Skip to content
BY 4.0 license Open Access Published by De Gruyter October 19, 2018

Ordinary K3 surfaces over a finite field

  • Lenny Taelman

Abstract

We give a description of the category of ordinary K3 surfaces over a finite field in terms of linear algebra data over 𝐙. This gives an analogue for K3 surfaces of Deligne’s description of the category of ordinary abelian varieties over a finite field, and refines earlier work by N.O. Nygaard and J.-D. Yu. Our main result is conditional on a conjecture on potential semi-stable reduction of K3 surfaces over p-adic fields. We give unconditional versions for K3 surfaces of large Picard rank and for K3 surfaces of small degree.

Introduction

Statement of the main results

A K3 surface X over a perfect field k of characteristic p is called ordinary if it satisfies the following equivalent conditions:

  1. the Hodge and Newton polygons of Hcrys2(X/W(k)) coincide,

  2. the Frobenius endomorphism of H2(X,𝒪X) is a bijection,

  3. the formal Brauer group of X (see [1]) has height 1.

If k is finite, then these are also equivalent with |X(k)|1modp. Building on [1] and [7], Nygaard [14] has shown that such an ordinary X has a canonical lift Xcan over the ring of Witt vectors W(k).

Choose an embedding ι:W(𝐅q)𝐂. Then with every ordinary K3 surface over 𝐅q we can associate a complex K3 surface Xcanι:=XcanW(𝐅q)𝐂 and an integral lattice

M:=H2(Xcanι,𝐙).

Using the Kuga–Satake construction, Nygaard [14] and Yu [20] have shown that there exists a (necessarily unique) endomorphism F of M𝐙𝐙[1p] such that for every p the canonical isomorphism

M𝐙Hét2(X𝐅¯q,𝐙)

matches F with the geometric Frobenius Frob on étale cohomology (see also Section 3.1). We have:

  1. the pairing -,- on M is unimodular, even, and of signature (3,19),

  2. Fx,Fy=q2x,y for every x,yM.

From Deligne’s proof of the Weil conjectures for K3 surfaces [6] we also know

  1. the endomorphism F of M𝐂 is semi-simple and all its eigenvalues have absolute value q.

Our first result is an integral p-adic property of the pair (M,F).

Theorem A.

The endomorphism F preserves the Z-module M and satisfies

  1. the 𝐙p[F]-module M𝐙p decomposes as M0M1M2 with

    1. FMs=qsMs for all s,

    2. M0, M1 and M2 are free 𝐙p-modules of rank 1, 20, 1 , respectively.

For a 𝐙-lattice M equipped with an endomorphism F satisfying (M1)–(M4) we denote by NS(M,F) the group

NS(M,F):={xMFdx=qdx for some d1}.

Using the fact that all line bundles on an ordinary K3 surface extend uniquely to its canonical lift, one shows that there is a natural bijection

PicX𝐅¯qNS(M,F),

and that the ample line bundles on X𝐅¯q span a real cone 𝒦M𝐑 satisfying

  1. 𝒦 is a connected component of

    {xNS(M,F)𝐑|x,x>0,x,δ0 for all δNS(M,F) with δ2=-2}

    satisfying F𝒦=𝒦.

See Section 3.2 for more details.

Definition (see [11]).

Let 𝒪K be a complete discrete valuation ring with fraction field K. We say that a K3 surface X over K satisfies () if there exists a finite extension KL and algebraic space 𝔛/𝒪L satisfying

  1. 𝔛LXL,

  2. 𝔛 is regular,

  3. the special fiber is a strict normal crossing divisor in 𝔛,

  4. the relative dualizing sheaf of 𝔛/𝒪L is trivial.

This is a strong form of “potential semi-stable reduction”. Over a complete dvr 𝒪K of residue characteristic 0, it is known that all K3 surfaces satisfy (). It is expected to hold in general, but currently only known under extra assumptions on X.

Our main result is the following description of the category of ordinary K3 surfaces over 𝐅q. It is an analogue for K3 surfaces of Deligne’s theorem [5] on ordinary abelian varieties over a finite field.

Theorem B.

Fix an embedding ι:W(Fq)C. The resulting functor X(M,F,K) is a fully faithful functor between the groupoids of

  1. ordinary K3 surfaces X over 𝐅q, and

  2. triples (M,F,𝒦) consisting of

    1. an integral lattice M,

    2. an endomorphism F of the 𝐙-module M, and

    3. a convex subset 𝒦M𝐑,

    satisfying (M1)–(M5).

If every K3 surface over FracW(Fq) satisfies (), then the functor is essentially surjective.

Fully faithfulness is essentially due to Nygaard [15] and Yu [20]. Our contribution is a description of the image of this functor.

Restricting to families for which () is known to hold, we also obtain unconditional equivalences of categories between ordinary K3 surfaces X/𝐅q satisfying one of the following additional conditions

  1. there is an ample PicX𝐅¯q with 2<p-4,

  2. PicX𝐅¯q contains a hyperbolic plane and p5,

  3. X has geometric Picard rank 12 and p5,

and triples (M,F,𝒦) satisfying the analogous constraints. See Theorem 4.6 for the precise statement.

About the proofs

A crucial ingredient in the proof of Theorems A and B is the following criterion which distinguishes the canonical lift amongst all lifts using p-adic étale cohomology.

Theorem C.

Let OK be a complete discrete valuation ring with perfect residue field k of characteristic p and fraction field K of characteristic 0. Let X be a projective K3 surface over OK and assume that Xk is ordinary. Then the following are equivalent:

  1. 𝔛 is the base change from W(k) to 𝒪K of the canonical lift of 𝔛k.

  2. Hét2(𝔛K¯,𝐙p)H0H1(-1)H2(-2) with Hi unramified 𝐙p[GalK]-modules, free of rank 1, 20, 1 over 𝐙p, respectively.

Here the (-1) and (-2) in (ii) denote Tate twists. This theorem is an integral refinement of a theorem of Yu [20] which characterizes quasi-canonical lifts by the splitting of étale cohomology with 𝐐p-coefficients.

The canonical lift of X is defined in terms of its enlarged formal Brauer group Ψ (a p-divisible group), and to prove Theorem C we need to compare the p-adic étale cohomology of the generic fiber of a K3 surface over 𝒪K to the Tate module of its enlarged formal Brauer group. With 𝐐p-coefficients, such a comparison has been shown by Artin and Mazur [1]. We give a different argument leading to an integral version, see Theorem 2.1. Once Theorem C is established, Theorem A is an almost formal consequence.

