We assume that all varieties are projective, normal, and defined over ℂ.

1 Introduction

Let X be a Fano variety with at most quotient singularities (a Fano orbifold).

Theorem 1.1

[37]

If dim(X)=2 and X is smooth, then

An important role in the proof of Theorem 1.1 is played by several holomorphic invariants, which are now known as α-invariants. Let us describe their algebraic counterparts.

Let D be an effective ℚ-divisor on the variety X. Then the number

is called the log canonical threshold of the divisor D (see [21, Definition 8.1]). Put

for every n∈ℕ. For small n, the number lct n (X) is usually not very hard to compute.

Example 1.2

[28]

If X is a smooth surface in ℙ3 of degree 3, then

The number lct n (X) is denoted by α n (X) in [38].

Remark 1.3

It follows from [27, Lemma 4.8] that the set

is finite (cf. [23]). Thus, there exists a divisor B∈|−nK X | such that lct n (X)=c(X,B/n)∈ℚ.

If the variety X is smooth, then it is proved by Demailly (see [6, Theorem A.3]) that

where α(X) is the α-invariant introduced by Tian in [36]. Put lct(X)=inf{lct n (X)∣n∈ℕ}.

Conjecture 1.4

[38, Question 1]

There is an n∈ℕ such that lct(X)=lct n (X).

The proof of Theorem 1.1 uses (at least implicitly) the following result.

Theorem 1.5

[10, 36]

The Fano orbifold X is Kähler–Einstein if

Note that there are many well-known obstructions to the existence of Kähler–Einstein metrics on smooth Fano manifolds and Fano orbifolds (see [14, 15, 25, 34]).

Example 1.6

If X≅ℙ(1,2,3), then X is not Kähler–Einstein (see [15, 34]).

Let us describe one more α-invariant that took its origin in [37].

Let \(\mathcal{M}\) be a linear system on the variety X. Then the number

is called the log canonical threshold of the linear system \(\mathcal{M}\) (cf. [21, Theorem 4.8]). Put

for every n∈ℕ. The number lct n,2(X) is denoted by α n,2(X) in [8] and [41]. Note that

$$ \mathrm{lct} (X )=\mathrm{inf} \bigl\{ \mathrm{lct}_{n,2} (X )\mid n\in\mathbb{N} \bigr\},$$
(1)

and it follows from [21, Theorem 4.8] that lct n (X)⩽lct n,2(X) for every n∈ℕ.

Remark 1.7

It follows from [27, Lemma 4.8] and [21, Theorem 4.8] that the set

is finite. Thus, there is a pencil \(\mathcal{B}\) in |−nK X | such that the equality \(\mathrm{lct}_{n,2}(X)=\mathrm{c}(X,\mathcal{B}/n)\). Then

if there exist at most finitely many effective ℚ-divisors D 1,D 2,…,D r on the variety X such that

and D 1 D 2⋯∼ D r K X .

The importance of the number lct n,2(X) is due to the following conjecture.

Conjecture 1.8

(cf. [8, Theorem 2], [41, Theorem 1])

Suppose that

for every n∈ℕ. Then X is Kähler–Einstein.

Note that Conjecture 1.8 is not much stronger than Theorem 1.5 by (1).

Example 1.9

Suppose that X is a smooth hypersurface in ℙm of degree m⩾3. Then

for every n∈ℕ by [2]. The equality lct n (X)=1−1/m holds ⇔ the hypersurface X contains a cone of dimension m−2 (see [2, Theorem 1.3], [2, Theorem 4.1], [13, Theorem 0.2]). Then

by Remark 1.7, [2, Remark 1.6], [2, Theorem 4.1], [2, Theorem 5.2], and [13, Theorem 0.2], because X contains at most finitely many cones by [9, Theorem 4.2]. If X is general, then

by [3, 5, 33]. Thus, if X is general, then it is Kähler–Einstein by Theorem 1.5.

The assertion of Conjecture 1.8 follows from [8, Theorem 2] and [41, Theorem 1] under an additional assumption that the Kähler–Ricci flow on X is tamed (see [8] and [41]).

Theorem 1.10

[8, 41]

If dim(X)=2, then the Kähler–Ricci flow on X is tamed.

Corollary 1.11

Suppose that dim(X)=2 and

for every n∈ℕ. Then X is Kähler–Einstein.

Two-dimensional Fano orbifolds are called del Pezzo surfaces.

Remark 1.12

Del Pezzo surfaces with quotient singularities are not classified (cf. [20]). But

  • del Pezzo surfaces with canonical singularities are classified (see [18]),

  • del Pezzo surfaces with 2-Gorenstein quotient singularities are classified (see [1]),

  • smoothable del Pezzo surfaces with quotient singularities of Picard rank 1 are classified (see [17]).

Del Pezzo surfaces with canonical singularities form a very natural class of del Pezzo surfaces.

Problem 1.13

Describe all Kähler–Einstein del Pezzo surface with canonical singularities.

Recall that if X is a del Pezzo surface with canonical singularities, then

  • either the inequality \(K_{X}^{2}\geqslant 5\) holds,

  • or one of the following possible cases occurs:

    • the equality \(K_{X}^{2}=1\) holds and X is a sextic surface in ℙ(1,1,2,3),

    • the equality \(K_{X}^{2}=2\) holds and X is a quartic surface in ℙ(1,1,1,2),

    • the equality \(K_{X}^{2}=3\) holds and X is a cubic surface in ℙ3,

    • the equality \(K_{X}^{2}=4\) holds and X is a complete intersection in ℙ4 of two quadrics.

Let us consider few examples to illustrate the expected answer to Problem 1.13.

Example 1.14

Suppose that X is a sextic surface in ℙ(1,1,2,3) such that its singular locus consists of singular points of type \(\mathbb{A}_{1}\) or \(\mathbb{A}_{2}\). Arguing as in the proof of [3, Lemma 4.1], we see that

for every n∈ℕ. Thus, the surface X is Kähler–Einstein by Corollary 1.11.

Example 1.15

Suppose that X is a quartic surface in ℙ(1,1,1,2) such that its singular locus consists of singular points of type \(\mathbb{A}_{1}\) or \(\mathbb{A}_{2}\). Then X is Kähler–Einstein by [16, Theorem 2].

Example 1.16

Suppose that X is a cubic surface in ℙ3 that is not a cone. Then

  • if X is smooth, then X is Kähler–Einstein by Theorem 1.1,

  • if Sing(X) consists of one point of type \(\mathbb{A}_{1}\), then it follows from [35, Theorem 5.1] that

    for every n∈ℕ, which implies that X is Kähler–Einstein by Corollary 1.11,

  • if the cubic surface X has a singular point that is not a singular point of type \(\mathbb{A}_{1}\) or \(\mathbb{A}_{2}\), then the surface X is not Kähler–Einstein by [11, Proposition 4.2].

Example 1.17

Suppose that X is a complete intersection in ℙ4 of two quadrics. Then

  • if X is smooth, then X is Kähler–Einstein by Theorem 1.1,

  • if X is Kähler–Einstein, then X has at most singular points of type \(\mathbb{A}_{1}\) (see [19]),

  • it follows from [24] or [16, Theorem 44] that X is Kähler–Einstein if it is given by

    and X has at most singular points of type \(\mathbb{A}_{1}\), where (λ 0:λ 1:λ 2:λ 3:λ 4)∈ℙ4.

Keeping in mind Examples 1.14, 1.15, 1.16, 1.17, [4, Example 1.12], and [26, Table 1], it is very natural to expect that the following answer to Problem 1.13 is true (cf. Example 1.6).

Conjecture 1.18

If the orbifold X is a del Pezzo surface with at most canonical singularities, then the surface X is Kähler–Einsteinit satisfies one of the following conditions:

  • \(K_{X}^{2}=1\) and Sing(X) consists of points of type \(\mathbb{A}_{1}\), \(\mathbb{A}_{2}\), \(\mathbb{A}_{3}\), \(\mathbb{A}_{4}\), \(\mathbb{A}_{5}\), \(\mathbb{A}_{6}\), \(\mathbb{A}_{7}\), or \(\mathbb{D}_{4}\),

  • \(K_{X}^{2}=2\) and Sing(X) consists of points of type \(\mathbb{A}_{1}\), \(\mathbb{A}_{2}\), or \(\mathbb{A}_{3}\),

  • \(K_{X}^{2}=3\) and Sing(X) consists of points of type \(\mathbb{A}_{1}\), or \(\mathbb{A}_{2}\),

  • \(K_{X}^{2}=4\) and Sing(X) consists of points of type \(\mathbb{A}_{1}\),

  • the surface X is smooth and \(6\geqslant K_{X}^{2}\geqslant 5\),

  • either X≅ℙ2 or X≅ℙ1×ℙ1.

In this paper, we prove the following result.

Theorem 1.19

Suppose that X is a sextic surface in ℙ(1,1,2,3). Then

for every n∈ℕ if Sing(X) consists of points of type \(\mathbb{A}_{1}\), \(\mathbb{A}_{2}\), \(\mathbb{A}_{3}\), \(\mathbb{A}_{4}\), \(\mathbb{A}_{5}\) or \(\mathbb{A}_{6}\).

Corollary 1.20

Suppose that X is a sextic surface in ℙ(1,1,2,3) such that its singular locus consists of singular points of type \(\mathbb{A}_{1}\), \(\mathbb{A}_{2}\), \(\mathbb{A}_{3}\), \(\mathbb{A}_{4}\), \(\mathbb{A}_{5}\), or \(\mathbb{A}_{6}\). Then X is Kähler–Einstein.

It should be pointed out that Corollary 1.20 and Examples 1.14, 1.15, 1.16, 1.17 illustrate a general philosophy that the existence of Kähler–Einstein metrics on Fano orbifolds is related to an algebro-geometric notion of stability (see [11, 12, 39, Theorem 4.1]).

