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

Automorphisms of a family of surfaces with pg=q=2p_{g}=q=2 and K2=7K^{2}=7

Matteo Penegini and Roberto Pignatelli
Abstract

We compute the automorphism group of all the elements of a family of surfaces of general type with pg=q=2p_{g}=q=2 and K2=7K^{2}=7, originally constructed by C. Rito in [Rit18]. We discuss the consequences of our results towards the Mumford-Tate conjecture.

In memoria di Gianfranco, un amico e un maestro

\Footnotetext

2020 Mathematics Subject Classification: 14J29, 14J10, 14B12

\Footnotetext

Keywords: Surface of general type, Albanese map, Automorphisms of varieties

\Footnotetext

Version: December 31, 2024

1 Introduction

The classification of general type surfaces with low numerical invariants is unanimously considered a very difficult problem to tackle. It is already difficult to construct new surfaces with low numerical invariants. Therefore, as soon as new examples are found, it is natural to test the famous conjectures on them. This short note stems from the question whether it is possible to verify the Mumford-Tate conjecture for surfaces SS of general type with pg=q=2p_{g}=q=2 and K2=7K^{2}=7 first constructed by Rito in [Rit18] and later studied by the authors in [PePi20].

These surfaces SS are obtained as a generically finite double covering of an abelian surface AA, which turns out to be the Albanese variety of SS, branched along a curve with a singular point of type (3,3)(3,3) and no other singularities. These surfaces give rise to three disjoint open subsets in the Gieseker moduli space 2, 2, 7can\mathcal{M}^{\mathrm{can}}_{2,\,2,\,7} which are all irreducible, generically smooth of dimension 2, that we shall denote by 1\mathcal{M}_{1}, 2\mathcal{M}_{2} and 4\mathcal{M}_{4}.

The Mumford-Tate conjecture for surfaces is still an open problem and only in very few examples it has been verified as true, see for example [Moo17a]. A strategy to prove the conjecture for surfaces with pg=q=2p_{g}=q=2 and of maximal Albanese dimension is outlined in the article [CoPe20]. The strategy reduces to finding geometric quotients XX of SS that are K3 surfaces whose weight 2 Hodge structure is a sub-Hodge structure of the weight 2 Hodge structure of SS orthogonal to the the sub-Hodge structure coming from the Albanese surface of SS. This strategy proved to be very successful in many cases, see [CoPe20]. The first question toward exploiting the strategy is to calculate the automorphism of the surfaces SS and then classify all possible quotients. This is the content of the main theorem of this note.

Theorem 1.1.

The automorphism group of the surface SS of general type with pg=q=2p_{g}=q=2 and K2=7K^{2}=7 constructed by Rito in [Rit18] is a product of cyclic groups

Aut(S)G×/2=G×σ,Aut(S)\cong G\times{\mathbb{Z}}/2{\mathbb{Z}}=G\times\langle\sigma\rangle,

where S/σA=Alb(S)S/\langle\sigma\rangle\cong A=Alb(S), while

  1. 1.0.1.

    GG is trivial if [S]4[S]\in\mathcal{M}_{4};

  2. 1.0.2.

    G/4G\cong{\mathbb{Z}}/4 if [S]1[S]\in\mathcal{M}_{1} and satisfies condition 11 of Proposition 3.6

  3. 1.0.3.

    G/2G\cong{\mathbb{Z}}/2 in all the other cases.

The second step of the strategy is to identify the quotients. We have at once

Corollary 1.2.

For all HAut(S)H\leq\textrm{Aut}(S), the quotients X=S/HX=S/H are irregular surfaces, i.e., q(X)1q(X)\geq 1.

This corollary tells us that in order to prove the Mumford-Tate conjecture for these surfaces a new strategy is needed.

Now, let us explain the way in which this paper is organized.

In the second section we recall the construction of the surfaces SS with the calculation of the invariants. The third section is devoted to the proof of the Main Theorem. The section four contains the calculation of the invariants of the quotient surfaces. Finally we include a section five where it is explained the strategy to prove the conjecture for surfaces with pg=q=2p_{g}=q=2 and of maximal Albanese dimension.

Acknowledgments. The first author was partially supported by GNSAGA-INdAM, by PRIN 2020KKWT53 003 - Progetto: Curves, Ricci flat Varieties and their Interactions and by the DIMA - Dipartimento di Eccellenza 2023-2027. The second author was partially supported by supported by the ”National Group for Algebraic and Geometric Structures, and their Applications” (GNSAGA - INdAM) and by the European Union under NextGenerationEU, PRIN 2022 Prot. n. 20223B5S8L.

Notation and conventions. We work over the field \mathbb{C} of complex numbers. By surface we mean a projective, non-singular surface SS, and for such a surface KSK_{S} denotes the canonical class, pg(S)=h0(S,KS)p_{g}(S)=h^{0}(S,\,K_{S}) is the geometric genus, q(S)=h1(S,KS)q(S)=h^{1}(S,\,K_{S}) is the irregularity and χ(𝒪S)=1q(S)+pg(S)\chi(\mathcal{O}_{S})=1-q(S)+p_{g}(S) is the Euler-Poincaré characteristic.

2 The surfaces

In this section we report, for the convenience of the reader, the construction of the families of surfaces under consideration. We use the notation of [PePi20], the main result in this direction is the following one.

Proposition 2.1.

[PePo13b] Let AA be an Abelian surface. Assume that AA contains a reduced curve whose class is 2-divisible in Pic(A)\text{\rm Pic}(A), whose self intersection is 1616, with a unique singular point of type (3,3)(3,3) and no other singularity. Then there exists a generically finite double cover SAS\rightarrow A branched along this curve. Moreover, the numerical invariants of SS are pg(S)=q(S)=2p_{g}(S)=q(S)=2 and KS2=7K_{S}^{2}=7.

In [Rit18] and later in [PePi20] the existence of the abelian surface AA, that has the properties as in Proposition 2.1, is proved. In particular, it is also shown that the double cover coincides with the Albanese map, hence AA is the Albanese variety associated to SS, we denote it by

α:SAlb(S)=A.\alpha\colon S\rightarrow\textrm{Alb}(S)=A.

We can be more precise, AA is isogenous to a product of two elliptic curves T1T_{1} and T2T_{2}. We denote by

ι:AT1×T2\iota\colon A\rightarrow T_{1}\times T_{2}

the isogeny, which is of degree 2. Clearly, AA carries a (1,2)-polarization LL which is a pull-back of a (product) principal polarization via the isogeny ι\iota. In addition, on AA we have two elliptic fibrations fj:ATjf_{j}\colon A\rightarrow T_{j} with fibres Λi\Lambda_{i} with Λi\Lambda_{i} isogenous to TiT_{i} by a degree two isogeny for i,j{1,2}i,j\in\{1,2\}. Notice that the isogeny is given by the restriction of ι\iota to the fibres.

The branching locus of α\alpha is an effective divisor with two irreducible components

C1+t|2L|,C_{1}+t\in|2L|, (1)

where C1=f21(b1)C_{1}=f_{2}^{-1}(b_{1}) is an element of |Λ2||\Lambda_{2}|, with b1T2b_{1}\in T_{2}. While, tt is a curve of geometric genus 33 with a tacnode tangent to C1C_{1} at a point pp. The situation is exemplified in the following Figure 1.

Refer to caption
Figure 1: The Branch Divisor of α\alpha and β\beta

We deduce that

C12=\displaystyle C_{1}^{2}= 0,\displaystyle 0, t2=\displaystyle t^{2}= 8,\displaystyle 8, C1t=\displaystyle C_{1}t= 4,\displaystyle 4, (2)

notice that (C1+t)2=16(C_{1}+t)^{2}=16.

