X}(D)|_{D})-2D^{3}-(-K_{X})^{2}\cdot D
Proof.

The assertions (1) and (2) follow from [TanI, Theorem 2.4] (cf. [Kaw21, Corollary 3.7]). The assertion (3) holds by χ(X,𝒪X)=1\chi(X,\mathcal{O}_{X})=1 and χ(X,𝒪X)=124(KXc2(X))\chi(X,\mathcal{O}_{X})=\frac{1}{24}(-K_{X}\cdot c_{2}(X)).

Let us show (4). By χ(X,𝒪X)=1\chi(X,\mathcal{O}_{X})=1, the Riemann–Roch theorem implies

χ(𝒪X(D))=\displaystyle\chi(\mathcal{O}_{X}(D))= 1+112((KX)2+c2(X))D+14(KXD2)+16D3,\displaystyle 1+\frac{1}{12}((-K_{X})^{2}+c_{2}(X))\cdot D+\frac{1}{4}(-K_{X}\cdot D^{2})+\frac{1}{6}D^{3},
χ(𝒪X(D))=\displaystyle\chi(\mathcal{O}_{X}(-D))= 1112((KX)2+c2(X))D+14(KXD2)16D3.\displaystyle 1-\frac{1}{12}((-K_{X})^{2}+c_{2}(X))\cdot D+\frac{1}{4}(-K_{X}\cdot D^{2})-\frac{1}{6}D^{3}.

Hence

χ(𝒪X(D))χ(𝒪X(D))=16((KX)2+c2(X))D+13D3.\chi(\mathcal{O}_{X}(D))-\chi(\mathcal{O}_{X}(-D))=\frac{1}{6}((-K_{X})^{2}+c_{2}(X))\cdot D+\frac{1}{3}D^{3}.

On the other hand, by the exact sequences

0𝒪X(D)𝒪X𝒪D00\longrightarrow\mathcal{O}_{X}(-D)\longrightarrow\mathcal{O}_{X}\longrightarrow\mathcal{O}_{D}\longrightarrow 0

and

0𝒪X𝒪X(D)𝒪D(D)0,0\longrightarrow\mathcal{O}_{X}\longrightarrow\mathcal{O}_{X}(D)\longrightarrow\mathcal{O}_{D}(D)\longrightarrow 0,

we have

χ(X,𝒪X(D))=χ(X,𝒪X)χ(D,𝒪D),χ(X,𝒪X(D))=χ(X,𝒪X)+χ(D,𝒪D(D)),\chi(X,\mathcal{O}_{X}(-D))=\chi(X,\mathcal{O}_{X})-\chi(D,\mathcal{O}_{D}),\quad\chi(X,\mathcal{O}_{X}(D))=\chi(X,\mathcal{O}_{X})+\chi(D,\mathcal{O}_{D}(D)),

where 𝒪D(D):=𝒪X(D)|D\mathcal{O}_{D}(D):=\mathcal{O}_{X}(D)|_{D}. Therefore, we obtain

χ(X,𝒪X(D))χ(X,𝒪X(D))=χ(D,𝒪D(D))+χ(D,𝒪D),\chi(X,\mathcal{O}_{X}(D))-\chi(X,\mathcal{O}_{X}(-D))=\chi(D,\mathcal{O}_{D}(D))+\chi(D,\mathcal{O}_{D}),

which implies

χ(D,𝒪D(D))+χ(D,𝒪D)=16((KX)2+c2(X))D+13D3,\chi(D,\mathcal{O}_{D}(D))+\chi(D,\mathcal{O}_{D})=\frac{1}{6}((-K_{X})^{2}+c_{2}(X))\cdot D+\frac{1}{3}D^{3},

as required. ∎

Proposition 3.9.

Let XX be a smooth projective threefold and let f:XYf:X\to Y be the contraction of a KXK_{X}-negative extremal ray RR of XX. Let \ell be a curve with []R[\ell]\in R. Then the following sequence

0Pic(Y)fPic(X)()0\longrightarrow{\operatorname{Pic}}(Y)\xrightarrow{f^{*}}{\operatorname{Pic}}(X)\xrightarrow{(-\cdot\ell)}\mathbb{Z}

is exact, where ()(D):=D(-\cdot\ell)(D):=D\cdot\ell for DPic(X)D\in{\operatorname{Pic}}(X).

Proof.

See [TanII, Proposition 3.12]. ∎

Corollary 3.10.

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2 and let RR be an extremal ray of NE(X){\operatorname{NE}}(X) with the corresponding contraction f:XYf\colon X\to Y. Then PicY{\operatorname{Pic}}\,Y\simeq\mathbb{Z}.

Proof.

The assertion follows from Pic(X)2{\operatorname{Pic}}(X)\simeq\mathbb{Z}^{2} (Lemma 3.8) and Proposition 3.9. ∎

Proposition 3.11.

Let XX be a smooth projective threefold and let f:XYf:X\to Y be the contraction of a KXK_{X}-negative extremal ray RR of XX. Then the following hold.

  1. (1)

    If dimY1\dim Y\neq 1, then Rif𝒪X=0R^{i}f_{*}\mathcal{O}_{X}=0 for any i>0i>0.

  2. (2)

    If XX is a Fano threefold, then Rif𝒪X=0R^{i}f_{*}\mathcal{O}_{X}=0 for any i>0i>0. Furthermore, Hj(X,𝒪X)Hj(Y,𝒪Y)H^{j}(X,\mathcal{O}_{X})\simeq H^{j}(Y,\mathcal{O}_{Y}) for any j0j\geq 0.

Proof.

(1) If dimXy1\dim X_{y}\leq 1 for any closed point yYy\in Y, then Rif𝒪X=0R^{i}f_{*}\mathcal{O}_{X}=0 holds by [Tan15, Theorem 0.5]. We may assume that dimXy2\dim X_{y}\geq 2 for some closed point yYy\in Y. In particular, dimY=0\dim Y=0 or dimY=3\dim Y=3, as we assume dimY1\dim Y\neq 1. If dimY=0\dim Y=0, then we obtain Rif𝒪X=0R^{i}f_{*}\mathcal{O}_{X}=0 by Lemma 3.8.

Assume that dimY=3\dim Y=3. By dimXy1\dim X_{y}\geq 1, f(E)f(E) is a point for E:=Ex(f)E:={\operatorname{Ex}}(f). Then we have an exact sequence for any r0r\in\mathbb{Z}_{\geq 0}:

0𝒪X((r+1)E)𝒪X(rE)𝒪X(rE)|E0.0\to\mathcal{O}_{X}(-(r+1)E)\to\mathcal{O}_{X}(-rE)\to\mathcal{O}_{X}(-rE)|_{E}\to 0.

Since 𝒪X(E)|E\mathcal{O}_{X}(-E)|_{E} is ample (Definition 3.3) and EE is a normal projective toric surface, we have that Hi(E,𝒪X(rE)|E)=0H^{i}(E,\mathcal{O}_{X}(-rE)|_{E})=0 for every i>0i>0. Therefore, we obtain surjections:

Rif𝒪X(mE))Rif𝒪X(E))Rif𝒪X.R^{i}f_{*}\mathcal{O}_{X}(-mE))\to\cdots\to R^{i}f_{*}\mathcal{O}_{X}(-E))\to R^{i}f_{*}\mathcal{O}_{X}.

By the Serre vanishing theorem, we have Rif𝒪X(mE))=0R^{i}f_{*}\mathcal{O}_{X}(-mE))=0 for some m>0m>0, as required.

(2) By (1), we may assume that dimY=1\dim Y=1. Pick general closed point Q1,,QnQ_{1},...,Q_{n} on YY. Set S1:=fQ1,,Sn:=fQnS_{1}:=f^{*}Q_{1},...,S_{n}:=f^{*}Q_{n}, and DY:=Q1++QnD_{Y}:=Q_{1}+\cdots+Q_{n}. By the Serre vanishing theorem

Hq(Y,Rif(f𝒪Y(DY)))=Hq(Y,Rif𝒪X𝒪Y(DY))=0.\displaystyle H^{q}(Y,R^{i}f_{*}(f^{*}\mathcal{O}_{Y}(D_{Y})))=H^{q}(Y,R^{i}f_{*}\mathcal{O}_{X}\otimes\mathcal{O}_{Y}(D_{Y}))=0.

for all q>0q>0 and n0n\gg 0. Hence, by the Leray spectral sequence, we have

H0(Y,(Rif𝒪X)𝒪Y(DY))Hi(X,f𝒪Y(DY)).H^{0}(Y,(R^{i}f_{*}\mathcal{O}_{X})\otimes\mathcal{O}_{Y}(D_{Y}))\simeq H^{i}(X,f^{*}\mathcal{O}_{Y}(D_{Y})).

Hence it is enough to show that Hi(X,f𝒪Y(DY))=Hi(X,𝒪X(S1++Sn))=0H^{i}(X,f^{*}\mathcal{O}_{Y}(D_{Y}))=H^{i}(X,\mathcal{O}_{X}(S_{1}+\cdots+S_{n}))=0 for every n>0n>0. We have an exact sequence

0𝒪X(=1nS)𝒪X=1n𝒪S0.0\longrightarrow\mathcal{O}_{X}(-\sum_{\ell=1}^{n}S_{\ell})\longrightarrow\mathcal{O}_{X}\longrightarrow\bigoplus_{\ell=1}^{n}\mathcal{O}_{S_{\ell}}\longrightarrow 0.

Since Q1++QnDYQ_{1}+\cdots+Q_{n}\sim D^{\prime}_{Y} for some divisor DYD^{\prime}_{Y} such that Q1,,QnSuppDYQ_{1},...,Q_{n}\not\in{\operatorname{Supp}}\,D^{\prime}_{Y}, it holds that

0𝒪X𝒪X(fDY)=1n𝒪S0.0\longrightarrow\mathcal{O}_{X}\longrightarrow\mathcal{O}_{X}(f^{*}D_{Y})\longrightarrow\bigoplus_{\ell=1}^{n}\mathcal{O}_{S_{\ell}}\longrightarrow 0.

Since each SiS_{i} is a canonical del Pezzo surface (Proposition 3.5(2)), we have Hi(S,𝒪S)=0H^{i}(S_{\ell},\mathcal{O}_{S_{\ell}})=0 for every \ell and every i>0i>0. It holds that Hi(X,𝒪X)=0H^{i}(X,\mathcal{O}_{X})=0 for any i>0i>0, which implies Hi(X,𝒪X(fDY))=0H^{i}(X,\mathcal{O}_{X}(f^{*}D_{Y}))=0 for any i>0i>0. ∎

Corollary 3.12.

Let XX be a Fano threefold with ρ(X)2\rho(X)\geq 2. For an extremal ray RR of NE(X){\operatorname{NE}}(X), let f:XYf\colon X\to Y be the contraction of RR. Let DYD_{Y} be an effective Cartier divisor on YY and set D:=fDYD:=f^{*}D_{Y}. Then we have

χ(D,𝒪D)=1χ(Y,𝒪Y(DY))andχ(D,𝒪X(D)|D)=χ(Y,𝒪Y(DY))1.\chi(D,\mathcal{O}_{D})=1-\chi(Y,\mathcal{O}_{Y}(-D_{Y}))\quad\text{and}\quad\chi(D,\mathcal{O}_{X}(D)|_{D})=\chi(Y,\mathcal{O}_{Y}(D_{Y}))-1.
Proof.

Set 𝒪D(D):=𝒪X(D)|D\mathcal{O}_{D}(D):=\mathcal{O}_{X}(D)|_{D}. We have the exact sequence

0𝒪X(D)𝒪X𝒪D0,0\longrightarrow\mathcal{O}_{X}(-D)\longrightarrow\mathcal{O}_{X}\longrightarrow\mathcal{O}_{D}\longrightarrow 0,

which yields

0𝒪X𝒪X(D)𝒪D(D)0.0\longrightarrow\mathcal{O}_{X}\longrightarrow\mathcal{O}_{X}(D)\longrightarrow\mathcal{O}_{D}(D)\longrightarrow 0.

Then the following hold:

χ(D,𝒪D)\displaystyle\chi(D,\mathcal{O}_{D}) =χ(X,𝒪X)χ(X,𝒪X(D)),\displaystyle=\chi(X,\mathcal{O}_{X})-\chi(X,\mathcal{O}_{X}(-D)),
χ(D,𝒪D(D))\displaystyle\chi(D,\mathcal{O}_{D}(D)) =χ(X,𝒪X(D))χ(X,𝒪X).\displaystyle=\chi(X,\mathcal{O}_{X}(D))-\chi(X,\mathcal{O}_{X}).

Recall that χ(X,𝒪X)=1\chi(X,\mathcal{O}_{X})=1 (Lemma 3.8). By the projection formula, we have

Rif𝒪X(D)Rif(f𝒪Y(DY))𝒪Y(DY)Rif𝒪XR^{i}f_{*}\mathcal{O}_{X}(D)\simeq R^{i}f_{*}(f^{*}\mathcal{O}_{Y}(D_{Y}))\simeq\mathcal{O}_{Y}(D_{Y})\otimes R^{i}f_{*}\mathcal{O}_{X}

for all i0i\geq 0. We obtain f𝒪X(D)𝒪Y(DY)f_{*}\mathcal{O}_{X}(D)\simeq\mathcal{O}_{Y}(D_{Y}), and Rif𝒪X(D)=0R^{i}f_{*}\mathcal{O}_{X}(D)=0 for all i>0i>0 (Proposition 3.11). Therefore, by the Leray spectral sequence, the following holds for every i0i\geq 0:

Hi(X,𝒪X(D))Hi(Y,f𝒪X(D))Hi(Y,𝒪Y(DY)).H^{i}(X,\mathcal{O}_{X}(D))\simeq H^{i}(Y,f_{*}\mathcal{O}_{X}(D))\simeq H^{i}(Y,\mathcal{O}_{Y}(D_{Y})).

Hence χ(X,𝒪X(D))=χ(Y,𝒪Y(DY))\chi(X,\mathcal{O}_{X}(D))=\chi(Y,\mathcal{O}_{Y}(D_{Y})), which implies

χ(D,𝒪D(D))=χ(Y,𝒪Y(DY))1.\chi(D,\mathcal{O}_{D}(D))=\chi(Y,\mathcal{O}_{Y}(D_{Y}))-1.

Similarly, we get χ(D,𝒪D)=1χ(Y,𝒪Y(DY))\chi(D,\mathcal{O}_{D})=1-\chi(Y,\mathcal{O}_{Y}(-D_{Y})). ∎

Lemma 3.13.

Let YY be a smooth projective threefold and let f:XYf:X\to Y be a blowup along a smooth curve BB on YY. Assume that KX-K_{X} is ample. Then (KX)3<(KY)3(-K_{X})^{3}<(-K_{Y})^{3}.

Proof.

Set E:=Ex(f)E:={\operatorname{Ex}}(f). Since KX-K_{X} is ample, we have (KX)2E>0(-K_{X})^{2}\cdot E>0. This, together with [TanII, Lemma 3.21], implies the following:

  1. (1)

    (KX)3=(KY)3+2KXB+2g(B)2(-K_{X})^{3}=(-K_{Y})^{3}+2K_{X}\cdot B+2g(B)-2.

  2. (2)

    KXB2g(B)+2=(KX)2E>0-K_{X}\cdot B-2g(B)+2=(-K_{X})^{2}\cdot E>0

By (1), it is enough to show KXB>g(B)1-K_{X}\cdot B>g(B)-1. Since KX-K_{X} is ample, this holds when g(B)1g(B)\leq 1. Hence we may assume that g(B)>1g(B)>1. In this case, the required inequality KXB>g(B)1-K_{X}\cdot B>g(B)-1 holds by (2): KXB>2g(B)2>g(B)1-K_{X}\cdot B>2g(B)-2>g(B)-1. ∎

3.4. Conic bundles

In this subsection, we recall some terminologies and results on conic bundles. For more details, we refer to [Tan-conic] and [TanII, Subsection 3.3].

Definition 3.14.

We say that f:XSf:X\to S is a conic bundle if f:XSf:X\to S is a flat projective morphism of noetherian schemes such that XsX_{s} is isomorphic to a conic on κ(s)2\mathbb{P}^{2}_{\kappa(s)}.

If XX is a smooth projective threefold and RR is an extremal ray of type CC, then its contraction f:XYf:X\to Y is a conic bundle to a smooth projective surface YY. For the definition of Δf\Delta_{f}, we refer to [TanII, 3.10].

Proposition 3.15.

Let f:XSf:X\to S be a conic bundle, where XX and SS are smooth varieties. Then the following hold.

  1. (1)

    f𝒪X=𝒪Sf_{*}\mathcal{O}_{X}=\mathcal{O}_{S}.

  2. (2)

    fωX1f_{*}\omega_{X}^{-1} is a locally free sheaf of rank 33.

  3. (3)

    ωX1\omega_{X}^{-1} is very ample over SS, and hence it defines a closed immersion ι:X(fωX1)\iota:X\hookrightarrow\mathbb{P}(f_{*}\omega_{X}^{-1}) over SS.

Proof.

See [Tan-conic, Lemma 2.5 and Proposition 2.7]. ∎

Proposition 3.16.

Let f:XSf:X\to S be a conic bundle, where XX is a smooth projective threefold and SS is a smooth projective surface. Then the following hold.

  1. (1)

    fKX2Sf_{*}K_{X}\sim-2S.

  2. (2)

    f(KX/S2)Δf-f_{*}(K_{X/S}^{2})\equiv\Delta_{f}.

  3. (3)

    For every divisor DD on SS, it holds that

    KX2fD=4KSDΔfD.K_{X}^{2}\cdot f^{*}D=-4K_{S}\cdot D-\Delta_{f}\cdot D.
Proof.

See [TanII, Proposition 3.11]. ∎

4. Basic properties of primitive Fano threefolds

4.1. Extremal rays on primitive Fano threefolds

Theorem 4.1 (Cone theorem).

Let XX be a Fano threefold. Then there exist finitely many extremal rays R1,,RnR_{1},...,R_{n} of NE(X){\operatorname{NE}}(X) such that

NE(X)=R1++Rn.{\operatorname{NE}}(X)=R_{1}+\cdots+R_{n}.

For each 1in1\leq i\leq n, there exists an extremal rational curve i\ell_{i} such that Ri=0[i]R_{i}=\mathbb{R}_{\geq 0}[\ell_{i}] (Definition 3.2).

Proof.

See [Mor82, Theorem 1.2]. ∎

4.1.1. Type C

Theorem 4.2.

Let XX be a Fano threefold. Assume that there exists a conic bundle f:XYf:X\to Y such that any fibre of ff is not smooth. Then the following hold.

  1. (1)

    p=2p=2.

  2. (2)

    One of the following holds.

    1. (a)

      Y2Y\simeq\mathbb{P}^{2} and XX is isomorphic to a prime divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,2)(1,2). Furthermore, ρ(X)=2\rho(X)=2 and XX is primitive.

    2. (b)

      Y1×1Y\simeq\mathbb{P}^{1}\times\mathbb{P}^{1} and XX is isomorphic to a prime divisor on P:=Y(𝒪(0,1)𝒪(1,0)𝒪)P:=\mathbb{P}_{Y}(\mathcal{O}(0,1)\oplus\mathcal{O}(1,0)\oplus\mathcal{O}) which is linearly equivalent to 𝒪P(1)2\mathcal{O}_{P}(1)^{\otimes 2}, where 𝒪P(1)\mathcal{O}_{P}(1) denotes the tautological bundle with respect to the 2\mathbb{P}^{2}-bundle structure PYP\to Y. Furthermore, ρ(X)=3\rho(X)=3 and XX is imprimitive.

In particular, if XX is a primitive Fano threefold and there exists an extremal ray of type CC whose contraction f:XYf:X\to Y has no smooth fibres, then (a) of (2) holds.

Proof.

See [MS03, Corollary 8 and Remark 10]. ∎

Lemma 4.3.

Let XX be a Fano threefold with an extremal ray RR of type CC. Let f:XYf:X\to Y be the contraction of RR. Then YY is a smooth rational surface.

Proof.

By Theorem 4.2, we may assume that ff is generically smooth. Note that YY is a smooth projective surface [Kol91, Main Theorem 1.1, (1.1.3.1)]. Since the geometric generic fibre Xη¯X_{\overline{\eta}} is smooth, (Xη¯,0)(X_{\overline{\eta}},0) is FF-pure. Then it follows from [Eji19, Corollary 4.10(2)] that KY-K_{Y} is big. In particular, the Kodaira dimension of YY is negative, i.e., 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 a smooth rational surface. ∎

Theorem 4.4.

Let XX be a Fano threefold and let RR be an extremal ray of type CC. Let f:XYf\colon X\to Y be the contraction of RR. Then ff is a conic bundle and YY is a smooth rational surface. Furthermore, if ff is generically smooth, then the following hold for the discriminant divisor Δf\Delta_{f}, the length of extremal ray μR\mu_{R}, and an extremal rational curve \ell of RR.

type of RR ff μR\mu_{R} \ell
C1C_{1} Δf\Delta_{f} is a nonzero effective divisor, if ff is generically smooth 11 a curve in a singular fibre
C2C_{2} Δf=0\Delta_{f}=0 and f is a 1\mathbb{P}^{1}-bundle 22 a fibre of ff
Proof.

We may assume that ff is generically smooth. Then general fibres of ff are 1\mathbb{P}^{1}.

Assume that ff is of type C1C_{1}, i.e., ff has a singular fibre. Then Δf0\Delta_{f}\neq 0. By KXf1(y)=2-K_{X}\cdot f^{-1}(y)=2, we have that KX=1-K_{X}\cdot\ell=1 and μR=1\mu_{R}=1.

Assume that ff is of type C2C_{2}, i.e., every fibre of ff is smooth. Then Δf=0\Delta_{f}=0 and any fibre is isomorphic to 1\mathbb{P}^{1}. Since YY is a smooth projective rational surface, we have Br(Y)=0{\operatorname{Br}}(Y)=0 (Proposition 2.6(2)). It follows from Proposition 2.7 that XX is isomorphic to (E)\mathbb{P}(E) over YY for some vector bundle EE of rank 22. In this case, we have μR=2\mu_{R}=2 and KX=2-K_{X}\cdot\ell=2 for any fibre \ell. ∎

Corollary 4.5.

Let XX be a Fano threefold. Let R=0[]R=\mathbb{R}_{\geq 0}[\ell] be an extremal ray of type CC with an extremal rational curve \ell. Let f:XYf\colon X\to Y be the contraction of CC. Then all the fibres of ff are numerically equivalent to (2/μR)(2/\mu_{R})\ell, where μR\mu_{R} denotes the length of the extremal ray RR.

Proof.

Assume that RR is of type C1C_{1}. By Theorem 4.4, we have μR=1\mu_{R}=1 and f1(y)2f^{-1}(y)\equiv 2\ell for any closed point yy. Note that this holds even if no fibre of ff is smooth.

Assume that RR is of type C2C_{2}. Then we have μR=2\mu_{R}=2 and f1(y)f^{-1}(y)\equiv\ell for any closed point yy. ∎

4.1.2. Type D

Theorem 4.6.

Let XX be a Fano threefold and let RR be an extremal ray of type DD. Let f:XYf\colon X\to Y be the contraction of RR. Fix a closed point yy and Xy:=f1(y)X_{y}:=f^{-1}(y) denotes the scheme-theoretic fibre over yy. Then Y1Y\simeq\mathbb{P}^{1} and XyX_{y} is a projective Goresntein surface on XX such that ωXy1\omega_{X_{y}}^{-1} is ample. Furthermore, the following hold for the length of extremal ray μR\mu_{R} and an extremal rational curve \ell of RR.

type of RR ff μR\mu_{R} \ell
D1D_{1} 1KXy271\leq K_{X_{y}}^{2}\leq 7 11
D2D_{2} KXy2=8K_{X_{y}}^{2}=8 22 a line on a fibre
D3D_{3} KXy2=9K_{X_{y}}^{2}=9 33 a line on a fibre
Proof.

By Proposition 3.11, we obtain Y1Y\simeq\mathbb{P}^{1}. The remaining assertions follow from Remark 3.4 and Proposition 3.5. ∎

4.1.3. Type E

Lemma 4.7.

Let XX be a Fano threefold and let f:XYf:X\to Y be the contraction of an extremal ray RR of type E1E_{1}. Note that C:=f(D)C:=f(D) is a smooth curve on YY, YY is a smooth projective threefold, and ff coincides with the blowup along CC. Assume that KY-K_{Y} is not ample. Then the following hold.

  1. (1)

    C1C\simeq\mathbb{P}^{1}.

  2. (2)

    𝒩C/Y𝒪1(1)𝒪1(1)\mathcal{N}^{*}_{C/Y}\simeq\mathcal{O}_{\mathbb{P}^{1}}(1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(1).

  3. (3)

    D1×1D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}.

  4. (4)

    𝒪X(KX)|D𝒪1×1(1,1)\mathcal{O}_{X}(K_{X})|_{D}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1) and 𝒪X(D)|D𝒪1×1(1,1)\mathcal{O}_{X}(D)|_{D}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1).

Proof.

It follows from Proposition 7.2(4) that

(4.7.1) KYC+22g(C)=(KX)2D>0,-K_{Y}\cdot C+2-2g(C)=(-K_{X})^{2}\cdot D>0,

where the inequality holds because KX-K_{X} is ample. For :=𝒩C/Y𝒪C(KY)\mathcal{E}:=\mathcal{N}^{*}_{C/Y}\otimes\mathcal{O}_{C}(-K_{Y}) and D:=()D^{\prime}:=\mathbb{P}(\mathcal{E}), it follows from [Har77, Ch. II, Lemma 7.9] that

  • π:D,φD𝜋C\pi^{\prime}\colon D^{\prime}\xrightarrow{\simeq,\varphi}D\xrightarrow{\pi}C, and

  • 𝒪D(1)φ𝒪D(1)(π)𝒪C(KY)\mathcal{O}_{D^{\prime}}(1)\simeq\varphi^{*}\mathcal{O}_{D}(1)\otimes(\pi^{\prime})^{*}\mathcal{O}_{C}(-K_{Y}).

It holds that

=𝒩C/Y𝒪C(KY)𝒩C/Y𝒪C(KC)2𝒩C/Y𝒪C(KC)𝒩C/Y.\mathcal{E}=\mathcal{N}^{*}_{C/Y}\otimes\mathcal{O}_{C}(-K_{Y})\simeq\mathcal{N}^{*}_{C/Y}\otimes\mathcal{O}_{C}(-K_{C})\otimes\bigwedge^{2}\mathcal{N}_{C/Y}\simeq\mathcal{O}_{C}(-K_{C})\otimes\mathcal{N}_{C/Y}.

where the first isomorphism holds by [Har77, Ch. II, Proposition 8.20] and the second one follows from [Har77, Ch. II, Exercise 5.16(b)]. Moreover,

𝒪D(1)\displaystyle\mathcal{O}_{D^{\prime}}(1) φ𝒪D(1)(π)𝒪C(KY)\displaystyle\simeq\varphi^{*}\mathcal{O}_{D}(1)\otimes(\pi^{\prime})^{*}\mathcal{O}_{C}(-K_{Y})
φ𝒪D(1)(φπ𝒪C(KY))\displaystyle\simeq\varphi^{*}\mathcal{O}_{D}(1)\otimes(\varphi^{*}\pi^{*}\mathcal{O}_{C}(-K_{Y}))
φ(𝒪D(1)π𝒪C(KY))\displaystyle\simeq\varphi^{*}(\mathcal{O}_{D}(1)\otimes\pi^{*}\mathcal{O}_{C}(-K_{Y}))
φ((𝒪X(D)f𝒪Y(KY))|D)\displaystyle\simeq\varphi^{*}((\mathcal{O}_{X}(-D)\otimes f^{*}\mathcal{O}_{Y}(-K_{Y}))|_{D})
φ(𝒪X(KX)|D),\displaystyle\simeq\varphi^{*}(\mathcal{O}_{X}(-K_{X})|_{D}),

where the fourth and fifth isomorphisms follow from (2) and (1) of Proposition 7.2, respectively. Since KX-K_{X} is ample, also 𝒪D(1)\mathcal{O}_{D^{\prime}}(1) is ample. Hence \mathcal{E} is an ample vector bundle on CC by [Har70, Ch. III, Theorem 1.1].

Claim.

It holds that KYC0-K_{Y}\cdot C\leq 0.

Proof of Claim.

Suppose the contrary, i.e., KYC>0-K_{Y}\cdot C>0. In order to derive a contraction, it suffices to prove that KY-K_{Y} is ample. To this end, it is enough to show (i) and (ii) below [Har70, Ch. I, Proposition 4.6].

  1. (i)

    KYZ>0-K_{Y}\cdot Z>0 for any curve ZZ on YY.

  2. (ii)

    KY-K_{Y} is semi-ample, i.e., |mKY||-mK_{Y}| is base point free for some m>0m\in\mathbb{Z}_{>0}.

Let us show (i). Fix a curve ZZ on YY. If Z=CZ=C, then KYZ>0-K_{Y}\cdot Z>0 holds by our assumption. Assume that ZCZ\neq C. Let Z~\widetilde{Z} be the proper transform of ZZ on XX. Since KX-K_{X} is ample and DD is an effective divisor with Z~SuppD\widetilde{Z}\not\subset{\operatorname{Supp}}\,D, we have

KYZ=fKYZ~=(KX+D)Z~>0.\displaystyle-K_{Y}\cdot Z=-f^{*}K_{Y}\cdot\widetilde{Z}=(-K_{X}+D)\cdot\widetilde{Z}>0.

Hence (i) holds.

Let us show (ii). By (i), KY-K_{Y} is nef. By KYC>0-K_{Y}\cdot C>0, 𝒪Y(KY)|C\mathcal{O}_{Y}(-K_{Y})|_{C} is ample. Hence π(𝒪Y(KY)|C)f(𝒪Y(KY))|D\pi^{*}(\mathcal{O}_{Y}(-K_{Y})|_{C})\simeq f^{*}(\mathcal{O}_{Y}(-K_{Y}))|_{D} is semi-ample. By f(KY)=KX+Df^{*}(-K_{Y})=-K_{X}+D, it follows from [CMM14, Theorem 3.2] that f(KY)f^{*}(-K_{Y}) is semi-ample. Since XX and YY are normal, KY-K_{Y} is semi-ample. Thus (ii) holds. This completes the proof of Claim. ∎

By (4.7.1) and Claim, we have g(C)=0g(C)=0. Hence (1) holds. Since =𝒪C(KC)𝒩C/Y\mathcal{E}=\mathcal{O}_{C}(-K_{C})\otimes\mathcal{N}_{C/Y} is a vector bundle of rank 2 on C1C\simeq\mathbb{P}^{1}, we can write

𝒪1(a)𝒪1(b).\mathcal{E}\simeq\mathcal{O}_{\mathbb{P}^{1}}(a)\oplus\mathcal{O}_{\mathbb{P}^{1}}(b).

for some a,ba,b\in\mathbb{Z}. Since \mathcal{E} is ample, we have a>0a>0 and b>0b>0 [Har70, Ch. III, Corollary 1.8]. By 𝒪D(1)φ(𝒪X(KX)|D)\mathcal{O}_{D^{\prime}}(1)\simeq\varphi^{*}(\mathcal{O}_{X}(-K_{X})|_{D}), we obtain

a+b\displaystyle a+b =degC()\displaystyle=\deg_{C}(\mathcal{E})
=c1(𝒪D(1))2\displaystyle=c_{1}(\mathcal{O}_{D^{\prime}}(1))^{2}
=c1(φ(𝒪X(KX)|D))2\displaystyle=c_{1}(\varphi^{*}(\mathcal{O}_{X}(-K_{X})|_{D}))^{2}
=(KX)2D\displaystyle=(-K_{X})^{2}\cdot D
=KYC+2g(C)\displaystyle=-K_{Y}\cdot C+2-g(C)
2,\displaystyle\leq 2,

where the second equality holds by Proposition 7.1(4) and the fifth one follows from Proposition 7.2(4). Hence we get a=b=1a=b=1 and 𝒪1(1)𝒪1(1)\mathcal{E}\simeq\mathcal{O}_{\mathbb{P}^{1}}(1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(1).

By 𝒪C(KC)𝒩C/Y\mathcal{E}\simeq\mathcal{O}_{C}(-K_{C})\otimes\mathcal{N}_{C/Y}, we have

𝒩C/YωC1(𝒪1(1)𝒪1(1))𝒪1(2)𝒪1(1)𝒪1(1).\displaystyle\mathcal{N}_{C/Y}^{*}\simeq\mathcal{E}^{*}\otimes\omega_{C}^{-1}\simeq(\mathcal{O}_{\mathbb{P}^{1}}(-1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(-1))\otimes\mathcal{O}_{\mathbb{P}^{1}}(2)\simeq\mathcal{O}_{\mathbb{P}^{1}}(1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(1).

Thus (2) holds. By Lemma 7.2(2), we obtain

D(𝒩C/Y)(𝒪1(1)𝒪1(1))1×1.D\simeq\mathbb{P}(\mathcal{N}_{C/Y}^{*})\simeq\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}}(1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(1))\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}.

Thus (3) holds.

Let us show (4). By D1×1D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}, we can write

𝒪X(D)|D𝒪1×1(m,n)\mathcal{O}_{X}(D)|_{D}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(m,n)

for some m,nm,n\in\mathbb{Z}. It holds that

2mn=c1(𝒪1×1(m,n))2=c1(𝒪X(D)|D)22mn=c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(m,n))^{2}=c_{1}(\mathcal{O}_{X}(D)|_{D})^{2}
=D3=degC(𝒩C/Y)=deg1(𝒪1(1)𝒪1(1))=2,=D^{3}=\deg_{C}(\mathcal{N}^{*}_{C/Y})=\deg_{\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}}(1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(1))=2,

where the fourth equality follows from Proposition 7.2(3). Hence we have (m,n)=(1,1)(m,n)=(1,1) or (m,n)=(1,1)(m,n)=(-1,-1). Since Dζ<0D\cdot\zeta<0 for an ff-exceptional curve ζD\zeta\subset D, we obtain (m,n)=(1,1)(m,n)=(-1,-1). By the adjunction formula 𝒪X(KX+D)|D𝒪D(KD)𝒪1×1(2,2)\mathcal{O}_{X}(K_{X}+D)|_{D}\simeq\mathcal{O}_{D}(K_{D})\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,-2), we obtain 𝒪X(KX)|D𝒪1×1(1,1)\mathcal{O}_{X}(K_{X})|_{D}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1). Thus (4) holds. ∎

Theorem 4.8.

Let XX be a primitive Fano threefold and let RR be an extremal ray of type EE. Let f:XYf\colon X\to Y be the contraction of RR. Then ff is a birational morphism to a projective normal threefold YY. Furthermore, the following hold for the length of extremal ray μR\mu_{R} and an extremal rational curve \ell of RR.

type of RR ff and DD μR\mu_{R} \ell
E1E_{1} YY is smooth, 11 P×1P\times\mathbb{P}^{1}
C=f(D)1C=f(D)\simeq\mathbb{P}^{1},
𝒩C/Y𝒪1(1)𝒪1(1),\mathcal{N}^{*}_{C/Y}\simeq\mathcal{O}_{\mathbb{P}^{1}}(1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(1),
D1×1,D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1},
𝒪D(D)𝒪1×1(1,1)\mathcal{O}_{D}(D)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)
E2E_{2} YY is smooth, 22 a line on DD
f(D)f(D) is a point,
D2D\simeq\mathbb{P}^{2},
𝒪D(D)𝒪2(1)\mathcal{O}_{D}(D)\simeq\mathcal{O}_{\mathbb{P}^{2}}(-1)
E3E_{3} f(D)f(D) is a point, 11 P×1P\times\mathbb{P}^{1} or 1×Q\mathbb{P}^{1}\times Q
D1×1D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}, on DD (P,Q1)(P,Q\in\mathbb{P}^{1})
𝒪D(D)𝒪1×1(1,1)\mathcal{O}_{D}(D)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1),
P×1P\times\mathbb{P}^{1} and 1×Q\mathbb{P}^{1}\times Q are numerically
equivalent on XX for all P,Q1P,Q\in\mathbb{P}^{1}
E4E_{4} f(D)f(D) is a point, 11 a line on DD
DD is isomorphic to
the singular quadric surface
in 3\mathbb{P}^{3},
𝒪D(D)𝒪D𝒪3(1)\mathcal{O}_{D}(D)\simeq\mathcal{O}_{D}\otimes\mathcal{O}_{\mathbb{P}^{3}}(-1)
E5E_{5} f(D)f(D) is a point, 11 a line on DD
D2D\simeq\mathbb{P}^{2},
𝒪D(D)𝒪2(2)\mathcal{O}_{D}(D)\simeq\mathcal{O}_{\mathbb{P}^{2}}(-2)

In particular, 𝒪X(D)|D\mathcal{O}_{X}(-D)|_{D} is ample.

Proof.

The columns μR\mu_{R} and \ell can be confirmed by (KX+D)=KD(K_{X}+D)\cdot\ell=K_{D}\cdot\ell. ∎

Corollary 4.9.

Let XX be a primitive Fano threefold with an extremal ray RR of type EE. Let f:XYf:X\to Y be the contraction of RR and set D:=Ex(f)D:={\operatorname{Ex}}(f). Then any effective divisor ZZ on DD is semi-ample.

Proof.

If D2D\simeq\mathbb{P}^{2} or D1×1D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}, then the assertion is clear. Hence we may assume that DD is a singular quadric surface in 3\mathbb{P}^{3} (Theorem 4.8). In this case, the assertion follows from the fact that DD is \mathbb{Q}-factorial and ρ(D)=1\rho(D)=1 (it is well known that DD is a projective toric surface which is obtained by contracting the (2)(-2)-curve on 1(𝒪1𝒪1(2))\mathbb{P}_{\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}}(2))). ∎

Corollary 4.10.

Let XX be a Fano threefold with an extremal ray RR of type E2,E3,E4E_{2},E_{3},E_{4}, or E5E_{5}. Let f:XYf:X\to Y be the contraction of RR and set D:=Ex(f)D:={\operatorname{Ex}}(f). Let g:X1g\colon X\to\mathbb{P}^{1} be a morphism. Then g(D)g(D) is a point.

Proof.

Assume that g(D)g(D) is not a point, i.e., g(D)=1g(D)=\mathbb{P}^{1}. Then there exist curves Γ1\Gamma_{1} and Γ2\Gamma_{2} on DD such that g(Γ1)g(\Gamma_{1}) is a point and g(Γ2)=1g(\Gamma_{2})=\mathbb{P}^{1}. We have g𝒪1(1)Γ1=0g^{*}\mathcal{O}_{\mathbb{P}^{1}}(1)\cdot\Gamma_{1}=0 and g𝒪1(1)Γ1>0g^{*}\mathcal{O}_{\mathbb{P}^{1}}(1)\cdot\Gamma_{1}>0. Since the numerical equivalence classes [Γ1][\Gamma_{1}] and [Γ2][\Gamma_{2}] lie on R{0}R\setminus\{0\}, we can find λ>0\lambda\in\mathbb{R}_{>0} satisfying Γ1λΓ2\Gamma_{1}\equiv\lambda\Gamma_{2}, which is a contradiction. ∎

4.2. Picard groups

Theorem 4.11.

Let XX be a Fano threefold with ρ(X)2\rho(X)\geq 2. Let RR an extremal ray of NE(X){\operatorname{NE}}(X) and let f:XYf\colon X\to Y be the contraction of RR. Pick an extremal rational curve \ell with R=0[]R=\mathbb{R}_{\geq 0}[\ell]. Then the sequence

0PicYfPicX()00\longrightarrow{\operatorname{Pic}}\,Y\stackrel{{\scriptstyle f^{*}}}{{\longrightarrow}}{\operatorname{Pic}}\,X\stackrel{{\scriptstyle(-\cdot\ell)}}{{\longrightarrow}}\mathbb{Z}\longrightarrow 0

is exact. In particular, ρ(X)=ρ(Y)+1\rho(X)=\rho(Y)+1.

Proof.

By Proposition 3.9, it is enough to find a divisor FF on XX such that F=1F\cdot\ell=1. If μR=1\mu_{R}=1, then we get KX=μR=1-K_{X}\cdot\ell=\mu_{R}=1. Therefore, we are done for the case when the type of RR is one of C1,D1,E1,E3,E4,E5C_{1},D_{1},E_{1},E_{3},E_{4},E_{5} (Remark 3.4). The remaining cases are C2,D2,D3C_{2},D_{2},D_{3}, and E2E_{2}.

Assume that RR is of type C2C_{2}. By Theorem 4.4, XX is isomorphic to a 1\mathbb{P}^{1}-bundle over YY, associated to some locally free sheaf of rank 22. Moreover, \ell is a fibre of this bundle (Theorem 4.4). Hence 𝒪X(1)=1\mathcal{O}_{X}(1)\cdot\ell=1 by Lemma 7.1(1).

Assume that RR is of type D3D_{3}. Then there exist Cartier divisors DD on XX and EE on YY such that KX3D+fEK_{X}\sim 3D+f^{*}E (Proposition 3.5(5)). Then 3=KX=3D3=-K_{X}\cdot\ell=3D\cdot\ell, i.e., D=1D\cdot\ell=1. By using Proposition 3.5(6), the same argument works for the case when RR is of type D2D_{2}.

Assume that RR is of type E2E_{2}. Then the exceptional divisor DD on XX satisfies D2,𝒪D(D)𝒪D(1)D\simeq\mathbb{P}^{2},\mathcal{O}_{D}(D)\simeq\mathcal{O}_{D}(-1), and \ell is a line on DD [TanII, Proposition 3.22]. Hence (D)=(D|D)D=(c1(𝒪2(1))c1(𝒪2(1)))2=1(-D\cdot\ell)=-(D|_{D}\cdot\ell)_{D}=-(c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(-1))\cdot c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(1)))_{\mathbb{P}^{2}}=1. ∎

Corollary 4.12.

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2. Then XX is primitive if and only if XX has no extremal ray of type E1E_{1}.

Proof.

If XX has no extremal ray of type E1E_{1}, then it is clear that XX is primitive. Suppose that XX is primitive and there exists an extremal ray RR of type E1E_{1}. It suffices to derive a contradiction. Let f:XYf\colon X\to Y be the contraction of RR. Then YY is a smooth projective threefold and ff is the blowup along a smooth curve CC. It suffices to prove that YY is Fano, because XX is a primitive Fano threefold.

By Corollary 3.10, there is an ample divisor LL on YY which generates PicY{\operatorname{Pic}}\,Y\simeq\mathbb{Z}. Hence we can write KYαL-K_{Y}\sim\alpha L for some α\alpha\in\mathbb{Z}. By the Bertini theorem, we can find a curve ZZ on YY with CZ=C\cap Z=\emptyset. In particular, its proper transform Z~\tilde{Z} on XX is disjoint from D:=Ex(f)D:={\operatorname{Ex}}(f). Then

(KX)Z~=(f(KY)D)Z~=KYZ=αLZ.\displaystyle(-K_{X})\cdot\tilde{Z}=(f^{*}(-K_{Y})-D)\cdot\tilde{Z}=-K_{Y}\cdot Z=\alpha L\cdot Z.

Since KX-K_{X} and LL are ample, we get α>0\alpha>0. Hence KYαL-K_{Y}\sim\alpha L is ample. ∎

4.3. Existence of conic bundle structures

Lemma 4.13.

Let XX be a primitive Fano threefold with ρ(X)2\rho(X)\geq 2. Then XX has an extremal ray of type CC or DD.

Proof.

Suppose the contrary. Let R1,,RnR_{1},\dots,R_{n} be all the extremal rays. We can find curve 1,,n\ell_{1},...,\ell_{n} such that

NE(X)=i=1nRi=i=1n0[i].{\operatorname{NE}}(X)=\sum_{i=1}^{n}R_{i}=\sum_{i=1}^{n}\mathbb{R}_{\geq 0}[\ell_{i}].

