Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-27T13:25:33.850Z Has data issue: false hasContentIssue false

Sixfolds of generalized Kummer type and K3 surfaces

Published online by Cambridge University Press:  05 January 2024

Salvatore Floccari*
Affiliation:
Institute of Algebraic Geometry, Leibniz University Hannover, Hannover 30167, Germany floccari@math.uni-hannover.de
Rights & Permissions [Opens in a new window]

Abstract

We prove that any hyper-Kähler sixfold $K$ of generalized Kummer type has a naturally associated manifold $Y_K$ of $\mathrm {K}3^{[3]}$ type. It is obtained as crepant resolution of the quotient of $K$ by a group of symplectic involutions acting trivially on its second cohomology. When $K$ is projective, the variety $Y_K$ is birational to a moduli space of stable sheaves on a uniquely determined projective $\mathrm {K}3$ surface $S_K$. As an application of this construction we show that the Kuga–Satake correspondence is algebraic for the K3 surfaces $S_K$, producing infinitely many new families of $\mathrm {K}3$ surfaces of general Picard rank $16$ satisfying the Kuga–Satake Hodge conjecture.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited. Compositio Mathematica is © Foundation Compositio Mathematica.
Copyright
© 2024 The Author(s)

1. Introduction

Together with manifolds of $\mathrm {K}3^{[n]}$ type, deformations of generalized Kummer varieties constitute the most well-studied hyper-Kähler manifolds. We refer to this deformation type in dimension $2n$ as to the $\mathrm {Kum}^n$ type. After Beauville [Reference BeauvilleBea83] gave the first examples of such hyper-Kähler manifolds, many more have been constructed from moduli spaces of stable sheaves on abelian surfaces, see [Reference YoshiokaYos01]. However, the varieties so obtained always have Picard rank at least $2$, and our understanding of a general projective variety of $\mathrm {Kum}^n$ type remains poor. In fact, for the time being, no construction of such variety is known (but see [Reference O'GradyO'Gr22] for some recent ideas).

In the present article we partially remedy this for hyper-Kähler sixfolds of generalized Kummer type, by associating to any $K$ of $\mathrm {Kum}^3$ type a hyper-Kähler manifold $Y_K$ of $\mathrm {K}3^{[3]}$ type. They are related by a dominant rational map

\[ K\dashrightarrow Y_K \]

of degree $2^5$, described as follows.

Any $K$ of $\mathrm {Kum}^3$ type admits an action of the group $( {\mathbb {Z}}/4 {\mathbb {Z}})^4\rtimes {\mathbb {Z}}/2 {\mathbb {Z}}$ by symplectic automorphisms, where $ {\mathbb {Z}}/2 {\mathbb {Z}}$ acts on the first factor as $-1$. In fact, this is the group $\operatorname {Aut}_0(K)$ of automorphisms of $K$ which act trivially on its second cohomology, which is deformation invariant by [Reference Hassett and TschinkelHT13]. It has been computed in [Reference Boissière, Nieper-Wißkirchen and SartiBNS11] for the generalized Kummer variety associated to an abelian surface. We let $G\subset \operatorname {Aut}_0(K)$ be the subgroup generated by the automorphisms whose fixed locus contains a $4$-dimensional component. We will show in Lemma 2.4 that

\[ G\cong ({\mathbb{Z}}/2{\mathbb{Z}})^5. \]

Theorem 1.1 Let $K$ be a manifold of $\mathrm {Kum}^3$ type. The quotient $K/G$ admits a resolution $Y_K\to K/G$ with $Y_K$ a manifold of $\mathrm {K}3^{[3]}$ type.

The resolution $Y_K\to K/G$ is obtained via a single blow-up of the reduced singular locus. It can also be described as the quotient by $G$ of the blow-up of $K$ at the union of the fixed loci of all non-trivial automorphisms in $G$. We show that there is an isometry of transcendental Hodge structures

\[ H_{\mathrm{tr}}^2(Y_K, {\mathbb{Q}}) \xrightarrow{\ \sim\ } H_{\mathrm{tr}}^2(K, {\mathbb{Q}})(2), \]

where on the right-hand side the form is multiplied by $2$. This parallels the classical construction of the Kummer $\mathrm {K}3$ surface $\operatorname {\mathrm {Km}}(A)$ associated to an abelian surface $A$, where $H^2_{\mathrm {tr}}(\operatorname {\mathrm {Km}}(A), {\mathbb {Z}})$ is Hodge isometric to $H^2_{\mathrm {tr}}(A, {\mathbb {Z}})(2)$. The study of the integral transcendental lattices of the manifolds $Y_K$ will be the subject of future work.

As a consequence, we obtain a well-defined $\mathrm {K}3$ surface associated to any projective sixfold of generalized Kummer type.

Theorem 1.2 Let $K$ be a projective variety of $\mathrm {Kum}^3$ type. There exists a unique (up to isomorphism) projective $\mathrm {K}3$ surface $S_K$ such that the variety $Y_K$ of $\mathrm {K}3^{[3]}$ type given by Theorem 1.1 is birational to a moduli space $M_{S_K, H}(v)$ of stable sheaves on $S_K$, for some primitive Mukai vector $v$ and a $v$-generic polarization $H$.

We call $S_K$ the $\mathrm {K}3$ surface associated to the sixfold $K$. It is characterized by the existence of a Hodge isometry $H^2_{\mathrm {tr}}(S_K, {\mathbb {Z}})\xrightarrow {\ \sim \ } H^2_{\mathrm {tr}}(Y_K, {\mathbb {Z}})$. These surfaces come in countably many $4$-dimensional families of general Picard rank $16$; up to isogeny, they are the K3 surfaces $S$ admitting an isometric embedding $H^2_{\mathrm {tr}}(S, {\mathbb {Q}})\hookrightarrow \Lambda _{\mathrm {Kum}^3}(2)\otimes _{ {\mathbb {Z}}} {\mathbb {Q}}$ of rational quadratic spaces. Here, $\Lambda _{\mathrm {Kum}^3}$ is the lattice which is the second cohomology of manifolds of $\mathrm {Kum}^3$ type; it was computed in [Reference BeauvilleBea83] that $\Lambda _{\mathrm {Kum}^3}=U^{\oplus 3}\oplus \langle -8 \rangle$, where we denote by $U$ the hyperbolic plane.

Applications

In a series of papers [Reference MarkmanMar20, Reference MarkmanMar22, Reference MarkmanMar23], Markman has proven striking results on the Hodge conjecture for hyper-Kähler varieties of $\mathrm {Kum}^n$ and $\mathrm {K}3^{[n]}$ type. He uses Verbitsky's theory of hyperholomorphic sheaves [Reference VerbitskyVer96] to produce very interesting algebraic cycles on general projective hyper-Kähler varieties via deformation. Through our construction, we are able to deduce more cases of the Hodge conjecture from his results.

Our main application is to the Kuga–Satake Hodge conjecture for $\mathrm {K}3$ surfaces [Reference van GeemenvGe00]. The Kuga–Satake construction [Reference DeligneDel71] associates via Hodge theory an abelian variety $\mathrm {KS}(S)$ to any projective K3 surface $S$, and the conjecture predicts the existence of an algebraic cycle inducing an embedding of the transcendental Hodge structure of the surface into the second cohomology of $\mathrm {KS}(S)\times \mathrm {KS}(S)$. It is known to hold in many cases for $\mathrm {K}3$ surfaces of Picard number at least $17$ (see [Reference MorrisonMor85]), but it is wide open otherwise. There are a couple of families of $\mathrm {K}3$ surfaces of general Picard rank $16$ for which the conjecture is known, namely, the family of double covers of the plane branched at six lines [Reference ParanjapePar88] and the family of K3 surfaces with $15$ nodes in $ {\mathbb {P}}^4$ (see [Reference Ingalls, Logan and PatashnickILP22]).

Theorem 1.3 gives infinitely many new families of K3 surfaces of general Picard rank $16$ for which the Kuga–Satake Hodge conjecture holds true.

Theorem 1.3 Let $S$ be a projective $\mathrm {K}3$ surface such that there exists an isometric embedding of $H^2_{\mathrm {tr}}(S, {\mathbb {Q}})$ into $\Lambda _{\mathrm {Kum}^3}(2)\otimes _{ {\mathbb {Z}}} {\mathbb {Q}}$, where $(2)$ indicates that the form is multiplied by $2$. Then, there exists an algebraic cycle $\gamma$ on $S\times \mathrm {KS}(S)\times \mathrm {KS}(S)$ inducing an embedding $\gamma _*\colon H^2_{\mathrm {tr}}(S, {\mathbb {Q}})\hookrightarrow H^2(\mathrm {KS}(S)\times \mathrm {KS}(S), {\mathbb {Q}})$ of rational Hodge structures.

Via Theorem 1.2, we deduce this statement from the validity of the Kuga–Satake Hodge conjecture for varieties of $\mathrm {Kum}^n$ type, established by Voisin [Reference VoisinVoi22] as a consequence of results of O'Grady [Reference O'GradyO'Gr21] and Markman [Reference MarkmanMar23]. Once Theorem 1.3 is proven, a result of Varesco [Reference VarescoVar22] implies that the Hodge conjecture holds for all powers of any K3 surface satisfying the assumption of the above theorem, see Corollary 5.8.

In the recent preprint [Reference MarkmanMar22], Markman proves that any rational Hodge isometry $H^2(X, {\mathbb {Q}})\xrightarrow { \ \sim \ } H^2(X', {\mathbb {Q}})$ between varieties of $\mathrm {K}3^{[n]}$ type is algebraic. He expects that an extension of his argument will lead to the analogous result for varieties of $\mathrm {Kum}^n$ type. In dimension $6$, we can obtain it from the $\mathrm {K}3^{[3]}$ case via Theorem 1.1.

Theorem 1.4 Let $K, K'$ be projective varieties of deformation type $\mathrm {Kum}^3$. Let

\[ f\colon H^2(K, {\mathbb {Q}})\xrightarrow { \ \sim \ } H^2(K', {\mathbb {Q}}) \]

be a Hodge isometry. Then $f$ is induced by an algebraic correspondence.

Overview of the contents

We sketch the proof of Theorem 1.1. First, we calculate the fixed loci $K^g$ of all the automorphisms $g\in G$: we show that $\bigcup _{g\neq 1 \in G} K^g$ is the union of $16$ hyper-Kähler manifolds of $\mathrm {K}3^{[2]}$ type, and we determine the various intersections of these components. Thanks to the deformation invariance of $G$, it is in fact enough to calculate these loci in the special case of the generalized Kummer sixfold on an abelian surface.

We then describe explicitly the singularities of $K/G$. It turns out that these are of a particular simple nature, modeled on products of ordinary double points on surfaces. A resolution $Y_K$ of $K/G$ is obtained via a single blow-up of the singular locus. The quotient $K/G$ is a primitive symplectic orbifold as studied by Fujiki [Reference FujikiFuj83] and Menet [Reference MenetMen20], and we use a criterion due to Fujiki to prove that $Y_K$ is a hyper-Kähler manifold. Moreover, the resolution can be performed in families, so that the $Y_K$ are deformation equivalent to each other. To complete the proof of Theorem 1.1 it is therefore sufficient to find a single $K$ of $\mathrm {Kum}^3$ type such that $Y_K$ is of $\mathrm {K}3^{[3]}$ type.

The specific example we study is a Beauville–Mukai system $K_J(v_3)$ on a general principally polarized abelian surface $\Theta \subset J$. It admits a Lagrangian fibration to the linear system $|2\Theta |= {\mathbb {P}}^3$, whose general fibres parametrize certain degree-$3$ line bundles supported on curves in the linear system. We show that the norm map for line bundles induces a dominant rational map of degree $2^5$ from $K_J(v_3)$ onto a Beauville–Mukai system $M_{\operatorname {\mathrm {Km}}(J)}(w_3)$, a moduli space of sheaves on the Kummer $\mathrm {K}3$ surface $\operatorname {\mathrm {Km}}(J)$ associated to $J$; the hyper-Kähler variety $M_{\operatorname {\mathrm {Km}}(J)}(w_3)$ is birational to ${\operatorname {\mathrm {Km}}(J)}^{[3]}$. We next prove that the norm map descends to a birational map

\[ K_J(v_3)/G \dashrightarrow M_{\operatorname{\mathrm{Km}}(J)}(w_3). \]

It follows that the hyper-Kähler manifold $Y_{K_J(v_3)}$ is birational to $\mathrm {Km}(J)^{[3]}$, and hence $Y_{K_J(v_3)}$ is of $\mathrm {K}3^{[3]}$ type by Huybrechts’ theorem [Reference HuybrechtsHuy99] that birational hyper-Kähler manifolds are deformation equivalent.

Once Theorem 1.1 is proven, the other results follow from it. When $K$ is projective, we define the associated $\mathrm {K}3$ surface $S_K$ as the unique (up to isomorphism) $\mathrm {K}3$ surface whose transcendental lattice is Hodge isometric to $H^2_{\mathrm {tr}}(Y_K, {\mathbb {Z}})$. This is justified by the fact that $H^2_{\mathrm {tr}}(Y_K, {\mathbb {Z}})$ has rank at most $6$ and, hence, it appears as transcendental lattice of some $\mathrm {K}3$ surface of Picard rank at least $16$. Moreover, for such Picard numbers, two $\mathrm {K}3$ surfaces with Hodge isometric transcendental lattices are isomorphic; therefore, $S_K$ is determined up to isomorphism. To prove our applications, we exploit the algebraic cycle in $K\times Y_K$ given by the rational map $K\dashrightarrow Y_K$.

The organization of this text is as follows. In § 2 we calculate the fixed loci of the automorphisms of $G$. In § 3 we study in detail the case of $K_J(v_3)$ as outlined above. In § 4 we prove Theorems 1.1 and 1.2. Finally, in § 5 we use our construction to prove Theorems 1.3 and 1.4.

2. Automorphisms trivial on the second cohomology

Let $K$ be a manifold of $\mathrm {Kum}^3$ type and let $\operatorname {Aut}_0(K)$ be the group of automorphisms of $K$ which act trivially on $H^2(K, {\mathbb {Z}})$. We let $K^h$ denote the fixed locus of an automorphism $h$ of $K$.

Definition 2.1 We denote by $G$ the subgroup of $\operatorname {Aut}_0(K)$ generated by the automorphisms $g\in \operatorname {Aut}_0(K)$ such that $K^g$ has a $4$-dimensional component.

We will see in Lemma 2.4 that $G\cong ( {\mathbb {Z}}/2 {\mathbb {Z}})^5$. The main result proven in this section is then the following.