Notice that the branch locus is singular in pp. Therefore, to get a smooth surface SS as a generically finite double cover of AA branched along C1+tC_{1}+t we have to blow up the point pp (see Figure 1) first.

  1. 2.0.1.

    First, we resolve the singularity in pp. To do that, we need to blow up AA twice, first in pp and then in a point infinitely near to pp. Let us denote these two blow ups by

    Bσ4Bσ3A.B^{\prime}\stackrel{{\scriptstyle\sigma_{4}}}{{\longrightarrow}}B\stackrel{{\scriptstyle\sigma_{3}}}{{\longrightarrow}}A.

    On BB^{\prime}, let us denote by FF the exceptional divisor relative to σ4\sigma_{4}, by EE^{\prime} the strict transform of the exceptional divisor EE relative to σ3\sigma_{3}, by C1C_{1} the strict transform of C1C_{1} and, finally, by RR the strict transform of tt (see Figure 1).

    In addition, one gathers the following information: E1E^{\prime}\cong{\mathbb{P}}^{1} and (E)2=2(E^{\prime})^{2}=-2, F1F\cong{\mathbb{P}}^{1} and F2=1F^{2}=-1, g(C1)=1g(C_{1})=1 and C12=2C_{1}^{2}=-2.

  2. 2.0.2.

    Second, we consider a double cover of β:SB\beta\colon S^{\prime}\longrightarrow B^{\prime} ramified over R+C1+ER+C_{1}+E^{\prime} (this is even since t+C1t+C_{1} is even on AA). The surface SS^{\prime} is a surface of general type, not minimal. Indeed, it contains a 1-1-curve, which is E^=β1(E)\hat{E}=\beta^{-1}(E^{\prime}). The ramification divisor is denoted R^+C1^+E^\hat{R}+\hat{C_{1}}+\hat{E}. Notice that C1^\hat{C_{1}} has genus 11 and C1^2=1\hat{C_{1}}^{2}=-1.

  3. 2.0.3.

    Finally, to get SS we contract the 1-1-curve E^\hat{E}.

We can summarize the construction of SS with the following diagram.

S\textstyle{S^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}β\scriptstyle{\beta}B\textstyle{B^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ4\scriptstyle{\sigma_{4}}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}B\textstyle{B\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ3\scriptstyle{\sigma_{3}}A\textstyle{A}

Moreover, the point pp is a [3,3] point, which is not a negligible singularities. A [3,3] point is a pair (x1,x2)(x_{1},\,x_{2}) such that x1x_{1} belongs to the first infinitesimal neighborhood of x2x_{2} and both are triple points for the curve. Thus, we may calculate the invariants of SS by using the formulae in [BHPV03, p. 237]. In those formulas x2x_{2} counts as a triple point (so m2=1m_{2}=1) and x1x_{1} as a quadruple point (so m1=2m_{1}=2). Then

2=2χ(𝒪S)=L2i=12mi(mi1),6=KS2=2L22i=12(mi1)2.2=2\chi(\mathcal{O}_{S^{\prime}})=L^{2}-\sum_{i=1}^{2}m_{i}(m_{i}-1),\quad 6=K_{S^{\prime}}^{2}=2L^{2}-2\sum_{i=1}^{2}(m_{i}-1)^{2}. (3)

Finally, once we contract the 1-1-curve on SS^{\prime}, we obtain hence KS2=7K_{S}^{2}=7 and χ(S)=1\chi(S)=1.

Considering the Abelian varieties A,T1,T2,T1×T2A,T_{1},T_{2},T_{1}\times T_{2} we choose the following points as neutral elements:

p\displaystyle p A,\displaystyle\in A, a3:=\displaystyle a_{3}:= f1(p)T1,\displaystyle f_{1}(p)\in T_{1}, b1:=\displaystyle b_{1}:= f2(p)T2,\displaystyle f_{2}(p)\in T_{2}, (a3,b1)\displaystyle(a_{3},b_{1}) T1×T2.\displaystyle\in T_{1}\times T_{2}.

With this particular choice ι,f1,f2\iota,f_{1},f_{2} are homomorphism of groups too.

The remaining 22-torsion points on each elliptic curve TjT_{j} will be denoted by

a1,a2,a4\displaystyle a_{1},a_{2},a_{4} T1[2],\displaystyle\in T_{1}[2], b2,b3,b4\displaystyle b_{2},b_{3},b_{4} T2[2].\displaystyle\in T_{2}[2].

This yields ι𝒪A𝒪T1(a4a3)𝒪T2(b2b1)\iota_{*}{\mathcal{O}}_{A}^{-}\cong{\mathcal{O}}_{T_{1}}(a_{4}-a_{3})\boxtimes{\mathcal{O}}_{T_{2}}(b_{2}-b_{1}) , where ι𝒪A\iota_{*}{\mathcal{O}}_{A}^{-} is the anti-invariant part of ι𝒪A\iota_{*}{\mathcal{O}}_{A}, see [PePi20, Lemma 3.4] for a detailed proof.

Remark 2.2.

Furthermore in [PePi20] it is proved that

f1(a4+a3)+f2(b3+b1)|2L|,f_{1}^{*}(a_{4}+a_{3})+f_{2}^{*}(b_{3}+b_{1})\in|2L|,

whence

Lf1(a¯)f2(b¯)L\cong f_{1}^{*}(\bar{a})\otimes f_{2}^{*}(\bar{b})

where b¯\bar{b} is a 4-torsion point such that b¯b¯b2\bar{b}\oplus\bar{b}\neq b_{2}. While for a¯\bar{a} we have three possible choices by [PePi20, Proposition 3.6]

  1. 2.0.1.

    a¯=a3\bar{a}=a_{3} (in this case b¯b¯=b4\bar{b}\oplus\bar{b}=b_{4});

  2. 2.0.2.

    a¯\bar{a} is a 2-torsion point such that a¯a4\bar{a}\neq a_{4} (in this case b¯b¯=b4\bar{b}\oplus\bar{b}=b_{4});

  3. 2.0.3.

    a¯\bar{a} is a 4-torsion point such that a¯a¯=a4\bar{a}\oplus\bar{a}=a_{4} (in this case b¯b¯=b3\bar{b}\oplus\bar{b}=b_{3}).

As just remarked all the choices are possible and to each choice corresponds a different irreducible component of the Gieseker moduli space 2,2,7can\mathcal{M}^{\textrm{can}}_{2,2,7} of the canonical models of the surfaces of general type with pg=q=2p_{g}=q=2 and K2=7K^{2}=7. We shall denote these components, following [PePi20, Definition 3.7], by i2,2,7can\mathcal{M}_{i}\subset\mathcal{M}^{\textrm{can}}_{2,2,7} with i{1,2,4}i\in\{1,2,4\} respectively. Note that the index ii equals the order of a¯\bar{a} as torsion point.

3 The automorphisms of the Rito’s surfaces

Consider the abelian variety AA. We know that there is an isogeny of degree 22 onto a product of elliptic curves T1×T2T_{1}\times T_{2}.

By taking the universal covers, we can write Tj:=/λjT_{j}:={\mathbb{C}}/\lambda_{j} where the λj2\lambda_{j}\cong{\mathbb{Z}}^{2} are lattices so that the origin maps to a3a_{3} respectively b1b_{1}. We choose generators e¯1,e¯2\bar{e}_{1},\bar{e}_{2} of λ1\lambda_{1}, e¯3,e¯4\bar{e}_{3},\bar{e}_{4} of λ2\lambda_{2} so that e¯12\frac{\bar{e}_{1}}{2} maps to a1a_{1}, e¯22\frac{\bar{e}_{2}}{2} maps to a2a_{2}, e¯32\frac{\bar{e}_{3}}{2} maps to b3b_{3}, e¯42\frac{\bar{e}_{4}}{2} maps to b4b_{4}.