For each i{1,,n}i\in\{1,...,n\}, let fi:XYif_{i}:X\to Y_{i} be the contraction of RiR_{i}, which is of type EE. Set Di:=Ex(fi)D_{i}:={\operatorname{Ex}}(f_{i}). Note that each DiD_{i} is a projective normal toric surface (Theorem 4.8).

Claim.

D1Di=D_{1}\cap D_{i}=\emptyset for all 2in2\leq i\leq n.

Proof of Claim.

Suppose that D1DiD_{1}\cap D_{i}\neq\emptyset for some 2in2\leq i\leq n. Since Di|D1D_{i}|_{D_{1}} is a nonzero effective Cartier divisor on D1D_{1}, there exists a curve ZZ such that ZD1DiZ\subset D_{1}\cap D_{i}.

Since 𝒪X(D1)|D1\mathcal{O}_{X}(-D_{1})|_{D_{1}} is ample (Theorem 4.8), we have

D1Z=𝒪X(D1)Z=(𝒪X(D1)|D1)Z<0.D_{1}\cdot Z=\mathcal{O}_{X}(D_{1})\cdot Z=(\mathcal{O}_{X}(D_{1})|_{D_{1}})\cdot Z<0.

On the other hand, it follows from Corollary 4.9 that there exist m>0m\in\mathbb{Z}_{>0} and an effective Cartier divisor ZZ^{\prime} on DiD_{i} such that mZZmZ\sim Z^{\prime} and SuppZ{\operatorname{Supp}}\,Z^{\prime} does not contain any irreducible component of D1DiD_{1}\cap D_{i}. Then it holds that

D1Z=(D1|Di)Z=1m(D1|Di)Z0,D_{1}\cdot Z=(D_{1}|_{D_{i}})\cdot Z=\frac{1}{m}(D_{1}|_{D_{i}})\cdot Z^{\prime}\geq 0,

which is a contradiction. This completes the proof of Claim. ∎

Fix s>0s\in\mathbb{Z}_{>0} such that |sKX||-sK_{X}| is very ample. Pick two general members H1,H2|sKX|H_{1},H_{2}\in|-sK_{X}| such that H1H2H_{1}\cap H_{2} is a smooth curve. By NE(X)=0[1]++0[n]\mathrm{NE}(X)=\mathbb{R}_{\geq 0}[\ell_{1}]+\cdots+\mathbb{R}_{\geq 0}[\ell_{n}], we have H1H2=H1H2i=1naiiH_{1}\cdot H_{2}=H_{1}\cap H_{2}\equiv\sum_{i=1}^{n}a_{i}\ell_{i} for some a1,,an0a_{1},...,a_{n}\in\mathbb{R}_{\geq 0}. We then obtain the following contradiction:

0<D1(sKX)2=D1H1H2=a1D11+i=2naiD1i=a1D110,0<D_{1}\cdot(-sK_{X})^{2}=D_{1}\cdot H_{1}\cdot H_{2}=a_{1}D_{1}\cdot\ell_{1}+\sum_{i=2}^{n}a_{i}D_{1}\cdot\ell_{i}=a_{1}D_{1}\cdot\ell_{1}\leq 0,

where Claim implies D12==D1n=0D_{1}\cdot\ell_{2}=\cdots=D_{1}\cdot\ell_{n}=0 and the inequality D11<0D_{1}\cdot\ell_{1}<0 follows from the ampleness of 𝒪X(D1)|D1\mathcal{O}_{X}(-D_{1})|_{D_{1}} (Theorem 4.8). ∎

Proposition 4.14.

Let XX be a primitive Fano threefold with ρ(X)2\rho(X)\geq 2. Then XX has an extremal ray of type CC.

Proof.

By Lemma 4.13, we may assume that there exists an extremal ray R1R_{1} is of type DD. Then we have ρ(X)=2\rho(X)=2 (Theorem 4.11). Let R2R_{2} be the other extremal ray. For each i{1,2}i\in\{1,2\}, let fi:XYif_{i}:X\to Y_{i} be the contraction of RiR_{i}, where Y1=1Y_{1}=\mathbb{P}^{1}. Suppose that R2R_{2} is of type EE or DD. It suffices to derive a contradiction. If R2R_{2} is of type DD, then Lemma 4.15(1) leads to a contradiction.

The problem is reduced to the case when R2R_{2} is of type EE. By Corollary 4.12, R2R_{2} is not of type E1E_{1}, and hence R2R_{2} is of type E2,E3,E4,E_{2},E_{3},E_{4}, or E5E_{5}. Set D2:=Ex(f2)D_{2}:={\operatorname{Ex}}(f_{2}). It follows from Corollary 4.10 that f1(D2)f_{1}(D_{2}) is a point. Pick a curve ZZ with ZD2Z\subset D_{2}. Then both f1(Z)f_{1}(Z) and f2(Z)f_{2}(Z) are points. Hence we get [Z]R1R2={0}[Z]\in R_{1}\cap R_{2}=\{0\}, which contradicts KXZ>0-K_{X}\cdot Z>0. ∎

Lemma 4.15.

Let XX be a primitive Fano threefold. Let R1R_{1} and R2R_{2} be two distinct extremal rays. For each i{1,2}i\in\{1,2\}, let fi:XYif_{i}:X\to Y_{i} be the contraction of RiR_{i}. Then the following hold.

  1. (1)

    The induced morphism f1×f2:XY1×Y2f_{1}\times f_{2}:X\to Y_{1}\times Y_{2} is a finite morphism.

  2. (2)

    If f1f_{1} is of type CC and f2f_{2} is of type EE, then the composite morphism

    f1|D:D𝑗Xf1Y1f_{1}|_{D}:D\overset{j}{\hookrightarrow}X\xrightarrow{f_{1}}Y_{1}

    is a finite surjective morphism, where D:=Ex(f2)D:={\operatorname{Ex}}(f_{2}) is the f2f_{2}-exceptional prime divisor and j:DXj:D\hookrightarrow X denotes the induced closed immersion.

Proof.

Let us show (1). Suppose that f1×f2:XY1×Y2f_{1}\times f_{2}:X\to Y_{1}\times Y_{2} is not a finite morphism. Then there exists a curve Γ\Gamma on XX such that (f1×f2)(Γ)(f_{1}\times f_{2})(\Gamma) is a point. Then both f1(Γ)f_{1}(\Gamma) and f2(Γ)f_{2}(\Gamma) are points. However, this implies [Γ]R1R2={0}[\Gamma]\in R_{1}\cap R_{2}=\{0\}, which is a contradiction. Thus (1) holds.

Let us show (2). By dimD=dimY1\dim D=\dim Y_{1}, it suffices to show that f1|Df_{1}|_{D} is a finite morphism. Suppose that f1|D:DY1f_{1}|_{D}:D\to Y_{1} is not a finite morphism. Then there exists a curve Γ\Gamma on XX such that ΓD\Gamma\subset D and f1(Γ)f_{1}(\Gamma) is a point. If R2R_{2} is of type E2,E3,E4E_{2},E_{3},E_{4}, or E5E_{5}, then also f2(Γ)f_{2}(\Gamma) is a point. This is a contradiction: [Γ]R1R2={0}[\Gamma]\in R_{1}\cap R_{2}=\{0\}. Hence R2R_{2} is of type E1E_{1}. Since 𝒪X(D)|D\mathcal{O}_{X}(-D)|_{D} is ample, we have that DΓ<0D\cdot\Gamma<0. Let Γ\Gamma^{\prime} be a general fibre of f1:XY1f_{1}:X\to Y_{1}, so that ΓD\Gamma^{\prime}\not\subset D. Hence we get DΓ0D\cdot\Gamma^{\prime}\geq 0. On the other hand, we have ΓλΓ\Gamma^{\prime}\equiv\lambda\Gamma for some λ>0\lambda\in\mathbb{R}_{>0}, which implies DΓ0D\cdot\Gamma\geq 0. This is a contradiction. ∎

4.4. ρ(X)3\rho(X)\leq 3

Theorem 4.16.

Let XX be a primitive Fano threefold. Let RR be an extremal ray of type CC and let f:XSf:X\to S be the contraction of RR. Then S2S\simeq\mathbb{P}^{2} or S1×1S\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}.

Proof.

If ff is a wild conic bundle, then the assertion follows from Theorem 4.2. Hence we may assume that ff is generically smooth. Note that SS is a smooth projective rational surface (Theorem 4.4). By the classification of smooth projective rational surfaces, it suffices to show that there exists no curve EE on SS such that E2<0E^{2}<0. Suppose that there is a curve EE on SS such that E2<0E^{2}<0. Let us derive a contradiction.

Fix a curve CC on XX such that Cf1(E)C\subset f^{-1}(E) and f(C)=Ef(C)=E. We have

NE(X)=i=1nRi=i=1n0[i],Ri=0[i]{\operatorname{NE}}(X)=\sum_{i=1}^{n}R_{i}=\sum_{i=1}^{n}\mathbb{R}_{\geq 0}[\ell_{i}],\qquad R_{i}=\mathbb{R}_{\geq 0}[\ell_{i}]

for the extremal rays R1=R,R2,,RnR_{1}=R,R_{2},...,R_{n} and curves 1,,n\ell_{1},...,\ell_{n} on XX. We then have

Ci=1naiiC\equiv\sum_{i=1}^{n}a_{i}\ell_{i}

for some a1,,an0a_{1},...,a_{n}\in\mathbb{R}_{\geq 0}.

Now, we can write fC=bEf_{*}C=bE for some b>0b\in\mathbb{Q}_{>0}. By E2<0E^{2}<0, we get

i=1naiEfi=Ei=1naifi=EfC=bE2<0.\sum_{i=1}^{n}a_{i}E\cdot f_{*}\ell_{i}=E\cdot\sum_{i=1}^{n}a_{i}f_{*}\ell_{i}=E\cdot f_{*}C=bE^{2}<0.

Hence, possibly after permuting the indices, we have Ef2<0E\cdot f_{*}\ell_{2}<0 (note that f1=0f_{*}\ell_{1}=0). Let f2:XY2f_{2}:X\to Y_{2} be the contraction of R2R_{2}. By Ef2<0E\cdot f_{*}\ell_{2}<0 and E2<0E^{2}<0, we get f(2)=Ef(\ell_{2})=E. By the projection formula, we obtain

(4.16.1) fE2=E(f2)<0.f^{*}E\cdot\ell_{2}=E\cdot(f_{*}\ell_{2})<0.

Assume that R2R_{2} is of type CC or DD. Then the fibres of f2:XY2f_{2}:X\to Y_{2} are curves or surfaces. By f1(E)Xf^{-1}(E)\neq X, there is a curve ZZ on XX such that

  1. (i)

    Zf1(E)Z\not\subset f^{-1}(E) and

  2. (ii)

    f2(Z)f_{2}(Z) is a point.

By (ii), we can write Zc2Z\equiv c\ell_{2} for some c>0c\in\mathbb{R}_{>0}. Therefore, we obtain

fE2=c1fEZ0,f^{*}E\cdot\ell_{2}=c^{-1}f^{*}E\cdot Z\geq 0,

where the inequality holds by (i). This contradicts (4.16.1).

Assume that R2R_{2} is of type EE. Set D2:=Ex(f2)D_{2}:={\operatorname{Ex}}(f_{2}). Since D2D_{2} is covered by curves ZZ with [Z]R2=0[2][Z]\in R_{2}=\mathbb{R}_{\geq 0}[\ell_{2}], it follows from (4.16.1) that D2f1(E)D_{2}\subset f^{-1}(E). By f(D2)f(f1(E))=Ef(D_{2})\subset f(f^{-1}(E))=E, there exists a curve CC^{\prime} on D2D_{2} such that f(C)f(C^{\prime}) is a point. We get [C]R[C^{\prime}]\in R. Since 𝒪X(D2)|D2\mathcal{O}_{X}(-D_{2})|_{D_{2}} is ample (Theorem 4.8), we have

D2C<0.D_{2}\cdot C^{\prime}<0.

On the other hand, D2D_{2} is disjoint from f1(s)f^{-1}(s) for a general closed point sSs\in S, and hence we get

D2f1(s)=0.D_{2}\cdot f^{-1}(s)=0.

This is a contradiction, because we have [C]R[C^{\prime}]\in R and [f1(s)]R[f^{-1}(s)]\in R. ∎

Theorem 4.17.

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

  1. (1)

    If f:XSf:X\to S is the contraction of an extremal ray of type CC, then the following hold.

    • ρ(X)=2\rho(X)=2 if and only if S2S\simeq\mathbb{P}^{2}.

    • ρ(X)=3\rho(X)=3 if and only if S1×1S\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}.

  2. (2)

    ρ(X)3.\rho(X)\leq 3.

  3. (3)

    Assume ρ(X)=2\rho(X)=2. Then XX has an extremal ray RR of type CC and we have S2S\simeq\mathbb{P}^{2} for the contraction f:XSf:X\to S of RR.

  4. (4)

    Assume ρ(X)=3\rho(X)=3. Then the following hold.

    1. (i)

      Let RR be an extremal ray of XX. Then RR is of type CC or E1E_{1}.

    2. (ii)

      XX has two extremal rays R1R_{1} and R2R_{2} such that R1R_{1} is of type CC and R2R_{2} is of type CC or E1E_{1}.

  5. (5)

    If XX has an extremal ray of type CC whose contraction f:XYf:X\to Y is not generically smooth, then p=2p=2, ρ(X)=2\rho(X)=2, and XX is a prime divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,2)(1,2).

Proof.

The assertion (1) follows from ρ(X)=ρ(S)+1\rho(X)=\rho(S)+1 and Theorem 4.16. Then (1) implies (2) and (3) by Proposition 4.14. The assertion (5) follows from Theorem 4.2.

Let us show (4). Since (i) implies (ii) (Proposition 4.14), it suffices to prove (i). Suppose that there exists an extremal ray RR which is not of type CC nor E1E_{1}. Let us derive a contradiction. Since RR is not of type DD, the type of RR is E2,E3,E4E_{2},E_{3},E_{4}, or E5E_{5}. Let f:XYf:X\to Y be the contraction of RR. Set D:=Ex(f)D:={\operatorname{Ex}}(f). There exists an extremal ray RR^{\prime} of type CC (Proposition 4.14). Let f:X1×1f^{\prime}:X\to\mathbb{P}^{1}\times\mathbb{P}^{1} be the contraction of RR^{\prime}. For each i{1,2}i\in\{1,2\}, consider

gi:Xf1×1pri1.g_{i}:X\xrightarrow{f^{\prime}}\mathbb{P}^{1}\times\mathbb{P}^{1}\xrightarrow{{\rm pr}_{i}}\mathbb{P}^{1}.

By Corollary 4.10, we have that gi(D)g_{i}(D) is a point for each i{1,2}i\in\{1,2\}. Hence also f(D)f^{\prime}(D) is a point. However, this is a contradiction, because any fibre of ff^{\prime} is one-dimensional. ∎

5. Fano threefolds with ρ(X)=2\rho(X)=2

In this section, we classify Fano threefolds with ρ(X)=2\rho(X)=2. If XX has an extremal ray of type C and its contraction is not generically smooth (i.e., a wild conic bunlde), then XX is a prime divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,2)(1,2) (Theorem 4.2). In what follows, we always assume that the contraction of type C is generically smooth.

Notation 5.1.

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2. By Theorem 4.1, NE(X)\mathrm{NE}(X) has exactly two extremal rays R1R_{1} and R2R_{2}. For each i{1,2}i\in\{1,2\}, let

fi:XYif_{i}:X\to Y_{i}

be the contraction of RiR_{i} and let i\ell_{i} be an extremal rational curve with Ri=0[i]R_{i}=\mathbb{R}_{\geq 0}[\ell_{i}]. Set μi:=KXi\mu_{i}:=-K_{X}\cdot\ell_{i}, which is the length of RiR_{i}. We have

NE(X)=R1+R2=0[1]+0[2].{\operatorname{NE}}(X)=R_{1}+R_{2}=\mathbb{R}_{\geq 0}[\ell_{1}]+\mathbb{R}_{\geq 0}[\ell_{2}].

It follows from Corollary 3.10 that Pic(Yi){\operatorname{Pic}}(Y_{i})\simeq\mathbb{Z} for each i{1,2}i\in\{1,2\}. Let LiL_{i} be an ample Cartier divisor on YiY_{i} such that 𝒪Yi(Li)\mathcal{O}_{Y_{i}}(L_{i}) generates PicYi{\operatorname{Pic}}\,Y_{i}. Set Hi:=fiLiH_{i}:=f_{i}^{*}L_{i}.

  • If RiR_{i} is of type CC, then we assume that fif_{i} is generically smooth and Δfi\Delta_{f_{i}} denotes its discriminant divisor (Subsection 3.4).

  • If RiR_{i} is of type DD, then di:=(KX)2Hid_{i}:=(-K_{X})^{2}\cdot H_{i}. Note that 1di91\leq d_{i}\leq 9 (Theorem 4.6).

  • Assume that RiR_{i} is of type EE. Set Di:=Ex(fi)D_{i}:={\operatorname{Ex}}(f_{i}). If RiR_{i} is of type E1E_{1}, then Bi:=fi(Di)B_{i}:=f_{i}(D_{i}). If RiR_{i} is of type E1,E3,E_{1},E_{3}, or E4E_{4} (resp. E2E_{2}, resp. E5E_{5}), then let rir_{i} be the largest positive integer which divides KX+Di-K_{X}+D_{i} (resp. KX+2Di-K_{X}+2D_{i}, resp. 2KX+Di-2K_{X}+D_{i}). Note that rir_{i} coincides with the index of the Fano threefold YiY_{i} when RiR_{i} is of type E1E_{1} or E2E_{2}.

5.1. The Picard group and the canonical divisor

Lemma 5.2.

We use Notation 5.1. Then, for each i{1,2}i\in\{1,2\}, HiH_{i} is a primitive element in PicX{\operatorname{Pic}}\,X, i.e., there exists no pair (s,H)(s,H) such that s2s\in\mathbb{Z}_{\geq 2}, HH is a Cartier divisor on XX, and HisHH_{i}\sim sH.

Proof.

Suppose that there exist a Cartier divisor HH on XX and an integer s2s\geq 2 such that HisHH_{i}\sim sH. By Hii=fiLii=0H_{i}\cdot\ell_{i}=f_{i}^{*}L_{i}\cdot\ell_{i}=0, we obtain Hi=0H\cdot\ell_{i}=0. By Theorem 4.11, there exists a Cartier divisor MM on YiY_{i} such that HfiMH\sim f_{i}^{*}M. Then we get 0HisHfiLisfiM0\sim H_{i}-sH\sim f_{i}^{*}L_{i}-sf_{i}^{*}M, which implies LisML_{i}\sim sM. This contradicts the fact that 𝒪Yi(Li)\mathcal{O}_{Y_{i}}(L_{i}) is a primitive element in PicYi{\operatorname{Pic}}\,Y_{i}. ∎

Lemma 5.3.

We use Notation 5.1. Fix i{1,2}i\in\{1,2\} and assume that RiR_{i} is of type CC. Then

Hi22μii.H_{i}^{2}\equiv\frac{2}{\mu_{i}}\ell_{i}.
Proof.

By Yi2Y_{i}\simeq\mathbb{P}^{2} (Theorem 4.17(1)), we have 𝒪X(Li)𝒪2(1)\mathcal{O}_{X}(L_{i})\simeq\mathcal{O}_{\mathbb{P}^{2}}(1). Hence Li2L_{i}^{2} is a point on 2\mathbb{P}^{2}. Then Hi2=fiLi2H_{i}^{2}=f_{i}^{*}L_{i}^{2} is a fibre of fif_{i}, which is numerically equivalent to 2μii\frac{2}{\mu_{i}}\ell_{i} (Corollary 4.5). ∎

We now compute c2(X)Hic_{2}(X)\cdot H_{i} for each i{1,2}i\in\{1,2\}.

Lemma 5.4.

We use Notation 5.1. Fix i{1,2}i\in\{1,2\}. Then the following hold.

  1. (1)

    Assume that RiR_{i} is of type CC. Recall that Δfi\Delta_{f_{i}} denotes the discriminant divisor of the generically smooth conic bundle fi:XYi=2f_{i}:X\to Y_{i}=\mathbb{P}^{2}. Then c2(X)Hic_{2}(X)\cdot H_{i} satisfies the following.

    type of RiR_{i} C1C_{1} C2C_{2}
    c2(X)Hic_{2}(X)\cdot H_{i} 6+degΔfi6+\deg\Delta_{f_{i}} 66

    Moreover, degΔfi1\deg\Delta_{f_{i}}\geq 1 if RiR_{i} is of type C1C_{1}.

  2. (2)

    Assume that RiR_{i} is of type DD. Then c2(X)Hic_{2}(X)\cdot H_{i} satisfies the following.

    type of RiR_{i} D1D_{1} D2D_{2} D3D_{3}
    c2(X)Hic_{2}(X)\cdot H_{i} 12(KX)2Hi12-(-K_{X})^{2}\cdot H_{i} 44 33
  3. (3)

    Assume that RiR_{i} is of type EE. Set Di:=Ex(fi)D_{i}:={\operatorname{Ex}}(f_{i}). Recall that if RiR_{i} is of type E1,E3E_{1},E_{3}, or E4E_{4} (resp. E2E_{2}, resp. E5E_{5}), then rir_{i} is defined as the largest positive integer that divides KX+Di-K_{X}+D_{i} (resp. KX+2Di-K_{X}+2D_{i}, resp. 2KX+Di-2K_{X}+D_{i}) in PicX{\operatorname{Pic}}\,X. Then c2(X)Hic_{2}(X)\cdot H_{i} satisfies the following:

    type of RiR_{i} E1E_{1} E2E_{2} E3E_{3} or E4E_{4} E5E_{5}
    c2(X)Hic_{2}(X)\cdot H_{i} 24ri+degBi\frac{24}{r_{i}}+\deg B_{i} 24ri\frac{24}{r_{i}} 24ri\frac{24}{r_{i}} 45ri\frac{45}{r_{i}}

    where we set Bi:=fi(Di)B_{i}:=f_{i}(D_{i}) and degBi:=KYiBri\deg B_{i}:=\frac{-K_{Y_{i}}\cdot B}{r_{i}} for the case when fif_{i} is of type E1E_{1}.

Proof.

Let us show (1). Assume that RiR_{i} is of type CC. Then fi:XYi=2f_{i}\colon X\to Y_{i}=\mathbb{P}^{2} is a conic bundle. Since LiL_{i} is an ample generator of Pic2{\operatorname{Pic}}\,\mathbb{P}^{2}, we may assume that LiL_{i} is a line on Yi=2Y_{i}=\mathbb{P}^{2}. Since Hi=fiLiH_{i}=f_{i}^{*}L_{i} is an effective divisor, it follows from Lemma 3.8(4) that

c2(X)Hi=6χ(Hi,𝒪Hi)+6χ(Hi,𝒪X(Hi)|Hi)2Hi3(KX)2Hi.c_{2}(X)\cdot H_{i}=6\chi(H_{i},\mathcal{O}_{H_{i}})+6\chi(H_{i},\mathcal{O}_{X}(H_{i})|_{H_{i}})-2H_{i}^{3}-(-K_{X})^{2}\cdot H_{i}.

By Corollary 3.12, it holds that

χ(Hi,𝒪Hi)=1χ(2,𝒪2(Li))=1χ(2,𝒪2(1))=1\chi(H_{i},\mathcal{O}_{H_{i}})=1-\chi(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(-L_{i}))=1-\chi(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(-1))=1

and

χ(Hi,𝒪X(Hi)|Hi)=χ(2,𝒪2(Li))1=χ(2,𝒪2(1))1=2.\chi(H_{i},\mathcal{O}_{X}(H_{i})|_{H_{i}})=\chi(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(L_{i}))-1=\chi(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(1))-1=2.

By Li3=0L_{i}^{3}=0, we have Hi3=fLi3=0H_{i}^{3}=f^{*}L_{i}^{3}=0. Since LiL_{i} is a line on 2\mathbb{P}^{2}, it follows from Proposition 3.16 that

(KX)2Hi=4K2LiΔfiLi=12deg(Δfi).(-K_{X})^{2}\cdot H_{i}=-4K_{\mathbb{P}^{2}}\cdot L_{i}-\Delta_{f_{i}}\cdot L_{i}=12-\deg(\Delta_{f_{i}}).

To summarise, it holds that

c2(X)Hi\displaystyle c_{2}(X)\cdot H_{i} =6+120(12deg(Δfi))\displaystyle=6+12-0-(12-\deg(\Delta_{f_{i}}))
=6+deg(Δfi).\displaystyle=6+\deg(\Delta_{f_{i}}).

Thus (1) holds.

Let us show (2). Assume that RiR_{i} is of type DD. Then we have fi:XYi=1f_{i}\colon X\to Y_{i}=\mathbb{P}^{1}. Since LiL_{i} is an ample generator of Pic1{\operatorname{Pic}}\,\mathbb{P}^{1}, we may assume that LiL_{i} is a point on Yi=1Y_{i}=\mathbb{P}^{1}. Applying Lemma 3.8(4), we get

c2(X)Hi=6χ(Hi,𝒪Hi)+6χ(Hi,𝒪X(Hi)|Hi)2Hi3(KX)2Hi.c_{2}(X)\cdot H_{i}=6\chi(H_{i},\mathcal{O}_{H_{i}})+6\chi(H_{i},\mathcal{O}_{X}(H_{i})|_{H_{i}})-2H_{i}^{3}-(-K_{X})^{2}\cdot H_{i}.

By Corollary 3.12, it holds that

χ(Hi,𝒪Hi)=1χ(1,𝒪1(Li))=1χ(𝒪1(1))=1\chi(H_{i},\mathcal{O}_{H_{i}})=1-\chi(\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}}(-L_{i}))=1-\chi(\mathcal{O}_{\mathbb{P}^{1}}(-1))=1

and

χ(Hi,𝒪X(Hi)|Hi)=χ(1,𝒪1(Li))1=χ(𝒪1(1))1=1.\chi(H_{i},\mathcal{O}_{X}(H_{i})|_{H_{i}})=\chi(\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}}(L_{i}))-1=\chi(\mathcal{O}_{\mathbb{P}^{1}}(1))-1=1.

We have Hi3=(fiLi)3=0H_{i}^{3}=(f_{i}^{*}L_{i})^{3}=0. To summarise, we obtain

c2(X)Hi=6+60(KX)2Hi=12(KX)2Hi.\displaystyle c_{2}(X)\cdot H_{i}=6+6-0-(-K_{X})^{2}\cdot H_{i}=12-(-K_{X})^{2}\cdot H_{i}.

Thus (2) holds.

Let us show (3). Assume that RiR_{i} is of type E1E_{1}. By KX=fiKXi+DiK_{X}=f_{i}^{*}K_{X_{i}}+D_{i}, we obtain

KX+DifiKXifi(riLi)=riHi.-K_{X}+D_{i}\sim-f_{i}^{*}K_{X_{i}}\sim f_{i}^{*}(r_{i}L_{i})=r_{i}H_{i}.

By Lemma 3.8(3), we obtain

c2(X)Hi\displaystyle c_{2}(X)\cdot H_{i} =1ri(c2(X)(KX)+c2(X)Di)\displaystyle=\frac{1}{r_{i}}(c_{2}(X)\cdot(-K_{X})+c_{2}(X)\cdot D_{i})
=1ri(24+c2(X)Di).\displaystyle=\frac{1}{r_{i}}(24+c_{2}(X)\cdot D_{i}).

Let g(Bi)g(B_{i}) be the genus of BiB_{i}. We have di:=degBi=(1riKXi)Bid_{i}:=\deg B_{i}=(-\frac{1}{r_{i}}K_{X_{i}})\cdot B_{i}. By [TanII, Lemma 3.21(2)], we get

  • (KX)2Di=ridi2g(Bi)+2(-K_{X})^{2}\cdot D_{i}=r_{i}d_{i}-2g(B_{i})+2,

  • (KX)Di2=2g(Bi)2(-K_{X})\cdot D_{i}^{2}=2g(B_{i})-2, and

  • Di3=ridi+22g(Bi)D_{i}^{3}=-r_{i}d_{i}+2-2g(B_{i}).

It holds that

χ(Di,𝒪Di)=1g(Bi),\chi(D_{i},\mathcal{O}_{D_{i}})=1-g(B_{i}),

and

χ(Di,𝒪X(Di)|Di)=χ(Di,𝒪Di)+12(Di|Di)(Di|DiKDi)\chi(D_{i},\mathcal{O}_{X}(D_{i})|_{D_{i}})=\chi(D_{i},\mathcal{O}_{D_{i}})+\frac{1}{2}(D_{i}|_{D_{i}})\cdot(D_{i}|_{D_{i}}-K_{D_{i}})
=1g(Bi)+12(KX)Di2=0.=1-g(B_{i})+\frac{1}{2}(-K_{X})\cdot D_{i}^{2}=0.

It follows from Lemma 3.8(4) that

c2(X)Di\displaystyle c_{2}(X)\cdot D_{i} =6χ(Di,𝒪Di)+6χ(Di,𝒪X(Di)|Di)2Di3(KX)2Di\displaystyle=6\chi(D_{i},\mathcal{O}_{D_{i}})+6\chi(D_{i},\mathcal{O}_{X}(D_{i})|_{D_{i}})-2D_{i}^{3}-(-K_{X})^{2}\cdot D_{i}
=6(1g(Bi))+02(ridi+22g(Bi))(ridi2g(Bi)+2)=ridi.\displaystyle=6(1-g(B_{i}))+0-2(-r_{i}d_{i}+2-2g(B_{i}))-(r_{i}d_{i}-2g(B_{i})+2)=r_{i}d_{i}.

To summarise, we obtain

c2(X)Hi=24ri+di,c_{2}(X)\cdot H_{i}=\frac{24}{r_{i}}+d_{i},

which completes the proof for the case when RiR_{i} is of type E1E_{1}.

Assume that RiR_{i} is of type E2E_{2}. By [TanII, Proposition 3.22], we get μi=2,Di2,Di3=1,(KX)2Di=4\mu_{i}=2,D_{i}\simeq\mathbb{P}^{2},D_{i}^{3}=1,(-K_{X})^{2}\cdot D_{i}=4, and 𝒪X(Di)|Di𝒪Di(1)\mathcal{O}_{X}(D_{i})|_{D_{i}}\simeq\mathcal{O}_{D_{i}}(-1). Moreover, i\ell_{i} is a line on DiD_{i}. By KXfiKYi+2DiK_{X}\sim f_{i}^{*}K_{Y_{i}}+2D_{i}, it holds that

KX+2DifiKYifi(riLi)=riHi.-K_{X}+2D_{i}\sim f_{i}^{*}K_{Y_{i}}\sim f_{i}^{*}(r_{i}L_{i})=r_{i}H_{i}.

By Lemma 3.8(3), we obtain

c2(X)Hi\displaystyle c_{2}(X)\cdot H_{i} =1ri(c2(X)(KX)+2c2(X)Di)\displaystyle=\frac{1}{r_{i}}(c_{2}(X)\cdot(-K_{X})+2c_{2}(X)\cdot D_{i})
=1r(24+2c2(X)Di).\displaystyle=\frac{1}{r}(24+2c_{2}(X)\cdot D_{i}).

By Di2D_{i}\simeq\mathbb{P}^{2} and 𝒪X(Di)|Di𝒪2(1)\mathcal{O}_{X}(D_{i})|_{D_{i}}\simeq\mathcal{O}_{\mathbb{P}^{2}}(-1), it holds that χ(Di,𝒪Di)=1\chi(D_{i},\mathcal{O}_{D_{i}})=1 and χ(Di,𝒪Di(Di))=0\chi(D_{i},\mathcal{O}_{D_{i}}(D_{i}))=0. It follows from Lemma 3.8(4) that

c2(X)Di\displaystyle c_{2}(X)\cdot D_{i} =6χ(Di,𝒪Di)+6χ(Di,𝒪X(Di)|Di)2Di3(KX)2Di\displaystyle=6\chi(D_{i},\mathcal{O}_{D_{i}})+6\chi(D_{i},\mathcal{O}_{X}(D_{i})|_{D_{i}})-2D_{i}^{3}-(-K_{X})^{2}\cdot D_{i}
=6+024=0.\displaystyle=6+0-2-4=0.

Thus we obtain c2(X)Hi=24ric_{2}(X)\cdot H_{i}=\frac{24}{r_{i}}, which completes the proof for the case when RiR_{i} is of type E2E_{2}.

Assume that RiR_{i} is of type E3E_{3} or E4E_{4}. It follows from [TanII, Proposition 3.22] that μi=1\mu_{i}=1, (KX)2Di=2(-K_{X})^{2}\cdot D_{i}=2, Di3=2D_{i}^{3}=2, DiD_{i} is isomorphic to a possibly singular quadric surface in 3\mathbb{P}^{3}, 𝒪X(Di)|Di𝒪3(1)|Di\mathcal{O}_{X}(D_{i})|_{D_{i}}\simeq\mathcal{O}_{\mathbb{P}^{3}}(-1)|_{D_{i}}, and i\ell_{i} is a line on 3\mathbb{P}^{3} contained in DiD_{i} (Theorem 4.8). In particular, Dii=1D_{i}\cdot\ell_{i}=-1. We have KX+DifiKYifi(riLi)=riHi-K_{X}+D_{i}\sim-f_{i}^{*}K_{Y_{i}}\sim f_{i}^{*}(r_{i}L_{i})=r_{i}H_{i}. We then get

c2(X)Hi=1ri(24+c2(X)Di).c_{2}(X)\cdot H_{i}=\frac{1}{r_{i}}(24+c_{2}(X)\cdot D_{i}).

Let us compute c2(X)Dic_{2}(X)\cdot D_{i}. By Lemma 3.8(4), we have

c2(X)Di=6χ(Di,𝒪Di)+6χ(Di,𝒪X(Di)|Di)2Di3(KX)2Di.c_{2}(X)\cdot D_{i}=6\chi(D_{i},\mathcal{O}_{D_{i}})+6\chi(D_{i},\mathcal{O}_{X}(D_{i})|_{D_{i}})-2D_{i}^{3}-(-K_{X})^{2}\cdot D_{i}.

It holds that χ(Di,𝒪Di)=h0(Di,𝒪Di)=1\chi(D_{i},\mathcal{O}_{D_{i}})=h^{0}(D_{i},\mathcal{O}_{D_{i}})=1 and χ(Di,𝒪Di(Di))=χ(Di,𝒪3(1)|Di)=0\chi(D_{i},\mathcal{O}_{D_{i}}(D_{i}))=\chi(D_{i},\mathcal{O}_{\mathbb{P}^{3}}(-1)|_{D_{i}})=0. Therefore, we obtain

c2(X)Di=6+042=0,c_{2}(X)\cdot D_{i}=6+0-4-2=0,

which implies c2(X)Hi=24ric_{2}(X)\cdot H_{i}=\frac{24}{r_{i}}. This completes the proof for the case when RiR_{i} is of type E3E_{3} or E4E_{4}.

Assume that RiR_{i} is of type E5E_{5}. It follows from [TanII, Proposition 3.22] that μi=1\mu_{i}=1, Di2D_{i}\simeq\mathbb{P}^{2}, Di3=4D_{i}^{3}=4, (KX)2Di=1(-K_{X})^{2}\cdot D_{i}=1, 𝒪X(Di)|Xi𝒪2(2)\mathcal{O}_{X}(D_{i})|_{X_{i}}\simeq\mathcal{O}_{\mathbb{P}^{2}}(-2), and i\ell_{i} is a line on DiD_{i}. Hence

(2KX+Di)i=2KXi+Dii=22=0,\displaystyle(-2K_{X}+D_{i})\cdot\ell_{i}=-2K_{X}\cdot\ell_{i}+D_{i}\cdot\ell_{i}=2-2=0,

which implies 2KX+DiriHi-2K_{X}+D_{i}\sim r_{i}H_{i}. As before, we can write

c2(X)Hi=1ri(224+c2(X)Di).c_{2}(X)\cdot H_{i}=\frac{1}{r_{i}}(2\cdot 24+c_{2}(X)\cdot D_{i}).

By Lemma 3.8(4), we have

c2(X)Di=6χ(Di,𝒪Di)+6χ(Di,𝒪X(Di)|Di)2Di3(KX)2Di.c_{2}(X)\cdot D_{i}=6\chi(D_{i},\mathcal{O}_{D_{i}})+6\chi(D_{i},\mathcal{O}_{X}(D_{i})|_{D_{i}})-2D_{i}^{3}-(-K_{X})^{2}\cdot D_{i}.

It holds that χ(Di,𝒪Di)=1\chi(D_{i},\mathcal{O}_{D_{i}})=1 and χ(Di,𝒪X(Di)|Di)=0\chi(D_{i},\mathcal{O}_{X}(D_{i})|_{D_{i}})=0. We obtain

c2(X)Di=6+081=3,\displaystyle c_{2}(X)\cdot D_{i}=6+0-8-1=-3,

and hence c2(X)Hi=45ric_{2}(X)\cdot H_{i}=\frac{45}{r_{i}}. Thus (3) holds. ∎

Lemma 5.5.

We use Notation 5.1. Fix i{1,2}i\in\{1,2\}.

  1. (1)

    If RiR_{i} is of type C1C_{1}, then 7c2(X)Hi177\leq c_{2}(X)\cdot H_{i}\leq 17.

  2. (2)

    If RiR_{i} is not of type E1E_{1}, then 1c2(X)Hi241\leq c_{2}(X)\cdot H_{i}\leq 24 or c2(X)Hi=45c_{2}(X)\cdot H_{i}=45.

  3. (3)

    If RiR_{i} is neither of type E1E_{1} nor E5E_{5}, then 1c2(X)Hi241\leq c_{2}(X)\cdot H_{i}\leq 24.

Proof.

Let us show (1). Assume that RiR_{i} is of type C1C_{1}. By the same argument as in the proof of Lemma 5.4(1),

c2(X)Hi=18(KX)2Hi.c_{2}(X)\cdot H_{i}=18-(-K_{X})^{2}\cdot H_{i}.

Since KX-K_{X} is ample, we get c2(X)Hi17c_{2}(X)\cdot H_{i}\leq 17. On the other hand, it follows from Lemma 5.4(1) that

c2(X)Hi=6+degΔfic_{2}(X)\cdot H_{i}=6+\deg\Delta_{f_{i}}

and degΔfi>0\deg\Delta_{f_{i}}>0. In particular, (c2(X)Hi)7(c_{2}(X)\cdot H_{i})\geq 7. Thus (1) holds.

Let us show (2). If RiR_{i} not of type D1D_{1}, then these assertions follow from (1) and Lemma 5.4. Assume that RiR_{i} is of type D1D_{1}. Then Lemma 5.4 implies

c2(X)Hi=12(KX)2Hi.c_{2}(X)\cdot H_{i}=12-(-K_{X})^{2}\cdot H_{i}.

Since 1(KX)2Hi71\leq(-K_{X})^{2}\cdot H_{i}\leq 7, we have 5c2(X)Hi115\leq c_{2}(X)\cdot H_{i}\leq 11. Thus (2) holds. Similarly, (3) holds. ∎

Lemma 5.6.

We use Notation 5.1. Set

a:=|Pic(X)/(H1H2)|,a:=|{\operatorname{Pic}}(X)/(\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2})|,

which is the cardinality of the set Pic(X)/(H1H2){\operatorname{Pic}}(X)/(\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2}). Then the following hold.

  1. (1)

    H1H_{1} and H2H_{2} are linearly equivalent over \mathbb{R} in (PicX)/({\operatorname{Pic}}\,X)\otimes_{\mathbb{Z}}\mathbb{R}/\equiv. Moreover, a>0a\in\mathbb{Z}_{>0}.

  2. (2)

    Pic(X)/(H1H2)/(H21)/(H12){\operatorname{Pic}}(X)/(\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2})\simeq\mathbb{Z}/(H_{2}\cdot\ell_{1})\mathbb{Z}\simeq\mathbb{Z}/(H_{1}\cdot\ell_{2})\mathbb{Z}.

  3. (3)

    a=H21=H12a=H_{2}\cdot\ell_{1}=H_{1}\cdot\ell_{2}.

  4. (4)

    aKXμ2H1+μ1H2.-aK_{X}\sim\mu_{2}H_{1}+\mu_{1}H_{2}.

  5. (5)

    24a=μ2H1c2(X)+μ1H2c2(X)24a=\mu_{2}H_{1}\cdot c_{2}(X)+\mu_{1}H_{2}\cdot c_{2}(X).

Proof.

Let us show (1). Assume

λ1H1+λ2H20\lambda_{1}H_{1}+\lambda_{2}H_{2}\equiv 0

for λ1,λ2\lambda_{1},\lambda_{2}\in\mathbb{R}. By H1=f1L1H_{1}=f^{*}_{1}L_{1}, we have H11=0H_{1}\cdot\ell_{1}=0. Hence we obtain λ2H21=0\lambda_{2}H_{2}\cdot\ell_{1}=0. Since f2(1)f_{2}(\ell_{1}) is a curve, we have H21>0H_{2}\cdot\ell_{1}>0. We then get λ2=0\lambda_{2}=0, which implies λ1=0\lambda_{1}=0. Therefore, H1H_{1} and H2H_{2} are linearly independent over \mathbb{R}. Since H1H2\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2} is a free subgroup of rank 22 of PicX2{\operatorname{Pic}}\,X\simeq\mathbb{Z}^{\oplus 2}, the quotient group

Pic(X)/(H1H2){\operatorname{Pic}}(X)/(\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2})

is a finite abelian group. Thus (1) holds.

Let us show (2) and (3). By H21>0H_{2}\cdot\ell_{1}>0 and H12>0H_{1}\cdot\ell_{2}>0, (2) implies (3). We now prove (2). The composite group homomorphism

φ:PicX(1)/(H21)\varphi:{\operatorname{Pic}}\,X\stackrel{{\scriptstyle(-\cdot\ell_{1})}}{{\longrightarrow}}\mathbb{Z}\longrightarrow\mathbb{Z}/(H_{2}\cdot\ell_{1})\mathbb{Z}

is surjective, because so is each map (Theorem 4.11). It suffices to show that Ker(φ)=H1H2{\operatorname{Ker}}(\varphi)=\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2}. The inclusion Ker(φ)H1H2{\operatorname{Ker}}(\varphi)\supset\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2} is clear. It is enough to show the opposite inclusion. Take HKer(φ)H\in{\operatorname{Ker}}(\varphi). Then H1(H21)H\cdot\ell_{1}\in(H_{2}\cdot\ell_{1})\mathbb{Z}, and hence there exists bb\in\mathbb{Z} such that

H1=b(H21).H\cdot\ell_{1}=b(H_{2}\cdot\ell_{1}).

Thus we obtain HbH2Ker(1)=Im(f1)=H1H-bH_{2}\in\mathrm{Ker}(-\cdot\ell_{1})=\mathrm{Im}(f^{*}_{1})=\mathbb{Z}H_{1} by Theorem 4.11. Then we can write HcH1+bH2H\sim cH_{1}+bH_{2} for some cc\in\mathbb{Z}. Thus (2) and (3) holds.

Let us show (4). Since H1H_{1} and H2H_{2} are linearly independent over \mathbb{R} by (1), {H1,H2}\{H_{1},H_{2}\} is a basis of N1(X)\mathrm{N}^{1}(X)_{\mathbb{R}}. Then we can write

KXβ1H1+β2H2-K_{X}\equiv\beta_{1}H_{1}+\beta_{2}H_{2}

for some β1,β2\beta_{1},\beta_{2}\in\mathbb{R}. By μ1=KX1\mu_{1}=-K_{X}\cdot\ell_{1} and H21=aH_{2}\cdot\ell_{1}=a, we get

μ1=KX1=β1(H11)+β2(H21)=β2a.\displaystyle\mu_{1}=-K_{X}\cdot\ell_{1}=\beta_{1}(H_{1}\cdot\ell_{1})+\beta_{2}(H_{2}\cdot\ell_{1})=\beta_{2}a.