Theorem 2.2 Let $Z\subset K$ be defined as $Z:= \bigcup _{g\neq 1\in G} K^{g}$. Then $Z$ is the union of $16$ fourfolds $Z_j$, each of which is a smooth hyper-Kähler manifold of $\mathrm {K}3^{[2]}$ type. Moreover, two distinct components intersect in a K3 surface, three distinct components intersect in four points, and four or more distinct components do not intersect.

In order to prove this, we will compute the fixed loci of all automorphisms in $G$. Thanks to the deformation invariance of automorphisms trivial on the second cohomology it will be sufficient to treat the case of the generalized Kummer variety associated to an abelian surface $A$, i.e. $K$ is the fibre over $0$ of the composition

\[ A^{[4]} \xrightarrow{ \ \nu\ } A^{(4)} \xrightarrow{ \ \sum\ } A \]

of the Hilbert–Chow morphism with the summation map. We denote by $A^{(4)}_0\subset A^{(4)}$ the fibre over $0$ of $\sum$. The restriction $\nu \colon K\to A^{(4)}_0$ is a crepant resolution.

In this case the group of automorphisms of $K$ acting trivially on the second cohomology has the natural description $\operatorname {Aut}_0(K) = A_4 \rtimes \langle -1 \rangle,$ and the action on $K$ and $A^{[4]}$ is induced by that on $A$, see [Reference Boissière, Nieper-Wißkirchen and SartiBNS11].

2.3 Notation

We let $A_n$ be the group of points of order $n$ of $A$. Via the isomorphism above, we write the elements of $\operatorname {Aut}_0(K^3(A))$ as $(\epsilon, \pm 1)$ for $\epsilon \in A_4$. For $\tau \in A_2$, we denote by $A_{2,\tau }$ the set $\{a\in A \ | \ 2a=\tau \}$, which consists of $16$ points. The quotient surface $A/\langle (\tau, -1) \rangle$ has $16$ nodes corresponding to points in $A_{2,\tau }$; its minimal resolution $\operatorname {\mathrm {Km}}^{\tau }(A)$ is isomorphic to the Kummer $\mathrm {K}3$ surface associated to $A$.

Lemma 2.4 Let $K$ be any manifold of $\mathrm {Kum}^3$ type. Then $G\cong ( {\mathbb {Z}}/2 {\mathbb {Z}})^5$.

Proof. Since automorphisms in $\operatorname {Aut}_0(K)$ deform with $K$, their fixed loci deform as well. Therefore, it suffices to prove the lemma for the generalized Kummer variety $K=K^3(A)$ on an abelian surface $A$. In this case we show that

\[ G=A_2\times \langle -1\rangle \cong ({\mathbb{Z}}/2{\mathbb{Z}})^5. \]

We have $\operatorname {Aut}_0(K^3(A))= A_4\rtimes \langle -1\rangle$. If $h\in \operatorname {Aut}_0(K^3(A))$ is induced by translation by $0\neq \epsilon \in A_4$, then, by [Reference OguisoOgu20, Lemma 3.5], the fixed locus $(K^3(A))^h$ is a union of K3 surfaces if $2\epsilon =0$ and consists of isolated point if $2\epsilon \neq 0$. Consider now $h=(\epsilon, -1)$. The fixed locus of $h=(\epsilon, -1)$ on $A^{(4)}$ consists of

\begin{align*} W_{\epsilon} = \{ (a, -a -\epsilon , b , - b - \epsilon), \ \text{for } a,b \in A \} \subset A^{(4)}, \end{align*}

and subvarieties of lower dimension. If $\epsilon \in A_2$, then $W_{\epsilon }$ is contained in $A^{(4)}_0$ and, hence, ${(\epsilon, -1)\in G}$. If $\epsilon \notin A_2$, then $W_{\epsilon }$ has empty intersection with $A^{(4)}_0$; the action of $h$ on $A^{(4)}_0$ fixes isolated points and the surfaces $\{ (\alpha, b, -b-\epsilon, - \alpha + \epsilon ),\ b\in A\}$, where $\alpha \in A$ satisfies $\alpha =-\alpha -\epsilon$. The general point of such a surface is supported on four distinct points. Therefore, the fixed locus $(K^3(A))^h$ consists of surfaces and isolated points, because it is a union of symplectic varieties and the fibres of $\nu$ have dimension at most $3$. We conclude that $G$ is generated by the $16$ involutions $(\tau,-1)$ with $\tau \in A_2$ and, hence, $G=A_2\times \langle -1\rangle$.

2.5 Relevant subvarieties

In what follows we let $K=K^3(A)$ be the generalized Kummer variety on an abelian surface $A$. We will calculate the fixed loci of automorphisms in $G=A_2\times \langle -1\rangle$ acting on $K$. The next definition introduces the relevant subvarieties. We denote by $\xi =(a,b,c,d)$ a point of $A_0^{(4)}$, and refer to $\{a,b,c,d\}$ as the support of $\xi$.

Definition 2.6 We define the following subvarieties of $A_0^{(4)}$:

  • for any $\tau \in A_{2}$, we let

    \[ W_{\tau}:= \{(a, b, -a+\tau, -b+\tau), \ \text{for } \ a,b\in A \}; \]
  • for a pair $(\tau, \theta )\in A_2 \times A_2$ with $\tau \neq 0$, we let

    \[ V_{\tau, \theta} := \{ (a, a+\tau, -a+\theta , -a+\tau+\theta), \ \text{for } \ a\in A\}. \]

We denote by $\overline {W_{\tau }}$ and $\overline {V_{\tau, \theta }}$ the subvarieties of $K$ obtained as strict transform of $W_{\tau }$ and $V_{\tau, \theta }$ under the Hilbert–Chow morphism, respectively.

Note that none among the $W_{\tau }$ and $V_{\tau, \theta }$ is contained in the exceptional locus of $\nu \colon K\to A^{(4)}_0$, so that $\overline {W_{\tau }}$ and $\overline {V_{\tau,\theta }}$ are well-defined. We will show in Lemma 2.9 that the $\overline {W_{\tau }}$ are hyper-Kähler varieties of $\mathrm {K}3^{[2]}$ type and that the $\overline {V_{\tau,\theta }}$ are K3 surfaces.

Remark 2.7 (i) For $\tau \in A_2$, the $G$-equivariant morphism $\psi _{\tau }\colon A\times A\to A^{(4)}$ such that $(a,b)\mapsto (a, b, -a+\tau, -b+\tau )$ has degree $8$ and induces an isomorphism

\[ (A/ \langle (\tau, -1) \rangle)^{(2)} \xrightarrow{\ \sim\ } W_{\tau}. \]

Moreover, $W_{\tau }\neq W_{\tau '}$ unless $\tau =\tau '$, as can be seen considering a point $(\epsilon, \epsilon, \epsilon, \epsilon )$ for some $\epsilon \in A_{2,\tau }$.

(ii) Given $0\neq \tau \in A_2$ and $\theta \in A_2$, the $G$-equivariant morphism $\psi _{\tau, \theta }\colon A\to A^{(4)}$ defined by $a \mapsto (a, a+\tau, -a+\theta, -a+\tau +\theta )$ induces an isomorphism

\[ A/\langle \tau, (\theta, -1) \rangle \xrightarrow{\ \sim\ } V_{\tau, \theta}. \]

For a given $\tau \in A_2$, the union $\bigcup _{\theta \in A_2} V_{\tau, \theta }$ has eight irreducible components. In fact, $V_{\tau, \theta }=V_{\tau, \theta '}$ if and only if $\theta ' \in \{\theta, \theta +\tau \}$, which can be seen considering a point $(a, a+\tau, a, a+\tau )$: it belongs to $V_{\tau, \theta }$ if and only if $a\in A_{2, \theta }$ or $a\in A_{2,\theta +\tau }$. Moreover, it can be easily checked that $V_{\tau, \theta }\neq V_{\tau ', \theta '}$ for $\tau \neq \tau '$. In total, we therefore have $120$ distinct surfaces $V_{\tau, \theta }$.

We recall a result of Kamenova, Mongardi and Oblomonkov which will be used in the sequel.

Proposition 2.8 [Reference Kamenova, Mongardi and OblomkovKMO22, Lemma B.1]

Let $\iota \colon {\mathbb {C}}^2\to {\mathbb {C}}^2$ be the involution given by $(x,y)\mapsto (-x, -y)$. Consider the rational map

\[ \Phi\colon (\mathrm{Bl}_{0}({\mathbb{C}}^2/\iota))^{[n]} \dashrightarrow ({\mathbb{C}}^2)^{[2n]} \]

which maps a general $\zeta ={(P_1, \ldots, P_{n})}$ to $\xi ={(P_1, \iota (P_1), \ldots, P_n, \iota (P_n))}$. Then $\Phi$ extends to a closed embedding, with image the fixed locus for the action of $\iota$ on $( {\mathbb {C}}^2)^{[2n]}$.

The next lemma describes the varieties $\overline {V_{\tau, \theta }}$ and $\overline {W_{\tau }}$ introduced above.

Lemma 2.9

  1. (i) For $\tau \neq 0$, $\theta \in A_2$, the rational map $\phi _{\tau,\theta }$ defined by the commutative diagram

    where $q$ is the natural map, extends to a regular morphism, which is an isomorphism onto its image $\overline {V_{\tau, \theta }}$.
  2. (ii) For $\tau \in A_2$, the rational map $\phi _{\tau }$ defined by the commutative diagram

    where $q$ is the composition of the Hilbert–Chow morphism with the natural map $(\operatorname {\mathrm {Km}}^{\tau }(A))^{(2)} \to (A/\langle (\tau, -1)\rangle )^{(2)}$, extends to a regular morphism, which is an isomorphism onto its image $\overline {W_{\tau }}$.

Proof. We prove part (i). As translations are fixed point free, for any $\alpha \in A/\langle \tau, (\theta, -1)\rangle$ the support of $\psi _{\tau, \theta } (\alpha )\in V_{\tau, \theta }$ consists of either two or four distinct points, according to whether $\alpha$ is a node or not. If $\alpha$ is not a node $q^{-1}(\alpha )$ is a single point at which $\phi _{\tau, \theta }$ is well-defined. Otherwise, $\psi _{\tau, \theta } (\alpha ) =(a, a+\tau, a, a+\tau )$ for some $a$ in $A_{2, \theta }$ or $A_{2,\tau +\theta }$. Then, there is a canonical identification $\nu ^{-1}(\psi _{\tau, \theta } (\alpha ))= {\mathbb {P}}(T_aA) \times {\mathbb {P}}(T_{a+\tau } A)$. The exceptional divisor in $\operatorname {\mathrm {Km}}^{\theta }(A/\langle \tau \rangle )$ corresponding to the node $\alpha$ is identified with $ {\mathbb {P}}(T_a A)$. Translation by $\tau$ gives an isomorphism $ {\mathbb {P}}(T_aA)\xrightarrow {\ \sim \ } {\mathbb {P}}(T_{a+\tau}A)$, and the morphism $\phi _{\tau, \theta }$ is extended via $\phi _{\tau, \theta } (t):= (t, \tau (t))$ for $t\in {\mathbb {P}}(T_aA)$.

The proof of part (ii) is similar. If $\psi _{\tau } (\alpha, \beta )$ consists of four distinct points, then $q^{-1}(\alpha, \beta )$ is a single point at which $\phi _{\tau }$ is well-defined. If the support of $\psi _{\tau }(\alpha, \beta )$ consists of three distinct points, then exactly one between $\alpha$ and $\beta$ is a node of $A/\langle (\tau, -1) \rangle$, and $\psi _{\tau }(\alpha, \beta )=(a, \epsilon, -a+\tau, \epsilon )$ for some $a\in A\setminus A_{2,\tau }$ and $\epsilon \in A_{2,\tau }$; we extend $\phi _{\tau }$ via the canonical identifications $q^{-1}(\alpha, \beta ) = {\mathbb {P}}(T_{\epsilon }A)=\nu ^{-1}(\psi _{\tau }(\alpha,\beta ))$. If the support of $\psi _{\tau }(\alpha, \beta )$ consists of two distinct points then either $\alpha =\beta$ for some smooth point or $\alpha \neq \beta$ with both $\alpha$ and $\beta$ nodes of $A/\langle (\tau,-1)\rangle$. In the first case there exists $a\in A\setminus A_{2,\tau }$ such that $\psi _{\tau }(\alpha,\beta )=(a, a, -a+\tau, -a+\tau )$. Then $q^{-1}(\alpha, \beta ) = \mathbb {P}(T_a A)$ and $\nu ^{-1}(\psi _{\tau } (\alpha, \beta ))= {\mathbb {P}}(T_{a}A)\times {\mathbb {P}}(T_{-a+\tau } A)$, and we define $\phi _{\tau }$ by $t\mapsto (t, (\tau, -1)(t))$. In the second case $\psi _{\tau }(\alpha, \beta )= (\epsilon _1, \epsilon _2, \epsilon _1, \epsilon _2)$ for some $\epsilon _1\neq \epsilon _2$ both in $A_{2,\tau }$, and we extend $\phi _{\tau }$ via the canonical isomorphisms $q^{-1}(\alpha, \beta ) = {\mathbb {P}}(T_{\epsilon _1}A)\times {\mathbb {P}}(T_{\epsilon _2}A) = \nu ^{-1}(\psi _{\tau }(\alpha,\beta ))$. Finally, consider the case when $\alpha =\beta$ and $\alpha$ is a node. Locally analytically around the node $\alpha$ of $A/\langle (\tau, -1)\rangle$, the morphism $q$ is isomorphic to the composition

\[ (\mathrm{Bl}_0({\mathbb{C}}^2/\iota))^{[2]}\to (\mathrm{Bl}_0({\mathbb{C}}^2/\iota))^{(2)} \to ({\mathbb{C}}^2/\iota)^{(2)}, \]

where $\iota (x,y)=(-x,-y)$ is the restriction of the involution $(\tau, -1)$. The rational map $\phi _{\tau }$ is identified with $\Phi \colon (\mathrm {Bl}_0( {\mathbb {C}}^2/\iota ))^{[2]} \dashrightarrow ( {\mathbb {C}}^2)^{[4]}$ which maps a general $\zeta =(\alpha,\beta )$ to the subscheme $\xi =(\alpha,\iota (\alpha ), \beta, \iota (\beta ))$. Then Proposition 2.8 implies that $\phi _\tau$ extends to a closed embedding $(\mathrm {Bl}_0( {\mathbb {C}}^2/\iota ))^{[2]}\hookrightarrow ( {\mathbb {C}}^2)^{[4]}$, completing the proof.