So, in 2{\mathbb{C}}^{2} -coordinates we have

e1=\displaystyle e_{1}= (e¯1,0),\displaystyle(\bar{e}_{1},0), e2=\displaystyle e_{2}= (e¯2,0),\displaystyle(\bar{e}_{2},0), e3=\displaystyle e_{3}= (0,e¯3),\displaystyle(0,\bar{e}_{3}), e4=\displaystyle e_{4}= (0,e¯4).\displaystyle(0,\bar{e}_{4}).

Since the universal cover of T1×T2T_{1}\times T_{2} factors through the isogeny ι\iota, we obtain A=2/λA={\mathbb{C}}^{2}/\lambda where λ\lambda is a sublattice of index 22 of the lattice

λ1λ2={tiei|ti}\lambda_{1}\oplus\lambda_{2}=\left\{\sum t_{i}e_{i}|t_{i}\in{\mathbb{Z}}\right\}
Lemma 3.1.

The lattice λ\lambda is the sublattice of λ1λ2\lambda_{1}\oplus\lambda_{2} of the elements whose sum ti\sum t_{i} is even.

Proof.

We set λ~:={mjej|mj is even}\tilde{\lambda}:=\left\{\sum m_{j}e_{j}|\sum m_{j}\text{ is even}\right\}, A~:=2/λ~\tilde{A}:={\mathbb{C}}^{2}/\tilde{\lambda}. The inclusion λ~λ1λ2\tilde{\lambda}\subset\lambda_{1}\oplus\lambda_{2} induces an isogeny ι~:A~T1×T2\tilde{\iota}\colon\tilde{A}\rightarrow T_{1}\times T_{2} of degree 22.

An isogeny of degree 22 is determined by the anti-invariant part of the direct image of the trivial bundle, that is a generator of the kernel of the pull-back map among the Picard groups. So we only need to prove that ι~(𝒪T1(a4a3)𝒪T2(b2b1))\tilde{\iota}^{*}\left({\mathcal{O}}_{T_{1}}(a_{4}-a_{3})\boxtimes{\mathcal{O}}_{T_{2}}(b_{2}-b_{1})\right) is the trivial sheaf of A~\tilde{A}.

I

Let λ1λ1\lambda^{\prime}_{1}\subset\lambda_{1} be the index 22 sublattice of the elements of the form m1e¯1+m2e¯2m_{1}\bar{e}_{1}+m_{2}\bar{e}_{2} with m1+m2m_{1}+m_{2} even. In the same way, let λ2λ2\lambda^{\prime}_{2}\subset\lambda_{2} be the index 22 sublattice of the elements of the form m3e¯3+m4e¯4m_{3}\bar{e}_{3}+m_{4}\bar{e}_{4} with m3+m4m_{3}+m_{4} even. These define an isogeny of degree 22, ι~i:Ti=/λiTi\tilde{\iota}_{i}\colon T^{\prime}_{i}={\mathbb{C}}/\lambda^{\prime}_{i}\rightarrow T_{i}, for i=1,2i=1,2. We can derive the following commutative diagram

T1×T2\textstyle{T^{\prime}_{1}\times T^{\prime}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T1\textstyle{T_{1}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι~1\scriptstyle{\tilde{\iota}_{1}}A~\textstyle{\tilde{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι~\scriptstyle{\tilde{\iota}}T2\textstyle{T^{\prime}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι~2\scriptstyle{\tilde{\iota}_{2}}T1×T2\textstyle{T_{1}\times T_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T1\textstyle{T_{1}}T2\textstyle{T_{2}} (4)

Notice that ι~2𝒪T2(b2b1)\tilde{\iota}_{2}^{*}{\mathcal{O}}_{T_{2}}(b_{2}-b_{1}) is trivial on /λ2{\mathbb{C}}/\lambda^{\prime}_{2}. This is standard; it can be show for example as follows.

The point b1b_{1} pulls back to the sum of two points, the classes modulo λ2\lambda^{\prime}_{2} of 0 and e¯3\bar{e}_{3}. The point b2b_{2} pulls back to the sum of the classes of 12(e¯3+e¯4)\frac{1}{2}(\bar{e}_{3}+\bar{e}_{4}) and e¯3+12(e¯3+e¯4)\bar{e}_{3}+\frac{1}{2}(\bar{e}_{3}+\bar{e}_{4}). Since

(12(e¯3+e¯4)+e¯3+12(e¯3+e¯4))(0+e¯3)=e¯3+e¯4λ2\left(\frac{1}{2}(\bar{e}_{3}+\bar{e}_{4})+\bar{e}_{3}+\frac{1}{2}(\bar{e}_{3}+\bar{e}_{4})\right)-(0+\bar{e}_{3})=\bar{e}_{3}+\bar{e}_{4}\in\lambda^{\prime}_{2}

the two divisors of degree 22 we have obtained are linearly equivalent.

In the same way we can prove that ι~1𝒪T1(a4a3)\tilde{\iota}_{1}^{*}{\mathcal{O}}_{T_{1}}(a_{4}-a_{3}) is trivial on /λ1{\mathbb{C}}/\lambda^{\prime}_{1}.

The 2-torsion line bundles on T1×T2T_{1}\times T_{2} that pull back to trivial bundles on T1×T2T^{\prime}_{1}\times T^{\prime}_{2} are the line bundles: 𝒪T1(a4a3)𝒪T2{\mathcal{O}}_{T_{1}}(a_{4}-a_{3})\boxtimes{\mathcal{O}}_{T_{2}}, 𝒪T1𝒪T2(b2b1){\mathcal{O}}_{T_{1}}\boxtimes{\mathcal{O}}_{T_{2}}(b_{2}-b_{1}) and 𝒪T1(a4a3)𝒪T2(b2b1){\mathcal{O}}_{T_{1}}(a_{4}-a_{3})\boxtimes{\mathcal{O}}_{T_{2}}(b_{2}-b_{1}). Exactly one of them pulls back to the trivial line bundle to A~\tilde{A}.

We can conclude the proof observing that if ι~(𝒪T1(a4a3)𝒪T2)\tilde{\iota}^{*}\left({\mathcal{O}}_{T_{1}}(a_{4}-a_{3})\boxtimes{\mathcal{O}}_{T_{2}}\right) were trivial on A~\tilde{A} than this would imply that there were a fibration form A~\tilde{A} onto T1T^{\prime}_{1} and this is absurd. In the same way we exclude the case 𝒪T1𝒪T2(b2b1){\mathcal{O}}_{T_{1}}\boxtimes{\mathcal{O}}_{T_{2}}(b_{2}-b_{1}). ∎

We will need the following general result for an abelian surface with a (1,2)(1,2)-polarization, the proof of which can be found in [Bar87, Section 1.2].

Remark 3.2.

The linear system |L||L| contains exactly two reducible divisors union of elements respectively of Λ1\Lambda_{1} and Λ2\Lambda_{2}, the curves f1a¯+f2b¯f_{1}^{*}\bar{a}+f_{2}^{*}\bar{b} and f1(a¯a4)+f2(b¯b2)f_{1}^{*}(\bar{a}\oplus a_{4})+f_{2}^{*}(\bar{b}\oplus b_{2}).

Since the Albanese morphism α:SA\alpha\colon S\longrightarrow A has degree 22, it determines an involution σ:SS\sigma\colon S\rightarrow S that is central in AutS\text{\rm Aut}\ S and an exact sequence

0/2=σAutSG00\rightarrow{\mathbb{Z}}/2{\mathbb{Z}}=\langle\sigma\rangle\rightarrow\text{\rm Aut}\ S\rightarrow G\rightarrow 0

where GG is the group of the self-biholomorphisms φ:AA\varphi\colon A\rightarrow A such that

  1. 3.0.1.