By symmetry, we obtain μ2=β1a\mu_{2}=\beta_{1}a. Hence (4) holds. Finally, (5) follows from (4) and (KX)c2(X)=24(-K_{X})\cdot c_{2}(X)=24 (Lemma 3.8(3)). ∎

Lemma 5.7.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1}. Then

degB1(r1μ2a)2L13\deg B_{1}\leq\left(r_{1}-\frac{\mu_{2}}{a}\right)^{2}\cdot L_{1}^{3}

for a:=|Pic(X)/(H1H2)|>0a:=|{\operatorname{Pic}}(X)/(\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2})|\in\mathbb{Z}_{>0} (cf. Lemma 5.6(1)).

Proof.

Since H1H_{1} and H2H_{2} are nef, we get H1H220H_{1}\cdot H_{2}^{2}\geq 0. By aKXμ2H1+H2-aK_{X}\sim\mu_{2}H_{1}+H_{2} (Lemma 5.6(4)) and KXD1=f1KY1r1H1K_{X}-D_{1}=f_{1}^{*}K_{Y_{1}}\sim-r_{1}H_{1}, the following holds:

0H1H22\displaystyle 0\leq H_{1}\cdot H_{2}^{2} =\displaystyle= H1(aKXμ2H1)2\displaystyle H_{1}\cdot(-aK_{X}-\mu_{2}H_{1})^{2}
=\displaystyle= H1((ar1μ2)H1aD1)2\displaystyle H_{1}\cdot((ar_{1}-\mu_{2})H_{1}-aD_{1})^{2}
=\displaystyle= (ar1μ2)2H132a(ar1μ2)H12D1+a2H1D12.\displaystyle(ar_{1}-\mu_{2})^{2}H^{3}_{1}-2a(ar_{1}-\mu_{2})H^{2}_{1}\cdot D_{1}+a^{2}H_{1}\cdot D_{1}^{2}.

Then the required inequality degB1(r1μ2a)2L13\deg B_{1}\leq\left(r_{1}-\frac{\mu_{2}}{a}\right)^{2}\cdot L_{1}^{3} follows from H13=L13,H12D1=0,H_{1}^{3}=L_{1}^{3},H_{1}^{2}\cdot D_{1}=0, and H1D12=degB1H_{1}\cdot D_{1}^{2}=-\deg B_{1}. ∎

Proposition 5.8.

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2. Let RR be an extremral ray of type E1E_{1} and let f:XYf:X\to Y be the contraction of RR. Then YY is a Fano threefold of index rY2r_{Y}\geq 2.

Proof.

Set R1:=RR_{1}:=R and a:=|Pic(X)/(H1H2)|>0a:=|{\operatorname{Pic}}(X)/(\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2})|\in\mathbb{Z}_{>0} (Lemma 5.6(1)). We use Notation 5.1. By KY=f(KX)-K_{Y}=f_{*}(-K_{X}), KY-K_{Y} is big. This, together with ρ(Y)=1\rho(Y)=1, implies that KY-K_{Y} is ample. Suppose rY=1r_{Y}=1. It is enough to derive a contradiction. Recall that |KY||-K_{Y}| is not very ample [TanII, Theorem 1.1]. By [TanI, Theorem 1.1], we obtain (KY)3{2,4}(-K_{Y})^{3}\in\{2,4\}. It follows from (KX)32(-K_{X})^{3}\in 2\mathbb{Z} and 0<(KX)3<(KY)30<(-K_{X})^{3}<(-K_{Y})^{3} (Lemma 3.13) that (KY)3=4(-K_{Y})^{3}=4 and (KX)3=2(-K_{X})^{3}=2. Let φ|KX|:Xh0(X,KX)1\varphi_{|-K_{X}|}:X\dashrightarrow\mathbb{P}^{h^{0}(X,-K_{X})-1} be the rational map induced by the complete linear system |KX||-K_{X}|. If dim(Imφ|KX|)2\dim({\operatorname{Im}}\,\varphi_{|-K_{X}|})\geq 2, then our condition ρ(X)=2\rho(X)=2 contradicts [TanI, Proposition 6.8]. Hence we may assume that dim(Imφ|KX|)=1\dim({\operatorname{Im}}\,\varphi_{|-K_{X}|})=1. In this case, the mobile part MM of |KX||-K_{X}| is base point free and the induced morphism φ|M|:X1\varphi_{|M|}:X\to\mathbb{P}^{1} is the contraction of the extremal ray R2R_{2} of type D with d2:=(KX)2H2=1d_{2}:=(-K_{X})^{2}\cdot H_{2}=1 [TanI, Proposition 4.1]. In particular, μ2=1\mu_{2}=1. Then Lemma 5.4 and Lemma 5.6 imply

24a=24r1+degB1+(12d2)=35+degB1.24a=\frac{24}{r_{1}}+\deg B_{1}+(12-d_{2})=35+\deg B_{1}.

By degB1(r1μ2a)2L13=(11a)24<4\deg B_{1}\leq(r_{1}-\frac{\mu_{2}}{a})^{2}\cdot L_{1}^{3}=(1-\frac{1}{a})^{2}\cdot 4<4 (Lemma 5.7), we obtain 24a=35+degB1<35+424a=35+\deg B_{1}<35+4, which implies a=1a=1. However, this leads to the following contradiction: degB1(11a)24=0\deg B_{1}\leq(1-\frac{1}{a})^{2}\cdot 4=0. ∎

Proposition 5.9.

We use Notation 5.1. Then the following hold.

  1. (1)

    {H1,H2}\{H_{1},H_{2}\} is a \mathbb{Z}-linear basis of PicX2{\operatorname{Pic}}\,X\simeq\mathbb{Z}^{\oplus 2}.

  2. (2)

    {1,2}\{\ell_{1},\ell_{2}\} is a dual basis of N1(X)\mathrm{N}_{1}(X)_{\mathbb{Z}}, i.e., H11=H22=0H_{1}\cdot\ell_{1}=H_{2}\cdot\ell_{2}=0 and H12=H21=1H_{1}\cdot\ell_{2}=H_{2}\cdot\ell_{1}=1. Here we set N1(X):=Z1(X)/\mathrm{N}_{1}(X)_{\mathbb{Z}}:={\rm Z}_{1}(X)/\equiv, where Z1(X):=CC{\rm Z}_{1}(X):=\bigoplus_{C}\mathbb{Z}C is the free \mathbb{Z}-module that is freely generated by all the curves CC on XX and \equiv denotes the numerical equivalence.

  3. (3)

    KXμ2H1+μ1H2.-K_{X}\sim\mu_{2}H_{1}+\mu_{1}H_{2}.

  4. (4)

    24=μ2H1c2(X)+μ1H2c2(X)24=\mu_{2}H_{1}\cdot c_{2}(X)+\mu_{1}H_{2}\cdot c_{2}(X).

Proof of Proposition 5.9 for the case when XX is primitive.

Set a:=|Pic(X)/(H1H2)|>0a:=|{\operatorname{Pic}}(X)/(\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2})|\in\mathbb{Z}_{>0} (Lemma 5.6(1)). Since XX is primitive, there is an extremal ray of type CC (Proposition 4.14). In what follows, we assume that R1R_{1} is of type CC. The proof consists of 7 steps.

Step 1.

In order to complete the proof of Proposition 5.9, it is enough to show that a=1a=1.

Proof of Step 1.

Assume a=1a=1. Then the assertion (1) follows from Lemma 5.6(1)(2). The assertions (3) and (4) hold by Lemma 5.6(4)(5). The assertion (2) follows from Lemma 5.6(3) and H11=H22=0H_{1}\cdot\ell_{1}=H_{2}\cdot\ell_{2}=0. This completes the proof of Step 1. ∎

Step 2.

If (μ1,μ2)=(1,1)(\mu_{1},\mu_{2})=(1,1), then a=1a=1.

Proof of Step 2.

By Lemma 5.6(5), we obtain

24a=H1c2(X)+H2c2(X).24a=H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X).

Since R1R_{1} is of type C1C_{1}, the following holds (Lemma 5.5(1)):

7c2(X)H117.7\leq c_{2}(X)\cdot H_{1}\leq 17.

If c2(X)H2=45c_{2}(X)\cdot H_{2}=45, then 5224a6252\leq 24a\leq 62, which is a contradiction. Hence 1c2(X)H2241\leq c_{2}(X)\cdot H_{2}\leq 24 (Lemma 5.5(2)), which implies

24a=H1c2(X)+H2c2(X)17+24=41.24a=H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)\leq 17+24=41.

Thus a=1a=1. This completes the proof of Step 2. ∎

Step 3.

If (μ1,μ2)=(1,2)(\mu_{1},\mu_{2})=(1,2), then a=1a=1.

Proof of Step 3.

By Lemma 5.6(5), we obtain

24a=2H1c2(X)+H2c2(X).24a=2H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X).

In this case, aa is odd. Indeed, if aa is even, then the linear equivalence

a(KX)2H1+H2a(-K_{X})\sim 2H_{1}+H_{2}

would imply that H2H_{2} is not a primitive element, which is a contradiction (Lemma 5.2). Since R1R_{1} is of type C1C_{1} and R2R_{2} is not of type E5E_{5} by μ2=2\mu_{2}=2, we get c2(X)H117c_{2}(X)\cdot H_{1}\leq 17 and c2(X)H224c_{2}(X)\cdot H_{2}\leq 24 (Lemma 5.5), which imply

24a=2H1c2(X)+H2c2(X)217+24=58.24a=2H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)\leq 2\cdot 17+24=58.

As aa is odd, we get a=1a=1. This completes the proof of Step 3. ∎

Step 4.

If (μ1,μ2)=(1,3)(\mu_{1},\mu_{2})=(1,3), then a=1a=1.

Proof of Step 4.

In this case, R1R_{1} is of type C1C_{1} and R2R_{2} is of type D3D_{3}. By Lemma 5.6(5), we obtain

24a=3H1c2(X)+H2c2(X).24a=3H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X).

We now show that aa is odd. By aKX3H1+H2-aK_{X}\sim 3H_{1}+H_{2} (Lemma 5.6(4)), we get

a3(KX)3\displaystyle a^{3}(-K_{X})^{3} =27H13+27H12H2+9H1H22+H23\displaystyle=27H_{1}^{3}+27H_{1}^{2}\cdot H_{2}+9H_{1}\cdot H_{2}^{2}+H_{2}^{3}
=(a)272(H21)\displaystyle\overset{{\rm(a)}}{=}27\cdot 2(H_{2}\cdot\ell_{1})
=(b)272a,\displaystyle\overset{{\rm(b)}}{=}27\cdot 2a,

where (b) holds by Lemma 5.6(3) and (a) follows from H13=0H^{3}_{1}=0, H220H_{2}^{2}\equiv 0, and H1221H_{1}^{2}\equiv 2\ell_{1} (Lemma 5.3). Hence we obtain a2(KX)3=227a^{2}(-K_{X})^{3}=2\cdot 27. By (KX)32(-K_{X})^{3}\in 2\mathbb{Z}, we get a2a\not\in 2\mathbb{Z}.

We have c2(X)H117c_{2}(X)\cdot H_{1}\leq 17 (Lemma 5.5) and c2(X)H2=3c_{2}(X)\cdot H_{2}=3 (Lemma 5.4), which imply

24a=3H1c2(X)+H2c2(X)317+3=54.24a=3H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)\leq 3\cdot 17+3=54.

As aa is odd, we get a=1a=1. This completes the proof of Step 4. ∎

In what follows, we treat the case when μ1=2\mu_{1}=2. In this case, R1R_{1} is of type C2C_{2} and H1c2(X)=6H_{1}\cdot c_{2}(X)=6 (Lemma 5.4).

Step 5.

If (μ1,μ2)=(2,1)(\mu_{1},\mu_{2})=(2,1), then a=1a=1.

Proof of Step 6.

By Lemma 5.6(5), we obtain

24a=H1c2(X)+2H2c2(X)=6+2H2c2(X).24a=H_{1}\cdot c_{2}(X)+2H_{2}\cdot c_{2}(X)=6+2H_{2}\cdot c_{2}(X).

By the same argument as in the case (μ1,μ2)=(1,2)(\mu_{1},\mu_{2})=(1,2) (Step 3), we see that aa is odd. If c2(X)H2=45c_{2}(X)\cdot H_{2}=45, then

24a=6+245=96,24a=6+2\cdot 45=96,

which contradicts the fact that aa is odd. Hence c2(X)H224c_{2}(X)\cdot H_{2}\leq 24 (Lemma 5.5), which implies

24a=6+2H2c2(X)6+224=54.24a=6+2H_{2}\cdot c_{2}(X)\leq 6+2\cdot 24=54.

As aa is odd, we get a=1a=1. This completes the proof of Step 5. ∎

Step 6.

If (μ1,μ2)=(2,2)(\mu_{1},\mu_{2})=(2,2), then a=1a=1.

Proof of Step 6.

By Lemma 5.6(5), we obtain

24a=2H1c2(X)+2H2c2(X)=12+2H2c2(X).24a=2H_{1}\cdot c_{2}(X)+2H_{2}\cdot c_{2}(X)=12+2H_{2}\cdot c_{2}(X).

Note that the type of R2R_{2} is E2,C2E_{2},C_{2}, or D2D_{2} (Remark 3.4). If R2R_{2} is of type E2E_{2}, then c2(X)H2=24/rc_{2}(X)\cdot H_{2}=24/r for some r>0r\in\mathbb{Z}_{>0} (Lemma 5.4). Hence

12a=6+H2c2(X)=6+24r.12a=6+H_{2}\cdot c_{2}(X)=6+\frac{24}{r}.

By 24/r624/r\in 6\mathbb{Z}, we get r{1,2,4}r\in\{1,2,4\}. Then it is easy to see that (a,r)=(1,4)(a,r)=(1,4) is a unique solution of 12a=6+24r12a=6+\frac{24}{r}.

We may assume that R2R_{2} is of type C2C_{2} or D2D_{2}. In this case, we have c2(X)H26c_{2}(X)\cdot H_{2}\leq 6 (Lemma 5.4), which implies

12a=6+H2c2(X)12.12a=6+H_{2}\cdot c_{2}(X)\leq 12.

Thus a=1a=1. This completes the proof of Step 6. ∎

Step 7.

If (μ1,μ2)=(2,3)(\mu_{1},\mu_{2})=(2,3), then a=1a=1.

Proof of Step 7.

In this case, R2R_{2} is of type D3D_{3}. By Lemma 5.6(5) and Lemma 5.4, we obtain

24a=3H1c2(X)+2H2c2(X)=18+6=24.24a=3H_{1}\cdot c_{2}(X)+2H_{2}\cdot c_{2}(X)=18+6=24.

Hence a=1a=1. This completes the proof of Step 7. ∎

Step 1–Step 7 complete the proof of Proposition 5.9 for the case when XX is primitive. ∎

Proof of Proposition 5.9 for the case when XX is imprimitive.

Since XX is imprimitive, XX has an extremal ray of type E1E_{1}. In what follows, we assume that R1R_{1} is of type E1E_{1}. Set a:=|Pic(X)/(H1H2)|>0a:=|{\operatorname{Pic}}(X)/(\mathbb{Z}H_{1}\oplus\mathbb{Z}H_{2})|\in\mathbb{Z}_{>0} (Lemma 5.6(1)). It is enough to show a=1a=1 (cf. Step 1 of the primitive case).

Claim 5.10.
  1. (1)

    H1c2(X)31H_{1}\cdot c_{2}(X)\leq 31.

  2. (2)

    H1c2(X){25(if a3)23(if if a2)17(if a=1).H_{1}\cdot c_{2}(X)\leq\begin{cases}25\qquad(\text{if }a\leq 3)\\ 23\qquad(\text{if }\text{if }a\leq 2)\\ 17\qquad(\text{if }a=1).\end{cases}

  3. (3)

    H1c2(X)20H_{1}\cdot c_{2}(X)\leq 20 if a3a\leq 3 and μ2=2\mu_{2}=2.

Proof of Claim 5.10.

By Lemma 5.4 and Lemma 5.7, we have

(5.10.1) H1c2(X)=24r1+degB124r1+(r1μ2a)2L13.H_{1}\cdot c_{2}(X)=\frac{24}{r_{1}}+\deg B_{1}\leq\frac{24}{r_{1}}+\left(r_{1}-\frac{\mu_{2}}{a}\right)^{2}\cdot L_{1}^{3}.

In what follows, we shall separate the argument depending on the value of r1r_{1}. Note that r1{2,3,4}r_{1}\in\{2,3,4\} (Proposition 5.8).

Assume r1=4r_{1}=4. In this case, we have L13=1L_{1}^{3}=1. This, together with (5.10.1), implies

H1c2(X)6+(4μ2a)2<6+42=22.H_{1}\cdot c_{2}(X)\leq 6+\left(4-\frac{\mu_{2}}{a}\right)^{2}<6+4^{2}=22.

If a=1a=1, then H1c2(X)6+(41)2=15H_{1}\cdot c_{2}(X)\leq 6+(4-1)^{2}=15. If a3a\leq 3 and μ2=2\mu_{2}=2, then H1c2(X)6+(423)2=6+1009<6+12=18H_{1}\cdot c_{2}(X)\leq 6+(4-\frac{2}{3})^{2}=6+\frac{100}{9}<6+12=18.

Assume r1=3r_{1}=3. In this case, we have L13=2L_{1}^{3}=2. This, together with (5.10.1), implies

H1c2(X)8+2(3μ2a)2<8+232=28.H_{1}\cdot c_{2}(X)\leq 8+2\left(3-\frac{\mu_{2}}{a}\right)^{2}<8+2\cdot 3^{2}=28.

If a3a\leq 3, then H1c2(X)8+2(313)2=8+1289<8+1359=8+15=23H_{1}\cdot c_{2}(X)\leq 8+2\left(3-\frac{1}{3}\right)^{2}=8+\frac{128}{9}<8+\frac{135}{9}=8+15=23. If a2a\leq 2, then H1c2(X)8+2(312)2=8+252<21H_{1}\cdot c_{2}(X)\leq 8+2\left(3-\frac{1}{2}\right)^{2}=8+\frac{25}{2}<21. If a=1a=1, then H1c2(X)8+2(31)2=16H_{1}\cdot c_{2}(X)\leq 8+2\left(3-1\right)^{2}=16. If a3a\leq 3 and μ2=2\mu_{2}=2, then H1c2(X)8+2(323)2=8+989<8+999=19H_{1}\cdot c_{2}(X)\leq 8+2\left(3-\frac{2}{3}\right)^{2}=8+\frac{98}{9}<8+\frac{99}{9}=19.

Assume r1=2r_{1}=2. In this case, we have 1L1351\leq L_{1}^{3}\leq 5. This, together with (5.10.1), implies

H1c2(X)12+5(2μ2a)2<12+54=32.H_{1}\cdot c_{2}(X)\leq 12+5\left(2-\frac{\mu_{2}}{a}\right)^{2}<12+5\cdot 4=32.

If a3a\leq 3, then H1c2(X)12+5(213)2=12+1259<12+1269=26H_{1}\cdot c_{2}(X)\leq 12+5\left(2-\frac{1}{3}\right)^{2}=12+\frac{125}{9}<12+\frac{126}{9}=26. If a2a\leq 2, then H1c2(X)12+5(212)2=12+454<24H_{1}\cdot c_{2}(X)\leq 12+5\left(2-\frac{1}{2}\right)^{2}=12+\frac{45}{4}<24. If a=1a=1, then H1c2(X)12+5(21)2=17H_{1}\cdot c_{2}(X)\leq 12+5\left(2-1\right)^{2}=17. If a3a\leq 3 and μ2=2\mu_{2}=2, then H1c2(X)12+5(223)2=12+809<21H_{1}\cdot c_{2}(X)\leq 12+5\left(2-\frac{2}{3}\right)^{2}=12+\frac{80}{9}<21. This completes the proof of Claim 5.10. ∎

We first treat the case when R2R_{2} is of type E1E_{1}. In this case, the same conclusion of Claim 5.10 holds for H2c2(X)H_{2}\cdot c_{2}(X). Claim 5.10(1) implies

24a=H1c2(X)+H2c2(x)31+31=62.24a=H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(x)\leq 31+31=62.

Hence we may assume that a=2a=2. By Claim 5.10(2), we get

48=24a=H1c2(X)+H2c2(x)23+23=46.48=24a=H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(x)\leq 23+23=46.

This is absurd. Thus R2R_{2} is not of type E2E_{2}.

We now consider the case when H2c2(X)=45H_{2}\cdot c_{2}(X)=45. Then R2R_{2} is of type E5E_{5} (Lemma 5.4, Lemma 5.5), and hence μ2=1\mu_{2}=1. Then Claim 5.10(1) implies

24a=H1c2(X)+H2c2(X)=H1c2(X)+4531+45=76.24a=H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)=H_{1}\cdot c_{2}(X)+45\leq 31+45=76.

Hence a3a\leq 3. By Claim 5.10(2), we obtain

24a=H1c2(X)+H2c2(X)=H1c2(X)+4525+45=70.24a=H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)=H_{1}\cdot c_{2}(X)+45\leq 25+45=70.

Hence we may assume that a=2a=2. We then get 48=24a=H1c2(X)+4548=24a=H_{1}\cdot c_{2}(X)+45, which implies

24r1<24r1+degB=H1c2(X)=3.\frac{24}{r_{1}}<\frac{24}{r_{1}}+\deg B=H_{1}\cdot c_{2}(X)=3.

This contradicts r1{2,3,4}r_{1}\in\{2,3,4\}. Thus H2c2(X)45H_{2}\cdot c_{2}(X)\neq 45, and hence 1H2c2(X)241\leq H_{2}\cdot c_{2}(X)\leq 24 (Lemma 5.5).

Assume that μ2=1\mu_{2}=1. Then Claim 5.10(1) implies

24a=H1c2(X)+H2c2(X)31+24=55.24a=H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)\leq 31+24=55.

Hence we may assume that a=2a=2. By Claim 5.10(2), we get

48=24a=H1c2(X)+H2c2(X)23+24=47,48=24a=H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)\leq 23+24=47,

which is absurd.

Assume that μ2=2\mu_{2}=2. Then Claim 5.10(1) implies

24a=2H1c2(X)+H2c2(X)231+24=86.24a=2H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)\leq 2\cdot 31+24=86.

Hence a3a\leq 3. By Claim 5.10(3), we get

24a=2H1c2(X)+H2c2(X)220+24=64.24a=2H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X)\leq 2\cdot 20+24=64.

Thus we may assume that a=2a=2. We then get

2KX2H1+H2,-2K_{X}\sim 2H_{1}+H_{2},

which contradicts the fact that H2H_{2} is primitive (Lemma 5.2).

Assume that μ2=3\mu_{2}=3. Then R2R_{2} is of type D3D_{3} and H2c2(X)=3H_{2}\cdot c_{2}(X)=3 (Lemma 5.4). Therefore, 24a=3H1c2(X)+H2c2(X)24a=3H_{1}\cdot c_{2}(X)+H_{2}\cdot c_{2}(X) (Lemma 5.6) implies the following (Claim 5.10(1)):

8a=H1c2(X)+132.8a=H_{1}\cdot c_{2}(X)+1\leq 32.

By aKX3H1+H2-aK_{X}\sim 3H_{1}+H_{2}, we obtain a3a\not\in 3\mathbb{Z}. Hence we may assume that a{2,4}a\in\{2,4\}.

Suppose a=4a=4. Then Lemma 5.4 and Lemma 5.7 imply

31=H1c2(X)=24r1+degB124r1+(r134)2L13.31=H_{1}\cdot c_{2}(X)=\frac{24}{r_{1}}+\deg B_{1}\leq\frac{24}{r_{1}}+\left(r_{1}-\frac{3}{4}\right)^{2}L_{1}^{3}.

If r1=4r_{1}=4, then L13=1L_{1}^{3}=1, and hence 316+(43a)2<6+16=2231\leq 6+(4-\frac{3}{a})^{2}<6+16=22, which is absurd. If r1=3r_{1}=3, then L13=2L_{1}^{3}=2, and hence 318+(33a)22<8+92=2631\leq 8+(3-\frac{3}{a})^{2}\cdot 2<8+9\cdot 2=26, which is impossible. If r1=2r_{1}=2, then 1L1351\leq L_{1}^{3}\leq 5, and hence 3112+(234)2L1312+25165<12+831\leq 12+(2-\frac{3}{4})^{2}\cdot L_{1}^{3}\leq 12+\frac{25}{16}\cdot 5<12+8, which is a contradiction. We then obtain a4a\neq 4.

Suppose a=2a=2. It suffices to derive a contradiction. We get 16=H1c2(X)+116=H_{1}\cdot c_{2}(X)+1. Hence Lemma 5.4 and Lemma 5.7 imply

15=H1c2(X)=24r1+degB124r1+(r132)2L13.15=H_{1}\cdot c_{2}(X)=\frac{24}{r_{1}}+\deg B_{1}\leq\frac{24}{r_{1}}+\left(r_{1}-\frac{3}{2}\right)^{2}\cdot L_{1}^{3}.

If r1=4r_{1}=4, then L13=1L_{1}^{3}=1 and 156+2541<6+715\leq 6+\frac{25}{4}\cdot 1<6+7, which is absurd. If r1=3r_{1}=3, then L13=2L_{1}^{3}=2 and 158+942<8+515\leq 8+\frac{9}{4}\cdot 2<8+5, which is a contradiction. If r1=2r_{1}=2, then 1L1351\leq L_{1}^{3}\leq 5 and 1512+14L1312+54<12+215\leq 12+\frac{1}{4}\cdot L_{1}^{3}\leq 12+\frac{5}{4}<12+2, which is absurd. This completes the proof of Proposition 5.9. ∎

Lemma 5.11.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1}. Then the following hold.

  1. (1)

    H12H2=(r1μ2)L13H_{1}^{2}\cdot H_{2}=(r_{1}-\mu_{2})L_{1}^{3}.

  2. (2)

    H1H22=(r1μ2)2L13degB1H_{1}\cdot H_{2}^{2}=(r_{1}-\mu_{2})^{2}L_{1}^{3}-\deg B_{1}.

Proof.

Let us show (1). Set D1:=Ex(f1)D_{1}:={\operatorname{Ex}}(f_{1}). We obtain

H12H2=(i)H12(KXμ2H1)=(ii)H12((r1μ2)H1D1)=(r1μ2)L13,H_{1}^{2}\cdot H_{2}\overset{{\rm(i)}}{=}H_{1}^{2}\cdot(-K_{X}-\mu_{2}H_{1})\overset{{\rm(ii)}}{=}H_{1}^{2}\cdot((r_{1}-\mu_{2})H_{1}-D_{1})=(r_{1}-\mu_{2})L_{1}^{3},

where (i) follows from KXμ2H1+H2-K_{X}\sim\mu_{2}H_{1}+H_{2} (Proposition 5.9(3)) and (ii) holds by KX=f1KX1+D1=r1H1+D1K_{X}=f_{1}^{*}K_{X_{1}}+D_{1}=-r_{1}H_{1}+D_{1}. Thus (1) holds. The assertion (2) follows from the displayed equation in the proof of Lemma 5.7 by using H12D1=0H_{1}^{2}\cdot D_{1}=0 and a=1a=1 (Proposition 5.9). ∎

5.2. Case CCC-C

Notation 5.12.

We use Notation 5.1. Assume that R1R_{1} is of type CC and R2R_{2} is of type CC. In particular, for each i{1,2}i\in\{1,2\}, fi:XYi=2f_{i}:X\to Y_{i}=\mathbb{P}^{2} is of type CC. Set

f:=f1×f2:X2×2.f:=f_{1}\times f_{2}\colon X\to\mathbb{P}^{2}\times\mathbb{P}^{2}.

Let f~:Xf(X)\widetilde{f}:X\to f(X) be the induced morphism, where f(X)f(X) denotes the scheme-theoretic image. For the ii-th projection πi:2×22\pi_{i}:\mathbb{P}^{2}\times\mathbb{P}^{2}\to\mathbb{P}^{2}, set Mi:=πLiM_{i}:=\pi^{*}L_{i}. Note that {M1,M2}\{M_{1},M_{2}\} is a free \mathbb{Z}-linear basis of Pic(2×2){\operatorname{Pic}}(\mathbb{P}^{2}\times\mathbb{P}^{2}).

Lemma 5.13.

We use Notation 5.12. Then

deg(f~)f(X)=f(X)2μ2M1+2μ1M2.\deg(\widetilde{f})\cdot f(X)=f_{*}(X)\sim\frac{2}{\mu_{2}}M_{1}+\frac{2}{\mu_{1}}M_{2}.

In particular, deg(f~)=1\deg(\widetilde{f})=1 or deg(f~)=2\deg(\widetilde{f})=2.

Proof.

Recall that f:XY1×Y2=2×2f:X\to Y_{1}\times Y_{2}=\mathbb{P}^{2}\times\mathbb{P}^{2} is a finite morphism (Lemma 4.15). We can write

f(X)a1M1+a2M2f_{*}(X)\sim a_{1}M_{1}+a_{2}M_{2}

for some a1,a2a_{1},a_{2}\in\mathbb{Z}. Pick a curve Γ1=(point×line)\Gamma_{1}=(\mathrm{point}\times\mathrm{line}) on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2}. Since M1(line×2)M_{1}\sim(\mathrm{line}\times\mathbb{P}^{2}) and M2(2×line)M_{2}\sim(\mathbb{P}^{2}\times\mathrm{line}), we have

M1Γ1=0andM2Γ1=1.M_{1}\cdot\Gamma_{1}=0\qquad\text{and}\qquad M_{2}\cdot\Gamma_{1}=1.

Hence we obtain f(X)Γ1=a2f_{*}(X)\cdot\Gamma_{1}=a_{2}. Now,

M12M2\displaystyle M_{1}^{2}\cdot M_{2} (point×2)(2×line)\displaystyle\equiv(\mathrm{point}\times\mathbb{P}^{2})\cdot(\mathbb{P}^{2}\times\mathrm{line})
(point×line)\displaystyle\equiv(\mathrm{point}\times\mathrm{line})
Γ1.\displaystyle\equiv\Gamma_{1}.

We have H21=1H_{2}\cdot\ell_{1}=1 (Proposition 5.9(2)) and H122μ11H_{1}^{2}\equiv\frac{2}{\mu_{1}}\ell_{1} (Lemma 5.3). We obtain

f(X)Γ1\displaystyle f_{*}(X)\cdot\Gamma_{1} =f(X)M12M2\displaystyle=f_{*}(X)\cdot M_{1}^{2}\cdot M_{2}
=(fM1)2fM2\displaystyle=(f^{*}M_{1})^{2}\cdot f^{*}M_{2}
=H12H2\displaystyle=H_{1}^{2}\cdot H_{2}
=2μ1H21\displaystyle=\frac{2}{\mu_{1}}H_{2}\cdot\ell_{1}
=2μ1.\displaystyle=\frac{2}{\mu_{1}}.

Hence a2=2μ1a_{2}=\frac{2}{\mu_{1}}. Similarly, we have a1=2μ2a_{1}=\frac{2}{\mu_{2}}. ∎

Lemma 5.14.

We use Notation 5.12. Then (KX)3=6(μ12+μ22)(-K_{X})^{3}=6(\mu_{1}^{2}+\mu_{2}^{2}).

Proof.

By Proposition 5.9, we have KXμ2H1+μ1H2-K_{X}\sim\mu_{2}H_{1}+\mu_{1}H_{2}. Since LiL_{i} is a line on Yi=2Y_{i}=\mathbb{P}^{2},

Hi3=fiLi3=0.H_{i}^{3}=f_{i}^{*}L_{i}^{3}=0.

By Hi22μiiH_{i}^{2}\equiv\frac{2}{\mu_{i}}\ell_{i} (Lemma 5.3) and H12=H21=1H_{1}\cdot\ell_{2}=H_{2}\cdot\ell_{1}=1 (Proposition 5.9), we obtain

(KX)3\displaystyle(-K_{X})^{3} =(μ2H1+μ1H2)3\displaystyle=(\mu_{2}H_{1}+\mu_{1}H_{2})^{3}
=μ23H13+3μ22μ1H12H2+3μ2μ12H1H22+μ13H23\displaystyle=\mu_{2}^{3}H_{1}^{3}+3\mu_{2}^{2}\mu_{1}H_{1}^{2}\cdot H_{2}+3\mu_{2}\mu_{1}^{2}H_{1}\cdot H_{2}^{2}+\mu_{1}^{3}H_{2}^{3}
=3μ22μ12μ1(H21)+3μ2μ122μ2(H12)\displaystyle=3\mu_{2}^{2}\mu_{1}\cdot\frac{2}{\mu_{1}}(H_{2}\cdot\ell_{1})+3\mu_{2}\mu_{1}^{2}\cdot\frac{2}{\mu_{2}}(H_{1}\cdot\ell_{2})
=6(μ12+μ22).\displaystyle=6(\mu_{1}^{2}+\mu_{2}^{2}).

5.2.1. Case deg(f~)=1\deg(\widetilde{f})=1

Lemma 5.15.

We use Notation 5.12. Assume that degf~=1\deg\widetilde{f}=1. Then the following hold.

  1. (1)

    f~:Xf(X)\widetilde{f}:X\to f(X) is an isomorphism.

  2. (2)

    If (μ1,μ2)=(1,1)(\mu_{1},\mu_{2})=(1,1), then X2×2X\subset\mathbb{P}^{2}\times\mathbb{P}^{2} is of bidegree (2,2)(2,2).

  3. (3)

    If (μ1,μ2)=(1,2)(\mu_{1},\mu_{2})=(1,2), then X2×2X\subset\mathbb{P}^{2}\times\mathbb{P}^{2} is of bidegree (1,2)(1,2).

  4. (4)

    If (μ1,μ2)=(2,2)(\mu_{1},\mu_{2})=(2,2), then X2×2X\subset\mathbb{P}^{2}\times\mathbb{P}^{2} is of bidegree (1,1)(1,1).

Proof.

If (1) holds, then (2), (3), and (4) hold. Thus it suffices to show (1).

Set Y:=f(X)Y:=f(X). Since f~:XY\widetilde{f}:X\to Y is a finite birational morphism from a smooth variety XX to a variety YY, f~\widetilde{f} is nothing but the normalisation of YY. As YY is a prime divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2}, YY is Gorenstein. By [Rei94, Proposition 2.3], we have

ωX𝒪X(C)f~ωY\omega_{X}\otimes\mathcal{O}_{X}(C)\simeq\widetilde{f}^{*}\omega_{Y}

for the conductor CC, which is an effective Cartier divisor on XX whose support coincides with the non-isomorphic locus Ex(f~){\operatorname{Ex}}(\widetilde{f}) of f~\widetilde{f}. By the adjunction formula, we have

ωY\displaystyle\omega_{Y} (ω2×2𝒪2×2(Y))𝒪Y\displaystyle\simeq(\omega_{\mathbb{P}^{2}\times\mathbb{P}^{2}}\otimes\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(Y))\otimes\mathcal{O}_{Y}
(𝒪2×2(3,3)𝒪2×2(2μ2,2μ1))𝒪Y\displaystyle\simeq\left(\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(-3,-3)\otimes\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}\left(\frac{2}{\mu_{2}},\frac{2}{\mu_{1}}\right)\right)\otimes\mathcal{O}_{Y}
𝒪2×2(2μ23,2μ13)𝒪Y,\displaystyle\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}\left(\frac{2}{\mu_{2}}-3,\frac{2}{\mu_{1}}-3\right)\otimes\mathcal{O}_{Y},

where the second isomorphims holds by Lemma 5.13. Therefore, we obtain

KX+C(2μ23)H1+(2μ13)H2.\displaystyle K_{X}+C\sim\left(\frac{2}{\mu_{2}}-3\right)H_{1}+\left(\frac{2}{\mu_{1}}-3\right)H_{2}.

It follows from KXμ2H1+μ1H2-K_{X}\sim\mu_{2}H_{1}+\mu_{1}H_{2} (Proposition 5.9) that

C(μ2+2μ23)H1+(μ1+2μ13)H2.C\sim\left(\mu_{2}+\frac{2}{\mu_{2}}-3\right)H_{1}+\left(\mu_{1}+\frac{2}{\mu_{1}}-3\right)H_{2}.

As we have μi{1,2}\mu_{i}\in\{1,2\} for each ii, we obtain C0C\sim 0, which implies C=0C=0. Hence f~\tilde{f} is an isomorphism. ∎

5.2.2. Case deg(f~)=2\deg(\widetilde{f})=2

Lemma 5.16.

Let WW be a nonzero effective Cartier divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,1)(1,1). Then the following hold.

  1. (1)

    WW is isomorphic to one of the following:

    • W1:={x0y0=0}2×2W_{1}:=\{x_{0}y_{0}=0\}\subset\mathbb{P}^{2}\times\mathbb{P}^{2}.

    • W2:={x0y0+x1y1=0}2×2W_{2}:=\{x_{0}y_{0}+x_{1}y_{1}=0\}\subset\mathbb{P}^{2}\times\mathbb{P}^{2}.

    • W3:={x0y0+x1y1+x2y2=0}2×2W_{3}:=\{x_{0}y_{0}+x_{1}y_{1}+x_{2}y_{2}=0\}\subset\mathbb{P}^{2}\times\mathbb{P}^{2}.

    Note that W1W_{1} is not irreducible, W2W_{2} is not smooth but irreducible, and W3W_{3} is smooth.

  2. (2)

    If WW2W\simeq W_{2}, then there exists a closed point Q2Q\in\mathbb{P}^{2} such that dimq11(Q)=2\dim q_{1}^{-1}(Q)=2, where q1:W2×2pr12q_{1}:W\hookrightarrow\mathbb{P}^{2}\times\mathbb{P}^{2}\xrightarrow{{\rm pr}_{1}}\mathbb{P}^{2}.

Proof.

Let us show (1). Since WW is defined by

fH0(2×2,𝒪2×2(1,1))H0(2,𝒪2(1))kH0(2,𝒪2(1)),f\in H^{0}(\mathbb{P}^{2}\times\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1))\simeq H^{0}(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(1))\otimes_{k}H^{0}(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(1)),

we have

W={0i,j2aijxiyj=0}2×2.W=\left\{\sum_{0\leq i,j\leq 2}a_{ij}x_{i}y_{j}=0\right\}\subset\mathbb{P}^{2}\times\mathbb{P}^{2}.

Applying the elementary operations to the 3×33\times 3-matrix (aij)(a_{ij}), we may assume that WW is one of W1,W2W_{1},W_{2}, and W3W_{3} (note that we can apply row and column elementray operations).

Let us show (2). By the proof of (1), WWiW\simeq W_{i} is obtained by applying a suitable Aut(2)×Aut(2){\rm Aut}(\mathbb{P}^{2})\times{\rm Aut}(\mathbb{P}^{2})-action on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2}. Hence we may assume that W=W2W=W_{2}. Then the assertion (2) holds for Q:=[0:0:1]Q:=[0:0:1]. ∎

Lemma 5.17.

We use Notation 5.12. Assume that degf~1\deg\widetilde{f}\neq 1. Set W:=f(X)W:=f(X). Then the following hold.

  1. (1)

    degf~=2\deg\widetilde{f}=2 and μ1=μ2=1\mu_{1}=\mu_{2}=1.

  2. (2)

    (KX)3=12(-K_{X})^{3}=12.

  3. (3)

    WW is a smooth prime divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,1)(1,1).

  4. (4)

    For :=(Coker(𝒪Wf~𝒪X))1\mathcal{L}:=({\operatorname{Coker}}(\mathcal{O}_{W}\to\widetilde{f}_{*}\mathcal{O}_{X}))^{-1}, it holds that 𝒪2×2(1,1)|W\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1)|_{W}.

  5. (5)

    f~:XW\widetilde{f}:X\to W is a split double cover.

Proof.

By Lemma 5.13, we have

deg(f~)f(X)=f(X)2μ2M1+2μ1M2.\deg(\tilde{f})\cdot f(X)=f_{*}(X)\sim\frac{2}{\mu_{2}}M_{1}+\frac{2}{\mu_{1}}M_{2}.

We then get deg(f~)=2deg(\tilde{f})=2 and μ1=μ2=1\mu_{1}=\mu_{2}=1. Hence (1) holds. By Lemma 5.14, (KX)3=6(μ12+μ22)=12.(-K_{X})^{3}=6(\mu_{1}^{2}+\mu_{2}^{2})=12. Thus (2) holds. Furthermore, we have

𝒪2×2(W)𝒪2×2(1,1).\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(W)\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1).

Then the assertion (3) follows from Lemma 5.16(2).

Let us show (4). Pick a Cartier divisor LL on WW with 𝒪W(L)\mathcal{L}\simeq\mathcal{O}_{W}(L). By Lemma 2.3, we have

(5.17.1) KXf~(KW+L).K_{X}\sim\widetilde{f}^{*}(K_{W}+L).

By the adjunction formula, we obtain

KW\displaystyle K_{W} (K2×2+W)|W\displaystyle\sim(K_{\mathbb{P}^{2}\times\mathbb{P}^{2}}+W)|_{W}
(3M13M2+M1+M2)|W\displaystyle\sim(-3M_{1}-3M_{2}+M_{1}+M_{2})|_{W}
2(M1+M2)|W.\displaystyle\sim-2(M_{1}+M_{2})|_{W}.

By (5.17.1) and KXH1H2K_{X}\sim-H_{1}-H_{2} (Lemma 5.9), we obtain

f~LKXf~KW(H1H2)f~(2(M1+M2)|W)=H1+H2.\widetilde{f}^{*}L\sim K_{X}-\widetilde{f}^{*}K_{W}\sim(-H_{1}-H_{2})-\widetilde{f}^{*}(-2(M_{1}+M_{2})|_{W})=H_{1}+H_{2}.

It holds that Lc1(𝒪2×2(1,1)|W)L\equiv c_{1}(\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1)|_{W}). Since WW is a smooth Fano threefold, we obtain Lc1(𝒪2×2(1,1)|W)L\sim c_{1}(\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1)|_{W}). Thus (4) holds. The assertion (5) follows from (3), (4), and Lemma 2.4(2). ∎

Theorem 5.18.

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2. Set W:={x0y0+x1y1+x2y2=0}2×2W:=\{x_{0}y_{0}+x_{1}y_{1}+x_{2}y_{2}=0\}\subset\mathbb{P}^{2}\times\mathbb{P}^{2}, which is a smooth hypersurface on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,1)(1,1). Let R1R_{1} and R2R_{2} be the distinct extremal rays of NE(X){\operatorname{NE}}(X). Assume that each of R1R_{1} and R2R_{2} is of type CC. Then, possibly after permuting R1R_{1} and R2R_{2}, one and only one of the following holds.

  1. (1)

    R1R_{1} is of type C1C_{1}, R2R_{2} is of type C1C_{1}, and one of the following holds.

    1. (a)

      XX is isomorphic to a hypersurface of 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (2,2)(2,2) and (KX)3=12(-K_{X})^{3}=12.

    2. (b)

      There exists a split double cover f:XWf:X\to W such that (f𝒪X/𝒪W)1𝒪2×2(1,1)|W(f_{*}\mathcal{O}_{X}/\mathcal{O}_{W})^{-1}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{2}}(1,1)|_{W} and (KX)3=12(-K_{X})^{3}=12.

  2. (2)

    R1R_{1} is of type C1C_{1}, R2R_{2} is of type C2C_{2}, (KX)3=30(-K_{X})^{3}=30, and XX is isomorphic to a hypersurface of 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,2)(1,2).

  3. (3)

    R1R_{1} is of type C2C_{2}, R2R_{2} is of type C2C_{2}, (KX)3=48(-K_{X})^{3}=48, and XX is isomorphic to WW.

Proof.

If one of the contractions is not generically smooth, then (2) holds by Proposition 4.17(5). Hence we may assume that each contraction is generically smooth. Then the assertion follows from Lemma 5.14, Lemma 5.15, and Lemma 5.17. ∎

5.3. Case CDC-D

Notation 5.19.

