Perverse–Hodge complexes for Lagrangian fibrations
-
-
scAbstract. Perverse–Hodge complexes are objects in the derived category of coherent sheaves obtained from Hodge modules associated with Saito’s decomposition theorem. We study perverse–Hodge complexes for Lagrangian fibrations and propose a symmetry between them. This conjectural symmetry categorifies the authors’ “perverse = Hodge” identity and specializes to Matsushita’s theorem on the higher direct images of the structure sheaf. We verify our conjecture in several cases by making connections with variations of Hodge structures, Hilbert schemes, and Looijenga–Lunts–Verbitsky Lie algebras.
scKeywords. Holomorphic symplectic varieties, Lagrangian fibrations, perverse sheaves, Hodge modules
sc2020 Mathematics Subject Classification. 14J42, 14D06, 14F10
-
-
cJune 2, 2023Received by the Editors on May 27, 2022.
Accepted on June 18, 2023.
Department of Mathematics, Yale University, New Haven, CT 06511, USA
sce-mail: [email protected]
BICMR, Peking University, Beijing 100871, China
sce-mail: [email protected]
J. S. is supported by the NSF grant DMS-2134315. Q. Y. is supported by the NSFC grants 11831013 and 11890661.
© by the author(s) This work is licensed under http://creativecommons.org/licenses/by-sa/4.0/
1. Introduction
1.1. Perverse–Hodge symmetry
For a compact irreducible symplectic variety(1)(1)(1)We say that is irreducible symplectic if it is a compact Kähler manifold such that is spanned by a nowhere degenerate -form. of dimension with a Lagrangian fibration , the decomposition theorem, cf. [BBD82],
(1.1) |
provides important invariants for the topology of . A perverse–Hodge symmetry was proven in [SY22], connecting the cohomology of the perverse sheaves with the Hodge numbers of .
Theorem 1.1 (cf. [SY22]).
We have
(1.2) |
Here stands for , and denotes the Hodge number.
The identity (1.2) governs the cohomology of the Lagrangian base, the invariant cohomology of a nonsingular fiber of , and the Gokapumar–Vafa theory of surfaces; we refer to [SY22, FSY22, HLS+21, HM22] for more discussions on Theorem 1.1 and its applications.
The purpose of this paper is to explore and propose a categorification of the perverse–Hodge symmetry. It suggests that Theorem 1.1 should conceptually be viewed as a cohomological shadow of a sheaf-theoretic symmetry for Lagrangian fibrations with possibly noncompact ambient spaces . It is a mysterious phenomenon since all existing proofs of (1.2), cf. [SY22, HLS+21, HM22], rely heavily on the global cohomological properties of compact irreducible symplectic manifolds, and they do not “explain” why such a categorification should exist. Our formulation uses perverse–Hodge complexes constructed from Hodge modules.
1.2. Perverse–Hodge complexes
Let be a nonsingular quasi-projective symplectic variety of dimension . Here is a closed nowhere degenerate holomorphic -form on . Let be a proper Lagrangian fibration onto a nonsingular base ; i.e., the restriction of the symplectic form to regular part of a fiber vanishes. Interesting examples of include Lagrangian fibrations of compact irreducible symplectic varieties, cf. [Bea91], and Hitchin’s integrable systems, cf. [Hit87a, Hit87b].
By Saito’s theory [Sai89, Sai90], the decomposition theorem (1.1) can be upgraded to an identity in the bounded derived category of Hodge modules. Let be the trivial Hodge module, i.e., the pure Hodge module associated with the shifted trivial local system . We have
(1.3) |
The Hodge module consists of a regular holonomic (left-)-module equipped with a good filtration ; it corresponds to the perverse sheaf under the Riemann–Hilbert correspondence. The increasing filtration induces an increasing filtration on the de Rham complex
The associated graded pieces are natural objects in the bounded derived category of coherent sheaves on ,
Up to re-indexing and shifting, we define
We call the perverse–Hodge complexes associated with the Lagrangian fibration ; here is the perverse degree, and is the Hodge degree. The object is nontrivial only if .
