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

The Looijenga–Lunts–Verbitsky Algebra for Primitive Symplectic Varieties with Isolated Singularities

Benjamin Tighe Department of Mathematics, University of Oregon, Eugene, OR 97403, USA bentighe@uoregon.edu
    • scAbstract. We extend results of Looijenga–Lunts and Verbitsky and show that the total Lie algebra g\mathfrak g for the intersection cohomology of a primitive symplectic variety XX with isolated singularities is isomorphic to

      gscso\mathfrak g\cong\mathfrak scso
      ((IH

      ^2(X, Q), Q_X)⊕h),

      whereQXistheintersectionBeauvilleBogomolovFujikiformandhisahyperbolicplane.Thisgivesanew,algebraicproofforirreducibleholomorphicsymplecticmanifoldswhichdoesnotrelyonthehyperkählermetric.Alongtheway,westudythestructureofIH(X,Q)asascgwhereQ_{X}istheintersectionBeauville--Bogomolov--Fujikiformand\mathfrak hisahyperbolicplane.Thisgivesanew,\textit{algebraic}proofforirreducibleholomorphicsymplecticmanifoldswhichdoesnotrelyonthehyperk\"{a}hlermetric.\par Alongtheway,westudythestructureof\operatorname{IH}^{*}(X,\mathbb Q)asa\mathfrak scg

      representationwithparticularemphasisontheVerbitskycomponent,multidimensionalKugaSatakeconstructions,andMumfordTatealgebrasandgivesomeimmediateapplicationsconcerningthe-representation---withparticularemphasisontheVerbitskycomponent,multidimensionalKuga--Satakeconstructions,andMumford--Tatealgebras---andgivesomeimmediateapplicationsconcerningtheP = Wconjectureforprimitivesymplecticvarieties.
      sc𝐊𝐞𝐲𝐰𝐨𝐫𝐝𝐬
      .
      conjectureforprimitivesymplecticvarieties.\vskip 3.0pt plus 1.0pt minus 1.0pt\par\vskip 12.0pt plus 4.0pt minus 4.0pt\par\noindent sc\bf Keywords.
      Hyperkähler manifolds, symplectic varieties, intersection cohomology, semismall morphisms

      sc2020 Mathematics Subject Classification. 32S35, 14C30, 32S60, 32S50 (primary); 14D06, 14B05, 14E15, 32S20 (secondary)

  •  
    cJuly 9, 2024Received by the Editors on August 25, 2023.
    Accepted on August 17, 2024.


    Department of Mathematics, University of Oregon, Eugene, OR 97403, USA

    sce-mail: bentighe@uoregon.edu


    © by the author(s) This work is licensed under http://creativecommons.org/licenses/by-sa/4.0/

1.  Introduction

1.1.  Background

Hyperkähler manifolds are distinguished in complex algebraic geometry due to their rich Hodge theory and form one of the three building blocks of the Beauville–Bogomolov decomposition of K-trivial varieties. Verbitsky’s global Torelli theorem, see [Ver13], states that hyperkähler manifolds are essentially determined by their second cohomology, along with its monodromy representation. One then expects the higher cohomology groups to be determined by the Hodge theory of H2H^{2}. This can be described using the Looijenga–Lunts–Verbitsky (LLV) algebra, a Lie algebra on the total cohomology H(X,Q)H^{*}(X,\mathbb Q) of a compact hyperkähler manifold XX, which was studied independently by Looijenga–Lunts [LL97] and by Verbitsky [Ver96] in his thesis.

We say a class ωH2(X,Q)\omega\in H^{2}(X,\mathbb Q) is HL if it satisfies the hard Lefschetz theorem: For every kk, the cupping morphism LωL_{\omega} on cohomology gives isomorphisms

Lωk:HdimXk(X,Q)HdimX+k(X,Q).L_{\omega}^{k}\colon H^{\dim X-k}(X,\mathbb Q)\overset{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}{\longrightarrow}H^{\dim X+k}(X,\mathbb Q).

Equivalently, ω\omega is HL if the nilpotent operator LωL_{\omega} completes to an sl2\mathfrak{sl}_{2}-triple gω=Lω,,ωH\mathfrak g_{\omega}=\langle L_{\omega},{}_{\omega},H\rangle, where H=[Lω,]ωH=[L_{\omega},{}_{\omega}] acts as (kdimX)id(k-\dim X)\operatorname{id} on Hk(X,Q)H^{k}(X,\mathbb Q). We define the LLV algebra of a hyperkähler manifold XX to be the Lie algebra generated by all possible sl2\mathfrak{sl}_{2}-algebras of the cohomology ring H(X)H^{*}(X) coming from hard Lefschetz operators:

g=Lω,:ωωisHL.\mathfrak{g}=\langle L_{\omega},{}_{\omega}:\omega\leavevmode\nobreak\ \mathrm{is\leavevmode\nobreak\ HL}\rangle.