We use Notation 5.1. Assume that R1R_{1} is of type CC and R2R_{2} is of type DD. In particular, f1:XY1=2f_{1}:X\to Y_{1}=\mathbb{P}^{2} is of type CC and f2:XY2=1f_{2}:X\to Y_{2}=\mathbb{P}^{1} is of type DD. If f2f_{2} is of type DiD_{i} for i{1,2,3}i\in\{1,2,3\}, then we have μ2=i\mu_{2}=i (Remark 3.4). Set

f:=f1×f2:X2×1.f:=f_{1}\times f_{2}\colon X\to\mathbb{P}^{2}\times\mathbb{P}^{1}.
Lemma 5.20.

We use Notation 5.19. Then (KX)3=6μ22(-K_{X})^{3}=6\mu_{2}^{2}.

Proof.

By Proposition 5.9, KXμ2H1+μ1H2-K_{X}\sim\mu_{2}H_{1}+\mu_{1}H_{2}. By L1𝒪2(1)L_{1}\simeq\mathcal{O}_{\mathbb{P}^{2}}(1) and L2𝒪1(1)L_{2}\simeq\mathcal{O}_{\mathbb{P}^{1}}(1), we have H13=f1L13=0H_{1}^{3}=f_{1}^{*}L_{1}^{3}=0 and H22=f2L220H_{2}^{2}=f_{2}^{*}L_{2}^{2}\equiv 0. Therefore, we obtain

(KX)3\displaystyle(-K_{X})^{3} =(μ2H1+μ1H2)3\displaystyle=(\mu_{2}H_{1}+\mu_{1}H_{2})^{3}
=μ23H13+3μ22μ1H12H2+3μ2μ12H1H22+μ23H23\displaystyle=\mu_{2}^{3}H_{1}^{3}+3\mu_{2}^{2}\mu_{1}H_{1}^{2}\cdot H_{2}+3\mu_{2}\mu_{1}^{2}H_{1}\cdot H_{2}^{2}+\mu_{2}^{3}H_{2}^{3}
=3μ22μ1(2μ11)H2\displaystyle=3\mu_{2}^{2}\mu_{1}\cdot\left(\frac{2}{\mu_{1}}\ell_{1}\right)\cdot H_{2}
=6μ22,\displaystyle=6\mu_{2}^{2},

where the third equality holds by H122μ11H_{1}^{2}\equiv\frac{2}{\mu_{1}}\ell_{1} (Lemma 5.3) and the fourth one follows from H21=1H_{2}\cdot\ell_{1}=1 (Proposition 5.9). ∎

Lemma 5.21.

We use Notation 5.19.

  1. (1)

    If R1R_{1} is of type C1C_{1}, then f:X2×1f:X\to\mathbb{P}^{2}\times\mathbb{P}^{1} is a double cover.

  2. (2)

    If R1R_{1} is of type C2C_{2}, then f:X2×1f:X\to\mathbb{P}^{2}\times\mathbb{P}^{1} is an isomorphism and R2R_{2} is of type D3D_{3}.

Proof.

By Lemma 4.15(1), ff is a finite morphism. It follows from L1𝒪2(1)L_{1}\simeq\mathcal{O}_{\mathbb{P}^{2}}(1) and L2𝒪1(1)L_{2}\simeq\mathcal{O}_{\mathbb{P}^{1}}(1) that degf=(f1L1)2f2L2\deg f=(f_{1}^{*}L_{1})^{2}\cdot f_{2}^{*}L_{2}. By Corollary 4.5, any fibre of f1:XY1=2f_{1}:X\to Y_{1}=\mathbb{P}^{2} is numerically equivalent to 2μ11\frac{2}{\mu_{1}}\ell_{1}, i.e., (f1L1)22μ11(f_{1}^{*}L_{1})^{2}\equiv\frac{2}{\mu_{1}}\ell_{1}. We then obtain

degf=(f1L1)2f2L2=2μ11H2=2μ1,\deg f=(f_{1}^{*}L_{1})^{2}\cdot f_{2}^{*}L_{2}=\frac{2}{\mu_{1}}\ell_{1}\cdot H_{2}=\frac{2}{\mu_{1}},

where the last equality holds by Proposition 5.9.

(1) Assume that R1R_{1} is of type C1C_{1}. We then have μ1=1\mu_{1}=1, and hence f:X2×1f:X\to\mathbb{P}^{2}\times\mathbb{P}^{1} is a double cover.

(2) Assume that R2R_{2} is of type C2C_{2}. Then we have μ1=1\mu_{1}=1 and degf=1\deg f=1. Hence, f:X2×1f:X\to\mathbb{P}^{2}\times\mathbb{P}^{1} is a finite birational morphism of normal varieties, which is automatically an isomorphism. In particular, R2R_{2} is of type D3D_{3}. ∎

Lemma 5.22.

We use Notation 5.19. Assume that R1R_{1} is of type C1C_{1}. Set :=(f𝒪X/𝒪2×1)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}})^{-1}. Then the following hold.

  1. (1)

    R2R_{2} is of type D1D_{1} or D2D_{2}.

  2. (2)

    𝒪2×1(3μ2,1)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(3-\mu_{2},1).

  3. (3)

    f:X2×1f:X\to\mathbb{P}^{2}\times\mathbb{P}^{1} is a split double cover.

Proof.

By μ1=1\mu_{1}=1, we have KXμ2H1H2K_{X}\sim-\mu_{2}H_{1}-H_{2}, i.e., ωXf𝒪2×1(μ2,1)\omega_{X}\simeq f^{*}\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(-\mu_{2},-1). Recall that we have ωXf(ω2×1)\omega_{X}\simeq f^{*}(\omega_{\mathbb{P}^{2}\times\mathbb{P}^{1}}\otimes\mathcal{L}) (Lemma 2.3). It holds that 𝒪2×1(b1,b2)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(b_{1},b_{2}) for some b1,b2b_{1},b_{2}\in\mathbb{Z}. Then

ωX\displaystyle\omega_{X} f(ω2×1)\displaystyle\simeq f^{*}\left(\omega_{\mathbb{P}^{2}\times\mathbb{P}^{1}}\otimes\mathcal{L}\right)
f(𝒪2×1(3,2)𝒪2×1(b1,b2))\displaystyle\simeq f^{*}\left(\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(-3,-2)\otimes\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(b_{1},b_{2})\right)
f𝒪2×1(b13,b22).\displaystyle\simeq f^{*}\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(b_{1}-3,b_{2}-2).

Hence b1=3μ2b_{1}=3-\mu_{2} and b2=1b_{2}=1. Thus (2) holds.

Let us show (3), i.e.,

0𝒪2×1f𝒪X00\to\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}\to f_{*}\mathcal{O}_{X}\to\mathcal{L}\to 0

splits. The extension class lies in

Ext2×11(,𝒪2×1)H1(2×1,1)H1(2×1,𝒪2×1(μ23,1))=0,{\rm Ext}^{1}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(\mathcal{L},\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}})\simeq H^{1}(\mathbb{P}^{2}\times\mathbb{P}^{1},\mathcal{L}^{-1})\simeq H^{1}(\mathbb{P}^{2}\times\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(\mu_{2}-3,-1))=0,

where the last equality holds by the Künneth formula. Thus (3) holds.

Let us show (1). Suppose that R2R_{2} is of type D3D_{3}. In this case, we have that XS×1X\simeq S\times\mathbb{P}^{1} for some smooth projective surface SS (Lemma 8.4). Then XX has an extremal ray of type C2C_{2}, which is a contraction. Thus (1) holds. ∎

Theorem 5.23.

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2. Let R1R_{1} and R2R_{2} be extremal rays of NE(X){\operatorname{NE}}(X). Assume that R1R_{1} is of type CC and R2R_{2} is of type DD. Then one and only one of the following holds.

  1. (1)

    R1R_{1} is of type C1C_{1}, R2R_{2} is of type D1D_{1}, (KX)3=6(-K_{X})^{3}=6, and there exists a split double cover f:X2×1f:X\to\mathbb{P}^{2}\times\mathbb{P}^{1} such that (f𝒪X/𝒪2×1)1𝒪2×1(2,1)(f_{*}\mathcal{O}_{X}/\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}})^{-1}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(2,1).

  2. (2)

    R1R_{1} is of type C1C_{1}, R2R_{2} is of type D2D_{2}, (KX)3=24(-K_{X})^{3}=24, and there exists a split double cover f:X2×1f:X\to\mathbb{P}^{2}\times\mathbb{P}^{1} such that (f𝒪X/𝒪2×1)1𝒪2×1(1,1)(f_{*}\mathcal{O}_{X}/\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}})^{-1}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(1,1).

  3. (3)

    R1R_{1} is of type C2C_{2}, R2R_{2} is of type D3D_{3}, (KX)3=54(-K_{X})^{3}=54, and X2×1X\simeq\mathbb{P}^{2}\times\mathbb{P}^{1}.

Proof.

The assertions follow from Lemma 5.20, Lemma 5.21, and Lemma 5.22. ∎

5.4. Case CEC-E (primitive)

Notation 5.24.

We use Notation 5.1. Assume that R1R_{1} is of type CC and R2R_{2} is of type E2,E3,E4E_{2},E_{3},E_{4}, or E5E_{5}. Set D:=Ex(f2:XY2)D:={\operatorname{Ex}}(f_{2}:X\to Y_{2}), which is a prime divisor on XX.

Let us consider the induced morphism f1|D:DY1=2f_{1}|_{D}\colon D\to Y_{1}=\mathbb{P}^{2}.

Lemma 5.25.

We use Notation 5.24. Then the following hold.

  1. (1)

    If R1R_{1} is of type C1C_{1}, then f1|D:DY1=2f_{1}|_{D}:D\to Y_{1}=\mathbb{P}^{2} is a double cover and R2R_{2} is of type E3E_{3} or E4E_{4}.

  2. (2)

    If R1R_{1} is of type C2C_{2}, then f1|D:DY1=2f_{1}|_{D}:D\to Y_{1}=\mathbb{P}^{2} is an isomorphism and R2R_{2} is of type E2E_{2} or E5E_{5}.

Proof.

By Lemma 4.15(2), f1|D:DY1=2f_{1}|_{D}:D\to Y_{1}=\mathbb{P}^{2} is a finite surjective morphism.

Step 1.

It holds that

deg(f1|D)=2μ1D1=1μ22(ωD𝒪D(D))2,\deg(f_{1}|_{D})=\frac{2}{\mu_{1}}D\cdot\ell_{1}=\frac{1}{\mu_{2}^{2}}(\omega_{D}\otimes\mathcal{O}_{D}(-D))^{2},

where 𝒪D(D):=𝒪X(D)|D\mathcal{O}_{D}(-D):=\mathcal{O}_{X}(-D)|_{D}.

Proof of Step 1.

By Corollary 4.5, a fibre of f1f_{1} is numerically equivalent to (2/μ1)1(2/\mu_{1})\ell_{1}. We then obtain

deg(f1|D)=2μ1D1=DH12,\deg(f_{1}|_{D})=\frac{2}{\mu_{1}}D\cdot\ell_{1}=D\cdot H_{1}^{2},

where the latter equality follows from Lemma 5.3. By Proposition 5.9(3), we have

H12\displaystyle H_{1}^{2} 1μ22(KXμ1H2)2\displaystyle\equiv\frac{1}{\mu_{2}^{2}}(-K_{X}-\mu_{1}H_{2})^{2}
=1μ22((KX)22μ1(KX)H2+μ12H22).\displaystyle=\frac{1}{\mu_{2}^{2}}((-K_{X})^{2}-2\mu_{1}(-K_{X})\cdot H_{2}+\mu_{1}^{2}H_{2}^{2}).

Since R2R_{2} is of type E2,E3,E4E_{2},E_{3},E_{4} or E5E_{5}, f2(D)f_{2}(D) is a point on YY. Therefore, we get

H2D=f2L2D0.H_{2}\cdot D=f^{*}_{2}L_{2}\cdot D\equiv 0.

Hence we obtain

deg(f1|D)=DH12=1μ22D((KX)22μ1(KX)H2+μ12H22)\deg(f_{1}|_{D})=D\cdot H_{1}^{2}=\frac{1}{\mu_{2}^{2}}D\cdot((-K_{X})^{2}-2\mu_{1}(-K_{X})\cdot H_{2}+\mu_{1}^{2}H_{2}^{2})
=1μ22D(KX)2=1μ22(ωD𝒪D(D))2,=\frac{1}{\mu_{2}^{2}}D\cdot(-K_{X})^{2}=\frac{1}{\mu_{2}^{2}}(\omega_{D}\otimes\mathcal{O}_{D}(-D))^{2},

where the last equality holds by the adjunction formula. This completes the proof of Step 1. ∎

Step 2.

The assertion (1) holds.

Proof of Step 2.

Assume that R1R_{1} is of type C1C_{1}. We have μ1=1\mu_{1}=1. Recall that R2R_{2} is of type E2,E3,E4,E_{2},E_{3},E_{4}, or E5E_{5}.

Assume that R2R_{2} is of type E2E_{2}. Then we have μ2=2,D2\mu_{2}=2,D\simeq\mathbb{P}^{2}, and 𝒪D(D)𝒪D(1)\mathcal{O}_{D}(D)\simeq\mathcal{O}_{D}(-1). By Step 1, we get

2D1=1μ22(ωD𝒪D(D))2=14(3+1)2=1.2D\cdot\ell_{1}=\frac{1}{\mu_{2}^{2}}(\omega_{D}\otimes\mathcal{O}_{D}(-D))^{2}=\frac{1}{4}(-3+1)^{2}=1.

This is absurd.

Assume that R2R_{2} is of type E5E_{5}. Then μ2=1,D2,\mu_{2}=1,D\simeq\mathbb{P}^{2}, and 𝒪D(D)𝒪2(2)\mathcal{O}_{D}(D)\simeq\mathcal{O}_{\mathbb{P}^{2}}(-2). By Step 1, we get

2D1=1μ22(ωD𝒪D(D))2=(𝒪2(3+2))2=1.2D\cdot\ell_{1}=\frac{1}{\mu_{2}^{2}}(\omega_{D}\otimes\mathcal{O}_{D}(-D))^{2}=(\mathcal{O}_{\mathbb{P}^{2}}(-3+2))^{2}=1.

This is a contradiction.

Assume that R2R_{2} is of type E3E_{3} or E4E_{4}. Then we have μ2=1\mu_{2}=1, DD is isomorphic to a possibly singular quadric surface in 3\mathbb{P}^{3}, and 𝒪D(D)𝒪3(1)|D\mathcal{O}_{D}(D)\simeq\mathcal{O}_{\mathbb{P}^{3}}(-1)|_{D}. We have

ωD(ω3𝒪3(D))|D𝒪3(4+2)|D𝒪3(2)|D.\omega_{D}\simeq(\omega_{\mathbb{P}^{3}}\otimes\mathcal{O}_{\mathbb{P}^{3}}(D))|_{D}\simeq\mathcal{O}_{\mathbb{P}^{3}}(-4+2)|_{D}\\ \simeq\mathcal{O}_{\mathbb{P}^{3}}(-2)|_{D}.

By Step 1, we get

deg(f1|D)=2D1=1μ22(ωD𝒪D(D))2=(𝒪3(2+1)|D)2=2.\deg(f_{1}|_{D})=2D\cdot\ell_{1}=\frac{1}{\mu_{2}^{2}}(\omega_{D}\otimes\mathcal{O}_{D}(-D))^{2}=(\mathcal{O}_{\mathbb{P}^{3}}(-2+1)|_{D})^{2}=2.

This completes the proof of Step 2. ∎

Step 3.

The assertion (2) holds.

Proof of Step 3.

Assume that R1R_{1} is of type C2C_{2}. We have μ1=2\mu_{1}=2.

Suppose that R2R_{2} is of type E3E_{3} or E4E_{4}. Then μ2=1\mu_{2}=1, and hence Proposition 5.9 implies

KXH1+2H2.-K_{X}\sim H_{1}+2H_{2}.

We then get

24\displaystyle 24 =c2(X)(KX)\displaystyle=c_{2}(X)\cdot(-K_{X})
=c2(X)H1+2c2(X)H2\displaystyle=c_{2}(X)\cdot H_{1}+2c_{2}(X)\cdot H_{2}
=6+48r\displaystyle=6+\frac{48}{r}

for some r>0r\in\mathbb{Z}_{>0} (Lemma 5.4). This is impossible.

Hence R2R_{2} is of type E2E_{2} or E5E_{5}. It holds that

deg(f1|D)=D1=1μ22(ωD𝒪D(D))2=1,\deg(f_{1}|_{D})=D\cdot\ell_{1}=\frac{1}{\mu_{2}^{2}}(\omega_{D}\otimes\mathcal{O}_{D}(-D))^{2}=1,

where the first two equalities are guaranteed by Step 1 and the last one follows from the same argument as in Step 2. Then f1|D:D2f_{1}|_{D}:D\to\mathbb{P}^{2} is a finite birational morphism of normal varieties, which is automatically an isomorphism. This completes the proof of Step 3. ∎

Step 2 and Step 3 complete the proof of Lemma 5.25. ∎

5.4.1. Case C1EC_{1}-E

Notation 5.26.

We use Notation 5.1. Assume that R1R_{1} is of type C1C_{1} and R2R_{2} is of type E2,E3,E4,E_{2},E_{3},E_{4}, or E5E_{5}. Set D:=Ex(f2:XY2)D:={\operatorname{Ex}}(f_{2}:X\to Y_{2}). In particular, f1:XY1=2f_{1}:X\to Y_{1}=\mathbb{P}^{2} is of type C1C_{1} and f2:XY2f_{2}:X\to Y_{2} is of type E3E_{3} or E4E_{4} (Lemma 5.25). Recall that DD is a (possibly singular) quadric surface in 3\mathbb{P}^{3} and set 𝒪D(n):=𝒪3(n)|D\mathcal{O}_{D}(n):=\mathcal{O}_{\mathbb{P}^{3}}(n)|_{D} for nn\in\mathbb{Z}. By Lemma 3.15, we have a diagram:

X{X}P:=((f1)𝒪X(KX)){P:=\mathbb{P}((f_{1})_{*}\mathcal{O}_{X}(-K_{X}))}2.{\mathbb{P}^{2}.}ι\scriptstyle{\iota}π\scriptstyle{\pi}

Since ι\iota is a closed immersion, we identify XX with the smooth prime divisor ι(X)\iota(X) on the 2\mathbb{P}^{2}-bundle PP over 2\mathbb{P}^{2}.

Lemma 5.27.

We use Notation 5.26. Then R1(f1)𝒪X(KXD)=0R^{1}(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)=0 and there exists an exact sequence:

0(f1)𝒪X(KXD)(f1)𝒪X(KX)(f1)𝒪D(KX)0.0\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\longrightarrow(f_{1})_{*}\mathcal{O}_{D}(-K_{X})\longrightarrow 0.
Proof.

Since (KXD)KX(-K_{X}-D)-K_{X} is f1f_{1}-ample (Lemma 5.25), the assertion follows from R1(f1)𝒪X(KXD)=0R^{1}(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)=0, which is guaranteed by [Tan15, Theorem 0.5]. ∎

Lemma 5.28.

We use Notation 5.26. Then (f1|D)𝒪2(n)𝒪D(n)(f_{1}|_{D})^{*}\mathcal{O}_{\mathbb{P}^{2}}(n)\simeq\mathcal{O}_{D}(n) for all nn\in\mathbb{Z}.

Proof.

We may assume n=1n=1.

Assume that R2R_{2} is of type E3E_{3}. Then D1×1D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1} and we can write

(f1|D)𝒪2(1)𝒪D(a,b)(f_{1}|_{D})^{*}\mathcal{O}_{\mathbb{P}^{2}}(1)\simeq\mathcal{O}_{D}(a,b)

for some a,ba,b\in\mathbb{Z}. Since f1|D:D2f_{1}|_{D}:D\to\mathbb{P}^{2} is a finite morphism, this invertible sheaf is ample, which implies a>0a>0 and b>0b>0. By the projection formula, we get

2ab\displaystyle 2ab =(𝒪D(a,b))2\displaystyle=(\mathcal{O}_{D}(a,b))^{2}
=deg(f1|D)(𝒪2(1))2\displaystyle=\deg(f_{1}|_{D})\cdot(\mathcal{O}_{\mathbb{P}^{2}}(1))^{2}
=2.\displaystyle=2.

Hence we have ab=1ab=1. We then get (a,b)=(1,1)(a,b)=(1,1) and

(f1|D)𝒪2(1)𝒪1×1(1,1)𝒪3(1)|D=𝒪D(1).(f_{1}|_{D})^{*}\mathcal{O}_{\mathbb{P}^{2}}(1)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)\simeq\mathcal{O}_{\mathbb{P}^{3}}(1)|_{D}=\mathcal{O}_{D}(1).

Assume that R2R_{2} is of type E4E_{4}. By [Har77, Ch. II, Exercise 6.3 and 6.5], the restriction homomorphism Pic(3)Pic(D){\operatorname{Pic}}(\mathbb{P}^{3})\to{\operatorname{Pic}}(D) is an isomorphism. Hence we have

(f1|D)𝒪2(1)𝒪D(a)(f_{1}|_{D})^{*}\mathcal{O}_{\mathbb{P}^{2}}(1)\simeq\mathcal{O}_{D}(a)

for some aa\in\mathbb{Z}. Since this invertible sheaf is ample, we get a>0a>0. By the projection formula,

a2c1(𝒪D(1))2=c1(𝒪D(a))2=deg(f1|D)c1(𝒪2(1))2=2.a^{2}c_{1}(\mathcal{O}_{D}(1))^{2}=c_{1}(\mathcal{O}_{D}(a))^{2}=\deg(f_{1}|_{D})\cdot c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(1))^{2}=2.

Hence a=1a=1. ∎

Proposition 5.29.

We use Notation 5.26. Then the following hold.

  1. (1)

    (f1)𝒪X(KXD)𝒪2(2).(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\simeq\mathcal{O}_{\mathbb{P}^{2}}(2).

  2. (2)

    (f1)𝒪D(KX)𝒪2𝒪2(1)(f_{1})_{*}\mathcal{O}_{D}(-K_{X})\simeq\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1).

  3. (3)

    (f1)𝒪X(KX)𝒪2𝒪2(1)𝒪2(2)(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\simeq\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2).

Proof.

Let us show (1). For a closed point t2t\in\mathbb{P}^{2} and the fibre XtX_{t} of f1:XY1=2f_{1}:X\to Y_{1}=\mathbb{P}^{2} over tt, we have KXXt=DXt=2-K_{X}\cdot X_{t}=D\cdot X_{t}=2. Hence h0(Xt,𝒪Xt(KXD))=1h^{0}(X_{t},\mathcal{O}_{X_{t}}(-K_{X}-D))=1 for all tY1=2t\in Y_{1}=\mathbb{P}^{2}. By Grauert’s theorem [Har77, Ch. III, Theorem 12.9], (f1)(𝒪X(KXD))(f_{1})_{*}(\mathcal{O}_{X}(-K_{X}-D)) is an invertible sheaf on 2\mathbb{P}^{2}. Hence we can write

(f1)𝒪X(KXD)𝒪2(a)(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\simeq\mathcal{O}_{\mathbb{P}^{2}}(a)

for some aa\in\mathbb{Z}. Again by [Har77, Ch. III, Theorem 12.9], we get

𝒪X(KXD)f1(f1)𝒪X(KXD)f1𝒪2(a).\mathcal{O}_{X}(-K_{X}-D)\simeq f^{*}_{1}(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\simeq f^{*}_{1}\mathcal{O}_{\mathbb{P}^{2}}(a).

By Lemma 5.28 and adjunction formula, we obtain

𝒪D(a)(f1|D)𝒪2(a)𝒪D(KXD)ωD1𝒪3(2)|D=𝒪D(2),\mathcal{O}_{D}(a)\simeq(f_{1}|_{D})^{*}\mathcal{O}_{\mathbb{P}^{2}}(a)\simeq\mathcal{O}_{D}(-K_{X}-D)\simeq\omega_{D}^{-1}\simeq\mathcal{O}_{\mathbb{P}^{3}}(2)|_{D}=\mathcal{O}_{D}(2),

which implies a=2a=2. Thus (1) holds.

Let us show (2). By 𝒪D(KXD)𝒪D(2)\mathcal{O}_{D}(-K_{X}-D)\simeq\mathcal{O}_{D}(2) and 𝒪D(D)𝒪D(1)\mathcal{O}_{D}(D)\simeq\mathcal{O}_{D}(-1), we obtain 𝒪D(KX)𝒪D(1)\mathcal{O}_{D}(-K_{X})\simeq\mathcal{O}_{D}(1). Then

(f1|D)𝒪D(KX)\displaystyle(f_{1}|_{D})_{*}\mathcal{O}_{D}(-K_{X}) (f1|D)𝒪D(1)\displaystyle\simeq(f_{1}|_{D})_{*}\mathcal{O}_{D}(1)
(f1|D)(𝒪D(f1|D)𝒪2(1))\displaystyle\simeq(f_{1}|_{D})_{*}(\mathcal{O}_{D}\otimes(f_{1}|_{D})^{*}\mathcal{O}_{\mathbb{P}^{2}}(1))
(f1|D)𝒪D𝒪2(1).\displaystyle\simeq(f_{1}|_{D})_{*}\mathcal{O}_{D}\otimes\mathcal{O}_{\mathbb{P}^{2}}(1).

Since f1|D:D2f_{1}|_{D}\colon D\to\mathbb{P}^{2} is a double cover (Lemma 5.25), we have an exact sequence

0𝒪2(f1|D)𝒪D𝒪2(b)00\to\mathcal{O}_{\mathbb{P}^{2}}\to(f_{1}|_{D})_{*}\mathcal{O}_{D}\to\mathcal{O}_{\mathbb{P}^{2}}(-b)\to 0

for some bb\in\mathbb{Z}. Since this sequence splits, we get (f1|D)𝒪D𝒪2𝒪2(b)(f_{1}|_{D})_{*}\mathcal{O}_{D}\simeq\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(-b). Furthermore, ωD(f1|D)(ω2𝒪2(b))\omega_{D}\simeq(f_{1}|_{D})^{*}(\omega_{\mathbb{P}^{2}}\otimes\mathcal{O}_{\mathbb{P}^{2}}(b)) (Lemma 2.3). By ωD𝒪D(2)\omega_{D}\simeq\mathcal{O}_{D}(-2), we obtain

𝒪D(2)\displaystyle\mathcal{O}_{D}(-2) ωD\displaystyle\simeq\omega_{D}
(f1|D)(𝒪2(3)𝒪2(b))\displaystyle\simeq(f_{1}|_{D})^{*}(\mathcal{O}_{\mathbb{P}^{2}}(-3)\otimes\mathcal{O}_{\mathbb{P}^{2}}(b))
(f1|D)(𝒪2(b3))\displaystyle\simeq(f_{1}|_{D})^{*}(\mathcal{O}_{\mathbb{P}^{2}}(b-3))
𝒪D(b3),\displaystyle\simeq\mathcal{O}_{D}(b-3),

where the last isomorphism holds by Lemma 5.28. Thus b=1b=1 and (2) holds.

Let us show (3). By (1), (2), and Lemma 5.27, we have the following exact sequence:

0𝒪2(2)(f1)𝒪X(KX)𝒪2𝒪2(1)0.0\longrightarrow\mathcal{O}_{\mathbb{P}^{2}}(2)\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\longrightarrow\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\longrightarrow 0.

By Ext1(𝒪2𝒪2(1),𝒪2(2))=0\mathrm{Ext}^{1}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1),\mathcal{O}_{\mathbb{P}^{2}}(2))=0, this exact sequence splits, and hence

(f1)𝒪X(KX)𝒪2𝒪2(1)𝒪2(2).(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\simeq\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2).

Thus (3) holds. ∎

Lemma 5.30.

We use Notation 5.26. Then 𝒪P(X)𝒪P(2)\mathcal{O}_{P}(X)\simeq\mathcal{O}_{P}(2), where 𝒪P(1)\mathcal{O}_{P}(1) denotes the tautological line bundle of the 2\mathbb{P}^{2}-bundle structure π:P=((f1)𝒪X(KX))2\pi:P=\mathbb{P}((f_{1})_{*}\mathcal{O}_{X}(-K_{X}))\to\mathbb{P}^{2} and we set 𝒪P(2):=𝒪P(1)2\mathcal{O}_{P}(2):=\mathcal{O}_{P}(1)^{\otimes 2}.

Proof.

Recall that we have the following diagram (Notation 5.26, Proposition 5.29):

X{X}P=(𝒪2𝒪2(1)𝒪2(2)){P=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))}2.{\mathbb{P}^{2}.}ι\scriptstyle{\iota}π\scriptstyle{\pi}

By PicPπPic2𝒪P(1){\operatorname{Pic}}\,P\simeq\pi^{*}{\operatorname{Pic}}\,\mathbb{P}^{2}\oplus\mathbb{Z}\mathcal{O}_{P}(1), we can write

(5.30.1) 𝒪P(X)π𝒪2(a)𝒪P(n)\mathcal{O}_{P}(X)\simeq\pi^{*}\mathcal{O}_{\mathbb{P}^{2}}(a)\otimes\mathcal{O}_{P}(n)

for some a,na,n\in\mathbb{Z}. It follows from the adjunction formula that

(5.30.2) 𝒪X(KX)𝒪P(KP+X)|X.\mathcal{O}_{X}(K_{X})\simeq\mathcal{O}_{P}(K_{P}+X)|_{X}.

By Proposition 7.1(2), it holds that

(5.30.3) 𝒪P(KP)𝒪P(3)π(ω2det(𝒪2𝒪2(1)𝒪2(2)))𝒪P(3).\mathcal{O}_{P}(K_{P})\simeq\mathcal{O}_{P}(-3)\otimes\pi^{*}(\omega_{\mathbb{P}^{2}}\otimes\det(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)))\simeq\mathcal{O}_{P}(-3).

Since the closed immersion ι:XP\iota\colon X\to P is induced by the surjection (f1)(f1)𝒪X(KX)𝒪X(KX)(f_{1})^{*}(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\to\mathcal{O}_{X}(-K_{X}), it follows from [Har77, Ch. II, the proof of Proposition 7.12] that

(5.30.4) 𝒪X(KX)ι𝒪P(1)𝒪P(1)|X.\mathcal{O}_{X}(-K_{X})\simeq\iota^{*}\mathcal{O}_{P}(1)\simeq\mathcal{O}_{P}(1)|_{X}.

By (5.30.1), (5.30.2), (5.30.3) and (5.30.4), we have

𝒪P(1)|X𝒪X(KX)𝒪P(KPX)|X(π𝒪2(a)𝒪P(3n))|X.\mathcal{O}_{P}(1)|_{X}\simeq\mathcal{O}_{X}(-K_{X})\simeq\mathcal{O}_{P}(-K_{P}-X)|_{X}\simeq(\pi^{*}\mathcal{O}_{\mathbb{P}^{2}}(-a)\oplus\mathcal{O}_{P}(3-n))|_{X}.

Taking the intersection number of this equation with a general fibre of π|X:X2\pi|_{X}\colon X\to\mathbb{P}^{2}, we have 1=3n1=3-n, i.e., n=2n=2. Hence we obtain 𝒪Xπ𝒪2(a)|X\mathcal{O}_{X}\simeq\pi^{*}\mathcal{O}_{\mathbb{P}^{2}}(a)|_{X}. We see that a=0a=0 by taking the intersection number of this equation with a curve CC on XX such that π(C)\pi(C) is a curve. ∎

Lemma 5.31.

We use Notation 5.26. Then (KX)3=14(-K_{X})^{3}=14.

Proof.

It follows from Proposition 7.1(5) that

(KX)3=\displaystyle(-K_{X})^{3}= 2c1()22c2()+4c1()c1(𝒪2)+6c1()K2\displaystyle 2c_{1}(\mathcal{E})^{2}-2c_{2}(\mathcal{E})+4c_{1}(\mathcal{E})\cdot c_{1}(\mathcal{O}_{\mathbb{P}^{2}})+6c_{1}(\mathcal{E})\cdot K_{\mathbb{P}^{2}}
+9c1(𝒪2)K2+6K22+3c1(𝒪2)2.\displaystyle+9c_{1}(\mathcal{O}_{\mathbb{P}^{2}})\cdot K_{\mathbb{P}^{2}}+6K_{\mathbb{P}^{2}}^{2}+3c_{1}(\mathcal{O}_{\mathbb{P}^{2}})^{2}.

Moreover, we have

c1()=c1(𝒪2)+c1(𝒪2(1))+c1(𝒪2(2))=c1(𝒪2(3))\displaystyle c_{1}(\mathcal{E})=c_{1}(\mathcal{O}_{\mathbb{P}^{2}})+c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(1))+c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(2))=c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(3))
c2()=c1(𝒪2(1))c1(𝒪2(2))=2\displaystyle c_{2}(\mathcal{E})=c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(1))\cdot c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(2))=2
c1(𝒪2)=0\displaystyle c_{1}(\mathcal{O}_{\mathbb{P}^{2}})=0

by [Har77, Appendix A, §3, C3 and C.5]. Hence we obtain

(KX)3=23222+40+63(3)+90+6(3)2+30=18454+54=14.(-K_{X})^{3}=2\cdot 3^{2}-2\cdot 2+4\cdot 0+6\cdot 3\cdot(-3)+9\cdot 0+6\cdot(-3)^{2}+3\cdot 0\\ =18-4-54+54=14.

Proposition 5.32.

We use Notation 5.26. Then there exist a split double cover

f:XV7f:X\to V_{7}

such that KXf(12KV7)K_{X}\sim f^{*}(\frac{1}{2}K_{V_{7}}), where V7V_{7} is the blowup of 3\mathbb{P}^{3} at a point and 12KV7\frac{1}{2}K_{V_{7}} is a Weil divisor on V7V_{7} satisfying 2(12KV7)KV72\cdot(\frac{1}{2}K_{V_{7}})\sim K_{V_{7}} (note that 12KV7\frac{1}{2}K_{V_{7}} is unique up to linear equivalence). In particular, for :=(f𝒪X/𝒪V7)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{V_{7}})^{-1}, we have 𝒪V7(12KV7)\mathcal{L}\simeq\mathcal{O}_{V_{7}}(-\frac{1}{2}K_{V_{7}}).

Proof.

Set SPS\subset P to be the section of π:P=2(𝒪2𝒪2(1)𝒪2(2))Z:=2\pi:P=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))\to Z:=\mathbb{P}^{2} corresponding to the natural projection:

𝒪2𝒪2(1)𝒪2(2)𝒪2.\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)\to\mathcal{O}_{\mathbb{P}^{2}}.

Let μ:P~P\mu:\widetilde{P}\to P be the blowup along SS.

Step 1.

There exists the following commutative diagram

(5.32.1) P~{\widetilde{P}}(𝒪2𝒪2(1)𝒪2(2))=P{\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))=P}(𝒪2(1)𝒪2(2))=V7{\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))=V_{7}}2=Z{\mathbb{P}^{2}=Z}μ\scriptstyle{\mu}π\scriptstyle{\pi}α\scriptstyle{\alpha}β\scriptstyle{\beta}

such that for any closed point z2=Zz\in\mathbb{P}^{2}=Z, the base change of (5.32.1) by z2=Zz\hookrightarrow\mathbb{P}^{2}=Z can be written as follows

(5.32.2) 𝔽1:=(𝒪1𝒪1(1))=P~z{\mathbb{F}_{1}:=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}}(1))=\widetilde{P}_{z}}2=Pz{\mathbb{P}^{2}=P_{z}}1=(V7)z{\mathbb{P}^{1}=(V_{7})_{z}}z{z}μz\scriptstyle{\mu_{z}}πz\scriptstyle{\pi_{z}}αz\scriptstyle{\alpha_{z}}βz\scriptstyle{\beta_{z}}

where μz\mu_{z} is the blowup at the closed point SzS_{z} and αz:𝔽11\alpha_{z}:\mathbb{F}_{1}\to\mathbb{P}^{1} is the 1\mathbb{P}^{1}-bundle.

Proof of Step 1.

We have the canonical injective graded 𝒪2\mathcal{O}_{\mathbb{P}^{2}}-algebra homomorhism

ι:Sym(𝒪2(1)𝒪2(2))Sym(𝒪2𝒪2(1)𝒪2(2)).\iota:{\rm Sym}(\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))\hookrightarrow{\rm Sym}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)).

Then ι\iota induces a dominant rational map φ\varphi over 2\mathbb{P}^{2} as follows:

P=2(𝒪2𝒪2(1)𝒪2(2)){P=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))}2(𝒪2(1)𝒪2(2))=V7{\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))=V_{7}}Z=2{Z=\mathbb{P}^{2}}π\scriptstyle{\pi}β\scriptstyle{\beta}φ\scriptstyle{\varphi}

Note that this construction commutes with base changes by open immersions ZZ=2Z^{\prime}\hookrightarrow Z=\mathbb{P}^{2}. Then the indeterminacies of φ\varphi are resolved by the blowup along SS, because this can be checked after taking an affine open cover of Z=2Z=\mathbb{P}^{2} trivialising 𝒪2(n)\mathcal{O}_{\mathbb{P}^{2}}(n). Therefore, we get a commutative diagram (5.32.1). This diagram fibrewisely induces the diagram (5.32.2), because we have the corresponding diagram over an arbitrary open subset of Z=2Z=\mathbb{P}^{2} which trivialises 𝒪2(n)\mathcal{O}_{\mathbb{P}^{2}}(n). This completes the proof of Step 1. ∎

Step 2.

It holds that XS=X\cap S=\emptyset. In particular, μ1(X)X\mu^{-1}(X)\simeq X.

Proof of Step 2.

Suppose XSX\cap S\neq\emptyset. Note that P=2(𝒪2𝒪2(1)𝒪2(2))P=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)) is covered by 𝔸4\mathbb{A}^{4}. Then it follows from [Har77, Ch. I, Proposition 7.1] that dim(XS)3+24=1\dim(X\cap S)\geq 3+2-4=1. Hence there exists a curve CC with CXSC\subset X\cap S. By 𝒪X(KX)𝒪P(1)|X\mathcal{O}_{X}(-K_{X})\simeq\mathcal{O}_{P}(1)|_{X} and 𝒪P(1)|S𝒪S\mathcal{O}_{P}(1)|_{S}\simeq\mathcal{O}_{S}, we obtain

KXC=(𝒪P(1)|X)C=𝒪P(1)C=(𝒪P(1)|S)C=𝒪SC=0.-K_{X}\cdot C=(\mathcal{O}_{P}(1)|_{X})\cdot C=\mathcal{O}_{P}(1)\cdot C=(\mathcal{O}_{P}(1)|_{S})\cdot C=\mathcal{O}_{S}\cdot C=0.

This contradicts the ampleness of KX-K_{X}. This completes the proof of Step 2. ∎

Step 3.

The induced morphism

f:Xμ1(X)α|μ1(X)V7f:X\xrightarrow{\simeq}\mu^{-1}(X)\xrightarrow{\alpha|_{\mu^{-1}(X)}}V_{7}

is a double cover.

Proof of Step 3.

Fix a closed point vV7v\in V_{7}. Set z:=β(v)Z=2z:=\beta(v)\in Z=\mathbb{P}^{2}. Take the base change by z2z\hookrightarrow\mathbb{P}^{2}. We then obtain the above diagram (5.32.2) and the fibre XzX_{z} is a conic in 2=Pz\mathbb{P}^{2}=P_{z}. It holds by Step 2 that XzX_{z} is disjoint from the (1)(-1)-curve on the blowup P~z=𝔽1\widetilde{P}_{z}=\mathbb{F}_{1} of Pz=2P_{z}=\mathbb{P}^{2} lying over the point SzS_{z}. We then have Xzζ=2X_{z}\cdot\zeta=2 for a fibre ζ\zeta of αz:P~z(V7)z\alpha_{z}:\widetilde{P}_{z}\to(V_{7})_{z}, because ζ\zeta is nothing but the proper transform of a line on 2=Pz\mathbb{P}^{2}=P_{z} passing through SzS_{z}. Since f1(v)f^{-1}(v) consists of at most two points, f:XV7f:X\to V_{7} is a finite surjective morphism. By Xzζ=2X_{z}\cdot\zeta=2, ff is a double cover. This completes the proof of Step 3. ∎

Step 4.

It holds that KXf(12KV7+βM)K_{X}\sim f^{*}(\frac{1}{2}K_{V_{7}}+\beta^{*}M) for some Cartier divisor MM on 2\mathbb{P}^{2}.

Proof of Step 4.

The above diagram (5.32.1) consists of smooth projective toric varieties and toric morphisms. Furthermore, each morphism is of relative Picard number one.

Set

N:=2(KP~+μX)+αKV7.N:=-2(K_{\widetilde{P}}+\mu^{*}X)+\alpha^{*}K_{V_{7}}.

Fix a closed point zZ=2z\in Z=\mathbb{P}^{2}. We then have the diagram (5.32.2). Take a fibre ζ1\zeta\simeq\mathbb{P}^{1} of the 1\mathbb{P}^{1}-bundle αz:𝔽1=P~z1=(V7)z\alpha_{z}:\mathbb{F}_{1}=\widetilde{P}_{z}\to\mathbb{P}^{1}=(V_{7})_{z} over a closed point. Note that μzXzζ=2\mu_{z}^{*}X_{z}\cdot\zeta=2 (cf. the proof of Step 3), and hence we get

(5.32.3) Nζ=(2(KP~+μX)+αKV7)ζ=2(2+2)+0=0N\cdot\zeta=(-2(K_{\widetilde{P}}+\mu^{*}X)+\alpha^{*}K_{V_{7}})\cdot\zeta=-2(-2+2)+0=0

and

(5.32.4) Nμ1(Xz)=(2(KP~+μX)+αKV7)μ1(Xz)=2degωXz+2degK(V7)z=44=0.N\cdot\mu^{-1}(X_{z})=(-2(K_{\widetilde{P}}+\mu^{*}X)+\alpha^{*}K_{V_{7}})\cdot\mu^{-1}(X_{z})=-2\deg\omega_{X_{z}}+2\deg K_{(V_{7})_{z}}=4-4=0.

By (5.32.3), there exists a Cartier divisor NN^{\prime} on V7V_{7} such that NαNN\sim\alpha^{*}N^{\prime}. It follows from (5.32.4) that we can find a Cartier divisor N′′N^{\prime\prime} on 2=Z\mathbb{P}^{2}=Z such that NβN′′N^{\prime}\sim\beta^{*}N^{\prime\prime}. For N′′′:=N′′N^{\prime\prime\prime}:=-N^{\prime\prime}, we get 2(KP~+μX)α(KV7+βN′′′)2(K_{\widetilde{P}}+\mu^{*}X)\sim\alpha^{*}(K_{V_{7}}+\beta^{*}N^{\prime\prime\prime}). Taking the restriction to μ1(X)(X)\mu^{-1}(X)(\simeq X), we have 2KXf(KV7+βN′′′)2K_{X}\sim f^{*}(K_{V_{7}}+\beta^{*}N^{\prime\prime\prime}). Then KXf(12KV7)K_{X}-f^{*}(\frac{1}{2}K_{V_{7}}) is numerically trivial over ZZ. Therefore, we can find a Cartier divisor MM on ZZ such that KX=f(12KV7+βM)K_{X}=f^{*}(\frac{1}{2}K_{V_{7}}+\beta^{*}M). This completes the proof of Step 4. ∎

Step 5.

It holds that KXf(12KV7)K_{X}\sim f^{*}(\frac{1}{2}K_{V_{7}}).

Proof of Step 5.

By Step 4, we have KXf(12KV7+βM)K_{X}\sim f^{*}(\frac{1}{2}K_{V_{7}}+\beta^{*}M).

We now show that degM0\deg M\leq 0, i.e., M-M is nef. For the exceptional divisor E2E\simeq\mathbb{P}^{2} of the blowup V73V_{7}\to\mathbb{P}^{3}, it holds that

𝒪V7(12KV7)|E𝒪E(1).\mathcal{O}_{V_{7}}\left(\frac{1}{2}K_{V_{7}}\right)\Big{|}_{E}\simeq\mathcal{O}_{E}(-1).