Our main proposal is the following conjectural symmetry between perverse–Hodge complexes.
Conjecture 1.2.
Let be a Lagrangian fibration. We have
1.3. Main results
We provide evidence for Conjecture 1.2 and verify it in several cases.
1.3.1. Smooth morphisms
Our first theorem verifies Conjecture 1.2 when is smooth. In fact, we obtain a stronger result in this case.
Theorem 1.3.
Assume that is smooth. The symplectic form on together with a polarization induces an isomorphism
at the level of complexes.
Theorem 1.3 is essentially a reformulation of a result of Donagi and Markman [DM96] on the polarized variation of Hodge structures associated with the family. Both complexes and have the same length, and a term-by-term isomorphism is constructed between them. The mystery of Conjecture 1.2 is an “extension” of this isomorphism to the singular fibers. As we see from Section 3.4, in general such an extension is complicated, and the derived category is essential for the formulation.
1.3.2. Hilbert schemes
Next, we consider the Lagrangian fibration
(1.4) |
induced by an elliptic fibration of a symplectic surface . Typical examples include:
-
(1)
is an elliptic , and is a Lagrangian fibration of the compact irreducible symplectic variety ; and
- (2)
The decomposition theorem associated with (1.4) has many supports besides the full base . In particular, the isomorphism of Conjecture 1.2 in this case is not merely an extension of the isomorphism of Theorem 1.3 for variations of Hodge structures. Semisimple objects of the decomposition theorem (1.3) supported on the “boundary” of contribute nontrivially.
1.3.3. Global cohomology
Lastly, we consider Lagrangian fibrations associated with compact irreducible symplectic varieties.(2)(2)(2)As the base is assumed to be nonsingular, by a result of Hwang [Hwa08] we know that is a projective space. However, this fact will not be used in this paper. Since is projective, the (hyper-)cohomology groups of the perverse–Hodge complexes are finite-dimensional.
The following theorem shows that in this case Conjecture 1.2 holds cohomologically.
Theorem 1.5.
Let be a Lagrangian fibration with a compact irreducible symplectic variety. Then we have
(1.5) |
We prove Theorem 1.5 following the ideas of [SY22], which connects the cohomology groups in (1.5) to the weight spaces of the Looijenga–Lunts–Verbitsky algebra; cf. [LL97, Ver90, Ver95, Ver96]. As a byproduct we deduce that (1.5) refines (1.2), which justifies that Conjecture 1.2 categorifies Theorem 1.1.
Acknowledgments
We are grateful to Davesh Maulik for his enthusiasm and for many helpful discussions. We also thank Bohan Fang, Mirko Mauri, Peng Shan, and the anonymous referee for useful comments and suggestions.
2. Smooth morphisms and variations of Hodge structures
Throughout this section, we assume that is smooth, so that the Hodge modules are variations of Hodge structures.
2.1. Variations of Hodge structures
As a consequence of the Arnold–Liouville theorem, a nonsingular fiber of a Lagrangian fibration is a complex torus. In particular, the smooth map is a family of abelian varieties. The key to understanding the topology of is the variation of Hodge structures
it is polarized of weight with associated holomorphic vector bundle . The integrable connection and the Hodge filtration
(2.1) |
are compatible via the Griffiths transversality relation
This yields an -linear map between the graded pieces of (2.1)
(2.2) |
Here is a vector bundle describing the variation of of the fibers with .
For our purpose, we also consider the variation of Hodge structures of weight . Its Hodge filtration is
where the piece is given by
We denote by the graded piece .
Lemma 2.1.
We have a canonical isomorphism of vector bundles
Proof.
The morphism is induced by the cup product and the compatibility with Hodge filtrations. It suffices to check that it is an isomorphism when restricting to each ; this follows from the fact that is an abelian variety, so that we have
2.2. Symplectic form
We discuss the interplay between the symplectic form and the variation of Hodge structures .