Finally, we briefly sketch the argument for the proof of Theorem B. Fully faithfulness was shown by Nygaard [15] and Yu [20] (see Section 3.3 for more details). The proof of essential surjectivity starts with the observation that the decomposition in (M4) induces (via the embedding ι:𝐙p𝐂) a Hodge structure on M, for which NS(M,F) consists precisely of the Hodge classes. The Torelli theorem for K3 surfaces then shows that there is a canonical K3 surface X/𝐂 with H2(X,𝐙)=M and whose ample cone 𝒦H2(X,𝐑) coincides with 𝒦M𝐑. This K3 surface has complex multiplication, and hence can be defined over a number field. Using the strong version of the main theorem of CM for K3 surfaces of [18], we show that we can find a model of X over K:=FracW(𝐅q)𝐂 such that

  1. the GalK-module Hét2(XK¯,𝐙p)=M𝐙p decomposes as in Theorem C,

  2. for p, the GalK-module Hét2(XK¯,𝐙)=M𝐙 is unramified, and Frobenius acts as F.

By assuming property (), it follows from the Néron–Ogg–Shafarevich criterion of Liedtke and Matsumoto [11] that X has good reduction over an unramified extension L of K. Using Theorem C, we show that XL is the canonical lift of its reduction, and deduce from this that X has already a smooth projective model 𝔛 over 𝒪K. By construction, its reduction 𝔛k maps under our functor to the given triple (M,F,𝒦).

A question

We end this introduction with an essentially lattice-theoretical question to which we do not know the answer:

Question.

Does there exist a triple (M,F,𝒦) satisfying (M1)–(M5) and the inequality 1+trF+q2<0?

By (M3) such a triple can only exist for small q. A positive answer to this question would imply that there exist K3 surfaces over p-adic fields that do not satisfy (). Indeed, if (M,F,𝒦) came from a K3 surface X/𝐅q as in Theorem B, we would have

|X(𝐅q)|=tr(Frob,H(X𝐅¯q,𝐐))=1+trF+q2<0,

which is absurd.

1 p-divisible groups associated to K3 surfaces

Let Λ be a complete noetherian local ring with perfect residue field k of characteristic p>0 and let 𝔛 be a K3 surface over SpecΛ. We recall (and complement) some of the main results of Artin and Mazur [1] on the formal Brauer group and enlarged formal Brauer group of 𝔛.

1.1 The formal Brauer group

Let ArtΛ be the category of Artinian local Λ-algebras (A,𝔪) with perfect residue field A/𝔪. For an (A,𝔪)ArtΛ we denote by 𝒰A the sheaf on 𝔛ét defined by the exact sequence

1𝒰A𝒪𝔛A×𝒪𝔛A/𝔪×1.

The formal Brauer group of 𝔛 is the functor

Br^(𝔛):ArtΛAb,AH2(𝔛ét,𝒰A).

By [1] it is representable by a one-dimensional formal group, and if 𝔛 is not supersingular, then Br^(𝔛) is a p-divisible group.

Lemma 1.1.

One has H(Xét,UA)=H(X,UA) and H1(Xét,UA)=0.

Here H(𝔛,-) denotes Zariski cohomology.

Proof.

The sheaf 𝒰A has a filtration whose graded pieces are 𝔪n/𝔪n+1A/𝔪𝒪𝔛A/𝔪. Since these are coherent, we have

H(𝔛ét,𝒰A)=H(𝔛,𝒰A).

Moreover, since H1(𝔛A/𝔪,𝒪𝔛A/𝔪) vanishes, we conclude that H1(𝔛ét,𝒰A)=0. ∎

Lemma 1.2.

For every (A,m)ArtΛ there is a natural exact sequence

0Br^(𝔛)[pr](A)Hfl2(𝔛A,μpr)Hfl2(𝔛A/𝔪,μpr)

of abelian groups.

Here Hfl denotes cohomology in the fppf topology.

Proof.

Consider the complex 𝒰Apr𝒰A on 𝔛ét in degrees 0 and 1. We have a short exact sequence

1H1(𝔛,𝒰A)𝐙/pr𝐙H2(𝔛,𝒰Apr𝒰A)H2(𝔛,𝒰A)[pr]1

and thanks to Lemma 1.1 we obtain canonical isomorphisms

Br^(𝔛)[pr](A)=H2(𝔛,𝒰A)[pr]=H2(𝔛,𝒰Apr𝒰A).

Now consider the short exact sequence

of length-two complexes concentrated in degrees 0 and 1. Using the Kummer sequence and the fact that 𝐆m is smooth, we see that

Hétn(𝔛,𝒪𝔛A×pr𝒪𝔛A×)=Hfln(𝔛A,𝐆mpr𝐆m)=Hfln(𝔛A,μpr)

and similarly for 𝔛A/𝔪. It follows that the above short exact sequence of complexes induces a long exact sequence of (hyper-)cohomology groups

Hfl1(𝔛A/𝔪,μpr)H2(𝔛,𝒰Apr𝒰A)Hfl2(𝔛A,μpr)Hfl2(𝔛A/𝔪,μpr).

Since A/𝔪 is perfect and Pic𝔛A/𝔪 is torsion-free, we have Hfl1(𝔛A/𝔪,μpr)=0 and we conclude

Br^(𝔛)[pr](A)=H2(𝔛,𝒰Apr𝒰A)=ker[Hfl2(𝔛A,μpr)Hfl2(𝔛A/𝔪,μpr)],

which is what we had to show. ∎

1.2 The enlarged formal Brauer group

We now assume that 𝔛k is ordinary. Denote by μp the sheaf colimrμpr on the fppf site. The enlarged formal Brauer group of 𝔛 is the functor

Ψ(𝔛):ArtΛAb,AHfl2(𝔛A,μp).

A priori this differs from the definition of Artin–Mazur [1, Section IV.1] in two ways. First, Artin and Mazur restrict to A with algebraically closed residue fields, and then use Galois descent to extend their definition to non-closed perfect residue fields, and second, they restrict to those classes in Hfl2(𝔛A,μp) that map to the p-divisible part of Hfl2(𝔛k¯,μp). Lemmas 1.4 and 1.5 below show that the above definition is equivalent to that of Artin–Mazur (under our standing condition that 𝔛k is an ordinary K3 surface). See also [14, Corollary 1.5].

The following lemma is well-known, and implicitly used in [1] and [14]. We include it for the sake of completeness.

Lemma 1.3.

For any quasi-compact quasi-separated scheme X there is a natural isomorphism

Hfl(X,μp)colimrHfl(X,μpr).

Proof.

This follows from [22, 0739], taking for the class of quasi-compact and quasi-separated schemes, and for Cov the fpqc covers consisting of finitely many affine schemes. ∎

Lemma 1.4 ([14, Corollary 1.4]).

The group Hfl2(Xk¯,μp) is p-divisible. ∎

Lemma 1.5.

One has Hfl2(Xk¯,μp)Gal(k¯/k)=Hfl2(Xk,μp).

Proof.

Since the p-th power map k¯×k¯× is a bijection, and since Pic𝔛k¯ is torsion-free, we have Hfli(𝔛k¯,μp)=colimrHfli(𝔛k¯,μpr)=0 for i{0,1}. It now follows from the Hochschild–Serre spectral sequence

E2s,t=Hs(Gal(k¯/k),Hflt(𝔛k¯,μp))Hfls+t(𝔛k,μp)

that Hfl2(𝔛k¯,μp)Gal(k¯/k)=Hfl2(𝔛k,μp). ∎

The following theorem summarizes the properties of the enlarged formal Brauer group that we will use.