Pick a line E\ell\subset E. Since (12KV7+βM)-(\frac{1}{2}K_{V_{7}}+\beta^{*}M) is ample, we obtain

0<(12KV7+βM)=1βM.0<-\left(\frac{1}{2}K_{V_{7}}+\beta^{*}M\right)\cdot\ell=1-\beta^{*}M\cdot\ell.

Since EZ=2E\to Z=\mathbb{P}^{2} is a finite surjective morphism, we get degM0\deg M\leq 0.

By KXf(12KV7+βM)K_{X}\sim f^{*}(\frac{1}{2}K_{V_{7}}+\beta^{*}M), we have

8KX3=(f(KV7+2βM))3=2(KV73+6KV72βM+12KV7(βM)2),8K_{X}^{3}=(f^{*}(K_{V_{7}}+2\beta^{*}M))^{3}=2(K_{V_{7}}^{3}+6K_{V_{7}}^{2}\cdot\beta^{*}M+12K_{V_{7}}\cdot(\beta^{*}M)^{2}),

where the last equality holds by (βM)3=0(\beta^{*}M)^{3}=0. It follows from KX3=14K_{X}^{3}=-14 (Lemma 5.31) and KV73=K33+8=56K_{V_{7}}^{3}=K_{\mathbb{P}^{3}}^{3}+8=-56 that

KV72βM+2KV7(βM)2=0.K_{V_{7}}^{2}\cdot\beta^{*}M+2K_{V_{7}}\cdot(\beta^{*}M)^{2}=0.

Suppose that M≁0M\not\sim 0. By degM0\deg M\leq 0, we have degM<0\deg M<0, i.e., M-M is ample. We then have

KV72βM=(KV7)2(βM)<0andKV7(βM)2=(KV7)(βM)2<0,K_{V_{7}}^{2}\cdot\beta^{*}M=-(-K_{V_{7}})^{2}\cdot(-\beta^{*}M)<0\quad{\rm and}\quad K_{V_{7}}\cdot(\beta^{*}M)^{2}=-(-K_{V_{7}})\cdot(-\beta^{*}M)^{2}<0,

which is a contraction. Therefore, we obtain M0M\sim 0. This completes the proof of Step 5. ∎

Step 6.

The induced morphism f:XV7f:X\to V_{7} is a split double cover.

Proof of Step 6.

For a Cartier divisor LL satisying

0𝒪V7f𝒪X𝒪V7(L)0,0\to\mathcal{O}_{V_{7}}\to f_{*}\mathcal{O}_{X}\to\mathcal{O}_{V_{7}}(-L)\to 0,

it suffices to show that Ext1(𝒪V7(L),𝒪V7)=0{\operatorname{Ext}}^{1}(\mathcal{O}_{V_{7}}(-L),\mathcal{O}_{V_{7}})=0, which is equivalent to H1(V7,𝒪V7(L))=0H^{1}(V_{7},\mathcal{O}_{V_{7}}(L))=0. We have KXf(KV7+L)K_{X}\sim f^{*}(K_{V_{7}}+L) (Lemma 2.3). By Step 5, we get L12KV7L\sim-\frac{1}{2}K_{V_{7}}. Since V7V_{7} is toric and LL is ample, we get H1(V7,𝒪V7(L))=0H^{1}(V_{7},\mathcal{O}_{V_{7}}(L))=0 [Fuj07, Theorem 1.6]. This completes the proof of Step 6. ∎

Step 5 and Step 6 complete the proof of Proposition 5.32. ∎

5.4.2. Case C2EC_{2}-E

Lemma 5.33.

We use Notation 5.1. Assume that R1R_{1} is of type C2C_{2} and R2R_{2} is of type E2,E3,E4E_{2},E_{3},E_{4}, or E5E_{5}. Then the following hold.

  1. (1)

    R2R_{2} is of type E2E_{2} or E5E_{5}.

  2. (2)

    If R2R_{2} is of type E2E_{2}, then X(𝒪2𝒪2(1))X\simeq\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)) and (KX)3=56(-K_{X})^{3}=56.

  3. (3)

    If R2R_{2} is of type E5E_{5}, then X(𝒪2𝒪2(2))X\simeq\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)) and (KX)3=62(-K_{X})^{3}=62.

Proof.

Set D:=Ex(f2:XY2)D:={\operatorname{Ex}}(f_{2}:X\to Y_{2}). Since f1:X2f_{1}\colon X\to\mathbb{P}^{2} is a 1\mathbb{P}^{1}-bundle, we can write

f1:X=2()2f_{1}\colon X=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{E})\to\mathbb{P}^{2}

for some locally free sheaf \mathcal{E} of rank 2 on 2\mathbb{P}^{2}. By Lemma 5.25, f1|D:D2f_{1}|_{D}\colon D\to\mathbb{P}^{2} is an isomorphism and R2R_{2} is of type E2E_{2} or E5E_{5}. In particular, (1) holds. We have 𝒪X(D)𝒪X(n)(f1)\mathcal{O}_{X}(D)\simeq\mathcal{O}_{X}(n)\otimes(f_{1})^{*}\mathcal{L} for some nn\in\mathbb{Z} and Pic2\mathcal{L}\in{\operatorname{Pic}}\,\mathbb{P}^{2}. Since DF=1D\cdot F=1 for a fibre F1F\simeq\mathbb{P}^{1} of f1:X=2()2f_{1}\colon X=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{E})\to\mathbb{P}^{2}, we obtain n=1n=1, i.e.,

(5.33.1) 𝒪X(D)𝒪X(1)(f1).\mathcal{O}_{X}(D)\simeq\mathcal{O}_{X}(1)\otimes(f_{1})^{*}\mathcal{L}.

We have the following exact sequence

0𝒪X𝒪X(D)j(𝒪X(D)|D)0,0\longrightarrow\mathcal{O}_{X}\longrightarrow\mathcal{O}_{X}(D)\longrightarrow j_{*}(\mathcal{O}_{X}(D)|_{D})\longrightarrow 0,

where j:DXj:D\hookrightarrow X denotes the induced closed immersion. By R1(f1)𝒪X=0R^{1}(f_{1})_{*}\mathcal{O}_{X}=0 (Proposition 3.11), we obtain the exact sequence

(5.33.2) 0(f1)𝒪X(f1)𝒪X(D)(f1)j(𝒪X(D)|D)0.0\longrightarrow(f_{1})_{*}\mathcal{O}_{X}\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(D)\longrightarrow(f_{1})_{*}j_{*}(\mathcal{O}_{X}(D)|_{D})\longrightarrow 0.

We have (f1)𝒪X=𝒪2(f_{1})_{*}\mathcal{O}_{X}=\mathcal{O}_{\mathbb{P}^{2}}. By (5.33.1) and (f1)𝒪X(1)(f_{1})_{*}\mathcal{O}_{X}(1)\simeq\mathcal{E} [Har77, Ch. II, Proposition 7.11], it holds that (f1)𝒪X(D)(f_{1})_{*}\mathcal{O}_{X}(D)\simeq\mathcal{E}\otimes\mathcal{L}. If R2R_{2} is of type E2E_{2} (resp. E5E_{5}), then we set e:=1e:=1 (resp. e:=2e:=2). By Theorem 4.8, we have 𝒪X(D)|D𝒪2(e)\mathcal{O}_{X}(D)|_{D}\simeq\mathcal{O}_{\mathbb{P}^{2}}(-e). Since f1j:D2f_{1}\circ j:D\to\mathbb{P}^{2} is an isomorphism, we get (f1)j(𝒪X(D)|D)𝒪2(e)(f_{1})_{*}j_{*}(\mathcal{O}_{X}(D)|_{D})\simeq\mathcal{O}_{\mathbb{P}^{2}}(-e). Hence, the exact sequence (5.33.2) can be written as follows:

0𝒪2𝒪2(e)0.0\longrightarrow\mathcal{O}_{\mathbb{P}^{2}}\longrightarrow\mathcal{E}\otimes\mathcal{L}\longrightarrow\mathcal{O}_{\mathbb{P}^{2}}(-e)\longrightarrow 0.

By Ext1(𝒪2(e),𝒪2)H1(2,𝒪2(e))=0\mathrm{Ext}^{1}(\mathcal{O}_{\mathbb{P}^{2}}(-e),\mathcal{O}_{\mathbb{P}^{2}})\simeq H^{1}(\mathbb{P}^{2},\mathcal{O}_{\mathbb{P}^{2}}(e))=0, we obtain 𝒪2𝒪2(e)\mathcal{E}\otimes\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(-e). Therefore, it hold that

X=2()2(𝒪2(e))2(𝒪2𝒪2(e)).X=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{E})\simeq\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{E}\otimes\mathcal{L}\otimes\mathcal{O}_{\mathbb{P}^{2}}(e))\simeq\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\otimes\mathcal{O}_{\mathbb{P}^{2}}(e)).

What is remaining is to compute (KX)3(-K_{X})^{3}. By Proposition 7.1(3), we have

(KX)3=2c1(𝒪2𝒪2(e))28c2(𝒪2𝒪2(e))+6K22.(-K_{X})^{3}=2c_{1}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(e))^{2}-8c_{2}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(e))+6K_{\mathbb{P}^{2}}^{2}.

By [Har77, Appendix A], it holds that

c1(𝒪2𝒪2(e))=c1(𝒪2)+c1(𝒪2(e))=c1(𝒪2(e)),\displaystyle c_{1}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(e))=c_{1}(\mathcal{O}_{\mathbb{P}^{2}})+c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(e))=c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(e)),
c2(𝒪2𝒪2(e))=c1(𝒪2)c1(𝒪2(e))=0.\displaystyle c_{2}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(e))=c_{1}(\mathcal{O}_{\mathbb{P}^{2}})\cdot c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(e))=0.

Therefore, we obtain

(KX)3\displaystyle(-K_{X})^{3} =2c1(𝒪2𝒪2(e))28c2(𝒪2𝒪2(e))+6K22\displaystyle=2c_{1}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(e))^{2}-8c_{2}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(e))+6K_{\mathbb{P}^{2}}^{2}
=2e280+69\displaystyle=2e^{2}-8\cdot 0+6\cdot 9
=2e2+54.\displaystyle=2e^{2}+54.

Theorem 5.34.

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2. Let R1R_{1} and R2R_{2} be extremal rays of NE(X){\operatorname{NE}}(X). Assume that R1R_{1} is of type CC and R2R_{2} is of type E2,E3,E4E_{2},E_{3},E_{4}, or E5E_{5}. Then one and only one of the following holds.

  1. (1)

    R1R_{1} is of type C1C_{1}, R2R_{2} is of type E3E_{3} or E4E_{4}, (KX)3=14(-K_{X})^{3}=14, and there exists a split double cover f:XV7f:X\to V_{7} such that f𝒪X/𝒪V7𝒪V7(12KV7)1f_{*}\mathcal{O}_{X}/\mathcal{O}_{V_{7}}\simeq\mathcal{O}_{V_{7}}(-\frac{1}{2}K_{V_{7}})^{-1}.

  2. (2)

    R1R_{1} is of type C2C_{2}, R2R_{2} is of type E2E_{2}, (KX)3=56(-K_{X})^{3}=56, and X(𝒪2𝒪2(1))X\simeq\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1)).

  3. (3)

    R1R_{1} is of type C2C_{2}, R2R_{2} is of type E5E_{5}, (KX)3=62(-K_{X})^{3}=62, and X(𝒪2𝒪2(2))X\simeq\mathbb{P}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2)).

Proof.

The assertion follows from Lemma 5.25, Proposition 5.32, and Lemma 5.33. ∎

5.5. Case E1CE_{1}-C

Lemma 5.35.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1}. Then

g(B1)=(KX)32(KY1)32+r1degB1+1g(B_{1})=\frac{(-K_{X})^{3}}{2}-\frac{(-K_{Y_{1}})^{3}}{2}+r_{1}\deg B_{1}+1
Proof.

See, e.g., [TanII, Lemma 3.21(2)(a)]. ∎

Lemma 5.36.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type CC. Then the following hold.

  1. (A)

    18=μ2(24r1+degB1)+degΔf218=\mu_{2}\left(\frac{24}{r_{1}}+\deg B_{1}\right)+\deg\Delta_{f_{2}}.

  2. (B)

    H12H2=(r1μ2)L13=1μ22(8degΔf2)H_{1}^{2}\cdot H_{2}=(r_{1}-\mu_{2})L_{1}^{3}=\frac{1}{\mu_{2}^{2}}(8-\deg\Delta_{f_{2}}).

  3. (C)

    H1H22=(r1μ2)2L13degB1=2μ2H_{1}\cdot H_{2}^{2}=(r_{1}-\mu_{2})^{2}L_{1}^{3}-\deg B_{1}=\frac{2}{\mu_{2}}.

Proof.

The assertion (A) follows from Lemma 5.4 and Proposition 5.9(4). The assertion (C) holds by Lemma 5.11 and

2=(KX)H22=(μ2H1+H2)H22=μ2H1H22.2=(-K_{X})\cdot H_{2}^{2}=(\mu_{2}H_{1}+H_{2})\cdot H_{2}^{2}=\mu_{2}H_{1}\cdot H_{2}^{2}.

The assertion (B) follows from Lemma 5.11 and

12degΔf2=(KX)2H2=(μ2H1+H2)2H2=μ22H12H2+2μ2H1H22=μ22H12H2+4,12-\deg\Delta_{f_{2}}=(-K_{X})^{2}\cdot H_{2}=(\mu_{2}H_{1}+H_{2})^{2}\cdot H_{2}=\mu_{2}^{2}H_{1}^{2}\cdot H_{2}+2\mu_{2}H_{1}\cdot H_{2}^{2}=\mu_{2}^{2}H_{1}^{2}\cdot H_{2}+4,

where the first equality is guaranteed by Proposition 3.16(3). ∎

Proposition 5.37.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type C1C_{1}. Then one of the following holds.

  1. (1)

    (KX)3=16,r1=4,degB1=7,g(B1)=5,degΔf2=5(-K_{X})^{3}=16,r_{1}=4,\deg B_{1}=7,g(B_{1})=5,\deg\Delta_{f_{2}}=5 (No. 2-9).

  2. (2)

    (KX)3=20,r1=3,degB1=6,g(B1)=2,degΔf2=4(-K_{X})^{3}=20,r_{1}=3,\deg B_{1}=6,g(B_{1})=2,\deg\Delta_{f_{2}}=4 (No. 2-13).

  3. (3)

    (KX)3=18,r1=2,(KX1)3=24,degB1=1,g(B1)=0,degΔf2=5(-K_{X})^{3}=18,r_{1}=2,(-K_{X_{1}})^{3}=24,\deg B_{1}=1,g(B_{1})=0,\deg\Delta_{f_{2}}=5 (No. 2-11).

  4. (4)

    (KX)3=22,r1=2,(KX1)3=32,degB1=2,g(B1)=0,degΔf2=4(-K_{X})^{3}=22,r_{1}=2,(-K_{X_{1}})^{3}=32,\deg B_{1}=2,g(B_{1})=0,\deg\Delta_{f_{2}}=4 (No. 2-16).

  5. (5)

    (KX)3=26,r1=2,(KX1)3=40,degB1=3,g(B1)=0,degΔf2=3(-K_{X})^{3}=26,r_{1}=2,(-K_{X_{1}})^{3}=40,\deg B_{1}=3,g(B_{1})=0,\deg\Delta_{f_{2}}=3 (No. 2-20).

Proof.

We have μ2=1\mu_{2}=1. By Lemma 5.36, the following hold.

  1. (A)

    18=24r1+degB1+degΔf218=\frac{24}{r_{1}}+\deg B_{1}+\deg\Delta_{f_{2}}.

  2. (B)

    H12H2=(r11)L13=8degΔf2H_{1}^{2}\cdot H_{2}=(r_{1}-1)L_{1}^{3}=8-\deg\Delta_{f_{2}}.

  3. (C)

    H1H22=(r11)2L13degB1=2H_{1}\cdot H_{2}^{2}=(r_{1}-1)^{2}L_{1}^{3}-\deg B_{1}=2.

Assume r1=4r_{1}=4, and hence L13=1L_{1}^{3}=1. Then (B) and (C) imply degΔf2=5\deg\Delta_{f_{2}}=5 and degB1=7\deg B_{1}=7, respectively. Hence the following hold (Lemma 5.35):

(KX)3=(H1+H2)3=H13+3H12H2+3H1H22+H23=L13+9+6+0=16(-K_{X})^{3}=(H_{1}+H_{2})^{3}=H_{1}^{3}+3H_{1}^{2}\cdot H_{2}+3H_{1}\cdot H_{2}^{2}+H_{2}^{3}=L_{1}^{3}+9+6+0=16
g(B1)=832+47+1=5.g(B_{1})=8-32+4\cdot 7+1=5.

Assume r1=3r_{1}=3, and hence L13=2L_{1}^{3}=2. Then (B) and (C) imply degΔf2=4\deg\Delta_{f_{2}}=4 and degB1=6\deg B_{1}=6, respectively. Hence the following hold (Lemma 5.35):

(KX)3=(H1+H2)3=H13+3H12H2+3H1H22+H23=2+12+6+0=20(-K_{X})^{3}=(H_{1}+H_{2})^{3}=H_{1}^{3}+3H_{1}^{2}\cdot H_{2}+3H_{1}\cdot H_{2}^{2}+H_{2}^{3}=2+12+6+0=20
g(B1)=1027+36+1=2.g(B_{1})=10-27+3\cdot 6+1=2.

Assume r1=2r_{1}=2, and hence 1L1351\leq L_{1}^{3}\leq 5. Then

  1. (A)

    6=degB1+degΔf26=\deg B_{1}+\deg\Delta_{f_{2}}.

  2. (B)

    H12H2=L13=8degΔf2H_{1}^{2}\cdot H_{2}=L_{1}^{3}=8-\deg\Delta_{f_{2}}.

  3. (C)

    H1H22=L13degB1=2H_{1}\cdot H_{2}^{2}=L_{1}^{3}-\deg B_{1}=2.

By (C), we get L13=2+degB13L_{1}^{3}=2+\deg B_{1}\geq 3.

Assume (r1,L13)=(2,3)(r_{1},L_{1}^{3})=(2,3). Then (B) and (C) imply degΔf2=5\deg\Delta_{f_{2}}=5 and degB1=1\deg B_{1}=1, respectively. In particular, g(B1)=0g(B_{1})=0. Hence the following hold:

(KX)3=(H1+H2)3=H13+3H12H2+3H1H22+H23=3+9+6+0=18(-K_{X})^{3}=(H_{1}+H_{2})^{3}=H_{1}^{3}+3H_{1}^{2}\cdot H_{2}+3H_{1}\cdot H_{2}^{2}+H_{2}^{3}=3+9+6+0=18

Assume (r1,L13)=(2,4)(r_{1},L_{1}^{3})=(2,4). Then (B) and (C) imply degΔf2=4\deg\Delta_{f_{2}}=4 and degB1=2\deg B_{1}=2, respectively. In particular, g(B1)=0g(B_{1})=0. Hence the following hold:

(KX)3=(H1+H2)3=H13+3H12H2+3H1H22+H23=4+12+6+0=22.(-K_{X})^{3}=(H_{1}+H_{2})^{3}=H_{1}^{3}+3H_{1}^{2}\cdot H_{2}+3H_{1}\cdot H_{2}^{2}+H_{2}^{3}=4+12+6+0=22.

Assume (r1,L13)=(2,5)(r_{1},L_{1}^{3})=(2,5). Then (B) and (C) imply degΔf2=3\deg\Delta_{f_{2}}=3 and degB1=3\deg B_{1}=3, respectively. Hence the following hold (Lemma 5.35):

(KX)3=(H1+H2)3=H13+3H12H2+3H1H22+H23=5+15+6+0=26.(-K_{X})^{3}=(H_{1}+H_{2})^{3}=H_{1}^{3}+3H_{1}^{2}\cdot H_{2}+3H_{1}\cdot H_{2}^{2}+H_{2}^{3}=5+15+6+0=26.
g(B1)=1320+23+1=0.g(B_{1})=13-20+2\cdot 3+1=0.

Proposition 5.38.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type C2C_{2}. Then one of the following holds.

  1. (1)

    (KX)3=38,r1=4,degB1=3,g(B1)=0(-K_{X})^{3}=38,r_{1}=4,\deg B_{1}=3,g(B_{1})=0 (No. 2-27).

  2. (2)

    (KX)3=46,r1=3,degB1=1,g(B1)=0(-K_{X})^{3}=46,r_{1}=3,\deg B_{1}=1,g(B_{1})=0 (No. 2-31).

Proof.

We have μ2=2\mu_{2}=2 and Δf2=0\Delta_{f_{2}}=0. By Lemma 5.36, the following hold.

  1. (A)

    18=48r1+2degB118=\frac{48}{r_{1}}+2\deg B_{1}.

  2. (B)

    H12H2=(r12)L13=2H_{1}^{2}\cdot H_{2}=(r_{1}-2)L_{1}^{3}=2.

  3. (C)

    H1H22=(r12)2L13degB1=1H_{1}\cdot H_{2}^{2}=(r_{1}-2)^{2}L_{1}^{3}-\deg B_{1}=1.

By (B), we obtain r1{3,4}r_{1}\in\{3,4\}. Moreover, (B) and (C) imply

(KX)3=(2H1+H2)3=8H13+12H12H2+6H1H22+H23=8L13+122+61+0=8L13+30.(-K_{X})^{3}=(2H_{1}+H_{2})^{3}=8H_{1}^{3}+12H_{1}^{2}\cdot H_{2}+6H_{1}\cdot H_{2}^{2}+H_{2}^{3}=8L_{1}^{3}+12\cdot 2+6\cdot 1+0=8L_{1}^{3}+30.

Assume r1=4r_{1}=4. Then L13=1L_{1}^{3}=1 and (KX)3=8L13+30=38(-K_{X})^{3}=8L_{1}^{3}+30=38. Moreover, (A) implies degB1=3\deg B_{1}=3. Hence the following hold (Lemma 5.35):

g(B1)=1932+43+1=0.g(B_{1})=19-32+4\cdot 3+1=0.

Assume r1=3r_{1}=3. Then L13=2L_{1}^{3}=2 and (KX)3=8L13+30=46(-K_{X})^{3}=8L_{1}^{3}+30=46. Moreover, (A) implies degB1=1\deg B_{1}=1. Hence the following hold (Lemma 5.35):

g(B1)=2327+31+1=0.g(B_{1})=23-27+3\cdot 1+1=0.

5.6. Case E1DE_{1}-D

Lemma 5.39.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type DD. Set d2:=(KX)2H2d_{2}:=(-K_{X})^{2}\cdot H_{2}. Then the following hold.

  1. (A)

    12=μ2(24r1+degB1)d212=\mu_{2}\left(\frac{24}{r_{1}}+\deg B_{1}\right)-d_{2}.

  2. (B)

    H12H2=(r1μ2)L13=d2μ22H_{1}^{2}\cdot H_{2}=(r_{1}-\mu_{2})L_{1}^{3}=\frac{d_{2}}{\mu_{2}^{2}}.

  3. (C)

    H1H22=(r1μ2)2L13degB1=0H_{1}\cdot H_{2}^{2}=(r_{1}-\mu_{2})^{2}L_{1}^{3}-\deg B_{1}=0.

Proof.

The assertion (A) follows from Lemma 5.4 and Proposition 5.9(4). The assertion (C) holds by Lemma 5.11 and H220H_{2}^{2}\equiv 0. The assertion (B) follows from Lemma 5.11 and

d2=(KX)2H2=(μ2H1+H2)2H2=μ22H12H2.d_{2}=(-K_{X})^{2}\cdot H_{2}=(\mu_{2}H_{1}+H_{2})^{2}\cdot H_{2}=\mu_{2}^{2}H_{1}^{2}\cdot H_{2}.

Proposition 5.40.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type D3D_{3}. Then (KX)3=54,r1=4,degB1=1,g(B1)=0(-K_{X})^{3}=54,r_{1}=4,\deg B_{1}=1,g(B_{1})=0 (No. 2-33).

Proof.

We have d2=(KX)2H2=9d_{2}=(-K_{X})^{2}\cdot H_{2}=9. By Lemma 5.39, the following hold:

  1. (A)

    7=24r1+degB17=\frac{24}{r_{1}}+\deg B_{1}.

  2. (B)

    H12H2=(r13)L13=1H_{1}^{2}\cdot H_{2}=(r_{1}-3)L_{1}^{3}=1.

  3. (C)

    H1H22=(r13)2L13degB1=0H_{1}\cdot H_{2}^{2}=(r_{1}-3)^{2}L_{1}^{3}-\deg B_{1}=0.

By (B), we get r1=4r_{1}=4 and L13=1L_{1}^{3}=1. Then (C) implies degB1=1\deg B_{1}=1. Hence the following holds:

(KX)3=(3H1+H2)3=27H13+27H12H2+9H1H22+H33=27+27+0+0=54.(-K_{X})^{3}=(3H_{1}+H_{2})^{3}=27H_{1}^{3}+27H_{1}^{2}\cdot H_{2}+9H_{1}\cdot H_{2}^{2}+H_{3}^{3}=27+27+0+0=54.

Proposition 5.41.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type D2D_{2}. Then one of the following holds.

  1. (1)

    (KX)3=32,r1=4,degB1=4,g(B1)=1(-K_{X})^{3}=32,r_{1}=4,\deg B_{1}=4,g(B_{1})=1 (No. 2-25).

  2. (2)

    (KX)3=40,r1=3,degB1=2,g(B1)=0(-K_{X})^{3}=40,r_{1}=3,\deg B_{1}=2,g(B_{1})=0 (No. 2-29).

Proof.

We have d2=8d_{2}=8. By Lemma 5.39, the following hold:

  1. (A)

    10=24r1+degB110=\frac{24}{r_{1}}+\deg B_{1}.

  2. (B)

    H12H2=(r12)L13=2H_{1}^{2}\cdot H_{2}=(r_{1}-2)L_{1}^{3}=2.

  3. (C)

    H1H22=(r12)2L13degB1=0H_{1}\cdot H_{2}^{2}=(r_{1}-2)^{2}L_{1}^{3}-\deg B_{1}=0.

By (B), r13r_{1}\geq 3. Moreover, we get

(KX)3=(2H1+H2)3=8H13+12H12H2+6H1H22+H23=8L13+24.(-K_{X})^{3}=(2H_{1}+H_{2})^{3}=8H_{1}^{3}+12H_{1}^{2}\cdot H_{2}+6H_{1}\cdot H_{2}^{2}+H_{2}^{3}=8L_{1}^{3}+24.

Assume r1=4r_{1}=4. We then get L13=1L_{1}^{3}=1 and (KX)3=8+24=32(-K_{X})^{3}=8+24=32. Moreover, (A) implies degB1=4\deg B_{1}=4. Hence the following hold (Lemma 5.35):

g(B1)=1632+44+1=1.g(B_{1})=16-32+4\cdot 4+1=1.

Assume r1=3r_{1}=3. We then get L13=2L_{1}^{3}=2 and (KX)3=16+24=40(-K_{X})^{3}=16+24=40. Moreover, (A) implies degB1=2\deg B_{1}=2. In particular, g(B1)=0g(B_{1})=0. ∎

Proposition 5.42.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type D1D_{1}. Then one of the following holds.

  1. (1)

    (KX)3=10,r1=4,degB1=9,g(B1)=10,(KX)2H2=3(-K_{X})^{3}=10,r_{1}=4,\deg B_{1}=9,g(B_{1})=10,(-K_{X})^{2}\cdot H_{2}=3 (No. 2-4).

  2. (2)

    (KX)3=14,r1=3,degB1=8,g(B1)=5,(KX)2H2=4(-K_{X})^{3}=14,r_{1}=3,\deg B_{1}=8,g(B_{1})=5,(-K_{X})^{2}\cdot H_{2}=4 (No. 2-7).

  3. (3)

    (KX)3=4,r1=2,(KY1)3=8,degB1=1,g(B1)=1,(KX)2H2=1(-K_{X})^{3}=4,r_{1}=2,(-K_{Y_{1}})^{3}=8,\deg B_{1}=1,g(B_{1})=1,(-K_{X})^{2}\cdot H_{2}=1 (No. 2-1).

  4. (4)

    (KX)3=8,r1=2,(KY1)3=16,degB1=2,g(B1)=1,(KX)2H2=2(-K_{X})^{3}=8,r_{1}=2,(-K_{Y_{1}})^{3}=16,\deg B_{1}=2,g(B_{1})=1,(-K_{X})^{2}\cdot H_{2}=2 (No. 2-3).

  5. (5)

    (KX)3=12,r1=2,(KY1)3=24,degB1=3,g(B1)=1,(KX)2H2=3(-K_{X})^{3}=12,r_{1}=2,(-K_{Y_{1}})^{3}=24,\deg B_{1}=3,g(B_{1})=1,(-K_{X})^{2}\cdot H_{2}=3 (No. 2-5).

  6. (6)

    (KX)3=16,r1=2,(KY1)3=32,degB1=4,g(B1)=1,(KX)2H2=4(-K_{X})^{3}=16,r_{1}=2,(-K_{Y_{1}})^{3}=32,\deg B_{1}=4,g(B_{1})=1,(-K_{X})^{2}\cdot H_{2}=4 (No. 2-10).

  7. (7)

    (KX)3=20,r1=2,(KY1)3=40,degB1=5,g(B1)=1,(KX)2H2=5(-K_{X})^{3}=20,r_{1}=2,(-K_{Y_{1}})^{3}=40,\deg B_{1}=5,g(B_{1})=1,(-K_{X})^{2}\cdot H_{2}=5 (No. 2-14).

Proof.

By Lemma 5.39, the following hold:

  1. (A)

    12+d2=24r1+degB112+d_{2}=\frac{24}{r_{1}}+\deg B_{1}.

  2. (B)

    H12H2=(r11)L13=d2H_{1}^{2}\cdot H_{2}=(r_{1}-1)L_{1}^{3}=d_{2}.

  3. (C)

    H1H22=(r11)2L13degB1=0H_{1}\cdot H_{2}^{2}=(r_{1}-1)^{2}L_{1}^{3}-\deg B_{1}=0.

It holds that

(KX)3=(H1+H2)3=H13+3H12H2+3H1H22+H33=L13+3d2.(-K_{X})^{3}=(H_{1}+H_{2})^{3}=H_{1}^{3}+3H_{1}^{2}\cdot H_{2}+3H_{1}\cdot H_{2}^{2}+H_{3}^{3}=L_{1}^{3}+3d_{2}.

Assume r1=4r_{1}=4, and hence L13=1L_{1}^{3}=1. Then (B) and (C) imply d2=3d_{2}=3 and degB1=9\deg B_{1}=9, respectively. In particular, (KX)3=1+33=10(-K_{X})^{3}=1+3\cdot 3=10. Hence the following hold (Lemma 5.35):

g(B1)=532+49+1=10.g(B_{1})=5-32+4\cdot 9+1=10.

Assume r1=3r_{1}=3, and hence L13=2L_{1}^{3}=2. Then (B) and (C) imply d2=4d_{2}=4 and degB1=8\deg B_{1}=8, respectively. In particular, (KX)3=2+34=14(-K_{X})^{3}=2+3\cdot 4=14. Hence the following hold (Lemma 5.35):

g(B1)=727+38+1=5.g(B_{1})=7-27+3\cdot 8+1=5.

Assume r1=2r_{1}=2, and hence 1L1351\leq L_{1}^{3}\leq 5. Then

  1. (A)

    d2=degB1d_{2}=\deg B_{1}.

  2. (B)

    H12H2=L13=d2H_{1}^{2}\cdot H_{2}=L_{1}^{3}=d_{2}.

  3. (C)

    H1H22=L13degB1=0H_{1}\cdot H_{2}^{2}=L_{1}^{3}-\deg B_{1}=0.

We get (KX)3=L13+3d2=4L13(-K_{X})^{3}=L_{1}^{3}+3d_{2}=4L_{1}^{3}. Then the following hold (Lemma 5.35):

g(B1)=2L134L13+2L13+1=1.g(B_{1})=2L_{1}^{3}-4L_{1}^{3}+2\cdot L_{1}^{3}+1=1.

Lemma 5.43.

Let σ:XY\sigma:X\to Y be a blowup along a smooth curve Γ\Gamma on a Fano threefold YY. Let π:X1\pi:X\to\mathbb{P}^{1} be a morphism with π𝒪X=𝒪1\pi_{*}\mathcal{O}_{X}=\mathcal{O}_{\mathbb{P}^{1}}. Take a Cartier divisor DD on YY. Assume that (i) and (ii) hold.

  1. (i)

    (KY)D2=(KY)Γ(-K_{Y})\cdot D^{2}=(-K_{Y})\cdot\Gamma.

  2. (ii)

    DσFD\sim\sigma_{*}F for a fibre FF of π:X1\pi:X\to\mathbb{P}^{1}.

Then Γ\Gamma is a complete intersection of two members of |D||D|.

Proof.

Take two fibres F1F_{1} and F2F_{2} of π:X1\pi:X\to\mathbb{P}^{1}. Set D1:=σF1D_{1}:=\sigma_{*}F_{1} and D2:=σF2D_{2}:=\sigma_{*}F_{2}. For each i{1,2}i\in\{1,2\}, we get Di=σFiσF(ii)DD_{i}=\sigma_{*}F_{i}\sim\sigma_{*}F\overset{{\rm(ii)}}{\sim}D. We then obtain Γ(D1D2)red\Gamma\supset(D_{1}\cap D_{2})_{{\operatorname{red}}}. Since D1D2D_{1}\cap D_{2} is of pure one-dimensional, we get (D1D2)red=Γ(D_{1}\cap D_{2})_{{\operatorname{red}}}=\Gamma or (D1D2)red=(D_{1}\cap D_{2})_{{\operatorname{red}}}=\emptyset. We then obtain (D1D2)red=Γ(D_{1}\cap D_{2})_{{\operatorname{red}}}=\Gamma, as otherwise the equality (D1D2)red=(D_{1}\cap D_{2})_{{\operatorname{red}}}=\emptyset would lead to the following contradiction:

0=(KY)(D1D2)=(KY)D1D2=(KY)D2=(i)(KY)Γ>0.0=(-K_{Y})\cdot(D_{1}\cap D_{2})=(-K_{Y})\cdot D_{1}\cdot D_{2}=(-K_{Y})\cdot D^{2}\overset{{\rm(i)}}{=}(-K_{Y})\cdot\Gamma>0.

It is enough to show the scheme-theoretic equality D1D2=ΓD_{1}\cap D_{2}=\Gamma. Recall that we have the equality (D1D2)red=Γ(D_{1}\cap D_{2})_{{\operatorname{red}}}=\Gamma. Since D1D2D_{1}\cap D_{2} is Cohen-Macaulay, it suffices to prove that D1D2D_{1}\cap D_{2} is R0R_{0}, i.e., 𝒪D1D2,ξ\mathcal{O}_{D_{1}\cap D_{2},\xi} is reduced for the generic point ξ\xi of D1D2D_{1}\cap D_{2}. This follows from (KY)(D1D2)=KYΓ(-K_{Y})\cdot(D_{1}\cap D_{2})=-K_{Y}\cdot\Gamma, because the effective 11-cycle associated with D1D2D_{1}\cap D_{2} is determined by its length at the generic point [Bad01, Lemma 1.18]. ∎

Proposition 5.44.

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2 which has extremal rays of type DD and E1E_{1}. Let σ:XY\sigma:X\to Y be the contraction of the extremal ray of type E1E_{1}. Then the blowup centre σ(Ex(σ))\sigma({\operatorname{Ex}}(\sigma)) of σ\sigma is a complete intersection D1D2D_{1}\cap D_{2} of two members D1D_{1} and D2D_{2} of |D||D| for some Cartier divisor DD on YY.

In the following proof, we use No. of XX given in Proposition 5.40, Proposition 5.41, and Proposition 5.42. In particular, No. of XX is one of 2-1, 2-3, 2-4, 2-5, 2-7, 2-10, 2-14, 2-25, 2-29, 2-33.

Proof.

Set Γ:=σ(Ex(σ))\Gamma:=\sigma({\operatorname{Ex}}(\sigma)). Let π:X1\pi:X\to\mathbb{P}^{1} be the contraction of the extremal ray of type DD. Take a fibre FF of π\pi. Set μ\mu to be the length of π\pi. Then π:X1\pi:X\to\mathbb{P}^{1} is of type DμD_{\mu} (Remark 3.4(2)). For each case, it is enough to find a Cartier divisor DD on YY satisfying (i) and (ii) of Lemma 5.43. In what follows, we treat the following three cases separately.

  1. (1)

    2-4, 2-25, 2-33.

  2. (2)

    2-7, 2-29.

  3. (3)

    2-1, 2-3, 2-5, 2-10, 2-14.

(1) In these cases, we have Y=3Y=\mathbb{P}^{3} and degΓ=s2\deg\Gamma=s^{2} for some 1s31\leq s\leq 3 (Proposition 5.40, Proposition 5.41, Proposition 5.42). Let DD be a Cartier divisor satisfying 𝒪3(D)𝒪3(s)\mathcal{O}_{\mathbb{P}^{3}}(D)\simeq\mathcal{O}_{\mathbb{P}^{3}}(s). Then Lemma 5.43(i) holds by (KY)D2=4s2=(KY)Γ(-K_{Y})\cdot D^{2}=4s^{2}=(-K_{Y})\cdot\Gamma. It holds that KXσ𝒪3(μ)+F-K_{X}\sim\sigma^{*}\mathcal{O}_{\mathbb{P}^{3}}(\mu)+F (Proposition 5.9(3)). Hence σFKY𝒪3(μ)𝒪3(4μ)\sigma_{*}F\sim-K_{Y}-\mathcal{O}_{\mathbb{P}^{3}}(\mu)\sim\mathcal{O}_{\mathbb{P}^{3}}(4-\mu). In order to prove Lemma 5.43(ii), it suffices to show s+μ=4s+\mu=4. By using the fact that π\pi is of type DμD_{\mu}, this can be checked for all the cases 2-4, 2-25, 2-33 (Proposition 5.40, Proposition 5.41, Proposition 5.42).

(2) In these cases, we have Y=QY=Q and degΓ=2s2\deg\Gamma=2s^{2} for some 1s21\leq s\leq 2 (Proposition 5.41, Proposition 5.42). Let DD be a Cartier divisor satisfying 𝒪Q(D)𝒪Q(s)\mathcal{O}_{Q}(D)\simeq\mathcal{O}_{Q}(s). Then Lemma 5.43(i) holds by (KY)D2=6s2=(KY)Γ(-K_{Y})\cdot D^{2}=6s^{2}=(-K_{Y})\cdot\Gamma. It holds that KXσ𝒪Q(μ)+F-K_{X}\sim\sigma^{*}\mathcal{O}_{Q}(\mu)+F (Proposition 5.9(3)). Hence σFKY𝒪Q(μ)𝒪Q(3μ)\sigma_{*}F\sim-K_{Y}-\mathcal{O}_{Q}(\mu)\sim\mathcal{O}_{Q}(3-\mu). In order to prove Lemma 5.43(ii), it suffices to show s+μ=3s+\mu=3. By using the fact that π\pi is of type DμD_{\mu}, this can be checked for both cases 2-7, 2-29 (Proposition 5.41, Proposition 5.42).

(3) In these cases, YY is a Fano threefold of index rY=2r_{Y}=2 and Γ\Gamma is an elliptic curve such that d=D3d=D^{3} and 1d51\leq d\leq 5 for d:=degΓd:=\deg\Gamma, where DD is a Cartier divisor satisfying KY2D-K_{Y}\sim 2D (Proposition 5.42). Recall that the degree degΓ\deg\Gamma is defined as 12(KY)Γ\frac{1}{2}(-K_{Y})\cdot\Gamma. Then Lemma 5.43(i) holds by (KY)D2=2D3=2d=(KY)Γ(-K_{Y})\cdot D^{2}=2D^{3}=2d=(-K_{Y})\cdot\Gamma. Since π\pi is of type D1D_{1} for all the cases 2-1, 2-3, 2-5, 2-10, 2-14 (Proposition 5.42), it holds that KXσ(μD)+FσD+F-K_{X}\sim\sigma^{*}(\mu D)+F\sim\sigma^{*}D+F (Proposition 5.9(3)). Hence σFKYDD\sigma_{*}F\sim-K_{Y}-D\sim D. Thus Lemma 5.43(ii) holds. ∎

5.7. Case E1EE_{1}-E

Lemma 5.45.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type E2,E3,E_{2},E_{3}, or E4E_{4}. Then the following hold.

  1. (A)

    24=μ2(24r1+degB1)+24r224=\mu_{2}(\frac{24}{r_{1}}+\deg B_{1})+\frac{24}{r_{2}}.

  2. (B)

    H12H2=(r21μ2)2L23=(r1μ2)L13H_{1}^{2}\cdot H_{2}=(\frac{r_{2}-1}{\mu_{2}})^{2}L_{2}^{3}=(r_{1}-\mu_{2})L_{1}^{3}.

  3. (C)

    H1H22=r21μ2L23=(r1μ2)2L13degB1H_{1}\cdot H_{2}^{2}=\frac{r_{2}-1}{\mu_{2}}L_{2}^{3}=(r_{1}-\mu_{2})^{2}L_{1}^{3}-\deg B_{1}.

Proof.

The assertion (A) follows from Lemma 5.4(3) and Proposition 5.9(4). For each i{1,2}i\in\{1,2\}, let aia_{i} be the positive integer satisfying

KXfiKYi+aiDi.K_{X}\sim f_{i}^{*}K_{Y_{i}}+a_{i}D_{i}.

Specifically, we set a1:=1a_{1}:=1 and a2:=2a_{2}:=2 (resp. a2:=1)a_{2}:=1) if R2R_{2} is of type E2E_{2} (resp. E3E_{3} or E4E_{4}) [TanII, Proposition 3.22]. For each i{1,2}i\in\{1,2\}. we have

μ2H1+H2KX=fiKXiaiDi=riHiaiDi.\mu_{2}H_{1}+H_{2}\sim-K_{X}=-f^{*}_{i}K_{X_{i}}-a_{i}D_{i}=r_{i}H_{i}-a_{i}D_{i}.

Then the assertion (B) holds by H12H2=(r1μ2)L13H_{1}^{2}\cdot H_{2}=(r_{1}-\mu_{2})L_{1}^{3} (Lemma 5.11) and

μ22H12H2=((r21)H2a2D2)2H2=(r21)2L23.\mu_{2}^{2}H_{1}^{2}\cdot H_{2}=((r_{2}-1)H_{2}-a_{2}D_{2})^{2}\cdot H_{2}=(r_{2}-1)^{2}L_{2}^{3}.

The assertion (C) follows from H1H22=(r1μ2)2L13degBH_{1}\cdot H_{2}^{2}=(r_{1}-\mu_{2})^{2}L_{1}^{3}-\deg B (Lemma 5.11) and

μ2H1H22=((r21)H2a2D2)H22=(r21)L23.\mu_{2}H_{1}\cdot H_{2}^{2}=((r_{2}-1)H_{2}-a_{2}D_{2})\cdot H_{2}^{2}=(r_{2}-1)L_{2}^{3}.

5.7.1. Case E1E2E_{1}-E_{2}

Proposition 5.46.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type E2E_{2}. Then (KX)3=46,r1=4,degB1=2,g(B1)=0,r2=3(-K_{X})^{3}=46,r_{1}=4,\deg B_{1}=2,g(B_{1})=0,r_{2}=3 (No. 2-30).

Proof.

For each i{1,2}i\in\{1,2\}, YiY_{i} is smooth and rir_{i} is the index of the Fano threefold YiY_{i}. In particular, ri{1,2,3,4}r_{i}\in\{1,2,3,4\}. We have μ2=2\mu_{2}=2. Hence Lemma 5.45 implies the following:

  1. (A)

    24=48r1+2degB1+24r224=\frac{48}{r_{1}}+2\deg B_{1}+\frac{24}{r_{2}}.