Looijenga–Lunts and Verbitsky show that the LLV algebra admits a natural isomorphism

(1.1) gso((H2(X,Q),qX)h),\mathfrak g\cong\mathfrak{so}\left(\left(H^{2}(X,\mathbb Q),q_{X}\right)\oplus\mathfrak h\right),

where qXq_{X} is the Beauville–Bogomolov–Fujiki form and h\mathfrak h a hyperbolic plane. The LLV algebra not only acts on H(X,Q)H^{*}(X,\mathbb Q), but the structure theorem shows that the algebra is dependent only on the pair (H2(X,Q),qX)(H^{2}(X,\mathbb Q),q_{X}) and is therefore a deformation invariant. In fact, the various Hodge structures of the higher cohomology groups corresponding to deformations of XX are detected by the representation theory of g\mathfrak g, as gR:=so((H2(X,R),qX)h)\mathfrak g_{\mathbb R}:=\mathfrak{so}((H^{2}(X,\mathbb R),q_{X})\oplus\mathfrak h) contains the Weil operators C=i(pq)idC=i(p-q)\operatorname{id} for all complex structures on XX.

In recent years, there has been progress in generalizing the Hodge theory of hyperkähler manifolds to the singular setting. A primitive symplectic variety is a normal compact Kähler variety XX such that H1(𝒪X)=0H^{1}(\mathscr{O}_{X})=0 and the regular locus UU admits a global holomorphic symplectic form σ\sigma which extends holomorphically across any resolution of singularities and satisfies H0(U,)U2=CσH^{0}(U,{}_{U}^{2})=\mathbb C\cdot\sigma. Such varieties also enjoy a rich Hodge theory: By work of Bakker–Lehn [BL22], the second cohomology group of a primitive symplectic variety XX carries a pure Hodge structure and admits a version of global Torelli, which for Q\mathbb Q-factorial terminal singularities says that XX is essentially recovered by its H2H^{2}. It is then natural to ask if there is a generalization of the LLV algebra for primitive symplectic varieties which encodes the Hodge theory.

1.2.  Main Results

1.2.1.  The LLV algebra for intersection cohomology

Constructing the LLV algebra for the ordinary cohomology of a primitive symplectic variety is difficult since H(X,Q)H^{*}(X,\mathbb Q), a priori, neither carries a pure Hodge structure nor satisfies the hard Lefschetz theorem. Instead, we work with the intersection cohomology groups.

Intersection cohomology was invented by Goresky–MacPherson [GM80] as a way of generalizing Poincaré duality to singular topological spaces. Beilinson–Bernstein–Deligne [BBD82] observed, using characteristic pp methods, that the intersection cohomology groups of a projective variety admit a decomposition theorem with respect to projective morphisms. As a consequence, the intersection cohomology groups carry pure Hodge structures and satisfy the hard Lefschetz theorem. This was also observed by Saito [Sai88] in greater generality using the theory of mixed Hodge modules, as well as work of de Cataldo–Migliorini [dCM05] using purely Hodge theoretic techniques. The goal of this paper is to understand the total Lie algebra with respect to intersection cohomology, which we define analogously as the Lie algebra generated by the sl2\mathfrak{sl}_{2}-operators corresponding to any HL class, i.e., those classes ωIH2(X,Q)\omega\in\operatorname{IH}^{2}(X,\mathbb Q) such that

Lωk:IHdimXk(X,Q)IHdimX+k(X,Q).L_{\omega}^{k}\colon\operatorname{IH}^{\dim X-k}(X,\mathbb Q)\overset{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}{\longrightarrow}\operatorname{IH}^{\dim X+k}(X,\mathbb Q).

To this end, we define a Beauville–Bogomolov–Fujiki (BBF) form QXQ_{X} on the intersection cohomology IH2(X,Q)\operatorname{IH}^{2}(X,\mathbb Q) of a primitive symplectic variety XX; see Section 5.2.1. It is compatible with the standard BBF form qXq_{X} on H2(X,Q)H^{2}(X,\mathbb Q) (see Definition 2.5) and satisfies

QX|H2(X,Q)=qXQ_{X}|_{H^{2}(X,\mathbb Q)}=q_{X}

corresponding to the natural inclusion H2(X,Q)IH2(X,Q)H^{2}(X,\mathbb Q)\subset\operatorname{IH}^{2}(X,\mathbb Q) (see Remark 2.9).