    φL=L\varphi^{*}L=L;

  2. 3.0.2.

    φ(C1+t)=C1+t\varphi(C_{1}+t)=C_{1}+t, equivalently φ(C1)=C1\varphi(C_{1})=C_{1}, φ(t)=t\varphi(t)=t, φ(p)=p\varphi(p)=p.

We have written an isomorphism A2/λA\cong{\mathbb{C}}^{2}/\lambda where the point pp is the image of the origin of 2{\mathbb{C}}^{2}. From φ(p)=p\varphi(p)=p it follows that the elements of GG are automorphisms of AA as a group with the group structure induced by 2{\mathbb{C}}^{2}. Now, we can see that one of the above assumption is not necessary, indeed we have the following lemma.

Lemma 3.3.

The group GG is the group of the automorphisms of the Abelian variety AA preserving the group structure induced by the identification 2/λ=A{\mathbb{C}}^{2}/\lambda=A such that φL=L\varphi^{*}L=L and φ(C1)=C1\varphi(C_{1})=C_{1}.

Proof.

The only nontrivial thing to prove is that, if φ:AA\varphi\colon A\rightarrow A is a group automorphism such that φL=L\varphi^{*}L=L and φ(C1)=C1\varphi(C_{1})=C_{1}, then φ(t)=t\varphi(t)=t. Notice that by hypothesis and equation (1) we have that tt is linearly equivalent to φ(t)\varphi(t).

By (2), the intersection number between tt and C1C_{1} is 44. More precisely, tt cuts on C1C_{1} the divisor 4p4p. Since φ(p)=p\varphi(p)=p and φ(C1)=C1\varphi(C_{1})=C_{1}, φ(t)\varphi(t) cuts 4p4p on C1C_{1} as well.

Assume by contradiction φ(t)t\varphi(t)\neq t, then the functions defining them span a subspace VH0(A,t)V\subset H^{0}(A,t) of dimension 22. By what we just said tt and φ(t)\varphi(t) cut on C1C_{1} the same divisor 4p4p, therefore the restriction map ρ:VH0(C1,t|C1)\rho\colon V\longrightarrow H^{0}(C_{1},t|_{C_{1}}) has rank 11. By dimension count the kernel of ρ\rho is a one dimensional subspace generated by say ss. Then C1{s=0}C_{1}\subset\{s=0\}, let us call D1D_{1} the residue curve. By definition D1D_{1} is an effective divisor in |tC1|=|2L2C1||t-C_{1}|=|2L-2C_{1}|.

Notice that LL is numerically equivalent to Λ1+Λ2\Lambda_{1}+\Lambda_{2} while C1C_{1} is numerically equivalent to Λ2\Lambda_{2}. Since D1D_{1} is linearly equivalent to 2L2C12L-2C_{1} if follows that it is numerically equivalent to 2Λ12\Lambda_{1}, hence the intersection product D1Λ1=0D_{1}\cdot\Lambda_{1}=0. Since any element in |Λ1||\Lambda_{1}| is irreducible we have that D1D_{1} is union of two of such elements. Let us denote these two elements them by AA and BB and we get numerically D1=A+BD_{1}=A+B.

Let DD be a further element of |Λ1||\Lambda_{1}|. Then DD is isomorphic to /λ2{\mathbb{C}}/\lambda_{2}^{\prime}. More precisely, we have the following diagram

A~\textstyle{\tilde{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι~\scriptstyle{\tilde{\iota}}T1×T2\textstyle{T_{1}\times T_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}/λ2\textstyle{{\mathbb{C}}/\lambda^{\prime}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ξ\scriptstyle{\xi}ι~2\scriptstyle{\tilde{\iota}_{2}}T2\textstyle{T_{2}} (5)

where ξ\xi maps isomorphically /λ2{\mathbb{C}}/\lambda^{\prime}_{2} onto DD. Notice that ξ\xi, being an isomorphism, allows us to see DD as a degree 22 étale cover of T2T_{2} via the composition with the isogeny ι~2\tilde{\iota}_{2}. Since AA or BB restricted to DD are trivial so is the restriction of D1D_{1}. A fortiori the restrictions to DD of 2L2L and 2C12C_{1} are linearly equivalent, which means that the restriction to DD of LL and C1C_{1} differ by 22-torsion.

The restriction to DD of C1C_{1} is ι~2b1\tilde{\iota}_{2}^{*}b_{1} since b1b_{1} is f2(p)f_{2}(p). Moreover the restriction to DD of LL is ι~2b¯\tilde{\iota}_{2}^{*}\bar{b} by Remark 2.2. Hence the 44-torsion point b1b¯b_{1}-\bar{b} in T2T_{2} lifts to a 22-torsion point in DD. Thus the line bundle 𝒪T2(2(b1b¯)){\mathcal{O}}_{T_{2}}(2(b_{1}-\bar{b})) is the only 22-torsion on T2T_{2} that lifts to the trivial bundle on DD. This implies (see also proof of Lemma 3.1) that b¯b¯=b2\bar{b}\oplus\bar{b}=b_{2} but this is absurd because by Remark 2.2 we have either b¯b¯=b3\bar{b}\oplus\bar{b}=b_{3} or b¯b¯=b4\bar{b}\oplus\bar{b}=b_{4}. ∎

The action of GG on AA may be uniquely lifted to an action on 2{\mathbb{C}}^{2} fixing the origin, so representing GG as a finite subgroup of the linear group GL2()GL_{2}({\mathbb{C}}) of the matrices preserving the lattice λ\lambda. We will then write φ\varphi as a matrix

φ=(φ11φ12φ21φ22)\varphi=\begin{pmatrix}\varphi_{11}&\varphi_{12}\\ \varphi_{21}&\varphi_{22}\end{pmatrix}
Lemma 3.4.

GG is a group of diagonal matrices. In other words φ12=φ21=0\varphi_{12}=\varphi_{21}=0.

Proof.

We consider 2{\mathbb{C}}^{2} with the natural coordinates (x,y)(x,y) given by the construction, so that the line x=0x=0 is the connected component through the origin of the preimage of C1C_{1} in 2{\mathbb{C}}^{2}. Since φ(C1)=C1\varphi(C_{1})=C_{1} the matrix ϕ\phi preserves x=0x=0, so φ12=0\varphi_{12}=0.

Now we use that φL=L\varphi^{*}L=L. Since by Remark 3.2, |L||L| contains exactly two reducible divisors, φ\varphi either preserves or exchange them. The preimage of both divisors on 2{\mathbb{C}}^{2} is a union of countably many ”horizontal” lines y=cy=c and countably many vertical lines x=cx=c. However, if φ210\varphi_{21}\neq 0 any horizontal line is mapped to a line that is neither horizontal nor vertical, a contradiction. ∎

In particular φ\varphi is given by two roots of the unity φjj\varphi_{jj}\in{\mathbb{C}} giving automorphisms of the two elliptic curves: each φjj\varphi_{jj} gives an automorphism of TjT_{j}.

We will now need the following well known facts on automorphisms of elliptic curves see for instance [Sil08, Section III.10].

Lemma 3.5.

Let ω\omega be a nontrivial automorphism of an elliptic curve of order nn. Then n=2,3,4n=2,3,4 or 66. Moreover

  1. 3.0.1.

    for every 44-torsion point pTp\in T, ω(p)p\omega(p)\neq p;

  2. 3.0.2.

    if n=3,6n=3,6, then for every 22-torsion point pTp\in T, ω(p)p\omega(p)\neq p;

  3. 3.0.3.

    if n=4n=4 then there is exactly one 22-torsion point pTp\in T such that ω(p)=p\omega(p)=p;

  4. 3.0.4.

    if n=2n=2 then ω(p)=p\omega(p)=p for all 22-torsion points pTp\in T.