Remark 1.21

If X is a sextic surface in ℙ(1,1,2,3) with canonical singularities, then either

or Sing(X) consists only of points of type \(\mathbb{A}_{1}\) and \(\mathbb{A}_{2}\) (see [40]).

What is known about α-invariants of del Pezzo surfaces with canonical singularities?

Theorem 1.22

[3]

If X is a smooth del Pezzo surface, then lct(X)=lct1(X).

Theorem 1.23

[3, 31]

If X is a del Pezzo surface with canonical singularities, then

in the case when \(K_{X}^{2}\geqslant 3\).

Theorem 1.24

[31]

If X is a quartic surface in ℙ(1,1,1,2) with canonical singularities, then

In this paper, we prove the following result (cf. Example 1.14).

Theorem 1.25

Suppose that X is a sextic surface in ℙ(1,1,2,3) with canonical singularities, let ω:X→ℙ(1,1,2) be a natural double cover, and let R be its branch curve in ℙ(1,1,2). Then

It should be pointed out that if X is a del Pezzo surface with at most canonical singularities, then all possible values of the number lct1(X) are computed in [2830].

Example 1.26

If X is a sextic surface in ℙ(1,1,2,3) with canonical singularities, then

  • lct1(X)=1/6 ⇔ the surface X has a singular point of type \(\mathbb{E}_{8}\),

  • lct1(X)=1/4 ⇔ the surface X has a singular point of type \(\mathbb{E}_{7}\),

  • lct1(X)=1/3 ⇔ the surface X has a singular point of type \(\mathbb{E}_{6}\),

  • lct1(X)=1/2 ⇔ the surface X has a singular point of type \(\mathbb{D}_{4}\), \(\mathbb{D}_{5}\), \(\mathbb{D}_{6}\), \(\mathbb{D}_{7}\) or \(\mathbb{D}_{8}\),

  • lct1(X)=2/3 ⇔ the following two conditions are satisfied:

    • the surface X has no singular points of type \(\mathbb{D}_{4}\), \(\mathbb{D}_{5}\), \(\mathbb{D}_{6}\), \(\mathbb{D}_{7}\), \(\mathbb{D}_{8}\), \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\),

    • there is a curve in |−K X | that has a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\),

  • lct1(X)=3/4 ⇔ the following three conditions are satisfied:

    • the surface X has no singular points of type \(\mathbb{D}_{4}\), \(\mathbb{D}_{5}\), \(\mathbb{D}_{6}\), \(\mathbb{D}_{7}\), \(\mathbb{D}_{8}\), \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\),

    • there is no curve in |−K X | that has a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\),

    • there is a curve in |−K X | that has a cusp at a point in Sing(X) of type \(\mathbb{A}_{1}\),

  • lct1(X)=5/6 ⇔ the following three conditions are satisfied:

    • the surface X has no singular points of type \(\mathbb{D}_{4}\), \(\mathbb{D}_{5}\), \(\mathbb{D}_{6}\), \(\mathbb{D}_{7}\), \(\mathbb{D}_{8}\), \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\),

    • there is no curve in |−K X | that have a cusp at a point in Sing(X),

    • there is a curve in |−K X | that has a cusp,

  • lct1(X)=1 ⇔ there are no cuspidal curves in |−K X |.

A crucial role in the proofs of both Theorems 1.25 and 1.19 is played by a new local inequality that we discovered. This inequality is a technical tool, but let us describe it now.

Let S be a surface, let D be an arbitrary effective ℚ-divisor on the surface S, let O be a smooth point of the surface S, let Δ1 and Δ2 be reduced irreducible curves on S such that

and the divisor Δ12 has a simple normal crossing singularity at the smooth point O∈Δ1∩Δ2, let a 1 and a 2 be some non-negative rational numbers. Suppose that the log pair

is not Kawamata log terminal at O, but (S,D+a 1Δ1+a 2Δ2) is Kawamata log terminal in a punctured neighborhood of the point O.

Theorem 1.27

Let A,B,M,N,α,β be non-negative rational numbers. Then

in the case when the following conditions are satisfied:

  • the inequality αa 1+βa 2⩽1 holds,

  • the inequalities A(B−1)⩾1⩾max(M,N) hold,

  • the inequalities α(A+M−1)⩾A 2(B+N−1)β and α(1−M)+A hold,

  • either the inequality 2M+AN⩽2 holds or

Corollary 1.28

Suppose that

for some integer m such that m⩾3. Then

Proof

To prove the required assertion, let us put

and let us check that all hypotheses of Theorem 1.27 are satisfied.

We have αa 1+βa 2⩽1 by assumption. We have

since m⩾3. We have

since m⩾3. We have α(1−M)+=2⩾2=A and 2M+AN=0⩽2.

Thus, we see that all hypotheses of Theorem 1.27 are satisfied. Then

by Theorem 1.27. □

For the convenience of the reader, we organize the paper in the following way:

  • in Sect. 2, we collect auxiliary results,

  • in Sect. 3, we prove Theorem 1.27,

  • in Sect. 4, we prove Theorem 4.1,

  • in Sect. 5, we prove Theorem 5.1,

  • in Sect. 6, we prove Theorem 6.1.

By Remark 1.21, both Theorems 1.19 and 1.25 follow from Theorems 4.1, 5.1, and 6.1.

2 Preliminaries

Let S be a surface with canonical singularities, and let D be an effective ℚ-divisor on S. Put

where D i is an irreducible curve, and a i ∈ℚ>0. We assume that D i D j ij.

Suppose that (S,D) is log canonical, but (S,D) is not Kawamata log terminal.

Remark 2.1

Let \(\bar{D}\) be an effective ℚ-divisor on the surface S such that

and the log pair \((S,\bar{D})\) is log canonical, where \(\bar{a}_{i}\) is a non-negative rational number. Put

where α is well defined and α⩽1. Then \(\alpha=1\iff D=\bar{D}\). Suppose that \(D\ne \bar{D}\). Put

and choose k∈{1,…,r} such that \(\alpha=a_{k}/\bar{a}_{k}\). Then \(D_{k}\not\subset\mathrm{Supp}(D^{\prime})\) and \(D^{\prime}\sim_{\mathbb{Q}} \bar{D}\sim_{\mathbb{Q}} D\), but the log pair (S,D′) is not Kawamata log terminal.

Let LCS(S,D) be the locus of log canonical singularities of the log pair (S,D) (see [6]).

Theorem 2.2

[22, Theorem 17.4]

If −(K S +D) is nef and big, then LCS(S,D) is connected.

Take a point P∈LCS(S,D). Suppose that LCS(S,D) contains no curves that pass through P.

Lemma 2.3

Suppose that \(P\not\in\mathrm{Sing}(S)\) and \(P\not\in\mathrm{Sing}(D_{1})\). Then

Proof

The log pair \((S, D_{1}+\sum_{i=2}^{r}a_{i}D_{i})\) is not log canonical at P, since a 1<1. Then

by [22, Theorem 17.6]. □

Let \(\pi\colon \bar{S}\to S\) be a birational morphism, and let \(\bar{D}\) be a proper transform of D via π. Then

where E i is an irreducible π-exceptional curve, and e i ∈ℚ. We assume that E i =E j i=j.

Suppose, in addition, that the birational morphism π induces an isomorphism

Remark 2.4

The log pair \((\bar{S},\bar{D}+\sum_{i=1}^{s}e_{i}E_{i})\) is not Kawamata log terminal at a point in \(\bigcup_{i=1}^{s}E_{i}\).

Suppose that S is singular at P, and either P is a singular point of type \(\mathbb{D}_{n}\) for some n∈ℕ⩾4, or the point P is a singular point of type \(\mathbb{E}_{m}\) for some m∈{6,7,8}.

Lemma 2.5

Suppose that \(E_{1}^{2}=E_{2}^{2}=\cdots=E_{s}^{2}=-2\). Then e 1=1 if

Proof

This follows from [32, Proposition 2.9], because (SP) is a weakly exceptional singularity (see [32, Example 4.7], [7, Example 3.4], [7, Theorem 3.15]). □

Lemma 2.6

Suppose that S is a sextic surface in ℙ(1,1,2,3) that has canonical singularities, and suppose that DK X . Let μ be a positive rational number such that either

or μ=2/3 and D is not a curve in |−K X | with a cusp at a point in Sing(S) of type \(\mathbb{A}_{2}\). Then

the locus LCS(S,μD) contains no points of type \(\mathbb{A}_{1}\) or \(\mathbb{A}_{2}\), and |LCS(S,μD)|⩽1.

Proof

This follows from Theorem 2.2 and the proof of [3, Lemma 4.1]. □

Most of the described results are valid in much more general settings (cf. [22] and [21]).

3 Local Inequality

The purpose of this section is to prove Theorem 1.27.

Let S be a surface, let D be an arbitrary effective ℚ-divisor on the surface S, let O be a smooth point of the surface S, let Δ1 and Δ2 be reduced irreducible curves on S such that

and the divisor Δ12 has a simple normal crossing singularity at the smooth point O∈Δ1∩Δ2, let a 1 and a 2 be some non-negative rational numbers. Suppose that the log pair

is not Kawamata log terminal at O, but (S,D+a 1Δ1+a 2Δ2) is Kawamata log terminal in a punctured neighborhood of the point O. In particular, we must have a 1<1 and a 2<1.

Let A,B,M,N,α,β be non-negative rational numbers such that

  • the inequality αa 1+βa 2⩽1 holds,

  • the inequalities A(B−1)⩾1⩾max(M,N) hold,

  • the inequalities α(A+M−1)⩾A 2(B+N−1)β and α(1−M)+A holds,

  • either the inequality 2M+AN⩽2 holds or

Lemma 3.1

The inequalities A+M⩾1 and B>1 holds. The inequality

holds. The inequality β(1−N)+B holds. The inequalities

and α(2−M)B+β(1−N)(A+1)⩾B(A+1) hold.

