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

Supersingular reduction of Kummer surfaces in residue characteristic 22

Yuya Matsumoto Department of Mathematics, Faculty of Science and Technology, Tokyo University of Science, 2641 Yamazaki, Noda, Chiba, 278-8510, Japan [email protected] [email protected]
(Date: 2023/02/19)
Abstract.

Given an abelian surface AA, defined over a discrete valuation field and having good reduction, does the attached Kummer surface Km⁑(A)\operatorname{Km}(A) also have good reduction? In this paper we give an affirmative answer in the extreme case, that is, when the abelian surface has supersingular reduction in characteristic 22.

2010 Mathematics Subject Classification:
14J28 (Primary) 14L15, 14J17 (Secondary)
This work was supported by JSPS KAKENHI Grant Number 16K17560 and 20K14296.

1. Introduction

Let KK be a Henselian discrete valuation field of characteristic 0 with valuation ring π’ͺ\mathcal{O}, maximal ideal 𝔭\mathfrak{p}, and residue field kk assumed to be perfect. We say that a K3 surface or an abelian surface XX over KK has good reduction if there exists an algebraic space 𝒳\mathcal{X} over π’ͺ\mathcal{O}, proper and smooth, whose generic fiber π’³βŠ—π’ͺK\mathcal{X}\otimes_{\mathcal{O}}K is isomorphic to XX. We say that XX has potential good reduction if there exists a finite extension Kβ€²/KK^{\prime}/K such that the base change XKβ€²X_{K^{\prime}} has good reduction.

The NΓ©ron–Ogg–Shafarevich criterion (Serre–Tate [Serre--Tate]*Theorem 1) shows that good reduction of an abelian variety AA is equivalent to the unramifiedness of the Galois action on its ll-adic cohomology group He´​t1​(AKΒ―,β„šl)H_{\mathrm{\acute{e}t}}^{1}(A_{\overline{K}},\mathbb{Q}_{l}). We ([Matsumoto:goodreductionK3]*Theorem 1.1, [Liedtke--Matsumoto]*Theorem 1.3) showed a similar criterion for K3 surfaces (using He´​t2H_{\mathrm{\acute{e}t}}^{2}), assuming the existence of semistable models with certain properties and the residue characteristic being large enough.

In this paper we focus on Kummer surfaces. For an abelian surface AA over a field FF, the Kummer surface Km⁑(A)\operatorname{Km}(A) is defined to be the minimal resolution of A/{Β±1}A/\{\pm 1\}. It is known that Km⁑(A)\operatorname{Km}(A) is a K3 surface if and only if char⁑Fβ‰ 2\operatorname{char}F\neq 2 or AA is non-supersingular (Proposition 3.1). From this geometric nature of Kummer surfaces, the following question arises naturally.

Question 1.1.

Let AA be an abelian surface over the discrete valuation field KK, and let X=Km⁑(A)X=\operatorname{Km}(A) be the attached Kummer surface, which is a K3 surface since we are assuming char⁑K=0\operatorname{char}K=0. Assume AA has good reduction, with proper smooth model (NΓ©ron model) π’œ\mathcal{A} over π’ͺ\mathcal{O}.

  1. (1)

    Does X=Km⁑(A)X=\operatorname{Km}(A) have potential good reduction?

  2. (2)

    Are smooth proper models of XX related to π’œ\mathcal{A} in a geometric way?

  3. (3)

    Is the special fiber 𝒳k\mathcal{X}_{k} isomorphic to Km⁑(π’œk)\operatorname{Km}(\mathcal{A}_{k})?

Let us say that we are in the extreme case if π’œk\mathcal{A}_{k} is supersingular of characteristic 22, and in the non-extreme case if otherwise (i.e.Β if char⁑kβ‰ 2\operatorname{char}k\neq 2, or char⁑k=2\operatorname{char}k=2 and π’œk\mathcal{A}_{k} is non-supersingular). If we are in the non-extreme cases, then it is known that the all questions have essentially affirmative answers (Proposition 3.4): Roughly, one can blow-up the β€œsingularity” of the quotient π’œ/{Β±1}\mathcal{A}/\{\pm 1\} of the NΓ©ron model π’œ\mathcal{A} of AA to obtain a smooth proper model of X=Km⁑(A)X=\operatorname{Km}(A).

If we are in the extreme case, then the above construction does not give a smooth proper model of X=Km⁑(A)X=\operatorname{Km}(A), mainly because the singularity of π’œk/{Β±1}\mathcal{A}_{k}/\{\pm 1\} is non-rational. In this case Km⁑(π’œk)\operatorname{Km}(\mathcal{A}_{k}) is a rational surface, and (3) cannot hold. What we prove in this paper is that (1) and (2) holds even in this extreme case.

Theorem 1.2.

Let KK and kk be as above. Suppose that char⁑K=0\operatorname{char}K=0 and char⁑k=2\operatorname{char}k=2, that an abelian surface AA over KK has good reduction with NΓ©ron model π’œ\mathcal{A}, and that the special fiber π’œk\mathcal{A}_{k} is supersingular. Then X=Km⁑(A)X=\operatorname{Km}(A) has potential good reduction.

We overview the method of our proof and its consequences. We give (in Section 4.3) a proper birational morphism Ο•:𝒳~β†’π’œ/{Β±1}\phi\colon\tilde{\mathcal{X}}\to\mathcal{A}/\{\pm 1\}, defined as the composite of explicit normalized blow-ups at singular points of the special fiber, from a proper model 𝒳~\tilde{\mathcal{X}} of XX that is strictly semistable in the broad sense (Definition 4.1). Then, according to Theorem 4.2, we can run a relative MMP and obtain a birational map f:𝒳~⇒𝒳f\colon\tilde{\mathcal{X}}\dashrightarrow\mathcal{X} to a smooth proper model 𝒳\mathcal{X} of XX, which achieves good reduction.

The special fiber 𝒳k\mathcal{X}_{k} of 𝒳\mathcal{X} is not isomorphic, nor even birational, to the special fiber π’œk/{Β±1}\mathcal{A}_{k}/\{\pm 1\} of π’œ/{Β±1}\mathcal{A}/\{\pm 1\}, since the latter is a rational surface. Hence the birational map fβˆ˜Ο•βˆ’1f\circ\phi^{-1} (resp.Β its inverse Ο•βˆ˜fβˆ’1\phi\circ f^{-1}) contracts π’œk/{Β±1}\mathcal{A}_{k}/\{\pm 1\} (resp. 𝒳k\mathcal{X}_{k}). We may understand that the β€œK3-ness” of the special fiber π’œ/{Β±1}\mathcal{A}/\{\pm 1\} lies in the local ring at the singular point of π’œk/{Β±1}\mathcal{A}_{k}/\{\pm 1\}, not on the smooth complement of the point in π’œk/{Β±1}\mathcal{A}_{k}/\{\pm 1\}.

Moreover, our construction shows that the special fiber 𝒳k\mathcal{X}_{k} is birational to the quotient of a certain rational surface Aβ€²A^{\prime} by a certain group scheme of length 22 (see Remark 4.7). We can modify Aβ€²A^{\prime} and make it an β€œabelian-like” surface, by which we mean a surface sharing numerical properties with abelian surfaces. Thus 𝒳k\mathcal{X}_{k} can be viewed as an example of an inseparable analogue of Kummer surfaces. We will study this subject in another paper.

The proof of Theorem 4.2 is essentially given in [Matsumoto:goodreductionK3]*Section 3, except that we use Takamatsu–Yoshikawa’s MMP [Takamatsu--Yoshikawa:mixed3fold]*Theorem 1.1 (which is applicable in the case of residue characteristic p=2p=2) in place of Kawamata’s (which requires pβ‰₯5p\geq 5).

2. Preliminaries

2.1. Some elliptic double points

In this subsection, we work over an algebraically closed field kk of characteristic 22.

Definition 2.1.

For (m,n)=(4,4),(2,8),(2,6)(m,n)=(4,4),(2,8),(2,6) and r=0,1r=0,1, we say that k​[[x,y,z]]/(z2+r​z​xm/2​yn/2+xm+1+yn+1)k[[x,y,z]]/(z^{2}+rzx^{m/2}y^{n/2}+x^{m+1}+y^{n+1}) is EDPm,n(r)\mathrm{EDP}_{m,n}^{(r)}.

Remark 2.2.

The following properties of EDPm,n(r)\mathrm{EDP}_{m,n}^{(r)} hold.

  • β€’

    They are all elliptic double points.

  • β€’

    The exceptional divisor of the minimal resolution of EDPm,n(r)\mathrm{EDP}_{m,n}^{(r)} is independent of rr and consists of:

    • –

      one cuspidal rational curve of arithmetic genus 11 for EDP2,6(r)\mathrm{EDP}_{2,6}^{(r)};

    • –

      smooth rational curves, with the dual graph shown in Figure 1, for EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)} and EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}.

    In the figures, a vertex with βˆ’m-m inside denotes a smooth rational curve of self-intersection βˆ’m-m, and the absence of a number denotes self-intersection βˆ’2-2. The symbols \raisebox{-0.9pt}{\small19}⃝0\raisebox{-0.9pt}{\small19}⃝_{0} and \raisebox{-0.9pt}{\small4}⃝0,11\raisebox{-0.9pt}{\small4}⃝_{0,1}^{1} are due to Wagreich [Wagreich:ellipticsingularities]*Theorem 3.8.

  • β€’

    EDPm,n(1)\mathrm{EDP}_{m,n}^{(1)} is a β„€/2​℀\mathbb{Z}/2\mathbb{Z}-quotient singularity by [Artin:wild2]*Theorem, and EDPm,n(0)\mathrm{EDP}_{m,n}^{(0)} is an Ξ±2\alpha_{2}-quotient singularity by [Matsumoto:k3alphap]*Theorem 3.8.

Remark 2.3.

We have alternative equations for EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)} and EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}.

  • β€’

    EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)} is isomorphic to k​[[x,y,z]]/(z2+r​z​x2​(x+y2)+x​(x+y2)2+x4​y)k[[x,y,z]]/(z^{2}+rzx^{2}(x+y^{2})+x(x+y^{2})^{2}+x^{4}y). Indeed, starting from this equation and letting xβ€²=x+y2x^{\prime}=x+y^{2}, zβ€²=z+x′​yz^{\prime}=z+x^{\prime}y, xβ€²=(xβ€²β€²+r​y5)​(1+r​z)x^{\prime}=(x^{\prime\prime}+ry^{5})(1+rz), we obtain z′⁣2​u1+r​z′​x′′​y4​u2+x′′⁣3​u3+y9​u4=0z^{\prime 2}u_{1}+rz^{\prime}x^{\prime\prime}y^{4}u_{2}+x^{\prime\prime 3}u_{3}+y^{9}u_{4}=0 for some units uiu_{i}, and by replacing xβ€²β€²,y,zβ€²x^{\prime\prime},y,z^{\prime} with suitable unit multiples we obtain the equation in Definition 2.1.

  • β€’

    EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)} is also isomorphic to k​[[x,y,z]]/(z2+r​z​x2​y2+x4​y+x​y4)k[[x,y,z]]/(z^{2}+rzx^{2}y^{2}+x^{4}y+xy^{4}). Indeed, starting from the equation of Definition 2.1, letting ΞΆ\zeta be a primitive 55-th root of 11 and v=x+΢​yv=x+\zeta y and w=΢​x+yw=\zeta x+y, we have x5+y5=(ΞΆ2+ΞΆ3)​(v4​w+v​w4)x^{5}+y^{5}=(\zeta^{2}+\zeta^{3})(v^{4}w+vw^{4}), and so on.

The EDPs with (m,n)=(4,4),(2,8)(m,n)=(4,4),(2,8) are relevant to us because of the following. (The remaining one with (m,n)=(2,6)(m,n)=(2,6) is used to describe possible degenerations.)

Proposition 2.4 (Katsura [Katsura:Kummer2]).

Suppose AA is a supersingular abelian surface in characteristic 22. If AA is superspecial (resp.Β non-superspecial), then the singularity of A/{Β±1}A/\{\pm 1\} consists of one point of type EDP4,4(1)\mathrm{EDP}_{4,4}^{(1)} (resp.Β EDP2,8(1)\mathrm{EDP}_{2,8}^{(1)}).

Here, a supersingular abelian surface is called superspecial if it is isomorphic (not only isogenous) to the product of two supersingular elliptic curves.

Proof.