The various intersections of the submanifolds $\overline {W_{\tau }}\subset K$ are as follows.

Lemma 2.10

  1. (i) For $\tau _1 \neq \tau _2$ in $A_2$ we have $\overline {W_{\tau _1}}\cap \overline {W_{\tau _2}}=\overline {V_{\tau _1+\tau _2, \tau _1}}$.

  2. (ii) For pairwise distinct $\tau _1$, $\tau _2$, $\tau _3$ in $A_2$, the intersection $\overline {W_{\tau _1}}\cap \overline {W_{\tau _2}}\cap \overline {W_{\tau _3}}$ consists of the four distinct points of the set

    \[ \nu^{-1}( \{ (a, a+\tau_1+\tau_2, -a+\tau_1, -a+\tau_2) \in A_0^{(4)}, \ a\in A_{2, \tau_1+\tau_2+\tau_3}\}) \subset K. \]
  3. (iii) The intersection of four or more distinct submanifolds $\overline {W_{\tau }}$ is empty.

Proof. (i) A point $\xi \in A_0^{(4)}$ belongs to ${W_{\tau _1}}\cap {W_{\tau _2}}$ if and only if it can be written as $\xi =(a,b,-a+\tau _1, -b+\tau _1)$ as well as $\xi =(a', b' , -a'+\tau _2, -b'+\tau _2)$, for some $a,b,a',b'$ in $A$. Then, we may assume that $a'=a$; since $\tau _1\neq \tau _2$, it follows that $-a+\tau _1 \neq -a'+\tau _2$. This forces either $b' =-a+\tau _1$ or $-b'+\tau _2= -a +\tau _1$; in both cases, we can write

\[ \xi = (a, -a+\tau_1, -a+\tau_2, a+\tau_1+\tau_2), \]

i.e. $\xi \in V_{\tau _1+\tau _2, \tau _1}$. Thus, ${W_{\tau _1}}\cap {W_{\tau _2}}=V_{\tau _1+\tau _2, \tau _1}$ and, hence, for their strict transforms we have $\overline {W_{\tau _1}}\cap \overline {W_{\tau _2}}=\overline {V_{\tau _1+\tau _2, \tau _1}}$.

(ii) By part (i), a point $\xi$ lies in the intersection $W_{\tau _1}\cap W_{\tau _2}\cap W_{\tau _2}$ if and only if it can be written as $\xi = (a, a+\tau _1+\tau _2, -a+\tau _1, -a+\tau _2)$ as well as $\xi =(a'',b'', -a''+\tau _3, -b''+\tau _3)$. Again, we may assume $a''= a$. Then we must have $-a+\tau _3= a+\tau _1+\tau _2$; equivalently, $a\in A_{2,\tau _1+\tau _2+\tau _3}$. Moreover, either we have $b''=-a+\tau _1$ or $b''=-a+\tau _2$. Note that, since $a\in A_{2, \tau _1+\tau _2+\tau _3}$, in the first case $-b''+\tau _3= -a+\tau _2$, while in the second case $-b''+\tau _3= - a+\tau _1$. This shows that

\[ W_{\tau_1}\cap W_{\tau_2}\cap W_{\tau_3} = \{ (a, a+\tau_1+\tau_2, -a+\tau_1, -a+\tau_2) \in A_0^{(4)}, \ a\in A_{2, \tau_1+\tau_2+\tau_3}\}. \]

This set consists of four distinct points. Indeed, $G$ acts transitively on it and the subgroup of $G$ generated by the $(\tau _i, -1)$, $i=1,2,3$ acts trivially. Moreover, it is easy to see that if $g\in G$ is not an element of this subgroup, then it has no fixed points in the above set. Hence, the intersection $W_{\tau _1}\cap W_{\tau _2}\cap W_{\tau _3}$ is in bijection with $( {\mathbb {Z}}/2 {\mathbb {Z}})^2$. No point of this set belongs to the exceptional locus of the Hilbert–Chow morphism, so that $\overline {W_{\tau _1}}\cap \overline {W_{\tau _2}}\cap \overline {W_{\tau _3}}$ consists of four distinct points.

(iii) Note that $W_{\tau _1}\cap W_{\tau _2}\cap W_{\tau _3}\cap W_{\tau _1+\tau _2+\tau _3}$ is empty. We have seen above that if $\theta \in A_2\setminus \{\tau _1, \tau _2, \tau _3, \tau _1+\tau _2+\tau _3\}$, the action of $(\theta, -1)$ on $W_{\tau _1}\cap W_{\tau _2}\cap W_{\tau _3}$ is free; as ${W_{\theta }}$ is clearly fixed by $(\theta, -1)$, it follows that $W_{\tau _1}\cap W_{\tau _2}\cap W_{\tau _3}\cap W_{\theta } =\emptyset$.

2.11 Fixed loci

We calculate the fixed locus of the automorphisms $g\in G$. Our results confirm those of Oguiso [Reference OguisoOgu20] and Kamenova, Mongardi and Oblomkov [Reference Kamenova, Mongardi and OblomkovKMO22].

Proposition 2.12 Let $g\neq 1\in G$. Then:

  • if $g=(\tau, 1)$, with $\tau \neq 0\in A_2$, the fixed locus $K^g$ is the disjoint union of the eight $\mathrm {K}3$ surfaces $\overline {V_{\tau, \theta }}$, for $\theta$ varying in $A_2$;

  • if $g=(\tau, -1)$ with $\tau \in A_2$, the fixed locus $K^g$ is the union of the $\mathrm {K}3^{[2]}$ type variety $\overline {W_{\tau }}$ and $140$ isolated points; the isolated points are given by the set

    \[ \biggl\{(\epsilon_1, \epsilon_2, \epsilon_3, \epsilon_4) , \ \epsilon_i \in A_{2,\tau} \text{ pairwise distinct and such that } \sum_{i=1}^4 \epsilon_i =0 \in A \biggr\}. \]

Proof. Let $g=(\tau, 1)$ with $\tau \neq 0\in A_2$. The action of $g$ on $\xi =(a,b,c,d)\in A^{(4)}$ is given by $\xi \mapsto (a+\tau, b+\tau, c+\tau, d+\tau )$. We see immediately that the surfaces $V_{\tau, \theta }$ are fixed by $g$. Conversely, if $\xi$ is fixed and $a$ belongs to its support, then also $a+\tau$ must be in the support. Therefore, $\xi =(a, a+\tau, b, b+\tau )$; if $\xi \in A^{(4)}_0$, then $2a+2b=0$, so that $\theta := a+b$ belongs to $A_2$ and $\xi \in V_{\tau, \theta }$. By the description of $\overline {V_{\tau, \theta }}$ given in Lemma 2.9(i), the fixed locus of $g$ in $\nu ^{-1}(V_{\tau, \theta })$ is precisely $\overline {V_{\tau, \theta }}$. Since the fixed locus is smooth, $K^g$ is the disjoint union of the eight K$3$ surfaces $\overline {V_{\tau, \theta }}$, for $\theta \in A_2$.

Assume now that $g=(\tau, -1)$. If $\xi =(a,b,c,d) \in A^{(4)}$, the action of $g$ is given by $\xi \mapsto (-a+\tau, -b+\tau, -c+\tau, -d+\tau )$. Note that $W_{\tau }$ is contained in the fixed locus of $g$ on $A^{(4)}_0$. Moreover any $\xi$ entirely supported at points of $A_{2,\tau }$ is fixed by $g$. Conversely, let $\xi \in A_0^{(4)}$ be fixed. Then we have the following possibilities. If the support of $\xi$ does not intersect $A_{2, \tau }$, we have $\xi = (a, -a+\tau, b, -b+\tau )$ for some $a,b\in A\setminus A_{2, \tau }$ and $\xi$ belongs to $W_{\tau }$. If the support of $\xi$ contains some $\epsilon \in A_{2,\tau }$ and some $a\notin A_{2,\tau }$, then $\xi = (a, -a+\tau, \epsilon, b)$ for some $b\in A$; as $\xi \in A_0^{(4)}$, we must have $b=\epsilon$, and hence $\xi$ belongs to $W_{\tau }$. Finally, assume that $\xi =( \epsilon _1, \epsilon _2, \epsilon _3, \epsilon _4)$ is entirely supported at $A_{2,\tau }$. As $\xi$ belongs to $A^{(4)}_0$ we have $\sum _i \epsilon _i =0$ in $A$. If the support of $\xi$ consists of four distinct points, then $\xi \notin W_{\tau }$. Otherwise, $\xi =(\epsilon _1, \epsilon _1, \epsilon _2, \epsilon _2)$ lies in $W_{\tau }$.

We conclude that the fixed locus of $g$ acting on $A^{(4)}_0$ consists of $W_{\tau }$ and the isolated points in the set

\[ \biggl\{(\epsilon_1, \epsilon_2, \epsilon_3, \epsilon_4) , \ \epsilon_i \in A_{2,\tau} \text{ pairwise distinct and such that } \sum_{i=1}^4 \epsilon_i =0 \in A \biggr\}. \]

There are $140$ isolated fixed points. Indeed, the total number of ordered sequences $(\epsilon _1, \epsilon _2, \epsilon _3, \epsilon _4)$ such that $\sum _i \epsilon _i = 0$ is $16^3$. There are $16$ such sequences with support a single point, and $6\binom{16}{2}=16 \cdot 45$ with support two distinct points. The number of ordered sequences as above supported on four distinct points is $16^3-16-16\cdot 45=16\cdot 210$; under the symmetric group, each of them has an orbit of cardinality $24$. Hence, the number of isolated fixed points is $\tfrac {210\cdot 16}{24} = 140$.

Therefore, $K^g$ consists of $140$ isolated fixed points and the fixed locus of $g$ acting on $\nu ^{-1}(W_\tau )$. To conclude we have to check that the latter coincides with $\overline {W_{\tau }}$. Using the description of $\overline {W_{\tau }}$ given in the proof of Lemma 2.9(ii) and the second assertion of Proposition 2.8 it is readily seen that $\nu ^{-1}(\xi ) \cap K^g = \nu ^{-1}(\xi ) \cap \overline {W_{\tau }}$ for any $\xi \in W_{\tau }$.

Remark 2.13 Let $h=(\epsilon, \pm 1)$ be an automorphism in $\operatorname {Aut}_0(K^3(A))=A_4\rtimes \langle -1\rangle$. One can see directly from Definition 2.6 that $h$ maps $\overline {W_{\tau }}$ to $\overline {W_{\tau +2\epsilon }}$, for any $\tau \in A_2$.

We can finally give the proof of the main result of this section.

Proof of Theorem 2.2 Let us first assume that $K$ is the generalized Kummer variety associated to an abelian surface $A$. Let $g= (\tau, 1)\in G$, with $\tau \neq 0\in A_2$. Consider a component $\overline {V_{\tau, \theta }}$ of $K^g$. The surface $V_{\tau, \theta }$ is clearly contained in $W_{\theta }$ and, hence, the same holds for their strict transforms. Now let $g= (\tau, -1)\in G$. Let $P$ be an isolated fixed point in $K^{g}$. Then we have $P=(\epsilon _1, \epsilon _2, \epsilon _3, \epsilon _4)$ for some pairwise distinct $\epsilon _i \in A_{2,\tau }$ summing up to $0$ in $A$. The last condition implies that we can write $P=(\epsilon _1, \epsilon _2, \epsilon _1+\theta, \epsilon _2+\theta )$ for some $0\neq \theta \in A_{2}$. This is the same as

\[ P = (\epsilon_1, \epsilon_2, -\epsilon_1+\theta+\tau, -\epsilon_2+\theta+\tau), \]

which lies in $\overline {W_{\tau +\theta }}$. By Lemma 2.12 we conclude that for a generalized Kummer variety the union of the fixed loci $K^g$ consists of the $16$ varieties $\overline {W_{\tau }}$ of $\mathrm {K}3^{[2]}$ type. The intersections of these components were computed in Lemma 2.10.

Now let $f\colon \mathcal {K}\to B$ be a smooth and proper family of $\mathrm {Kum}^3$ manifolds over a connected base $B$, with a fibre $\mathcal {K}_0$ isomorphic to the generalized Kummer variety on an abelian surface $A$. By the result of Hassett and Tschinkel [Reference Hassett and TschinkelHT13, Theorem 2.1], up to a finite étale base change, there is a fibrewise action of $G=A_2\times \langle -1 \rangle$ on $\mathcal {K}$. By [Reference FujikiFuj83, Lemma 3.10], this action is locally trivial with respect to $f$, i.e. any $x\in \mathcal {K}$ has a neighborhood $U_x=(f^{-1}f(U_x)\cap \mathcal {K}_{f(x)} )\times f(U_x)$ such that the action on $U_x$ of the stabilizer subgroup of $x$ is induced by that on the fibre $\mathcal {K}_{f(x)}$. Therefore, for any $\tau \in A_{2}$ the fixed locus $\mathcal {K}^g$ of $g=(\tau, -1)$ contains a smooth family $\overline {\mathcal {W}_{\tau }}\to B$ of manifolds of $\mathrm {K}3^{[2]}$ type as the unique component of $\mathcal {K}^g$ with positive-dimensional intersection with the fibres of $f$. We conclude that for any $b\in B$ we have

\[ \bigcup_{1\neq g\in G} (\mathcal{K}_b)^g = \bigcup_{\tau\in A_2} \overline{\mathcal{W}_{\tau}}_{\lvert_{\mathcal{K}_b}}, \]

and that the components $\overline {\mathcal {W}_{\tau }}_{\lvert _{\mathcal {K}_b}}$ intersect as claimed.

3. A Lagrangian fibration

Throughout this section, $J$ denotes a general principally polarized abelian surface. We fix a symmetric theta divisor $\Theta \subset J$, which is unique up to translation by a point of $J_2$. For any integer $d$, we consider the Mukai vector $v_d:= (0, 2\Theta, d-4)$ on $J$ and the moduli space $M_J(v_d)$ of $\Theta$-semistable sheaves with Mukai vector $v_d$ ([Reference Huybrechts and LehnHL10]). The Albanese fibration

\[ M_J(v_d) \xrightarrow{ \ \mathrm{alb} \ } \operatorname{\mathrm{Pic}}^0(J)\times J \]

is isotrivial [Reference YoshiokaYos01]. We denote the fibre of $\mathrm {alb}$ over $(\mathcal {O}_J, 0 )$ by $K_J(v_d)$. For $d$ even, $K_J(v_d)$ is singular and admits a crepant resolution of $\mathrm {OG}6$ type, by [Reference O'GradyO'Gr03] and [Reference Lehn and SorgerLS06]. When $d$ is odd, $K_J(v_d)$ is smooth, and it is a hyper-Kähler sixfold of $\mathrm {Kum}^3$ type [Reference YoshiokaYos01]. In this case, we consider the group $G \subset \operatorname {Aut}_0(K_{J}(v_d))$ introduced in Definition 2.1.