Proof

The inequality B>1 follows from the inequality A(B−1)⩾1. Then

because 2BB+1. Similarly, we see that A+M⩾1, because

and B+N−1⩾0. The inequality β(1−N)+B follows from the inequalities

because A+1⩾2−M.

Let us show that the inequality

holds. Let L 1 be the line in ℝ2 given by the equation

and let L 2 be the line that is given by the equation

where (x,y) are coordinates on ℝ2. Then L 1 intersects the line y=0 at the point

and L 2 intersects the line y=0 at the point (A/(1−M),0). But

which implies that α(2−M)B+β(1−N)(A+1)⩾B(A+1) if

where (α 0,β 0) is the intersection point of the lines L 1 and L 2. But

where Δ=2ABABMA+AM−1+M+NANAM+NNM. But

because A(B−1)⩾1, which implies that A 2 β 0(B+N−1)⩾α 0(A+M−1).

Finally, let us show that the inequality

holds. Let \(L^{\prime}_{1}\) be the line in ℝ2 given by the equation

where (x,y) are coordinates on ℝ2. Then \(L^{\prime}_{1}\) intersects the line y=0 at the point

and L 2 intersects the line y=0 at the point (A/(1−M),0). But

which implies that α(B+1−MBN)+β(A+1−ANM)⩾AB−1 if

where (α 1,β 1) is the intersection point of the lines \(L^{\prime}_{1}\) and L 2. Note that

where Δ′=AB−1−ABM+AM+2MNAMM 2.

To complete the proof, it is enough to show that the inequality

holds. This inequality is equivalent to the inequality

which is true, because M⩽1 and AN+2M−2⩽0. □

Let us prove Theorem 1.27 by reductio ad absurdum. Suppose that the inequalities

hold. Let us show that this assumption leads to a contradiction.

Lemma 3.2

The inequalities a 1>(1−M)/A and a 2>(1−N)/B hold.

Proof

It follows from Lemma 2.3 that

which implies that a 1>(1−M)/A. Similarly, we see that a 2>(1−N)/B. □

Put m 0=mult O (D). Then m 0 is a positive rational number.

Remark 3.3

The inequalities m 0<M+Aa 1a 2 and m 0<N+Ba 2a 1 hold.

Lemma 3.4

The inequality m 0+a 1+a 2<2 holds.

Proof

We know that m 0+a 1+a 2<M+(A+1)a 1 and m 0+a 1+a 2<N+(B+1)a 2. Then

which implies that m 0+a 1+a 2<2 by Lemma 3.1. □

Let π 1:S 1S be the blow-up of the point O, and let F 1 be the π 1-exceptional curve. Then

where D 1, \(\Delta^{1}_{1}\), \(\Delta^{1}_{2}\) are proper transforms of the divisors D, Δ1, Δ2 via π 1, respectively. Then

is not Kawamata log terminal at some point O 1F 1 (see Remark 2.4), where m 0+a 1+a 2⩾1.

Lemma 3.5

Either \(O_{1}=F_{1}\cap\Delta^{1}_{1}\) or \(O_{1}=F_{1}\cap\Delta^{1}_{2}\).

Proof

Suppose that \(O_{1}\not \in\Delta^{1}_{1}\cup\Delta^{1}_{2}\). Then m 0=D 1F 1>1 by Lemma 2.3. But

because m 0<M+Aa 1a 2 and m 0<N+Ba 2a 1. On the other hand, we have

because αa 1+βa 2⩽1 and AB−1>0. But we already proved that m 0>1. Thus, we see that

which is impossible by Lemma 3.1. □

Lemma 3.6

The inequality \(O_{1}\ne F_{1}\cap\Delta^{1}_{1}\) holds.

Proof

Suppose that \(O_{1}=F_{1}\cap\Delta^{1}_{1}\). It follows from Lemma 2.3 that

which implies that a 1>(2−M)/(A+1). Then

because a 2>(1−N)/B by Lemma 3.2. Thus, we see that

which is impossible by Lemma 3.1. □

Therefore, we see that \(O_{1}=F_{1}\cap\Delta^{1}_{2}\). Then the log pair

is not Kawamata log terminal at the point O 1. We know that 1>m 0+a 1+a 2−1⩾0.

We have a blow-up π 1:S 1S. For any n∈ℕ, consider a sequence of blow-ups

such that π i+1:S i+1S i is a blow-up of the point \(F_{i}\cap\Delta^{i}_{2}\) for every i∈{1,…,n−1}, where

  • we denote by F i the exceptional curve of the morphism π i ,

  • we denote by \(\Delta^{i}_{2}\) the proper transform of the curve Δ2 on the surface S i .

For every k∈{1,…,n} and for every i∈{1,…,k}, let D k, \(\Delta^{k}_{1}\), and \(F^{k}_{i}\) be the proper transforms on the surface S k of the divisors D, Δ1 and F i , respectively. Then

where π=π n ∘⋯∘π 2π 1 and \(m_{i}=\mathrm{mult}_{O_{i}}(D^{i})\) for every i∈{1,…,n}. Then the log pair

$$ \Biggl(S_{n},D^{n}+a_{1} \Delta^{n}_{1}+a_{2}\Delta^{n}_{2}+ \sum_{i=1}^{n} \Biggl(a_{1}+ia_{2}-i+ \sum_{j=0}^{i-1}m_{j} \Biggr)F^{n}_{i} \Biggr) $$
(2)

is not Kawamata log terminal at some point of the set \(F^{n}_{1}\cup F^{n}_{2}\cup\cdots\cup F^{n}_{n}\) (see Remark 2.4).

Put \(O_{k}=F_{k}\cap\Delta^{k}_{2}\) for every k∈{1,…,n}.

Lemma 3.7

For every i∈{1,…,n}, we have

and (2) is Kawamata log terminal at every point of the set \((F^{n}_{1}\cup F^{n}_{2}\cup\cdots\cup F^{n}_{n})\setminus O_{n}\).

Since mult O (D⋅Δ2)<N+Ba 2a 1 by assumption, it follows from Lemma 3.7 that

which implies that n⩽(N+Ba 2)/(1−a 2). On the other hand, the assertion of Lemma 3.7 holds for arbitrary n∈ℕ. So, taking any n>(N+Ba 2)/(1−a 2), we obtain a contradiction.

We see that to prove Theorem 1.27, it is enough to prove Lemma 3.7.

Let us prove Lemma 3.7 by induction on n∈ℕ. The case n=1 is already done.

We may assume that n⩾2. For every k∈{1,…,n−1}, we may assume that

the singularities of the log pair

are Kawamata log terminal along \((F^{k}_{1}\cup F^{k}_{2}\cup\cdots\cup F^{k}_{k})\setminus O_{k}\) and not Kawamata log terminal at O k .

Lemma 3.8

The inequality a 2>(nN)/(B+n−1) holds.

Proof

The singularities of the log pair

are not Kawamata log terminal at the point O n−1. Then it follows from Lemma 2.3 that

which implies that a 2>(nN)/(B+n−1). □

Lemma 3.9

The inequalities \(1>a_{1}+na_{2}-n+\sum_{j=0}^{n-1}m_{j}\geqslant 0\) hold.

Proof

The inequality \(a_{1}+na_{2}-n+\sum_{j=0}^{n-1}m_{j}\geqslant 0\) follows from the fact that the log pair

is not Kawamata log terminal at the point O n−1.

Suppose that \(a_{1}+na_{2}-n+\sum_{j=0}^{n-1}m_{j}\geqslant 1\). Let us derive a contradiction.

It follows from Remark 3.3 that m 0+a 2M+Aa 1. Then

which implies that a 1⩾(n+1−Mn)/(nA+1). But a 2>(nN)/(B+n−1) by Lemma 3.8. Then

where α(1−M)/A+β⩾1 by assumption. Therefore, we see that

where n⩾2. But A+M>1 and B+N>1 by Lemma 3.2, since a 1<1 and a 2<1. Then

but A 2(B+N−1)βα(A+M−1) by assumption. Then

which implies that βA(B+N−1)>α(B−1)(A+M−1). Then

because A 2(B+N−1)βα(A+M−1) by assumption. Then we have α≠0 and A(B−1)<1, which is impossible, because A(B−1)⩾1 by assumption. □

Lemma 3.10

The log pair (2) is Kawamata log terminal at every point of the set

Proof

Suppose that there is a point QF n such that

but (2) is not Kawamata log terminal at the point Q. Then the log pair

is not Kawamata log terminal at the point Q as well. Then

by Lemma 2.3, because \(a_{1}+na_{2}-n+\sum_{j=0}^{n-1}m_{j}<1\) by Lemma 3.9. Then

because m 0<M+Aa 1a 2 and m 0<N+Ba 2a 1 by Remark 3.3. We have

because αa 1+βa 2⩽1 and AB−1>0. But m 0>1. Thus, we see that

which contradicts our initial assumptions. □

Lemma 3.11

The log pair (2) is Kawamata log terminal at the point \(F_{n}\cap F^{n}_{n-1}\).

Proof

Suppose that (2) is not Kawamata log terminal at \(F_{n}\cap F^{n}_{n-1}\). Then the log pair

is not Kawamata log terminal at the point \(F_{n}\cap F^{n}_{n-1}\) as well. Then

by Lemma 2.3, because \(a_{1}+(n-1)a_{2}-(n-1)+\sum_{j=0}^{n-2}m_{j}<1\). Note that

which implies that m 0+a 2<Aa 1+M. Then

which gives a 1>(n+1−nM)/(An+1).

Now arguing as in the proof of Lemma 3.9, we obtain a contradiction. □

The assertion of Lemma 3.7 is proved. The assertion of Theorem 1.27 is proved.