In the superspecial case, this is [Katsura:Kummer2]*Proposition 8. In the non-superspecial case, [Katsura:Kummer2]*Lemma 12 gives an equation similar to the one given in Remark 2.3. ∎

βˆ’3-3 βˆ’3-3
EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}: \raisebox{-0.9pt}{\small19}⃝0\raisebox{-0.9pt}{\small19}⃝_{0} EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)}: \raisebox{-0.9pt}{\small4}⃝0,11\raisebox{-0.9pt}{\small4}⃝_{0,1}^{1}
Figure 1. Dual graphs of minimal resolutions of EDPs

2.2. Quotients by group schemes of length 22 in mixed characteristic

We recall Tate–Oort’s classification [Tate--Oort:groupschemes]*Theorem 2 of finite group schemes GG of length 22. For our purposes it suffices to consider GG over a base ring π’ͺ\mathcal{O} that is a PID (possibly a field). Then, such group schemes are of the form GΞ±,Ξ²G_{\alpha,\beta}, defined as follows, with parameters Ξ±,β∈π’ͺ\alpha,\beta\in\mathcal{O} satisfying α​β=2\alpha\beta=2 (which are denoted by a,ba,b in [Tate--Oort:groupschemes]).

The underlying scheme is GΞ±,Ξ²=Spec⁑RG_{\alpha,\beta}=\operatorname{Spec}R, R=π’ͺ​[X]/(X2βˆ’Ξ±β€‹X)R=\mathcal{O}[X]/(X^{2}-\alpha X). The comultiplication map is given by Rβ†’RβŠ—π’ͺR:X↦XβŠ—1+1βŠ—Xβˆ’Ξ²β€‹XβŠ—XR\to R\otimes_{\mathcal{O}}R\colon X\mapsto X\otimes 1+1\otimes X-\beta X\otimes X. Actions of GΞ±,Ξ²G_{\alpha,\beta} on affine schemes Spec⁑S\operatorname{Spec}S correspond, via Sβ†’RβŠ—π’ͺS:s↦1βŠ—s+XβŠ—Ξ΄β€‹(s)S\to R\otimes_{\mathcal{O}}S\colon s\mapsto 1\otimes s+X\otimes\delta(s), to π’ͺ\mathcal{O}-linear maps Ξ΄:Sβ†’S\delta\colon S\to S satisfying the conditions δ​(s​t)=δ​(s)​t+s​δ​(t)+α​δ​(s)​δ​(t)\delta(st)=\delta(s)t+s\delta(t)+\alpha\delta(s)\delta(t) and δ​(δ​(s))=βˆ’Ξ²β€‹Ξ΄β€‹(s)\delta(\delta(s))=-\beta\delta(s). In particular, id+α​δ\mathrm{id}+\alpha\delta is an automorphism of the π’ͺ\mathcal{O}-algebra SS of order dividing 22.

For a unit e∈π’ͺβˆ—e\in\mathcal{O}^{*}, the group schemes GΞ±,Ξ²G_{\alpha,\beta} and Ge​α,eβˆ’1​βG_{e\alpha,e^{-1}\beta} are isomorphic under the obvious maps.

Example 2.5.

Three important cases of GΞ±,Ξ²G_{\alpha,\beta} are the following.

  • β€’

    If (Ξ±,Ξ²)=(1,2)(\alpha,\beta)=(1,2), then GΞ±,Ξ²G_{\alpha,\beta} is isomorphic to the constant group scheme β„€/2​℀\mathbb{Z}/2\mathbb{Z}. The map Ξ΄\delta is equal to gβˆ’idg-\mathrm{id}, where gg is (the action on SS of) the nontrivial element of β„€/2​℀\mathbb{Z}/2\mathbb{Z}.

  • β€’

    If (Ξ±,Ξ²)=(2,1)(\alpha,\beta)=(2,1), then GΞ±,Ξ²G_{\alpha,\beta} is isomorphic to ΞΌ2\mu_{2}. Actions on Spec⁑S\operatorname{Spec}S also correspond to β„€/2​℀\mathbb{Z}/2\mathbb{Z}-gradings S=⨁iβˆˆβ„€/2​℀SiS=\bigoplus_{i\in\mathbb{Z}/2\mathbb{Z}}S_{i} on the π’ͺ\mathcal{O}-algebra SS, and then Ξ΄\delta is equal to (βˆ’1)(-1) times the projection to S1S_{1}. If moreover 2=02=0 in π’ͺ\mathcal{O}, then the maps Ξ΄\delta are precisely the derivations of multiplicative type.

  • β€’

    If (Ξ±,Ξ²)=(0,0)(\alpha,\beta)=(0,0) (which implies 2=02=0 in π’ͺ\mathcal{O}), then GΞ±,Ξ²G_{\alpha,\beta} is isomorphic to Ξ±2\alpha_{2}. The maps Ξ΄\delta are precisely the derivations of additive type.

If π’ͺ\mathcal{O} is a field of characteristic 22, then any GG is isomorphic to exactly one of above.

If 22 is invertible in π’ͺ\mathcal{O}, then any GG is isomorphic to both β„€/2​℀\mathbb{Z}/2\mathbb{Z} and ΞΌ2\mu_{2}.

The fixed locus Fix⁑(G)\operatorname{Fix}(G) of an action on Spec⁑S\operatorname{Spec}S is the closed subscheme, or the closed subset, corresponding to the ideal of SS generated by Im⁑(δ)\operatorname{Im}(\delta). In other words, its complement is the largest open subscheme where GG acts freely.

Proposition 2.6.

Let π’ͺ\mathcal{O} be a local PID (possibly a field) with maximal ideal 𝔭\mathfrak{p} and residue field kk. Let G=GΞ±,Ξ²G=G_{\alpha,\beta} be a group scheme of length 22 over π’ͺ\mathcal{O}, acting on π’ͺ​[[u,v]]\mathcal{O}[[u,v]], and assume Fix⁑(G)βŠƒ(u=v=0)\operatorname{Fix}(G)\supset(u=v=0) (as subsets of Spec⁑π’ͺ​[[u,v]]\operatorname{Spec}\mathcal{O}[[u,v]]) and Fix⁑(Gkβ†·k​[[u,v]])=(u=v=0)\operatorname{Fix}(G_{k}\curvearrowright k[[u,v]])=(u=v=0) (as subsets of Spec⁑k​[[u,v]]\operatorname{Spec}k[[u,v]]). Then, the elements

x\displaystyle x :=u​(u+α​δ​(u)),\displaystyle:=u(u+\alpha\delta(u)), a\displaystyle a :=β​u+δ​(u),\displaystyle:=\beta u+\delta(u),
y\displaystyle y :=v​(v+α​δ​(v)),\displaystyle:=v(v+\alpha\delta(v)), b\displaystyle b :=β​v+δ​(v),\displaystyle:=\beta v+\delta(v),
z\displaystyle z :=β​u​v+δ​(u)​v+u​δ​(v),\displaystyle:=\beta uv+\delta(u)v+u\delta(v),

are GG-invariant, and the invariant subalgebra π’ͺ​[[u,v]]G\mathcal{O}[[u,v]]^{G} is equal to π’ͺ​[[x,y,z]]/(F)\mathcal{O}[[x,y,z]]/(F), F=z2βˆ’Ξ±β€‹a​b​z+a2​y+b2​xβˆ’Ξ²2​x​yF=z^{2}-\alpha abz+a^{2}y+b^{2}x-\beta^{2}xy.

Suppose char⁑(Frac⁑π’ͺ)=0\operatorname{char}(\operatorname{Frac}\mathcal{O})=0 and char⁑k=2\operatorname{char}k=2. Let Ο„\tau be the Tyurina number of k​[[x,y,z]]/(F)k[[x,y,z]]/(F). If GβŠ—π’ͺk=Ξ±2G\otimes_{\mathcal{O}}k=\alpha_{2}, then Ο„=2​deg⁑Fix⁑(Gβ†·k​[[u,v]])=2​deg⁑Fix⁑(Gβ†·(Frac⁑π’ͺ)βŠ—π’ͺπ’ͺ​[[u,v]])\tau=2\deg\operatorname{Fix}(G\curvearrowright k[[u,v]])=2\deg\operatorname{Fix}(G\curvearrowright(\operatorname{Frac}\mathcal{O})\otimes_{\mathcal{O}}\mathcal{O}[[u,v]]). If GβŠ—π’ͺk=ΞΌ2G\otimes_{\mathcal{O}}k=\mu_{2}, then Ο„=2\tau=2 and deg⁑Fix⁑(Gβ†·k​[[u,v]])=deg⁑Fix⁑(Gβ†·(Frac⁑π’ͺ)βŠ—π’ͺπ’ͺ​[[u,v]])=1\deg\operatorname{Fix}(G\curvearrowright k[[u,v]])=\deg\operatorname{Fix}(G\curvearrowright(\operatorname{Frac}\mathcal{O})\otimes_{\mathcal{O}}\mathcal{O}[[u,v]])=1.

If GΞ±,Ξ²=G1,2=β„€/2​℀G_{\alpha,\beta}=G_{1,2}=\mathbb{Z}/2\mathbb{Z}, then the formulas simplify to x=u​g​(u)x=ug(u), y=v​g​(v)y=vg(v), z=u​g​(v)+g​(u)​vz=ug(v)+g(u)v, a=u+g​(u)a=u+g(u), b=v+g​(v)b=v+g(v), where gg is the non-trivial element of the group.

Proof.

It is straightforward to check that x,y,z,a,bx,y,z,a,b are invariant. Letting zβ€²:=z+α​δ​(u)​δ​(v)z^{\prime}:=z+\alpha\delta(u)\delta(v), we observe z+zβ€²=α​a​bz+z^{\prime}=\alpha ab and z​zβ€²=a2​y+b2​xβˆ’Ξ²2​x​yzz^{\prime}=a^{2}y+b^{2}x-\beta^{2}xy, which implies F​(x,y,z)=0F(x,y,z)=0. Thus it remains to show that x,y,zx,y,z generates π’ͺ​[[u,v]]G\mathcal{O}[[u,v]]^{G}. Since this can be checked modulo the maximal ideals, we may assume that π’ͺ=k\mathcal{O}=k is a field. Then GG is isomorphic to one in Example 2.5 (i.e.Β β„€/2​℀\mathbb{Z}/2\mathbb{Z}, ΞΌ2\mu_{2}, or Ξ±2\alpha_{2}).

Note that in any case the ideal generated by Im⁑(Ξ΄)\operatorname{Im}(\delta) is generated by two elements δ​(u),δ​(v)\delta(u),\delta(v).

Suppose G=ΞΌ2G=\mu_{2} or G=Ξ±2G=\alpha_{2}. Then x=u2,y=v2x=u^{2},y=v^{2}. It follows from [Matsumoto:k3alphap]*Theorem 3.8 that k​[[u,v]]Gk[[u,v]]^{G} has a k​[[x,y]]k[[x,y]]-basis of the form 1,w1,w. Write w=c00+c10​u+c01​v+c11​u​vw=c_{00}+c_{10}u+c_{01}v+c_{11}uv with ci​j∈k​[[u2,v2]]=k​[[x,y]]c_{ij}\in k[[u^{2},v^{2}]]=k[[x,y]]. Write z=d+e​wz=d+ew with d,e∈k​[[x,y]]d,e\in k[[x,y]]. It suffices to show e∈k​[[x,y]]βˆ—e\in k[[x,y]]^{*}. We have z2=d2+e2​w2=d2+e2​c002+e2​c102​x+e2​c012​y+e2​c112​x​yz^{2}=d^{2}+e^{2}w^{2}=d^{2}+e^{2}c_{00}^{2}+e^{2}c_{10}^{2}x+e^{2}c_{01}^{2}y+e^{2}c_{11}^{2}xy. We also have z2=b2​x+a2​yβˆ’Ξ²2​x​yz^{2}=b^{2}x+a^{2}y-\beta^{2}xy. It suffices to show that a2,b2,Ξ²2{a^{2},b^{2},\beta^{2}} have no common divisor. By the assumption on the fixed locus, the ideal (δ​(u),δ​(v))=(a,b)(\delta(u),\delta(v))=(a,b) is supported on the closed point, and so is the ideal (a2,b2)(a^{2},b^{2}).

Now suppose G=β„€/2​℀G=\mathbb{Z}/2\mathbb{Z}. In this case x=N​(u)x=N(u), y=N​(v)y=N(v), and z=Tr⁑(u​g​(v))z=\operatorname{Tr}(ug(v)), and it follows from Lemma 2.7 that x,y,zx,y,z generates π’ͺ​[[u,v]]G\mathcal{O}[[u,v]]^{G}.