By [Mat05, Lemma 2.6], the symplectic form and a polarization on induce an isomorphism
(2.3) |
which further yields
Lemma 2.2 (Donagi–Markman).
The morphism is symmetric with respect to the two factors of the target.
Proof.
More generally, for any we consider the morphism
where the first map is induced by the Gauss–Manin connection and the second identity is given by Lemma 2.1.
Corollary 2.3.
The composition
is symmetric with respect to the second and third factors of the target.
Proof.
We proceed by induction on . The induction base is Lemma 2.1. Now assume that the statement holds for . We have
We consider a local section of which can be written as with and local sections of and , respectively. The image consists of two terms and . We obtain from the induction hypothesis and the induction base that both of them are local sections of . This completes the induction. ∎
2.3. Proof of Theorem 1.3
The main ingredients of the proof are
We first note that by Lemma 2.1 we have a canonical isomorphism
(2.4) |
Hence (2.3) induces an isomorphism of vector bundles
(2.5) |
by switching the second and third factors of the right-hand side of (2.4).
Secondly, for any , the Gauss–Manin connection of induces an -linear morphism
The following proposition shows the compatibility between the isomorphisms and the morphisms ; it relies heavily on ingredient 2.
Proposition 2.4.
We have a commutative diagram
Proof.
To simplify the notation, we write the morphism of the top horizontal arrow as
(2.6) |
where we suppress the isomorphisms induced by (2.3) and Lemma 2.1. In particular, for a local section
(2.7) |
the image is a local section of . Similarly, for the same element (2.7) we have
To prove the commutativity of the diagram, it suffices to show that and coincide under the natural isomorphism switching the second and the third factors
By Griffiths transversality, the morphism of (2.6) is linear with respect to the second factor on the left-hand side as it represents . Therefore, we have
Here
(2.8) |
and the wedge product with and are on the second and third factors, respectively. Hence the desired property is a consequence of Corollary 2.3, which states that (2.8) is symmetric with respect to the second and third factors. ∎
Lastly, we show that Theorem 1.3 follows from the commutative diagram of Proposition 2.4. Since is smooth, the -module in the Hodge module is the variation of Hodge structures , and the filtration is described as
In particular, the de Rham complex of is
and the associated perverse–Hodge complexes are
We prove Theorem 1.3 by showing that the two complexes and match term by term via the isomorphisms (2.5). For convenience we may assume . As for , we find
(2.9) |
On the other hand, we have for by the fact that is a family of abelian varieties of dimension . Consequently, the complex is of the form
(2.10) |
We see that the complexes (2.9) and (2.10) are of the same length and are both concentrated in degrees . Then Proposition 2.4 yields an isomorphism
where the commutative diagram guarantees that it is indeed an isomorphism of complexes. This completes the proof. ∎
3. Hodge modules
In order to extend the isomorphisms established in Section 2 to the singular fibers, we need to use Saito’s theory of Hodge modules [Sai89, Sai90]. We begin with some relevant properties of Hodge modules. Then we recall Matsushita’s result [Mat05] on the higher direct images of ; we show in Proposition 3.3 that Matsushita’s result is equivalent to the case of Conjecture 1.2.
3.1. Hodge modules
Recall that a variation of Hodge structures of weight on a nonsingular variety is a triple
(3.1) |
where is a (-)local system, is an increasing filtration, and , such that the restriction to each is a pure Hodge structure of weight .(3)(3)(3)Traditionally the Hodge filtration is a decreasing filtration; the relation with the increasing filtration here is . We say that is polarizable if it admits a morphism
inducing a polarization on each stalk .
Pure Hodge modules introduced by Saito [Sai89] are vast generalizations of variations of Hodge structures. As in (3.1), a pure Hodge module on is a triple
(3.2) |
where is a regular holonomic -module, is a good filtration on , and is a perverse sheaf on , such that corresponds to via the Riemann–Hilbert correspondence.(4)(4)(4)For convenience, we sometimes only use the pair to denote a pure Hodge module. Such triples satisfy a number of technical conditions; in particular, one can define the notions of weight and polarization.