4 One Cyclic Singular Point

Let X be a sextic surface in ℙ(1,1,2,3) with canonical singularities such that |Sing(X)|=1, let ω:X→ℙ(1,1,2) be the natural double cover, let R be its ramification curve in ℙ(1,1,2), and suppose that Sing(X) consists of one singular point of type \(\mathbb{A}_{m}\), where m∈{1,…,8}.

Theorem 4.1

The following equality holds:

and if lct(X)=2/3, then there is a unique effective ℚ-divisor D on X such that DK X and

By Theorem 1.5, Corollary 1.11, and Remark 1.7, we obtain the following two corollaries.

Corollary 4.2

If m⩽6, then lct n,2(X)>2/3 for every n∈ℕ.

Corollary 4.3

If m⩽6, then X is Kähler–Einstein.

In the rest of this section, we will prove Theorem 4.1.

Let D be an arbitrary effective ℚ-divisor on the surface X such that

and put μ=c(X,D). To prove Theorem 4.1, it is enough to show that

and if μ=lct(X)=2/3, then D is uniquely defined. Note that lct1(X)⩾5/6 if m⩾3 (see [30]).

Let us prove Theorem 4.1. By Lemma 2.6, we may assume that m⩾3 and μ<lct1(X). Then

by Lemma 2.6. Put P=Sing(X).

Let \(\pi\colon\bar{X}\to X\) be a minimal resolution, let E 1,E 2,…,E m be π-exceptional curves such that

let C be the curve in |−K X | such that PC, and let \(\bar{C}\) be it proper transform on \(\bar{X}\). Then

and the curve C is irreducible. We may assume that DC, because μ⩾lct1(X) if D=C.

By Remark 2.1, we may assume that \(C\not\subset\mathrm{Supp}(D)\).

Let \(\bar{D}\) be the proper transform of the divisor D on the surface \(\bar{X}\). Then

where a i is a non-negative rational number. Then the log pair

$$ \Biggl(\bar{X},\mu\bar{D}+\sum _{i=1}^{m}\mu a_iE_i \Biggr) $$
(3)

is not Kawamata log terminal (by Remark 2.4). On the other hand, we have

where all intersections \(\bar{D}\cdot E_{1}, \bar{D}\cdot E_{2}, \ldots, \bar{D}\cdot E_{m}\) are non-negative. Moreover, we have

where the intersection \(\bar{D}\cdot\bar{C}\) is non-negative, since \(C\not\subset\mathrm{Supp}(D)\) by assumption. Hence, we have

$$ \left\{\begin{array}{l@{\quad}l} \displaystyle a_1\geqslant \frac{a_2}{2}, \\[7pt] a_2\geqslant \displaystyle\frac{a_1+a_3}{2}, \\[7pt] a_3\geqslant \displaystyle\frac{a_2+a_4}{2}, \\[7pt] \vdots \\[7pt] a_{m-1}\geqslant\displaystyle \frac{a_{m-2}+a_m}{2}, \\[7pt] a_m\geqslant\displaystyle\frac{a_{m-1}}{2}, \\[7pt] 1\geqslant a_1+a_m. \\ \end{array} \right. $$
(4)

It should be pointed out that at least one inequality in (4) must be strict, since \(\bar{D}\cdot E_{i}>0\) for at least one i∈{1,…,m}, because P∈Supp(D). Then a i >0 for some i∈{1,…,m}.

Note that a 1a 2/2 by (4). Similarly, it follows from (4) that

which implies that a 2⩾2a 3/3. Arguing in the same way, we see that

for every k∈{1,…,m−1} (use (4) and induction on k). Using symmetry, we see that

for every k∈{1,…,m−1}. In particular, the inequality a k >0 holds for every k∈{1,…,m}, since we already know that a i >0 for some i∈{1,…,m}.

Lemma 4.4

Suppose that μa i <1 for every i∈{1,…,m}. Then

  • there exists a point

    such that the log pair (3) is not Kawamata log terminal at Q,

  • the log pair (3) is Kawamata log terminal outside of the point Q,

  • if μ<(m+1)/(2m−2), then QE 1E 2 and QE m−1E m .

Proof

It follows from Remark 2.4 and Theorem 2.2 that there is a point \(Q\in \bigcup_{i=1}^{m}E_{i}\) such that the log pair (3) is not Kawamata log terminal at Q and is Kawamata log terminal elsewhere.

Suppose that QE 1 and \(Q\not\in E_{2}\). Then

by Lemma 2.3. Taking (4) into account, we get

and adding all these inequalities together we get

which implies that a 1+a m >1. However, the latter is impossible, since a 1+a m ⩽1 by (4).

We see that if QE 1, then Q=E 1E 2. Similarly, we see that Q=E m−1E m if QE m .

Suppose that QE i and \(Q\not\in E_{j}\) for every ji. Then i≠1 and im. We have

by Lemma 2.3. Taking (4) into account, we get

and adding all these inequalities together we get

which implies that a 1+a m >1. However, the latter is impossible, since a 1+a m ⩽1 by (4).

Thus, we see that there is k∈{1,…,m−1} such that Q=E k E k+1.

Suppose that μ<(m+1)/(2m−2). Let us show that k≠1 and km−1.

Due to symmetry, it is enough to show that k≠1. Recall that m⩾3.

Suppose that k=1. Then Q=E 1E 2. Take \(\bar{\mu}\in\mathbb{Q}\) such that \((m+1)/(2m-2)>\bar{\mu}>\mu\) and

is not Kawamata log terminal at Q and is Kawamata log terminal outside of the point Q. Then

by (4), since a 1≠0 and a 2≠0. On the other hand, we have

since \(\mu<\bar{\mu}\). Therefore, it follows from Corollary 1.28 that

which implies that a 2(m−2)>a 3(m−1), since \(\mu<\bar{\mu}\). But we proved earlier that

which is impossible, since a 2(m−2)>a 3(m−1). Thus, we see that k≠1. □

If m=3, then it follows from (4) that a 1⩽3/4, a 2⩽1, a 3⩽3/4.

Corollary 4.5

If m=3, then μ⩾lct1(X)⩾5/6.

Lemma 4.6

Suppose that m=4. Then μ⩾lct2(X)=4/5.

Proof

There is a unique smooth irreducible curve \(\bar{Z}\subset\bar{X}\) such that

and \(E_{2}\cap E_{3}\in\bar{Z}\) (cf. the proof of Lemma 6.8). Put \(Z=\pi(\bar{Z})\). Then

To complete the proof, it is enough to show that μ⩾4/5. Suppose that μ<4/5.

By Remark 2.1, we may assume that \(Z\not\subset\mathrm{Supp}(D)\), because Z is irreducible.

It follows from (4) that a 1⩽4/5, a 2⩽6/5, a 3⩽6/5, a 4⩽4/5.

Put Q=E 2E 3. Then it follows from Lemma 4.4 that (3) is not Kawamata log terminal at the point Q and is Kawamata log terminal outside of the point Q. Then

by Lemma 2.3. Similarly, we see that

which implies that a 2>5/6 and a 3>5/6.

Let \(\xi\colon\tilde{X}\to\bar{X}\) be a blow-up of the point Q, let E be the exceptional curve of the blow-up ξ, and let \(\tilde{D}\) be the proper transform of the divisor \(\bar{D}\) on the surface \(\tilde{X}\). Put \(\delta=\mathrm{mult}_{Q}(\bar{D})\).

Let \(\tilde{E}_{1}\), \(\tilde{E}_{2}\), \(\tilde{E}_{3}\), \(\tilde{E}_{4}\) be the proper transforms on \(\tilde{X}\) of E 1, E 2, E 3, E 4, respectively. Then

$$ \bigl(\tilde{X},\mu\tilde{D}+\mu a_2\tilde{E}_2+\mu a_3\tilde{E}_3+ (\mu a_2+\mu a_3+\mu \delta-1 )E \bigr)$$
(5)

is not Kawamata log terminal at some point OE.

Let \(\tilde{Z}\) be the proper transform on \(\tilde{X}\) of the curve \(\bar{Z}\). Then

which implies that δ+a 2+a 3⩽2. We have μa 2+μa 3+μδ−1⩽2μ−1⩽3/5, which implies that (5) is Kawamata log terminal outside of the point O by Theorem 2.2. We have

which implies that δ⩽1. If \(O\not\in\tilde{E}_{2}\cup\tilde{E}_{3}\), then

by Lemma 2.3. Thus, we see that either \(O=\tilde{E}_{2}\cap E\) or \(O=\tilde{E}_{3}\cap E\).

Without loss of generality, we may assume that \(O=\tilde{E}_{2}\cap E\). By Lemma 2.3, one has

since δ+a 2+a 3⩽2 and a 3>5/6. The obtained contradiction concludes the proof. □

Let τ be a biregular involution of the surface \(\bar{X}\) that is induced by the double cover ω.

Lemma 4.7

Suppose that m=5. Then there exists a unique curve Z∈|−2K X | such that

and either D=Z/2 or μ>2/3.

Proof

Let \(\alpha\colon\bar{X}\to\breve{X}\) be a contraction of the curves \(\bar{C}\), E 5, E 4, E 3. Then

and \(\breve{X}\) is a smooth del Pezzo surface such that \(K_{\breve{X}}^{2}=5\), which implies that there is a smooth irreducible rational curve \(\breve{L}_{2}\) on the surface \(\breve{X}\) such that \(\breve{L}_{2}\cdot\alpha(E_{2})=1\) and \(\breve{L}_{2}\cdot \breve{L}_{2}=-1\).

Let \(\bar{L}_{2}\) be the proper transform of the curve \(\breve{L}_{2}\) on the surface \(\bar{X}\). Then \(\bar{L}_{2}\cdot \bar{L}_{2}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{2}=E_{3}\cdot \bar{L}_{2}=E_{4}\cdot \bar{L}_{2}=E_{5}\cdot \bar{L}_{2}=\bar{C}\cdot\bar{L}_{2}=0\).