Latter assertion. Suppose Gk=Ξ±2G_{k}=\alpha_{2}. Then we have deg⁑Fix⁑(Gβ†·k​[[u,v]])=deg⁑k​[[u,v]]/(a,b)\deg\operatorname{Fix}(G\curvearrowright k[[u,v]])=\deg k[[u,v]]/(a,b), and Ο„=dimkk​[[x,y,z]]/(a2,b2,z2)=2​dimkk​[[x,y]]/(a2,b2)=2​dimkk​[[u,v]]/(a,b)\tau=\dim_{k}k[[x,y,z]]/(a^{2},b^{2},z^{2})=2\dim_{k}k[[x,y]]/(a^{2},b^{2})=2\dim_{k}k[[u,v]]/(a,b) since x=u2x=u^{2} and y=v2y=v^{2} in this case.

Suppose Gk=ΞΌ2G_{k}=\mu_{2}. In this case we can linearize the action and then δ​(u)=u\delta(u)=u and δ​(v)=v\delta(v)=v, hence deg⁑Fix⁑(Gβ†·k​[[u,v]])=1\deg\operatorname{Fix}(G\curvearrowright k[[u,v]])=1. ∎

Lemma 2.7.

Suppose kk is a field of characteristic 22, and G=β„€/2​℀={1,g}G=\mathbb{Z}/2\mathbb{Z}=\{1,g\} acts on k​[[u,v]]k[[u,v]], freely outside the closed point. Let x=N​(u)=u​g​(u)x=N(u)=ug(u), y=N​(v)=v​g​(v)y=N(v)=vg(v), and z=Tr⁑(u​g​(v))=u​g​(v)+g​(u)​vz=\operatorname{Tr}(ug(v))=ug(v)+g(u)v. Then k​[[u,v]]G=k​[[x,y,z]]/(z2+a​b​z+a2​y+b2​x)k[[u,v]]^{G}=k[[x,y,z]]/(z^{2}+abz+a^{2}y+b^{2}x), where a=Tr⁑(u)=u+g​(u)a=\operatorname{Tr}(u)=u+g(u) and b=Tr⁑(v)=v+g​(v)b=\operatorname{Tr}(v)=v+g(v).

Proof.

It is straightforward to check z2+a​b​z+a2​y+b2​x=0z^{2}+abz+a^{2}y+b^{2}x=0 (cf.Β the first paragraph of the proof of Proposition 2.6). It remains to show that x,y,zx,y,z generate the maximal ideal 𝔫′\mathfrak{n}^{\prime} of k​[[u,v]]Gk[[u,v]]^{G}.

By [Artin:wild2]*Theorem, there exist local coordinates uβ€²,vβ€²u^{\prime},v^{\prime} with the following properties: letting xβ€²=N​(uβ€²)x^{\prime}=N(u^{\prime}), yβ€²=N​(vβ€²)y^{\prime}=N(v^{\prime}), zβ€²=Tr⁑(u′​g​(vβ€²))z^{\prime}=\operatorname{Tr}(u^{\prime}g(v^{\prime})), and R0=k​[[xβ€²,yβ€²]]R_{0}=k[[x^{\prime},y^{\prime}]],

  • β€’

    k​[[u,v]]k[[u,v]] has an R0R_{0}-basis 1,uβ€²,vβ€²,u′​g​(vβ€²)1,u^{\prime},v^{\prime},u^{\prime}g(v^{\prime}) and

  • β€’

    k​[[u,v]]Gk[[u,v]]^{G} has an R0R_{0}-basis 1,zβ€²1,z^{\prime}.

Write

(uv)=(p00p10p01p11q00q10q01q11)​(1uβ€²vβ€²u′​g​(vβ€²)),\begin{pmatrix}u\\ v\end{pmatrix}=\begin{pmatrix}p_{00}&p_{10}&p_{01}&p_{11}\\ q_{00}&q_{10}&q_{01}&q_{11}\end{pmatrix}\begin{pmatrix}1\\ u^{\prime}\\ v^{\prime}\\ u^{\prime}g(v^{\prime})\end{pmatrix},

where pi​j,qi​j∈R0p_{ij},q_{ij}\in R_{0}, p00,q00∈(xβ€²,yβ€²)p_{00},q_{00}\in(x^{\prime},y^{\prime}), and det(p10p01q10q01)∈R0βˆ—\det\begin{pmatrix}p_{10}&p_{01}\\ q_{10}&q_{01}\end{pmatrix}\in R_{0}^{*}. Then a straightforward computation yields

(xyz)≑(p102p012p10​p01q102q012q10​q0100p10​q01+p01​q10)​(xβ€²yβ€²zβ€²)(mod𝔫′⁣2).\begin{pmatrix}x\\ y\\ z\end{pmatrix}\equiv\begin{pmatrix}p_{10}^{2}&p_{01}^{2}&p_{10}p_{01}\\ q_{10}^{2}&q_{01}^{2}&q_{10}q_{01}\\ 0&0&p_{10}q_{01}+p_{01}q_{10}\end{pmatrix}\begin{pmatrix}x^{\prime}\\ y^{\prime}\\ z^{\prime}\end{pmatrix}\pmod{\mathfrak{n}^{\prime 2}}.

Since the matrix is invertible, x,y,zx,y,z generate 𝔫′/𝔫′⁣2\mathfrak{n}^{\prime}/\mathfrak{n}^{\prime 2}. ∎

3. Non-extreme cases

The good reduction problem of Kummer surfaces in the non-extreme cases are solved by Ito (see [Matsumoto:SIP]*Section 4) if char⁑kβ‰ 2\operatorname{char}k\neq 2, and by Lazda–Skorobogatov [Lazda--Skorobogatov:reductionofkummer] if char⁑k=2\operatorname{char}k=2 and π’œk\mathcal{A}_{k} is non-supersingular. For the readers’ convenience, we include a sketch of the proof for these cases.

3.1. Singularities of Kummer surfaces in the non-extreme cases

We already mentioned the result for the extreme case in Proposition 2.4. In the non-extreme cases, we have the following.

Proposition 3.1.

Let AA be an abelian surface over a field FF. If char⁑Fβ‰ 2\operatorname{char}F\neq 2, or char⁑F=2\operatorname{char}F=2 and AA is non-supersingular, then the minimal resolution of A/{Β±1}A/\{\pm 1\} is a K3 surface. Moreover, if all points of A​[2]A[2] are FF-rational, then A/{Β±1}A/\{\pm 1\} has only RDPs as singularities, and the number and the type of RDPs are

  • β€’

    16​A116A_{1} if char⁑Fβ‰ 2\operatorname{char}F\neq 2,

  • β€’

    4​D414D_{4}^{1} if char⁑F=2\operatorname{char}F=2 and pβˆ’rank⁑(A)=2\operatorname{\mathit{p}-rank}(A)=2,

  • β€’

    2​D822D_{8}^{2} if char⁑F=2\operatorname{char}F=2 and pβˆ’rank⁑(A)=1\operatorname{\mathit{p}-rank}(A)=1.

For the assertion on quotient singularities, see [Katsura:Kummer2]*Proposition 3 for a proof (for the determination of the coindices, see [Artin:wild2]*Examples).

3.2. Good reduction of Kummer surfaces in the non-extreme cases

Let K,π’ͺ,kK,\mathcal{O},k be as in Introduction.

Proposition 3.2.

Suppose π’œβ†’Spec⁑π’ͺ\mathcal{A}\to\operatorname{Spec}\mathcal{O} is an abelian scheme over π’ͺ\mathcal{O}. Then the fibers of π’œ/{Β±1}\mathcal{A}/\{\pm 1\} are isomorphic to the quotients of the fibers.

Proof.

See [Lazda--Skorobogatov:reductionofkummer]*Proposition 4.6. ∎

Proposition 3.3.

Suppose 𝒳→Spec⁑π’ͺ\mathcal{X}\to\operatorname{Spec}\mathcal{O} is a flat morphism of finite type, with both fibers 𝒳k\mathcal{X}_{k} and 𝒳K\mathcal{X}_{K} normal surfaces. Suppose π’΅βŠ‚π’³\mathcal{Z}\subset\mathcal{X} is a section (i.e. the composite 𝒡→𝒳→Spec⁑π’ͺ\mathcal{Z}\to\mathcal{X}\to\operatorname{Spec}\mathcal{O} is an isomorphism) such that 𝒡k\mathcal{Z}_{k} and 𝒡K\mathcal{Z}_{K} are RDPs of the fibers. Then Bl𝒡(𝒳)Kβ‰…Bl𝒡K(𝒳K)\operatorname{Bl}_{\mathcal{Z}}(\mathcal{X})_{K}\cong\operatorname{Bl}_{\mathcal{Z}_{K}}(\mathcal{X}_{K}) and Bl𝒡(𝒳)kβ‰…Bl𝒡k(𝒳k)\operatorname{Bl}_{\mathcal{Z}}(\mathcal{X})_{k}\cong\operatorname{Bl}_{\mathcal{Z}_{k}}(\mathcal{X}_{k}).

Proof.

See [Lazda--Skorobogatov:reductionofkummer]*Proposition 4.1 or [Overkamp:Kummer]*Lemma 3.10. ∎

Proposition 3.4.

Let KK and kk be as in the introduction (with char⁑K=0\operatorname{char}K=0). Suppose an abelian surface AA over KK has good reduction with NΓ©ron model π’œ\mathcal{A}. Assume either char⁑kβ‰ 2\operatorname{char}k\neq 2, or char⁑k=2\operatorname{char}k=2 and π’œk\mathcal{A}_{k} is non-supersingular. Then X=Km⁑(A)X=\operatorname{Km}(A) has potential good reduction.

Proof.

Let 𝒴:=π’œ/{Β±1}\mathcal{Y}:=\mathcal{A}/\{\pm 1\}. By Proposition 3.2, we have 𝒴K=A/{Β±1}\mathcal{Y}_{K}=A/\{\pm 1\} and 𝒴k=Ak/{Β±1}\mathcal{Y}_{k}=A_{k}/\{\pm 1\}. By replacing KK with a finite extension, we may assume that π’œK​[2]\mathcal{A}_{K}[2] and π’œk​[2]\mathcal{A}_{k}[2] consist respectively of KK-rational points and kk-rational points. Then, by Proposition 3.1, Sing⁑(𝒴K)\operatorname{Sing}(\mathcal{Y}_{K}) is 16​A116A_{1} and Sing⁑(𝒴k)\operatorname{Sing}(\mathcal{Y}_{k}) is one of 16​A116A_{1}, 4​D414D_{4}^{1}, or 2​D822D_{8}^{2}.

Choose an RDP of 𝒴K\mathcal{Y}_{K}, take its closure 𝒡\mathcal{Z} in 𝒴\mathcal{Y}, and take the blow-up 𝒴(1)=Bl𝒡⁑𝒴\mathcal{Y}^{(1)}=\operatorname{Bl}_{\mathcal{Z}}\mathcal{Y}. Then 𝒡k\mathcal{Z}_{k} should be a singular point of 𝒴k\mathcal{Y}_{k}, hence one of the RDPs. By Proposition 3.3, each fiber of 𝒴(1)\mathcal{Y}^{(1)} is the blow-up at an RDP. Repeating this 1616 times, we obtain a scheme 𝒴(16)\mathcal{Y}^{(16)} whose fibers are Km⁑(π’œK)\operatorname{Km}(\mathcal{A}_{K}) and Km⁑(π’œk)\operatorname{Km}(\mathcal{A}_{k}). ∎

4. Proof of main theorem

Let π’ͺ\mathcal{O} be as before (i.e.Β a Henselian discrete valuation ring with residue field kk). In this section, kk is of characteristic 22, and the maximal ideal of π’ͺ\mathcal{O} is denoted by 𝔭=(Ο–)\mathfrak{p}=(\varpi).

We first show in Section 4.1 that, in order to prove good reduction of a K3 surface XX, it suffices to construct a model of XX satisfying a kind of semistability. Then we give an explicit construction of such a model in Sections 4.2–4.3. We summarize the proof of Theorem 1.2 in Section 4.4.

4.1. Strict semistability in the broad sense

We need the following variant of the strict semistability.

Definition 4.1.

We say that a flat scheme 𝒳\mathcal{X} of finite type over π’ͺ\mathcal{O} is strictly semistable in the broad sense if it satisfies the following conditions.

  • β€’

    The generic fiber XX of 𝒳\mathcal{X} is smooth.

  • β€’

    Every irreducible component of 𝒳k\mathcal{X}_{k} is smooth.

  • β€’

    For each closed point PP of the special fiber 𝒳k\mathcal{X}_{k}, π’ͺ𝒳,P\mathcal{O}_{\mathcal{X},P} is Γ©tale over π’ͺ​[x1,…,xn]/(x1​…​xrβˆ’Ο΅)\mathcal{O}[x_{1},\dots,x_{n}]/(x_{1}\dots x_{r}-\epsilon) for some 1≀r≀n1\leq r\leq n and some Ο΅βˆˆπ”­\epsilon\in\mathfrak{p}.