Theorem 1.6.

Let X be a formal K3 surface over Λ with Xk ordinary. Then the enlarged formal Brauer group Ψ(X) is representable by a p-divisible group over Λ. Its étale-local exact sequence

0Ψ(𝔛)Ψ(𝔛)Ψét(𝔛)0

satisfies

  1. Ψ(𝔛) is a connected p-divisible group of height 1 , with

    Ψ(𝔛)[pr](A)=ker(Hfl2(𝔛A,μpr)Hfl2(𝔛A/𝔪,μpr))

    for all (A,𝔪)ArtΛ. It is canonically isomorphic to Br^(𝔛),

  2. Ψ(𝔛)[pr](A)=Hfl2(𝔛A,μpr) for all (A,𝔪)ArtΛ,

  3. Ψét(𝔛)[pr](A)=Hfl2(𝔛A/𝔪,μpr) for all (A,𝔪)ArtΛ.

Proof.

The representability and (i) are shown in [1, Proposition IV.1.8].

To prove (ii), note that since H1(𝔛A,𝐆m)=Pic𝔛A is torsion-free, we have a natural isomorphism

Hfl1(𝔛A,μpr)=H0(𝔛A,𝐆m)𝐙/pr𝐙.

Taking the colimit over r, we obtain a natural isomorphism

Hfl1(𝔛A,μp)=colimrHfl1(𝔛A,μpr)=H0(𝔛A,𝐆m)(𝐐p/𝐙p),

and in particular we see that Hfl1(𝔛A,μp) is p-divisible. Now the long exact sequence associated to

1μprμpprμp1

induces an isomorphism

Hfl2(𝔛A,μpr)Hfl2(𝔛A,μp)[pr],

which proves (ii).

A similar argument shows (iii) for A with A/𝔪 algebraically closed, after which Lemma 1.5 implies the general case. ∎

1.3 Canonical lifts

Let X/k be an ordinary K3 surface. Since formal groups of height 1 are rigid, the p-divisible group Ψ(X) over k extends uniquely to a p-divisible group Ψ(X)can over Λ. Also the étale p-divisible group Ψét(X) over k extends uniquely to a p-divisible group Ψét(X)can over Λ.

To every lift 𝔛/Λ of X/k we then have an associated short exact sequence of p-divisible groups

(1.1)0Ψ(X)canΨ(𝔛)Ψét(X)can0

over Λ. In analogy with Serre–Tate theory, we have the following theorem.

Theorem 1.7 (Nygaard [14, Theorem 1.6]).

The map

{formal lifts 𝔛/Λ of X/k}ExtΛ1(Ψét(X)can,Ψ(X)can),𝔛Ψ(𝔛)

is a bijection.

It follows that there exists a unique lift 𝔛/Λ for which the sequence (1.1) splits. This 𝔛 is unique up to unique isomorphism, and is called the canonical lift of X. We denote it by Xcan.

Proposition 1.8 ([14, Proposition 1.8]).

The map PicXcanPicX is a bijection.

Corollary 1.9 ([14, Proposition 1.8]).

The lift Xcan is algebraizable and projective.

2 p-adic étale cohomology

Let 𝒪K be a complete discrete valuation ring whose residue field k is perfect of characteristic p and whose fraction field K is of characteristic 0.

2.1 p-adic étale cohomology and the enlarged formal Brauer group

Theorem 2.1.

Let X be a projective K3 surface over OK. Assume that Xk is ordinary. Then there is a natural injective map of GalK-modules

TpΨ(𝔛)K¯Hét2(XK¯,𝐙p(1))

whose cokernel is a free Zp-module of rank 1.

Recall that if 𝔛 is ordinary, then TpΨ(𝔛)K¯ has rank 21. Up to possible torsion in the cokernel, Theorem 2.1 is shown in [1, Section IV.2]. The proof of Artin and Mazur is based on Lefschetz pencils, reducing the problem on H2 to a statement about H1 and torsors. We give a proof working directly with the H2 and their relation to Brauer groups to obtain the finer “integral” statement above. This is made possible by the theorem of Gabber and de Jong [4] asserting that the Brauer group and the cohomological Brauer group of a quasi-projective scheme coincide.

Let 𝔛 be a formal K3 surface over 𝒪K. We denote by 𝔛n the truncation 𝔛𝒪K/𝔪n.

Lemma 2.2.

For all i the natural map

Hfli(𝔛,μpr)limnHfli(𝔛n,μpr)

is an isomorphism.

Proof.

As in the proof of Lemma 1.2, we have

Hfli(𝔛n,μpr)=Héti(𝔛,𝒪𝔛n×pr𝒪𝔛n×)

and similarly

Hfli(𝔛,μpr)=Héti(𝔛,𝒪𝔛×pr𝒪𝔛×).

Let 𝒰n be the kernel of 𝒪𝔛n×𝒪𝔛1× and 𝒰 the kernel of 𝒪𝔛×𝒪𝔛1×. Then by the usual dévissage arguments the lemma reduces to showing that

Héti(𝔛,𝒰)limnHéti(𝔛,𝒰n)

is an isomorphism for all i.

Since the maps 𝒰n+1𝒰n are surjective, we have Rlimn𝒰n=𝒰. Since 𝒰 has a filtration with graded pieces isomorphic to 𝒪𝔛1, it has cohomology concentrated in degrees 0 and 2. These two facts imply

RΓét(𝔛,Rlimn𝒰n)=Hét0(𝔛,𝒰)Hét2(𝔛,𝒰)[-2]

in 𝒟(Ab). Similarly, we have

RlimnRΓét(𝔛,𝒰n)=limnHét0(𝔛,𝒰n)limnHét2(𝔛,𝒰n)[-2]

in 𝒟(Ab). As RΓét commutes with Rlim, the lemma follows. ∎

Corollary 2.3.

If Xk is ordinary, then Ψ(X)(K)[pr]=Hfl2(X,μpr).

Proof.

Indeed, we have

Ψ(𝔛)(K)[pr]=limnΨ(𝔛)[pr](𝒪K/𝔪n)=limnHfl2(𝔛n,μpr),

so the corollary follows from Lemma 2.2. ∎

Proposition 2.4.

If X is a projective K3 surface over OK, then for all r the natural map Hfl2(X,μpr)Hfl2(XK,μpr) is injective.

Proof.

The Kummer sequence gives a commutative diagram with exact rows

Since 𝔛 is projective, we have Br=Br for 𝔛 and 𝔛K.

The left arrow in the diagram is an isomorphism since the special fiber 𝔛k is a principal divisor in 𝔛, so that Pic𝔛Pic𝔛K is an isomorphism. By [8, Corollary 1.8] the natural maps of Br𝔛 and Br𝔛K to BrK(𝔛K) are injective, so that also the right arrow in the diagram is injective. We conclude that the middle map is injective. ∎

Proof of Theorem 2.1.