Let \(\beta\colon\bar{X}\to\check{X}\) be a contraction of the curves \(\bar{L}_{2}\), \(\bar{C}\), E 5, E 4. Then

and \(\check{X}\) is a smooth del Pezzo surface such that \(K_{\check{X}}^{2}=5\), which implies that there is an irreducible smooth curve \(\check{L}_{3}\subset\check{X}\) such that \(\check{L}_{3}\cdot\beta(E_{3})=1\) and \(\check{L}_{3}\cdot\check{L}_{3}=-1\) (cf. the proof of Lemma 6.7).

Let \(\bar{L}_{3}\) be the proper transform of the curve \(\check{L}_{3}\) on the surface \(\bar{X}\). Then \(\bar{L}_{3}\cdot \bar{L}_{3}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{3}=E_{2}\cdot \bar{L}_{3}=E_{4}\cdot \bar{L}_{3}=E_{5}\cdot \bar{L}_{3}=\bar{C}\cdot\bar{L}_{3}=0\).

If \(\tau(\bar{L}_{3})=\bar{L}_{3}\), then \(2\pi(\bar{L}_{3})\sim -2K_{X}\), but \(\pi(\bar{L}_{3})\) is not a Cartier divisor.

Put \(Z=\pi(\bar{L}_{3}+\tau(\bar{L}_{3}))\). Then Z∼−2K X and c(X,Z)=1/3. We see that lct2(X)⩽2/3.

Suppose that DZ/2. To complete the proof, it is enough to show that μ>2/3.

Suppose that μ⩽2/3. Let us derive a contradiction. It follows from (4) that

By Remark 2.1, without loss of generality we may assume that \(\pi(\bar{L}_{3})\not\subset\mathrm{Supp}(D)\). Then

which implies that a 3⩽1.

Put Q=E 2E 3. By Lemma 4.4, we may assume that (3) is not Kawamata log terminal at the point Q and is Kawamata log terminal outside of the point Q. Then

by Lemma 2.3, which implies that a 3>9/8 by (4). But a 3⩽1. □

Lemma 4.8

Suppose that m=6. Then there exists a unique curve Z∈|−2K X | such that

and either D=Z/2 or μ>2/3.

Proof

Let \(\alpha\colon\bar{X}\to\breve{X}\) be a contraction of the curves \(\bar{C}\), E 6, E 5, E 4, and E 3. Then

and \(\breve{X}\) is a smooth del Pezzo surface such that \(K_{\breve{X}}^{2}=6\), which implies that there is a smooth irreducible rational curve \(\breve{L}_{2}\) on the surface \(\breve{X}\) such that \(\breve{L}_{2}\cdot\alpha(E_{2})=1\) and \(\breve{L}_{2}\cdot \breve{L}_{2}=-1\).

Let \(\bar{L}_{2}\) be the proper transform of the curve \(\breve{L}_{2}\) on the surface \(\bar{X}\). Then \(\bar{L}_{2}\cdot \bar{L}_{2}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{2}=E_{3}\cdot \bar{L}_{2}=E_{4}\cdot \bar{L}_{2}=E_{5}\cdot \bar{L}_{2}=E_{6}\cdot \bar{L}_{2}=\bar{C}\cdot\bar{L}_{2}=0\).

Let \(\beta\colon\bar{X}\to\check{X}\) be a contraction of the curves \(\bar{L}_{2}\), \(\bar{C}\), E 6, E 5, and E 4. Then

and \(\check{X}\) is a smooth del Pezzo surface such that \(K_{\check{X}}^{2}=6\), which implies that there are irreducible smooth rational curves \(\check{L}_{3}\) and \(\check{L}_{2}^{\prime}\) on the surface \(\check{X}\) such that

and \(\check{L}_{3}\cdot\check{L}_{3}=\check{L}_{2}^{\prime}\cdot\check{L}_{2}^{\prime}=-1\). Let \(\bar{L}_{3}\) and \(\bar{L}_{2}^{\prime}\) be the proper transforms of the curves \(\check{L}_{3}\) and \(\check{L}_{2}^{\prime}\) on the surface \(\bar{X}\), respectively. Then \(\bar{L}_{3}\cdot \bar{L}_{3}=\bar{L}_{2}^{\prime}\cdot \bar{L}_{2}^{\prime}=-1\) and

which implies that \(\bar{C}\cdot\bar{L}_{3}=\bar{C}\cdot\bar{L}_{2}^{\prime}=0\), and \(E_{i}\cdot \bar{L}_{3}=E_{j}\cdot \bar{L}_{2}^{\prime}=0\) for every i≠3 and j≠2,

Put \(\bar{L}_{4}=\tau(\bar{L}_{3})\), \(\bar{L}_{5}=\tau(\bar{L}_{2})\), \(\bar{L}_{5}^{\prime}=\tau(\bar{L}_{2}^{\prime})\). Then \(\bar{C}\cdot\bar{L}_{4}=\bar{C}\cdot\bar{L}_{5}=\bar{C}\cdot\bar{L}_{5}^{\prime}=0\) and

which implies that \(E_{i}\cdot\bar{L}_{5}=E_{i}\cdot\bar{L}_{5}^{\prime}=E_{j}\cdot \bar{L}_{4}=0\) for every i≠5 and j≠4.

Put \(L_{3}=\pi(\bar{L}_{3})\), \(L_{4}=\pi(\bar{L}_{4})\), \(L_{2}=\pi(\bar{L}_{2})\), \(L_{2}^{\prime}=\pi(\bar{L}_{2}^{\prime})\), \(L_{5}=\pi(\bar{L}_{5})\), \(L_{5}^{\prime}=\pi(\bar{L}_{5}^{\prime})\). Then

and c(X,L 3+L 4)=1/3, which implies that lct2(X)⩽2/3.

Note that \(\mathrm{c}(X,L_{2}+L_{5})=\mathrm{c}(X,L_{2}^{\prime}+L_{5}^{\prime})=1/2\).

Suppose that D≠(L 3+L 4)/2. To complete the proof, it is enough to show that μ>2/3.

Suppose that μ⩽2/3. Let us derive a contradiction.

It follows from (4) that a 1⩽6/7, a 2⩽10/7, a 3⩽12/7, a 4⩽12/7, a 5⩽10/7, a 6⩽6/7.

By Remark 2.1, without loss of generality we may assume that \(\bar{L}_{4}\not\subset\mathrm{Supp}(D)\). Then

which gives us a 4⩽1. Similarly, we may assume that either \(\bar{L}_{2}\not\subset\mathrm{Supp}(D)\) or \(\bar{L}_{5}\nobreak\not\subset\nobreak\mathrm{Supp}(D)\), which implies that either a 2⩽1 or a 5⩽1, respectively.

Let us show that \(L_{2}+L_{2}^{\prime}+L_{3}\sim -3K_{X}\). We can easily see that

which implies that \(L_{2}+L_{2}^{\prime}+L_{3}\sim_{\mathbb{Q}} -3K_{X}\), since Pic(X)≅ℤ3 and

but \(L_{2}+L_{2}^{\prime}+L_{3}\) is a Cartier divisor, which implies that \(L_{2}+L_{2}^{\prime}+L_{3}\sim -3K_{X}\).

Since \(\mathrm{c}(X,L_{2}+L_{2}^{\prime}+L_{3})=1/4\), we may assume that Supp(D) does not contain at least one curve among L 2, \(L_{2}^{\prime}\), and L 3 by Remark 2.1, which implies that either a 2⩽1 or a 3⩽1.

It follows from (4) and a 4⩽2 that μa i <1 for every i. By Lemma 4.4, there exists a point

such that (3) is not Kawamata log terminal at the point \(Q\in\bar{X}\), but it is Kawamata log terminal elsewhere. Take k∈{2,3,4} such that Q=E k E k+1. It follows from Lemma 2.3 that

which is impossible by (4), since a 4⩽1, and either a 2⩽1 or a 3⩽1. □

Lemma 4.9

Suppose that m=7. Then the following conditions are equivalent:

  • the curve R is irreducible,

  • the surface \(\bar{X}\) contains an irreducible curve \(\bar{L}_{4}\) such that \(\bar{L}_{4}\cdot \bar{L}_{4}=-1\) and \(\bar{L}_{4}\cdot E_{4}=1\).

  • the surface \(\bar{X}\) contains an irreducible curve \(\bar{L}_{4}\) such that \(\bar{L}_{4}\cdot \bar{L}_{4}=-1\), \(\bar{L}_{4}\cdot E_{4}=1\) and

Proof

Suppose that \(\bar{X}\) has an irreducible curve \(\bar{L}_{4}\) such that \(\bar{L}_{4}\cdot \bar{L}_{4}=-1\) and \(\bar{L}_{4}\cdot E_{4}=1\). Then

where \(L_{4}=\pi(\bar{L}_{4})\). Then \(\tau(\bar{L}_{4})=\bar{L}_{4}\) and ω(L 4)⊂Supp(R), because

Suppose now that the curve R is reducible. Let us show that the surface \(\bar{X}\) contains an irreducible curve \(\bar{L}_{4}\) such that \(\bar{L}_{4}\cdot \bar{L}_{4}=-1\) and \(\bar{L}_{4}\cdot E_{4}=1\).

Let \(\eta\colon\bar{X}\to\bar{X}^{\prime}\) be a contraction of the curve \(\bar{C}\). Then there is a commutative diagram

where π′ is a minimal resolution, ϕ is an anticanonical embedding, ψ is a projection from ϕω(P), and ω′ is a double cover branched at ψϕ(R). Note that X′ is a del Pezzo surface and \(K_{X^{\prime}}^{2}=2\).