Recall that 𝒳\mathcal{X} is strictly semistable if we can take Ο΅\epsilon to be a uniformizer of π’ͺ\mathcal{O}.

The existence of such a model is quite useful for the good reduction problem:

Theorem 4.2.

Let XX be a K3 surface over KK. Suppose that the Galois representation He´​t2​(XKΒ―,β„šl)H_{\mathrm{\acute{e}t}}^{2}(X_{\overline{K}},\mathbb{Q}_{l}) is unramified for some auxiliary prime lβ‰ pl\neq p after replacing KK with a finite extension, and that XX admits a proper model 𝒳\mathcal{X} over π’ͺ\mathcal{O} that is strictly semistable in the broad sense. Then XX has potential good reduction.

Here, a representation of Gal⁑(K¯/K)\operatorname{Gal}(\overline{K}/K) is unramified if the inertia subgroup of Gal⁑(K¯/K)\operatorname{Gal}(\overline{K}/K) acts trivially.

Proof.

By replacing KK with a finite extension (which does not affect the semistability in the broad sense), we may assume that He´​t2​(XKΒ―,β„šl)H_{\mathrm{\acute{e}t}}^{2}(X_{\overline{K}},\mathbb{Q}_{l}) is unramified.

By an argument similar to [Saito:logsmooth]*Lemma 1.7, 𝒳\mathcal{X} is log smooth over π’ͺ\mathcal{O}. By [Saito:logsmooth]*Theorem 1.8, there exist a finite extension π’ͺβ€²\mathcal{O}^{\prime} of π’ͺ\mathcal{O} and a morphism π’³β€²β†’π’³βŠ—π’ͺπ’ͺβ€²\mathcal{X}^{\prime}\to\mathcal{X}\otimes_{\mathcal{O}}\mathcal{O}^{\prime} that is isomorphic above the interior of 𝒳\mathcal{X} such that 𝒳′\mathcal{X}^{\prime} is proper and strictly semistable over π’ͺβ€²\mathcal{O}^{\prime}. Here the interior of 𝒳\mathcal{X} is the complement of the union of intersections of two or more components of 𝒳k\mathcal{X}_{k} (hence the interior contains the generic fiber). Thus we may assume that 𝒳\mathcal{X} is strictly semistable.

Then we can apply the arguments given in [Matsumoto:goodreductionK3]*Section 3, with the use of Kawamata’s MMP replaced with Takamatsu–Yoshikawa’s [Takamatsu--Yoshikawa:mixed3fold]*Theorem 1.1, which, unlike Kawamata’s, has no restriction on the residue characteristic. ∎

If X=Km⁑(A)X=\operatorname{Km}(A) is the Kummer surface associated to an abelian surface AA with good reduction, then, by the NΓ©ron–Ogg–Shafarevich criterion, the Galois representation He´​t1​(AKΒ―,β„šl)H_{\mathrm{\acute{e}t}}^{1}(A_{\overline{K}},\mathbb{Q}_{l}) is unramified, and then by the isomorphism

He´​t2(Km(A)KΒ―,β„šl)≅⨁A​[2]β„šl(βˆ’1)βŠ•β‹€2He´​t1(AKΒ―,β„šl),H_{\mathrm{\acute{e}t}}^{2}(\operatorname{Km}(A)_{\overline{K}},\mathbb{Q}_{l})\cong\bigoplus_{A[2]}\mathbb{Q}_{l}(-1)\oplus\bigwedge^{2}H_{\mathrm{\acute{e}t}}^{1}(A_{\overline{K}},\mathbb{Q}_{l}),

He´​t2(Km(A)KΒ―,β„šl)H_{\mathrm{\acute{e}t}}^{2}(\operatorname{Km}(A)_{\overline{K}},\mathbb{Q}_{l}) is also unramified after replacing KK with a finite extension. Thus, our goal is to construct a strictly semistable model in the broad sense of the Kummer surface X=Km⁑(A)X=\operatorname{Km}(A).

4.2. Overview of the blow-up construction

The main idea is to perform a (weighted) blow-up with respect to suitable coordinates.

Suppose 𝒴\mathcal{Y} is a smooth affine scheme over π’ͺ\mathcal{O} of relative dimension 22 and G=GΞ±,Ξ²G=G_{\alpha,\beta} is a group scheme of length 22 (see Section 2.2). Suppose GG acts on 𝒴\mathcal{Y}, with the quotient morphism denoted by Ο€:𝒴→𝒴/G=𝒳\pi\colon\mathcal{Y}\to\mathcal{Y}/G=\mathcal{X}, and the following properties hold.

  • β€’

    Fix⁑(Gk↷𝒴k)={y}\operatorname{Fix}(G_{k}\curvearrowright\mathcal{Y}_{k})=\{y\}, and the quotient singularity π​(y)βˆˆπ’³k\pi(y)\in\mathcal{X}_{k} is either EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)} or EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)}.

  • β€’

    All points of Fix⁑(GK↷𝒴K)\operatorname{Fix}(G_{K}\curvearrowright\mathcal{Y}_{K}) specializes to yy.

Let 𝒴′→𝒴\mathcal{Y}^{\prime}\to\mathcal{Y} be the weighted blow-up specified below, with center supported on yy. The GG-action extends to 𝒴′\mathcal{Y}^{\prime}. Let 𝒳′:=𝒴′/G\mathcal{X}^{\prime}:=\mathcal{Y}^{\prime}/G. We list the properties of 𝒴′\mathcal{Y}^{\prime} and 𝒳′\mathcal{X}^{\prime} to be established. The construction and the proof will be given in Section 4.3.

Claim 4.3.

The following holds.

  1. (1)

    𝒳′\mathcal{X}^{\prime} is a normalized (weighted) blow-up of 𝒳\mathcal{X}.

  2. (2)

    𝒳kβ€²\mathcal{X}^{\prime}_{k} consists of two components: the strict transform ZZ of 𝒳k\mathcal{X}_{k} and the exceptional divisor EE.

  3. (3)

    The GG-action on π’΄β€²βˆ–Ο€βˆ’1​(Z)\mathcal{Y}^{\prime}\setminus\pi^{-1}(Z) induces an action of Gβ€²=GΞ±β€²,Ξ²β€²G^{\prime}=G_{\alpha^{\prime},\beta^{\prime}} with isolated fixed locus, and the quotient is π’³β€²βˆ–Z\mathcal{X}^{\prime}\setminus Z.

  4. (4)

    The exceptional divisor EE is a normal projective surface and satisfies H1​(π’ͺE)=0H^{1}(\mathcal{O}_{E})=0 and KE=0K_{E}=0. In particular, if Sing⁑(E)\operatorname{Sing}(E) consists only of RDPs, then EE is an RDP K3 surface.

  5. (5)

    The image of Sing⁑(𝒳Kβ€²)\operatorname{Sing}(\mathcal{X}^{\prime}_{K}) by the specialization map is contained in Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z.

Here, an RDP K3 surface is a proper surface with only rational double point singularities (if any) whose minimal resolution is a K3 surface.

Claim 4.4.

After replacing kk with a finite separable extension, we have the following description of Sing⁑(Z)\operatorname{Sing}(Z) and Sing⁑(E)\operatorname{Sing}(E).

  1. (1)

    If yy is EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}, then there are P1,…,P5∈Z∩EP_{1},\dots,P_{5}\in Z\cap E and the following hold.

    1. (a)

      Sing⁑(Z)={P1,…,P5}\operatorname{Sing}(Z)=\{P_{1},\dots,P_{5}\}, all RDPs of type A1A_{1}.

    2. (b)

      Sing⁑(E)∩Z={P1,…,P5}\operatorname{Sing}(E)\cap Z=\{P_{1},\dots,P_{5}\}, all RDPs of type A1A_{1}.

    3. (c)

      Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z is one of 16​A116A_{1}, 4​D404D_{4}^{0}, 2​D802D_{8}^{0}, 2​E802E_{8}^{0}, 1​D1601D_{16}^{0}, or 1​E​D​P2,8(0)1\mathrm{EDP}_{2,8}^{(0)}.

  2. (2)

    If yy is EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)}, then there are Q1,…,Q4∈Z∩EQ_{1},\dots,Q_{4}\in Z\cap E and the following hold.

    1. (a)

      Sing⁑(Z)={Q1,…,Q4}\operatorname{Sing}(Z)=\{Q_{1},\dots,Q_{4}\}, Q1Q_{1} is a quotient singularity of type (1,1)/3{(1,1)}/{3}, and all others are RDPs of type A1A_{1}.

    2. (b)

      Sing⁑(E)∩Z={Q1,…,Q4}\operatorname{Sing}(E)\cap Z=\{Q_{1},\dots,Q_{4}\}, Q1Q_{1} is an RDP of type A2A_{2}, and all others are RDPs of type A1A_{1}.

    3. (c)

      Every element of Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z is one of the following: A1A_{1}, D40D_{4}^{0}, D80D_{8}^{0}, D120D_{12}^{0}, D160D_{16}^{0}, E80E_{8}^{0}, or EDP2,6(0)\mathrm{EDP}_{2,6}^{(0)}.

    4. (d)

      If all singularities of Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z are of the same type, then Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z is one of 16​A116A_{1}, 4​D404D_{4}^{0}, 2​D802D_{8}^{0}, 2​E802E_{8}^{0}, or 1​D1601D_{16}^{0}.

Let 𝒳′′→𝒳′\mathcal{X}^{\prime\prime}\to\mathcal{X}^{\prime} be the blow-up specified below (with center supported on Sing⁑(Z)\operatorname{Sing}(Z)).

Claim 4.5.

𝒳′′\mathcal{X}^{\prime\prime} is strictly semistable in the broad sense outside Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z.

Remark 4.6.

For our purpose (Theorem 1.2), what we need directly are resolutions of EDPs of type EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)} and EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)} with r=1r=1. However we also consider the case r=0r=0 because we need the resolution of EDP2,8(0)\mathrm{EDP}_{2,8}^{(0)} for the proof of the case of EDP4,4(1)\mathrm{EDP}_{4,4}^{(1)}. The arguments for r=0r=0 and r=1r=1 are parallel.

Remark 4.7.

Our construction also shows that the (affine) surface Eβˆ–ZE\setminus Z is the quotient of Ο€βˆ’1​(Eβˆ–Z)\pi^{-1}(E\setminus Z) by Gβ€²βˆˆ{ΞΌ2,Ξ±2}G^{\prime}\in\{\mu_{2},\alpha_{2}\}. Although we will not show it in this paper, this can be extended to a quotient morphism Aβ€²β†’EA^{\prime}\to E by Gβ€²G^{\prime} from a proper non-normal surface Aβ€²A^{\prime} that satisfies Ο‰Aβ€²β‰…π’ͺAβ€²\omega_{A^{\prime}}\cong\mathcal{O}_{A^{\prime}} and hi​(π’ͺAβ€²)=1,2,1h^{i}(\mathcal{O}_{A^{\prime}})=1,2,1 for i=0,1,2i=0,1,2. We say that such Aβ€²A^{\prime} is an abelian-like surface and EE is an inseparable analogue of Kummer surface. Details will be given in another paper.

4.3. Explicit computation of the blow-up

4.3.1. Case of EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}

First we consider the case where yβˆˆπ’΄k/Gy\in\mathcal{Y}_{k}/G is EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}. Let u,v∈π’ͺ𝒴,yu,v\in\mathcal{O}_{\mathcal{Y},y} be local coordinates satisfying (u=v=0)βŠ‚Fix⁑(G↷𝒴)(u=v=0)\subset\operatorname{Fix}(G\curvearrowright\mathcal{Y}). Let x,y,z,a,b∈π’ͺ𝒴,yGx,y,z,a,b\in\mathcal{O}_{\mathcal{Y},y}^{G} as in Proposition 2.6. Then, by the proposition, the maximal ideal of π’ͺ𝒴,yG\mathcal{O}_{\mathcal{Y},y}^{G} is equal to (x,y,z)+𝔭(x,y,z)+\mathfrak{p}, and we have F=0F=0, where

F=z2βˆ’Ξ±β€‹z​a​b+a2​y+b2​xβˆ’Ξ²2​x​y.F=z^{2}-\alpha zab+a^{2}y+b^{2}x-\beta^{2}xy.