The proof is now formal. By Corollary 2.3 and Proposition 2.4 we have for every r and every finite extension KL a canonical injection

Ψ(𝔛)[pr](L)Hfl2(𝔛L,μpr)=Hét2(𝔛L,𝐙/pr𝐙(1)).

Taking the colimit over all L, we obtain a GalK-equivariant injective map

ρr:Ψ(𝔛)[pr](K¯)Hét2(𝔛K¯,𝐙/pr𝐙(1)),

and taking the limit over r, we obtain a GalK-equivariant injective map

ρ:TpΨ(𝔛)K¯Hét2(𝔛K¯,𝐙p(1)).

Denote the cokernel of ρ by Q. Tensoring ρ with 𝐙/p𝐙 yields an exact sequence

0Tor(Q,𝐙/p𝐙)Ψ(𝔛)[p](K¯)ρ𝐙/p𝐙Hét2(𝔛K¯,𝐙/p𝐙(1))Q𝐙/p𝐙0.

Since ρ1=ρ𝐙/p𝐙 is injective, we see that Tor(Q,𝐙/p𝐙) vanishes and that Q is torsion-free. ∎

2.2 Canonical lifts and p-adic étale cohomology

In this subsection we prove Theorem C, characterizing the canonical lift in terms of p-adic étale cohomology.

Lemma 2.5.

Let U be a free Zp-module of rank 2 and b:U×UZp a non-degenerate symmetric bilinear form. Let LU be a totally isotropic rank-1 submodule. If L is saturated in

U:={x𝐐p𝐙pUb(x,U)𝐙p},

then U=U.

Proof.

Since L is saturated in U, it is also saturated in UU and we may choose a basis (e,f) for U with L=e. Set d:=b(e,f). Since b(e,e)=0, the determinant of b is -d2. Since e/d lies in U, we must have that d is a unit and therefore U=U. ∎

Proof of Theorem C.

Assume that (ii) holds. Then we have

Hét2(𝔛K¯,𝐙p(1))=H0(1)H1H2(-1)

with the Hi unramified. Since the Tate module of a p-divisible group is Hodge–Tate of weights 0 and -1, we have Hom(TpΨ(𝔛)K¯,H2(-1))=0, and by Theorem 2.1 we see that

TpΨ(𝔛)K¯=H0(1)H1.

By Tate’s theorem [19, Theorem 4] this implies that

Ψ(𝔛)=Ψ0(𝔛)Ψét(𝔛)

with TpΨ0(𝔛)K¯=H0(1) and TpΨét(𝔛)K¯=H1. It follows that 𝔛 is the base change of the canonical lift of 𝔛k to 𝒪K.

Conversely, assume that 𝔛 is the base change of the canonical lift of 𝔛k to 𝒪K. Let H1 be the image of the direct summand TpΨét(𝔛)K¯ under the embedding

TpΨ(𝔛)K¯Hét2(𝔛K¯,𝐙p(1))

of Theorem 2.1. It is a primitive submodule, and considering Hodge–Tate weights, we see that the restriction of the bilinear form on H2 to H1 is non-degenerate. Let UHét2(𝔛K¯,𝐙p(1)) be its orthogonal complement. Then U is a rank 2 lattice over 𝐙p. The inclusions of H1 and U as mutual orthogonal complements inside the self-dual lattice Hét2(𝔛K¯,𝐙p(1)) induce an isomorphism

α:U/U(H1)/H1

and an identification

Hét2(𝔛K¯,𝐙p(1)):={(x,y)U(H1)α(x)=y}.

Consider the unramified GalK-module H0:=TpΨ(𝔛)K¯(-1). We have that H0(1) is a totally isotropic line in U. We claim that it is saturated in U. Indeed, if xU satisfies pxH0(1) then (x,α(x)) defines a p-torsion element in the cokernel of the embedding TpΨ(𝔛)K¯Hét2(𝔛K¯,𝐙p(1)), which must be trivial by Theorem 2.1. By Lemma 2.5 we conclude that U=U and that Hét2(𝔛K¯,𝐙p(1))=UH1.

Now U is a unimodular 𝐙p-lattice of rank 2 containing an isotropic line. Moreover, since the intersection pairing on Hét2(𝔛K¯,𝐙p(1)) is even, so is the lattice U. It follows that there is a unique isotropic line H2(-1)U with U=H0(1)H2(-1) and with H0 and H2 dual unramified representations. We find

Hét2(𝔛K¯,𝐙p(1))=H0(1)H1H2(-1),

as claimed. ∎

Remark 2.6.

Using the results on integral p-adic Hodge theory by Bhatt, Morrow, and Scholze [2], one can show that the splitting of Hét2(𝔛K¯,𝐙p) as in Theorem C implies an analogous splitting of the filtered crystal Hcrys2(𝔛/W). If p>2, then the splitting of Hcrys2(𝔛/W) implies that 𝔛 is the canonical lift of 𝔛k (see [7] and [14, Lemma 1.11, Theorem 1.12]). For p=2, however, the splitting of Hcrys2(𝔛/W) is a weaker condition than the splitting of Hét2(𝔛K¯,𝐙p), see also [7, Section 2.1.16.b].

3 The functor X(M,F,𝒦)

Let 𝐅q be a finite field with q=pa elements. Let W be the ring of Witt vectors of 𝐅q, and K its fraction field. Fix an embedding ι:K¯𝐂. By Section 1.3, every ordinary K3 surface X over 𝐅q has a canonical lift Xcan over W. We will denote by Xcanι the complex K3 surface obtained by base changing Xcan along ι:W𝐂.

3.1 Construction of a pair (M,F)

Let X be an ordinary K3 surface over 𝐅q. The following theorem, due to Nygaard and Yu, says that the Frobenius on X can be lifted to an endomorphism of the Betti cohomology of Xcanι, at least after inverting p. The proof relies on the Kuga–Satake construction.

Theorem 3.1 (Nygaard [14, Section 3], Yu [20, Lemma 2.3]).

There is a unique endomorphism F of H2(Xcanι,Z[1p]) such that

  1. for every p the map F corresponds under the comparison isomorphism

    H2(Xcanι,𝐙[1p])𝐙Hét2(X𝐅¯q,𝐙)

    to the geometric Frobenius Frob on étale cohomology,

  2. the map F corresponds under the comparison isomorphism

    H2(Xcanι,𝐙[1p])BcrysHcrys2(X/W)WBcrys

    to the endomorphism ϕaid, where ϕ denotes the crystalline Frobenius.

Moreover, F preserves the Hodge structure on H2(Xcanι,Q). ∎

For later use in the proofs of Theorems A and B, we record some well-known properties of Tate twists of unramified p-adic Galois representations.

Lemma 3.2.

Let V be an unramified GalK-representation over Qp, and let n be an integer. Let Frob be the geometric Frobenius endomorphism of V, relative to Fq. Consider the GalK-module V(-n):=VQp(-n).