The main result of this section is the following.

Theorem 3.1 The quotient $K_J(v_3)/ G$ is birational to $\operatorname {\mathrm {Km}}(J)^{[3]}$.

We will, in fact, show that for any odd $d$ the quotient $K_J(v_d)/G$ is birational to a variety of $\mathrm {K}3^{[3]}$ type. Before proceeding with the proof we need to fix some notation.

3.2 Preliminaries

Let $C$ be a smooth projective curve of genus $g$. Its Picard variety $\operatorname {\mathrm {Pic}}^{\bullet }(C)=\bigsqcup _d \operatorname {\mathrm {Pic}}^d(C)$ is the group of line bundles modulo isomorphism. Tensor product gives multiplication maps $m\colon \operatorname {\mathrm {Pic}}^{a}(C) \times \operatorname {\mathrm {Pic}}^{b}(C) \to \operatorname {\mathrm {Pic}}^{a+b}(C)$. We denote by $[n]\colon \operatorname {\mathrm {Pic}}^d(C)\to \operatorname {\mathrm {Pic}}^{nd}(C)$ the $n$th power map. The Jacobian of $C$ is the abelian variety $\operatorname {\mathrm {Pic}}^0(C)$; for $d\neq 0$, the variety $\operatorname {\mathrm {Pic}}^d(C)$ is a torsor under $\operatorname {\mathrm {Pic}}^0(C)$.

Assume now that $\pi \colon \tilde {C}\to C$ is an étale double cover, so that $\tilde {C}$ is of genus $2g-1$. In this situation, for each integer $d$, we have:

  • the pull-back map $\pi ^*\colon \operatorname {\mathrm {Pic}}^{d}(C)\to \operatorname {\mathrm {Pic}}^{2d}(\tilde {C})$;

  • the norm map $\mathrm {Nm}_\pi \colon \operatorname {\mathrm {Pic}}^d(\tilde {C})\to \operatorname {\mathrm {Pic}}^d({C})$;

  • the covering involution $\sigma \colon \tilde {C}\to \tilde {C}$, inducing $\sigma ^*\colon \operatorname {\mathrm {Pic}}^{d}(\tilde {C})\to \operatorname {\mathrm {Pic}}^d(\tilde {C})$.

The composition $\mathrm {Nm}_\pi \circ \pi ^*\colon \operatorname {\mathrm {Pic}}^d(C)\to \operatorname {\mathrm {Pic}}^{2d}(C)$ coincides with multiplication by $2$, while $\pi ^*\circ \mathrm {Nm}_\pi \colon \operatorname {\mathrm {Pic}}^d(\tilde {C})\to \operatorname {\mathrm {Pic}}^{2d}(\tilde {C})$ is given by $L\mapsto L\otimes \sigma ^*(L)$.

We refer to Mumford's paper [Reference MumfordMum74] for the following. The étale double cover $\pi$ is uniquely determined by a $2$-torsion line bundle $\eta$ on $C$, which is the only non-trivial element in $\ker (\pi ^*\colon \operatorname {\mathrm {Pic}}^{0}(C)\to \operatorname {\mathrm {Pic}}^0(\tilde {C}))$. Moreover, the image of the pull-back map $\pi ^*\colon \operatorname {\mathrm {Pic}}^{d}(C)\to \operatorname {\mathrm {Pic}}^{2d}(\tilde {C})$ is precisely the fixed locus of $\sigma ^*$. The involution $\sigma ^*$ acts as the inverse $L\mapsto L^{\vee }$ on the kernel of the norm map $\mathrm {Nm}_\pi \colon \operatorname {\mathrm {Pic}}^0(\tilde {C})\to \operatorname {\mathrm {Pic}}^0(C)$. This kernel has two connected components; the one containing the neutral element is the Prym variety $P(\pi )$ of the cover, an abelian variety of dimension $g-1$. The fibres of $\mathrm {Nm}_\pi \colon \operatorname {\mathrm {Pic}}^d(\tilde {C})\to \operatorname {\mathrm {Pic}}^d(C)$ are torsors under $P(\pi )\times {\mathbb {Z}}/2 {\mathbb {Z}}$.

3.3 Geometric set-up

Let $\Theta \subset J$ be the general principally polarized abelian surface of Picard rank $1$, with $\Theta$ a symmetric theta divisor. By Riemann–Roch $H^0(J, \mathcal {O}_J(2\Theta ))=4$; we will identify the complete linear system $\lvert 2\Theta \rvert$ with $ {\mathbb {P}}^3$. It is classically known that this linear system is base-point free and induces an embedding of $J/\pm 1$ into $ {\mathbb {P}}^{3,\vee }$ as a quartic surface with $16$ nodes. Blowing-up the nodes of this quartic surface one obtains the Kummer K3 surface $\operatorname {\mathrm {Km}}(J)$ associated to $J$. We denote by $H$ the divisor on $\operatorname {\mathrm {Km}}(J)$ obtained as pull-back of a hyperplane section of $J/\pm 1\subset {\mathbb {P}}^{3,\vee }$. The linear system $|H|$ is naturally identified with $\lvert 2\Theta \rvert = {\mathbb {P}}^3$.

Consider the universal family of genus-$5$ curves $\tilde {\mathcal {C}}\to \lvert 2\Theta \rvert$. The involution $-1$ of $J$ acts trivially on $\lvert 2\Theta \rvert$, and we consider the quotient family $\mathcal {C}\to {\mathbb {P}}^3$, which is identified with the hyperplane linear system on $J/\pm 1\subset {\mathbb {P}}^{3,\vee }$. There is a degree $2$ morphism $\pi \colon \tilde {\mathcal {C}}\to \mathcal {C}$ over $ {\mathbb {P}}^3$. We denote by $\mathcal {D}\to \lvert H\rvert$ the universal family of curves over the linear system $\lvert H\rvert$ on $\operatorname {\mathrm {Km}}(J)$; clearly, $\mathcal {C}_b=\mathcal {D}_b$ if and only if $\mathcal {C}_b$ does not pass through any of the nodes of $J/\pm 1$.

A detailed study of the linear system $\lvert 2\Theta \rvert$ and the map $\pi \colon \tilde {\mathcal {C}}\to \mathcal {C}$ can be found in Verra's article [Reference VerraVer87]. In particular, he shows that, for any $b\in {\mathbb {P}}^3$, the curve $\tilde {\mathcal {C}}_b$ is smooth if and only if $\mathcal {C}_b$ is so, and in this case the Prym variety of the étale double cover $\pi _b \colon \tilde {\mathcal {C}}_b \to \mathcal {C}_b$ is the abelian surface $J$. Moreover, $\mathcal {C}_b$ is smooth if and only if it does not contain any node of $J/\pm 1$ and it does not lie on a tangent hyperplane.

3.4 Beauville–Mukai systems

Taking the Jacobian of the smooth curves in the families introduced above yields families of abelian varieties over a Zariski open subvariety of $ {\mathbb {P}}^3$. The total space of these families can be compactified considering suitable moduli spaces of stable sheaves on $J$ and $\mathrm {Km}(J)$. The construction leads to certain hyper-Kähler varieties equipped with Lagrangians fibrations, called Beauville–Mukai systems [Reference BeauvilleBea91, Reference MukaiMuk84].

Consider the Mukai vector $v_d=(0, 2\Theta, d-4)$ on $J$, for an integer $d$. The moduli space $M_J(v_d)$ parametrizes pure dimension $1$ sheaves on $J$ which are push-forward of semistable and torsion-free sheaves of rank $1$ and degree $d$ supported on curves algebraically equivalent to $2\Theta$. Mapping a sheaf to its support [Reference Le PotierLeP93] gives a morphism

\[ M_J(v_d) \xrightarrow{\ \mathrm{supp} \ } \operatorname{\mathrm{Pic}}^0(J)\times {\mathbb{P}}^3. \]

We define the degree-$d$ relative compactified Jacobian as

\[ \overline{\operatorname{\mathrm{Pic}}}^d(\tilde{\mathcal{C}}/ {\mathbb{P}}^3) := \mathrm{supp}^{-1}(\{ \mathcal{O}_J \} \times {\mathbb{P}}^3). \]

If $b\in {\mathbb {P}}^3$ corresponds to a smooth curve in $|2\Theta |$, the fibre over $b$ of the support morphism $\mathrm {supp}\colon \overline {\operatorname {\mathrm {Pic}}}^d(\tilde {\mathcal {C}}/ {\mathbb {P}}^3) \to {\mathbb {P}}^3$ is the degree-$d$ Picard variety of the curve. The Albanese morphism $\mathrm {alb}\colon \overline {\operatorname {\mathrm {Pic}}}^d(\tilde {\mathcal {C}}/ {\mathbb {P}}^3) \to J$ is an isotrivial fibration with fibre $K_J(v_d)$, and it maps a sheaf $F\in K_J(v_d)$ to the sum $\sum c_2(F)\in J$, see [Reference WieneckWie18, § 6].

A similar construction can be done for the family of curves $\mathcal {D}$ on $\operatorname {\mathrm {Km}}(J)$. Given an integer $d$, we consider the Mukai vector $w_d= (0, H, d-2)$ on $\operatorname {\mathrm {Km}}(J)$. The degree $d$ relative compactified Jacobian of curves in $|H|$ is the moduli space

\[ \overline{\operatorname{\mathrm{Pic}}}^d(\mathcal{D}/ {\mathbb{P}}^3) := M_{\operatorname{\mathrm{Km}}(J)}(w_d), \]

of semistable sheaves with Mukai vector $w_d$, with respect to a fixed $w_d$-generic polarization. It is a hyper-Kähler variety of $\mathrm {K}3^{[3]}$ type by [Reference O'GradyO'Gr97] and [Reference YoshiokaYos99]. The moduli space $M_{\operatorname {\mathrm {Km}}(J)}(w_d)$ parametrizes pure dimension-$1$ sheaves on $\operatorname {\mathrm {Km}}(J)$ which are push-forward of torsion-free sheaves of rank $1$ and degree $d$ on curves in the linear system $\lvert H \rvert$. There is a morphism $\mathrm {supp} \colon \overline {\operatorname {\mathrm {Pic}}}^d(\mathcal {D}/ {\mathbb {P}}^3)\to {\mathbb {P}}^3$ which is a Lagrangian fibration, whose general fibres are the degree-$d$ Picard varieties of the smooth curves in $\lvert H \rvert$.

3.5 The relative norm map

We denote by $B\subset {\mathbb {P}}^3$ the locus parametrizing smooth curves. The maps introduced in § 3.2 give morphisms over $B$. Hence, we obtain rational maps

\begin{align*} \mathrm{Nm}_\pi & \colon \overline{\operatorname{\mathrm{Pic}}}^d(\tilde{\mathcal{C}}/{\mathbb{P}}^3) \dashrightarrow \overline{\operatorname{\mathrm{Pic}}}^{d}({\mathcal{D}}/{\mathbb{P}}^3);\\ \pi^* & \colon \overline{\operatorname{\mathrm{Pic}}}^d(\mathcal{D}/{\mathbb{P}}^3) \dashrightarrow \overline{\operatorname{\mathrm{Pic}}}^{2d}(\tilde{\mathcal{C}}/{\mathbb{P}}^3). \end{align*}

The pull-back $\pi ^*$ has been studied by Rapagnetta [Reference RapagnettaRap07] and by Mongardi, Rapagnetta and Saccà [Reference Mongardi, Rapagnetta and SaccàMRS18]. They show that it gives a degree-$2$ dominant rational map

\[ \overline{\operatorname{\mathrm{Pic}}}^d(\mathcal{D}/{\mathbb{P}}^3) \dashrightarrow K_J(v_{2d}), \]

exhibiting a variety of OG6 type as the quotient of a variety of $\mathrm {K}3^{[3]}$ type by a birational symplectic involution. We consider instead the norm map $\mathrm {Nm}_{\pi }$.

Lemma 3.6 For each $d$, the restriction $\mathrm {Nm}_\pi \colon K_J(v_d) \dashrightarrow \overline {\operatorname {\mathrm {Pic}}}^{d}({\mathcal {D}}/ {\mathbb {P}}^3)$ is dominant and generically finite of degree $2^5$.

Proof. For each $k$, the $k$th power map $[k]\colon \overline {\operatorname {\mathrm {Pic}}}^d(\tilde {\mathcal {C}}/ {\mathbb {P}}^3) \dashrightarrow \overline {\operatorname {\mathrm {Pic}}}^{kd}(\tilde {\mathcal {C}}/ {\mathbb {P}}^3)$ restricts to a rational map $[k]\colon K_J(v_d) \dashrightarrow K_J(v_{kd})$. This can be easily checked at smooth curves $\tilde {\mathcal {C}_b}$ in the family using the description of the Albanese map $L\mapsto \sum c_2(L)$.

Consider the following commutative diagram.

Both the horizontal maps are dominant and generically finite of the same degree $2^6$. Indeed, they preserve the fibres of the respective support morphisms, which, generically, are torsors under abelian threefolds. Hence, the degree of $[2]$ is the number of points of order $2$ on an abelian threefold.

It thus suffices to prove the lemma for $\mathrm {Nm}_\pi \colon K_J(v_{2d}) \dashrightarrow \overline {\operatorname {\mathrm {Pic}}}^{2d}({\mathcal {D}}/ {\mathbb {P}}^3)$. Since $K_J(v_{2d})$ is the closure of the image of $\pi ^*$, the involution $\sigma ^*$ is the identity on $K_J(v_{2d})$. Hence, the rational map

coincides with multiplication by $[2]$; this composition is therefore dominant and generically finite of degree $2^6$. Since $\pi ^*\colon \overline {\operatorname {\mathrm {Pic}}}^{2d}({\mathcal {D}}/ {\mathbb {P}}^3)\dashrightarrow K_J(v_{4d})$ is generically finite of degree $2$, it follows that $\mathrm {Nm}_\pi$ is dominant and it has degree $2^5$.