The following theorem by Saito [Sai90] provides a concrete description of polarizable pure Hodge modules on as extensions of polarizable variations of Hodge structures.
Theorem 3.1 (Saito).
-
(1)
The category of polarizable pure Hodge modules of weight on is abelian and semisimple.
-
(2)
For any closed subvariety , a simple polarizable variation of Hodge structures of weight on a nonsingular open subset of can be uniquely extended to a simple object in .
-
(3)
All simple objects in arise this way.
From now on, all Hodge modules are assumed to be pure and polarizable. For any simple Hodge module (3.2), we define its support to be the support of the simple perverse sheaf .
Set . The de Rham complex of a Hodge module is
and is concentrated in degrees . The filtration induces an increasing filtration
whose graded piece is the complex of -modules
Note that for when is given by a variation of Hodge structures, but this is not true for general Hodge modules. The functor extends naturally to the bounded derived category of Hodge modules taking values in .
3.2. Decomposition theorem, Saito’s formula, and duality
Let be a projective morphism between nonsingular varieties. For a Hodge module , Saito’s decomposition theorem [Sai89] states that there is a decomposition in the bounded derived category of Hodge modules on ,
(3.3) |
with . Its compatibility with the functor is given by the following formula (often known as Saito’s formula; see [Sai89, Section 2.3.7]):
(3.4) |
The functor is also compatible with Serre duality. Recall that . For a Hodge module , we have
(3.5) |
where is the dualizing sheaf of ; see [Sch16, Lemma 7.4].
Now we consider a Lagrangian fibration with . For our purpose, we study the direct image of the trivial Hodge module
A direct calculation yields
(3.6) |
Here conventionally for . Formulas (3.3) and (3.4) then read
and
(3.7) |
Finally, since and all the fibers of have dimension , we have for . Applying the duality (3.5), we also have for .
3.3. Matsushita’s theorem revisited
Let be a Lagrangian fibration with . Matsushita [Mat05] calculated the higher direct images of .
Theorem 3.2 (Matsushita).
For , we have
(3.8) |
Proof.
Matsushita’s proof assumes that is projective. However, as he explained in [Mat05, Remark 2.10], the projectivity for is only used for Kollár’s decomposition [Kol86]
Since the decomposition now holds for any projective morphism as a consequence of Saito’s theory of Hodge modules [Sai91], we may safely remove the projectivity assumption for in Matsushita’s theorem. ∎
Proposition 3.3.
Proof.
Remark 3.4.
In view of (3.7), Conjecture 1.2 would provide a new recipe for calculating the higher direct images of for all , extending Matsushita’s work. We could use the perverse–Hodge symmetry to trade the contributions of for all for the contributions of for all . The latter has the advantage of only involving a single Hodge module .
3.4. An example
To illustrate the subtleties when extending Theorem 1.3 to the singular fibers, we consider the following basic example.
Let be an elliptic fibration of a symplectic surface. We assume that only has singular fibers with a double point, over a finite set . Let be the open embedding.
We look at the symmetry
(3.11) |
proposed by Conjecture 1.2. Since the Hodge module is the trivial Hodge module , the left-hand side of (3.11) is . For the right-hand side, let denote the variation of Hodge structures on . Then the Hodge module on is the minimal extension , which can be described concretely using Deligne’s canonical extension.