With these facts in mind we can prove

Proposition 3.6.

We have the following possibilities for the group GG according to the cases in Remark 2.2.

  • »

    In case 11: if T1T_{1} has an automorphism of order 44, GG is cyclic of order 44 generated by the automorphism given by φ11=i\varphi_{11}=i, φ22=1\varphi_{22}=1, that is

    (x,y)(ix,y).(x,y)\mapsto(ix,y).
  • »

    In case 33: GG is the trivial group of order 11.

  • »

    In the remaining cases, GG is a cyclic group of order 22 generated by the automorphism given by φ11=1\varphi_{11}=-1, φ22=1\varphi_{22}=1, the involution

    (x,y)(x,y)(x,y)\mapsto(-x,y)
Proof.

By Lemma 3.4 φ\varphi acts on the fibrations fj:ATjf_{j}\colon A\rightarrow T_{j} acting on the codomain by φjj\varphi_{jj}.

Since by Remark 3.2, |L||L| contains exactly two reducible divisors, f1a¯+f2b¯f_{1}^{*}\bar{a}+f_{2}^{*}\bar{b} and f1(a¯a4)+f2(b¯b2)f_{1}^{*}(\bar{a}\oplus a_{4})+f_{2}^{*}(\bar{b}\oplus b_{2}), φ\varphi either preserves or exchange them. So, on T1×T2T_{1}\times T_{2}, the matrix φ\varphi maps (a¯,b¯)(\bar{a},\bar{b}) either to (a¯,b¯)(\bar{a},\bar{b}) or to (a¯a4,b¯b2)(\bar{a}\oplus a_{4},\bar{b}\oplus b_{2}).

We now show φ22=1\varphi_{22}=1. In fact in both cases φ222(b¯)=b¯\varphi_{22}^{2}(\bar{b})=\bar{b}. Since b¯\bar{b} is a 4-torsion point, by Lemma 3.5, part 1, φ222=1\varphi_{22}^{2}=1. Moreover, if φ221\varphi_{22}\neq 1 (so φ22=1\varphi_{22}=-1) φ22(b¯)=b¯b2\varphi_{22}(\bar{b})=\bar{b}\oplus b_{2} that implies b¯b¯=b2\bar{b}\oplus\bar{b}=b_{2}, a contradiction. So φ22=1\varphi_{22}=1.

As a first consequence, φ11(a¯)=a¯\varphi_{11}(\bar{a})=\bar{a}.

Now recall that the matrix φ\varphi preserves the lattice λ\lambda. Since φ22=1\varphi_{22}=1, then φ11\varphi_{11} preserves λ1\lambda_{1} and the index 22 sublattice λ1λ1\lambda^{\prime}_{1}\subset\lambda_{1} of the elements of the form m1e¯1+m2e¯2m_{1}\bar{e}_{1}+m_{2}\bar{e}_{2} with m1+m2m_{1}+m_{2} even. This implies φ11(e¯1+e¯2)(e¯1+e¯2)2λ1\varphi_{11}(\bar{e}_{1}+\bar{e}_{2})-(\bar{e}_{1}+\bar{e}_{2})\in 2\lambda_{1}. Dividing by 22 we obtain φ11(a4)=a4\varphi_{11}(a_{4})=a_{4}.

Now we distinguish the three cases according to Remark 2.2.

In case 33, a¯\bar{a} is a 44-torsion point. Then by φ11(a¯)=a¯\varphi_{11}(\bar{a})=\bar{a} and Lemma 3.5, part 1, φ11=1\varphi_{11}=1.

In case 22, a¯\bar{a} and a4a_{4} are distinct 22-torsion points fixed by φ11\varphi_{11}. Then by Lemma 3.5, part 2 and 3, φ11\varphi_{11} has order 11 or 22, so φ11=±1\varphi_{11}=\pm 1. On the other hand the map (x,y)(x,y)(x,y)\mapsto(-x,y) preserves λ\lambda so it defines an automorphism of AA that defines an element of GG.

Finally, case 11. In this case φ11(a¯)=a¯\varphi_{11}(\bar{a})=\bar{a} holds indipendently by the choice of the complex number of φ11\varphi_{11}. Still we have the condition φ11(a4)=a4\varphi_{11}(a_{4})=a_{4} that by Lemma 3.5, part 2, forces φ114=1\varphi_{11}^{4}=1. If T1T_{1} has no automorphisms of order 44, then we obtain φ11=±1\varphi_{11}=\pm 1 and we conclude as in case 2. Else analogous argument shows that (x,y)(ix,y)(x,y)\mapsto(ix,y) generates GG. ∎

Then we can compute Aut(S)\text{\rm Aut}(S).

Proposition 3.7.
Aut(S)G×/2Aut(S)\cong G\times{\mathbb{Z}}/2{\mathbb{Z}}
Proof.

Choose any element gGg\in G. Then by 3.6 gg acts as (x,y)(kx,y)(x,y)\mapsto(kx,y) in the coordinates considered there, where kk is a complex number with k4=1k^{4}=1. In those coordinates C1C_{1} is defined by yy.

Let ZZ be the finite double cover of AA birational to SS, and let qZq\in Z be the unique point over p=(0,0)p=(0,0). In a neighbourhood of qq, ZZ has equation z2=f(x,y)z^{2}=f(x,y) where ff is an equation of the branch locus, that is geometrically gg-invariant. Then there is a constant cc such that gf=cfg^{*}f=cf. From f(x,y)=y3+O(4)f(x,y)=y^{3}+O(4) we deduce c=1c=1 and gf=fg^{*}f=f.

The involution on ZZ induced by the Albanese morphism of SS acts near qq as (x,y,z)(x,y,z)(x,y,z)\rightarrow(x,y,-z); gg acts (x,y)(kx,y)(x,y)\rightarrow(kx,y), so the liftings of gg act as (x,y,z)(kx,y,±z)(x,y,z)\mapsto(kx,y,\pm z). So the liftings acting locally trivially on the variable zz form a splitting map GAut(S)G\rightarrow\text{\rm Aut}(S) mapping to a subgroup that commutes with the Albanese involution. ∎

4 The quotients of SS

In this brief section we shall prove Corollary 1.2. By Proposition 3.7 Aut(S)=σ×G\text{\rm Aut}(S)=\langle\sigma\rangle\times G let us consider HAut(S)H\leq\text{\rm Aut}(S), then we have the following diagram

S\textstyle{S^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A\textstyle{A}XS/H\textstyle{X\cong S/H}

A natural question to address is the classification of quotient surfaces S/HS/H.

A first step in studying the quotient surfaces is to determine their numerical invariants. To this end we study the induced action of the group HH on the cohomology groups of SS. Recall that the the global sections of H1,0(S)H^{1,0}(S) comes from the one forms on AA that we denote by dx,dydx,dy. Moreover, one of the two generators of the global sections of H2,0(S)H^{2,0}(S) can be identified with dxdydx\wedge dy, and of course being pg(S)=2p_{g}(S)=2 we have a global 2-forms ω\omega not coming from AA. To summarize this we can write

H2,0(S)dxdy,ω,\displaystyle H^{2,0}(S)\cong\langle dx\wedge dy,\omega\rangle, H1,0(S)dx,dy.\displaystyle H^{1,0}(S)\cong\langle dx,dy\rangle.

Let us denote by gg the generator of the cyclic group GG, which has order 22 or 44 according to the three cases of Proposition 3.6. The same proposition describes completely the induced action on the generators of the cohomology groups in each case. In particular, we have

dxgdx={dx,idx\displaystyle dx\mapsto g^{*}dx=\begin{cases}-dx,\\ idx\end{cases} dygdy=dy.\displaystyle dy\mapsto g^{*}dy=dy.