Since (u=v=0)βŠ‚Fix⁑(G)(u=v=0)\subset\operatorname{Fix}(G), we have a,b∈(x,y,z)a,b\in(x,y,z) (not only a,b∈(x,y,z)+𝔭a,b\in(x,y,z)+\mathfrak{p}). By the description of the blow-up of EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)} at the origin, we have aΒ―,b¯∈π”ͺ2\overline{a},\overline{b}\in\mathfrak{m}^{2} (where π”ͺ=(x,y,z)\mathfrak{m}=(x,y,z)) and moreover

a2​y+b2​x¯≑c50​x5+c41​x4​y+c14​x​y4+c05​y5(mod((x,y)6+(z)))\overline{a^{2}y+b^{2}x}\equiv c_{50}x^{5}+c_{41}x^{4}y+c_{14}xy^{4}+c_{05}y^{5}\pmod{((x,y)^{6}+(z))}

for some ci​j∈kc_{ij}\in k. Here the overline denotes the reduction modulo 𝔭\mathfrak{p}. By a coordinate change (on u,vu,v), and after replacing kk with a finite separable extension, we may assume a¯≑x2\overline{a}\equiv x^{2}, b¯≑y2(mod((x,y)3+(z)))\overline{b}\equiv y^{2}\pmod{((x,y)^{3}+(z))}, and furthermore

a\displaystyle a ≑x2+a10​x+a01​y,\displaystyle\equiv x^{2}+a_{10}x+a_{01}y,
b\displaystyle b ≑y2+b10​x+b01​y(mod((x,y)3+(z)))\displaystyle\equiv y^{2}+b_{10}x+b_{01}y\pmod{((x,y)^{3}+(z))}

with ai​j,bi​jβˆˆπ”­a_{ij},b_{ij}\in\mathfrak{p}. After replacing π’ͺ\mathcal{O} with a finite extension if necessary, we take Ο΅βˆˆπ”­\epsilon\in\mathfrak{p} satisfying the following condition:

a10Ο΅2,a01Ο΅2,b10Ο΅2,b01Ο΅2,Ξ²2Ο΅6∈π’ͺ,and moreover at least one∈π’ͺβˆ—.\frac{a_{10}}{\epsilon^{2}},\frac{a_{01}}{\epsilon^{2}},\frac{b_{10}}{\epsilon^{2}},\frac{b_{01}}{\epsilon^{2}},\frac{\beta^{2}}{\epsilon^{6}}\in\mathcal{O},\quad\text{and moreover at least one}\in\mathcal{O}^{*}.

Let 𝒴′→𝒴\mathcal{Y}^{\prime}\to\mathcal{Y} be the blow-up at the ideal (u,v,Ο΅)(u,v,\epsilon). Then 𝒳′:=𝒴′/G\mathcal{X}^{\prime}:=\mathcal{Y}^{\prime}/G is the normalized weighted blow-up at (x,y,z,Ο΅)(x,y,z,\epsilon) with respect to the weight (2,2,5,1)(2,2,5,1).

We have 𝒳′=𝒳ϡ′βˆͺ𝒳xβ€²βˆͺ𝒳yβ€²\mathcal{X}^{\prime}=\mathcal{X}^{\prime}_{\epsilon}\cup\mathcal{X}^{\prime}_{x}\cup\mathcal{X}^{\prime}_{y}, where 𝒳?β€²=Spec⁑A?\mathcal{X}^{\prime}_{?}=\operatorname{Spec}A_{?} are the affine open subschemes of the blow-up 𝒳′\mathcal{X}^{\prime} with the obvious meaning. We first consider 𝒳ϡ′=Spec⁑AΟ΅\mathcal{X}^{\prime}_{\epsilon}=\operatorname{Spec}A_{\epsilon}, whose special fiber is equal to Eβˆ–ZE\setminus Z. The π’ͺ\mathcal{O}-algebra AΟ΅A_{\epsilon} is Γ©tale over the subalgebra generated by

xβ€²:=xΟ΅2,yβ€²:=yΟ΅2,zβ€²:=zΟ΅5,x^{\prime}:=\frac{x}{\epsilon^{2}},\quad y^{\prime}:=\frac{y}{\epsilon^{2}},\quad z^{\prime}:=\frac{z}{\epsilon^{5}},

subject to Fβ€²=0F^{\prime}=0, where Fβ€²F^{\prime} is naively β€œΟ΅βˆ’10​F\epsilon^{-10}F”, i.e.,

Fβ€²=z′⁣2βˆ’Ο΅3​α​a′​b′​zβ€²+a′⁣2​yβ€²+b′⁣2​xβ€²βˆ’(Ο΅βˆ’6​β2)​x′​yβ€²F^{\prime}=z^{\prime 2}-\epsilon^{3}\alpha a^{\prime}b^{\prime}z^{\prime}+a^{\prime 2}y^{\prime}+b^{\prime 2}x^{\prime}-(\epsilon^{-6}\beta^{2})x^{\prime}y^{\prime}

where aβ€²:=Ο΅βˆ’4​aa^{\prime}:=\epsilon^{-4}a and bβ€²:=Ο΅βˆ’4​bb^{\prime}:=\epsilon^{-4}b. Let ai​jβ€²=Ο΅βˆ’2​ai​j¯∈ka_{ij}^{\prime}=\overline{\epsilon^{-2}a_{ij}}\in k and bi​jβ€²=Ο΅βˆ’2​bi​j¯∈kb_{ij}^{\prime}=\overline{\epsilon^{-2}b_{ij}}\in k. Here, again, the overline denotes the reduction modulo 𝔭\mathfrak{p}. We have AΟ΅βŠ—π’ͺk=k​[xβ€²,yβ€²,zβ€²]/(Fβ€²Β―)A_{\epsilon}\otimes_{\mathcal{O}}k=k[x^{\prime},y^{\prime},z^{\prime}]/(\overline{F^{\prime}}), Fβ€²Β―=z′⁣2+a′⁣2​yβ€²+b′⁣2​xβ€²βˆ’Ο΅βˆ’6​β2¯​x′​yβ€²\overline{F^{\prime}}=z^{\prime 2}+a^{\prime 2}y^{\prime}+b^{\prime 2}x^{\prime}-\overline{\epsilon^{-6}\beta^{2}}x^{\prime}y^{\prime}, aβ€²=x′⁣2+a10′​xβ€²+a01′​yβ€²a^{\prime}=x^{\prime 2}+a^{\prime}_{10}x^{\prime}+a^{\prime}_{01}y^{\prime}, bβ€²=y′⁣2+b10′​xβ€²+b01′​yβ€²b^{\prime}=y^{\prime 2}+b^{\prime}_{10}x^{\prime}+b^{\prime}_{01}y^{\prime}.

Suppose Ο΅βˆ’6​β2∈π’ͺβˆ—\epsilon^{-6}\beta^{2}\in\mathcal{O}^{*}, hence its image Ο΅βˆ’6​β2¯∈k\overline{\epsilon^{-6}\beta^{2}}\in k is nonzero. Then since Fβ€²Β―x′​yβ€²=Ο΅βˆ’6​β2Β―β‰ 0\overline{F^{\prime}}_{x^{\prime}y^{\prime}}=\overline{\epsilon^{-6}\beta^{2}}\neq 0, every singularity of Eβˆ–ZE\setminus Z is of type A1A_{1}, and considering the degrees of Fβ€²Β―xβ€²\overline{F^{\prime}}_{x^{\prime}} and Fβ€²Β―yβ€²\overline{F^{\prime}}_{y^{\prime}} we conclude that the singularity of Eβˆ–ZE\setminus Z is 16​A116A_{1}.

Suppose Ο΅βˆ’6​β2βˆˆπ”­\epsilon^{-6}\beta^{2}\in\mathfrak{p}. Then we have

aβ€²Β―\displaystyle\overline{a^{\prime}} =x′⁣2+a10′​xβ€²+a01′​yβ€²,\displaystyle=x^{\prime 2}+a^{\prime}_{10}x^{\prime}+a^{\prime}_{01}y^{\prime},
bβ€²Β―\displaystyle\overline{b^{\prime}} =y′⁣2+b10′​xβ€²+b01′​yβ€²\displaystyle=y^{\prime 2}+b^{\prime}_{10}x^{\prime}+b^{\prime}_{01}y^{\prime}

with ai​jβ€²,bi​jβ€²βˆˆka^{\prime}_{ij},b^{\prime}_{ij}\in k, not all 0 by the definition of Ο΅\epsilon. Then

Fβ€²Β―=z′⁣2+x′⁣4​yβ€²+x′​y′⁣4+b10′⁣2​x′⁣3+a10′⁣2​x′⁣2​yβ€²+b01′⁣2​x′​y′⁣2+a01′⁣2​y′⁣3.\overline{F^{\prime}}=z^{\prime 2}+x^{\prime 4}y^{\prime}+x^{\prime}y^{\prime 4}+b_{10}^{\prime 2}x^{\prime 3}+a_{10}^{\prime 2}x^{\prime 2}y^{\prime}+b_{01}^{\prime 2}x^{\prime}y^{\prime 2}+a_{01}^{\prime 2}y^{\prime 3}.

To describe the singularities, we consider c:=(b10β€²:a10β€²:b01β€²:a01β€²)βˆˆβ„™3c:=(b^{\prime}_{10}:a^{\prime}_{10}:b^{\prime}_{01}:a^{\prime}_{01})\in\mathbb{P}^{3} and the following subsets of β„™3\mathbb{P}^{3}:

Q\displaystyle Q ={(t0:t1:t2:t3)∣t1t2βˆ’t0t3=0},\displaystyle=\{(t_{0}:t_{1}:t_{2}:t_{3})\mid t_{1}t_{2}-t_{0}t_{3}=0\},
Ξ“1\displaystyle\Gamma_{1} ={(t0:t1:t2:t3)∣t1t2βˆ’t0t3=t12βˆ’t0t2=t22βˆ’t1t3=0},\displaystyle=\{(t_{0}:t_{1}:t_{2}:t_{3})\mid t_{1}t_{2}-t_{0}t_{3}=t_{1}^{2}-t_{0}t_{2}=t_{2}^{2}-t_{1}t_{3}=0\},
Ξ“2\displaystyle\Gamma_{2} ={(t0:t1:t2:t3)∣t1t2βˆ’t0t3=t22βˆ’t0t1=t12βˆ’t2t3=0}.\displaystyle=\{(t_{0}:t_{1}:t_{2}:t_{3})\mid t_{1}t_{2}-t_{0}t_{3}=t_{2}^{2}-t_{0}t_{1}=t_{1}^{2}-t_{2}t_{3}=0\}.

QQ is the image of the Segre embedding of β„™1Γ—β„™1\mathbb{P}^{1}\times\mathbb{P}^{1} to β„™3\mathbb{P}^{3}, and Ξ“1\Gamma_{1} and Ξ“2\Gamma_{2} are twisted cubic curves contained in QQ. Then it is straightforward to check that the singularity of Eβˆ–ZE\setminus Z is as shown in Figure 2 in the following sense: a configuration of singularities (e.g.Β 2​D802D_{8}^{0}) occurs if and only if cc belongs to the corresponding locus (QQ) but not to its subspaces (Ξ“1\Gamma_{1}, Ξ“2\Gamma_{2}, and Ξ“1βˆ©Ξ“2\Gamma_{1}\cap\Gamma_{2}). We also note that Ξ“1βˆ©Ξ“2=Ξ“1​(𝔽4)=Ξ“2​(𝔽4)\Gamma_{1}\cap\Gamma_{2}=\Gamma_{1}(\mathbb{F}_{4})=\Gamma_{2}(\mathbb{F}_{4}).