Theorem 1.1.

Let XX be a primitive symplectic variety with isolated singularities and b25b_{2}\geq 5 and g\mathfrak g the algebra generated by all sl2\mathfrak{sl}_{2}-triples corresponding to HL classes in IH2(X,Q)\operatorname{IH}^{2}(X,\mathbb Q). There are isomorphisms

gso((IH2(X,Q),QX)h),gRso(B22,4),\mathfrak g\cong\mathfrak{so}\left(\left(\operatorname{IH}^{2}(X,\mathbb Q),Q_{X}\right)\oplus\mathfrak h\right),\quad\mathfrak g_{\mathbb R}\cong\mathfrak{so}(B_{2}-2,4),

where B2=dimIH2(X,Q)B_{2}=\dim\operatorname{IH}^{2}(X,\mathbb Q).

Moreover, a Hodge structure on IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q) is determined by a Hodge structure on IH2(X,Q)\operatorname{IH}^{2}(X,\mathbb Q) and the action of  g\mathfrak g on IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q).

The assumption on b2b_{2} is due to our use of the global moduli theory of Bakker–Lehn [BL22]. We note that the case b24b_{2}\leq 4 holds assuming the surjectivity of the period map and other special cases (see Section 5.4). We emphasize the fact that our proof of Theorem 1.1 gives an algebraic proof of (1.1). We expect our methods to generalize to any primitive symplectic variety.

1.2.2.  Symplectic symmetry on the intersection cohomology groups

One of the key features of the cohomology of a hyperkähler manifold XX is its structure as an irreducible holomorphic symplectic manifold. The holomorphic symplectic form σ\sigma on XX induces isomorphisms XnpXn+p{}_{X}^{n-p}\xrightarrow{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}{}_{X}^{n+p} by wedging, where 2n2n is the (complex) dimension of XX. Passing to cohomology, we get the symplectic hard Lefschetz theorem

Lσp:Hnp,q(X)Hn+p,q(X),L_{\sigma}^{p}\colon H^{n-p,q}(X)\overset{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}{\longrightarrow}H^{n+p,q}(X),

which induces the extra symmetry on the Hodge diamond of XX. An interesting observation is that, by deforming a compact hyperkähler manifold XX, we can see that the hard Lefschetz theory of H(X,Q)H^{*}(X,\mathbb Q) is related to its symplectic hard Lefschetz theory due to Verbitsky’s global Torelli theorem. With this in mind, we first show that the intersection cohomology groups of primitive symplectic varieties with isolated singularities also admit this symplectic symmetry.

Theorem 1.2.

Let XX be a primitive symplectic variety of dimension 2n2n with isolated singularities, and let IHp,q(X)IHk(X,C)\operatorname{IH}^{p,q}(X)\subset\operatorname{IH}^{k}(X,\mathbb C) be the (p,q)(p,q)-part of the canonical Hodge structure on IHk(X,Q)\operatorname{IH}^{k}(X,\mathbb Q). There is a cupping morphism Lσ:IHp,q(X)IHp+2,q(X)L_{\sigma}\colon\operatorname{IH}^{p,q}(X)\to\operatorname{IH}^{p+2,q}(X) on the Hodge pieces of the intersection cohomology which induced isomorphisms

Lσp:IHnp,q(X)IHn+p,q(X).L_{\sigma}^{p}\colon\operatorname{IH}^{n-p,q}(X)\overset{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}{\longrightarrow}\operatorname{IH}^{n+p,q}(X).

To prove Theorem 1.2, we study the Hodge theory of the (compactly supported) cohomology of the regular locus U:=XregU:=X_{\mathrm{reg}}. We prove the following useful theorem, giving an analog of [Ara90, Theorem 2]

Theorem 1.3.

Let XX be a primitive symplectic variety of dimension 2n2n with regular locus UU. Suppose that the singular locus of  XX is smooth. The Hodge-to-de Rham spectral sequence

E1p,q=Hq(U,)UpHp+q(U,C)E_{1}^{p,q}=H^{q}\left(U,{}_{U}^{p}\right)\Longrightarrow H^{p+q}(U,\mathbb C)

degenerates at E1E_{1} for p+q<2n1p+q<2n-1.