If HH is trivial or H<σ>H\cong<\sigma> then we know that the quotients are respectively SS or birational to AA. Else HG{1}H\cap G\neq\{1\} and this yields at once that q(X)=1q(X)=1. More precisely, H1(X)<dy>H^{1}(X)\cong<dy>, and thus we have proved Corollary 1.2. We can remark already that no quotient XX can be a K3 surface. Moreover, we have

g(dxdy)={dxdy,idxdy.g^{*}(dx\wedge dy)=\begin{cases}-dx\wedge dy,\\ idx\wedge dy.\end{cases}

Therefore we have that pg(X)p_{g}(X) is either 0 or 11 according to the HH-invariance of ω\omega .

5 Towards the Mumford-Tate conjecture

This section is devoted to explain the strategy used up to know for proving the Mumford Tate conjecture for surfaces with pg(S)=q(S)=2p_{g}(S)=q(S)=2 and of maximal Albanese dimension.

Let SS be a smooth projective complex surface with invariants pg(S)=q(S)=2p_{g}(S)=q(S)=2, and assume that the Albanese morphism α:SA\alpha\colon S\to A is surjective. We can make the following general observations (see also [CoPe20]). It holds:

  1. 5.0.1.

    The induced map on cohomology α:H(A,)H(S,)\alpha^{*}\colon H^{*}(A,{\mathbb{Z}})\to H^{*}(S,{\mathbb{Z}}) is injective. The orthogonal complement Hnew2=H(A,)H(S,)H^{2}_{\textnormal{new}}=H^{*}(A,{\mathbb{Z}})^{\perp}\subset H^{*}(S,{\mathbb{Z}}) is a Hodge structure of weight 22 with Hodge numbers (1,n,1)(1,n,1), where n=h1,1(S)4n=h^{1,1}(S)-4. Such a Hodge structure is said to be of K3 type.

  2. 5.0.2.

    Let SS^{\prime} be a smooth projective complex surface with invariant pg(S)=1p_{g}(S^{\prime})=1. Then Morrison [Mor87] showed that there exists a K3 surface XX^{\prime} together with an isomorphism ι:H2(S,)traH2(X,)tra\iota^{\prime}\colon H^{2}(S^{\prime},{\mathbb{Q}})^{\textnormal{tra}}\to H^{2}(X^{\prime},{\mathbb{Q}})^{\textnormal{tra}} that preserves the Hodge structure, the integral structure, and the intersection pairing. (Here (_)tra(\_)^{\textnormal{tra}} denotes the transcendental part of the Hodge structure, that is, the orthogonal complement of the Hodge classes.)

We now look closely to our surfaces SS with pg(S)=q(S)=2p_{g}(S)=q(S)=2, for which we know that the Albanese map is a generically finite cover.

Then we have

Proposition 5.1.

Let SS be a smooth projective complex surface with invariants pg(S)=q(S)=2p_{g}(S)=q(S)=2, and assume that the Albanese morphism α:SA\alpha\colon S\to A is surjective. Then there exists a K3 surface XX and an isomorphism of Hodge structures

ι:(Hnew2(S,))traH2(X,)tra.\iota\colon(H^{2}_{\textnormal{new}}(S,{\mathbb{Q}}))^{\textnormal{tra}}\to H^{2}(X,{\mathbb{Q}})^{\textnormal{tra}}.

This is a direct consequence of [Mor87]. Notice that the surface XX is related only Hodge theoretically to SS. Therefore, this is not enough to prove the conjecture, to this end we have to address the following question:

Do there exist XX and ι\iota as above, such that ι\iota is motivated in the sense of Andrè?

Let us briefly explain and recall some facts on categories of motives, and for the reader convenience we state the motivic Mumford–Tate conjecture, for a more detailed introduction on the subject see [Moo17b, Moo17a]. First we recall some facts about Chow motives and André motives of surfaces. We do not need full generality, so let KK be a subfield of {\mathbb{C}}.

Given smooth and projective varieties XX and YY over a field KK (i.e., objects in the category SmPr/K\textnormal{SmPr}_{/K}) of dimension dXd_{X} and dYd_{Y} respectively, a correspondence of degree kk from XX to YY is an element γ\gamma of AdX+k(X×Y)A^{d_{X}+k}(X\times Y). Then γ\gamma induces a map A(X)A+k(Y)A^{\cdot}(X)\rightarrow A^{\cdot+k}(Y) by the formula

γ(β):=π2(γπ1(β)),\gamma_{*}(\beta):=\pi_{2*}(\gamma\cdot\pi^{*}_{1}(\beta)),

where π1:X×YX\pi_{1}\colon X\times Y\rightarrow X and π2:X×YY\pi_{2}\colon X\times Y\rightarrow Y denote the projections. The category rat\mathcal{M}_{\textnormal{rat}} of Chow motives (with rational coefficients) over KK is defined as follows:

  • »

    the objects of rat\mathcal{M}_{\textnormal{rat}} are triples (X,p,n)(X,p,n) such that XSmPrK,pAdX(X×X)X\in\textnormal{SmPr}_{K},\,p\in A^{d_{X}}(X\times X) is an idempotent correspondence (i.e. pp=pp_{*}\circ p_{*}=p_{*}) and nn is an integer;

  • »

    the morphisms in rat\mathcal{M}_{\textnormal{rat}} from (X,p,n)(X,p,n) to (Y,q,m)(Y,q,m) are correspondences f:XYf\colon X\rightarrow Y of degree nmn-m, such that fp=f=qff\circ p=f=q\circ f.

We recall that rat\mathcal{M}_{\textnormal{rat}} is an additive, {\mathbb{Q}}-linear, pseudoabelian category, see [Sch91, Theorem 1.6].

We consider from here only the cases in which we are interested hence let us suppose K=K={\mathbb{C}}, There exists a functor

h:SmPr/opratsuch that h:X(X,id,0)h\colon\textnormal{SmPr}_{/{\mathbb{C}}}^{\textnormal{op}}\to\mathcal{M}_{\textnormal{rat}}\ \ \mbox{such that }h:X\mapsto(X,\textrm{id},0)

from the opposite category of smooth projective varieties over {\mathbb{C}} to the category of Chow motives.

We denote also with Ai(M)A^{i}(M) the ii-th Chow group of a motive MratM\in\mathcal{M}_{\textnormal{rat}}. In general, it is not known whether the Künneth projectors πi\pi_{i} are algebraic, so it does not (yet) make sense to speak of the summand hi(X)h(X)h^{i}(X)\subset h(X) for an arbitrary smooth projective variety X/X/{\mathbb{C}}. However, a so-called Chow–Künneth decomposition does exist for curves [Man68], for surfaces [Mur90], and for abelian varieties [DeMu91]. For algebraic surfaces there is in fact the following theorem, which strengthens the decomposition of the Chow motive. Statement is copied from [Lat19, Theorem 2.2].

Theorem 5.2.

Let SS be a smooth projective surface over {\mathbb{C}}. There exists a self-dual Chow–Künneth decomposition {πi}\{\pi_{i}\} of SS, with the further property that there is a splitting

π2=π2alg+π2traA2(S×S)\pi_{2}=\pi_{2}^{\textnormal{alg}}+\pi_{2}^{\textnormal{tra}}\quad\in A^{2}(S\times S)

in orthogonal idempotents, defining a splitting h2(S)=halg2(S)htra2(S)h^{2}(S)=h^{2}_{\textnormal{alg}}(S)\oplus h^{2}_{\textnormal{tra}}(S) with Chow groups

Ai(halg2(S))={NS(S)if i=1,0otherwise,andAi(htra2(S))={AAJ2(S)if i=2,0otherwise.A^{i}(h^{2}_{\textnormal{alg}}(S))=\begin{cases}\textrm{NS}(S)&\text{if $i=1$,}\\ 0&\text{otherwise,}\end{cases}\qquad\text{and}\quad A^{i}(h^{2}_{\textnormal{tra}}(S))=\begin{cases}A^{2}_{\textnormal{AJ}}(S)&\text{if $i=2$,}\\ 0&\text{otherwise.}\end{cases}

Here AAJ2(S)A^{2}_{\textnormal{AJ}}(S) denotes the kernel of the Abel–Jacobi map.