4​D40{4D_{4}^{0}}β„™3{\mathbb{P}^{3}}2​D80{2D_{8}^{0}}Q=β„™1Γ—β„™1{Q=\mathbb{P}^{1}\times\mathbb{P}^{1}}2​E80{2E_{8}^{0}}1​D160{1D_{16}^{0}}Ξ“1{\Gamma_{1}}Ξ“2{\Gamma_{2}}1​E​D​P2,8(0){1\mathrm{EDP}_{2,8}^{(0)}}Ξ“1βˆ©Ξ“2{\Gamma_{1}\cap\Gamma_{2}}
Figure 2. Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z in the case of EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}, assuming Ο΅βˆ’6​β2βˆˆπ”­\epsilon^{-6}\beta^{2}\in\mathfrak{p}

Consider the assertion on the action (Claim 4.3(3)). The action of G=GΞ±,Ξ²G=G_{\alpha,\beta} on 𝒴\mathcal{Y} corresponds to the map Ξ΄:π’ͺ𝒴→π’ͺ𝒴\delta\colon\mathcal{O}_{\mathcal{Y}}\to\mathcal{O}_{\mathcal{Y}} as in Section 2.2. Since δ​(u)=a\delta(u)=a has weight β‰₯4=3+wt⁑(u)\geq 4=3+\operatorname{wt}(u) and δ​(v)=b\delta(v)=b has weight β‰₯4=3+wt⁑(v)\geq 4=3+\operatorname{wt}(v) (i.e.Β a,b∈(u,v)4a,b\in(u,v)^{4}), it extends to Ξ΄:π’ͺ𝒴′→π’ͺ𝒴′​(βˆ’3​E𝒴)\delta\colon\mathcal{O}_{\mathcal{Y}^{\prime}}\to\mathcal{O}_{\mathcal{Y}^{\prime}}(-3E_{\mathcal{Y}}), where Eπ’΄βŠ‚π’΄β€²E_{\mathcal{Y}}\subset\mathcal{Y}^{\prime} is the exceptional divisor. On π’΄β€²βˆ–Ο€βˆ’1​(Z)\mathcal{Y}^{\prime}\setminus\pi^{-1}(Z), the subsheaf π’ͺπ’΄β€²βˆ–Ο€βˆ’1​(Z)​(βˆ’3​E𝒴)βŠ‚π’ͺπ’΄β€²βˆ–Ο€βˆ’1​(Z)\mathcal{O}_{\mathcal{Y}^{\prime}\setminus\pi^{-1}(Z)}(-3E_{\mathcal{Y}})\subset\mathcal{O}_{\mathcal{Y}^{\prime}\setminus\pi^{-1}(Z)} is equal to Ο΅3​π’ͺπ’΄β€²βˆ–Ο€βˆ’1​(Z)\epsilon^{3}\mathcal{O}_{\mathcal{Y}^{\prime}\setminus\pi^{-1}(Z)}, hence Ξ΄β€²:=Ο΅βˆ’3​δ\delta^{\prime}:=\epsilon^{-3}\delta defines a morphism π’ͺπ’΄β€²βˆ–Ο€βˆ’1​(Z)β†’π’ͺπ’΄β€²βˆ–Ο€βˆ’1​(Z)\mathcal{O}_{\mathcal{Y}^{\prime}\setminus\pi^{-1}(Z)}\to\mathcal{O}_{\mathcal{Y}^{\prime}\setminus\pi^{-1}(Z)} and it corresponds to an action of Gβ€²=GΞ±β€²,Ξ²β€²G^{\prime}=G_{\alpha^{\prime},\beta^{\prime}}, (Ξ±β€²,Ξ²β€²)=(Ο΅3​α,Ο΅βˆ’3​β)(\alpha^{\prime},\beta^{\prime})=(\epsilon^{3}\alpha,\epsilon^{-3}\beta). Since δ′​(Ο΅βˆ’1​u)=Ο΅βˆ’4​a=aβ€²\delta^{\prime}(\epsilon^{-1}u)=\epsilon^{-4}a=a^{\prime} and δ′​(Ο΅βˆ’1​v)=bβ€²\delta^{\prime}(\epsilon^{-1}v)=b^{\prime}, the fixed locus of the action of Gβ€²G^{\prime} on 𝒴ϡ′\mathcal{Y}^{\prime}_{\epsilon} is (aβ€²Β―=bβ€²Β―=0)(\overline{a^{\prime}}=\overline{b^{\prime}}=0). This descrpition also shows that the image under the specialization map of Sing⁑(XK)\operatorname{Sing}(X_{K}) is equal to Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z.

Next we consider singularities of ZZ and EE contained in Z∩EZ\cap E, which is contained in 𝒳xβ€²βˆͺ𝒳yβ€²\mathcal{X}^{\prime}_{x}\cup\mathcal{X}^{\prime}_{y}. By symmetry, it suffices to consider 𝒳xβ€²=Spec⁑Ax\mathcal{X}^{\prime}_{x}=\operatorname{Spec}A_{x}. Define elements of AxA_{x} by

yβ€²=yx,zβ€²=zx2,zβ€²β€²=ϡ​zx3,q=Ο΅2x,aβ€²=ax2,bβ€²=bx2.y^{\prime}=\frac{y}{x},\quad z^{\prime}=\frac{z}{x^{2}},\quad z^{\prime\prime}=\frac{\epsilon z}{x^{3}},\quad q=\frac{\epsilon^{2}}{x},\quad a^{\prime}=\frac{a}{x^{2}},\quad b^{\prime}=\frac{b}{x^{2}}.

Also let

s=z2x5=z′⁣2x=α​a′​b′​z′​xβˆ’(b′⁣2+a′⁣2​yβ€²βˆ’q3​(Ο΅βˆ’6​β2)​yβ€²).s=\frac{z^{2}}{x^{5}}=\frac{z^{\prime 2}}{x}=\alpha a^{\prime}b^{\prime}z^{\prime}x-(b^{\prime 2}+a^{\prime 2}y^{\prime}-q^{3}(\epsilon^{-6}\beta^{2})y^{\prime}).

Then the π’ͺ\mathcal{O}-algebra AxA_{x} is Γ©tale over the subalgebra generated by x,yβ€²,zβ€²,zβ€²β€²,qx,y^{\prime},z^{\prime},z^{\prime\prime},q, subject to

(szβ€²zβ€²β€²xΟ΅q)∈V2:={(T02T0​T1T0​T2T12T1​T2T22)},\begin{pmatrix}s\\ z^{\prime}\quad z^{\prime\prime}\\ x\quad\epsilon\quad q\end{pmatrix}\in V_{2}:=\biggl{\{}\begin{pmatrix}T_{0}^{2}\\ T_{0}T_{1}\quad T_{0}T_{2}\\ T_{1}^{2}\quad T_{1}T_{2}\quad T_{2}^{2}\end{pmatrix}\biggr{\}},

where V2βŠ‚π”Έ6V_{2}\subset\mathbb{A}^{6} is the cone over the image of the second Veronese map β„™2β†’β„™5\mathbb{P}^{2}\to\mathbb{P}^{5}. From the above description, it is straightforward to check the following assertions:

  • β€’

    𝒳x∩Z=(Ο–=q=zβ€²β€²=0)\mathcal{X}_{x}\cap Z=(\varpi=q=z^{\prime\prime}=0),

  • β€’

    𝒳x∩E=(Ο–=x=zβ€²=0)\mathcal{X}_{x}\cap E=(\varpi=x=z^{\prime}=0),

  • β€’

    𝒳x∩Sing⁑(Z)∩E=𝒳x∩Sing⁑(E)∩Z=𝒳x∩E∩Z∩(s=0)\mathcal{X}_{x}\cap\operatorname{Sing}(Z)\cap E=\mathcal{X}_{x}\cap\operatorname{Sing}(E)\cap Z=\mathcal{X}_{x}\cap E\cap Z\cap(s=0),

  • β€’

    Sing⁑(Z)∩E=Sing⁑(E)∩Z\operatorname{Sing}(Z)\cap E=\operatorname{Sing}(E)\cap Z consists of 55 points {P1,…,P5}\{P_{1},\dots,P_{5}\} on which (x:y)βˆˆβ„™1(𝔽4)(x:y)\in\mathbb{P}^{1}(\mathbb{F}_{4}),

  • β€’

    each PiP_{i} is an RDP of type A1A_{1} in both ZZ and EE,

Define 𝒳′′→𝒳′\mathcal{X}^{\prime\prime}\to\mathcal{X}^{\prime} to be the blow-up at the ideal (s,zβ€²β€²,x,zβ€²,Ο΅,q)(s,z^{\prime\prime},x,z^{\prime},\epsilon,q) over 𝒳xβ€²\mathcal{X}^{\prime}_{x}, and similar over 𝒳yβ€²\mathcal{X}^{\prime}_{y}. Then 𝒳′′\mathcal{X}^{\prime\prime} is strictly semistable in the broad sense above each PiP_{i}. Each exceptional divisor of 𝒳′′→𝒳′\mathcal{X}^{\prime\prime}\to\mathcal{X}^{\prime} is β„™2\mathbb{P}^{2}, and the local equations at the triple points are, for example, zβ€²s​zβ€²β€²s​sβˆ’Ο΅\frac{z^{\prime}}{s}\frac{z^{\prime\prime}}{s}s-\epsilon.

By construction, EE is a hypersurface of degree 10=5+2+2+110=5+2+2+1 in ℙ​(5,2,2,1)\mathbb{P}(5,2,2,1), hence we obtain H1​(π’ͺE)=0H^{1}(\mathcal{O}_{E})=0 and KE=0K_{E}=0.

Eβˆ–ZE\setminus Z is the quotient of Eπ’΄βˆ–Ο€βˆ’1​(Z)E_{\mathcal{Y}}\setminus\pi^{-1}(Z) by Gkβ€²=GΞ±β€²Β―,Ξ²β€²Β―=G0,Ξ²β€²Β―G^{\prime}_{k}=G_{\overline{\alpha^{\prime}},\overline{\beta^{\prime}}}=G_{0,\overline{\beta^{\prime}}}, which is isomorphic to either ΞΌ2\mu_{2} or Ξ±2\alpha_{2} (respectively if Ξ²β€²Β―β‰ 0\overline{\beta^{\prime}}\neq 0 or Ξ²β€²Β―=0\overline{\beta^{\prime}}=0).

4.3.2. Case of EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)}

Next we consider the case where yy is EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)}.

Let u,v,x,y,z,a,b,Fu,v,x,y,z,a,b,F as in the previous case. According to the equation given in Definition 2.1, we may assume

aΒ―\displaystyle\overline{a} ≑y4(mod(x,y5,z)),\displaystyle\equiv y^{4}\pmod{(x,y^{5},z)},
bΒ―\displaystyle\overline{b} ≑x(mod(x2,x​y,y3,z)).\displaystyle\equiv x\pmod{(x^{2},xy,y^{3},z)}.

We can assume moreover (by coordinate changes like uβ€²=u+Ο΅1​vu^{\prime}=u+\epsilon_{1}v and vβ€²=v+Ο΅2​yv^{\prime}=v+\epsilon_{2}y for Ο΅1,Ο΅2βˆˆπ”­\epsilon_{1},\epsilon_{2}\in\mathfrak{p}) that

a\displaystyle a ≑y4+a3​y3+a2​y2+a1​y,\displaystyle\equiv y^{4}+a_{3}y^{3}+a_{2}y^{2}+a_{1}y,
b\displaystyle b ≑x+b1​y(mod(x2,x​y,y5,z)),\displaystyle\equiv x+b_{1}y\pmod{(x^{2},xy,y^{5},z)},

a1,a2,a3,b1βˆˆπ”­a_{1},a_{2},a_{3},b_{1}\in\mathfrak{p}. We take Ο΅βˆˆπ”­\epsilon\in\mathfrak{p} satisfying the following condition (again, after replacing π’ͺ\mathcal{O} if necessary):

aiΟ΅8βˆ’2​i,biΟ΅6βˆ’2​i,Ξ²2Ο΅10∈π’ͺ,and moreover at least one∈π’ͺβˆ—.\frac{a_{i}}{\epsilon^{8-2i}},\frac{b_{i}}{\epsilon^{6-2i}},\frac{\beta^{2}}{\epsilon^{10}}\in\mathcal{O},\quad\text{and moreover at least one}\in\mathcal{O}^{*}.

Let 𝒴′→𝒴\mathcal{Y}^{\prime}\to\mathcal{Y} be the weighted blow-up at the ideal (u,v,Ο΅)(u,v,\epsilon) with respect to the weight (3,1,1)(3,1,1). Then 𝒳′:=𝒴′/G\mathcal{X}^{\prime}:=\mathcal{Y}^{\prime}/G is the normalized weighted blow-up at (x,y,z,Ο΅)(x,y,z,\epsilon) with respect to the weight (6,2,9,1)(6,2,9,1).