Recall that the canonical extension depends on a real interval or where the eigenvalues of the residue endomorphism should lie. In our situation, the monodromy around each point of is unipotent (given by the matrix in local coordinates), so the eigenvalues are necessarily integers. Let be the canonical extension of with respect to either or ; it is locally free of rank on . Schmid’s theorem says that is a filtration by locally free subsheaves. By [Sai90, Section 3.10], we have
where the -action is induced by Deligne’s meromorphic connection on , and
It follows that for the right-hand side of (3.11), we have
(3.12) |
In particular, as a complex it has two nontrivial terms. But is clearly surjective; to see its kernel, we do a calculation in local coordinates. Let be the local coordinate of near , and let be a local trivialization of . Since the monodromy matrix around is , the residue matrix for is . In other words, we have
We also have
where are holomorphic functions with nonvanishing. From this we see that
hence
(3.13) |
On the other hand, the map induces an isomorphism
(3.14) |
sending to . Combining (3.13) and (3.14), we deduce that the kernel of in (3.12) is
Finally, by [Kol86, Theorem 2.6] and (3.8), we have
which yields the desired isomorphism (3.11) only(!) in the derived category .
Note that the proof in Section 4.1 works for all symplectic surfaces with and does not rely on information about the singular fibers.
4. Hilbert schemes of points
In this section we prove Theorem 1.4; it is completed by a series of compatibility results regarding the perverse–Hodge symmetry and natural geometric operations.
4.1. Surfaces
We first verify Theorem 1.4 for , where the Lagrangian fibration is an elliptic surface . It suffices to prove Conjecture 1.2 for
The cases when were covered by Proposition 3.3. Therefore, it remains to show the symmetry (3.11) for , whose left-hand side is
The right-hand side can be computed via the duality (3.5):
where the last isomorphism follows from (3.10).
4.2. Closed embeddings and finite morphisms
Let be an irreducible quasi-projective variety of dimension . Let
(4.1) |
be a finite sequence of Hodge modules on . We say that the Hodge modules in (4.1) are perverse–Hodge symmetric (PHS for short) on if for any we have
Clearly Conjecture 1.2 is equivalent to the statement that the Hodge modules given by the decomposition theorem (1.3) are PHS on . In general we say that a morphism is PHS if the trivial Hodge module is pure on and the Hodge modules obtained from the decomposition theorem of are PHS on .
The following proposition shows the compatibility between the perverse–Hodge symmetry and push-forwards along closed embeddings and finite morphisms.
Proposition 4.1.
Assume that the Hodge modules are PHS on .
-
(1)
Let be a closed embedding of codimension . Then the Hodge modules
are PHS on .
-
(2)
If is a finite surjective morphism with , then the Hodge modules
are PHS on .
4.3. External products
Let and be quasi-projective varieties, and let
be Hodge modules on and , respectively. We recall the following standard lemma concerning the external product
on .
Lemma 4.2.
We have
Proof.
This follows from the fact that the (filtered) de Rham functor is compatible with taking external product (cf. [MSS11, Equation (1.4.1)]). ∎
Now we consider projective morphisms
(4.2) |
with nonsingular. For each , we have the Hodge modules obtained from the decomposition theorem
We show the compatibility between the perverse–Hodge symmetry and products of varieties.
Proposition 4.3.
Proof.
Since
we have
Therefore, from the decomposition theorem
we obtain each summand
By Lemma 4.2 this further yields
Using this decomposition and the fact that the are PHS, we see that there is a one-to-one correspondence between the summands in the decompositions of
respectively. This completes the proof. ∎
Remark 4.4.
Remark 4.5.
We note that the only use of the nonsingular assumption in Proposition 4.3 is that the trivial Hodge modules on are pure.
4.4. Symmetric products
Let be a Hodge module on . Its symmetric product was introduced in [MSS11], which defines a Hodge module on the symmetric product of the variety. Furthermore, such an operation is extended to the bounded derived category of Hodge modules on .
Proposition 4.6.
If the are PHS on , then the are PHS on .
Proof.