Proof.

For the proof see [Lat19, Theorem 2.2] and references therein. ∎

The idea to construct such K3 surface XX is to exploit the automorphism group GG of SS and prove that a quotient S/HS/H by some subgroup HGH\leq G is birational to XX. Of course this will give us a weak answer to the previous question. Nevertheless, it will suffices to prove that – using the notion of motivated cycles introduced by André [And96] – there exist XX and ι\iota as above, such that ι\iota is motivated.

To speak about motivic Mumford–Tate conjecture we need to introduce the notion of motivated cycles (for a brief introduction see e.g. [Moo17b, Section 3.1]).

Definition 5.3.

Let KK be a subfield of {\mathbb{C}}, and let XX be a smooth projective variety over KK. A class γ\gamma in H2i(X)H^{2i}(X) is called a motivated cycle of degree ii if there exists an auxiliary smooth projective variety YY over KK such that γ\gamma is of the form π(αβ)\pi_{*}(\alpha\cup\star\beta), where π:X×YX\pi\colon X\times Y\to X is the projection, α\alpha and β\beta are algebraic cycle classes in H(X×Y)H^{*}(X\times Y), and β\star\beta is the image of β\beta under the Hodge star operation.

Every algebraic cycle is motivated, and under the Lefschetz standard conjecture the converse holds as well. The set of motivated cycles naturally forms a graded {\mathbb{Q}}-algebra. The category of motives over KK, denoted MotK\textnormal{Mot}_{K}, consists of objects (X,p,m)(X,p,m), where XX is a smooth projective variety over KK, pp is an idempotent motivated cycle on X×XX\times X, and mm is an integer. A morphism (X,p,m)(Y,q,n)(X,p,m)\to(Y,q,n) is a motivated cycle γ\gamma of degree nmn-m on Y×XY\times X such that qγp=γq\gamma p=\gamma. We denote with Hmot(X)\mathrm{H}_{\textnormal{mot}}(X) the object (X,Δ,0)(X,\Delta,0), where Δ\Delta is the class of the diagonal in X×XX\times X. The Künneth projectors πi\pi_{i} are motivated cycles, and we denote with Hmoti(X)\mathrm{H}_{\textnormal{mot}}^{i}(X) the object (X,πi,0)(X,\pi_{i},0). Observe that Hmot(X)=iHmoti(X)\mathrm{H}_{\textnormal{mot}}(X)=\bigoplus_{i}\mathrm{H}_{\textnormal{mot}}^{i}(X). This gives contravariant functors Hmot(_)\mathrm{H}_{\textnormal{mot}}(\_) and Hmoti(_)\mathrm{H}_{\textnormal{mot}}^{i}(\_) from the category of smooth projective varieties over KK to MotK\textnormal{Mot}_{K}.

Theorem 5.4.

The category MotK\textnormal{Mot}_{K} is Tannakian over {\mathbb{Q}}, semisimple, graded, and polarised. Every classical cohomology theory of smooth projective varieties over KK factors via MotK\textnormal{Mot}_{K}.

Proof.

See théorème 0.4 of [And96]. ∎

Definition 5.5.

Let KK be a subfield of {\mathbb{C}}. An abelian motive over KK is an object of the Tannakian subcategory of MotK\textnormal{Mot}_{K} generated by objects of the form Hmot(X)\mathrm{H}_{\textnormal{mot}}(X) where XX is either an Abelian variety or X=Spec(L)X=\textnormal{Spec}(L) for some finite extension L/KL/K, with LL\subset{\mathbb{C}}.

We denote the category of abelian motives over KK with AbMotK\textnormal{AbMot}_{K}.

Finally we need the following theorem

Theorem 5.6.

The Hodge realization functor H()H(-) restricted to the subcategory of abelian motives is a full functor.

Proof.

See théorème 0.6.2 of [And96]. ∎

By Theorem 5.4, the singular cohomology and \ell-adic cohomology functors factor via MotK\textnormal{Mot}_{K}. This means that if MM is a motive, then we can attach to it a Hodge structure H(M)H(M) and an \ell-adic Galois representation H(M)H_{\ell}(M). The Artin comparison isomorphism between singular cohomology and a \ell-adic cohomology extends to an isomorphism of vector spaces H(M)H(M)H_{\ell}(M)\cong H(M)\otimes{\mathbb{Q}}_{\ell} that is natural in the motive MM.

We can state the motivated Mumford–Tate conjecture following [Moo17b, Section 3.2]. We shall write GMT(M)G_{\textnormal{MT}}(M) for the Mumford–Tate group GMT(H(M))G_{\textnormal{MT}}(H(M)). Similarly, we write G(M)G_{\ell}(M) (resp. G(M)G_{\ell}^{\circ}(M)) for G(H(M))G_{\ell}(H_{\ell}(M)) (resp. (G(H(M)(G_{\ell}^{\circ}(H_{\ell}(M)) for the Tate group. The Mumford–Tate conjecture extends to motives: for the motive MM it asserts that the comparison isomorphism H(M)H(M)H_{\ell}(M)\cong H(M)\otimes{\mathbb{Q}}_{\ell} induces an isomorphism

G(M)GMT(M).G_{\ell}^{\circ}(M)\cong G_{\textnormal{MT}}(M)\otimes{\mathbb{Q}}_{\ell}.

The discussion we have given here enable us to prove the following.

Proposition 5.7.

Let SS be a surface of general type as above if there exists a subgroup HH of the automorphisms group GG of SS, and there exist a K3 surface XX birational to S/HS/H and ι\iota as in (H) is motivated (in the sense of André) then the Tate and Mumford–Tate conjectures hold for SS. That is the Tate and Mumford–Tate conjectures hold for those SS that are deformation equivalent to such a surface SS.

Proof.

The proof of this theorem is contained in[CoPe20, Section 5]. We illustrate only the demonstration strategy in the realm of motives.

The main idea in [CoPe20] is that for surfaces SS with pg=2p_{g}=2 it is sometimes possible to decompose the weight 22 Hodge structure into two Hodge substructures of K3 type and see that these Hodge substructures are indeed the Hodge structures of either Abelian surfaces or K3 surfaces which are (birational) quotients of SS. This geometric construction makes possible to consider the theory of motivated cycles introduced by André, and to decompose the motive of SS into two abelian motives of K3 type. For these motives the Mumford–Tate conjecture is known. This, together with the main results of [Com16] and [Com19] allows to prove the Mumford-Tate conjecture for SS. In [CoPe20] it was used the fact that the surfaces studied are of maximal Albanese dimension hence there is naturally an Abelian surface as a quotient surface.

Let us summarize here, in the form of a table, the classification of the minimal surfaces of general type with pg=q=2p_{g}=q=2 and of maximal Albanese dimension. Moreover, for each family we will indicate whether the Mumford–Tate conjecture has been proved or not.