  2. (B)

    H12H2=(r212)2L23=(r12)L13H_{1}^{2}\cdot H_{2}=(\frac{r_{2}-1}{2})^{2}L_{2}^{3}=(r_{1}-2)L_{1}^{3}.

  3. (C)

    H1H22=r212L23=(r12)2L13degB1H_{1}\cdot H_{2}^{2}=\frac{r_{2}-1}{2}L_{2}^{3}=(r_{1}-2)^{2}L_{1}^{3}-\deg B_{1}.

Since each of H1H_{1} and H2H_{2} is nef and big, we have H12H2>0H_{1}^{2}\cdot H_{2}>0 and H1H22>0H_{1}\cdot H_{2}^{2}>0. By (B) and (C), we obtain r1>2r_{1}>2 and r2>1r_{2}>1. In particular, r1{3,4}r_{1}\in\{3,4\} and r2{2,3,4}r_{2}\in\{2,3,4\}.

Assume r1=4r_{1}=4, and hence L13=1L_{1}^{3}=1. By (B) and (C), the following hold:

r2=3,L23=2,H12H2=2,H1H22=2,degB1=2.r_{2}=3,\quad L_{2}^{3}=2,\quad H_{1}^{2}\cdot H_{2}=2,\quad H_{1}\cdot H_{2}^{2}=2,\quad\deg B_{1}=2.

Then g(B1)=0g(B_{1})=0 and

(KX)3=(2H1+H2)3=8+24+12+2=46.(-K_{X})^{3}=(2H_{1}+H_{2})^{3}=8+24+12+2=46.

Assume r1=3r_{1}=3, and hence L13=2L_{1}^{3}=2. By (B) and (C), the following hold:

r2=3,L23=2,H12H2=2,H1H22=2,degB1=0.r_{2}=3,\quad L_{2}^{3}=2,\quad H_{1}^{2}\cdot H_{2}=2,\quad H_{1}\cdot H_{2}^{2}=2,\quad\deg B_{1}=0.

This is absurd. ∎

5.7.2. Cases E1E3E_{1}-E_{3} and E1E4E_{1}-E_{4}

Proposition 5.47.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type E3E_{3} or E4E_{4}. Then one of the following holds.

  1. (1)

    (KX)3=22,r1=4,degB1=6,g(B1)=4,r2=2,L23=3(-K_{X})^{3}=22,r_{1}=4,\deg B_{1}=6,g(B_{1})=4,r_{2}=2,L_{2}^{3}=3 (No. 2-15).

  2. (2)

    (KX)3=30,r1=3,degB1=4,g(B1)=1,r2=2,L23=4(-K_{X})^{3}=30,r_{1}=3,\deg B_{1}=4,g(B_{1})=1,r_{2}=2,L_{2}^{3}=4 (No. 2-23).

Proof.

Note that Y2Y_{2} is not smooth. We have μ2=1\mu_{2}=1. Hence Lemma 5.45 implies the following:

  1. (A)

    24=24r1+degB1+24r224=\frac{24}{r_{1}}+\deg B_{1}+\frac{24}{r_{2}}.

  2. (B)

    H12H2=(r21)2L23=(r11)L13H_{1}^{2}\cdot H_{2}=(r_{2}-1)^{2}L_{2}^{3}=(r_{1}-1)L_{1}^{3}.

  3. (C)

    H1H22=(r21)L23=(r11)2L13degB1H_{1}\cdot H_{2}^{2}=(r_{2}-1)L_{2}^{3}=(r_{1}-1)^{2}L_{1}^{3}-\deg B_{1}.

Since each of H1H_{1} and H2H_{2} is nef and big, we have H12H2>0H_{1}^{2}\cdot H_{2}>0 and H1H22>0H_{1}\cdot H_{2}^{2}>0. By (B) and (C), we obtain r12r_{1}\geq 2 and r22r_{2}\geq 2.

Assume r1=4r_{1}=4, and hence L13=1L_{1}^{3}=1. By (B) and (C), the following hold:

r2=2,L23=3,H12H2=3,H1H22=3,degB1=6.r_{2}=2,\quad L_{2}^{3}=3,\quad H_{1}^{2}\cdot H_{2}=3,\quad H_{1}\cdot H_{2}^{2}=3,\quad\deg B_{1}=6.

It follows from Lemma 5.35 that

(KX)3=(H1+H2)3=1+9+9+3=22,(-K_{X})^{3}=(H_{1}+H_{2})^{3}=1+9+9+3=22,
g(B1)=1132+24+1=4.g(B_{1})=11-32+24+1=4.

Assume r1=3r_{1}=3, and hence L13=2L_{1}^{3}=2. By (B), we get (r21)2L23=4(r_{2}-1)^{2}L_{2}^{3}=4. Thus (r2,L23){(2,4),(3,1)}(r_{2},L_{2}^{3})\in\{(2,4),(3,1)\}. Suppose that (r2,L23)=(3,1)(r_{2},L_{2}^{3})=(3,1). Then (B) and (C) imply H12H22H_{1}^{2}\cdot H_{2}\in 2\mathbb{Z} and H1H222H_{1}\cdot H_{2}^{2}\in 2\mathbb{Z}. We get the following contradiction: (KX)3=(H1+H2)3L13+L230mod2(-K_{X})^{3}=(H_{1}+H_{2})^{3}\equiv L_{1}^{3}+L_{2}^{3}\not\equiv 0\mod 2. Thus (r2,L23)=(2,4)(r_{2},L_{2}^{3})=(2,4). Then (B) and (C) imply the following:

H12H2=4,H1H22=4,degB1=4.H_{1}^{2}\cdot H_{2}=4,\quad H_{1}\cdot H_{2}^{2}=4,\quad\deg B_{1}=4.

Then the following hold (Lemma 5.35):

(KX)3=(H1+H2)3=2+12+12+4=30,(-K_{X})^{3}=(H_{1}+H_{2})^{3}=2+12+12+4=30,
g(B1)=1527+12+1=1.g(B_{1})=15-27+12+1=1.

Assume r1=2r_{1}=2, and hence 1L1351\leq L_{1}^{3}\leq 5. Then (A)-(C) can be rewritten as follows:

  1. (A)

    12=degB1+24r212=\deg B_{1}+\frac{24}{r_{2}}.

  2. (B)

    H12H2=(r21)2L23=L13H_{1}^{2}\cdot H_{2}=(r_{2}-1)^{2}L_{2}^{3}=L_{1}^{3}.

  3. (C)

    H1H22=(r21)L23=L13degB1H_{1}\cdot H_{2}^{2}=(r_{2}-1)L_{2}^{3}=L_{1}^{3}-\deg B_{1}.

By 1L1351\leq L_{1}^{3}\leq 5 and (B), we obtain r2{2,3}r_{2}\in\{2,3\}. Since the case r2=2r_{2}=2 contradicts (A), we get r2=3r_{2}=3. Then (B) and 1L1351\leq L_{1}^{3}\leq 5 imply L23=1L_{2}^{3}=1 and L13=4L_{1}^{3}=4. By (B) and (C), we get H12H22H_{1}^{2}\cdot H_{2}\in 2\mathbb{Z} and H1H222H_{1}\cdot H_{2}^{2}\in 2\mathbb{Z}, respectively. We then obtain the following contradiction: (KX)3=(H1+H2)3L13+L230mod2(-K_{X})^{3}=(H_{1}+H_{2})^{3}\equiv L_{1}^{3}+L_{2}^{3}\not\equiv 0\mod 2. ∎

5.7.3. Case E1E5E_{1}-E_{5}

Proposition 5.48.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type E5E_{5}. Then (KX)3=40,r1=4,degB1=3,g(B1)=1,r2=3,L23=12(-K_{X})^{3}=40,r_{1}=4,\deg B_{1}=3,g(B_{1})=1,r_{2}=3,L_{2}^{3}=12 (No. 2-28).

Proof.

By Lemma 5.4, we obtain μ2=1\mu_{2}=1 and H2c2(X)=45/r2H_{2}\cdot c_{2}(X)=45/r_{2}. It holds that

H1+H2KX=fKX1D1=r1H1D1,H_{1}+H_{2}\sim-K_{X}=-f^{*}K_{X_{1}}-D_{1}=r_{1}H_{1}-D_{1},
H1+H2KX=fKX212D2=r22H212D2.H_{1}+H_{2}\sim-K_{X}=-f^{*}K_{X_{2}}-\frac{1}{2}D_{2}=\frac{r_{2}}{2}H_{2}-\frac{1}{2}D_{2}.

We then get the following:

H12H2=H12((r11)H1D1)=(r11)L13.H_{1}^{2}\cdot H_{2}=H_{1}^{2}\cdot((r_{1}-1)H_{1}-D_{1})=(r_{1}-1)L_{1}^{3}.
H12H2=((r221)H212D2)2H2=(r221)2L23.H_{1}^{2}\cdot H_{2}=((\frac{r_{2}}{2}-1)H_{2}-\frac{1}{2}D_{2})^{2}\cdot H_{2}=(\frac{r_{2}}{2}-1)^{2}L_{2}^{3}.
H1H22=((r221)H212D2)H22=(r221)L23.H_{1}\cdot H_{2}^{2}=((\frac{r_{2}}{2}-1)H_{2}-\frac{1}{2}D_{2})\cdot H_{2}^{2}=(\frac{r_{2}}{2}-1)L_{2}^{3}.
H1H22=H1((r11)H1D1)2=(r11)2L13+H1D12=(r11)2L13degB.H_{1}\cdot H_{2}^{2}=H_{1}\cdot((r_{1}-1)H_{1}-D_{1})^{2}=(r_{1}-1)^{2}L_{1}^{3}+H_{1}\cdot D_{1}^{2}=(r_{1}-1)^{2}L_{1}^{3}-\deg B.

Therefore, the following hold, where (A) is guaranteed by Lemma 5.4(3) and Proposition 5.9(4):

  1. (A)

    24=24r1+degB1+45r224=\frac{24}{r_{1}}+\deg B_{1}+\frac{45}{r_{2}}.

  2. (B)

    H12H2=(r221)2L23=(r11)L13H_{1}^{2}\cdot H_{2}=(\frac{r_{2}}{2}-1)^{2}L_{2}^{3}=(r_{1}-1)L_{1}^{3}. In particular, L23=4(r11)(r22)2L13L_{2}^{3}=\frac{4(r_{1}-1)}{(r_{2}-2)^{2}}L_{1}^{3}.

  3. (C)

    H1H22=(r221)L23=(r11)2L13degB1H_{1}\cdot H_{2}^{2}=(\frac{r_{2}}{2}-1)L_{2}^{3}=(r_{1}-1)^{2}L_{1}^{3}-\deg B_{1}.

By (A), we have r12r_{1}\geq 2 and r23r_{2}\geq 3.

Assume r1=4r_{1}=4, and hence L13=1L_{1}^{3}=1. By (B), we get L23=4(r11)(r22)2L13=12(r22)2\mathbb{Z}\ni L_{2}^{3}=\frac{4(r_{1}-1)}{(r_{2}-2)^{2}}L_{1}^{3}=\frac{12}{(r_{2}-2)^{2}}. Hence r2{3,4}r_{2}\in\{3,4\}. Since r24r_{2}\neq 4 by (A), we obtain r2=3r_{2}=3. Then (A)-(C) imply the following:

L23=12,degB1=3,H12H2=3,H1H22=6.L_{2}^{3}=12,\quad\deg B_{1}=3,\quad H_{1}^{2}\cdot H_{2}=3,\quad H_{1}\cdot H_{2}^{2}=6.

Then the following hold (Lemma 5.35):

(KX)3=(H1+H2)3=1+9+18+12=40(-K_{X})^{3}=(H_{1}+H_{2})^{3}=1+9+18+12=40
g(B1)=2032+12+1=1.g(B_{1})=20-32+12+1=1.

Assume r1=3r_{1}=3, and hence L13=2L_{1}^{3}=2. By (A), we have r2{3,5,9,15,45}r_{2}\in\{3,5,9,15,45\}. This, together with >0L23=4(r11)(r22)2L13=16(r22)2\mathbb{Z}_{>0}\ni L_{2}^{3}=\frac{4(r_{1}-1)}{(r_{2}-2)^{2}}L_{1}^{3}=\frac{16}{(r_{2}-2)^{2}}, implies r2=3r_{2}=3 and L23=16L_{2}^{3}=16. By (C), we get the following contradiction:

degB1=(r11)2L13(r221)L23=88=0.\deg B_{1}=(r_{1}-1)^{2}L_{1}^{3}-(\frac{r_{2}}{2}-1)L_{2}^{3}=8-8=0.

Assume r1=2r_{1}=2, and hence 1L1351\leq L_{1}^{3}\leq 5. Then (A)-(C) can be rewritten as follows:

  1. (A)

    12=degB1+45r212=\deg B_{1}+\frac{45}{r_{2}}.

  2. (B)

    H12H2=(r221)2L23=L13H_{1}^{2}\cdot H_{2}=(\frac{r_{2}}{2}-1)^{2}L_{2}^{3}=L_{1}^{3}. In particular, L23=4(r22)2L13L_{2}^{3}=\frac{4}{(r_{2}-2)^{2}}L_{1}^{3}.

  3. (C)

    H1H22=(r221)L23=L13degB1H_{1}\cdot H_{2}^{2}=(\frac{r_{2}}{2}-1)L_{2}^{3}=L_{1}^{3}-\deg B_{1}.

By (A), we have r2{5,9,15,45}r_{2}\in\{5,9,15,45\}. However, this contradicts >0L23=4(r22)2L13\mathbb{Z}_{>0}\ni L_{2}^{3}=\frac{4}{(r_{2}-2)^{2}}L_{1}^{3} and 1L1351\leq L_{1}^{3}\leq 5. ∎

5.7.4. Case E1E1E_{1}-E_{1}

Proposition 5.49.

We use Notation 5.1. Assume that R1R_{1} is of type E1E_{1} and R2R_{2} is of type E1E_{1}. Moreover, suppose r1r2r_{1}\geq r_{2}. Then one of the following holds possibly after permuting R1R_{1} and R2R_{2}.

  1. (1)

    (KX)3=20,Y1Y23,g(B1)=g(B2)=3,degB1=degB2=6(-K_{X})^{3}=20,Y_{1}\simeq Y_{2}\simeq\mathbb{P}^{3},g(B_{1})=g(B_{2})=3,\deg B_{1}=\deg B_{2}=6 (No. 2-12).

  2. (2)

    (KX)3=24,Y13,g(B1)=1,degB1=5,Y2Q,g(B2)=1,degB2=5(-K_{X})^{3}=24,Y_{1}\simeq\mathbb{P}^{3},g(B_{1})=1,\deg B_{1}=5,Y_{2}\simeq Q,g(B_{2})=1,\deg B_{2}=5 (No. 2-17).

  3. (3)

    (KX)3=26,Y13,g(B1)=2,degB1=5,r2=2,L23=4,g(B2)=0,degB2=1(-K_{X})^{3}=26,Y_{1}\simeq\mathbb{P}^{3},g(B_{1})=2,\deg B_{1}=5,r_{2}=2,L_{2}^{3}=4,g(B_{2})=0,\deg B_{2}=1 (No. 2-19).

  4. (4)

    (KX)3=30,Y13,g(B1)=0,degB1=4,r2=2,L23=5,g(B2)=0,degB2=2(-K_{X})^{3}=30,Y_{1}\simeq\mathbb{P}^{3},g(B_{1})=0,\deg B_{1}=4,r_{2}=2,L_{2}^{3}=5,g(B_{2})=0,\deg B_{2}=2 (No. 2-22).

  5. (5)

    (KX)3=28,Y1Y2Q,g(B1)=g(B2)=0,degB1=degB2=4(-K_{X})^{3}=28,Y_{1}\simeq Y_{2}\simeq Q,g(B_{1})=g(B_{2})=0,\deg B_{1}=\deg B_{2}=4 (No. 2-21).

  6. (6)

    (KX)3=34,Y1Q,g(B1)=0,degB1=3,r2=2,L23=5,g(B2)=0,degB2=1(-K_{X})^{3}=34,Y_{1}\simeq Q,g(B_{1})=0,\deg B_{1}=3,r_{2}=2,L_{2}^{3}=5,g(B_{2})=0,\deg B_{2}=1 (No. 2-26).

Proof.

We may apply Lemma 5.11 for both the extremal rays R1R_{1} and R2R_{2}, although we need to permute the indices when we apply it for R2R_{2}. Then the following hold (Lemma 5.4(3), Proposition 5.9(4)):

  1. (A)

    24=24r1+degB1+24r2+degB224=\frac{24}{r_{1}}+\deg B_{1}+\frac{24}{r_{2}}+\deg B_{2}.

  2. (B)

    H12H2=(r11)L13=(r21)2L23degB2H_{1}^{2}\cdot H_{2}=(r_{1}-1)L_{1}^{3}=(r_{2}-1)^{2}L_{2}^{3}-\deg B_{2}.

  3. (C)

    H1H22=(r21)L23=(r11)2L13degB1H_{1}\cdot H_{2}^{2}=(r_{2}-1)L_{2}^{3}=(r_{1}-1)^{2}L_{1}^{3}-\deg B_{1}.

Since each of H1H_{1} and H2H_{2} is nef and big, we have H12H2>0H_{1}^{2}\cdot H_{2}>0 and H1H22>0H_{1}\cdot H_{2}^{2}>0. By (B) and (C), we get r12r_{1}\geq 2 and r22r_{2}\geq 2, respectively. By r1r2r_{1}\geq r_{2} and (A), we obtain r13r_{1}\geq 3. In particular,

(r1,r2){(4,4),(4,3),(4,2),(3,3),(3,2)}.(r_{1},r_{2})\in\{(4,4),(4,3),(4,2),(3,3),(3,2)\}.

In what follows, we shall use

  • (KX)3=(H1+H2)3=L13+3H12H2+3H1H22+L23(-K_{X})^{3}=(H_{1}+H_{2})^{3}=L_{1}^{3}+3H_{1}^{2}\cdot H_{2}+3H_{1}\cdot H_{2}^{2}+L_{2}^{3}, and

  • g(Bi)=(KX)32(KYi)32+ridegBi+1g(B_{i})=\frac{(-K_{X})^{3}}{2}-\frac{(-K_{Y_{i}})^{3}}{2}+r_{i}\deg B_{i}+1 (Lemma 5.35).

(1) Assume (r1,r2)=(4,4)(r_{1},r_{2})=(4,4). Then (L13,L23)=(1,1)(L_{1}^{3},L_{2}^{3})=(1,1) and the following hold:

  1. (A)

    12=degB1+degB212=\deg B_{1}+\deg B_{2}.

  2. (B)

    H12H2=3=9degB1H_{1}^{2}\cdot H_{2}=3=9-\deg B_{1}.

  3. (C)

    H1H22=3=9degB2H_{1}\cdot H_{2}^{2}=3=9-\deg B_{2}.

Hence degB1=degB2=6\deg B_{1}=\deg B_{2}=6,

(KX)3=1+9+9+1=20,g(B1)=g(B2)=1032+24+1=3.(-K_{X})^{3}=1+9+9+1=20,\qquad g(B_{1})=g(B_{2})=10-32+24+1=3.

(2) Assume (r1,r2)=(4,3)(r_{1},r_{2})=(4,3). Then (L13,L23)=(1,2)(L_{1}^{3},L_{2}^{3})=(1,2) and the following hold:

  1. (A)

    10=degB1+degB210=\deg B_{1}+\deg B_{2}.

  2. (B)

    H12H22=3=8degB2H_{1}^{2}\cdot H_{2}^{2}=3=8-\deg B_{2}.

  3. (C)

    H1H22=4=9degB1H_{1}\cdot H_{2}^{2}=4=9-\deg B_{1}.

Hence (degB1,degB2,H12H2,H1H22)=(5,5,3,4)(\deg B_{1},\deg B_{2},H_{1}^{2}\cdot H_{2},H_{1}\cdot H_{2}^{2})=(5,5,3,4),

(KX)3=1+9+12+2=24,(-K_{X})^{3}=1+9+12+2=24,
g(B1)=1232+20+1=1,g(B2)=1227+15+1=1.g(B_{1})=12-32+20+1=1,\qquad g(B_{2})=12-27+15+1=1.

(3), (4) Assume (r1,r2)=(4,2)(r_{1},r_{2})=(4,2). Then L13=1L_{1}^{3}=1, 1L2351\leq L_{2}^{3}\leq 5, and the following hold:

  1. (A)

    6=degB1+degB26=\deg B_{1}+\deg B_{2}.

  2. (B)

    H12H22=3=L23degB2H_{1}^{2}\cdot H_{2}^{2}=3=L_{2}^{3}-\deg B_{2}.

  3. (C)

    H1H22=L23=9degB1H_{1}\cdot H_{2}^{2}=L_{2}^{3}=9-\deg B_{1}.

By 1L2351\leq L_{2}^{3}\leq 5, L23degB2=3L_{2}^{3}-\deg B_{2}=3, and degB2>0\deg B_{2}>0, we have two solutions: (degB2,L23){(1,4),(2,5)}(\deg B_{2},L_{2}^{3})\in\{(1,4),(2,5)\}.

Assume (degB2,L23)=(1,4)(\deg B_{2},L_{2}^{3})=(1,4). Then we get (degB1,H12H2,H1H22)=(5,3,4)(\deg B_{1},H_{1}^{2}\cdot H_{2},H_{1}\cdot H_{2}^{2})=(5,3,4). Then

(KX)3=1+9+12+4=26,(-K_{X})^{3}=1+9+12+4=26,
g(B1)=1332+20+1=2,g(B2)=0.g(B_{1})=13-32+20+1=2,\qquad g(B_{2})=0.

Assume (degB2,L23)=(2,5)(\deg B_{2},L_{2}^{3})=(2,5). Then we get (degB1,H12H2,H1H22)=(4,3,5)(\deg B_{1},H_{1}^{2}\cdot H_{2},H_{1}\cdot H_{2}^{2})=(4,3,5). Then

(KX)3=1+9+15+5=30,(-K_{X})^{3}=1+9+15+5=30,
g(B1)=1532+16+1=0,g(B2)=0.g(B_{1})=15-32+16+1=0,\qquad g(B_{2})=0.

(5) Assume (r1,r2)=(3,3)(r_{1},r_{2})=(3,3). Then L13=L23=2L_{1}^{3}=L_{2}^{3}=2, and the following hold:

  1. (A)

    8=degB1+degB28=\deg B_{1}+\deg B_{2}.

  2. (B)

    H12H22=4=8degB2H_{1}^{2}\cdot H_{2}^{2}=4=8-\deg B_{2}.

  3. (C)

    H1H22=4=8degB1H_{1}\cdot H_{2}^{2}=4=8-\deg B_{1}.

Hence degB1=degB2=H12H2=H1H22=4\deg B_{1}=\deg B_{2}=H_{1}^{2}\cdot H_{2}=H_{1}\cdot H_{2}^{2}=4,

(KX)3=2+12+12+2=28,g(B1)=g(B2)=1427+12+1=0.(-K_{X})^{3}=2+12+12+2=28,\qquad g(B_{1})=g(B_{2})=14-27+12+1=0.

(6) Assume (r1,r2)=(3,2)(r_{1},r_{2})=(3,2). Then L13=2L_{1}^{3}=2, 1L2351\leq L_{2}^{3}\leq 5, and the following hold:

  1. (A)

    4=degB1+degB24=\deg B_{1}+\deg B_{2}.

  2. (B)

    H12H22=4=L23degB2H_{1}^{2}\cdot H_{2}^{2}=4=L_{2}^{3}-\deg B_{2}.

  3. (C)

    H1H22=L23=8degB1H_{1}\cdot H_{2}^{2}=L_{2}^{3}=8-\deg B_{1}.

By 1L2351\leq L_{2}^{3}\leq 5, L23degB2=4L_{2}^{3}-\deg B_{2}=4, and degB2>0\deg B_{2}>0, we obtain degB2=1\deg B_{2}=1 and L23=5L_{2}^{3}=5. Hence we get (degB1,H12H2,H1H22)=(3,4,5)(\deg B_{1},H_{1}^{2}\cdot H_{2},H_{1}\cdot H_{2}^{2})=(3,4,5). Then

(KX)3=2+12+15+5=34,(-K_{X})^{3}=2+12+15+5=34,
g(B1)=1727+9+1=0,g(B2)=1720+2+1=0.g(B_{1})=17-27+9+1=0,\qquad g(B_{2})=17-20+2+1=0.

6. Primitive Fano threefolds with ρ(X)=3\rho(X)=3

In this section, we classify primitive Fano threefolds with ρ(X)=3\rho(X)=3. The goal of this section is to prove the following theorem.

Theorem 6.1 (Theorem 6.7, Theorem 6.17).

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Then one and only one of the following holds.

  1. (1)

    (KX)3=12(-K_{X})^{3}=12 and there exists a split double cover f:X1×1×1f:X\to\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} such that f𝒪X/𝒪1×1×1𝒪1×1×1(1,1,1)1f_{*}\mathcal{O}_{X}/\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1,1)^{-1}.

  2. (2)

    (KX)3=14(-K_{X})^{3}=14 and XX is isomorphic to a prime divisor XX^{\prime} on P:=(𝒪1×1𝒪1×1(1,1)2)P:=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2}) such that 𝒪P(X)𝒪P(2)π𝒪1×1(2,3)\mathcal{O}_{P}(X^{\prime})\simeq\mathcal{O}_{P}(2)\otimes\pi^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,3), where π:P=(𝒪1×1𝒪1×1(1,1)2)1×1\pi:P=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2})\to\mathbb{P}^{1}\times\mathbb{P}^{1} denotes the natural projection.

  3. (3)

    (KX)3=48(-K_{X})^{3}=48 and X1×1×1X\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}.

  4. (4)

    (KX)3=52(-K_{X})^{3}=52 and X(𝒪1×1𝒪1×1(1,1))X\simeq\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)).

By Theorem 4.17, there are the following two cases, which we shall treat separately:

  • All the extremal rays are of type CC (Subsection 6.1).

  • There exist extremal rays R1R_{1} and R2R_{2} such that R1R_{1} is of type CC and R2R_{2} is of type E1E_{1} (Subsection 6.2).

6.1. Case CCCC-C-C

Lemma 6.2.

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Assume that all the extremal rays are of type CC. Let FF be a two-dimensional extremral face generated by extremal rays R1R_{1} and R2R_{2}. Let f1:XY1f_{1}:X\to Y_{1}, f2:XY2f_{2}:X\to Y_{2}, and π:XB\pi:X\to B be the contractions of R1R_{1}, R2R_{2}, and FF, respectively. Then we have the following commutative diagram

X{X}Y1{Y_{1}}Y2{Y_{2}}B{B}f1\scriptstyle{f_{1}}f2\scriptstyle{f_{2}}π\scriptstyle{\pi}g1\scriptstyle{g_{1}}g2\scriptstyle{g_{2}}

such that the following hold.

  1. (1)

    Y1=B×C1Y_{1}=B\times C_{1}.

  2. (2)

    Y2=B×C2Y_{2}=B\times C_{2}.

  3. (3)

    B=C1=C2=1B=C_{1}=C_{2}=\mathbb{P}^{1}.

  4. (4)

    Y1BY_{1}\to B and Y2BY_{2}\to B are the first projections.

Proof.

Each extremal ray RR is an intersection of two two-dimensional extremal faces FF and FF^{\prime}. This implies that there are exactly two non-trivial contractions XB1X\to B_{1} and XB2X\to B_{2} that factor through the contraction XYX\to Y of RR. By Y=1×1Y=\mathbb{P}^{1}\times\mathbb{P}^{1}, these must coinside the ones induced by the projections. ∎

Remark 6.3.

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Assume that all the extremal rays are of type CC. Let FF be an extremal face of NE(X){\operatorname{NE}}(X) and let h:XZh:X\to Z be the contraction of FF. Then the following hold.

  1. (1)

    The following are equivalent.

    • dimF=1\dim F=1, i.e., FF is an extremal ray.

    • dimZ=2\dim Z=2.

    • Z1×1Z\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}.

  2. (2)

    The following are equivalent.

    • dimF=2\dim F=2.

    • dimZ=1\dim Z=1.

    • Z1Z\simeq\mathbb{P}^{1}.

Proposition 6.4.

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Assume that all the extremal rays are of type CC. Then the number of the extremal rays of XX is three.

Proof.

Suppose that XX has at least four extrmemal rays. Then we can find two two-dimensional extremal faces FF and FF^{\prime} such that FF={0}F\cap F^{\prime}=\{0\}. Let π:XB\pi:X\to B and π:XB\pi^{\prime}:X\to B^{\prime} be the contractions of FF and FF^{\prime}, respectively. By Remark 6.3, we get dimB=dimB=1\dim B=\dim B^{\prime}=1. Then there exists a curve CC on XX such that (π×π)(C)(\pi\times\pi^{\prime})(C) is a point for π×π:XB×B\pi\times\pi^{\prime}:X\to B\times B^{\prime}. This implies that π(C)\pi(C) and π(C)\pi^{\prime}(C) are points, i.e., [C]FF={0}[C]\in F\cap F^{\prime}=\{0\}, which is absurd. ∎

Notation 6.5.

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Assume that all the extremal rays are of type CC. Then there exist exactly three two-dimensional extremal faces F1,F2,F3F_{1},F_{2},F_{3} (Proposition 6.4). We have also exactly three extremal rays R12,R23,R13R_{12},R_{23},R_{13}, which are given as follows:

R12:=F1F2,R23:=F2F3,R13:=F1F3.R_{12}:=F_{1}\cap F_{2},\qquad R_{23}:=F_{2}\cap F_{3},\qquad R_{13}:=F_{1}\cap F_{3}.

Corresponding to these extremal faces, we have the following contraction morphisms:

  • πiX:XBi:=1\pi^{X}_{i}:X\to B_{i}:=\mathbb{P}^{1} for all 1i31\leq i\leq 3.

  • fij:XYij:=Bi×Bj=1×1f_{ij}:X\to Y_{ij}:=B_{i}\times B_{j}=\mathbb{P}^{1}\times\mathbb{P}^{1} for all 1i<j31\leq i<j\leq 3.

Note that fijf_{ij} and πi\pi_{i} are compatible, i.e.,

πiX:XfijYij=Bi×Bjpr1Bi,πjX:XfijYij=Bi×Bjpr2Bj.\pi^{X}_{i}:X\xrightarrow{f_{ij}}Y_{ij}=B_{i}\times B_{j}\xrightarrow{{\rm pr}_{1}}B_{i},\qquad\pi^{X}_{j}:X\xrightarrow{f_{ij}}Y_{ij}=B_{i}\times B_{j}\xrightarrow{{\rm pr}_{2}}B_{j}.

Set

f:=π1X×π2X×π3X:XZ:=B1×B2×B3=1×1×1.f:=\pi_{1}^{X}\times\pi_{2}^{X}\times\pi_{3}^{X}:X\to Z:=B_{1}\times B_{2}\times B_{3}=\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}.

Let πiZ:ZBi=1\pi_{i}^{Z}:Z\to B_{i}=\mathbb{P}^{1} be the ii-th projection. For each i{1,2,3}i\in\{1,2,3\}, we set

  • Hi:=(πiX)𝒪1(1)H_{i}:=(\pi^{X}_{i})^{*}\mathcal{O}_{\mathbb{P}^{1}}(1), and

  • HiZ:=(πiZ)𝒪1(1)H^{Z}_{i}:=(\pi^{Z}_{i})^{*}\mathcal{O}_{\mathbb{P}^{1}}(1).

For 1i<j31\leq i<j\leq 3, let ij\ell_{ij} be an extremal rational curve of the extremral ray RijR_{ij}. Note that we have Hiij=Hjij=0H_{i}\cdot\ell_{ij}=H_{j}\cdot\ell_{ij}=0.

Lemma 6.6.

We use Notation 6.5. Then f:XZ=1×1×1f:X\to Z=\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} is a finite surjective morphism.

Proof.

It suffices to show that ff is a finite morphism. Suppose that ff is not a finite morphism. Then there exists a curve CC on XX such that f(C)f(C) is a point. By f=π1X×π2X×π3Xf=\pi^{X}_{1}\times\pi^{X}_{2}\times\pi^{X}_{3}, all of π1X(C),π2X(C)\pi^{X}_{1}(C),\pi^{X}_{2}(C), and π3X(C)\pi^{X}_{3}(C) are points. Then [C]F1F2F3={0}[C]\in F_{1}\cap F_{2}\cap F_{3}=\{0\}, which is a contradiction. ∎

Theorem 6.7.

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Assume that all the extremal rays of NE(X){\operatorname{NE}}(X) are of type CC. Let R1R_{1} and R2R_{2} be distinct extremal rays of NE(X){\operatorname{NE}}(X). Then one and only one of the following holds.

  1. (1)

    There exists a split double cover f:X1×1×1f:X\to\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} such that f𝒪X/𝒪1×1×1𝒪1×1×1(1,1,1)1f_{*}\mathcal{O}_{X}/\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1,1)^{-1}. Furthermore, 𝒪X(KX)f𝒪1×1×1(1,1,1)\mathcal{O}_{X}(-K_{X})\simeq f^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1,1), (KX)3=12(-K_{X})^{3}=12, there are exactly three extremal rays of NE(X){\operatorname{NE}}(X), all the extremal rays are of type C1C_{1}, and Δφ\Delta_{\varphi} is of bidegree (4,4)(4,4) for the contraction φ:X1×1\varphi:X\to\mathbb{P}^{1}\times\mathbb{P}^{1} of an arbitrary extremal ray.

  2. (2)

    X1×1×1X\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}. Furthermore, (KX)3=48(-K_{X})^{3}=48, there are exactly three extremal rays R1,R2,R3R_{1},R_{2},R_{3} of NE(X){\operatorname{NE}}(X), and all the extremal rays are of type C2C_{2}.

Proof.

We use Notation 6.5. Set d:=degfd:=\deg f. For each i{1,2,3}i\in\{1,2,3\}, we have morphisms:

πiX:X𝑓Z=1×1×1πiZ1.\pi^{X}_{i}:X\xrightarrow{f}Z=\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}\xrightarrow{\pi_{i}^{Z}}\mathbb{P}^{1}.

By

H1ZH2ZH3Z=(π1Z)𝒪1(1)(π2Z)𝒪1(1)(π3Z)𝒪1(1)=1,H_{1}^{Z}\cdot H_{2}^{Z}\cdot H_{3}^{Z}=(\pi_{1}^{Z})^{*}\mathcal{O}_{\mathbb{P}^{1}}(1)\cdot(\pi_{2}^{Z})^{*}\mathcal{O}_{\mathbb{P}^{1}}(1)\cdot(\pi_{3}^{Z})^{*}\mathcal{O}_{\mathbb{P}^{1}}(1)=1,

we have

d=degf=(fH1Z)(fH2Z)(fH3Z)=H1H2H3.d=\deg f=(f^{*}H_{1}^{Z})\cdot(f^{*}H_{2}^{Z})\cdot(f^{*}H_{3}^{Z})=H_{1}\cdot H_{2}\cdot H_{3}.

Since H1H2H_{1}\cdot H_{2} is a fibre of f12:XY12=B1×B2=1×1f_{12}:X\to Y_{12}=B_{1}\times B_{2}=\mathbb{P}^{1}\times\mathbb{P}^{1}, we obtain H1H22μ1212H_{1}\cdot H_{2}\equiv\frac{2}{\mu_{12}}\ell_{12}, where μ12\mu_{12} denotes the length of R12R_{12}. Therefore, we get

d=H1H2H3=2μ12H312,d=H_{1}\cdot H_{2}\cdot H_{3}=\frac{2}{\mu_{12}}H_{3}\cdot\ell_{12},

which implies

KX12=μ12=2dH312.-K_{X}\cdot\ell_{12}=\mu_{12}=\frac{2}{d}H_{3}\cdot\ell_{12}.

By symmetry, we have that dKX12=2H312,dKX23=2H123,dKX13=2H213.-dK_{X}\cdot\ell_{12}=2H_{3}\cdot\ell_{12},-dK_{X}\cdot\ell_{23}=2H_{1}\cdot\ell_{23},-dK_{X}\cdot\ell_{13}=2H_{2}\cdot\ell_{13}. By the exact sequence

0PicY12PicX12,0\to{\operatorname{Pic}}\,Y_{12}\to{\operatorname{Pic}}\,X\xrightarrow{\cdot\ell_{12}}\mathbb{Z},

H1,H2,H3H_{1},H_{2},H_{3} form a \mathbb{Q}-linear basis of (PicX)({\operatorname{Pic}}\,X)\otimes_{\mathbb{Z}}\mathbb{Q}. Hence we obtain

dKX2H1+2H2+2H3.-dK_{X}\sim 2H_{1}+2H_{2}+2H_{3}.

For the positive integer g>0g\in\mathbb{Z}_{>0} satisying (KX)3=2g2(-K_{X})^{3}=2g-2, we get

d3(2g2)=d3(KX)3=8(H1+H2+H3)3=48degf=48d.d^{3}(2g-2)=d^{3}(-K_{X})^{3}=8(H_{1}+H_{2}+H_{3})^{3}=48\deg f=48d.

Hence d2(g1)=24=233d^{2}(g-1)=24=2^{3}\cdot 3, which implies d{1,2}d\in\{1,2\}.

Assume d=1d=1. Then f:XZ=1×1×1f:X\to Z=\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} is a finite birational morphism (Lemma 6.6). Since ZZ is normal, ff is an isomorphism. We have (KX)3=𝒪1×1×1(2,2,2)3=48(-K_{X})^{3}=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,2,2)^{3}=48. Furthermore, each projection X1×1X\to\mathbb{P}^{1}\times\mathbb{P}^{1} is clearly of type C2C_{2}. Hence (2) holds.

Assume d=2d=2. Then KXH1+H2+H3-K_{X}\sim H_{1}+H_{2}+H_{3} and f:XZ=1×1×1f:X\to Z=\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} is a double cover (Lemma 6.6). This double cover is split (Lemma 2.4), because 1×1×1\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} is FF-split and H1(1×1×1,𝒪1×1×1(E))=0H^{1}(\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(E))=0 for any effective divisor EE on 1×1×1\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}. We have (KX)3=degf𝒪1×1×1(1,1,1)3=12(-K_{X})^{3}=\deg f\cdot\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1,1)^{3}=12.

Let us show that f12:XB1×B2f_{12}:X\to B_{1}\times B_{2} is of type C1C_{1}. We can write

𝒪1×1(Δf12)=𝒪1×1(a,b)\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\Delta_{f_{12}})=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(a,b)

for some a,ba,b\in\mathbb{Z}. We obtain

4=2d=2H1H2H3=(KX)2H1=()4K1×1𝒪1×1(1,0)Δf12𝒪1×1(1,0)4=2d=2H_{1}\cdot H_{2}\cdot H_{3}=(-K_{X})^{2}\cdot H_{1}\overset{(\star)}{=}-4K_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\cdot\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)-\Delta_{f_{12}}\cdot\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)
=𝒪1×1(8a,8b)𝒪1×1(1,0)=8b.=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(8-a,8-b)\cdot\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)=8-b.

where (\star) follows from Proposition 3.16. Therefore, b=4b=4. Similarly, a=4a=4. Hence f12f_{12} is of type C1C_{1}. By symmetry, fij:XBi×Bjf_{ij}:X\to B_{i}\times B_{j} is of type C1C_{1} for all 1i<j31\leq i<j\leq 3. Thus (1) holds. ∎

Remark 6.8.

By the above proof, the split double cover f:X1×1×1f:X\to\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} in Theorem 6.7(1) can be chosen to be ff as in Notation 6.5.

6.2. Case CEC-E

What is remaining is the following case (cf. Theorem 4.17):

Notation 6.9.

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Let R1R_{1} and R2R_{2} be two distinct extremal rays. Assume that

  • R1R_{1} is of type CC, and

  • R2R_{2} is of type E1E_{1}.

For each i{1,2}i\in\{1,2\}, let

fi:XYif_{i}:X\to Y_{i}

be the contraction of RiR_{i}, where Y1=1×1Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1}. Set D:=Ex(f2)D:={\operatorname{Ex}}(f_{2}), which is a prime divisor on XX such that D1×1D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1} (Theorem 4.8). Let i\ell_{i} be an extremal rational curve with Ri=0[i]R_{i}=\mathbb{R}_{\geq 0}[\ell_{i}]. Set μi:=KXi\mu_{i}:=-K_{X}\cdot\ell_{i}, which is the length of RiR_{i}.

Lemma 6.10.

We use Notation 6.9. For the induced morphism

f1|D:D1×11×1=Y1,f_{1}|_{D}\colon D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}\to\mathbb{P}^{1}\times\mathbb{P}^{1}=Y_{1},

the following hold.

  1. (1)

    If R1R_{1} is of type C1C_{1}, then f1|Df_{1}|_{D} is a double cover.

  2. (2)

    If R2R_{2} is of type C2C_{2}, then f1|Df_{1}|_{D} is an isomorphism.

Proof.

By Lemma 4.15(2), f1|D:DY1f_{1}|_{D}:D\to Y_{1} is a finite surjective morphism. Fix a fibre ζ\zeta of f1:XY1f_{1}:X\to Y_{1} over a closed point of Y1Y_{1}. By ζ2μ11\zeta\equiv\frac{2}{\mu_{1}}\ell_{1} (Corollary 4.5), it holds that

deg(f1|D)=Dζ=2μ1D1.\deg(f_{1}|_{D})=D\cdot\zeta=\frac{2}{\mu_{1}}D\cdot\ell_{1}.

Recall that we have the following exact sequence (Theorem 4.11):

0Pic(1×1)f1PicX10.0\longrightarrow{\operatorname{Pic}}\,(\mathbb{P}^{1}\times\mathbb{P}^{1})\stackrel{{\scriptstyle f_{1}^{*}}}{{\to}}{\operatorname{Pic}}\,X\stackrel{{\scriptstyle\cdot\ell_{1}}}{{\to}}\mathbb{Z}\longrightarrow 0.

Fix a Cartier divisor EE on XX with E1=1E\cdot\ell_{1}=1, whose existence is guaranteed by the surjectivity of the last map EE1E\mapsto E\cdot\ell_{1}. Then there exist aa\in\mathbb{Z} and a Cartier divisor LL on Y1=1×1Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1} such that

Df1L+aE.D\sim f_{1}^{*}L+aE.

By a=D1a=D\cdot\ell_{1}, we have a>0a\in\mathbb{Z}_{>0}. Since LL is a divisor on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, we have L3=0L^{3}=0 and 𝒪1×1(L)𝒪1×1(b,c)\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(L)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(b,c) for some b,cb,c\in\mathbb{Z}. We have L22bcQL^{2}\equiv 2bcQ for a closed point QQ on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, which implies f1L22bcζ4bcμ11f_{1}^{*}L^{2}\equiv 2bc\zeta\equiv\frac{4bc}{\mu_{1}}\ell_{1} and f1L2E=4bcμ1f_{1}^{*}L^{2}\cdot E=\frac{4bc}{\mu_{1}}. Therefore, we obtain

D3\displaystyle D^{3} =(f1L+aE)3\displaystyle=(f^{*}_{1}L+aE)^{3}
=f1L3+3f1L2(aE)+3f1L(a2E2)+a3E3\displaystyle=f_{1}^{*}L^{3}+3f_{1}^{*}L^{2}\cdot(aE)+3f^{*}_{1}L\cdot(a^{2}E^{2})+a^{3}E^{3}
=3a4bcμ1+3a2(f1LE2)+a3E3.\displaystyle=3a\cdot\frac{4bc}{\mu_{1}}+3a^{2}(f^{*}_{1}L\cdot E^{2})+a^{3}E^{3}.

On the other hand, for C:=f2(D)C:=f_{2}(D), it holds that

D3=degC(𝒩C/Y2)=deg1(𝒪1(1)𝒪1(1))=2,D^{3}=\deg_{C}(\mathcal{N}^{*}_{C/Y_{2}})=\deg_{\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}}(1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(1))=2,

where the first and second equalities follow from Lemma 7.2(3) and Lemma 4.7(2), respectively. We then obtain

(6.10.1) 2=D3=a(12bcμ1+3a(f1LE2)+a2E3).2=D^{3}=a\left(\frac{12bc}{\mu_{1}}+3a(f_{1}^{*}L\cdot E^{2})+a^{2}E^{3}\right).