Let be the -quotient map. For an object , we similarly consider the symmetric product
Now by Proposition 4.3, we know that the Hodge modules are PHS on . Proposition 4.1(2) further implies that the Hodge modules are PHS on . To prove the corresponding property for
on , it suffices to show that the natural isomorphism
is equivariant with respect to the -actions. It follows from [MSS11] that the (filtered) de Rham functor is compatible with the symmetric group action on ; more precisely, see [MSS11, proof of Proposition 1.5]. ∎
4.5. Proof of Theorem 1.4
For our purpose, we describe the decomposition theorem associated with the morphism
The first map of the composition is semismall, and the associated decomposition theorem is calculated in [GS93], which we now review. For a partition
(4.3) |
of , we use to denote the variety
We consider the stratification of the target variety by the combinatorial types of the points:
there is a canonical finite surjective morphism
see [GS93, Section 3]. We use to denote the length of the partition (4.3). Then we have
The main result of [GS93] is the decomposition theorem
Composing with the symmetric product map induced by , we have
(4.4) |
Now we consider the commutative diagram
where is defined analogously to , is induced by , and the right vertical arrow is the restriction of . Since
the right-hand side of (4.4) can be expressed as
(4.5) |
We complete the proof of Theorem 1.4 by showing that the Hodge modules given by each term in (4.5) are PHS on .
Since is PHS, by Proposition 4.3, is PHS for any . Combining with Proposition 4.6, we obtain that each symmetric product is PHS, which further yields that is PHS by taking products. Equivalently, the Hodge modules obtained from the decomposition theorem
are PHS on . Finally, we push forward the Tate-twisted Hodge modules
(4.6) |
along the composition of the finite surjective map with the closed embedding . By Proposition 4.1, the Hodge modules (4.6) are PHS on as desired, where the Tate twist in (4.6) is crucial. ∎
5. Global cohomology and LLV algebras
We specialize to Lagrangian fibrations associated with compact irreducible symplectic varieties. We prove Theorem 1.5, which is the perverse–Hodge symmetry at the level of global cohomology. Our main tool is the Looijenga–Lunts–Verbitsky (LLV for short) Lie algebra [Ver95, Ver96, LL97], a structure that is unique to compact irreducible symplectic varieties.
5.1. LLV (sub)algebras
Let be a compact irreducible symplectic variety or, equivalently, a projective hyper-Kähler manifold. Assume . We call an element of Lefschetz type if for any , cupping with gives an isomorphism
Such a class induces an -triple acting on . The LLV algebra is generated by all -triples associated with Lefschetz type classes. As is shown in [Ver95, Ver96] and [LL97] independently, there is a natural isomorphism
where is the second Betti number of .
Two subalgebras of played a crucial role in establishing Theorem 1.1. The first is Verbitsky’s generated by the -triples associated with the three Kähler classes . By [Ver90], the weight decomposition of under Verbitsky’s coincides with the Hodge decomposition. Consider the Cartan subalgebra of this spanned by
Then the Hodge decomposition
(5.1) |
satisfies
The second is the perverse introduced in [SY22], which concerns a Lagrangian fibration . Let be the pull-back of an ample class on , and let be a relatively ample class satisfying ; here is the Beauville–Bogomolov–Fujiki form on . The perverse is generated by the -triples associated with
and a third element satisfying
here is the bilinear form associated with .
5.2. Perverse–Hodge algebra
Let be a Lagrangian fibration with a compact irreducible symplectic variety of dimension . Since is of Picard rank , the ample class admits a unique decomposition with as in the previous section. We also have , where is the symplectic form; hence . In particular, the perverse can be generated by the -triples associated with , and .
We now consider the subalgebra generated by the -triples associated with
(5.3) |
We call it the perverse–Hodge algebra; it is naturally isomorphic to by the description of in [Ver95, Theorem 11.1]. A Cartan subalgebra is spanned by
(5.4) |
We have the weight decomposition
(5.5) |
so that
The perverse–Hodge algebra contains both the perverse and Verbitsky’s as subalgebras. Comparing (5.5) with (5.2) and (5.1), we find
where is the Hodge filtration on the pure Hodge structure .