KS2K^{2}_{S} deg(α)\deg(\alpha) #\# dim\dim Name mtc pq/ms Reference
1 88 22 22 020^{2} No [PRR20]
2 88 {2,4,6}\{2,4,6\} 44 33,43^{3},4 SIP Yes [Pen11]
3 77 33 11 33 PP7 Yes [PiPo17, CaFr18]
4 77 22 3 33 ? [Rit18, PePi20]
5 66 44 11 44 PP4 Yes [PePo14]
6 66 33 ?? 33 AC3 ? [AlCa22, CaSe22]
7 66 22 11 33 PP2 ? [PePo13b]
8 66 22 22 424^{2} PP2 Yes [PePo13b] [Pig20]
9 55 33 11 44 CHPP Yes [ChHa06, PePo13a]
10 44 22 11 44 Catanese Yes [Pen11, CiML02]
Table 1: State of the art of the classification of minimal complex algebraic surfaces with invariants pg=q=2p_{g}=q=2 and with maximal Albanese dimension.

In Table 1 have indicated, where possible, the number of families (#\#) and the dimensions of the irreducible component containing them (dim\dim). Moreover, we point out if some members of the family are product-quotient surfaces (pq) or mixed surfaces (ms), in particular for some PP2 surfaces the proof is given by the second author in [Pig20], while for the remaining ones is unknown. In the last column, we give references to more detailed descriptions of the class. Finally, we put a checkmark in the column mtc if the strategy given in [CoPe20] is enough to prove the Tate and Mumford–Tate conjectures for a class. As one can see, up to now, it is the only way to prove Mumford–Tate conjecture for these surfaces.

Finally, we explain here the meaning of the half-line between the surfaces with KS27K^{2}_{S}\geq 7 and KS26K^{2}_{S}\leq 6. This line is due to the classification theorem of [DJZ23], who recently proved that the classification of surfaces with pg=q=2p_{g}=q=2 and KS26K^{2}_{S}\leq 6 is complete.

References

  • [AlCa22] M. Alessandro, F. Catanese On the components of the Main Stream of the moduli space of surfaces of general type with pg=q=2p_{g}=q=2, preprint available at arXiv:2212.14872
  • [And96] Y. André Pour une théorie inconditionnelle des motifs. In: Institut des Hautes Études Scientifiques. Publications Mathématiques 83 (1996), pp. 5 – 49.
  • [Bar87] W. Barth: Abelian surfaces with (1,2)(1,2)-Polarization. Algebraic Geometry, Sendai, 1985, 41–84, Adv. Stud. Pure Math. 10, North-Holland, Amsterdam, 1987.
  • [BHPV03] W. Barth, K. Hulek, C.A.M. Peters, A. Van de Ven, Compact Complex Surfaces. Grundlehren der Mathematischen Wissenschaften, Vol 4, Second enlarged edition, Springer-Verlag, Berlin, 2003.
  • [CaFr18] N. Cancian, D. Frapporti, On semi-isogenous mixed surfaces, Math. Nachr. 291, 264–283 (2018).
  • [CaSe22] F. Catanese, E. Sernesi, The Hesse pencil and polarizations of type (1,3) on Abelian surfaces, preprint available at arXiv:2212.14877.
  • [ChHa06] J. A. Chen, C. D. Hacon, A surface of general type with pg=q=2p_{g}=q=2 and KX2=5K_{X}^{2}=5, Pacif. J. Math. 223 (2006), 219–228.
  • [CiML02] C.Ciliberto, M. Mendes Lopes, On surfaces with pg=q=2p_{g}=q=2 and non-birational bicanonical maps, Advances in Geometry 2.3 (2002), 281–300.
  • [Com16] J. Commelin, The Mumford–Tate conjecture for the product of an abelian surface and a K3 surface, Documenta Mathematica 21 (2016), 1691–1713.
  • [Com19] J. Commelin, The Mumford-Tate conjecture for products of abelian varieties, Algebr. Geom. 6 (2019), 650–677.
  • [CoPe20] J. Commelin, M. Penegini, On the cohomology of surfaces with pg=q=2p_{g}=q=2 and maximal Albanese dimension, Transactions of the American Mathematical Society, 2020, 373, 1749–1773.
  • [DeMu91] C. Deninger and J. P. Murre, Motivic decomposition of abelian schemes and the Fourier transform, in: J. Reine und Angew. Math. 422 (1991), 201–219.
  • [DJZ23] J. Du, Z. Jiang, G. Zhang, Cohomological rank functions and surfaces of general type with pg=q=2, preprint available at ArXiv:2309.05097
  • [Lat19] R. Laterveer, Algebraic cycles and triple K3 burgers Ark. Mat. 57(1) (2019), 157–189.
  • [Man68] J. I. Manin, Correspondences, motifs and monoidal transformations. In: Mat. Sb. (N.S.) 77 (1968), 475–507.
  • [Moo17a] B. Moonen, On the Tate and Mumford–Tate conjectures in codimension 1 for varieties with h2,0=1h^{2,0}=1. Duke Mathematical Journal 166 (2017), 739–799.
  • [Moo17b] B. Moonen, Families of Motives and the Mumford–Tate Conjecture, Milan J. Math. 85, 257–307 (2017).
  • [Mor87] D. R. Morrison. Isogenies between algebraic surfaces with geometric genus one. In: Tokyo Journal of Mathematics 10.1 (1987), pp. 179–187.
  • [Mur90] J. P. Murre, On the motive of an algebraic surface, in: J. Reine und Angew. Math. 409 (1990), 190–204.
  • [Pen11] M. Penegini, The classification of isotrivially fibred surfaces with pg=q=2p_{g}=q=2. Collect. Math. 62, 239–274 (2011).
  • [PePi20] M. Penegini, R. Pignatelli, A note on a family of surfaces with pg=q=2p_{g}=q=2 and K2=7K^{2}=7, Boll. Unione Mat. Ital. 15, 305–331 (2022).
  • [PePo13a] M. Penegini, F. Polizzi, On surfaces with pg=q=2,K2=5p_{g}=q=2,\ K^{2}=5 and Albanese map of degree 33, Osaka J. Math. 50, 643–686 (2013).
  • [PePo13b] M. Penegini, F. Polizzi, On surfaces with pg=q=2,K2=6p_{g}=q=2,K^{2}=6 and Albanese map of degree 22, Canad. J. Math. 65 (2013), 195–221
  • [PePo14] M. Penegini, F. Polizzi, A new family of surfaces with pg=q=2p_{g}=q=2 and K2=6K^{2}=6 whose Albanese map has degree 44, J. London Math. Soc. 90, 741–762 (2014).
  • [Pig20] R. Pignatelli, Quotients of the square of a curve by a mixed action, further quotients and Albanese morphisms. Revista Matematica Complutense 33 (2020), 911–931.
  • [PiPo17] R. Pignatelli, F. Polizzi A family of surfaces with pg=q=2p_{g}=q=2, K2=7K^{2}=7 and Albanese map of degree 33 Mathematische Nachrichten, 2017, 290, pp. 2684–2695
  • [PRR20] F. Polizzi, C. Rito, X. Roulleau, A pair of rigid surfaces with pg=q=2p_{g}=q=2 and K2=8K^{2}=8 whose universal cover is not the bidisk, Int. Math. Res. Not. IMRN 2020, 11, 3453–3493 (2020).
  • [Rit18] C. Rito New surfaces with K2=7K^{2}=7 and pg=q2p_{g}=q\leq 2, Asian J. Math. 22 (2018), no. 6, 1117–1126.
  • [Sch91] A. J. Scholl Classical motives. In: Motives (Seattle, WA, 1991). Vol. 55. Proc. Sympos. Pure Math. Amer. Math. Soc., Providence, RI, 1994, pp. 163–187.
  • [Sil08] J.H. Silverman, The Arithmetic of Elliptic Curves, 2nd Edition, Springer GTM vol 106, (2008).

Matteo Penegini, Università degli Studi di Genova, DIMA Dipartimento di Matematica, I-16146 Genova, Italy
e-mail [email protected]

Roberto Pignatelli Università degli Studi di Trento, Dipartimento di Matematica, I-38123 Trento, Italy
e-mail [email protected]