By a>0,b,ca\in\mathbb{Z}_{>0},b,c\in\mathbb{Z}, and μ1{1,2}\mu_{1}\in\{1,2\}, it holds that a=1a=1 or a=2a=2. If a=2a=2, then the right hand side of (6.10.1) would be contained in 44\mathbb{Z}, which is absurd. Then we obtain a=1a=1, which implies D1=1D\cdot\ell_{1}=1 and deg(f1|D)=2μ1\deg(f_{1}|_{D})=\frac{2}{\mu_{1}}.

(1) Assume that R1R_{1} is of type C1C_{1}. Then we have deg(f1|D)=2\deg(f_{1}|_{D})=2, i.e., f1|Df_{1}|_{D} is a double cover. Hence (1) holds.

(2) Assume that R1R_{1} is of type C2C_{2}. Then we have deg(f1|D)=1\deg(f_{1}|_{D})=1, i.e., f1|Df_{1}|_{D} is a finite birational morphism. Since f1|D:D=1×1Y1=1×1f_{1}|_{D}:D=\mathbb{P}^{1}\times\mathbb{P}^{1}\to Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1} is a finite birational morphism of normal varieties, f1|Df_{1}|_{D} is an isomorphism. Thus (2) holds. ∎

6.2.1. Case C2E1C_{2}-E_{1}

Proposition 6.11.

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Assume that there exist two extremal rays R1R_{1} and R2R_{2} such that R1R_{1} is of type C2C_{2} and R2R_{2} is of type E1E_{1}. Then

X1×1(𝒪1×1𝒪1×1(1,1))X\simeq\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))

and (KX)3=52(-K_{X})^{3}=52.

Proof.

We use Notation 6.9. Recall that we have X1×1()X\simeq\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{E}) and =(f1)𝒪X(1)\mathcal{E}=(f_{1})_{*}\mathcal{O}_{X}(1). Set 𝒪D(D):=𝒪X(D)|D\mathcal{O}_{D}(D):=\mathcal{O}_{X}(D)|_{D}.

Step 1.

There is the following exact sequence:

0(f1)𝒪X(f1)𝒪X(D)(f1)𝒪D(D)0.0\to(f_{1})_{*}\mathcal{O}_{X}\to(f_{1})_{*}\mathcal{O}_{X}(D)\to(f_{1})_{*}\mathcal{O}_{D}(D)\to 0.
Proof of Step 1.

We have an exact sequence

0𝒪X𝒪X(D)𝒪D(D)0,0\to\mathcal{O}_{X}\to\mathcal{O}_{X}(D)\to\mathcal{O}_{D}(D)\to 0,

which induces another exact sequence

0(f1)𝒪X(f1)𝒪X(D)(f1)𝒪D(D)R1(f1)𝒪X.0\to(f_{1})_{*}\mathcal{O}_{X}\to(f_{1})_{*}\mathcal{O}_{X}(D)\to(f_{1})_{*}\mathcal{O}_{D}(D)\to R^{1}(f_{1})_{*}\mathcal{O}_{X}.

By Proposition 3.11, R1(f1)𝒪X=0R^{1}(f_{1})_{*}\mathcal{O}_{X}=0. This completes the proof of Step 1. ∎

Step 2.

The following hold.

  1. (1)

    (f1)𝒪X(D)(f_{1})_{*}\mathcal{O}_{X}(D)\simeq\mathcal{E}\otimes\mathcal{L} for some Pic(1×1)\mathcal{L}\in{\operatorname{Pic}}(\mathbb{P}^{1}\times\mathbb{P}^{1}).

  2. (2)

    (f1)𝒪D(D)𝒪1×1(1,1)(f_{1})_{*}\mathcal{O}_{D}(D)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1).

Proof of Step 2.

Let us show (1). Since f1:XY1=1×1f_{1}:X\to Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1} is a 1\mathbb{P}^{1}-bundle, it holds that

𝒪X(D)𝒪X(n)f1\mathcal{O}_{X}(D)\simeq\mathcal{O}_{X}(n)\otimes f_{1}^{*}\mathcal{L}

for some nn\in\mathbb{Z} and Pic(1×1)\mathcal{L}\in{\operatorname{Pic}}(\mathbb{P}^{1}\times\mathbb{P}^{1}). Fix a fibre ζ\zeta of f1:XY1=1×1f_{1}:X\to Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1}. Since f1|D:DY1f_{1}|_{D}:D\to Y_{1} is an isomorphism (Lemma 6.10), we have Dζ=1D\cdot\zeta=1. By 𝒪X(1)ζ=1\mathcal{O}_{X}(1)\cdot\zeta=1, we obtain n=1n=1. It holds that

(f1)𝒪X(D)(f1)(𝒪X(1)f1)((f1)𝒪X(1)).(f_{1})_{*}\mathcal{O}_{X}(D)\simeq(f_{1})_{*}(\mathcal{O}_{X}(1)\otimes f_{1}^{*}\mathcal{L})\simeq((f_{1})_{*}\mathcal{O}_{X}(1))\otimes\mathcal{L}\simeq\mathcal{E}\otimes\mathcal{L}.

Thus (1) holds.

Let us show (2). Since f1|D:DY1=1×1f_{1}|_{D}:D\to Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1} is an isomorphism, Lemma 4.7(4) implies (f1)𝒪D(D)𝒪1×1(1,1)(f_{1})_{*}\mathcal{O}_{D}(D)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1). This completes the proof of Step 2. ∎

Step 3.

X1×1(𝒪1×1𝒪1×1(1,1))X\simeq\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)).

Proof of Step 3.

By Step 1and Step 2, we have the following exact sequence for some Pic(1×1)\mathcal{L}\in{\operatorname{Pic}}\,(\mathbb{P}^{1}\times\mathbb{P}^{1}):

0𝒪1×1𝒪1×1(1,1)0.0\to\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\to\mathcal{E}\otimes\mathcal{L}\to\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)\to 0.

It holds that

Ext1×11(𝒪1×1(1,1),𝒪1×1)H1(1×1,𝒪1×1(1,1))=0,{\operatorname{Ext}}^{1}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1),\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}})\simeq H^{1}(\mathbb{P}^{1}\times\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))=0,

which implies

𝒪1×1𝒪1×1(1,1).\mathcal{E}\otimes\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1).

Therefore,

X1×1()1×1(𝒪1×1(1,1))1×1(𝒪1×1𝒪1×1(1,1)).X\simeq\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{E})\simeq\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{E}\otimes\mathcal{L}\otimes\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))\simeq\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)).

This completes the proof of Step 3. ∎

Step 4.

(KX)3=52(-K_{X})^{3}=52.

Proof of Step 4.

By Proposition 7.1(2), we have

(KX)3=2c1(𝒪1×1𝒪1×1(1,1))28c2(𝒪1×1𝒪1×1(1,1))+6K1×12.(-K_{X})^{3}=2c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))^{2}-8c_{2}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))+6K_{\mathbb{P}^{1}\times\mathbb{P}^{1}}^{2}.

By [Har77, Appendix A, §3, C3 and C5], we obtain

c1(𝒪1×1𝒪1×1(1,1))\displaystyle c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)) =c1(𝒪1×1)+c1(𝒪1×1(1,1))=c1(𝒪1×1(1,1)),\displaystyle=c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}})+c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))=c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)),
c2(𝒪1×1𝒪1×1(1,1))\displaystyle c_{2}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)) =c1(𝒪1×1)c1(𝒪1×1(1,1))=0.\displaystyle=c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}})\cdot c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))=0.

Hence it holds that (KX)3=2280+68=52.(-K_{X})^{3}=2\cdot 2-8\cdot 0+6\cdot 8=52. This completes the proof of Step 4. ∎

Step 3 and Step 4 complete the proof of Proposition 6.11

6.2.2. Case C1E1C_{1}-E_{1}

Notation 6.12.

We use Notation 6.9. Assume that R1R_{1} is of type C1C_{1}. In particular, XX is a primitive Fano threefold with ρ(X)=3\rho(X)=3, R1R_{1} is of type C1C_{1}, and R2R_{2} is of type E1E_{1}. By Lemma 3.15, we have morphisms

X{X}P:=(){P^{\prime}:=\mathbb{P}(\mathcal{E}^{\prime})}Y1=1×1{Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1}}ι\scriptstyle{\iota^{\prime}}π\scriptstyle{\pi^{\prime}}

where :=(f1)𝒪X(KX)\mathcal{E}^{\prime}:=(f_{1})_{*}\mathcal{O}_{X}(-K_{X}) and ι\iota^{\prime} is a closed immersion. We identify XX with the smooth prime divisor ι(X)\iota^{\prime}(X) on PP^{\prime}.

Lemma 6.13.

We use Notation 6.12. For the induced morphism

f1|D:D=1×1Y1=1×1,f_{1}|_{D}:D=\mathbb{P}^{1}\times\mathbb{P}^{1}\to Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1},

we set :=((f1|D)𝒪D/𝒪Y1)1\mathcal{L}:=((f_{1}|_{D})_{*}\mathcal{O}_{D}/\mathcal{O}_{Y_{1}})^{-1}, which is an invertible sheaf (cf. Remark 2.2, Lemma 6.10(1)). Then, after possibly permuting the direct product factors of Y1=1×1Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1}, the following hold.

  1. (1)

    𝒪1×1(0,1)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(0,1).

  2. (2)

    f1|Df_{1}|_{D} is a split double cover.

  3. (3)

    There exists a double cover h:11h:\mathbb{P}^{1}\to\mathbb{P}^{1} such that f1|D=id×h:D=1×1Y1=1×1f_{1}|_{D}={\rm id}\times h:D=\mathbb{P}^{1}\times\mathbb{P}^{1}\to Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1}.

  4. (4)

    (f1|D)𝒪D(0,1)𝒪1×1(0,1)𝒪1×1(0,1)(f_{1}|_{D})_{*}\mathcal{O}_{D}(0,-1)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(0,-1)\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(0,-1).

Proof.

Since f1|D:D1×11×1f_{1}|_{D}\colon D\simeq\mathbb{P}^{1}\times\mathbb{P}^{1}\to\mathbb{P}^{1}\times\mathbb{P}^{1} is a double cover (Lemma 6.10(1)), Lemma 2.3 implies

ωD(f1|D)(ω1×1).\omega_{D}\simeq(f_{1}|_{D})^{*}(\omega_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\otimes\mathcal{L}).

We can write 𝒪1×1(a,b)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(a,b) for some a,ba,b\in\mathbb{Z}. We then get

(6.13.1) 𝒪D(2,2)(f1|D)(𝒪1×1(a2,b2)).\mathcal{O}_{D}(-2,-2)\simeq(f_{1}|_{D})^{*}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(a-2,b-2)).

Hence a<2a<2 and b<2b<2. Furthermore, we get

8\displaystyle 8 =(c1(𝒪D(2,2)))2\displaystyle=(c_{1}(\mathcal{O}_{D}(-2,-2)))^{2}
=deg(f1|D)c1(𝒪1×1(a2,b2))2\displaystyle=\deg(f_{1}|_{D})\cdot c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(a-2,b-2))^{2}
=4(a2)(b2),\displaystyle=4(a-2)(b-2),

i.e., (a2)(b2)=2(a-2)(b-2)=2. Therefore, after possibly permuting the direct product factors of Y1=1×1Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1}, we obtain (a,b)=(0,1)(a,b)=(0,1) and 𝒪1×1(0,1)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(0,1). Thus (1) holds. Lemma 2.4 and Lemma 8.4 imply (2) and (3), respectively.

Let us show (4). It is easy to see that h𝒪1(1)h_{*}\mathcal{O}_{\mathbb{P}^{1}}(1) is a locally free sheaf of rank 2 on 1\mathbb{P}^{1}. Hence we can write h𝒪1(1)𝒪1(c)𝒪1(d)h_{*}\mathcal{O}_{\mathbb{P}^{1}}(1)\simeq\mathcal{O}_{\mathbb{P}^{1}}(c)\oplus\mathcal{O}_{\mathbb{P}^{1}}(d) for some c,dc,d\in\mathbb{Z}. We obtain

h0(𝒪1(c))+h0(𝒪1(d))\displaystyle h^{0}(\mathcal{O}_{\mathbb{P}^{1}}(c))+h^{0}(\mathcal{O}_{\mathbb{P}^{1}}(d)) =\displaystyle= h0(1,h𝒪1(1))=h0(1,𝒪1(1))=2\displaystyle h^{0}(\mathbb{P}^{1},h_{*}\mathcal{O}_{\mathbb{P}^{1}}(1))=h^{0}(\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}}(1))=2
h0(𝒪1(c1))+h0(𝒪1(d1))\displaystyle h^{0}(\mathcal{O}_{\mathbb{P}^{1}}(c-1))+h^{0}(\mathcal{O}_{\mathbb{P}^{1}}(d-1)) =\displaystyle= h0(1,𝒪1(1)h𝒪1(1))=h0(1,𝒪1(1))=0.\displaystyle h^{0}(\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}}(-1)\otimes h_{*}\mathcal{O}_{\mathbb{P}^{1}}(1))=h^{0}(\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}}(-1))=0.

Therefore, we get c=d=0c=d=0 and

h𝒪1(1)𝒪1(1)h𝒪1(1)𝒪1(1)𝒪1(1).h_{*}\mathcal{O}_{\mathbb{P}^{1}}(-1)\simeq\mathcal{O}_{\mathbb{P}^{1}}(-1)\otimes h_{*}\mathcal{O}_{\mathbb{P}^{1}}(1)\simeq\mathcal{O}_{\mathbb{P}^{1}}(-1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(-1).

This, together with the flat base change theorem, implies the following:

(f1|D)𝒪D(0,1)\displaystyle(f_{1}|_{D})_{*}\mathcal{O}_{D}(0,-1) (id×h)𝒪D(0,1)\displaystyle\simeq({\rm id}\times h)_{*}\mathcal{O}_{D}(0,-1)
(id×h)(pr2)𝒪1(1)\displaystyle\simeq({\rm id}\times h)_{*}({\rm pr}_{2})^{*}\mathcal{O}_{\mathbb{P}^{1}}(-1)
(pr2)h𝒪1(1)\displaystyle\simeq({\rm pr}_{2})^{*}h_{*}\mathcal{O}_{\mathbb{P}^{1}}(-1)
(pr2)(𝒪1(1)𝒪1(1))\displaystyle\simeq({\rm pr}_{2})^{*}(\mathcal{O}_{\mathbb{P}^{1}}(-1)\oplus\mathcal{O}_{\mathbb{P}^{1}}(-1))
𝒪1×1(0,1)𝒪1×1(0,1).\displaystyle\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(0,-1)\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(0,-1).

Thus (4) holds. ∎

Lemma 6.14.

We use Notation 6.12. Then, after possibly permuting the direct product factors of Y1=1×1Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1}, the following holds:

(f1)𝒪X(KX)𝒪1×1(2,1)𝒪1×1(1,0)2.(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,1)\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)^{\oplus 2}.
Proof.
Step 1.

The following hold.

  1. (1)

    (KXD)ζ=0(-K_{X}-D)\cdot\zeta=0 for a fibre ζ\zeta of f1:XY1f_{1}:X\to Y_{1}.

  2. (2)

    There exists the following exact sequence:

    0(f1)𝒪X(KXD)(f1)𝒪X(KX)(f1)𝒪D(KX)0.0\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\longrightarrow(f_{1})_{*}\mathcal{O}_{D}(-K_{X})\longrightarrow 0.
Proof of Step 1.

The assertion (1) follows from KXζ=2-K_{X}\cdot\zeta=2 and Dζ=2D\cdot\zeta=2 (Lemma 6.10). Let us show (2). We have an exact sequence

0𝒪X(KXD)𝒪X(KX)𝒪D(KX)0,0\longrightarrow\mathcal{O}_{X}(-K_{X}-D)\longrightarrow\mathcal{O}_{X}(-K_{X})\longrightarrow\mathcal{O}_{D}(-K_{X})\longrightarrow 0,

which induces another exact sequence

0(f1)𝒪X(KXD)(f1)𝒪X(KX)(f1)𝒪D(KX)R1(f1)𝒪X(KXD).\displaystyle 0\to(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\to(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\to(f_{1})_{*}\mathcal{O}_{D}(-K_{X})\to R^{1}(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D).

By (1) and [Tan15, Theorem 0.5], we obtain R1f1𝒪X(KXD)=0R^{1}f_{1*}\mathcal{O}_{X}(-K_{X}-D)=0. Thus (2) holds. This completes the proof of Step 1. ∎

Step 2.

(f1)𝒪X(KXD)𝒪1×1(2,1)(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,1).

Proof of Step 2.

By Step 1(1), we have (KXD)ζ=0(-K_{X}-D)\cdot\zeta=0. Therefore, we can write

𝒪X(KXD)f1𝒪1×1(s,t)\mathcal{O}_{X}(-K_{X}-D)\simeq f_{1}^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(s,t)

for some s,ts,t\in\mathbb{Z} (Theorem 4.11). We obtain

𝒪D(2,2)\displaystyle\mathcal{O}_{D}(2,2) 𝒪D(KD)\displaystyle\simeq\mathcal{O}_{D}(-K_{D})
𝒪X(KXD)|D\displaystyle\simeq\mathcal{O}_{X}(-K_{X}-D)|_{D}
(f1|D)𝒪1×1(s,t)\displaystyle\simeq(f_{1}|_{D})^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(s,t)
(id×h)𝒪1×1(s,t)\displaystyle\simeq({\rm id}\times h)^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(s,t)
𝒪D(s,2t),\displaystyle\simeq\mathcal{O}_{D}(s,2t),

where the fourth and fifth isomorphisms follow from Lemma 6.13. Hence we get s=2s=2 and t=1t=1. Then it holds that

(f1)𝒪X(KXD)(f1)f1𝒪1×1(s,t)𝒪1×1(s,t)𝒪1×1(2,1).(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\simeq(f_{1})_{*}f_{1}^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(s,t)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(s,t)\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,1).

This completes the proof of Step 2. ∎

Step 3.

(f1)𝒪D(KX)𝒪1×1(1,0)2(f_{1})_{*}\mathcal{O}_{D}(-K_{X})\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)^{\oplus 2}.

Proof of Step 3.

Recall that we have

(f1)𝒪D(KX)=(f1|D)(𝒪X(KX)|D).(f_{1})_{*}\mathcal{O}_{D}(-K_{X})=(f_{1}|_{D})_{*}(\mathcal{O}_{X}(-K_{X})|_{D}).

By 𝒪X(KXD)|D𝒪D(2,2)\mathcal{O}_{X}(-K_{X}-D)|_{D}\simeq\mathcal{O}_{D}(2,2) and 𝒪X(D)|D𝒪D(1,1)\mathcal{O}_{X}(D)|_{D}\simeq\mathcal{O}_{D}(-1,-1) (Theorem 4.8), it holds that

(f1|D)(𝒪X(KX)|D)(f1|D)𝒪D(1,1).(f_{1}|_{D})_{*}(\mathcal{O}_{X}(-K_{X})|_{D})\simeq(f_{1}|_{D})_{*}\mathcal{O}_{D}(1,1).

By Lemma 6.13, we obtain

(f1)𝒪D(KX)\displaystyle(f_{1})_{*}\mathcal{O}_{D}(-K_{X}) (f1|D)𝒪D(1,1)\displaystyle\simeq(f_{1}|_{D})_{*}\mathcal{O}_{D}(1,1)
(id×h)(𝒪D(1,2)𝒪D(0,1))\displaystyle\simeq({\rm id}\times h)_{*}(\mathcal{O}_{D}(1,2)\otimes\mathcal{O}_{D}(0,-1))
(id×h)(((id×h)𝒪1×1(1,1))𝒪D(0,1))\displaystyle\simeq({\rm id}\times h)_{*}((({\rm id}\times h)^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))\otimes\mathcal{O}_{D}(0,-1))
𝒪1×1(1,1)(id×h)𝒪D(0,1)\displaystyle\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)\otimes({\rm id}\times h)_{*}\mathcal{O}_{D}(0,-1)
𝒪1×1(1,0)2.\displaystyle\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)^{\oplus 2}.

This completes the proof of Step 3. ∎

Step 4.

It holds that

(f1)𝒪X(KX)𝒪1×1(2,1)𝒪1×1(1,0)2.(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,1)\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)^{\oplus 2}.
Proof of Step 4.

By Step 1, we have the following exact sequence:

0(f1)𝒪X(KXD)(f1)𝒪X(KX)(f1)𝒪D(KX)0.0\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(-K_{X}-D)\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\longrightarrow(f_{1})_{*}\mathcal{O}_{D}(-K_{X})\longrightarrow 0.

By Step 2 and Step 3, this exact sequence can be written as follows:

0𝒪1×1(2,1)(f1)𝒪X(KX)𝒪1×1(1,0)20.0\longrightarrow\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,1)\longrightarrow(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\longrightarrow\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)^{\oplus 2}\longrightarrow 0.

Hence it suffices to show that this exact sequence splits, which follows from the following computation:

Ext1(𝒪1×1(1,0)2,𝒪1×1(2,1))H1(1×1,𝒪1×1(1,1))2=0.\mathrm{Ext}^{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)^{\oplus 2},\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,1))\simeq H^{1}(\mathbb{P}^{1}\times\mathbb{P}^{1},\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1))^{\oplus 2}=0.

This completes the proof of Step 4. ∎

Step 4 completes the proof of Lemma 6.14. ∎

Lemma 6.15.

We use Notation 6.12. Then it holds that

𝒪P(X)𝒪P(2)π𝒪1×1(2,1).\mathcal{O}_{P^{\prime}}(X)\simeq\mathcal{O}_{P^{\prime}}(2)\otimes\pi^{\prime*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1).
Proof.

We have the following morphisms:

f1:XιP=()πY1=1×1,f_{1}:X\overset{\iota^{\prime}}{\hookrightarrow}P^{\prime}=\mathbb{P}(\mathcal{E}^{\prime})\xrightarrow{\pi^{\prime}}Y_{1}=\mathbb{P}^{1}\times\mathbb{P}^{1},

where =(f1)𝒪X(KX)=𝒪1×1(2,1)𝒪1×1(1,0)2\mathcal{E}^{\prime}=(f_{1})_{*}\mathcal{O}_{X}(-K_{X})=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,1)\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)^{\oplus 2} (Lemma 6.14). We identify XX with ι(X)\iota^{\prime}(X), which is a smooth prime divisor on PP^{\prime}. Since ι:XP\iota^{\prime}\colon X\hookrightarrow P^{\prime} is the closed immersion induced by the natural surjection f1(f1)𝒪X(KX)𝒪X(KX)f_{1}^{*}(f_{1})_{*}\mathcal{O}_{X}(-K_{X})\to\mathcal{O}_{X}(-K_{X}), it follows from [Har77, Ch. II, the proof of Propsition 7.12] that

(6.15.1) 𝒪X(KX)ι𝒪P(1)=𝒪P(1)|X.\mathcal{O}_{X}(-K_{X})\simeq\iota^{\prime*}\mathcal{O}_{P^{\prime}}(1)=\mathcal{O}_{P^{\prime}}(1)|_{X}.

By Proposition 7.1(2),

𝒪P(KP)\displaystyle\mathcal{O}_{P^{\prime}}(-K_{P^{\prime}}) 𝒪P(3)π(ω1×1det)1\displaystyle\simeq\mathcal{O}_{P^{\prime}}(3)\otimes\pi^{\prime*}(\omega_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\otimes\det\mathcal{E}^{\prime})^{-1}
𝒪P(3)π(𝒪1×1(2,2)𝒪1×1(4,1))1\displaystyle\simeq\mathcal{O}_{P^{\prime}}(3)\otimes\pi^{\prime*}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,-2)\otimes\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(4,1))^{-1}
(6.15.2) 𝒪P(3)π𝒪1×1(2,1).\displaystyle\simeq\mathcal{O}_{P^{\prime}}(3)\otimes\pi^{\prime*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1).

By (6.15.1) and (6.2.2), we obtain

𝒪P(X)|X𝒪P(KP)|X𝒪X(KX)𝒪P(2)|Xπ𝒪1×1(2,1)|X.\mathcal{O}_{P^{\prime}}(X)|_{X}\simeq\mathcal{O}_{P^{\prime}}(-K_{P^{\prime}})|_{X}\otimes\mathcal{O}_{X}(K_{X})\simeq\mathcal{O}_{P^{\prime}}(2)|_{X}\otimes\pi^{\prime*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1)|_{X}.

Set

𝒩:=𝒪P(X)1𝒪P(2)π𝒪1×1(2,1).\mathcal{N}:=\mathcal{O}_{P^{\prime}}(X)^{-1}\otimes\mathcal{O}_{P^{\prime}}(2)\otimes\pi^{\prime*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1).

We have 𝒩|X𝒪X\mathcal{N}|_{X}\simeq\mathcal{O}_{X}. For a fibre ζ\zeta of f1:X1×1f_{1}:X\to\mathbb{P}^{1}\times\mathbb{P}^{1}, we get 𝒩ζ=(𝒩|X)ζ=0\mathcal{N}\cdot\zeta=(\mathcal{N}|_{X})\cdot\zeta=0, which implies 𝒩π𝒩\mathcal{N}\simeq\pi^{\prime*}\mathcal{N}^{\prime} for some 𝒩Pic(1×1)\mathcal{N}^{\prime}\in{\operatorname{Pic}}(\mathbb{P}^{1}\times\mathbb{P}^{1}). Since the pullback of 𝒩\mathcal{N}^{\prime} to XX is trivial: f1𝒩𝒩|X𝒪Xf_{1}^{*}\mathcal{N}^{\prime}\simeq\mathcal{N}|_{X}\simeq\mathcal{O}_{X}, it holds that 𝒩𝒪1×1\mathcal{N}^{\prime}\simeq\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}. Therefore, 𝒩π𝒩𝒪P\mathcal{N}\simeq\pi^{\prime*}\mathcal{N}^{\prime}\simeq\mathcal{O}_{P^{\prime}}, as required. ∎

Proposition 6.16.

We use Notation 6.12. Then there exists a closed immersion

ι:XP:=1×1(𝒪1×1𝒪1×1(1,1)2)\iota:X\hookrightarrow P:=\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2})

such that

𝒪P(ι(X))𝒪P(2)π𝒪1×1(2,3),\mathcal{O}_{P}(\iota(X))\simeq\mathcal{O}_{P}(2)\otimes\pi^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,3),

where π:1×1(𝒪1×1𝒪1×1(1,1)2)1×1\pi:\mathbb{P}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2})\to\mathbb{P}^{1}\times\mathbb{P}^{1} denotes the induced projection. Furthermore, it holds that (KX)3=14(-K_{X})^{3}=14.

Proof.

By Lemma 6.14 and Lemma 6.15, we may assume that

  • =𝒪1×1(2,1)𝒪1×1(1,0)2\mathcal{E}^{\prime}=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,1)\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,0)^{\oplus 2},

  • P=()P^{\prime}=\mathbb{P}(\mathcal{E}^{\prime}),

  • XX is a smooth prime divisor on PP^{\prime}, and

  • 𝒪P(X)𝒪P(2)π𝒪1×1(2,1)\mathcal{O}_{P^{\prime}}(X)\simeq\mathcal{O}_{P^{\prime}}(2)\otimes\pi^{\prime*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1).

Set :=𝒪1×1(2,1)\mathcal{M}:=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,-1) and

:==𝒪1×1𝒪1×1(1,1)2.\mathcal{E}:=\mathcal{E}^{\prime}\otimes\mathcal{M}=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2}.

We have the induced morphisms:

π:Pψ,Pπ1×1\pi:P\xrightarrow{\psi,\simeq}P^{\prime}\xrightarrow{\pi^{\prime}}\mathbb{P}^{1}\times\mathbb{P}^{1}
X{X}P=(𝒪1×1𝒪1×1(1,1)2){P=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2})}1×1.{\mathbb{P}^{1}\times\mathbb{P}^{1}.}ι:=ψ1ι\scriptstyle{\iota:=\psi^{-1}\circ\iota^{\prime}}π:=πψ\scriptstyle{\pi:=\pi^{\prime}\circ\psi}

By 𝒪P(1)ψ𝒪P(1)π\mathcal{O}_{P}(1)\simeq\psi^{*}\mathcal{O}_{P^{\prime}}(1)\otimes\pi^{*}\mathcal{M} [Har77, Ch. II, Lemma 7.9], it holds that

𝒪P(ι(X))\displaystyle\mathcal{O}_{P}(\iota(X)) ψ𝒪P(X)\displaystyle\simeq\psi^{*}\mathcal{O}_{P^{\prime}}(X)
ψ𝒪P(2)ψπ𝒪1×1(2,1)\displaystyle\simeq\psi^{*}\mathcal{O}_{P^{\prime}}(2)\otimes\psi^{*}\pi^{\prime*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1)
(ψ𝒪P(1))2ψπ𝒪1×1(2,1)\displaystyle\simeq(\psi^{*}\mathcal{O}_{P^{\prime}}(1))^{\otimes 2}\otimes\psi^{*}\pi^{\prime*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1)
(𝒪P(1)π1)2ψπ𝒪1×1(2,1)\displaystyle\simeq(\mathcal{O}_{P}(1)\otimes\pi^{*}\mathcal{M}^{-1})^{\otimes 2}\otimes\psi^{*}\pi^{\prime*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1)
𝒪P(2)π2π𝒪1×1(2,1)\displaystyle\simeq\mathcal{O}_{P}(2)\otimes\pi^{*}\mathcal{M}^{-2}\otimes\pi^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-2,1)
𝒪P(2)π𝒪1×1(2,3).\displaystyle\simeq\mathcal{O}_{P}(2)\otimes\pi^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,3).

What is remaining is to compute (KX)3(-K_{X})^{3}. By Lemma 7.1(5), we have

(KX)3=2c1()2\displaystyle(-K_{X})^{3}=2c_{1}(\mathcal{E})^{2}- 2c2()+4(c1()c1())+6(c1()K1×1)\displaystyle 2c_{2}(\mathcal{E})+4(c_{1}(\mathcal{E})\cdot c_{1}(\mathcal{F}))+6(c_{1}(\mathcal{E})\cdot K_{\mathbb{P}^{1}\times\mathbb{P}^{1}})
+9(c1()K1×1)+6K1×12+3c1()2,\displaystyle+9(c_{1}(\mathcal{F})\cdot K_{\mathbb{P}^{1}\times\mathbb{P}^{1}})+6K_{\mathbb{P}^{1}\times\mathbb{P}^{1}}^{2}+3c_{1}(\mathcal{F})^{2},

where :=𝒪1×1(2,3)\mathcal{F}:=\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,3). By [Har77, Appendix A, §3, C.3 and C.5], it holds that

c1()\displaystyle c_{1}(\mathcal{E}) =2c1(𝒪1×1(1,1))and\displaystyle=2c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1))\quad{\rm and}
c2()\displaystyle c_{2}(\mathcal{E}) =c1(𝒪1×1(1,1))2=2.\displaystyle=c_{1}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1))^{2}=2.

Hence we obtain

(KX)3=2822+4(10)+68+9(10)+68+312=14.\displaystyle(-K_{X})^{3}=2\cdot 8-2\cdot 2+4\cdot(-10)+6\cdot 8+9\cdot(-10)+6\cdot 8+3\cdot 12=14.

Theorem 6.17.

Let XX be a primitive Fano threefold with ρ(X)=3\rho(X)=3. Let R1R_{1} and R2R_{2} be extremal rays of NE(X){\operatorname{NE}}(X). Assume that R1R_{1} is of type CC and R2R_{2} is of type EE. Then one and only one of the following holds.

  1. (1)

    R1R_{1} is of type C1C_{1}, R2R_{2} is of type E1E_{1}, (KX)3=14(-K_{X})^{3}=14, and XX is isomorphic to a prime divisor XX^{\prime} on P:=(𝒪1×1𝒪1×1(1,1)2)P:=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2}) such that 𝒪P(X)𝒪P(2)π𝒪1×1(2,3)\mathcal{O}_{P}(X^{\prime})\simeq\mathcal{O}_{P}(2)\otimes\pi^{*}\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,3), where π:P=(𝒪1×1𝒪1×1(1,1)2)1×1\pi:P=\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(-1,-1)^{\oplus 2})\to\mathbb{P}^{1}\times\mathbb{P}^{1} denotes the natural projection.

  2. (2)

    R1R_{1} is of type C2C_{2}, R2R_{2} is of type E1E_{1}, (KX)3=52(-K_{X})^{3}=52, and X(𝒪1×1𝒪1×1(1,1))X\simeq\mathbb{P}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}\oplus\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,1)).

Proof.

See Proposition 6.11 and Proposition 6.16. ∎

7. Appendix: Computation of Chern classes

In this section, we compute some intersection numbers on projective space bundles (Proposition 7.1) and blowups (Proposition 7.2). We include the proofs for the sake of completeness, although both results are obtained just by applying standard results on Chern classes.

Proposition 7.1.

Let YY be an nn-dimensional smooth variety and let \mathcal{E} be a locally free sheaf of rank rr on YY. Let π:X=()Y\pi\colon X=\mathbb{P}(\mathcal{E})\to Y be the projection induced by the r1\mathbb{P}^{r-1}-bundle structure. Then the following hold.

  1. (1)

    If r=2r=2, then it holds that 𝒪X(1)F=1\mathcal{O}_{X}(1)\cdot F=1 for any fibre FF of π:XY\pi:X\to Y over a closed point.

  2. (2)

    The following holds:

    ωX𝒪X(r)π(ωYdet).\omega_{X}\simeq\mathcal{O}_{X}(-r)\otimes\pi^{*}(\omega_{Y}\otimes\det\mathcal{E}).
  3. (3)

    If n=2n=2 and r=2r=2, then

    (KX)3=2c1()28c2()+6KY2.(-K_{X})^{3}=2c_{1}(\mathcal{E})^{2}-8c_{2}(\mathcal{E})+6K_{Y}^{2}.
  4. (4)

    If n=1n=1 and r=2r=2, then

    c1(𝒪X(1))2=degY().c_{1}(\mathcal{O}_{X}(1))^{2}=\deg_{Y}(\mathcal{E}).
  5. (5)

    If n=2n=2, r=3,r=3, and DD is a smooth prime divisor on XX such that 𝒪X(D)𝒪X(2)π\mathcal{O}_{X}(D)\simeq\mathcal{O}_{X}(2)\otimes\pi^{*}\mathcal{F} for some invertible sheaf \mathcal{F} on YY, then

    (KD)3=\displaystyle(-K_{D})^{3}= 2c1()22c2()+4c1()c1()+6c1()KY\displaystyle 2c_{1}(\mathcal{E})^{2}-2c_{2}(\mathcal{E})+4c_{1}(\mathcal{E})\cdot c_{1}(\mathcal{F})+6c_{1}(\mathcal{E})\cdot K_{Y}
    +9c1()KY+6KY2+3c1()2.\displaystyle+9c_{1}(\mathcal{F})\cdot K_{Y}+6K_{Y}^{2}+3c_{1}(\mathcal{F})^{2}.
Proof.

Fix a Cartier divisor ξ\xi on XX such that 𝒪X(1)𝒪X(ξ)\mathcal{O}_{X}(1)\simeq\mathcal{O}_{X}(\xi). The assertion (1) holds by

𝒪X(1)F=degF(𝒪X(1)|F)=deg1(𝒪1(1))=1.\mathcal{O}_{X}(1)\cdot F=\deg_{F}(\mathcal{O}_{X}(1)|_{F})=\deg_{\mathbb{P}^{1}}(\mathcal{O}_{\mathbb{P}^{1}}(1))=1.

Let us show (2). Since π\pi is locally trivial, it follows from [Har77, Ch. II, Proposition 8.11 and Exercise 8.3(a)] that we have an exact sequence

0πΩYΩXΩX/Y0.0\to\pi^{*}\Omega_{Y}\longrightarrow\Omega_{X}\longrightarrow\Omega_{X/Y}\longrightarrow 0.

By [Har77, Ch. II, Exercise 5.16(d),(e)], it holds that

ωX\displaystyle\omega_{X} n+r1ΩX\displaystyle\simeq\bigwedge^{n+r-1}\Omega_{X}
nπΩYr1ΩX/Y\displaystyle\simeq\bigwedge^{n}\pi^{*}\Omega_{Y}\otimes\bigwedge^{r-1}\Omega_{X/Y}
π(nΩY)r1ΩX/Y\displaystyle\simeq\pi^{*}\left(\bigwedge^{n}\Omega_{Y}\right)\otimes\bigwedge^{r-1}\Omega_{X/Y}
πωYωX/Y.\displaystyle\simeq\pi^{*}\omega_{Y}\otimes\omega_{X/Y}.

Moreover, we have ωX/Y𝒪X(r)πdet\omega_{X/Y}\simeq\mathcal{O}_{X}(-r)\otimes\pi^{*}\det\mathcal{E} by [Har77, Ch. III, Exercise 8.4(b)]. Thus (2) holds.

Let us show (3). Set :=ωYdet\mathcal{L}:=\omega_{Y}\otimes\det\mathcal{E}. It follows from (2) that ωX𝒪X(2)π\omega_{X}\simeq\mathcal{O}_{X}(-2)\otimes\pi^{*}\mathcal{L}. We then obtain

(7.1.1) (KX)3=(2ξc1(π))3=8ξ312ξ2πc1()+6ξπc1()2.\displaystyle(-K_{X})^{3}=(2\xi-c_{1}(\pi^{*}\mathcal{L}))^{3}=8\xi^{3}-12\xi^{2}\cdot\pi^{*}c_{1}(\mathcal{L})+6\xi\cdot\pi^{*}c_{1}(\mathcal{L})^{2}.

By [Har77, Appendix A, §3, page 429], we have an equation

πc0()ξ2πc1()ξ+πc2()=0.\pi^{*}c_{0}(\mathcal{E})\cdot\xi^{2}-\pi^{*}c_{1}(\mathcal{E})\cdot\xi+\pi^{*}c_{2}(\mathcal{E})=0.

By c0()=1c_{0}(\mathcal{E})=1, this can be written as

(7.1.2) ξ2=πc1()ξπc2().\xi^{2}=\pi^{*}c_{1}(\mathcal{E})\cdot\xi-\pi^{*}c_{2}(\mathcal{E}).

We then get

ξ3\displaystyle\xi^{3} =πc1()ξ2πc2()ξ\displaystyle=\pi^{*}c_{1}(\mathcal{E})\cdot\xi^{2}-\pi^{*}c_{2}(\mathcal{E})\cdot\xi
=πc1()(πc1()ξπc2())πc2()ξ\displaystyle=\pi^{*}c_{1}(\mathcal{E})\cdot(\pi^{*}c_{1}(\mathcal{E})\cdot\xi-\pi^{*}c_{2}(\mathcal{E}))-\pi^{*}c_{2}(\mathcal{E})\cdot\xi
(7.1.3) =(πc1())2ξπc1()πc2()πc2()ξ\displaystyle=(\pi^{*}c_{1}(\mathcal{E}))^{2}\cdot\xi-\pi^{*}c_{1}(\mathcal{E})\cdot\pi^{*}c_{2}(\mathcal{E})-\pi^{*}c_{2}(\mathcal{E})\cdot\xi
=π(c1()2c2())ξπ(c1()c2())\displaystyle=\pi^{*}(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))\cdot\xi-\pi^{*}(c_{1}(\mathcal{E})\cdot c_{2}(\mathcal{E}))
=π(c1()2c2())ξ,\displaystyle=\pi^{*}(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))\cdot\xi,

where the first and second equalities follows from (7.1.2) and the last one holds, because dimY=2\dim Y=2 implies c1()c2()=0c_{1}(\mathcal{E})\cdot c_{2}(\mathcal{E})=0. By (7.1.1), (7.1.2), and (7), we obtain

(KX)3\displaystyle(-K_{X})^{3} =8π(c1()2c2())ξ12(πc1()ξπc2())πc1()+6ξπc1()2\displaystyle=8\pi^{*}(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))\cdot\xi-12(\pi^{*}c_{1}(\mathcal{E})\cdot\xi-\pi^{*}c_{2}(\mathcal{E}))\cdot\pi^{*}c_{1}(\mathcal{L})+6\xi\cdot\pi^{*}c_{1}(\mathcal{L})^{2}
=8π(c1()2c2())ξ12π(c1()c1())ξ+6πc1()2ξ,\displaystyle=8\pi^{*}(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))\cdot\xi-12\pi^{*}(c_{1}(\mathcal{E})\cdot c_{1}(\mathcal{L}))\cdot\xi+6\pi^{*}c_{1}(\mathcal{L})^{2}\cdot\xi,
=8(c1()2c2())12c1()c1()+6c1()2.\displaystyle=8(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))-12c_{1}(\mathcal{E})\cdot c_{1}(\mathcal{L})+6c_{1}(\mathcal{L})^{2}.

where the second equality holds by c2()c1()=0c_{2}(\mathcal{E})\cdot c_{1}(\mathcal{L})=0 and the third one follows from πZξ=degZ\pi^{*}Z\cdot\xi=\deg Z for any 0-cycle ZA2(Y)Z\in\mathrm{A}^{2}(Y) (note that deg\deg is dropped by abuse of notation). By c1()=c1(ωY)+c1(det)=KY+c1()c_{1}(\mathcal{L})=c_{1}(\omega_{Y})+c_{1}(\det\mathcal{E})=K_{Y}+c_{1}(\mathcal{E}), we obtain

(KX)3\displaystyle(-K_{X})^{3} =8(c1()2c2())12c1()(KY+c1())+6(KY+c1())2\displaystyle=8(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))-12c_{1}(\mathcal{E})\cdot(K_{Y}+c_{1}(\mathcal{E}))+6(K_{Y}+c_{1}(\mathcal{E}))^{2}
=2c1()28c2()+6KY2.\displaystyle=2c_{1}(\mathcal{E})^{2}-8c_{2}(\mathcal{E})+6K_{Y}^{2}.

Thus (3) holds.

Let us show (4). It follows from [Har77, Appendix A, §3, page 429] that

ξ2=πc1()ξπc2().\xi^{2}=\pi^{*}c_{1}(\mathcal{E})\cdot\xi-\pi^{*}c_{2}(\mathcal{E}).

By dimY=1\dim Y=1, we have c2()=0c_{2}(\mathcal{E})=0, which implies

c1(𝒪X(1))2=ξ2=πc1()ξ=degY().c_{1}(\mathcal{O}_{X}(1))^{2}=\xi^{2}=\pi^{*}c_{1}(\mathcal{E})\cdot\xi=\deg_{Y}(\mathcal{E}).

Thus (4) holds.

Let us show (5). By the adjunction formula,

(KD)3=((KXD)|D)3=(KXD)3D.\displaystyle(-K_{D})^{3}=((-K_{X}-D)|_{D})^{3}=(-K_{X}-D)^{3}\cdot D.