We have 𝒳′=𝒳ϡ′βˆͺ𝒳xβ€²βˆͺ𝒳yβ€²\mathcal{X}^{\prime}=\mathcal{X}^{\prime}_{\epsilon}\cup\mathcal{X}^{\prime}_{x}\cup\mathcal{X}^{\prime}_{y}. We first consider 𝒳ϡ′=Spec⁑AΟ΅\mathcal{X}^{\prime}_{\epsilon}=\operatorname{Spec}A_{\epsilon}, whose special fiber is equal to Eβˆ–ZE\setminus Z. The π’ͺ\mathcal{O}-algebra AΟ΅A_{\epsilon} is Γ©tale over the subalgebra generated by

xβ€²:=xΟ΅6,yβ€²:=yΟ΅2,zβ€²:=zΟ΅9,x^{\prime}:=\frac{x}{\epsilon^{6}},\quad y^{\prime}:=\frac{y}{\epsilon^{2}},\quad z^{\prime}:=\frac{z}{\epsilon^{9}},

subject to Fβ€²=0F^{\prime}=0, where

Fβ€²=z′⁣2βˆ’Ο΅5​a′​b′​zβ€²+a′⁣2​yβ€²+b′⁣2​xβ€²βˆ’(Ο΅βˆ’10​β2)​x′​yβ€²F^{\prime}=z^{\prime 2}-\epsilon^{5}a^{\prime}b^{\prime}z^{\prime}+a^{\prime 2}y^{\prime}+b^{\prime 2}x^{\prime}-(\epsilon^{-10}\beta^{2})x^{\prime}y^{\prime}

is naively β€œΟ΅βˆ’18​F\epsilon^{-18}F”, where aβ€²:=Ο΅βˆ’8​aa^{\prime}:=\epsilon^{-8}a and bβ€²:=Ο΅βˆ’6​bb^{\prime}:=\epsilon^{-6}b. Let aiβ€²=Ο΅βˆ’(8βˆ’2​i)​ai¯∈ka_{i}^{\prime}=\overline{\epsilon^{-(8-2i)}a_{i}}\in k and biβ€²=Ο΅βˆ’(6βˆ’2​i)​bi¯∈kb_{i}^{\prime}=\overline{\epsilon^{-(6-2i)}b_{i}}\in k. We have AΟ΅βŠ—k=k​[xβ€²,yβ€²,zβ€²]/(Fβ€²Β―)A_{\epsilon}\otimes k=k[x^{\prime},y^{\prime},z^{\prime}]/(\overline{F^{\prime}}), Fβ€²Β―=z′⁣2+a′⁣2​yβ€²+b′⁣2​xβ€²βˆ’Ο΅βˆ’10​β2¯​x′​yβ€²\overline{F^{\prime}}=z^{\prime 2}+a^{\prime 2}y^{\prime}+b^{\prime 2}x^{\prime}-\overline{\epsilon^{-10}\beta^{2}}x^{\prime}y^{\prime}, aβ€²=y′⁣4+a3′​y′⁣3+a2′​y′⁣2+a1′​y′⁣1a^{\prime}=y^{\prime 4}+a^{\prime}_{3}y^{\prime 3}+a^{\prime}_{2}y^{\prime 2}+a^{\prime}_{1}y^{\prime 1}, bβ€²=xβ€²+b1′​yβ€²b^{\prime}=x^{\prime}+b^{\prime}_{1}y^{\prime}.

If Ξ²2β€‹Ο΅βˆ’10∈π’ͺβˆ—\beta^{2}\epsilon^{-10}\in\mathcal{O}^{*} then, as in the case of EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}, we observe that Sing⁑(E)βˆ–Z=16​A1\operatorname{Sing}(E)\setminus Z=16A_{1}.

Suppose Ο΅βˆ’10​β2βˆˆπ”­\epsilon^{-10}\beta^{2}\in\mathfrak{p}. Then we have

aβ€²Β―\displaystyle\overline{a^{\prime}} =y′⁣4+a3′​y′⁣3+a2′​y′⁣2+a1′​yβ€²,\displaystyle=y^{\prime 4}+a^{\prime}_{3}y^{\prime 3}+a^{\prime}_{2}y^{\prime 2}+a^{\prime}_{1}y^{\prime},
bβ€²Β―\displaystyle\overline{b^{\prime}} =xβ€²+b1′​yβ€²\displaystyle=x^{\prime}+b^{\prime}_{1}y^{\prime}

with aiβ€²,biβ€²βˆˆka^{\prime}_{i},b^{\prime}_{i}\in k, not all 0 by the definition of Ο΅\epsilon. Then

Fβ€²Β―=z′⁣2+x′⁣3+b1′⁣2​x′​y′⁣2+a1′⁣2​y′⁣3+a2′⁣2​y′⁣5+a3′⁣2​y′⁣7+y′⁣9.\overline{F^{\prime}}=z^{\prime 2}+x^{\prime 3}+b_{1}^{\prime 2}x^{\prime}y^{\prime 2}+a_{1}^{\prime 2}y^{\prime 3}+a_{2}^{\prime 2}y^{\prime 5}+a_{3}^{\prime 2}y^{\prime 7}+y^{\prime 9}.

Then the singularity at the origin is as in Figure 3.

D40{D_{4}^{0}}(){()}D80{D_{8}^{0}}(a1β€²=0){(a^{\prime}_{1}=0)}D120{D_{12}^{0}}E80{E_{8}^{0}}(a1β€²=a2β€²=0){(a^{\prime}_{1}=a^{\prime}_{2}=0)}(a1β€²=b1β€²=0){(a^{\prime}_{1}=b^{\prime}_{1}=0)}D160{D_{16}^{0}}EDP2,6(0){\mathrm{EDP}_{2,6}^{(0)}}(a1β€²=a2β€²=a3β€²=0){(a^{\prime}_{1}=a^{\prime}_{2}=a^{\prime}_{3}=0)}(a1β€²=a2β€²=b1β€²=0){(a^{\prime}_{1}=a^{\prime}_{2}=b^{\prime}_{1}=0)}
Figure 3. Singularity of EE at the origin in the case of EDP2,8(r)\mathrm{EDP}_{2,8}^{(r)}, assuming Ο΅βˆ’10​β2βˆˆπ”­\epsilon^{-10}\beta^{2}\in\mathfrak{p}

Again we have dimk​[xβ€²,yβ€²,zβ€²]/IΟ„=32\dim k[x^{\prime},y^{\prime},z^{\prime}]/I_{\tau}=32, where IΟ„=(a′⁣2,b′⁣2,z′⁣2)I_{\tau}=(a^{\prime 2},b^{\prime 2},z^{\prime 2}) is the Tyurina ideal. If all singularities are of the same type, then we conclude that Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z is one of 4​D40,2​D80,2​E80,1​D1604D_{4}^{0},2D_{8}^{0},2E_{8}^{0},1D_{16}^{0}. In particular, D120D_{12}^{0} and EDP2,6(0)\mathrm{EDP}_{2,6}^{(0)} do not appear under this assumption.

The assertion on the extension of the action of GG is checked similarly to the case of EDP4,4(r)\mathrm{EDP}_{4,4}^{(r)}. Since δ​(u)=a\delta(u)=a has weight β‰₯8=5+wt⁑(u)\geq 8=5+\operatorname{wt}(u) and δ​(v)=b\delta(v)=b has weight β‰₯6=5+wt⁑(v)\geq 6=5+\operatorname{wt}(v), the morphism Ξ΄:π’ͺ𝒴→π’ͺ𝒴\delta\colon\mathcal{O}_{\mathcal{Y}}\to\mathcal{O}_{\mathcal{Y}} extends to Ξ΄:π’ͺ𝒴′→π’ͺ𝒴′​(βˆ’5​E)\delta\colon\mathcal{O}_{\mathcal{Y}^{\prime}}\to\mathcal{O}_{\mathcal{Y}^{\prime}}(-5E). We argue similarly (with 33 replaced with 55).

Next we consider singularities of ZZ and EE contained in Z∩EZ\cap E, which is contained in 𝒳xβ€²βˆͺ𝒳yβ€²\mathcal{X}^{\prime}_{x}\cup\mathcal{X}^{\prime}_{y}. We check 𝒳xβ€²=Spec⁑Ax\mathcal{X}^{\prime}_{x}=\operatorname{Spec}A_{x} and omit 𝒳yβ€²\mathcal{X}^{\prime}_{y} (which does not give any more singularities). Define elements of AxA_{x} by

ti=yi​ϡ6βˆ’2​ix​(i=0,1,2,3),wi=z​yi​ϡ3βˆ’2​ix2​(i=0,1),v=z​y2x2,zβ€²=zx,\displaystyle t_{i}=\frac{y^{i}\epsilon^{6-2i}}{x}\;(i=0,1,2,3),\quad w_{i}=\frac{zy^{i}\epsilon^{3-2i}}{x^{2}}\;(i=0,1),\quad v=\frac{zy^{2}}{x^{2}},\quad z^{\prime}=\frac{z}{x},
aβ€²=ax,bβ€²=bx,s=z2x3=α​z′​a′​bβ€²βˆ’(b′⁣2+y​a′⁣2xβˆ’Ξ²2Ο΅10​t1​t0).\displaystyle a^{\prime}=\frac{a}{x},\quad b^{\prime}=\frac{b}{x},\quad s=\frac{z^{2}}{x^{3}}=\alpha z^{\prime}a^{\prime}b^{\prime}-(b^{\prime 2}+\frac{ya^{\prime 2}}{x}-\frac{\beta^{2}}{\epsilon^{10}}t_{1}t_{0}).

Then the π’ͺ\mathcal{O}-algebra AxA_{x} is Γ©tale over the subalgebra generated by

x,y,t0,t1,t2,t3,w0,w1,v,zβ€²,x,y,t_{0},t_{1},t_{2},t_{3},w_{0},w_{1},v,z^{\prime},

subject to

rank⁑(w0Ο΅2t0t1t2w1yt1t2t3)≀1,rank⁑(w1Ο΅t0t1t2vyϡ​t1ϡ​t2ϡ​t3)≀1,\displaystyle\operatorname{rank}\begin{pmatrix}w_{0}&\epsilon^{2}&t_{0}&t_{1}&t_{2}\\ w_{1}&y&t_{1}&t_{2}&t_{3}\\ \end{pmatrix}\leq 1,\operatorname{rank}\begin{pmatrix}w_{1}&\epsilon&t_{0}&t_{1}&t_{2}\\ v&y&\epsilon t_{1}&\epsilon t_{2}&\epsilon t_{3}\\ \end{pmatrix}\leq 1,
rank⁑(w0Ο΅3t0zβ€²xΟ΅3)≀1,rank⁑(w1ϡ​yt1zβ€²xΟ΅3)≀1,rank⁑(vy2t2t3zβ€²xΟ΅2y)≀1,\displaystyle\operatorname{rank}\begin{pmatrix}w_{0}&\epsilon^{3}&t_{0}\\ z^{\prime}&x&\epsilon^{3}\\ \end{pmatrix}\leq 1,\operatorname{rank}\begin{pmatrix}w_{1}&\epsilon y&t_{1}\\ z^{\prime}&x&\epsilon^{3}\\ \end{pmatrix}\leq 1,\operatorname{rank}\begin{pmatrix}v&y^{2}&t_{2}&t_{3}\\ z^{\prime}&x&\epsilon^{2}&y\\ \end{pmatrix}\leq 1,

and

(sw0w1zβ€²vt0t1t2Ο΅3ϡ​t2y​ϡϡ​t3xy2y​t3)∈V4\displaystyle\begin{pmatrix}s\\ w_{0}\quad w_{1}\quad z^{\prime}\quad v\\ t_{0}\quad t_{1}\quad t_{2}\quad\epsilon^{3}\quad\epsilon t_{2}\quad y\epsilon\quad\epsilon t_{3}\quad x\quad y^{2}\quad yt_{3}\end{pmatrix}\in V_{4}
:={(T02T0​T1T0​T2T0​T3T0​T4T12T1​T2T22T1​T3T1​T4T2​T3T2​T4T32T3​T4T42)},\displaystyle:=\biggl{\{}\begin{pmatrix}T_{0}^{2}\\ T_{0}T_{1}\quad T_{0}T_{2}\quad T_{0}T_{3}\quad T_{0}T_{4}\\ T_{1}^{2}\quad T_{1}T_{2}\quad T_{2}^{2}\quad T_{1}T_{3}\quad T_{1}T_{4}\quad T_{2}T_{3}\quad T_{2}T_{4}\quad T_{3}^{2}\quad T_{3}T_{4}\quad T_{4}^{2}\end{pmatrix}\biggr{\}},

where V4βŠ‚π”Έ15V_{4}\subset\mathbb{A}^{15} is the cone over the image of the second Veronese map β„™4β†’β„™14\mathbb{P}^{4}\to\mathbb{P}^{14}.

We have 𝒳x∩Z=(Ο–=t0=t1=t2=w0=w1=0)\mathcal{X}_{x}\cap Z=(\varpi=t_{0}=t_{1}=t_{2}=w_{0}=w_{1}=0) and 𝒳x∩E=(Ο–=x=zβ€²=y=v=0)\mathcal{X}_{x}\cap E=(\varpi=x=z^{\prime}=y=v=0).

Let Q2,Q3,Q4Q_{2},Q_{3},Q_{4} be the points on 𝒳x∩Z∩E\mathcal{X}_{x}\cap Z\cap E where ss (=1+t33+…=1+t_{3}^{3}+\dots) vanish. Around these points, t3t_{3} is a unit. and hence the following elements can be eliminated: x=t3βˆ’1​y3x=t_{3}^{-1}y^{3}, t1=t3βˆ’1​t22t_{1}=t_{3}^{-1}t_{2}^{2}, t0=t3βˆ’2​t23t_{0}=t_{3}^{-2}t_{2}^{3}, w0=t3βˆ’1​t2​w1w_{0}=t_{3}^{-1}t_{2}w_{1}, zβ€²=t3βˆ’1​y​vz^{\prime}=t_{3}^{-1}yv. Thus the maximal ideals of the local rings are generated by v,w1,y,t2,s,Ο–v,w_{1},y,t_{2},s,\varpi, subject to (s,w1,v,t2,ϡ​t3,y​t3)∈V2(s,w_{1},v,t_{2},\epsilon t_{3},yt_{3})\in V_{2}, where V2V_{2} is (as in Section 4.3.1) the cone over the image of the Veronese map β„™2β†’β„™5\mathbb{P}^{2}\to\mathbb{P}^{5}. It is straightforward to check that they are RDPs of type A1A_{1} on both components Z=(w1=Ο–=t2=0)Z=(w_{1}=\varpi=t_{2}=0) and E=(v=y=Ο–=0)E=(v=y=\varpi=0), and that the blow-up of 𝒳′\mathcal{X}^{\prime} at (s,v,y,w1,Ο΅,t2)(s,v,y,w_{1},\epsilon,t_{2}) is strictly semistable in the broad sense above these points.