  1. The map

    K×GL(V(-n)),xFrobv(x)qnv(x)NmK/𝐐p(x)-n

    factors over the reciprocity map K×GalKab and induces the action of GalK on V(-n).

  2. Dcrys(qnFrob)=ϕa as endomorphisms of Dcrys(V(-n)).

Proof.

The first statement follows from Lubin–Tate theory, see for example [17, Section 3.1, Theorem 2]. For the second, one uses that the functor Dcrys commutes with Tate twists to reduce to the case n=0. In this case, the statement follows from the observation that Frob1=1ϕa as endomorphisms of (V𝐙pW(𝐅¯q))Gal𝐅q, where the action of Gal𝐅q is the diagonal one. ∎

Theorem A is now an almost immediate consequence of Theorem C.

Proof of Theorem A.

By its definition (in Theorem 3.1), it is clear that the endomorphism F of H2(Xcanι,𝐐)𝐐p=Hét2(XK¯,𝐐p) satisfies

Dcrys(F)=ϕa

on Hcrys2(X/W)[1p].

By Theorem C we have a decomposition

Hét2(XK¯,𝐙p)=H0H1(-1)H2(-2)

with Hi unramified and by Lemma 3.2, also the endomorphism

F:=FrobH0qFrobH1q2FrobH2

of Hét2(XK¯,𝐐p) satisfies

Dcrys(F)=ϕa.

Since Dcrys is fully faithful, we must have F=F. But it then follows immediately that F preserves the 𝐙p-lattice Hét2(XK¯,𝐙p), and that Hét2(XK¯,𝐙p) decomposes as described in the theorem. ∎

We thus have constructed from X/𝐅q an integral lattice M:=H2(Xcanι,𝐙), equipped with an endomorphism F, satisfying (M1)–(M4). We end this paragraph by relating the p-adic decomposition in (M4) to the Hodge decomposition for Xcanι.

Lemma 3.3.

Let (M,F) be a pair satisfying (M1)–(M4). Then complex conjugation on MC maps the subspace MsZp,ιC to M2-sZp,ιC.

In other words: the decomposition in (M4) induces under ι:𝐙p𝐂 a 𝐙-Hodge structure on M.

Proof.

By property (M4), the one-dimensional subspace M0𝐙p,ι𝐂 of M𝐂 is the unique eigenspace for the endomorphism F corresponding to an eigenvalue u𝐐p𝐂 with vp(u)=0. By (M3) we have uu¯=q2, and hence also the eigenvalue u¯=q2u lies in 𝐐p𝐂. Since vp(u¯)=vp(q2), we see that the corresponding eigenspace is M2𝐙p,ι𝐂. Similarly, complex conjugation maps M2 to M0. By (M2), the subspace M1 is the orthogonal complement of M0M2, and hence is preserved by complex conjugation. ∎

Proposition 3.4.

Let (M,F) be the pair associated to an ordinary K3 surface X over Fq. Then we have MsZp,ιC=Hs,2-s(Xcanι) as subspaces of MC=H2(Xcanι,C).

Proof.

Indeed, under the “Hodge–Tate” comparison isomorphism

HdR2(Xcan,K/K)K𝐂pHét2(XK¯,𝐐p)𝐐p𝐂pM𝐙p𝐂p

the subspace (FiliHdR2(Xcan,K/K))K𝐂p is mapped to siMs𝐙p𝐂p. Extending ι to an embedding 𝐂p𝐂, we see that the Hodge filtration on H2(Xcanι,𝐂) agrees with the filtration on M𝐂 induced by the decomposition on M𝐙p, and hence by Lemma 3.3 we have Ms𝐙p𝐂=Hs,2-s(Xcanι). ∎

3.2 Line bundles and ample cone

Let X be an ordinary K3 surface over 𝐅q. Recall from Section 1.3 that line bundles on X extend uniquely to Xcan. We obtain isomorphisms

PicXPicXcanPicXcan,K.

Let Knr be the maximal unramified extension of K.

Proposition 3.5.

We have natural isomorphisms

PicX𝐅¯qPicXcan,KnrPicXcan,K¯,

and a class λPicXF¯q is ample if and only if its image in PicXcan,K¯ is ample.

Proof.

The first isomorphism follows from the fact that canonical lifts commute with finite unramified extensions. The second isomorphism follows from the triviality of the action of Gal(K/Knr) on PicXcan,K¯Hét2(Xcan,K¯,𝐐(1)) and the vanishing of BrKnr.

It remains to show that the isomorphism

PicX𝐅¯qPicXcan,K¯

restricts to a bijection between the subsets of ample classes. Fix an ample line bundle H on X. Then by the structure theorem on the ample cone of a K3 surface over an algebraically closed field ([9, Section 8.1]) we have that a line bundle L on X𝐅¯q is ample if and only if

  1. L2>0,

  2. for every DPicX𝐅¯q with D2=-2 we have LD0 and LD has the same sign as HD,

and similarly for line bundles on Xcan,K¯. But the bijection PicX𝐅¯qPicXcan,K¯ is an isometry, and the canonical lift Hcan of the ample line bundle H is itself ample, so we conclude that the bijection preserves ample classes. ∎

Proposition 3.6.

For every d1 the map

PicX𝐅qd{λH2(Xcanι,𝐙)Fdλ=qdλ}

is an isomorphism.

Proof.

Injectivity is clear, it suffices to show that the map is surjective. Without loss of generality we may assume that d=1.

By Proposition 3.4 any λH2(Xcanι,𝐙) satisfying Fλ=qλ is a Hodge class and by Theorem 3.1 we see that λ defines a GalK-invariant element of PicXcan,K¯. By Proposition 3.5, λ corresponds to a Gal𝐅q-invariant class in PicX𝐅¯q, which defines a line bundle on X since the Brauer group of 𝐅q vanishes. We conclude that the map is surjective as claimed. ∎

Proposition 3.7.

The real cone KMZR spanned by the classes of ample line bundles on PicXF¯q satisfies (M5).

Proof.

This follows immediately from Propositions 3.5 and 3.6, and the structure of the ample cone of a complex K3 surface. ∎

3.3 Fully faithfulness

In Section 3.1 and Section 3.2 we have constructed a functor X(M,F,𝒦) from ordinary K3 surfaces over 𝐅q to triples satisfying (M1)–(M5). We end this section by showing that this functor is fully faithful.

Proof of fully faithfulness in Theorem B.

This is shown in [15] and [20, Theorem 3.3] for K3 surfaces equipped with an ample line bundle. The same argument works here, we repeat it for the convenience of the reader.

Faithfulness. Assume that f,g:X1X2 are morphisms between ordinary K3 surfaces inducing the same maps H2(X2,canι,𝐙)H2(X1,canι,𝐙). Then fcanι=gcanι as maps from X1,canι to X2,canι and therefore fcan=gcan and f=g.

Fullness. Let X1 and X2 be ordinary K3 surfaces over 𝐅q. Let

φ:H2(X2,canι,𝐙)H2(X1,canι,𝐙)

be an isometry commuting with F and respecting ample cones. By the description of the ample cones of X1 and X2, we may choose ample line bundles 1 and 2 on X/𝐅q such that φ maps c1(2,canι) to c1(1,canι).