By (2) and our assumption, we get KX3ξ+πc1(ωYdet)K_{X}\sim-3\xi+\pi^{*}c_{1}(\omega_{Y}\otimes\det\mathcal{E}) and D2ξ+πc1()D\sim 2\xi+\pi^{*}c_{1}(\mathcal{F}), respectively. Thus

KX+Dξ+πc1(ωYdet)=ξ+πc1(𝒢)K_{X}+D\sim-\xi+\pi^{*}c_{1}(\omega_{Y}\otimes\det\mathcal{E}\otimes\mathcal{F})=-\xi+\pi^{*}c_{1}(\mathcal{G})
for𝒢:=ωYdet\text{for}\qquad\mathcal{G}:=\omega_{Y}\otimes\det\mathcal{E}\otimes\mathcal{F}

Then

(KD)3\displaystyle(-K_{D})^{3} =(KXD)3D\displaystyle=(-K_{X}-D)^{3}\cdot D
=(ξπc1(𝒢))3(2ξ+πc1())\displaystyle=(\xi-\pi^{*}c_{1}(\mathcal{G}))^{3}\cdot(2\xi+\pi^{*}c_{1}(\mathcal{F}))
(7.1.4) =(ξ33πc1(𝒢)ξ2+3πc1(𝒢)2ξ(πc1(𝒢))3)(2ξ+πc1())\displaystyle=(\xi^{3}-3\pi^{*}c_{1}(\mathcal{G})\cdot\xi^{2}+3\pi^{*}c_{1}(\mathcal{G})^{2}\cdot\xi-(\pi^{*}c_{1}(\mathcal{G}))^{3})\cdot(2\xi+\pi^{*}c_{1}(\mathcal{F}))
=(ξ33πc1(𝒢)ξ2+3πc1(𝒢)2ξ)(2ξ+πc1())\displaystyle=(\xi^{3}-3\pi^{*}c_{1}(\mathcal{G})\cdot\xi^{2}+3\pi^{*}c_{1}(\mathcal{G})^{2}\cdot\xi)\cdot(2\xi+\pi^{*}c_{1}(\mathcal{F}))
=2ξ4+(6πc1(𝒢)+πc1())ξ3+(6πc1(𝒢)23π(c1(𝒢)c1()))ξ2,\displaystyle=2\xi^{4}+(-6\pi^{*}c_{1}(\mathcal{G})+\pi^{*}c_{1}(\mathcal{F}))\cdot\xi^{3}+(6\pi^{*}c_{1}(\mathcal{G})^{2}-3\pi^{*}(c_{1}(\mathcal{G})\cdot c_{1}(\mathcal{F})))\cdot\xi^{2},

where the fourth equality holds by c1(𝒢)3=0c_{1}(\mathcal{G})^{3}=0 and the last one follows from c1(𝒢)2c1()=0c_{1}(\mathcal{G})^{2}\cdot c_{1}(\mathcal{F})=0. By [Har77, Appendix A, §3, page 429], we have an equation

πc0()ξ3πc1()ξ2+πc2()ξπc3()=0.\pi^{*}c_{0}(\mathcal{E})\cdot\xi^{3}-\pi^{*}c_{1}(\mathcal{E})\cdot\xi^{2}+\pi^{*}c_{2}(\mathcal{E})\cdot\xi-\pi^{*}c_{3}(\mathcal{E})=0.

By dimY=2\dim Y=2, we have c3()=0c_{3}(\mathcal{E})=0. It holds by c0()=1c_{0}(\mathcal{E})=1 that

(7.1.5) ξ3=πc1()ξ2πc2()ξ.\xi^{3}=\pi^{*}c_{1}(\mathcal{E})\cdot\xi^{2}-\pi^{*}c_{2}(\mathcal{E})\cdot\xi.

Moreover, we obtain

(7.1.6) ξ4\displaystyle\xi^{4} =πc1()ξ3πc2()ξ2\displaystyle=\pi^{*}c_{1}(\mathcal{E})\cdot\xi^{3}-\pi^{*}c_{2}(\mathcal{E})\cdot\xi^{2}
=πc1()(πc1()ξ2πc2()ξ)πc2()ξ2\displaystyle=\pi^{*}c_{1}(\mathcal{E})\cdot(\pi^{*}c_{1}(\mathcal{E})\cdot\xi^{2}-\pi^{*}c_{2}(\mathcal{E})\cdot\xi)-\pi^{*}c_{2}(\mathcal{E})\cdot\xi^{2}
=π(c1()2c2())ξ2,\displaystyle=\pi^{*}(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))\cdot\xi^{2},

where we used c1()c2()=0c_{1}(\mathcal{E})\cdot c_{2}(\mathcal{E})=0, which is guaranteed by dimY=2\dim Y=2. By (7), (7.1.5) and (7.1.6),

(KD)3\displaystyle(-K_{D})^{3} =2ξ4+(6πc1(𝒢)+πc1())ξ3+(6πc1(𝒢)23π(c1(𝒢)c1())ξ2\displaystyle=2\xi^{4}+(-6\pi^{*}c_{1}(\mathcal{G})+\pi^{*}c_{1}(\mathcal{F}))\cdot\xi^{3}+(6\pi^{*}c_{1}(\mathcal{G})^{2}-3\pi^{*}(c_{1}(\mathcal{G})\cdot c_{1}(\mathcal{F}))\cdot\xi^{2}
=2π(c1()2c2())ξ2\displaystyle=2\pi^{*}(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))\cdot\xi^{2}
+(6πc1(𝒢)+πc1())(πc1()ξ2πc2()ξ)\displaystyle+(-6\pi^{*}c_{1}(\mathcal{G})+\pi^{*}c_{1}(\mathcal{F}))\cdot(\pi^{*}c_{1}(\mathcal{E})\cdot\xi^{2}-\pi^{*}c_{2}(\mathcal{E})\cdot\xi)
(7.1.7) +(6πc1(𝒢)23π(c1(𝒢)c1())ξ2\displaystyle+(6\pi^{*}c_{1}(\mathcal{G})^{2}-3\pi^{*}(c_{1}(\mathcal{G})\cdot c_{1}(\mathcal{F}))\cdot\xi^{2}
=2π(c1()2c2())ξ2\displaystyle=2\pi^{*}(c_{1}(\mathcal{E})^{2}-c_{2}(\mathcal{E}))\cdot\xi^{2}
+(6πc1(𝒢)+πc1())πc1()ξ2\displaystyle+(-6\pi^{*}c_{1}(\mathcal{G})+\pi^{*}c_{1}(\mathcal{F}))\cdot\pi^{*}c_{1}(\mathcal{E})\cdot\xi^{2}
+(6πc1(𝒢)23π(c1(𝒢)c1())ξ2\displaystyle+(6\pi^{*}c_{1}(\mathcal{G})^{2}-3\pi^{*}(c_{1}(\mathcal{G})\cdot c_{1}(\mathcal{F}))\cdot\xi^{2}

where the last equality holds by c1(𝒢)c2()=c1()c2()=0c_{1}(\mathcal{G})\cdot c_{2}(\mathcal{E})=c_{1}(\mathcal{F})\cdot c_{2}(\mathcal{E})=0.

Since we have 𝒪X(1)|F=𝒪2(1)\mathcal{O}_{X}(1)|_{F}=\mathcal{O}_{\mathbb{P}^{2}}(1) for every fibre F2F\simeq\mathbb{P}^{2} of π\pi, we have ξ2F=(ξ|F)2=c1(𝒪2(1))2=1\xi^{2}\cdot F=(\xi|_{F})^{2}=c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(1))^{2}=1, which implies πZξ2=degZ\pi^{*}Z\cdot\xi^{2}=\deg Z for any 0-cycle ZA2(Y)Z\in\mathrm{A}^{2}(Y). It follows from (7) that

(KD)3\displaystyle(-K_{D})^{3} =2c1()22c2()+(6c1(𝒢)+c1())c1()+(6c1(𝒢)23c1(𝒢)c1())\displaystyle=2c_{1}(\mathcal{E})^{2}-2c_{2}(\mathcal{E})+(-6c_{1}(\mathcal{G})+c_{1}(\mathcal{F}))\cdot c_{1}(\mathcal{E})+(6c_{1}(\mathcal{G})^{2}-3c_{1}(\mathcal{G})\cdot c_{1}(\mathcal{F}))

By c1(𝒢)=KY+c1()+c1()c_{1}(\mathcal{G})=K_{Y}+c_{1}(\mathcal{E})+c_{1}(\mathcal{F}), we obtain

6c1(𝒢)c1()+6c1(𝒢)23c1(𝒢)c1()\displaystyle-6c_{1}(\mathcal{G})\cdot c_{1}(\mathcal{E})+6c_{1}(\mathcal{G})^{2}-3c_{1}(\mathcal{G})\cdot c_{1}(\mathcal{F})
=\displaystyle= c1(𝒢)(6c1()+6c1(𝒢)3c1())\displaystyle c_{1}(\mathcal{G})\cdot(-6c_{1}(\mathcal{E})+6c_{1}(\mathcal{G})-3c_{1}(\mathcal{F}))
=\displaystyle= (KY+c1()+c1())(6KY+3c1()).\displaystyle(K_{Y}+c_{1}(\mathcal{E})+c_{1}(\mathcal{F}))\cdot(6K_{Y}+3c_{1}(\mathcal{F})).

Therefore,

(KD)3\displaystyle(-K_{D})^{3} =2c1()22c2()+c1()c1()\displaystyle=2c_{1}(\mathcal{E})^{2}-2c_{2}(\mathcal{E})+c_{1}(\mathcal{F})\cdot c_{1}(\mathcal{E})
+(KY+c1()+c1())(6KY+3c1())\displaystyle+(K_{Y}+c_{1}(\mathcal{E})+c_{1}(\mathcal{F}))\cdot(6K_{Y}+3c_{1}(\mathcal{F}))
=2c1()22c2()+4c1()c1()\displaystyle=2c_{1}(\mathcal{E})^{2}-2c_{2}(\mathcal{E})+4c_{1}(\mathcal{F})\cdot c_{1}(\mathcal{E})
+6KY2+9KYc1()+6KYc1()+3c1()2.\displaystyle+6K_{Y}^{2}+9K_{Y}\cdot c_{1}(\mathcal{F})+6K_{Y}\cdot c_{1}(\mathcal{E})+3c_{1}(\mathcal{F})^{2}.

Thus (5) holds. ∎

Proposition 7.2.

Let YY be a smooth projective threefold and let CC be a smooth curve on YY. Let f:XYf\colon X\to Y be the blowup along CC. Set D:=Ex(f)D:={\operatorname{Ex}}(f). Then the following hold:

  1. (1)

    KXfKY+DK_{X}\sim f^{*}K_{Y}+D.

  2. (2)

    The induced morphism f|D:DCf|_{D}\colon D\to C coincides with the 1\mathbb{P}^{1}-bundle (𝒩C/Y)\mathbb{P}(\mathcal{N}^{*}_{C/Y}) over CC, and 𝒪X(D)|D𝒪(𝒩C/Y)(1)\mathcal{O}_{X}(-D)|_{D}\simeq\mathcal{O}_{\mathbb{P}(\mathcal{N}^{*}_{C/Y})}(1).

  3. (3)

    D3=degC(𝒩C/Y)D^{3}=\deg_{C}(\mathcal{N}^{*}_{C/Y}).

  4. (4)

    (KX)2D=KYC+22g(C)(-K_{X})^{2}\cdot D=-K_{Y}\cdot C+2-2g(C).

Proof.

The assertions (1) and (2) follow from [Har77, Ch. II, Exercise 8.5(b)] and and [Har77, II, Theorem 8.24(b),(c)], respectively. Let us show (3). By (2), we have D(𝒩C/Y)D\simeq\mathbb{P}(\mathcal{N}^{*}_{C/Y}). We then get

D3\displaystyle D^{3} =(𝒪X(D)|D)(𝒪X(D)|D)\displaystyle=(\mathcal{O}_{X}(-D)|_{D})\cdot(\mathcal{O}_{X}(-D)|_{D})
=c1(𝒪D(1))2\displaystyle=c_{1}(\mathcal{O}_{D}(1))^{2}
=degC(𝒩C/Y),\displaystyle=\deg_{C}(\mathcal{N}^{*}_{C/Y}),

where the second equality holds by (2) and the last one follows from Proposition 7.1(4). Thus (3) holds.

Let us show (4). Set π:=f|D:D=(𝒩C/Y)C\pi:=f|_{D}\colon D=\mathbb{P}(\mathcal{N}^{*}_{C/Y})\to C and :=(ωCdet𝒩C/Y)1\mathcal{L}:=(\omega_{C}\otimes\det\mathcal{N}^{*}_{C/Y})^{-1}. Fix a divisor ξ\xi on DD such that 𝒪D(ξ)=𝒪D(1)\mathcal{O}_{D}(\xi)=\mathcal{O}_{D}(1). Then

(KX)2D=(𝒪X(KX)|D)2.(-K_{X})^{2}\cdot D=(\mathcal{O}_{X}(-K_{X})|_{D})^{2}.

It holds that

𝒪X(KX)|D\displaystyle\mathcal{O}_{X}(-K_{X})|_{D} 𝒪D(KD)𝒪X(D)|D\displaystyle\simeq\mathcal{O}_{D}(-K_{D})\otimes\mathcal{O}_{X}(D)|_{D}
𝒪D(2)π(ωCdet𝒩C/Y)1𝒪D(1)\displaystyle\simeq\mathcal{O}_{D}(2)\otimes\pi^{*}(\omega_{C}\otimes\det\mathcal{N}^{*}_{C/Y})^{-1}\otimes\mathcal{O}_{D}(-1)
𝒪D(1)π,\displaystyle\simeq\mathcal{O}_{D}(1)\otimes\pi^{*}\mathcal{L},

where the first isomorphism holds by the adjunction formula and the second one follows from (2) and Proposition 7.1(2). Hence

(KX)2D\displaystyle(-K_{X})^{2}\cdot D =(ξ+c1(π))2\displaystyle=(\xi+c_{1}(\pi^{*}\mathcal{L}))^{2}
=ξ2+2ξπc1()+πc1()2\displaystyle=\xi^{2}+2\xi\cdot\pi^{*}c_{1}(\mathcal{L})+\pi^{*}c_{1}(\mathcal{L})^{2}
=ξ2+2πc1()ξ,\displaystyle=\xi^{2}+2\pi^{*}c_{1}(\mathcal{L})\cdot\xi,

where c1()2=0c_{1}(\mathcal{L})^{2}=0 holds by dimC=1\dim C=1. By [Har77, Appendix A, §3, page 429], we have ξ2=πc1(𝒩C/Y)ξ\xi^{2}=\pi^{*}c_{1}(\mathcal{N}^{*}_{C/Y})\cdot\xi. Therefore, we get

(KX)2D\displaystyle(-K_{X})^{2}\cdot D =πc1(𝒩C/Y)ξ+2πc1()ξ\displaystyle=\pi^{*}c_{1}(\mathcal{N}^{*}_{C/Y})\cdot\xi+2\pi^{*}c_{1}(\mathcal{L})\cdot\xi
=π(c1(𝒩C/Y)+2c1())ξ\displaystyle=\pi^{*}(c_{1}(\mathcal{N}^{*}_{C/Y})+2c_{1}(\mathcal{L}))\cdot\xi
=degC(c1(𝒩C/Y)+2c1())\displaystyle=\deg_{C}(c_{1}(\mathcal{N}^{*}_{C/Y})+2c_{1}(\mathcal{L}))
=degC(c1(𝒩C/Y)2KC)\displaystyle=\deg_{C}(-c_{1}(\mathcal{N}^{*}_{C/Y})-2K_{C})
=degC(c1(ωY|C))KC)\displaystyle=\deg_{C}(-c_{1}(\omega_{Y}|_{C}))-K_{C})
=KYC+22g(C),\displaystyle=-K_{Y}\cdot C+2-2g(C),

where the fourth equality holds by =(ωCdet𝒩C/Y)1\mathcal{L}=(\omega_{C}\otimes\det\mathcal{N}^{*}_{C/Y})^{-1} and the fifth equality follows from ωCωY|C(det𝒩C/Y)\omega_{C}\simeq\omega_{Y}|_{C}\otimes(\det\mathcal{N}_{C/Y}) [Har77, Ch. II, Proposition 8.20]. ∎

8. Appendix: Description of split double covers

Definition 8.1.

Let YY be a normal variety and let \mathcal{L} be an invertible sheaf. Set

𝕍:=SpecYi0i,\mathbb{V}:={\operatorname{Spec}}_{Y}\bigoplus_{i\geq 0}\mathcal{L}^{\otimes-i},

where i0i\bigoplus_{i\geq 0}\mathcal{L}^{\otimes-i} denotes the natural graded 𝒪Y\mathcal{O}_{Y}-algebra. Let π:𝕍Y\pi:\mathbb{V}\to Y be the projection. Take the tautological section zH0(𝕍,π)z\in H^{0}(\mathbb{V},\pi^{*}\mathcal{L}), which corresponds to the element (0,1,0,0,)i0H0(Y,1i)(0,1,0,0,...)\in\bigoplus_{i\geq 0}H^{0}(Y,\mathcal{L}^{1-i}) via

H0(𝕍,π)H0(Y,ππ)i0H0(Y,1i).H^{0}(\mathbb{V},\pi^{*}\mathcal{L})\simeq H^{0}(Y,\pi_{*}\pi^{*}\mathcal{L})\simeq\bigoplus_{i\geq 0}H^{0}(Y,\mathcal{L}^{1-i}).
  1. (1)

    For sH0(Y,2)s\in H^{0}(Y,\mathcal{L}^{\otimes 2}), the closed subscheme

    X:={z2π(s)=0}𝕍X:=\{z^{2}-\pi^{*}(s)=0\}\subset\mathbb{V}

    and the induced morphism XYX\to Y are called the simple μ2\mu_{2}-cover associated to (,s)(\mathcal{L},s).

  2. (2)

    Assume p=2p=2. For sH0(X,)s\in H^{0}(X,\mathcal{L}) and tH0(X,2)t\in H^{0}(X,\mathcal{L}^{\otimes 2}), the closed subscheme

    X:={z2π(s)z+π(t)=0}𝕍X:=\{z^{2}-\pi^{*}(s)\cdot z+\pi^{*}(t)=0\}\subset\mathbb{V}

    and the induced morphism XYX\to Y are called the split α,s\alpha_{\mathcal{L},s}-torsor associated to (,s,t)(\mathcal{L},s,t). Note that the split α,0\alpha_{\mathcal{L},0}-torsor associated to (,0,t)(\mathcal{L},0,t) coincides with the simple μ2\mu_{2}-cover associated to (,t)(\mathcal{L},-t).

Lemma 8.2.

Assume that p2p\neq 2. Let f:XYf:X\to Y be a double cover of smooth projective varieties. Set :=(f𝒪X/𝒪Y)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1}, which is an invertible sheaf (Remark 2.2). Then ff is a simple μ2\mu_{2}-cover associated to (,s)(\mathcal{L},s) for some sH0(Y,2)s\in H^{0}(Y,\mathcal{L}^{\otimes 2})

Proof.

See [CD89, Page 10]. ∎

Lemma 8.3.

Assume p=2p=2. Let f:XYf:X\to Y be a double cover of smooth projective varieties. Set :=(f𝒪X/𝒪Y)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1}, which is an invertible sheaf (Remark 2.2). Then ff is a spit double cover if and only if ff is a split α,s\alpha_{\mathcal{L},s}-torsor associated to some (,s,t)(\mathcal{L},s,t).

Proof.

If ff is the split α,s\alpha_{\mathcal{L},s}-torsor associated to some (,s,t)(\mathcal{L},s,t), then it is clear that ff is a split double cover. Conversely, assume that ff is a split double cover. By [CD89, Page 10-11], ff is given by a splittable admissible triple in the sense of [CD89, Page 11], which is nothing but a split α,s\alpha_{\mathcal{L},s}-torsor associated to some (,s,t)(\mathcal{L},s,t). ∎

Lemma 8.4.

Let f:XYf:X\to Y be a split double cover of smooth projective varieties. Set :=(f𝒪X/𝒪Y)1\mathcal{L}:=(f_{*}\mathcal{O}_{X}/\mathcal{O}_{Y})^{-1}, which is an invertible sheaf (Remark 2.2). Let π:YY\pi:Y\to Y^{\prime} be a morphism of smooth projective varieties such that g𝒪Y=𝒪Yg_{*}\mathcal{O}_{Y}=\mathcal{O}_{Y^{\prime}} and g\mathcal{L}\simeq g^{*}\mathcal{L}^{\prime} for some invertible sheaf \mathcal{L}^{\prime} on YY^{\prime}. Then there exist a split double cover f:XYf^{\prime}:X^{\prime}\to Y^{\prime} and a morphism πX:XX\pi_{X}:X\to X^{\prime} such that the following diagram is cartesian:

X{X}X{X^{\prime}}Y{Y}Y.{Y^{\prime}.}πX\scriptstyle{\pi_{X}}f\scriptstyle{f}f\scriptstyle{f^{\prime}}π\scriptstyle{\pi}

In particular, if Y=Y×kY′′Y=Y^{\prime}\times_{k}Y^{\prime\prime} and π:YY\pi:Y\to Y^{\prime} is the first projection in addition to the above assumptions, then XX×kY′′X\simeq X^{\prime}\times_{k}Y^{\prime\prime} and πX:XX\pi_{X}:X\to X^{\prime} coincides with the first projection.

Proof.

Assume p=2p=2. By Lemma 8.3, there exist sH0(Y,)s\in H^{0}(Y,\mathcal{L}) and tH0(Y,2)t\in H^{0}(Y,\mathcal{L}^{\otimes 2}) such that ff is the split α,s\alpha_{\mathcal{L},s}-torsor associated to (,s,t)(\mathcal{L},s,t). By π𝒪Y=𝒪Y\pi_{*}\mathcal{O}_{Y}=\mathcal{O}_{Y^{\prime}} and g\mathcal{L}\simeq g^{*}\mathcal{L}^{\prime}, we have the following corresponding elements:

  • H0(Y,)H0(Y,),ssH^{0}(Y,\mathcal{L})\simeq H^{0}(Y^{\prime},\mathcal{L}^{\prime}),\qquad s\mapsto s^{\prime}.

  • H0(Y,2)H0(Y,2),ttH^{0}(Y,\mathcal{L}^{\otimes 2})\simeq H^{0}(Y^{\prime},\mathcal{L}^{\prime\otimes 2}),\qquad t\mapsto t^{\prime}.

Set f:XYf^{\prime}:X^{\prime}\to Y^{\prime} to be the split α,s\alpha_{\mathcal{L}^{\prime},s^{\prime}} torsor associated to (,s,t)(\mathcal{L}^{\prime},s^{\prime},t^{\prime}).

Fix an affine open cover Y=iIYiY^{\prime}=\bigcup_{i\in I}Y^{\prime}_{i} such that |Yi𝒪Yi\mathcal{L}^{\prime}|_{Y^{\prime}_{i}}\simeq\mathcal{O}_{Y^{\prime}_{i}} for every iIi\in I. Then XX^{\prime} is given as follows:

X=iISpec(𝒪Yi(Yi)[zi]/(zi2+sizi+ti)),X^{\prime}=\bigcup_{i\in I}{\operatorname{Spec}}\,(\mathcal{O}_{Y^{\prime}_{i}}(Y^{\prime}_{i})[z^{\prime}_{i}]/(z^{\prime 2}_{i}+s^{\prime}_{i}z^{\prime}_{i}+t^{\prime}_{i})),

where si:=s|Yis^{\prime}_{i}:=s^{\prime}|_{Y^{\prime}_{i}} and ti:=t|Yit^{\prime}_{i}:=t^{\prime}|_{Y^{\prime}_{i}}. By taking an affine open cover of YY which refines Y=iIπ1(Yi)Y=\bigcup_{i\in I}\pi^{-1}(Y^{\prime}_{i}), we can directly check that the induced morphism XX×YYX\to X^{\prime}\times_{Y^{\prime}}Y is an isomorphism (more explicitly, if Y~\widetilde{Y} is an affine open subset contained in π1(Yi)\pi^{-1}(Y^{\prime}_{i}) for some ii, then both XX and X×YYX^{\prime}\times_{Y^{\prime}}Y are given by 𝒪Y~(Y~)[z]/(z2+(s|Y~)z+(t|Y~)\mathcal{O}_{\widetilde{Y}}(\widetilde{Y})[z]/(z^{2}+(s|_{\widetilde{Y}})z+(t|_{\widetilde{Y}})). If p2p\neq 2, the same argument works by using Lemma 8.2 instead of Lemma 8.3. ∎

9. Classification table of Fano threefolds with ρ=2\rho=2

Let XX be a Fano threefold with ρ(X)=2\rho(X)=2. Then one of No. 2-1, …, No. 2-36 holds. In addition to (1)–(6) of Theorem 1.2, we use the follwoing notation and terminologies:

  1. (7)

    If an extremal ray is of type CC, then degΔ\deg\Delta denotes the degree of the discriminant bundle Δ\Delta, which is an invertible sheaf on 2\mathbb{P}^{2}. When the conic bundle is generically smooth, then Δ\Delta coincides with the invertible sheaf associated with the discriminant divisor.

  2. (8)

    If an extremal ray is of type DD, then XtX_{t} denotes a fibre over a closed point t1t\in\mathbb{P}^{1}. In particular, (KX)2Xt(-K_{X})^{2}\cdot X_{t} coincides with (KXK)2(-K_{X_{K}})^{2} for the generic fibre XKX_{K}.

Table 5. Fano threefolds with ρ(X)=2\rho(X)=2
No. (KX)3(-K_{X})^{3} descriptions and extremal rays
2-1 44 D1:(KX)2Xt=1D_{1}:(-K_{X})^{2}\cdot X_{t}=1
E1:E_{1}: blowup of V1V_{1} along an elliptic curve of degree 11 which is a complete intersection of two members of |12KV1||-\frac{1}{2}K_{V_{1}}|
2-2 66 XX is a split double cover of 2×1\mathbb{P}^{2}\times\mathbb{P}^{1} with 𝒪(2,1)\mathcal{L}\simeq\mathcal{O}(2,1)
C1:degΔ=8C_{1}:\deg\Delta=8, X2:12×1pr12X\xrightarrow{{\rm 2:1}}\mathbb{P}^{2}\times\mathbb{P}^{1}\xrightarrow{{\rm pr}_{1}}\mathbb{P}^{2}
D1:(KX)2Xt=2D_{1}:(-K_{X})^{2}\cdot X_{t}=2
2-3 88 D1:(KX)2Xt=2D_{1}:(-K_{X})^{2}\cdot X_{t}=2
E1:E_{1}: is a blowup of V2V_{2} along an elliptic curve of degree 22 which is a complete intersection of two members of |12KV2||-\frac{1}{2}K_{V_{2}}|
2-4 1010 D1:(KX)2Xt=3D_{1}:(-K_{X})^{2}\cdot X_{t}=3
E1:E_{1}: blowup of 3\mathbb{P}^{3} along a curve of genus 1010 and degree 99 which is a complete intersection of two cubic surfaces
2-5 1212 D1:(KX)2Xt=3D_{1}:(-K_{X})^{2}\cdot X_{t}=3
E1:E_{1}: blowup of V3V_{3} along an elliptic curve of degree 33 which is a complete intersection of two members of |12KV3||-\frac{1}{2}K_{V_{3}}|
2-6 1212 a smooth divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (2,2)(2,2), or a split double cover of WW with 2ωW1\mathcal{L}^{\otimes 2}\simeq\omega_{W}^{-1}
C1:degΔ=6C_{1}:\deg\Delta=6, X2×2pr12X\to\mathbb{P}^{2}\times\mathbb{P}^{2}\xrightarrow{{\rm pr_{1}}}\mathbb{P}^{2}
C1:degΔ=6C_{1}:\deg\Delta=6, X2×2pr22X\to\mathbb{P}^{2}\times\mathbb{P}^{2}\xrightarrow{{\rm pr_{2}}}\mathbb{P}^{2}
2-7 1414 D1:(KX)2Xt=4D_{1}:(-K_{X})^{2}\cdot X_{t}=4
E1:E_{1}: blowup of QQ along a curve of genus 55 and degree 88 which is a complete intersection of two members of |𝒪Q(2)||\mathcal{O}_{Q}(2)|
2-8 1414 a split double cover of V7V_{7} with 2ωV71\mathcal{L}^{\otimes 2}\simeq\omega_{V_{7}}^{-1}
C1:degΔ=6C_{1}:\deg\Delta=6, X2:1V7=2(𝒪2𝒪2(1))pr2X\xrightarrow{{\rm 2:1}}V_{7}=\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1))\xrightarrow{{\rm pr}}\mathbb{P}^{2}
E3orE4E_{3}\,{\rm or}\,E_{4}
2-9 1616 C1:degΔ=5C_{1}:\deg\Delta=5
E1:E_{1}: blowup of 3\mathbb{P}^{3} along a curve of genus 55 and degree 77
2-10 1616 D1:D_{1}: (KX)2Xt=4(-K_{X})^{2}\cdot X_{t}=4
E1:E_{1}: blowup of V4V_{4} along an elliptic curve of degree 44 which is a complete intersection of two members of |12KV4||-\frac{1}{2}K_{V_{4}}|
2-11 1818 C1:C_{1}: degΔ=5\deg\Delta=5
E1:E_{1}: blowup of V3V_{3} along a line
2-12 2020 E1:E_{1}: blowup of 3\mathbb{P}^{3} along a curve of genus 33 and degree 66
E1:E_{1}: blowup of 3\mathbb{P}^{3} along a curve of genus 33 and degree 66
2-13 2020 C1:C_{1}: degΔ=4\deg\Delta=4
E1:E_{1}: blowup of QQ along a curve of genus 22 and degree 66
2-14 2020 D1:D_{1}: (KX)2Xt=5(-K_{X})^{2}\cdot X_{t}=5
E1:E_{1}: blowup of V5V_{5} along an elliptic curve of degree 55 which is a complete intersection of two members of |12KV5||-\frac{1}{2}K_{V_{5}}|
2-15 2222 E1:E_{1}: blowup of 3\mathbb{P}^{3} along a curve of genus 33 and degree 66
E3orE4E_{3}\,{\rm or}\,E_{4}
2-16 2222 C1:degΔ=4C_{1}:\deg\Delta=4
E1:E_{1}: blowup of V4V_{4} along a conic
2-17 2424 E1:E_{1}: blowup of 3\mathbb{P}^{3} along an elliptic curve of degree 55
E1:E_{1}: blowup of QQ along an elliptic curve of degree 55
2-18 2424 a split double cover of 2×1\mathbb{P}^{2}\times\mathbb{P}^{1} with 𝒪2×1(1,1)\mathcal{L}\simeq\mathcal{O}_{\mathbb{P}^{2}\times\mathbb{P}^{1}}(1,1)
C1:degΔ=4C_{1}:\deg\Delta=4, X2:12×1pr12X\xrightarrow{{\rm 2:1}}\mathbb{P}^{2}\times\mathbb{P}^{1}\xrightarrow{{\operatorname{pr}}_{1}}\mathbb{P}^{2}
D2:(KX)2Xt=8D_{2}:(-K_{X})^{2}\cdot X_{t}=8
2-19 2626 E1:E_{1}: blowup of 3\mathbb{P}^{3} along a curve of genus 22 and degree 55
E1:E_{1}: blowup of V4V_{4} along a line
2-20 2626 C1:degΔ=3C_{1}:\deg\Delta=3
E1:E_{1}: blowup of V5V_{5} along a cubic rational curve
2-21 2828 E1:E_{1}: blowup of QQ along a rational curve of degree 44
2828 E1:E_{1}: blowup of QQ along a rational curve of degree 44
2-22 3030 E1:E_{1}: blowup of 3\mathbb{P}^{3} along a rational curve of degree 44
E1:E_{1}: blowup of V5V_{5} along a conic
2-23 3030 E1:E_{1}: blowup of QQ along an elliptic curve of degree 44
E3orE4E_{3}\,{\rm or}\,E_{4}
2-24 3030 a divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,2)(1,2)
C1:degΔ=3C_{1}:\deg\Delta=3
C2C_{2}
2-25 3232 D2:(KX)2Xt=8D_{2}:(-K_{X})^{2}\cdot X_{t}=8
E1:E_{1}: blowup of 3\mathbb{P}^{3} along an elliptic curve of degree 44 which is a complete intersection of two quadric surfaces
2-26 3434 E1:E_{1}: blowup of QQ along a cubic rational curve
E1:E_{1}: blowup of V5V_{5} along a line
2-27 3838 C2C_{2}
E1:E_{1}: blowup of 3\mathbb{P}^{3} along a cubic rational curve
2-28 4040 E1:E_{1}: blowup of 3\mathbb{P}^{3} along an elliptic curve of degree 33
E3orE4E_{3}\,{\rm or}\,E_{4}
2-29 4040 D2:(KX)2Xt=8D_{2}:(-K_{X})^{2}\cdot X_{t}=8
E1:E_{1}: blowup of QQ along a conic which is a complete intersection of two members of |𝒪Q(1)||\mathcal{O}_{Q}(1)|
2-30 4646 E1:E_{1}: blowup of 3\mathbb{P}^{3} along a conic
E2:E_{2}: blowup of QQ at a point
2-31 4646 C2C_{2}
E1:E_{1}: blowup of QQ along a line
2-32 4848 WW, a divisor on 2×2\mathbb{P}^{2}\times\mathbb{P}^{2} of bidegree (1,1)(1,1)
C2:W2×2pr12C_{2}:W\hookrightarrow\mathbb{P}^{2}\times\mathbb{P}^{2}\xrightarrow{{\operatorname{pr}}_{1}}\mathbb{P}^{2}
C2:W2×2pr12C_{2}:W\hookrightarrow\mathbb{P}^{2}\times\mathbb{P}^{2}\xrightarrow{{\operatorname{pr}}_{1}}\mathbb{P}^{2}
2-33 5454 D3:(KX)2Xt=9D_{3}:(-K_{X})^{2}\cdot X_{t}=9
E1:E_{1}: blowup of 3\mathbb{P}^{3} along a line
2-34 5454 2×1\mathbb{P}^{2}\times\mathbb{P}^{1}
C2:C_{2}: the projection 2×12\mathbb{P}^{2}\times\mathbb{P}^{1}\to\mathbb{P}^{2}
D3:D_{3}: the projection 2×11\mathbb{P}^{2}\times\mathbb{P}^{1}\to\mathbb{P}^{1}
2-35 5656 V7V_{7}, i.e., 2(𝒪2𝒪2(1))\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1))
C2:C_{2}: the projection 2(𝒪2𝒪2(1))2\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(1))\to\mathbb{P}^{2}
E2:E_{2}: blowup of 3\mathbb{P}^{3} at a point
2-36 6262 2(𝒪2𝒪2(2))\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))
C2:C_{2}: the projection 2(𝒪2𝒪2(2))2\mathbb{P}_{\mathbb{P}^{2}}(\mathcal{O}_{\mathbb{P}^{2}}\oplus\mathcal{O}_{\mathbb{P}^{2}}(2))\to\mathbb{P}^{2}
E5:E_{5}: blowup at the singular point of the cone over the Veronese surface
Proof.

(of Table LABEL:table-pic2) We use Notation 5.1. Almost all the parts follow from Section 5. It is enough to compute degΔ\deg\Delta (type CC) and (KX)2Xt(-K_{X})^{2}\cdot X_{t} (type DD).

If f1:X2f_{1}:X\to\mathbb{P}^{2} is of type CC, then we have H13=0H_{1}^{3}=0 and H12H2=(2μ11)H2=2μ1H_{1}^{2}\cdot H_{2}=(\frac{2}{\mu_{1}}\ell_{1})\cdot H_{2}=\frac{2}{\mu_{1}} (Lemma 5.3, Proposition 5.9), which imply the following:

degΔ=(a)12(KX)2H1=(b)12(μ2H1+μ1H2)2H1=124μ2μ12H1H22,\deg\Delta\overset{{\rm(a)}}{=}12-(-K_{X})^{2}\cdot H_{1}\overset{{\rm(b)}}{=}12-(\mu_{2}H_{1}+\mu_{1}H_{2})^{2}\cdot H_{1}=12-4\mu_{2}-\mu_{1}^{2}H_{1}\cdot H_{2}^{2},

where (a) and (b) follow from Proposition 3.16 and Proposition 5.9, respectively. If f1:X1f_{1}:X\to\mathbb{P}^{1} is of type DD, then we have H13=H12H2=0H_{1}^{3}=H_{1}^{2}\cdot H_{2}=0, which imply the following:

(KX)2Xt=(KX)2H1=(μ2H1+μ1H2)2H1=μ12H1H22.(-K_{X})^{2}\cdot X_{t}=(-K_{X})^{2}\cdot H_{1}=(\mu_{2}H_{1}+\mu_{1}H_{2})^{2}\cdot H_{1}=\mu_{1}^{2}H_{1}\cdot H_{2}^{2}.

Therefore, it is enough to compute one of (KX)2H1(-K_{X})^{2}\cdot H_{1} and H1H22H_{1}\cdot H_{2}^{2}. If R2R_{2} is type DD, then we have H220H_{2}^{2}\equiv 0, and hence H1H22=0H_{1}\cdot H_{2}^{2}=0. If R2R_{2} is of type CC, then we get H1H22=2μ2H_{1}\cdot H_{2}^{2}=\frac{2}{\mu_{2}}.

In what follows, we assume that XX is a Fano threefold of No. 2-8 in order to explain how to compute degΔ\deg\Delta and (KX)2H1(-K_{X})^{2}\cdot H_{1} for the case when R1R_{1} is of type CC and R2R_{2} is of type EE. It holds that KXH1+H2-K_{X}\sim H_{1}+H_{2} (Proposition 5.9). For D:=Ex(f2)D:={\operatorname{Ex}}(f_{2}), we have KXf2KY+DK_{X}\sim f_{2}^{*}K_{Y}+D and ((KX)2D,(KX)D2,D3)=(2,2,2)((-K_{X})^{2}\cdot D,(-K_{X})\cdot D^{2},D^{3})=(2,-2,2) [TanII, Proposition 3.22]. By Lemma 5.25(1), we have D1=1D\cdot\ell_{1}=1 (note that ζ21\zeta\equiv 2\ell_{1} for a fibre ζ\zeta of f1:X2f_{1}:X\to\mathbb{P}^{2}). Hence we obtain f2KY1=KX1+D1=2-f_{2}^{*}K_{Y}\cdot\ell_{1}=-K_{X}\cdot\ell_{1}+D\cdot\ell_{1}=2. By H21=1H_{2}\cdot\ell_{1}=1 (Proposition 5.9), we get f2KY2H2-f_{2}^{*}K_{Y}\equiv 2H_{2}. Then

(KX)2H1=(KX)2(KXH2)=(KX)2(KXKX+D2)(-K_{X})^{2}\cdot H_{1}=(-K_{X})^{2}\cdot(-K_{X}-H_{2})=(-K_{X})^{2}\cdot\left(-K_{X}-\frac{-K_{X}+D}{2}\right)
=(KX)3(KX)2D2=1422=6.=\frac{(-K_{X})^{3}-(-K_{X})^{2}\cdot D}{2}=\frac{14-2}{2}=6.

Thus degΔ=12(KX)2H1=126=6\deg\Delta=12-(-K_{X})^{2}\cdot H_{1}=12-6=6. ∎

We close this paper with the following proposition. Although this result is not used in this paper, we will need it in the classification of Fano threefolds with ρ3\rho\geq 3.

Proposition 9.1.
  1. (1)

    Let XX be a Fano threefold of No. 2-15. Let f:X3f:X\to\mathbb{P}^{3} be a blowup along a smooth curve BB of degree 66. Then BB is contained in a (possibly singular) quadric surface SS on 3\mathbb{P}^{3}.

  2. (2)

    Let XX be a Fano threefold of No. 2-23. Let f:XQf:X\to Q be a blowup along a smooth curve BB of degree 44. Then BB is contained in a (possibly singular) quadric surface SS on QQ, i.e., SS is a hyperplane section of Q4Q\subset\mathbb{P}^{4}.

Proof.

We prove (1) and (2) simultaneously. If XX is 2-15 (resp. 2-23), then we set Y:=3Y:=\mathbb{P}^{3} (resp. Y:=QY:=Q) and s:=4s:=4 (resp. s:=3s:=3). Then we have a blowup f:XYf:X\to Y and YY is a Fano threefold of index ss. We have the contraction g:XZg:X\to Z of the other extremal ray, which is of type E3E_{3} or E4E_{4}. In particular, T:=Ex(g)T:={\operatorname{Ex}}(g) is a (possibly singular) quadric surface. Set S:=f(T)S:=f(T). It is enough to show that BSB\subset S and the induced composite morphism

fT:TX𝑓Sf_{T}:T\hookrightarrow X\xrightarrow{f}S

is an isomorphism. Set HfH_{f} (resp. HgH_{g}) to be the pullback on XX of the ample generator of YY (resp. ZZ). It holds that KXHf+Hg-K_{X}\sim H_{f}+H_{g} (Proposition 5.9). Let rr be the positive integer satisfying gKYrHg-g^{*}K_{Y}\sim rH_{g}.

Step 1.

BSB\subset S.

Proof of Step 1.

Suppose BSB\not\subset S. Then T=fST=f^{*}S. Pick a curve CC on SS disjoint from SBS\cap B. Set CX:=f1(C)CC_{X}:=f^{-1}(C)\xrightarrow{\simeq}C. Since SS is ample, we have SC>0S\cdot C>0. On the other hand, 𝒪X(T)|T\mathcal{O}_{X}(-T)|_{T} is ample, and hence TCX=𝒪X(T)|TC<0T\cdot C_{X}=\mathcal{O}_{X}(T)|_{T}\cdot C<0. By T=fST=f^{*}S, we get the following contradiction:

0<SC=fSCX=TCX<0.0<S\cdot C=f^{*}S\cdot C_{X}=T\cdot C_{X}<0.

This completes the proof of Step 1. ∎

Step 2.

Tζ=r1T\cdot\zeta=r-1 for a one-dimensional fibre ζ\zeta of f:XYf:X\to Y.

Proof of Step 2.

We have that

KXHf+HgKX+Es+KX+Tr,-K_{X}\sim H_{f}+H_{g}\equiv\frac{-K_{X}+E}{s}+\frac{-K_{X}+T}{r},

which implies

(srrs)KXrE+sT.-(sr-r-s)K_{X}\sim rE+sT.

Taking the intersection with ζ\zeta, we get srrs=r+sTζsr-r-s=-r+sT\cdot\zeta, which implies Tζ=r1T\cdot\zeta=r-1. This completes the proof of Step 2. ∎

Step 3.

It holds that r=2r=2 and Tζ=1T\cdot\zeta=1.

Proof of Step 3.

By [TanII, Proposition 3.22], we get

KZ3=(gKZ)3=(KXT)3=KX33KX2T+3KXT2T3=KX332+322,K_{Z}^{3}=(g^{*}K_{Z})^{3}=(K_{X}-T)^{3}=K_{X}^{3}-3K_{X}^{2}\cdot T+3K_{X}\cdot T^{2}-T^{3}=K_{X}^{3}-3\cdot 2+3\cdot 2-2,

and hence r3Hg3=(KZ)3=(KX)3+2r^{3}H_{g}^{3}=(-K_{Z})^{3}=(-K_{X})^{3}+2. If XX is of No. 2-15, then we get r3Hg3=22+2=24r^{3}H_{g}^{3}=22+2=24. If XX is of No. 2-23, then we get r3Hg3=30+2=32r^{3}H_{g}^{3}=30+2=32. Since the inequality r2r\geq 2 holds by Step 2, we obtain r=2r=2. Again by Step 2, we get Tζ=r1=1T\cdot\zeta=r-1=1. This completes the proof of Step 3. ∎

Step 4.

fT:TSf_{T}:T\to S is an isomorphism.

Proof of Step 4.

Let us show that SS is normal. Consider an exact sequence

0𝒪X(T)𝒪X𝒪T0.0\to\mathcal{O}_{X}(-T)\to\mathcal{O}_{X}\to\mathcal{O}_{T}\to 0.

By (TKX)ζ=1+1=0(-T-K_{X})\cdot\zeta=-1+1=0 (Step 3), we get R1f𝒪X(T)=0R^{1}f_{*}\mathcal{O}_{X}(-T)=0 [Tan15, Theorem 0.5]. Then 𝒪S=f𝒪T\mathcal{O}_{S}=f_{*}\mathcal{O}_{T}, i.e., SS is normal.

Note that any curve ζ\zeta on TT is not contracted by ff, as otherwise we would get a contradiction: ζR1R2={0}\zeta\in R_{1}\cap R_{2}=\{0\}. Hence TEx(f)T\neq{\operatorname{Ex}}(f), which implies that f:TSf:T\to S is birational. Since TT is a (possibly singular) quadric surface, f|T:TSf|_{T}:T\to S is an isomorphism. This completes the proof of Step 4. ∎

Step 1 and Step 4 complete the proof of Proposition 9.1. ∎

References