3.7 The action of $G$

Assume now that $d$ is odd, so that $K_J(v_d)$ is a smooth variety of $\mathrm {Kum}^3$ type and the support map gives a Lagrangian fibration on it. In this case the group $\operatorname {Aut}_0(K_J(v_d))$ has been explicitly identified by Kim in [Reference KimKim21].

Proposition 3.8 Let $d$ be odd. Then $G=\operatorname {\mathrm {Pic}}^0(J)_2 \times \langle -1\rangle$, where $-1$ acts on $K_J(v_d)$ via the pull-back $F\mapsto (-1)^*(F)$ of sheaves, and the action of $L\in \operatorname {\mathrm {Pic}}^0(J)_2$ is given by $F\mapsto F\otimes L$. Any element of $G$ preserves the fibres of the support fibration.

Proof. We first show that $\operatorname {\mathrm {Pic}}^0(J)_2\times \langle -1\rangle$ acts on $K_J(v_d)$ via automorphisms trivial on the second cohomology and which preserve the Lagrangian fibration given by the support morphism.

Since $-1\colon J\to J$ acts trivially on the cohomology of $J$ in even degrees, the pull-back of sheaves defines an automorphism $(-1)^*$ of $K_J(v_d)$; this automorphism is trivial for $d$ even but not for $d$ odd. By [Reference Mongardi and WandelMW15, Lemma 2.34], the action of $(-1)^*$ is symplectic, i.e. the induced action on the transcendental cohomology $H^2_{\mathrm {tr}}(K_J(v_d), {\mathbb {Z}})$ is the identity. Since all curves in $\lvert 2\Theta \rvert$ are stable under $-1$, the automorphism $(-1)^*$ preserves all fibres of the support fibration. The Picard rank of $K_J(v_d)$ is $2$, because $J$ is a general abelian surface by assumption (see [Reference YoshiokaYos01]). Hence, by Lemma 3.10, the induced action of $(-1)^*$ on $\operatorname {\mathrm {NS}}(K_J(v_d))$ is also the identity. Since $H^2(K_J(v_d), {\mathbb {Z}})$ is torsion free, we conclude that $(-1)^*$ acts trivially on the second cohomology.

Now let $L\in \operatorname {\mathrm {Pic}}^0(J)_2$. Let $B\subset \lvert 2\Theta \rvert$ be the locus of smooth curves, and consider the universal curve $j\colon \tilde {\mathcal {C}}_B\hookrightarrow J\times B$. By [Reference WieneckWie18, Lemma 6.9], the corresponding pull-back $j^*\colon \operatorname {\mathrm {Pic}}^0(J)\times B \to \overline {\operatorname {\mathrm {Pic}}}^0(\tilde {\mathcal {C}}/ {\mathbb {P}}^3)_B$ is an injection of abelian schemes. Note that the covering involution $\sigma$ for the cover $\pi \colon \tilde {\mathcal {C}}_B\to {\mathcal {C}}_B$ is the restriction of $-1\times \mathrm {id}_B$. Hence, the induced involution $\sigma ^*$ on $\overline {\operatorname {\mathrm {Pic}}}^0(\tilde {\mathcal {C}}/ {\mathbb {P}}^3)_B$ is the inverse map on $j^*_b(\operatorname {\mathrm {Pic}}^0(J))$, for each $b\in B$. As $K_J(v_0)$ is the fixed locus of this involution, the $2$-torsion line bundle $L$ on $J$ corresponds uniquely to a $2$-torsion section $s_L$ of $K_J(v_0)_B\to B$. For any $d$, the variety $K_J(v_d)_B$ is a torsor under $K_J(v_0)_B$ and, hence, $F\mapsto F\otimes L$ induces a birational automorphism $g_L$ of $K_J(v_d)$, which restricts to a translation on the smooth fibres of the support morphism. This implies that $g_L$ is symplectic. We deduce from Lemma 3.10 that $g_L$ extends to a regular automorphism of $K_J(v_d)$, whose induced action is trivial on $\operatorname {\mathrm {NS}}(K_J(v_d))$ as well.

According to [Reference KimKim21, Theorem 5.1] we have described all automorphisms in $\operatorname {Aut}_0(K_J(v_d))$ preserving the fibration $\mathrm {supp}\colon K_J(v_d)\to {\mathbb {P}}^3$. We claim now that an automorphism in $\operatorname {Aut}_0(K_J(v_d))$ which does not preserve the fibration cannot fix a component of dimension $4$ on $K_J(v_d)$; this will show that $G=\operatorname {\mathrm {Pic}}^0(J)_2\times \langle -1\rangle$.

Let $D = \mathrm {supp}^* (\mathcal {O}_{ {\mathbb {P}}^n}(1)) \in \operatorname {\mathrm {Pic}}(K_J(v_d))$ be the line bundle inducing the fibration. Since $\operatorname {Aut}_0(K_J(v_d))$ acts trivially on $H^2(K_J(v_d), {\mathbb {Z}})$, it acts on $|D|= {\mathbb {P}}^3=|2\Theta |$. We obtain an action of $\operatorname {Aut}_0(K_J(v_d))/(\operatorname {\mathrm {Pic}}^0(J)_2\times \langle -1\rangle ) \cong ( {\mathbb {Z}}/2 {\mathbb {Z}})^4$ on $|2\Theta |$, which, up to conjugation by an automorphism of $ {\mathbb {P}}^3$, is identified with the action generated by

\begin{align*} (x,y,z,w) \mapsto (z,w,x,y),&\quad (x,y,z,w) \mapsto (y,x,w,z), \\ (x,y,z,w) \mapsto (x,y,-z,-w),&\quad (x,y,z,w) \mapsto (x, -y, z, -w); \end{align*}

see [Reference Gonzalez-DorregoGon94, Lemma 1.52, Note 1.4]. Thus, $h\in \operatorname {Aut}_0(K_J(v_d))$ either acts trivially on $|2\Theta |$ or it fixes a pair of skew lines. Assume by contradiction that $h\in \operatorname {Aut}_0(K_J(v_d))$ fixes a fourfold but does not preserve the support fibration. Then $(K_J(v_d))^h$ contains one of the varieties ${Z_{j}}$ of $\mathrm {K}3^{[2]}$ type from Theorem 2.2, and the image of $Z_j$ via the fibration must be a line $R \subset |2\Theta |$ fixed by $h$. But, by a theorem of Matsushita [Reference MatsushitaMat99], the $\mathrm {K}3^{[2]}$ variety $Z_j$ does not admit any non-constant morphism to $ {\mathbb {P}}^1$.

Remark 3.9 By [Reference VerraVer87], the relative Prym variety $\mathcal {P}\to B$ of the double cover $\pi \colon \tilde {\mathcal {C}}_B\to \mathcal {C}_B$ is an isotrivial family over $B$ with fibre $J$. Since the involution $\sigma ^*$ acts as the inverse on the image of the pull-back $j^*$, the constant abelian scheme $j^*(\operatorname {\mathrm {Pic}}^0(J)\times B)$ is contained into, and hence equal to, the relative Prym variety.

Lemma 3.10 Let $f\colon X\to {\mathbb {P}}^n$ be a Lagrangian fibration, where $X$ is a $2n$-dimensional projective hyper-Kähler variety of Picard rank $2$. Let $g\colon X\dashrightarrow X$ be a birational automorphism such that $f\circ g = f$. Then $g$ extends to a regular automorphism of $X$, and $g^*_{\lvert _{ \operatorname {\mathrm {NS}}(X)}}\colon \operatorname {\mathrm {NS}}(X)\to \operatorname {\mathrm {NS}}(X)$ is the identity.

Proof. The Néron–Severi group $\operatorname {\mathrm {NS}}(X)$ is a rank $2$ lattice of signature $(1,1)$. Any birational automorphism $g$ induces an isometry $g^*$ of $H^2(X, {\mathbb {Z}})$ which restricts to an isometry of $\operatorname {\mathrm {NS}}(X)$. Denote by $D\in \operatorname {\mathrm {NS}}(X)$ the class of the line bundle $f^*(\mathcal {O}_{ {\mathbb {P}}^n}(1))$, and pick $E\in \operatorname {\mathrm {NS}}(X)$ such that $D$ and $E$ generate $\operatorname {\mathrm {NS}}(X)\otimes {\mathbb {Q}}$. Since $f\circ g=f$, we have $g^*(D)=D$. Using that $D$ is isotropic one sees that this forces $g^*(E)=E$ as well, so that $g^*$ is the identity on $\operatorname {\mathrm {NS}}(X)$; by [Reference FujikiFuj81], $g$ extends to an automorphism of $X$.

We can now complete the proof of the main result of this section.

Proof of Theorem 3.1 Using the identification of $G$ given in Proposition 3.8 we will show that the norm map descends to a birational map $K_J(v_d)/G\dashrightarrow \overline {\operatorname {\mathrm {Pic}}}^d(\mathcal {D}/ {\mathbb {P}}^3)$, for any odd $d$. The varieties $\overline {\operatorname {\mathrm {Pic}}}^d(\mathcal {D}/ {\mathbb {P}}^3)$ are of $\mathrm {K}3^{[3]}$ type, and $\overline {\operatorname {\mathrm {Pic}}}^3(\mathcal {D}/ {\mathbb {P}}^3)$ is birational to $\operatorname {\mathrm {Km}}(J)^{[3]}$, see [Reference BeauvilleBea99, Proposition 1.3].

For a smooth curve $\tilde {\mathcal {C}}_b \in \lvert 2\Theta \rvert$ the norm map is induced by the map of divisors which sends $\sum _i a_i[P_i]$ to $\sum _i a_i [\pi (P_i)]$. It is then clear that $\mathrm {Nm}_\pi ((-1)^*(F))= \mathrm {Nm}_\pi (F)$ for any $F\in K_J(v_d)_B$. Instead let $L\in \operatorname {\mathrm {Pic}}^0(J)_2$, and let $s_L\colon {\mathbb {P}}^3 \dashrightarrow K_J(v_0)$ be the rational section defined by $s_L(b)= j_b^*(L)$, where $j\colon \tilde {\mathcal {C}}_B\hookrightarrow J\times B$ denotes the natural inclusion. By Remark 3.9, the section $s_L$ is contained in the relative Prym variety; in particular, the composition $\mathrm {Nm}_\pi \circ s_L$ gives the zero section of $\overline {\operatorname {\mathrm {Pic}}}^0(\mathcal {D}/ {\mathbb {P}}^3)_B\to {\mathbb {P}}^3$.

Therefore, for any $g\in G$ and $y\in K_J(v_d)_B$ we have $\mathrm {Nm}_\pi (g(y))= \mathrm {Nm}_\pi (y)$, which means that $\mathrm {Nm}_\pi \colon K_J(v_d)\dashrightarrow \overline {\operatorname {\mathrm {Pic}}}^d(\mathcal {D}/ {\mathbb {P}}^3)$ descends to a rational map

\[ \overline{\mathrm{Nm}}_\pi\colon K_J(v_d)/G \dashrightarrow \overline{\operatorname{\mathrm{Pic}}}^d(\mathcal{D}/{\mathbb{P}}^3). \]

This map is, in fact, birational, because $\mathrm {Nm}_\pi \colon K_J(v_d)\dashrightarrow \overline {\operatorname {\mathrm {Pic}}}^d(\mathcal {D}/ {\mathbb {P}}^3)$ is generically finite of degree $2^5$ by Lemma 3.6, and $G$ has also order $2^5$.

4. The K3 surface associated to a $\mathrm {Kum}^3$ variety

In this section we give the proof of Theorems 1.1 and 1.2. We let $K$ be a manifold of $\mathrm {Kum}^3$ type and consider the quotient $K/G$ by the group $G\cong ( {\mathbb {Z}}/2 {\mathbb {Z}})^5$ of Definition 2.1. We will show that the blow-up of the singular locus of $K/G$ yields a hyper-Kähler manifold $Y_K$ of $\mathrm {K}3^{[3]}$ type.

4.1 The singularities of $K/G$

By Theorem 2.2, the locus $Z=\bigcup _{1\neq g\in G} K^g$ is the union of $16$ irreducible components $Z_i$, for $i=1,\ldots, 16$. Denote by $X_i \subset K/G$, the image of $Z_i$ via the quotient map $q\colon K\to K/G$. We introduce the following stratification of $K/G$ into closed subspaces:

\[ X^j := \{x \in K/G \ \vert\ x \text{ belongs to at least } j \text{ components } X_i\}; \]

clearly, $X^0=K/G$.

Proposition 4.2 The subspace $X^j$ is empty for $j\geqslant 4$. For $j<4$, a point $x\in X^j\setminus X^{j+1}$ has a neighborhood $U_x\subset K/G$ analytically isomorphic to

\[ ({\mathbb{C}}^2/\iota)^{j} \times ({\mathbb{C}}^2)^{3-j}, \]

where $\iota$ is the involution $(x,y)\mapsto (-x,-y)$.

Proof. Theorem 2.2 implies immediately that $X^j$ is empty for $j\geqslant 4$. The group $G$ acts freely on $q^{-1}(X^0\setminus X^1)$ and, hence, $X^0\setminus X^1$ is smooth.

If $\mathcal {K}\to B$ is a smooth proper family of $\mathrm {Kum}^3$ manifolds, the quotient $\mathcal {K}/G \to B$ is a locally trivial family (see [Reference FujikiFuj83, Lemma 3.10]): any point $x\in \mathcal {K}/G$ has a neighborhood $U_x$ of the form to $f(U_x)\times (f^{-1}(f(x))\cap U_x)$. Therefore, to prove the proposition we may assume that $K$ is the generalized Kummer sixfold on an abelian surface $A$.

In this case the components $Z_{i}$ are the explicit $\overline {W_{\tau }}$, for $\tau \in A_2$ (see Definition 2.6). Recall that $\overline {W_{\tau }}\subset K$ is the unique positive-dimensional component of the fixed locus of $(\tau, -1)\in G$, and that the induced action of $G/\langle (\tau,-1) \rangle$ on $\overline {W_{\tau }}$ is faithful. For any $z\in \overline {W_{\tau }}$ there is a decomposition $T_z(K)=N_{\overline {W_{\tau }}\vert K, z}\oplus T_z(\overline {W_{\tau }})$, where the first factor is the normal space. The action of $(\tau, -1)$ on $T_z(K)$ is $(-1, 1)$. Now let $x\in X^1 \setminus X^2$, and let $z\in K$ be a preimage of $x$. By definition, $z$ belongs to exactly one component $\overline {W_{\tau }}$. The stabilizer is $G_z=\langle (\tau,-1) \rangle$. By the above, there exists a neighborhood $V_z$ of $z\in K$ such that $g(V_z)\cap V_z$ is empty for any $g \notin G_z$, and $V_z=( {\mathbb {C}}^2)^3$ with $(\tau, -1)_{\vert _{V_z}} =(\iota, \mathrm {id}, \mathrm {id})$. The image of $V_z$ under the quotient map is thus a neighborhood $U_x=( {\mathbb {C}}^2/\iota )\times ( {\mathbb {C}}^2)^2$ of $x$ in $K/G$.