By Proposition 3.4 the map φ respects the Hodge structures, and therefore the Torelli theorem shows that there is a unique isomorphism f:X1,canιX2,canι with f=φ. Since fF2=F1f, and since the étale cohomology of the Xi,can,K¯ is unramified, we have that

f:Hét2(X2,can,K¯,𝐐)Hét2(X1,can,K¯,𝐐)

is GalK-equivariant, and hence f descends to a morphism of polarized K3 surfaces over K. By Matsusaka–Mumford [13, Theorem 2] this extends to an isomorphism f:X1,canX2,can and we conclude that φ comes from an isomorphism f:X1X2 over 𝐅q. ∎

4 Essential surjectivity

4.1 Models of K3 surfaces with complex multiplication

We briefly recall a few facts about complex K3 surfaces with complex multiplication. We refer to [21] for proofs. Let X/𝐂 be a K3 surface. Its (𝐐-)transcendental lattice VX is defined as the orthogonal complement of NS(X)𝐙𝐐 in H2(X,𝐐(1)). The endomorphism algebra E of the 𝐐-Hodge structure VX is a field, and we say that X has complex multiplication (by E) if VX is one-dimensional as an E-vector space. In this case, E is necessarily a CM-field. Denote its complex conjugation by σ:EE. The Mumford–Tate group of the Hodge structure VX is the algebraic torus T/𝐐 defined by

T(A)={x(AE)×xσ(x)=1}

for all 𝐐-algebras A. Note that T is an algebraic subgroup of SO(VX).

If X/𝐂 is a K3 surface with CM by E, then it can be defined over a number field. In [18, Theorem 2] we classified the models 𝒳 of X over finite extensions F of E in terms of their Galois representations on Hét2(𝒳F¯,𝐙^). We will deduce from that result a version for models over local fields. In the statement we will need the composition

rec:GalEab𝐀E,f×/E×T(𝐀f)/T(𝐐),

where the isomorphism is given by global class field theory (note that E has no real places), and the second map is given by zzσ(z).

Theorem 4.1.

Let X/C be a K3 surface with CM by E. Let K be a p-adic field containing E, and fix an embedding ι:KC extending the embedding EC given by the action of E on H2,0(X). Let

ρ:GalKO(H2(X,𝐙^(1)))

be a continuous homomorphism. Assume that for every σGalK we have

  1. ρ(σ) stabilizes NS(X)H2(X,𝐙^(1)) and 𝒦XNS(X)𝐑,

  2. the restriction of ρ(σ) to the transcendental lattice lands in the subgroup T(𝐀f) of O(VX𝐀f) and its image in T(𝐀f)/T(𝐐) is rec(σ).

Then there exists a model X/K of the K3 surface X so that the resulting action of GalK on Hét2(XK¯,Z^(1))=H2(X,Z^(1)) coincides with ρ.

A profinite group that acts continuously on NS(X) and stabilizes 𝒦X must stabilize an ample class in NS(X). This allows us to reformulate the two conditions on ρ as follows. For an ample λNS(X), denote by ΓλO(H2(X,𝐙^(1))) the subgroup consisting of those g satisfying

  1. g stabilizes NS(X)H2(X,𝐙^(1)) and λNS(X),

  2. the induced action on VX𝐀f factors over T(𝐀f).

Then ρ:GalKO(H2(X,𝐙^(1))) satisfies the conditions in the theorem if and only if there exists an ample class λ such that ρ factors over Γλ and makes the square

commute. In particular, we can think of ρ as a lift of the map rec. This point of view will be useful in the proof of Theorem 4.1.

Lemma 4.2.

The map δ:ΓλT(Af)/T(Q) is a continuous open homomorphism of profinite groups with finite kernel.

Proof.

Let ΓΓλ be the open subgroup of finite index consisting of those elements that act trivially on NS(X), so that Γ is naturally a subgroup of T(𝐀f). Consider the 𝐙-transcendental lattice Λ:=TXH2(X,𝐙(1)). Then Γ consists of those elements gT(𝐀f) that satisfy

  1. g stabilizes Λ𝐙𝐙^TX𝐐𝐀f,

  2. g acts trivially on the discriminant module Δ(Λ)=Λ/Λ.

In particular, Γ is a compact open subgroup of T(𝐀f), and hence of finite index in the maximal compact open subgroup 𝒦T(𝐀f).

It therefore suffices to show that the map δ𝒦:𝒦T(𝐀f)/T(𝐐) is a continuous open homomorphism of profinite groups with finite kernel. We have

𝒦T(𝐐)={x𝒪E×xσ(x)=1},

which is finite because E is a CM-field with complex conjugation σ. It follows that T(𝐐) is discrete in T(𝐀f), that the map δ𝒦 is open, and that kerδ𝒦 is finite. To see that T(𝐀f)/T(𝐐) is profinite, it suffices to show that cokerδ𝒦=T(𝐐)\T(𝐀f)/𝒦 is finite. This is a property of arbitrary tori over 𝐐, see [16, Proposition 9 and Theorem 2]. ∎

Lemma 4.3.

Let δ:G0G1 be an open continuous homomorphism of profinite groups with finite kernel. Let H be a closed subgroup of G1, and ρ:HG0 a continuous homomorphism making the triangle

commute. Then there exists an open subgroup UG1 containing H, and a continuous homomorphism ρ:UG0 making the square

commute.

Proof.

Since the kernel of δ is finite, there exists an open subgroup H0G0 with kerδH0={1}. The normalizer of H0 has finite index in G0, so the intersection

N0:=gG0gH0g-1

is a normal open subgroup on which δ is injective. Denote by N1G1 its image, and by s:N1N0 the inverse isomorphism. Note that N1G1 is open, and normalized by HG.

Consider the continuous function

N1HG0,gs(g)ρ(g)-1.

It takes values in the finite subset kerδG0, and maps 1 to 1. The collection of subgroups of the form N1H (for varying normal open N0G0) is a basis for the topology on H, so shrinking N0 if necessary, we may without loss of generality assume that the above map is constant. We then have s(g)=ρ(g) for all gN1H.

The product U:=N1H is an open subgroup of G containing H, and by the above the map

ρ:UG0,ghs(g)ρ(h)(gN1,hH)

is a well-defined homomorphism satisfying the required properties. ∎

Proof of Theorem 4.1.

Let λNS(X) be an ample class fixed by ρ(GalK) and let G0 be the topological group defined by the cartesian square

Note that δ is open with finite kernel, and hence that G0 is a profinite group. The map ρ induces a commutative triangle

By Lemma 4.3 there exists an intermediate field EFK with F finite over E, and a continuous homomorphism ρ′′:GalFG0 making the diagram

commute. Now [18, Theorem 2] guarantees the existence of a model 𝒳 over F whose Galois action on Hét2(𝒳F¯,𝐙^(1))=H2(X,𝐙^(1)) is given by the composition

GalFρ′′G0Γλ,

and hence the base change of 𝒳 to K fulfils the requirements. ∎

4.2 Criteria of good reduction

Let 𝒪K be a discrete valuation ring with fraction field K and perfect residue field k. In the introduction we defined a property () for K3 surfaces over K.