Let Q1Q_{1} be the point (ti=x=y=zβ€²=v=wi=0)(t_{i}=x=y=z^{\prime}=v=w_{i}=0). Around this point ss is a unit, hence the maximal ideal of the local ring is generated by w0,w1,t3,v,zβ€²,y,Ο–w_{0},w_{1},t_{3},v,z^{\prime},y,\varpi, subject to (t3,v,y​s,z′​s,w1,ϡ​s,w0​s)∈Vβ€²(t_{3},v,ys,z^{\prime}s,w_{1},\epsilon s,w_{0}s)\in V^{\prime}, where

Vβ€²={(T03,T02​T1,T0​T12,T13,T0​T2,T1​T2,T23)}βŠ‚π”Έ7V^{\prime}=\{(T_{0}^{3},T_{0}^{2}T_{1},T_{0}T_{1}^{2},T_{1}^{3},T_{0}T_{2},T_{1}T_{2},T_{2}^{3})\}\subset\mathbb{A}^{7}

is the toric variety attached to the monoid MβŠ‚β„•3M\subset\mathbb{N}^{3} generated by the vectors

(3,0,0),(2,1,0),(1,2,0),(0,3,0),(1,0,1),(0,1,1),(0,0,3).(3,0,0),(2,1,0),(1,2,0),(0,3,0),(1,0,1),(0,1,1),(0,0,3).

(In other words, the elements (t3,…,w0​s)(t_{3},\dots,w_{0}s) define a morphism k​[M]β†’π’ͺ𝒳′,Q1k[M]\to\mathcal{O}_{\mathcal{X}^{\prime},Q_{1}} from the monoid ring.) We can check that, around this point,

  • β€’

    the component ZZ is (Ο–=w1=w0=0)(\varpi=w_{1}=w_{0}=0) and π’ͺZ,Q1\mathcal{O}_{Z,Q_{1}} is (1,1)3\frac{(1,1)}{3},

  • β€’

    the component EE is (Ο–=v=y=zβ€²=0)(\varpi=v=y=z^{\prime}=0) and π’ͺE,Q1\mathcal{O}_{E,Q_{1}} is an RDP of type A2A_{2}.

Let 𝒳′′→𝒳′\mathcal{X}^{\prime\prime}\to\mathcal{X}^{\prime} be the blow-up at (t3,v,y,zβ€²,w1,Ο΅,w0)(t_{3},v,y,z^{\prime},w_{1},\epsilon,w_{0}). Then 𝒳′′\mathcal{X}^{\prime\prime} is strictly semistable in the broad sense. More precisely, there are 44 components Z~,E~,E1,E2\tilde{Z},\tilde{E},E_{1},E_{2} over this point, Z~\tilde{Z} and E~\tilde{E} are the minimal resolutions of ZZ and EE at Q1Q_{1} respectively, E1β‰…β„™2E_{1}\cong\mathbb{P}^{2} and E2≅𝔽2E_{2}\cong\mathbb{F}_{2} are exceptional divisors, and the local equations at triple points are vt3​w1t3​t3βˆ’Ο΅β€‹s\frac{v}{t_{3}}\frac{w_{1}}{t_{3}}t_{3}-\epsilon s at E~∩Z~∩E2\tilde{E}\cap\tilde{Z}\cap E_{2} and w0​sw1​t3w1​ϡ​sw1βˆ’Ο΅β€‹s\frac{w_{0}s}{w_{1}}\frac{t_{3}}{w_{1}}\frac{\epsilon s}{w_{1}}-\epsilon s at E~∩E1∩E2\tilde{E}\cap E_{1}\cap E_{2}.

The assertion on H1​(π’ͺE)H^{1}(\mathcal{O}_{E}) and KEK_{E} follows also similarly since EE is a hypersurface of degree 18=9+6+2+118=9+6+2+1 in ℙ​(9,6,2,1)\mathbb{P}(9,6,2,1).

4.4. End of Proof of Theorem 1.2

Let π’œ\mathcal{A} be the NΓ©ron model of the abelian surface AA. By Proposition 2.4, the singularity of π’œk/{Β±1}\mathcal{A}_{k}/\{\pm 1\} is an elliptic double point of type EDP4,4(1)\mathrm{EDP}_{4,4}^{(1)} or EDP2,8(1)\mathrm{EDP}_{2,8}^{(1)}, according to the supersingular abelian surface π’œk\mathcal{A}_{k} being superspecial or not. We apply the normalized weighted blow-up 𝒳′′→𝒳′→𝒳=π’œ/{Β±1}\mathcal{X}^{\prime\prime}\to\mathcal{X}^{\prime}\to\mathcal{X}=\mathcal{A}/\{\pm 1\} described in Sections 4.2–4.3. Since A​[2]A[2] acts on itself (by translation) transitively, all singularity of Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z are of the same type. From the classification of possible singularities under this assumption (Claim 4.4 (1c) and (2d)), either Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z consists of RDPs, or it is 1​E​D​P2,8(0)1\mathrm{EDP}_{2,8}^{(0)}. In the latter case, since we know that the affine surface Eβˆ–ZE\setminus Z is the quotient by Gβ€²G^{\prime}, we can apply the blow-up construction once more. Thus, in each case, we obtain a model 𝒳′′\mathcal{X}^{\prime\prime} that is strictly semistable in the broad sense outside RDPs on the exceptional component. In each case, moreover, again by Claim 4.4 (1c) and (2d), the configuration of the RDPs on the special fiber is one of 16​A1,4​D40,2​D80,2​E80,1​D16016A_{1},4D_{4}^{0},2D_{8}^{0},2E_{8}^{0},1D_{16}^{0}. We apply Proposition 3.3 (16 times) to 16​A116A_{1} on the generic fiber and obtain a proper model that is strictly semistable in the broad sense.

Applying Theorem 4.2, we conclude that X=Km⁑(A)X=\operatorname{Km}(A) has potential good reduction.

5. Examples

We give explicit examples of abelian surface for which 16​A116A_{1}, 4​D404D_{4}^{0}, 2​D802D_{8}^{0} occur as Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z. Also we show that Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z may differ between isogenous abelian surfaces.

Let π’ͺ\mathcal{O} be a discrete valuation ring as before (in particular char⁑K=0\operatorname{char}K=0 and char⁑k=2\operatorname{char}k=2). The 𝔭\mathfrak{p}-adic valuation is denoted (additively) by v:π’ͺβ†’β„šβ‰₯0βˆͺ{∞}v\colon\mathcal{O}\to\mathbb{Q}_{\geq 0}\cup\{\infty\}.

Elliptic curves over KK having good reduction can be written, after replacing π’ͺ\mathcal{O}, in the form Y2+c​X​Y+Y=X3Y^{2}+cXY+Y=X^{3} with c∈π’ͺc\in\mathcal{O}. Let us write this curve EcE_{c}. The reduction of EcE_{c} is defined by the same equation (with coefficients considered modulo 𝔭\mathfrak{p}), and is supersingular if and only if cβˆˆπ”­c\in\mathfrak{p}. Using coordinate change u=XYu=\frac{X}{Y} and uΒ―=βˆ’XY+c​X+1\overline{u}=-\frac{X}{Y+cX+1}, we obtain the form u+uΒ―+c​u​uΒ―βˆ’(u​uΒ―)2=0u+\overline{u}+cu\overline{u}-(u\overline{u})^{2}=0, with origin at u=uΒ―=0u=\overline{u}=0, and the inversion map given by u↔uΒ―u\leftrightarrow\overline{u}.

Let c1,c2βˆˆπ”­c_{1},c_{2}\in\mathfrak{p} and consider Ac1,c2:=Ec1Γ—Ec2A_{c_{1},c_{2}}:=E_{c_{1}}\times E_{c_{2}}, u+uΒ―+c1​u​uΒ―βˆ’(u​uΒ―)2=v+vΒ―+c2​v​vΒ―βˆ’(v​vΒ―)2=0u+\overline{u}+c_{1}u\overline{u}-(u\overline{u})^{2}=v+\overline{v}+c_{2}v\overline{v}-(v\overline{v})^{2}=0. This abelian surface have good superspecial reduction. Using the notations of Section 4.3.1, we have a=βˆ’c1​x+x2a=-c_{1}x+x^{2} and b=βˆ’c2​y+y2b=-c_{2}y+y^{2}. Then Ο΅\epsilon is chosen so that v​(Ο΅)=min⁑{12​v​(c1),12​v​(c2),13​v​(2)}v(\epsilon)=\min\{\frac{1}{2}v(c_{1}),\frac{1}{2}v(c_{2}),\frac{1}{3}v(2)\}. We observe that Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z is

  • β€’

    16​A116A_{1} if 4β€‹Ο΅βˆ’6∈π’ͺβˆ—4\epsilon^{-6}\in\mathcal{O}^{*}, equivalently if both v​(c1)v(c_{1}) and v​(c2)v(c_{2}) are β‰₯23​v​(2)\geq\frac{2}{3}v(2),

  • β€’

    4​D404D_{4}^{0} if 4β€‹Ο΅βˆ’6βˆˆπ”­4\epsilon^{-6}\in\mathfrak{p} and both Ο΅βˆ’2​c1∈π’ͺβˆ—\epsilon^{-2}c_{1}\in\mathcal{O}^{*} and Ο΅βˆ’2​c2∈π’ͺβˆ—\epsilon^{-2}c_{2}\in\mathcal{O}^{*}, equivalently if v​(c1)=v​(c2)<23​v​(2)v(c_{1})=v(c_{2})<\frac{2}{3}v(2),

  • β€’

    2​D802D_{8}^{0} if 4β€‹Ο΅βˆ’6βˆˆπ”­4\epsilon^{-6}\in\mathfrak{p} and either Ο΅βˆ’2​c1βˆˆπ”­\epsilon^{-2}c_{1}\in\mathfrak{p} or Ο΅βˆ’2​c2βˆˆπ”­\epsilon^{-2}c_{2}\in\mathfrak{p}, equivalently if v​(c1)v(c_{1}) and v​(c2)v(c_{2}) are different and at least one is <23​v​(2)<\frac{2}{3}v(2),

Next, we will see that isogenous abelian varieties may result in different configuration of singularities on Eβˆ–ZE\setminus Z. Consider another elliptic curve Ec1β€²E_{c_{1}^{\prime}} that admits an isogeny to Ec1E_{c_{1}} of degree 22 (if c1c_{1} is generic, then there exist exactly 33 such elliptic curves up to isomorphism). Using the formula j​(Ec)=c3​(c3βˆ’24)3​(c3βˆ’27)βˆ’1j(E_{c})=c^{3}(c^{3}-24)^{3}(c^{3}-27)^{-1} and the explicit form of the modular polynomial Ξ¦2​(X,Y)\Phi_{2}(X,Y), we observe that

  • β€’

    if v​(c1)<23​v​(2)v(c_{1})<\frac{2}{3}v(2), then there exists such c1β€²c_{1}^{\prime} with v​(c1β€²)=12​v​(c1)v(c_{1}^{\prime})=\frac{1}{2}v(c_{1}),

  • β€’

    if v​(c1)β‰₯23​v​(2)v(c_{1})\geq\frac{2}{3}v(2), then there exists such c1β€²c_{1}^{\prime} with v​(c1β€²)=13​v​(2)v(c_{1}^{\prime})=\frac{1}{3}v(2).

We conclude that, while Ac1,c2A_{c_{1},c_{2}} and Ac1β€²,c2A_{c_{1}^{\prime},c_{2}} are isogenous, the resulting configurations of singularities on Sing⁑(E)βˆ–Z\operatorname{Sing}(E)\setminus Z differ.

Acknowledgments

I thank Hiroyuki Ito, Kazuhiro Ito, Tetsushi Ito, Teruhisa Koshikawa, Ippei Nagamachi, Hisanori Ohashi, Teppei Takamatsu, and Fuetaro Yobuko for helpful comments and discussions.

References