The LLV structure theorem follows from the symplectic hard Lefschetz theory. Theorem 1.2 shows that there are operators Lσ,σL_{\sigma},{}_{\sigma} which complete to an sl2\mathfrak{sl}_{2}-triple

sσ=Lσ,,σHσ,\mathfrak s_{\sigma}=\langle L_{\sigma},{}_{\sigma},H_{\sigma}\rangle,

where HσH_{\sigma} acts as the holomorphic weight operator Hσ(α)=(pn)αH_{\sigma}(\alpha)=(p-n)\alpha for an intersection (p,q)(p,q)-class. Similarly, by conjugation we get a second sl2\mathfrak{sl}_{2}-triple

sσ¯=Lσ¯,,σ¯Hσ¯s_{\overline{\sigma}}=\langle L_{\overline{\sigma}},{}_{\overline{\sigma}},H_{\overline{\sigma}}\rangle

corresponding to the antiholomorphic symplectic form σ¯\overline{\sigma}, where Hσ¯(α)=(qn)αH_{\overline{\sigma}}(\alpha)=(q-n)\alpha. This generates an sl2×sl2\mathfrak{sl}_{2}\times\mathfrak{sl}_{2}-structure on the total intersection cohomology IH(X)\operatorname{IH}^{*}(X). The key observation is that the Lefschetz operators for σ\sigma and σ¯\overline{\sigma} commute:

[Lσ,Lσ¯]=[,σ]σ¯=0.[L_{\sigma},L_{\overline{\sigma}}]=[{}_{\sigma},{}_{\overline{\sigma}}]=0.

The representation theory of this sl2×sl2\mathfrak{sl}_{2}\times\mathfrak{sl}_{2}-action, along with the monodromy representation of H2(X,C)H^{2}(X,\mathbb C), describes the LLV algebra g\mathfrak g of intersection cohomology completely. This will lead to the proof of Theorem 1.1.

Remark 1.4.

We expect our methods to generalize to any primitive symplectic variety, although this will require a better understanding of the Hodge theory of the intersection cohomology groups. One case where our methods generalize is the case of symplectic orbifolds (see Proposition 5.13), although the LLV structure theorem should be known to experts due to the existence of hyperkähler metrics.

1.3.  Representation Theory and Hodge Theory of the LLV Algebra

The structure of the cohomology ring H(X,Q)H^{*}(X,\mathbb Q) of a compact hyperkähler manifold XX as a g\mathfrak g-representation has been studied in recent years, leading to many interesting results and conjectures concerning these varieties. We extend some of these results to the intersection cohomology module of a primitive symplectic variety with isolated singularities.

1.3.1.  LLV decomposition

Extending a result of Verbitsky [Ver96], we show that the module generated by IH2(X,C)\operatorname{IH}^{2}(X,\mathbb C) in IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) is a g\mathfrak g-module, called the Verbitsky component V(n)V_{(n)}. In [GKLR22], the Verbitsky component and the LLV decomposition were studied for the known examples of compact hyperkähler manifolds. In general, it is expected that V(n)V_{(n)} (along with the LLV decomposition) puts restrictive conditions on the cohomology of a compact hyperkähler manifold, and we expect the same to be true for primitive symplectic varieties.

1.3.2.  Kuga–Satake construction

Recall that the classical Kuga–Satake construction, see [KS67], associates to a K3 surface SS an abelian variety AA and an embedding H2(S)H1(A)H1(A)H^{2}(S)\hookrightarrow H^{1}(A)\otimes H^{1}(A)^{*} of polarized weight 2 Hodge structures (see for example [Huy16, Section 2.6]). In [KSV19], this construction was extended to compact hyperkähler manifolds and the LLV algebra: If XX is compact hyperkähler, there exist an abelian variety AA and an embedding gXgA\mathfrak g_{X}\hookrightarrow\mathfrak g_{A}, where we note that gAso(H1(A)H1(A))\mathfrak g_{A}\cong\mathfrak{so}(H^{1}(A)\otimes H^{1}(A)^{*}) by [LL97, Proposition (3.3)]. We outline how the same construction associates to a primitive symplectic variety with isolated singularities a complex torus and an embedding of the total Lie algebras.

1.3.3.  Mumford–Tate algebras

The Mumford–Tate algebra of a pure Hodge structure HH is the smallest Q\mathbb Q-algebraic subalgebra m\mathfrak m of gl(H)\mathfrak{gl}(H) for which mR\mathfrak m_{\mathbb R} contains the Weil operator. There is a relationship between the Mumford–Tate algebra and the LLV algebra, which was studied thoroughly in [GKLR22]. We make similar observations in the isolated singularities case and show that the Mumford–Tate algebra sits naturally inside the LLV algebra for intersection cohomology. Moreover, we show that the degree of transcendence of the total intersection cohomology IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) over the special Mumford–Tate group is equal to that of IH2(X,C)\operatorname{IH}^{2}(X,\mathbb C), which is the second statement of Theorem 1.1.