In the other cases we proceed similarly. Let $x\in X^2\setminus X^3$, and let $z\in K$ be one of its preimages. Then $z$ belongs to two distinct components $\overline {W_{\tau _1}}$ and $\overline {W_{\tau _2}}$. The stabilizer $G_z\cong ( {\mathbb {Z}}/2 {\mathbb {Z}})^2$ is the subgroup $\langle (\tau _1,-1), (\tau _2,-1)\rangle$ of $G$. There is a decomposition $T_z(K) = N_{\overline {W_{\tau _1}}\vert K, z} \oplus N_{\overline {W_{\tau _2}}\vert K, z} \oplus T_z(\overline {W_{\tau _1}}\cap \overline {W_{\tau _2}})$, and the action of $G_z$ is generated by $(-1, 1, 1)$ and $(1, -1, 1)$. This implies that the image in $K/G$ of a sufficiently small neighborhood of $z$ in $K$ is isomorphic to $( {\mathbb {C}}^2/\iota )^2\times ( {\mathbb {C}}^2)$.

Finally, let $x\in X^3$ and $z\in K$ a preimage of it. Then $z$ belongs to exactly three components $\overline {W_{\tau _i}}$, for $i=1,2,3$. The stabilizer $G_z$ is the subgroup generated by the involutions $(\tau _i, -1)$, for $i=1,2,3$, and there is a decomposition $T_z(K)=\bigoplus _{i=1}^3 N_{\overline {W_{\tau _i}}\vert K,z}$ of the tangent space. The action of $(\tau _i,-1)$ on $T_{z}(K)$ is $-1$ on the $i$th summand and the identity on the complement. We conclude that there exists a neighborhood $V_z$ of $z$ in $K$ whose image in $K/G$ is isomorphic to $( {\mathbb {C}}^2/\iota )^3$.

4.3 The symplectic resolution of $K/G$

We will now conclude the proof of Theorem 1.1 in several steps. First we show that the blow-up of the singular locus of $K/G$ resolves the singularities.

Proposition 4.4 Let $Y_K:= \mathrm {Bl}_{X^1}(K/G)$ be the blow-up of the singular locus with reduced structure. Then $Y_K$ is a smooth manifold. It is identified with the quotient by $G$ of the blow-up of $K$ at $Z=\bigcup _{1\neq g\in G} K^g$.

The proposition will be reduced to the following elementary statement.

Lemma 4.5 Let $\iota \colon {\mathbb {C}}^2 \to {\mathbb {C}}^2$ be the involution $(x,y)\mapsto (-x, -y)$, let $j,k$ be positive integers. Then the blow-up of $( {\mathbb {C}}^2/\iota )^j\times ( {\mathbb {C}}^2)^k$ along its singular locus is isomorphic to ${( \mathrm {Bl}_0 ( {\mathbb {C}}^2/\iota ) )^j\times ( {\mathbb {C}}^2)^k}$. It is thus smooth, and identified with the quotient $(\mathrm {Bl}_0( {\mathbb {C}}^2)/\iota )^j\times ( {\mathbb {C}}^2)^k$ of $(\mathrm {Bl}_0( {\mathbb {C}}^2))^j\times ( {\mathbb {C}}^2)^k$.

Proof. Given noetherian schemes $T_1$ and $T_2$ and closed subschemes $V_1\subsetneq T_1$ and $V_2\subsetneq T_2$, we have

\[ \mathrm{Bl}_{(V_1\times T_2) \cup ( T_1\times V_2)} (T_1\times T_2) = \mathrm{Bl}_{V_1}(T_1) \times \mathrm{Bl}_{V_2}(T_2). \]

It is easy to see this when $T_1$ and $T_2$ are affine schemes, using the definition of blowing-up via the Proj construction [Reference HartshorneHar77, II, § 7]. The singular locus of $( {\mathbb {C}}^2/\iota )^j\times ( {\mathbb {C}}^2)^k$ is the union of $\operatorname {pr}_i^{-1}(\{0\})$ for $i$ from $1$ to $j$, where $\operatorname {pr}_i\colon ( {\mathbb {C}}^2/\iota )^j\times ( {\mathbb {C}}^2)^k \to ( {\mathbb {C}}^2/\iota )$ is the projection onto the $i$th factor. Via the above observation, the statement is reduced to the case $j=1$, $k=0$, which is the well-known minimal resolution of an isolated $A_1$-singularity of a surface.

Proof of Proposition 4.4 By Proposition 4.2, any point $x\in K/G$ has a neighborhood of the form $( {\mathbb {C}}^2/\iota )^j\times ( {\mathbb {C}}^2)^k$. Therefore, Lemma 4.5 immediately implies that the blow-up $Y_K$ of the singular locus is smooth. It also implies that the natural birational map $\mathrm {Bl}_Z(K)/G\dashrightarrow Y_K$ extends to an isomorphism.

We will now use a criterion due to Fujiki [Reference FujikiFuj83] to show that $Y_K$ is symplectic, i.e. that it admits a nowhere degenerate holomorphic $2$-form.

Proposition 4.6 The manifold $Y_K$ is hyper-Kähler.

Proof. The singular space $K/G$ is a primitively symplectic V-manifold in the sense of Fujiki. According to his [Reference FujikiFuj83, Proposition 2.9], to show that $Y_K$ admits a symplectic form it suffices to check that the following two conditions are satisfied:

  1. (i) for each component $X_j$ of the singular locus of $K/G$, a general point $x\in X_j$ has a neighborhood $U_x = A \times V_x$ in $K/G$, where $A$ is a surface and $V_x=U_x\cap X_j$ is a smooth neighborhood of $x$ in $X_j$;

  2. (ii) the restriction of $p\colon Y_K\to K/G$ to the preimage of $U_x$ is the product

    \[ p'\times \mathrm{id}\colon \tilde{A}\times V_x \to A\times V_x, \]
    where $p'$ is the minimal resolution of $A$.

We have already shown in the course of proof of Propositions 4.2 and 4.4 that these conditions hold, with $A$ a neighborhood of the ordinary node $0\in {\mathbb {C}}^2/\iota$, whose minimal resolution is the blow-up of the singular point.

Hence, $Y_K$ is symplectic. From its description as the quotient $\mathrm {Bl}_Z(K)/G$ we obtain ${h^{2,0}(Y_K)=1}$ (see also § 4.7). Finally, $Y_K$ is simply connected by [Reference FujikiFuj83, Lemma 1.2].

We can now complete the proof of our main result.

Proof of Theorem 1.1 Let $\mathcal {K}\to B$ be a smooth proper family of manifolds of $\mathrm {Kum}^3$ type. Up to a finite étale base-change, $G$ acts fibrewise on $\mathcal {K}$, and the fixed loci of automorphisms in $G$ are smooth families over $B$. By Theorem 2.2, the union $\mathcal {Z}$ of $\mathcal {K}^g$ for $1\neq g\in G$ has $16$ components $\mathcal {Z}_i$, each of which is a smooth family of manifolds of $\mathrm {K}3^{[2]}$ type over $B$. The blow-up of $\mathcal {K}$ along $\mathcal {Z}$ is a smooth family $\widetilde {\mathcal {K}}$ over $B$, and the action of $G$ extends to a fibrewise action on $\widetilde {\mathcal {K}}$. By Propositions 4.4 and 4.6, the quotient $\mathcal {Y}=\widetilde {\mathcal {K}}/G$ is a smooth proper family of hyper-Kähler manifolds over $B$, with fibre over the point $b\in B$ the manifold $Y_{\mathcal {K}_b}$. It thus suffices to find a single $K$ of $\mathrm {Kum}^3$ type such that $Y_K$ is hyper-Kähler of $\mathrm {K}3^{[3]}$ type.

Consider the sixfold $K_J(v_3)$ introduced in § 3, which is constructed from a Beauville–Mukai system on a principally polarized abelian surface $J$. By Theorem 3.1, in this case $Y_{K_J(v_3)}$ is birational to the Hilbert scheme $\operatorname {\mathrm {Km}}(J)^{[3]}$ on the Kummer K3 surface associated to $J$. By [Reference HuybrechtsHuy99, Theorem 4.6], birational hyper-Kähler manifolds are deformation equivalent, and therefore $Y_{K_J(v_3)}$ is of $\mathrm {K}3^{[3]}$ type.

4.7 The associated $\mathrm {K}3$ surface

The following computation is entirely analogous to [Reference FloccariFlo23, § 3]. If $Z$ is a compact Kähler manifold, we let $H^2_{\mathrm {tr}}(Z, {\mathbb {Z}})$ denote the smallest sub-Hodge structure of $H^2(Z, {\mathbb {Z}})$ whose complexification contains $H^{2, 0}(Z)$. Let $K$ be a manifold of $\mathrm {Kum}^3$ type, and denote by $\widetilde {K}$ the blow-up of $K$ along $\bigcup _{1\neq g\in G} K^g$. By Proposition 4.4 we have a commutative diagram

where $p, p'$ are blow-up maps and $q, q'$ are the quotient maps for the action of $G$.

The pull-back gives an isomorphism ${p'}^*\colon H^2_{\mathrm {tr}}(K, {\mathbb {Z}}) \to H^2_{\mathrm {tr}}(\widetilde {K}, {\mathbb {Z}})$. Since $G$ acts trivially on $H^2(K, {\mathbb {Z}})$, the push-forward $q'_*\colon H^2_{\mathrm {tr}}(\widetilde {K}, {\mathbb {Z}})\to H^2_{\mathrm {tr}}(Y_K, {\mathbb {Z}})$ is injective, and ${q'}^*q'_*$ is multiplication by $2^5$ on $H^2_{\mathrm {tr}}(\widetilde {K}, {\mathbb {Z}})$. It follows that

\[ r_*\colon H^2_{\mathrm{tr}}(K, {\mathbb{Z}}) \to H^2_{\mathrm{tr}}(Y_K, {\mathbb{Z}}) \]

becomes an isomorphism of Hodge structures after tensoring with $ {\mathbb {Q}}$. Denote by $q_K$ and $q_{Y_K}$ the Beauville–Bogomolov forms on $H^2(K, {\mathbb {Z}})$ and $H^2(Y_K, {\mathbb {Z}})$, respectively.

Lemma 4.8 For any $x\in H_{\mathrm {tr}}^2(K, {\mathbb {Z}})$ we have $q_{Y_K} (r_*(x), r_*(x)) = 2^9 q_K(x, x)$. Therefore,

\[ \frac{1}{16}r_*\colon H^2_{\mathrm{tr}}(K, {\mathbb{Q}})(2) \xrightarrow{ \ \sim\ } H^2_{\mathrm{tr}}(Y_K, {\mathbb{Q}}), \]

is a rational Hodge isometry, where $(2)$ indicates that the form is multiplied by $2$.

Proof. Let $c_K$ and $c_Y$ be the Fujiki constants of $K$ and $Y$, respectively [Reference FujikiFuj87]. This means that we have $\int _K x^6 = c_K q_K(x,x)^3$ for any $x\in H^2(K, {\mathbb {Z}})$, and similarly for $Y_K$. Let $x\in H_{\mathrm {tr}}^2(K, {\mathbb {Z}})$; since $x$ is $G$-invariant, we have $r^*(r_* x) = 2^5 x$. We compute

\begin{align*} q_{Y_K}(r_*x, r_*x)^3 & = \frac{1}{c_{Y_K}}\int_{Y_K} (r_*x)^6 \\ &= \frac{1}{2^5 c_{Y_K}} \int_K (r^*r_* x)^6\\ & = \frac{1}{2^5 c_{Y_K}} \int_K (2^5 x)^6 \\ & =\frac{2^{25}c_K }{c_{Y_K}} q_K(x,x)^3. \end{align*}

By [Reference RapagnettaRap08], we have $c_{Y_K}= 15$ and $c_K=60$. Hence, $q_{Y_K}(r_*x, r_*x)=2^{9} q_K(x,x)$, so $r_*/2^4$ multiplies the form by $2$ and yields the claimed rational Hodge isometry.

We now prove that every projective $K$ of $\mathrm {Kum}^3$ type has a naturally associated K3 surface.

Proof of Theorem 1.2 Let $K$ be a projective variety of $\mathrm {Kum}^3$ type and let $Y_K$ be the crepant resolution of $K/G$. By the above lemma, the transcendental lattice $H^2_{\mathrm {tr}}(Y_K, {\mathbb {Z}})$ is an even lattice of signature $(2, k)$, and rank at most $6$. By [Reference MorrisonMor84, Corollary 2.10], there exists a $\mathrm {K}3$ surface $S_K$ such that $H^2_{\mathrm {tr}}(S_K, {\mathbb {Z}})$ is Hodge isometric to $H^2_{\mathrm {tr}}(Y_K, {\mathbb {Z}})$. A criterion independently due to Mongardi and Wandel [Reference Mongardi and WandelMW15] and Addington [Reference AddingtonAdd16, Proposition 4] ensures that $Y_K$ is birational to $M_{S_{K},H}(v)$, for some primitive Mukai vector $v$ and a $v$-generic polarization $H$ on $S_K$. The surface $S_K$ is uniquely determined up to isomorphism, because two $\mathrm {K}3$ surfaces of Picard rank at least $12$ with Hodge isometric transcendental lattices are isomorphic, see [Reference HuybrechtsHuy16, Chapter 16, Corollary 3.8]

Remark 4.9 The K3 surfaces $S_K$ come in countably many $4$-dimensional families. A projective $K_0$ of $\mathrm {Kum}^3$ type with Picard rank $1$ gives such a family, consisting of the K3 surfaces $S$ such that $H^2_{\mathrm {tr}}(S, {\mathbb {Z}})$ is a sublattice of $H^2_{\mathrm {tr}}(S_{K_0}, {\mathbb {Z}})$.

Up to isogeny, the K3 surfaces obtained are easily characterized as follows. Recall that $\Lambda _{\mathrm {Kum}^3}=U^{\oplus 3}\oplus \langle -8\rangle$ is the lattice $H^2(K, {\mathbb {Z}})$ for $K$ of $\mathrm {Kum}^3$ type.