The morphism π′ contracts the smooth curves η(E 2), η(E 3), η(E 4), η(E 5), and η(E 6). But

and X′ has a singularity of type \(\mathbb{A}_{5}\) at the point η(E 2). Put P′=η(E 2).

Put R′=ψϕ(R). Then R′ is reducible, since R is reducible.

Since \(\mathrm{Sing}(\mathbb{P}(1,1,2))\not\in R\), one of the following cases holds:

  • either ϕ(R) is a union of a smooth conic and an irreducible quartic,

  • or the curve ϕ(R) is a union of three different smooth conics.

The case when the curve ϕ(R) consists of a union of three different smooth conics is impossible, since the surface X′ has a singularity of type \(\mathbb{A}_{5}\) at the point P′=Sing(X′).

We see that the curve ϕ(R) is a union of a smooth conic and an irreducible quartic curve, which easily implies that R′ is a union of a line L and an irreducible cubic curve Z. Then

because X′ has a singularity of type \(\mathbb{A}_{5}\) at the point P′. Then \(\bar{X}\) contains a curve \(\bar{L}_{4}\) such that

and \(\bar{L}_{4}\) is irreducible. Then \(\bar{L}_{4}\cdot \bar{L}_{4}=-1\) and \(\bar{L}_{4}\cdot E_{4}=1\). □

The proof of Lemma 4.9 can be simplified using the results obtained in [31, Sect. 2].

Lemma 4.10

Suppose that m=7 and R is irreducible. Then μ⩾lct3(X)=3/5.

Proof

Arguing as in the proofs of Lemmas 4.7 and 4.8, we see that there is an irreducible smooth rational curve \(\bar{L}_{2}\) on the surface \(\bar{X}\) such that \(\bar{L}_{2}\cdot \bar{L}_{2}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{2}=E_{3}\cdot \bar{L}_{2}=E_{4}\cdot \bar{L}_{2}=E_{5}\cdot \bar{L}_{2}=E_{6}\cdot \bar{L}_{2}=E_{7}\cdot \bar{L}_{2}=\bar{C}\cdot\bar{L}_{2}=0\).

Put \(\bar{L}_{5}=\tau(\bar{L}_{2})\). Then \(\bar{L}_{5}\cdot \bar{L}_{5}=-1\) and \(-K_{\bar{X}}\cdot \bar{L}_{5}=E_{5}\cdot\bar{L}_{5}=1\), which implies that

Since the branch curve R is reducible by Lemma 4.9, one can show that there exists an irreducible smooth rational curve \(\bar{L}_{3}\) on the surface \(\bar{X}\) such that \(\bar{L}_{3}\cdot \bar{L}_{3}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{3}=E_{2}\cdot \bar{L}_{3}=E_{4}\cdot \bar{L}_{3}=E_{5}\cdot \bar{L}_{3}=E_{6}\cdot \bar{L}_{3}=E_{7}\cdot \bar{L}_{3}=\bar{C}\cdot\bar{L}_{3}=0\).

Put \(\bar{L}_{6}=\tau(\bar{L}_{2})\), \(\bar{L}_{5}=\tau(\bar{L}_{3})\), \(L_{2}=\pi(\bar{L}_{2})\), \(L_{3}=\pi(\bar{L}_{4})\), \(L_{5}=\pi(\bar{L}_{5})\) and \(L_{6}=\pi(\bar{L}_{6})\). Then

which implies that L 2+2L 3∼−3K X . Indeed, we have L 2+2L 3−3K X , since

and Pic(X)≅ℤ3. But L 2+2L 3 is a Cartier divisor, which implies that L 2+2L 3∼−3K X .

We have c(X,L 2+2L 3)=3/15 and L 2+2L 3∼−3K X , which implies that lct3(X)⩽3/5.

To complete the proof, it is enough to show that μ⩾3/5.

Suppose that μ<3/5. Let us derive a contradiction.

By Remark 2.1, we may assume that the support of the divisor \(\bar{D}\) does not contain at least one component of every curve \(\bar{L}_{2}+\bar{L}_{6}\), \(\bar{L}_{2}+2\bar{L}_{3}\), \(\bar{L}_{3}+\bar{L}_{5}\). But

which implies that a i ⩽1 if \(\bar{L}_{i}\not\subset\mathrm{Supp}(\bar{D})\). Therefore, either a 3⩽1 or a 2⩽1 and a 5⩽1.

If a 3⩽1, then it follows from (4) that

If a 2⩽1 and a 5⩽1, then it follows from (4) that

By Lemma 4.4, there exists k∈{2,3,4,5} such that (3) is not Kawamata log terminal at the point E k E k+1 and is Kawamata log terminal outside of E k E k+1.

Put Q=E k E k+1. Then it follows from Lemma 2.3 that

which is impossible by (4), since we assume that either a 3⩽1 or a 2⩽1 and a 5⩽1. □

Lemma 4.11

Suppose that m=7 and R is reducible. Then μ⩾lct2(X)=1/2.

Proof

By Lemma 4.9, the surface X contains an irreducible curve \(\bar{L}_{4}\) such that

and \(-\bar{L}_{4}\cdot \bar{L}_{4}=\bar{L}_{4}\cdot E_{4}=1\). Then \(-K_{\bar{X}}\cdot\bar{L}_{4}=1\), which implies that

Put \(L_{4}=\pi(\bar{L}_{4})\). Then 2L 4∼−2K X and

which implies that lct2(X)⩽c(X,L 4)=1/2.

To complete the proof, it is enough to show that μ⩾1/2.

Suppose that μ<1/2. Let us derive a contradiction.

By Remark 2.1, we may assume that \(L_{4}\not\subset\mathrm{Supp}(D)\). Then

which implies that a 4⩽1. Thus, it follows from (4) that

It follows from Lemma 4.4 that there exists a point

such that \(\mathrm{LCS}(\bar{X},\mu\bar{D}+\sum_{i=1}^{7}\mu a_{i}E_{i})=Q\).

Without loss of generality, we may assume that either Q=E 2E 3 or Q=E 3E 4.

If Q=E 3E 4, then it follows from Lemma 2.3 that

which together with (4) imply that a 4>1, which is a contradiction.

If Q=E 2E 3, then it follows from Lemma 2.3 that

which together with (4) immediately leads to a contradiction. □

Lemma 4.12

Suppose that m=8. Then μ⩾lct3(X)=1/2.

Proof

Arguing as in the proofs of Lemmas 4.7 and 4.8, we see that there is an irreducible smooth rational curve \(\bar{L}_{3}\) on the surface \(\bar{X}\) such that \(\bar{L}_{3}\cdot \bar{L}_{3}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{3}=E_{2}\cdot \bar{L}_{3}=E_{4}\cdot \bar{L}_{3}=E_{5}\cdot \bar{L}_{3}=E_{6}\cdot \bar{L}_{3}=E_{7}\cdot \bar{L}_{3}=\bar{C}\cdot\bar{L}_{3}=0\).

Put \(\bar{L}_{6}=\tau(\bar{L}_{3})\). Then \(\bar{L}_{6}\cdot \bar{L}_{6}=-1\) and \(-K_{\bar{X}}\cdot \bar{L}_{6}=E_{6}\cdot\bar{L}_{6}=1\), which implies that

Put \(L_{3}=\pi(\bar{L}_{3})\) and \(L_{6}=\pi(\bar{L}_{6})\). Then 3L 3∼3L 6∼−3K X . On the other hand, we have

which implies c(X,L 3)=c(X,L 6)=1/2. Then lct3(X)⩽1/2.

To complete the proof, it is enough to show that μ⩾1/2.

Suppose that μ<1/2. Let us derive a contradiction.

By Remark 2.1, we may assume that \(\mathrm{Supp}(\bar{D})\) does not contain \(\bar{L}_{3}\) and \(\bar{L}_{6}\). Then

which implies that a 3⩽1. Similarly, we have a 6⩽1. Then it follows from (4) that

By Lemma 4.4, there exists k∈{2,3,4,5,6} such that (3) is not Kawamata log terminal at the point E k E k+1 and is Kawamata log terminal outside of the point E k E k+1.

Put Q=E k E k+1. Then it follows from Lemma 2.3 that

which is impossible by (4), since a 3⩽1 and a 6⩽1. □

The assertion of Theorem 4.1 is proved.

5 One Non-Cyclic Singular Point

Let X be a sextic surface in ℙ(1,1,2,3) with canonical singularities such that |Sing(X)|=1, and Sing(X) consists of a singular point of type \(\mathbb{D}_{4}\), \(\mathbb{D}_{5}\), \(\mathbb{D}_{6}\), \(\mathbb{D}_{7}\), \(\mathbb{D}_{8}\), \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\).

Theorem 5.1

The following equality holds:

Corollary 5.2

The inequality lct(X)⩽1/2 holds.

In the rest of this section, we will prove Theorem 5.1.

Let D be an effective ℚ-divisor on X such that DK X . We must show that

To prove Theorem 5.1, put μ=c(X,D).

Suppose that μ<lct1(X). Then LCS(X,μD)=Sing(X) by Lemma 2.6. Put P=Sing(X).

Let \(\pi\colon\bar{X}\to X\) be a minimal resolution, let E 1,E 2,…,E m be irreducible π-exceptional curves, let C be the curve in |−K X | such that PC, and let \(\bar{C}\) be its proper transform on \(\bar{X}\). Then

where n i ∈ℕ. Without loss of generality, we may assume that E 3⋅∑ i≠3 E i =3. Then

By Remark 2.1, we may assume that \(C\not\subset\mathrm{Supp}(D)\), since the curve C is irreducible.

Let \(\bar{D}\) be the proper transform of the divisor D on the surface \(\bar{X}\). Then

where a i is a non-negative rational number. Then

which implies that \((\bar{X},\mu\bar{D}+\sum_{i=1}^{m}\mu a_{i}E_{i})\) is not Kawamata log terminal (see Remark 2.4).