Theorem 4.4 (Liedtke–Matsumoto [11]).

Let X be a K3 surface over K satisfying (). If for some different from the characteristic of k the action of GalK on Hét2(XK¯,Z) is unramified, then there exists a finite unramified extension KK and a proper smooth algebraic space X over OK with XKXK.∎

Analyzing the proof in the case where the specialization map on Picard groups is bijective, one obtains a stronger conclusion.

Proposition 4.5.

Let X be a K3 surface over K, let KK be an unramified extension, and let X over OK be a proper smooth algebraic space with XKXK. If the reduction map PicXK¯PicXk¯ is bijective, then there exists a smooth projective X over OK with XKX.

Proof.

The map Pic𝔛K¯Pic𝔛k¯ identifies the (-2)-classes on generic and special fiber, and hence induces a bijection between the ample cones in Pic𝔛K¯ and Pic𝔛k¯ (see also the proof of Proposition 3.5). Now choose an ample line bundle on X. It induces an ample line bundle on X:=XK, which extends to a relatively ample line bundle on 𝔛. In particular, the canonical RDP model P(X,) over 𝒪K of Liedtke and Matsumoto [11, Theorem 1.3] is non-singular. It follows from the construction of this model that

P(X,)=P(X,)𝒪K𝒪K

(see [3, end of Section 6]), and hence P(X,) is a smooth projective model 𝔛 over 𝒪K. ∎

Alternatively, one can verify that under the hypothesis of Proposition 4.5 the group 𝒲X,nr occurring in [3, Theorem 1.4] vanishes.

4.3 Proof of Theorem B

In Section 3.3 we have established that the functor

X/𝐅q(M,F,𝒦)

is fully faithful. To finish the proof of Theorem B, it remains to show that the functor is essentially surjective, assuming () holds for K3 surfaces over p-adic fields.

Proof of essential surjectivity in Theorem B.

Let (M,F,𝒦) be a triple satisfying properties (M1)–(M5). We will show that it lies in the essential image of our functor by constructing a suitable K3 surface over 𝐅q. We divide the construction in several steps.

Construction of a complex K3 surface. By Lemma 3.3 the decomposition

M𝐂=M2,0M1,1M0,2

with Ms,2-s:=Ms𝐙p,ι𝐂 defines a 𝐙-Hodge structure on M. The submodule NS(M,F) consists precisely of the Hodge classes in this Hodge structure. By the Torelli theorem for complex K3 surfaces, there is a projective K3 surface X and a Hodge isometry f:H2(X,𝐙)M mapping the ample cone of X to 𝒦. The pair (X,f) is unique up to unique isomorphism.

The complex K3 surface X has complex multiplication. Let VXH2(X,𝐐(1)) be the transcendental lattice. Note that F respects the decomposition

H2(X,𝐐(1))=NS(X)𝐐VX.

Every 𝐐-linear endomorphism of VX that commutes with F will respect the Hodge structure on VX, and since the endomorphism algebra of the 𝐐-Hodge structure VX is a field, we conclude that VX is a cyclic 𝐐[F]-module, that E:=EndVX is generated by F, and that dimEVX=1. In particular, X has complex multiplication by E, the field E is then a CM field, and if we denote the complex conjugation on E by σ, then the Mumford–Tate group T of VX satisfies

T(𝐐)={xE×xσ(x)=1}.

Observe that σ(F)=q2/F on VX, and hence F/q defines an element of T(𝐐).

The number field E has a unique place vp satisfying v(F/q)>0. We have Ev=𝐐p. Let K be the fraction field of W(𝐅q), considered as a subfield of 𝐂 via ι.

Descent to KC. For every p consider the unramified GalK-representation

ρ:GalKGL(M𝐙)

given by letting the geometric Frobenius Frob act as F. We also define a p-adic GalK-representation

ρp:GalKGL(M𝐙p)=GL(sMs)

by declaring that the Tate twisted 𝐙p[GalK]-modules Ms(s) are unramified with geometric Frobenius Frob acting as F/qs. The ρ and ρp assemble into an action of GalK on M𝐙^. Denote by M𝐙^(1) its Tate twist. The resulting map

ρ:GalKGL(M𝐙^(1))=GL(H2(X,𝐙^(1))

satisfies

  1. the image of ρ is contained in O(H2(X,𝐙^(1))),

  2. the image of ρ preserves PicX and 𝒦X(PicX)𝐑.

We claim that ρ also satisfies the reciprocity condition in Theorem 4.1.

Indeed, observe that the action of GalK on M𝐙^(1) is abelian. Let xK×. Using Lemma 3.2, we see that the action of the corresponding τ=τ(x)GalKab on M𝐙^(1) satisfies

  1. τ acts on M𝐙(1) by (F/q)v(x) (for p),

  2. τ acts on M0(1)M𝐙p(1) by (NmK/𝐐px)(F/q)v(x),

  3. τ acts on M1(1)M𝐙p(1) by (F/q)v(x),

  4. τ acts on M2(1)M𝐙p(1) by (NmK/𝐐px)-1(F/q)v(x).

(Note that by property (M4) these actions indeed preserve the 𝐙p-lattices Ms(1).)

On the other hand, the decomposition of M𝐙p induces a decomposition

VX𝐐p=V-1V0V1

with dimV-1=dimV1=1. The group E𝐐p acts on V-1VX𝐐p through the factor Ev×𝐐p×, and on V1 through Eσv×𝐐p×. The inclusion of Ev××Eσv×(E𝐐p)× defines a subgroup

Tv,σv={(xv,xσv)Ev××Eσv×xvxσv=1}T(𝐐p).

The compatibility between local and global class field theory and the definition of rec (see Section 4.1) implies that the diagram

in which the map K×Tv,σv maps x to (NmK/𝐐p(x),NmK/𝐐p(x)-1) commutes. We conclude that

K×GalKabGalEabrecT(𝐀f)/T(𝐐)

maps an xK× to the class of the element α=α(x)T(𝐀f) satisfying

  1. α=1 for all p,

  2. αp acts on VX𝐐p=V-1V0V1 by (NmK/𝐐px,1,(NmK/𝐐px)-1).

Since F/q lies in T(𝐐), we see that α(x) and τ(x) define the same element in T(𝐀f)/T(𝐐). This shows that ρ satisfies the requirements of Theorem 4.1, and we conclude that there is a model 𝒳/K of X whose GalK-action on Hét2(𝒳K¯,𝐙^)=M𝐙^ is the prescribed one.

Extension to OK and reduction to k. By construction, the action of GalK on Hét2(XK¯,𝐙) is unramified. By Theorem 4.4, and since we are assuming 𝒳 satisfies (), there exists a finite unramified extension KK so that 𝒳:=𝒳K has good reduction, and hence extends to a proper smooth 𝔛 over 𝒪K.