Lemma 4.10 Let $S$ be a projective $\mathrm {K}3$ surface. The following are equivalent:

  • there exists an isometric embedding of $H^2_{\mathrm {tr}}(S, {\mathbb {Q}})$ into $\Lambda _{\mathrm {Kum}^3}(2)\otimes _{ {\mathbb {Z}}} {\mathbb {Q}}$;

  • there exist a projective variety $K$ of $\mathrm {Kum}^3$ type with associated $\mathrm {K}3$ surface $S_K$ and a rational Hodge isometry $H^2_{\mathrm {tr}}(S, {\mathbb {Q}}) \xrightarrow {\ \sim \ } H^2_{\mathrm {tr}}(S_K, {\mathbb {Q}})$.

Proof. The second assertion implies the first thanks to Lemma 4.8. Conversely, let $\Phi \colon H^2_{\mathrm {tr}}(S, {\mathbb {Q}}) \hookrightarrow \Lambda _{\mathrm {Kum}^3}(2)\otimes _{ {\mathbb {Z}}} {\mathbb {Q}}$ be an isometric embedding. Choose a primitive sublattice $T\subset \Lambda _{\mathrm {Kum}^3}$ such that $T(2)\otimes _{ {\mathbb {Z}}} {\mathbb {Q}}$ coincides with the image of $\Phi$. Equip $T$ with the Hodge structure induced by that on $H^2_{\mathrm {tr}}(S, {\mathbb {Q}})$ via $\Phi$. By the surjectivity of the period map [Reference HuybrechtsHuy99], there exists a manifold $K$ of $\mathrm {Kum}^3$ type such that $H^2_{\mathrm {tr}}(K, {\mathbb {Z}})$ is Hodge isometric to $T$, and Lemma 4.8 gives a Hodge isometry $H^2_{\mathrm {tr}}(S, {\mathbb {Q}})\xrightarrow {\ \sim \ } H^2_{\mathrm {tr}}(S_K, {\mathbb {Q}})$. Since the signature of $T$ is necessarily $(2, k)$ with $k\leqslant 4$, Huybrechts’ projectivity criterion [Reference HuybrechtsHuy99] implies that $K$ is projective.

5. Applications to the Hodge conjecture

In this section we prove Theorems 1.3 and 1.4. Throughout, all cohomology groups are taken with rational coefficients, which are thus suppressed from the notation.

We start with a simple observation. Let $X, Y$ be smooth projective varieties and let $\phi \colon H^2(X)\to H^2(Y)$ be a morphism of Hodge structures.

Lemma 5.1 The morphism $\phi$ is induced by an algebraic correspondence if and only if its restriction to $H^{2}_{\mathrm {tr}}(X)$ is so.

Proof. There is a decomposition $H^2(X)=H^2_{\mathrm {tr}}(X) \oplus H^2_{\mathrm {alg}}(X)$, where $H^2_{\mathrm {alg}}(X)$ is spanned by cycle classes of divisors. Similarly, $H^2(Y)= H^2_{\mathrm {tr}}(Y) \oplus H^2_{\mathrm {alg}}(Y)$. Any morphism $H^2_{\mathrm {tr}}(X)\to H^2_{\mathrm {alg}}(Y)$ or $H^2_{\mathrm {alg}}(X)\to H^2_{\mathrm {tr}}(Y)$ of Hodge structures is trivial. Hence, $\phi$ gives a Hodge class

\[ \phi_{\mathrm{tr}} \oplus \phi_{\mathrm{alg}} \ \in\ \bigl(H^2_{\mathrm{tr}}(X)^{\vee} \otimes H^2_{\mathrm{tr}}(Y) \bigr) \oplus \bigl(H^2_{\mathrm{alg}}(X)^{\vee} \otimes H^2_{\mathrm{alg}}(Y)\bigr). \]

The lemma follows because the second summand consists of algebraic classes.

Proof of Theorem 1.4 Let $K$ and $K'$ be projective varieties of $\mathrm {Kum}^3$ type and assume that $f\colon H^2(K) \xrightarrow { \ \sim \ } H^2(K')$ is a rational Hodge isometry. By the above lemma, it suffices to show that the component $f_{\mathrm {tr}}\colon H_{\mathrm {tr}}^2(K) \xrightarrow {\ \sim \ } H_{\mathrm {tr}}^2(K')$ is algebraic.

Let $Y$, $Y'$ be the varieties of $\mathrm {K}3^{[3]}$ type given by Theorem 1.1 applied to $K$ and $K'$, respectively, and let $r\colon K\dashrightarrow Y$ and $r'\colon K'\dashrightarrow Y'$ be the corresponding rational maps. They induce isomorphisms of Hodge structures $r_*\colon H^2_{\mathrm {tr}}(K) \xrightarrow {\ \sim \ } H_{\mathrm {tr}}^2(Y)$ and $r'_*\colon H^2_{\mathrm {tr}}(K') \xrightarrow {\ \sim \ } H_{\mathrm {tr}}^2(Y')$. Let $\bar {f}_{\mathrm {tr}}$ be defined as

\[ \bar{f}_{\mathrm{tr}} := r'_* \circ f_{\mathrm{tr}}\circ (r_*)^{-1} \colon H^2_{\mathrm{tr}}(Y) \to H^2_{\mathrm{tr}}(Y'). \]

The inverse of $r_*$ is $r^*/2^5$, and similarly for $r'$. It follows that $f_{\mathrm {tr}}$ is algebraic if and only if $\bar {f}_{\mathrm {tr}}$ is so. By Lemma 4.8, the map $\bar {f}_{\mathrm {tr}}$ is a rational Hodge isometry and, hence, it is algebraic by Markman's theorem in [Reference MarkmanMar22].

5.2 The Kuga–Satake correspondence

The Kuga–Satake construction associates an abelian variety to any polarized Hodge structure of $\mathrm {K}3$ type. We briefly recall this construction, referring to [Reference van GeemenvGe00] and [Reference HuybrechtsHuy16, Chapter 4] for more details.

Let $(V, q)$ be an effective polarized $ {\mathbb {Q}}$-Hodge structure of weight $2$ with $h^{2,0}(V)=1$. The Clifford algebra $C(V)$ is defined as the quotient

\[ C(V):= \bigoplus_{k\geqslant 0} V^{\otimes k}/\langle v\otimes v - q(v,v) 1 \rangle_{v\in V}. \]

As a $ {\mathbb {Q}}$-vector space, $C(V)\cong \bigwedge ^{\bullet }(V)$; hence, $\dim (C(V))=2^{\mathrm {rk}(V)}$. The Clifford algebra is $ {\mathbb {Z}}/2 {\mathbb {Z}}$-graded, $C(V)=C^+(V)\oplus C^-(V)$. In [Reference DeligneDel71], Deligne shows that the Hodge structure of $V$ induces a Hodge structure of weight $1$ on $C^+(V)$ of type $(1,0), (0,1)$, which admits a polarization and, hence, defines an abelian variety $\mathrm {KS}(V)$ up to isogeny. This $2^{\mathrm {rk}(V)-2}$-dimensional abelian variety is called the Kuga–Satake variety of $V$. Upon fixing some $v_0\in V$, the action of $V$ on $C(V)$ via left multiplication induces an embedding of weight $0$ rational Hodge structures

\[ V(1) \hookrightarrow \operatorname{End}(C^+(V)) = H^1(\mathrm{KS}(V))^{\vee}\otimes H^1(\mathrm{KS}(V)), \]

which maps $v$ to the endomorphism $w\mapsto vwv_0$.

Remark 5.3 If $V'\subset V$ is a sub-Hodge structure such that ${V'}^{\bot }$ consists of Hodge classes, then $\mathrm {KS}(V)$ is isogenous to a power of $\mathrm {KS}(V')$. Replacing the form $q$ with a non-zero rational multiple results in isogenous Kuga–Satake varieties. See [Reference HuybrechtsHuy16, Chapter 4, Example 2.4 and Proof of Proposition 3.3]

Let $X$ be a projective hyper-Kähler variety of dimension $2n$. The Kuga–Satake variety $\mathrm {KS}(X)$ of $X$ is the abelian variety obtained from $(H^2_{\mathrm {tr}}(X), q_X)$ via the Kuga–Satake construction, where $q_X$ is the restriction of the Beauville–Bogomolov form. Identifying $H^1(\mathrm {KS}(X))$ with its dual, there exists an embedding of Hodge structures of $H^2_{\mathrm {tr}}(X)$ into $H^2(\mathrm {KS}(X)\times \mathrm {KS}(X) )$. According to the Hodge conjecture this embedding should be induced by an algebraic cycle.

Conjecture 5.4 (Kuga–Satake Hodge conjecture)

Let $X$ be a projective hyper-Kähler variety. There exists an algebraic cycle $\zeta$ on $X\times \mathrm {KS}(X) \times \mathrm {KS}(X)$ such that the associated correspondence induces an embedding of Hodge structures

\[ \zeta_*\colon H^2_{\mathrm{tr}}(X) \hookrightarrow H^2(\mathrm{KS}(X)\times \mathrm{KS}(X) ). \]

Remark 5.5 For a K3 surface $X$, the above form of the conjecture is equivalent to [Reference van GeemenvGe00, § 10.2]. To see this, let $\mathsf {Mot}$ be the category of Grothendieck motives over $ {\mathbb {C}}$, in which morphisms are given by algebraic cycles modulo homological equivalence, see [Reference AndréAnd04]. The motive $\operatorname {\mathfrak {h}}(X)\in \mathsf {Mot}$ of $X$ decomposes as the sum of its transcendental part $\operatorname {\mathfrak {h}}_{\mathrm {tr}}(X)$ and some motives of Hodge–Tate type. Since the standard conjectures hold for $X$ and $\mathrm {KS}(X)$, the tensor subcategory of $\mathsf {Mot}$ generated by their motives is abelian and semisimple [Reference ArapuraAra06, Theorem 4.1]. Therefore, Conjecture 5.4 and the formulation of van Geemen [Reference van GeemenvGe00, § 10.2] are both equivalent to $\operatorname {\mathfrak {h}}^2_{\mathrm {tr}}(X)$ being a direct summand of $\operatorname {\mathfrak {h}}^2(\mathrm {KS}(X)\times \mathrm {KS}(X))$ in the category $\mathsf {Mot}$.

We will use the following easy lemma.

Lemma 5.6 Let $X$ and $Z$ be projective hyper-Kähler varieties. Assume that there exists an algebraic cycle $\gamma$ on $Z\times X$ which induces a rational Hodge isometry

\[ \gamma_*\colon H^2_{\mathrm{tr}}(Z)\xrightarrow{\ \sim\ } H^2_{\mathrm{tr}}(X)(k), \]

for some non-zero $k\in {\mathbb {Q}}$. Then, if Conjecture 5.4 holds for $X$, it holds for $Z$ as well.

Proof. By Remark 5.3 there exists an isogeny $\phi \colon \mathrm {KS}(X)\to \mathrm {KS}(Z)$ of Kuga–Satake varieties. It induces an isomorphism

\[ \phi_*\colon H^2(\mathrm{KS}(X)\times \mathrm{KS}(X)) \xrightarrow{\ \sim\ } H^2(\mathrm{KS}(Z)\times \mathrm{KS}(Z)) \]

of rational Hodge structures. If Conjecture 5.4 holds for $X$, there exists an algebraic cycle $\zeta$ on $X\times \mathrm {KS}(X) \times \mathrm {KS}(X)$ giving an embedding of Hodge structures

\[ \zeta_*\colon H^2_{\mathrm{tr}}(X) \hookrightarrow H^2(\mathrm{KS}(X)\times \mathrm{KS}(X)). \]

It follows that the embedding of Hodge structure given by the composition

\[ \phi_* \circ \zeta_* \circ \gamma_* \colon H^2_{\mathrm{tr}}(Z) \hookrightarrow H^2(\mathrm{KS}(Z)\times \mathrm{KS}(Z)) \]

is induced by an algebraic cycle on $Z\times \mathrm {KS}(Z)\times \mathrm {KS}(Z)$.

For convenience of the reader we restate Theorem 1.3.

Theorem 5.7 Let $S$ be a projective $\mathrm {K}3$ surface such that there exists an isometric embedding of $H^2_{\mathrm {tr}}(S)$ into $\Lambda _{\mathrm {Kum}^3}(2)\otimes _{ {\mathbb {Z}}} {\mathbb {Q}}$. Then Conjecture 5.4 holds for $S$.

Proof. By Lemma 4.10, there exists a projective variety $K$ of $\mathrm {Kum}^3$ type with associated K3 surface $S_K$ and a rational Hodge isometry $t_0\colon H^2_{\mathrm {tr}}(S)\xrightarrow {\ \sim \ } H^2_{\mathrm {tr}}(S_K)$. Denote by $Y_K$ the crepant resolution of $K/G$ given by Theorem 1.1 and let $v, H$ be given by Theorem 1.2, so that $Y_K$ is birational to the moduli space $M_{S_K,H}(v)$. Denote by $r\colon K\dashrightarrow Y_K$ the natural rational map of degree $2^5$.

By [Reference BuskinBus19, Reference HuybrechtsHuy19], the isometry $t_0$ is algebraic. Next, by [Reference MukaiMuk87], there exists a quasi-tautological sheaf $U$ over $S_K \times M_{S_K,H}(v)$, which means that there exists an integer $\rho$ such that for any $F\in M_{S_K,H}(v)$ the restriction of $U$ to $S_K\times \{F\}$ is $F^{\oplus \rho }$. Consider the algebraic class

\[ \gamma:= \frac{1}{\rho}\mathrm{ch}(U)\cdot \mathrm{pr}^* \sqrt{\mathrm{td}_{S_K}}\in H^{\bullet}(S_K\times M_{S_K, H}(v)), \]

where $\mathrm {pr}\colon S_K\times M_{S_K, H}(v) \to S_K$ is the projection. By [Reference O'GradyO'Gr97], its Künneth component $\gamma _3\in H^6 (S_K\times M_{S_K, H}(v))$ induces a Hodge isometry

\[ t_1\colon H^2_{\mathrm{tr}}(S_K) \xrightarrow{\ \sim\ } H^2_{\mathrm{tr}}(M_{S_K, H}(v)). \]

By [Reference Charles and MarkmanCM13], the standard conjectures holds for $M_{S_K, H}(v)$ (this also follows from [Reference BüllesBül20] via the arguments of [Reference ArapuraAra06]). Hence, all Künneth components of $\gamma$ are algebraic. Now let $f\colon M_{S_K, H}(v)\dashrightarrow Y_K$ be a birational map. Then the push-forward $f_*\colon H^2_{\mathrm {tr}}(M_{S_K, H}(v))\xrightarrow {\ \sim \ } H^2_{\mathrm {tr}}(Y_K)$ is a Hodge isometry. By Lemma 4.8, a multiple of the composition $r^* \circ f_*$ gives a Hodge isometry $t_2\colon H^2_{\mathrm {tr}}(M_{S_K, H}(v)) \xrightarrow {\ \sim \ } H^2_{\mathrm {tr}}(K)(2)$.