Lemma 5.3

The equality μa 3=1 holds.

Proof

The equality μa 3=1 follows from Lemma 2.5. □

Lemma 5.4

Suppose that P is not a point of type \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\). Then

and P is either a point of type \(\mathbb{D}_{7}\) or is a point of type \(\mathbb{D}_{8}\).

Proof

Without loss of generality, we may assume that the diagram

shows how the π-exceptional curves intersect each other. Then

which implies that \(\bar{C}\cdot E_{m-1}=1\) and \(\bar{C}\cdot E_{i}=0\iff i\ne m-1\). Then

$$ \left\{\begin{aligned} &1-a_{m-1}=\bar{D}\cdot \bar{C}\geqslant 0, \\ &2a_1-a_3=\bar{D}\cdot E_{1}\geqslant 0, \\ &2a_2-a_3=\bar{D}\cdot E_{2}\geqslant 0, \\ &2a_3-a_1-a_2-a_3=\bar{D}\cdot E_{3}\geqslant 0, \\[-3pt] &\vdots \\[-3pt] &2a_{m-1}-a_{m-2}-a_m=\bar{D}\cdot E_{m-1}\geqslant 0, \\ &2a_m-a_{m-1}=\bar{D}\cdot E_{m}\geqslant 0, \\ \end{aligned} \right. $$
(6)

which easily implies that a 3⩽2 if m⩽6. But μa 3=1 and μ<lct1(X)=1/2 by Lemma 5.3, which implies that either m=7 or m=8.

Arguing as in the proofs of Lemmas 4.7 and 4.8, we may assume that there is an irreducible smooth rational curve \(\bar{L}_{1}\) on the surface \(\bar{X}\) such that \(\bar{L}_{1}\cdot \bar{L}_{1}=-1\) and

which implies that \(\bar{C}\cdot\bar{L}_{1}=0\) and \(E_{i}\cdot \bar{L}_{1}=0\iff i\ne 1\).

Let ω:X→ℙ(1,1,2) be the natural double cover given by |−2K X |, and let τ be a biregular involution of the surface \(\bar{X}\) that is induced by ω. Put \(\bar{L}_{2}=\tau(\bar{L}_{1})\). If m=7, then

and \(\bar{L}_{2}\cdot \bar{L}_{2}=-1\), which implies that \(\bar{C}\cdot\bar{L}_{2}=0\) and \(E_{i}\cdot \bar{L}_{2}=0\iff i\ne 2\).

Put \(L_{1}=\pi(\bar{L}_{1})\) and \(L_{2}=\pi(\bar{L}_{2})\). Then L 1+L 2∼−2K X . If m=7, then

which implies that c(X,L 1+L 2)=1/5 and lct2(X)⩽2/5. If m=7, then

by (6). But μa 3=1 by Lemma 5.3. Then μ⩾2/5 if m=7, which is exactly what we need.

We may assume that m=8. Then \(\bar{L}_{2}=\bar{L}_{1}\) and

which implies that lct2(X)⩽c(X,L 1)=1/3. But a 3⩽1/3 by (6) and μa 3=1 by Lemma 5.3, which implies that μ⩾1/3, which completes the proof since lct2(X)⩾lct(X). □

To complete the proof of Theorem 5.1, we may assume that P is a point of type \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\).

Without loss of generality, we may assume that the diagram

shows how the π-exceptional curves intersect each other. It is well known (cf. [29, 30]) that

  • if m=6, then \(\bar{C}\cdot E_{4}=1\), which implies that and \(\bar{C}\cdot E_{i}=0\iff i\ne 4\),

  • if m=7, then \(\bar{C}\cdot E_{1}=1\), which implies that and \(\bar{C}\cdot E_{i}=0\iff i\ne 1\),

  • if m=8, then \(\bar{C}\cdot E_{8}=1\), which implies that and \(\bar{C}\cdot E_{i}=0\iff i\ne 8\).

Put k=4 if m=6, put k=1 if m=7, put k=8 if m=8. Then

$$ \left\{\begin{aligned} &1-a_{k}=\bar{D}\cdot \bar{C}\geqslant 0, \\ &2a_1-a_3=\bar{D}\cdot E_{1}\geqslant 0, \\ &2a_2-a_3-a_1=\bar{D}\cdot E_{2} \geqslant 0, \\ &2a_3-a_2-a_4-a_5=\bar{D}\cdot E_{3}\geqslant 0, \\ &2a_4-a_3=\bar{D}\cdot E_{4}\geqslant 0, \\ &2a_5-a_3-a_6=\bar{D}\cdot E_{5}\geqslant 0, \\ &\vdots \\ &2a_{m-1}-a_{m-2}-a_m=\bar{D}\cdot E_{m-1}\geqslant 0, \\ &2a_m-a_{m-1}=\bar{D}\cdot E_{m}\geqslant 0, \\ \end{aligned} \right. $$
(7)

which implies that a 3<n 3. But n 3=1/lct1(X) and μa 3=1 by Lemma 5.3. Then μ⩾lct1(X).

The assertion of Theorem 5.1 is proved.

6 Many Singular Points

Let X be a sextic surface in ℙ(1,1,2,3) with canonical singularities such that |Sing(X)|⩾2.

Theorem 6.1

The following equality holds:

and if there exists an effective ℚ-divisor D on the surface X such that DK X and

then either D is an irreducible curve in |−K X | with a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\), or the divisor D is uniquely defined and it can be explicitly described.

Let D be an arbitrary effective ℚ-divisor on the surface X such that

and put μ=c(X,D). To prove Theorem 6.1, it is enough to show that

and if μ=lct(X)=2/3, then we have the following two possibilities:

  • either D is a curve in |−K X | with a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\),

  • or the divisor D is uniquely defined and it can be explicitly described.

Lemma 6.2

If Sing(X) has a point of type \(\mathbb{D}_{4}\), \(\mathbb{D}_{5}\), \(\mathbb{D}_{6}\), \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\), then μ⩾lct1(X).

Proof

Suppose that Sing(X) has a point of type \(\mathbb{D}_{4}\), \(\mathbb{D}_{5}\), \(\mathbb{D}_{6}\), \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\), but μ<lct1(X). Then

and LCS(X,μD) consists of a point in Sing(X) that is not of type \(\mathbb{A}_{1}\) or \(\mathbb{A}_{2}\) by Lemma 2.6.

If the locus LCS(X,μD) is a singular point of the surface X of type \(\mathbb{D}_{4}\), \(\mathbb{D}_{5}\), \(\mathbb{D}_{6}\), \(\mathbb{E}_{6}\), \(\mathbb{E}_{7}\), or \(\mathbb{E}_{8}\), then arguing as in the proof of Theorem 5.1, we immediately obtain a contradiction.

By Remark 1.21, the locus LCS(X,μD) must be a singular point of the surface X of type \(\mathbb{A}_{3}\), and we can easily obtain a contradiction arguing as in the proof of Corollary 4.5. □

Lemma 6.3

Suppose that Sing(X) consists of points of type \(\mathbb{A}_{1}\), \(\mathbb{A}_{2}\), or \(\mathbb{A}_{3}\). Then μ⩾lct1(X). If

then D is a curve in |−K X | with a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\).

Proof

This follows from Lemma 2.6 and the proof of Corollary 4.5. □

By Remark 1.21 and Lemmas 6.2 and 6.2, we may assume that

which implies that there is a point P∈Sing(X) that is a point of type \(\mathbb{A}_{m}\) for m∈{4,5,6,7}.

Let \(\pi\colon\bar{X}\to X\) be a minimal resolution, let E 1,E 2,…,E m be π-exceptional curves such that

and π(E i )=P for every i∈{1,…,m}, let C be the unique curve in |−K X | such that PC, and let \(\bar{C}\) be the proper transform of the curve C on the surface \(\bar{X}\). Then

and \(\bar{C}\cdot E_{2}=\bar{C}\cdot E_{3}=\cdots=\bar{C}\cdot E_{m-1}=0\). Note that \(\bar{C}\cong\mathbb{P}^{1}\) and \(\bar{C}\cdot\bar{C}=-1\).

Let \(\bar{D}\) be the proper transform of D on the surface \(\bar{X}\). Then

where a i is a non-negative rational number. Then

$$ \left\{\begin{aligned} &1-a_1-a_m= \bar{D}\cdot\bar{C}\geqslant 0, \\ &2a_1 - a_2=\bar{D}\cdot E_{1}\geqslant 0, \\ &\vdots \\ &2a_{m-1}-a_{m-2}-a_m=\bar{D}\cdot E_{m-1}\geqslant 0, \\ &2a_m - a_{m-1}=\bar{D}\cdot E_{m}\geqslant 0. \\ \end{aligned} \right. $$
(8)

Let \(\eta\colon\bar{X}\to\bar{X}^{\prime}\) be a contraction of the curve \(\bar{C}\). Then there is a commutative diagram

where ω and ω′ are natural double covers π′ is a minimal resolution, ϕ is an anticanonical embedding, and ψ is a projection from ϕω(P). Put P′=η(E 2). Then P′∈Sing(X′).

Remark 6.4

The birational morphism π′ contracts the smooth curves η(E 2),η(E 3),…,η(E m−1), and π′∘η contracts all π-exceptional curves that are different from the curves E 1,E 2,…,E m .

Let R be the branch curve in ℙ(1,1,2) of the double cover ω. Put R′=ψϕ(R).

Lemma 6.5

Suppose that m=7. Then μ⩾lct2(X)=1/2.