On the other hand, by Saito’s formula (3.4) applied to the Hodge module under the projection , we have
The left-hand side is precisely the cohomology of the perverse–Hodge complex . We conclude that
(5.6) |
5.3. Proof of Theorem 1.5
Via (5.6), we have identified the cohomology groups with the weight spaces of under the perverse–Hodge algebra . Theorem 1.5 is then equivalent to the symmetry of the weight spaces
(5.7) |
The symmetry is a feature of -representations. More concretely, consider the subspace
equipped with the quadratic form . Set
By [Ver95, Theorem 11.1] we have a natural isomorphism . Up to renormalization, we may assume
so that form a standard isotropic basis of . Under this basis, the three elements in (5.4) are precisely (twice) the basis of described in [FH91, Section 18.1]. Let denote the dual basis of . By [FH91, Section 18.1, p. 271], the Weyl group of is isomorphic to , where the symmetric group permutes the three coordinate axes of and the generators of act diagonally by and . In particular, there is an element of the Weyl group exchanging while fixing . Consequently, we obtain the symmetry of the weight spaces (5.7) through the Weyl group action. ∎
Remark 5.1.
Remark 5.2.
Remark 5.3.
One can collect the numbers and depict them in a -space. We call it the perverse–Hodge diamond of . For example, the -plane has the shape
To simplify the discussion we assume . Then the perverse–Hodge algebra can be upgraded to an by adding one more -triple, namely the one associated with an element which is orthogonal to the four classes in (5.3) with respect to and shares the same norm. The Weyl group of is the full symmetry group of the regular octahedron, and as such it acts on the perverse–Hodge diamond (whereas the Weyl group of acts as the subgroup of rotational symmetries). Meanwhile, it is expected that the perverse–Hodge diamond has precisely the shape of a regular octahedron, meaning that no nonzero numbers lie outside the convex hull of the six vertices
See also [GKL+22, Conjecture 1.19] for an even stronger conjecture. This expectation is verified for all known families of compact irreducible symplectic varieties in [GKL+22, Theorem 1.23]. For even cohomology , the expectation is shown in [GKL+22, Theorem 5.2] and [HM23, Corollary 3.4] to be equivalent to Nagai’s conjecture for type II degenerations of hyper-Kähler manifolds deformation equivalent to . It is verified for in [GKL+22, Proposition 6.5]; more generally, the -plane and -plane cases are proven in [HM23, Theorem 1.2]. (Alternatively, one can show the two cases using octahedral symmetry and the knowledge that the border of the Hodge diamond of only has ’s and ’s.)
References
- [Bea91] A. Beauville, Systèmes hamiltoniens complètement intégrables associés aux surfaces , in: Problems in the theory of surfaces and their classification (Cortona, 1988), pp. 25–31, Sympos. Math., XXXII, Academic Press, London, 1991.
- [BBD82] A. A. Beĭlinson, J. Bernstein, and P. Deligne, Faisceaux pervers, in: Analysis and topology on singular spaces, I (Luminy, 1981), pp. 5–171, Astérisque 100, 1982.
- [DM96] R. Donagi and E. Markman, Spectral covers, algebraically completely integrable, Hamiltonian systems, and moduli of bundles, in: Integrable systems and quantum groups (Montecatini Terme, 1993), pp. 1–119, Lecture Notes in Math., vol. 1620, Springer, Berlin, 1996.
- [FSY22] C. Felisetti, J. Shen, and Q. Yin, On intersection cohomology and Lagrangian fibrations of irreducible symplectic varieties, Trans. Amer. Math. Soc. 375 (2022), no. 4, 2987–3001.
- [FH91] W. Fulton and J. Harris, Representation theory. A first course, Grad. Texts in Math., vol. 129, Readings in Mathematics, Springer-Verlag, New York, 1991.
- [GS93] L. Göttsche and W. Sörgel, Perverse sheaves and the cohomology of Hilbert schemes of smooth algebraic surfaces, Math. Ann. 296 (1993), 235–245.