It follows that the Hodge isometry

\[ t_2\circ t_1\circ t_0 \colon H^2_{\mathrm{tr}}(S)\xrightarrow{\ \sim\ } H^2_{\mathrm{tr}}(K)(2) \]

is induced by an algebraic cycle on $S\times K$. The Kuga–Satake Hodge conjecture holds for $K$ by [Reference VoisinVoi22]. By Lemma 5.6, Conjecture 5.4 holds for the K3 surface $S$ as well.

For the general projective variety $K$ of $\mathrm {Kum}^3$ type, $H^2_{\mathrm {tr}}(K, {\mathbb {Q}})$ is a rank-$6$ Hodge structure with Hodge numbers $(1, 4, 1)$ and the explicit knowledge of the quadratic form allows one to show that the Kuga–Satake variety is $A^4$, where $A$ is an abelian fourfold of Weil type (cf. [Reference van GeemenvGe00, Theorem 9.2]). O'Grady moreover shows in [Reference O'GradyO'Gr21] that $H^1(A, {\mathbb {Q}})\cong H^3(K, {\mathbb {Q}})$. This is a crucial ingredient in the subsequent work of Markman [Reference MarkmanMar23] and Voisin [Reference VoisinVoi22]. O'Grady's theorem further allows us to apply a result of Varesco [Reference VarescoVar22] to prove the Hodge conjecture for all powers of the K3 surfaces appearing in Theorem 1.3.

Corollary 5.8 Let $S$ be a projective $\mathrm {K}3$ surface such that there exists an isometric embedding of $H^2_{\mathrm {tr}}(S)$ into $\Lambda _{\mathrm {Kum}^3}(2)\otimes _{ {\mathbb {Z}}} {\mathbb {Q}}$. Then the Hodge conjecture holds for all powers of $S$.

Proof. By Remark 4.9, there exists a $4$-dimensional family $\mathcal {S}\to B$ of projective K3 surfaces of general Picard rank $16$ such that: for each $b\in B$ the fibre $\mathcal {S}_b$ is of the form $S_{K_b}$ for some $K_b$ of $\mathrm {Kum}^3$ type and for some $0\in B$ there exists a rational Hodge isometry $H^2_{\mathrm {tr}}(S) \xrightarrow {\ \sim \ } H^2_{\mathrm {tr}}(\mathcal {S}_0)$. By [Reference HuybrechtsHuy19], it is sufficient to prove the Hodge conjecture for all powers of $\mathcal {S}_0$.

To this end, it will be enough to check that $\mathcal {S}\to B$ satisfies the two assumptions of [Reference VarescoVar22, Theorem 0.2]. As explained in [Reference VarescoVar22, Theorem 4.1], the first of these assumptions is that the Kuga–Satake variety of a general K3 surface in the family is isogenous to $A^4$ for an abelian fourfold $A$ of Weil type with trivial discriminant. This holds by O'Grady's theorem in [Reference O'GradyO'Gr21] as already mentioned. The other assumption is that the Kuga–Satake Hodge conjecture holds for the surfaces $\mathcal {S}_b$ for all $b\in B$, which is the content of Theorem 5.7.

Remark 5.9 The conclusions of Theorem 5.7 and Corollary 5.8 were known for the two $4$-dimensional families of K3 surfaces studied in [Reference ParanjapePar88] and [Reference Ingalls, Logan and PatashnickILP22]. The general fibre has transcendental lattice $U^{\oplus 2}\oplus \langle -2\rangle ^{\oplus 2}$ in the first case and $U^{\oplus 2} \oplus \langle -6\rangle \oplus \langle -2\rangle$ in the second. The reader may check that the K3 surfaces studied in [Reference ParanjapePar88] satisfy the assumption of our Theorem 5.7, while those appearing in [Reference Ingalls, Logan and PatashnickILP22] do not.

Acknowledgements

It is a pleasure to thank Ben Moonen, Kieran O'Grady, Claudio Pedrini and Stefan Schreieder for their interest in this project. I am grateful to Lie Fu and Giovanni Mongardi for their comments on a first version of this article, and to Grégoire Menet for useful discussions. I thank Mauro Varesco for suggesting to me Corollary 5.8. I thank Giacomo Mezzedimi and Domenico Valloni for many conversations related to these topics, and the anonymous referee for his/her comments.

Conflicts of Interest

None.

References

Addington, N., On two rationality conjectures for cubic fourfolds, Math. Res. Lett. 23 (2016), 113.CrossRefGoogle Scholar
André, Y., Une introduction aux motifs (motifs purs, motifs mixtes, périodes), Panoramas et Synthèses, vol. 17 (Société Mathématique de France, Paris, 2004).Google Scholar
Arapura, D., Motivation for Hodge cycles, Adv. Math. 207 (2006), 762781.CrossRefGoogle Scholar
Beauville, A., Variétés Kählériennes dont la première classe de Chern est nulle, J. Differential Geom. 18 (1983), 755782.CrossRefGoogle Scholar
Beauville, A., Systémes hamiltoniens complétement intégrables associés aux surfaces K3, in Problems in the theory of surfaces and their classification (Cortona, 1988), Symposia Mathematica, vol. XXXII (Academic Press, London, 1991), 2531.Google Scholar
Beauville, A., Counting rational curves on K3 surfaces, Duke Math. J. 97 (1999), 99108.CrossRefGoogle Scholar
Boissière, S., Nieper-Wißkirchen, M. and Sarti, A., Higher dimensional Enriques varieties and automorphisms of generalized Kummer varieties, J. Math. Pures Appl. 95 (2011), 553563.CrossRefGoogle Scholar
Bülles, T.-H., Motives of moduli spaces on K3 surfaces and of special cubic fourfolds, Manuscripta Math. 161 (2020), 109124.CrossRefGoogle Scholar
Buskin, N., Every rational Hodge isometry between two K3 surfaces is algebraic, J. Reine Angew. Math. 2019 (2019), 127150.CrossRefGoogle Scholar
Charles, F. and Markman, E., The standard conjectures for holomorphic symplectic varieties deformation equivalent to Hilbert schemes of K3 surfaces, Compos. Math. 149 (2013), 481494.CrossRefGoogle Scholar
Deligne, P., La conjecture de Weil pour les surfaces K3, Invent. Math. 15 (1971), 206226.CrossRefGoogle Scholar
Floccari, S., On the motive of O'Grady's six dimensional hyper-Kähler varieties, Épijournal Géom. Algébrique 7 (2023), article no. 4.Google Scholar
Fujiki, A., A theorem on bimeromorphic maps of Kähler manifolds and its applications, Publ. Res. Inst. Math. Sci. 17 (1981), 735754.CrossRefGoogle Scholar
Fujiki, A., On primitively symplectic compact Kähler V-manifolds of dimension four, in Classification of algebraic and analyitic manifolds, Progress in Mathematics, ed. K. Ueno, vol. 39 (Birkhäuser, Basle, 1983), 71250.Google Scholar
Fujiki, A., On the de Rham cohomology group of a compact Kähler symplectic manifold, in Algebraic Geometry, Sendai, 1985, ed. T. Oda (Mathematical Society of Japan, 1987), 105165.CrossRefGoogle Scholar
Gonzalez-Dorrego, M. d. R., $(16, 6)$ Configurations and geometry of Kummer surfaces in $\mathbb {P}^3$, Mem. Amer. Math. Soc. 107(512) (1994).Google Scholar
Hartshorne, R., Algebraic geometry, Graduate Texts in Mathematics, vol. 52 (Springer, 1977).CrossRefGoogle Scholar
Hassett, B. and Tschinkel, Y., Hodge theory and Lagrangian planes on generalized Kummer fourfolds, Mosc. Math. J. 13 (2013), 3356.CrossRefGoogle Scholar
Huybrechts, D., Compact hyper-Kähler manifolds: basic results, Invent. Math. 135 (1999), 63113.CrossRefGoogle Scholar
Huybrechts, D., Lectures on K3 surfaces, vol. 158 (Cambridge University Press, 2016).CrossRefGoogle Scholar
Huybrechts, D., Motives of isogenous K3 surfaces, Comment. Math. Helv. 94 (2019), 445458.CrossRefGoogle Scholar
Huybrechts, D. and Lehn, M., The geometry of moduli spaces of sheaves, second edition (Cambridge University Press, 2010).CrossRefGoogle Scholar
Ingalls, C., Logan, A. and Patashnick, O., Explicit coverings of families of elliptic surfaces by squares of curves, Math. Z. 302 (2022), 11911238.CrossRefGoogle Scholar
Kamenova, L., Mongardi, G. and Oblomkov, A., Symplectic involutions of K$3^{n}$ type and Kummer $n$ type manifolds, Bull. Lond. Math. Soc. 54 (2022), 894909.CrossRefGoogle Scholar
Kim, Y.-J., The dual Lagrangian fibration of known hyper-Kähler manifolds, Preprint (2021), arXiv:2109.03987.Google Scholar
Le Potier, J., Systèmes cohérents et structures de niveau, Astérisque 214 (1993).Google Scholar
Lehn, M. and Sorger, C., La singularité de O'Grady, J. Algebraic Geom. 15 (2006), 753770.CrossRefGoogle Scholar
Markman, E., The Beauville–Bogomolov class as a characteristic class, J. Algebraic Geom. 29 (2020), 199245.CrossRefGoogle Scholar
Markman, E., Rational Hodge isometries of hyper-Kahler varieties of K3[n]-type are algebraic, Preprint (2022), arXiv:2204.00516.Google Scholar
Markman, E., The monodromy of generalized Kummer varieties and algebraic cycles on their intermediate Jacobians, J. Eur. Math. Soc. (JEMS) 25 (2023), 231321.CrossRefGoogle Scholar
Matsushita, D., On fibre space structures of a projective irreducible symplectic manifold, Topology 38 (1999), 7983.CrossRefGoogle Scholar
Menet, G., Global Torelli theorem for irreducible symplectic orbifolds, J. Math. Pures Appl. 137 (2020), 213237.CrossRefGoogle Scholar
Mongardi, G., Rapagnetta, A. and Saccà, G., The Hodge diamond of O'Grady's six-dimensional example, Compos. Math. 154 (2018), 9841013.CrossRefGoogle Scholar
Mongardi, G. and Wandel, M., Induced automorphisms on irreducible symplectic manifolds, J. Lond. Math. Soc. 92 (2015), 123143.CrossRefGoogle Scholar
Morrison, D., On K3 surfaces with large Picard number, Invent. Math. 75 (1984), 105122.CrossRefGoogle Scholar
Morrison, D., The Kuga–Satake variety of an abelian surface, J. Algebra 92 (1985), 454476.CrossRefGoogle Scholar
Mukai, S., Symplectic structure of the moduli space of sheaves on an abelian or $K3$ surface, Invent. Math. 77 (1984), 101116.CrossRefGoogle Scholar
Mukai, S., On the moduli space of bundles on K3 surfaces I, in Vector bundles on algebraic varieties (Bombay, 1984), Tata Institute of Fundamental Research Studies in Mathematics, vol. 11 (Clarendon Press, 1987), 341413.Google Scholar
Mumford, D., Prym varieties I, in Contributions to analysis, eds L. V. Ahlfors, I. Kra, B. Maskitt and L. Nirenberg (Academic Press, 1974), 325350.CrossRefGoogle Scholar
O'Grady, K. G., The weight-two Hodge structure of moduli spaces of sheaves on a $K3$ surface, J. Algebraic Geom. 6 (1997), 599644.Google Scholar
O'Grady, K. G., A new six-dimensional irreducible symplectic variety, J. Algebraic Geom. 12 (2003), 435505.CrossRefGoogle Scholar
O'Grady, K. G., Compact tori associated to hyperkähler manifolds of Kummer type, Int. Math. Res. Not. IMRN 2021 (2021), 1235612419.CrossRefGoogle Scholar
O'Grady, K. G., Theta groups and projective models of hyperkähler varieties, Preprint (2022), arXiv:2204.09582.Google Scholar
Oguiso, K., No cohomologically trivial nontrivial automorphism of generalized Kummer manifolds, Nagoya Math. J. 239 (2020), 110122.CrossRefGoogle Scholar
Paranjape, K., Abelian varieties associated to certain K3 surfaces, Compos. Math. 68 (1988), 1122.Google Scholar
Rapagnetta, A., Topological invariants of O'Grady's six dimensional irreducible symplectic variety, Math. Z. 256 (2007), 134.CrossRefGoogle Scholar
Rapagnetta, A., On the Beauville form of the known irreducible symplectic varieties, Math. Ann. 340 (2008), 7795.CrossRefGoogle Scholar
van Geemen, B., Kuga–Satake varieties and the Hodge conjecture, in The arithmetic and geometry of algebraic cycles, eds B. B. Gordon, J. D. Lewis, S. Müller-Stach, S. Saito and N. Yui (Springer, Netherlands, 2000), 5182.CrossRefGoogle Scholar
Varesco, M., The Hodge conjecture for powers of K3 surfaces of Picard number 16, Preprint (2022), arXiv:2203.09778.Google Scholar
Verbitsky, M., Hyperholomorphic bundles over a hyperkähler manifold, J. Algebraic Geom. 5 (1996), 633669.Google Scholar
Verra, A., The fibre of the Prym map in genus three, Math. Ann. 276 (1987), 433448.CrossRefGoogle Scholar
Voisin, C., Footnotes to papers of O'Grady and Markman, Math. Z. 300 (2022), 34053416.CrossRefGoogle Scholar
Wieneck, B., Monodromy invariants and polarization types of generalized Kummer fibrations, Math. Z. 290 (2018), 347378.CrossRefGoogle Scholar
Yoshioka, K., Some examples of Mukai's reflections on K3 surfaces, J. Reine Angew. Math. 1999 (1999), 97123.CrossRefGoogle Scholar
Yoshioka, K., Moduli spaces of stable sheaves on abelian surfaces, Math. Ann. 321 (2001), 817884.CrossRefGoogle Scholar