Proof

Let \(\alpha\colon\bar{X}\to\breve{X}\) be a contraction of the irreducible curves \(\bar{C}\), E 7, E 6, E 5, E 4, E 3, and E 2, and let F be the π-exceptional curve such that π(F) is a point of type \(\mathbb{A}_{1}\). Then

Let \(\breve{L}_{2}\) be the fiber of the projection \(\breve{X}\to\mathbb{P}^{1}\) such that \(\alpha(\bar{C})\in\breve{L}_{2}\), and let \(\bar{L}_{2}\) be the proper transform of the curve \(\breve{L}_{2}\) on the surface \(\bar{X}\) via α. Then \(\bar{L}_{2}\cdot \bar{L}_{2}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{2}=E_{3}\cdot \bar{L}_{2}=E_{4}\cdot \bar{L}_{2}=E_{5}\cdot \bar{L}_{2}=E_{6}\cdot \bar{L}_{2}=E_{7}\cdot \bar{L}_{2}=\bar{C}\cdot\bar{L}_{2}=0\).

Let \(\beta\colon\bar{X}\to\check{X}\) be a contraction of the curves \(\bar{L}_{2}\), E 2, \(\bar{C}\), E 7, E 6, E 5, E 4. Then

and \(\check{X}\) is a smooth del Pezzo surface such that \(K_{\check{X}}^{2}=8\). Then \(\check{X}\cong\mathbb{P}^{1}\times \mathbb{P}^{1}\).

Let \(\check{L}_{4}\) be the curve in |β(F)| such that \(\beta(E_{4})\in \check{L}_{4}\), and let \(\bar{L}_{3}\) be its proper transform on the surface \(\bar{X}\) via β. Then one can easily check that \(\bar{L}_{4}\cdot \bar{L}_{4}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{4}=E_{2}\cdot \bar{L}_{4}=E_{3}\cdot \bar{L}_{4}=E_{5}\cdot \bar{L}_{4}=E_{6}\cdot \bar{L}_{4}=E_{7}\cdot \bar{L}_{4}=\bar{C}\cdot\bar{L}_{4}=F\cdot\bar{L}_{4}=0\).

Put \(L_{4}=\pi(\bar{L}_{4})\). Then one can easily check that

which implies that c(X,L 4)=1/2. But 2L 4∼−2K X , which implies that lct2(X)⩽1/2.

Arguing as in the proof of Lemma 4.9, we see that ω(L 4)⊂Supp(R).

Arguing as in the proof of Lemma 4.11 and using (8), we see that μ⩾lct2(X)=1/2. □

Lemma 6.6

Suppose that m=6. Then μ⩾lct2(X)=2/3, and if μ=2/3, then

  • either D is a curve in |−K X | with a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\),

  • or the divisor D is uniquely defined and can be explicitly described.

Proof

Let \(\alpha\colon\bar{X}\to\breve{X}\) be a contraction of the curves \(\bar{C}\), E 6, E 5, E 4, E 3, E 2. Then \(\breve{X}\) is a smooth surface such that \(K_{\breve{X}}^{2}=7\), and −K X is nef. There is a birational morphism \(\gamma\colon\breve{X}\to\hat{X}\) such that

and γ is a blow-down of a smooth irreducible rational curve that does not contain the point \(\alpha(\bar{C})\).

Let \(\hat{L}_{2}\) be the fiber of the projection \(\hat{X}\to\mathbb{P}^{1}\) such that \(\gamma\circ\alpha(\bar{C})\in\hat{L}_{2}\), and let \(\bar{L}_{2}\) be the proper transform of the curve \(\hat{L}_{2}\) on the surface \(\bar{X}\) via γα. Then \(\bar{L}_{2}\cdot \bar{L}_{2}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{2}=E_{3}\cdot \bar{L}_{2}=E_{4}\cdot \bar{L}_{2}=E_{5}\cdot \bar{L}_{2}=E_{6}\cdot \bar{L}_{2}=\bar{C}\cdot\bar{L}_{2}=0\).

Let \(\beta\colon\bar{X}\to\check{X}\) be a contraction of the curves \(\bar{L}_{2}\), \(\bar{C}\), E 6, E 5, E 4, and let F be the π-exceptional curve such that π(F) is a point of type \(\mathbb{A}_{1}\). Then

and \(\check{X}\) is a smooth del Pezzo surface such that \(K_{\check{X}}^{2}=6\). Thus, there exists an irreducible smooth rational curve \(\check{L}_{3}\) on the surface \(\check{X}\) such that \(\check{L}_{3}\cdot\check{L}_{3}=-1\), \(\check{L}_{3}\cdot\beta(E_{3})=1\), and \(\check{L}_{3}\cdot\beta(F)=0\).

Let \(\bar{L}_{3}\) be the proper transform of the curve \(\check{L}_{3}\) on the surface \(\bar{X}\). Then \(\bar{L}_{3}\cdot \bar{L}_{3}=-1\) and

which implies that \(E_{1}\cdot \bar{L}_{3}=E_{2}\cdot \bar{L}_{3}=E_{4}\cdot \bar{L}_{3}=E_{5}\cdot \bar{L}_{3}=E_{6}\cdot \bar{L}_{3}=\bar{C}\cdot\bar{L}_{3}=F\cdot \bar{L}_{3}=0\).

Put \(\bar{L}_{4}=\tau(\bar{L}_{3})\) and \(\bar{L}_{5}=\tau(\bar{L}_{2})\). Then \(\bar{C}\cdot\bar{L}_{4}=\bar{C}\cdot\bar{L}_{5}=0\) and

which implies that \(E_{i}\cdot\bar{L}_{5}=E_{j}\cdot \bar{L}_{4}=0\) for every i≠5 and j≠4.

Put \(L_{3}=\pi(\bar{L}_{3})\), \(L_{4}=\pi(\bar{L}_{4})\), \(L_{2}=\pi(\bar{L}_{2})\) and \(L_{5}=\pi(\bar{L}_{5})\). Then

which implies that c(X,L 3+L 4)=1/3 and c(X,L 2+L 5)=1/2. Then lct2(X)⩽2/3. But

which implies that c(X,2L 2+L 3)=1/4. Then 2L 2+L 3−3K X , since Pic(X)≅ℤ2 and

but 2L 2+L 3 is a Cartier divisor, which implies that 2L 2+L 3∼−3K X .

If D is not a curve in |−K X | and D≠(L 3+L 4)/2, then arguing as in the proof of Lemma 4.8, we easily see that μ>2/3, since we can use (8). The lemma is proved (see Example 1.26). □

Lemma 6.7

Suppose that m=5. Then μ⩾lct2(X)=2/3, and if μ=2/3, then

  • either D is a curve in |−K X | with a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\),

  • or the divisor D is uniquely defined and can be explicitly described.

Proof

The curve R′ has an ordinary tacnodal singularity at the point ω′(P′), which implies that there exists a line L′⊂ℙ2 such that either L′⊂Supp(R′) or \(L^{\prime}\not\subset\mathrm{Supp}(R^{\prime})\) and

There are irreducible smooth rational curves \(L^{\prime}_{3}\) and \(L^{\prime}_{4}\) on the surface X′ such that

and \(L^{\prime}_{3}=L^{\prime}_{4}\iff L^{\prime}\subset\mathrm{Supp}(R^{\prime})\). Note that neither \(L^{\prime}_{3}\) nor \(L^{\prime}_{4}\) contains a point in Sing(X′)∖R′.

Let \(\bar{L}^{\prime}_{3}\) be the proper transform of the curve \(L^{\prime}_{3}\) on the surface \(\bar{X}^{\prime}\). Then

and \(\bar{L}^{\prime}_{3}\cdot\eta(E_{3})=1\). Let \(\bar{L}^{\prime}_{4}\) be the proper transform of the curve \(L^{\prime}_{4}\) on the surface \(\bar{X}^{\prime}\). Then

and \(\bar{L}^{\prime}_{4}\cdot\eta(E_{3})=1\). One can also check that \(\bar{L}^{\prime}_{3}\cap\bar{L}^{\prime}_{4}=\varnothing\) if \(\bar{L}^{\prime}_{3}\ne\bar{L}^{\prime}_{4}\).

Let \(\bar{L}_{3}\) and \(\bar{L}_{4}\) be the proper transforms of the curves \(\bar{L}^{\prime}_{3}\) and \(\bar{L}^{\prime}_{4}\) on the surface \(\bar{X}\), respectively, and let us put \(L_{3}=\pi(\bar{L}_{3})\) and \(L_{4}=\pi(\bar{L}_{4})\). Then

and \(\mathrm{c}(X, \bar{L}_{3}+\bar{L}_{4})=1/3\), which implies that lct2(X)⩽2/3.

If \(D\ne (\bar{L}_{3}+\bar{L}_{4})/2\), then (8), the proof of Lemma 4.7, and Lemma 2.6 imply that

and if μ=2/3, then D is a curve in |−K X | with a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\). □

Lemma 6.8

Suppose that m=4. Then

and if μ=2/3, then D is a curve in |−K X | with a cusp at a point in Sing(X) of type  \(\mathbb{A}_{2}\).

Proof

The point ω′(P′) is an ordinary cusp of the curve R′. Then there is a line L′⊂ℙ2 such that

Let Z′ be a curve in X′ such that ω′(Z′)=L′ and −K XZ′=2. Then

and Z′ is irreducible curve that has an ordinary cusp at the point R′.

Let \(\bar{Z}^{\prime}\) be the proper transform of the curve Z′ on the surface \(\bar{X}^{\prime}\). Then Z′ is smooth and

Let \(\bar{Z}\) be the proper transform of the curve \(\bar{Z}^{\prime}\) on the surface \(\bar{X}\). Put \(Z=\pi(\bar{Z})\). Then

and E 2E 3Z. Then c(X,Z)=2/5, which implies that lct2(X)⩽4/5.

Arguing as in the proof of Lemma 4.6 and using Lemma 2.6 and (8), we see that

and if μ=2/3, then D is a curve in |−K X | with a cusp at a point in Sing(X) of type \(\mathbb{A}_{2}\). □

The assertion of Theorem 6.1 is proved.