By Theorem C, this model 𝔛 is the canonical lift of its reduction, and hence by Proposition 1.8 the map Pic𝔛K¯Pic𝔛𝐅¯q is surjective. We conclude with Proposition 4.5 that X/K has a canonical smooth projective model 𝔛/𝒪K. Again Theorem C guarantees that 𝔛 is the canonical lift of its reduction 𝔛k, and we see that the functor of Theorem B maps 𝔛k to the given triple (M,F,𝒦). ∎

4.4 Unconditional results

As above, we fix an embedding ι:W(𝐅q)𝐂.

Theorem 4.6.

The functor X(M,F,K) restricts to an equivalence between the sub-groupoids consisting of

  1. X/𝐅q for which there is an ample PicX𝐅¯q with 2<p-4,

  2. (M,F,𝒦) for which there exists a λM𝒦 satisfying λ2<p-4.

Assuming p5 it also restricts to an equivalence between

  1. X/𝐅q for which PicX𝐅¯q contains a hyperbolic lattice,

  2. (M,F,𝒦) for which NS(M,F) contains a hyperbolic lattice,

and between

  1. X/𝐅q with rkPicX𝐅¯q12,

  2. (M,F,𝒦) with rkNS(M,F)12.

Proof.

In view of Theorem B and Proposition 3.5, we only need to verify that any triple (M,F,𝒦) as in (ii) lies in the essential image of the functor X(M,F,𝒦) on ordinary K3 surfaces. It suffices to show that the relevant X over K=FracW(𝐅q) occurring in the proof in Section 4.3 satisfy ().

By [12, Theorem 1.1] and [10, Section 2] we know that any K3 surface over X with unramified Hét2(XK¯,𝐐), and satisfying one of

  1. there is an ample PicXK¯ with 2<p-4,

  2. PicXK¯ contains a hyperbolic plane and p5,

  3. PicXK¯ has rank 12 and p5,

has potentially good reduction. In particular, any such K3 surface satisfies hypothesis (). In all three cases the argument of Section 4.3 goes through unconditionally. ∎

References

[1] M. Artin and B. Mazur, Formal groups arising from algebraic varieties, Ann. Sci. Éc. Norm. Supér. (4) 10 (1977), no. 1, 87–131. 10.24033/asens.1322Search in Google Scholar

[2] B. Bhatt, M. Morrow and P. Scholze, Integral p-adic Hodge theory, preprint (2016), https://arxiv.org/abs/1602.03148. 10.1007/s10240-019-00102-zSearch in Google Scholar

[3] B. Chiarellotto, C. Lazda and C. Liedtke, A Néron–Ogg–Shafarevich criterion for K3 surfaces, preprint (2017), https://arxiv.org/abs/1701.02945. 10.1112/plms.12237Search in Google Scholar

[4] J. de Jong, A result of Gabber, (2004), http://www.math.columbia.edu/~dejong/papers/2-gabber.pdf. Search in Google Scholar

[5] P. Deligne, Variétés abéliennes ordinaires sur un corps fini, Invent. Math. 8 (1969), 238–243. 10.1007/BF01406076Search in Google Scholar

[6] P. Deligne, La conjecture de Weil pour les surfaces K3, Invent. Math. 15 (1972), 206–226. 10.1007/BF01404126Search in Google Scholar

[7] P. Deligne and L. Illusie, Cristaux ordinaires et coordonnées canoniques, Surfaces algébriques (Orsay 1976–78), Lecture Notes in Math. 868, Springer, Berlin (1981), 80–137. 10.1007/BFb0090647Search in Google Scholar

[8] A. Grothendieck, Le groupe de Brauer. II. Théorie cohomologique, Dix exposés sur la cohomologie des schémas, Adv. Stud. Pure Math. 3, North-Holland, Amsterdam (1968), 67–87. Search in Google Scholar

[9] D. Huybrechts, Lectures on K3 surfaces, Cambridge Stud. Adv. Math. 158, Cambridge University Press, Cambridge 2016. 10.1017/CBO9781316594193Search in Google Scholar

[10] K. Ito, Unconditional construction of K3 surfaces over finite fields with given L-function in large characteristic, Manuscripta Math. (2018), 10.1007/s00229-018-1066-4. 10.1007/s00229-018-1066-4Search in Google Scholar

[11] C. Liedtke and Y. Matsumoto, Good reduction of K3 surfaces, Compos. Math. 154 (2018), no. 1, 1–35. 10.1112/S0010437X17007400Search in Google Scholar

[12] Y. Matsumoto, Good reduction criterion for K3 surfaces, Math. Z. 279 (2015), no. 1–2, 241–266. 10.1007/s00209-014-1365-8Search in Google Scholar

[13] T. Matsusaka and D. Mumford, Two fundamental theorems on deformations of polarized varieties, Amer. J. Math. 86 (1964), 668–684. 10.2307/2373030Search in Google Scholar

[14] N. O. Nygaard, The Tate conjecture for ordinary K3 surfaces over finite fields, Invent. Math. 74 (1983), no. 2, 213–237. 10.1007/BF01394314Search in Google Scholar

[15] N. O. Nygaard, The Torelli theorem for ordinary K3 surfaces over finite fields, Arithmetic and geometry. Vol. I, Progr. Math. 35, Birkhäuser, Boston (1983), 267–276. 10.1007/978-1-4757-9284-3_11Search in Google Scholar

[16] T. Ono, On some arithmetic properties of linear algebraic groups, Ann. of Math. (2) 70 (1959), 266–290. 10.2307/1970104Search in Google Scholar

[17] J.-P. Serre, Local class field theory, Algebraic number theory (Brighton 1965), Thompson, Washington (1967), 128–161. Search in Google Scholar

[18] L. Taelman, Complex multiplication and Shimura stacks, preprint (2017), https://arxiv.org/abs/1707.01236. Search in Google Scholar

[19] J. T. Tate, p-divisible groups, Proceedings of a conference on local fields (Driebergen 1966), Springer, Berlin (1967), 158–183. 10.1007/978-3-642-87942-5_12Search in Google Scholar

[20] J.-D. Yu, Special lifts of ordinary K3 surfaces and applications, Pure Appl. Math. Q. 8 (2012), no. 3, 805–824. 10.4310/PAMQ.2012.v8.n3.a11Search in Google Scholar

[21] Y. G. Zarhin, Hodge groups of K3 surfaces, J. reine angew. Math. 341 (1983), 193–220. 10.1515/crll.1983.341.193Search in Google Scholar

[22] The Stacks Project Authors – Stacks Project, http://stacks.math.columbia.edu, 2017. Search in Google Scholar

Received: 2018-01-03
Revised: 2018-07-12
Published Online: 2018-10-19
Published in Print: 2020-04-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 29.5.2024 from https://www.degruyter.com/document/doi/10.1515/crelle-2018-0023/html
Scroll to top button