- [GKL+22] M. Green, Y.-J. Kim, R. Laza, and C. Robles, The LLV decomposition of hyper-Kähler cohomology (the known cases and the general conjectural behavior), Math. Ann. 382 (2022), no. 3-4, 1517–1590.
- [Grö14] M. Gröchenig, Hilbert schemes as moduli of Higgs bundles and local systems, Int. Math. Res. Not. 2014 (2014) (23), 6523–6575.
- [HLS+21] A. Harder, Z. Li, J. Shen, and Q. Yin, P = W for Lagrangian fibrations and degenerations of hyper-Kähler manifolds, Forum Math. Sigma 9 (2021), e50.
- [Hit87a] N. J. Hitchin, The self-duality equations on a Riemann surface, Proc. London Math. Soc. (3) 55 (1987), no. 1, 59–126.
- [Hit87b] by same author, Stable bundles and integrable systems, Duke Math. J. 54 (1987), no. 1, 91–114.
- [HM22] D. Huybrechts and M. Mauri, Lagrangian fibrations, Milan J. Math. 90 (2022), no. 2, 459–483.
- [HM23] by same author, On type II degenerations of hyperkähler manifolds, Math. Res. Lett. 30 (2023), no. 1, 125–141.
- [Hwa08] J.-M. Hwang, Base manifolds for fibrations of projective irreducible symplectic manifolds, Invent. Math. 174 (2008), no. 3, 625–644.
- [Kol86] J. Kollár, Higher direct images of dualizing sheaves. II, Ann. of Math. (2) 124 (1986), no. 1, 171–202.
- [LL97] E. Looijenga and V. A. Lunts, A Lie algebra attached to a projective variety, Invent. Math. 129 (1997), no. 2, 361–412.
- [Mat05] D. Matsushita, Higher direct images of dualizing sheaves of Lagrangian fibrations, Amer. J. Math. 127 (2005), no. 2, 243–259.
- [MSS11] L. Maxim, M. Saito, and J. Schürmann, Symmetric products of mixed Hodge modules, J. Math. Pures Appl. (9) 96 (2011), no. 5, 462–483.
- [Sai89] M. Saito, Modules de Hodge polarisables, Publ. Res. Inst. Math. Sci. 24 (1988), no. 6, 849–995 (1989).
- [Sai90] by same author, Mixed Hodge modules, Publ. Res. Inst. Math. Sci. 26 (1990), no. 2, 221–333.
- [Sai91] by same author, On Kollár’s conjecture, in: Several complex variables and complex geometry (Santa Cruz, CA, 1989), pp. 509–517, Proc. Sympos. Pure Math. vol. 52, Part 2, Amer. Math. Soc., Providence, RI, 1991.
- [Sch16] C. Schnell, On Saito’s vanishing theorem, Math. Res. Lett. 23 (2016), no. 2, 499–527.
- [SY22] J. Shen and Q. Yin, Topology of Lagrangian fibrations and Hodge theory of hyper-Kähler manifolds (with an appendix by C. Voisin), Duke Math. J. 171 (2022), no. 1, 209–241.
- [Ver90] M. S. Verbitskiĭ, Action of the Lie algebra of on the cohomology of a hyper-Kähler manifold, Funct. Anal. Appl. 24 (1990), no. 3, 229–230.
- [Ver95] M. Verbitsky, Cohomology of compact hyperkaehler manifolds, Ph.D. thesis, Harvard University, 1995, arXiv:alg-geom/9501001.
- [Ver96] by same author, Cohomology of compact hyper-Kähler manifolds and its applications, Geom. Funct. Anal. 6 (1996), no. 4, 601–611.
- [Voi18] C. Voisin, Torsion points of sections of Lagrangian torus fibrations and the Chow ring of hyper-Kähler manifolds, in: Geometry of moduli, pp. 295–326, Abel Symp. vol. 14, Springer, Cham, 2018.
- [Zha17] Z. Zhang, Multiplicativity of perverse filtration for Hilbert schemes of fibered surfaces, Adv. Math. 312 (2017), 636–679.