1.4.  P=W\boldsymbol{P=W} for Primitive Symplectic Varieties

The P=WP=W conjecture for compact hyperkähler manifolds asserts that the perverse filtration on the cohomology H(X,C)H^{*}(X,\mathbb C) induced by a Lagrangian fibration agrees with the weight filtration of the limit mixed Hodge structure of a type III degeneration(1)(1)(1)Recall that a degeneration 𝒳\mathscr{X}\to\Delta is of type III if the degeneration has maximally unipotent monodromy operator TT and the nilpotent log-monodromy operator NN is of index 3. of XX, which was shown to exist in [Sol20]. In [HLSY21], the two filtrations were shown to agree by showing their corresponding weight operators define the same element in the LLV algebra.

More generally, let XX be a primitive symplectic variety. We show that there is a good notion of degeneration for primitive symplectic varieties via locally trivial deformations, which respects Schmid’s nilpotent orbit theory; see [Sch73]. We also show that, corresponding to a degeneration of primitive symplectic varieties, there is a limit mixed Hodge structure on the intersection cohomology of the central fiber (see Section 7.2.1). Following [Sol20, HLSY21], we show the following.

Theorem 1.5.

Let XX be a primitive symplectic variety with isolated singularities and b25b_{2}\geq 5.

  1. (1)

    There exists a type III degeneration 𝒳\mathscr{X}\to\Delta of  XX whose logarithmic monodromy operator NN has index 3.

  2. (2)

    If  XX admits the structure of a Lagrangian fibration f:XBf:X\to B, then the perverse filtration PβP_{\beta} on IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) associated to the pullback of an ample class on BB agrees with the weight filtration of the limit mixed Hodge structure of the degeneration 𝒳\mathscr{X}\to\Delta.

1.5.  Outline

The paper is organized as follows. In Section 2, we review some results of primitive symplectic varieties which will be used throughout this paper, as well as the relevant properties of intersection cohomology and mixed Hodge structures. We prove two auxiliary results that will allow us to simplify our assumptions: The first says that bimeromorphic morphisms of primitive symplectic varieties are semismall, a generalization of Kaledin’s result for symplectic resolutions [Kal06]. The second, which is most likely known to experts, is a criterion for Q\mathbb Q-factoriality for a primitive symplectic variety XX in terms of the inclusion H2(X,Q)IH2(X,Q)H^{2}(X,\mathbb Q)\hookrightarrow\operatorname{IH}^{2}(X,\mathbb Q); see Proposition 2.17.

In Section 3, we prove that the intersection cohomology groups satisfy a symplectic hard Lefschetz theorem. We do this by studying the extension of differential forms across singularities and the Hodge-to-de Rham spectral sequence on the regular locus.

In Section 4, we use the symplectic symmetry from Section 3 to construct dual Lefschetz operators with respect to the symplectic forms σ,σ¯\sigma,\overline{\sigma} and study their commutator relations. We then define non-isotropic classes (γ,γ)(\gamma,\gamma^{\prime}) satisfying qX(γ,γ)=0q_{X}(\gamma,\gamma^{\prime})=0 and [,γ]γ=0[{}_{\gamma},{}_{\gamma^{\prime}}]=0, a key component in the structure of the LLV algebra.

In Section 5, we show that the intersection cohomology groups satisfy the LLV algebra structure theorem, which is a consequence of the previous sections as well as the monodromy density theorem of Bakker–Lehn [BL22, Theorem 1.1].

In Section 6, we discuss some representation-theoretic aspects of the LLV algebra. We construct the Verbitsky component generated by IH2\operatorname{IH}^{2}, extend the Kuga–Satake construction to intersection cohomology, and study the Mumford–Tate algebra in this setting.

In Section 7, we describe a singular version of the P=WP=W theorem.

1.6.  Notation

We work in the complex analytic category, and all varieties should be considered as complex analytic varieties unless otherwise stated.

If EX~E\subset\widetilde{X} is a simple normal crossing divisor of a smooth complex manifold X~\widetilde{X}, we set

(logE)X~p(E):=(logE)X~p𝒪X~E,{}_{\widetilde{X}}^{p}(\log E)(-E):={}_{\widetilde{X}}^{p}(\log E)\otimes_{\mathscr{O}_{\widetilde{X}}}\mathscr{I}_{E},

where (logE)X~p{}_{\widetilde{X}}^{p}(\log E) is the sheaf of logarithmic pp-forms and E\mathscr{I}_{E} is the ideal sheaf of EE.

When we speak of intersection cohomology, we always mean with respect to the middle perversity.

Finally, if (X,σ)(X,\sigma) is a primitive symplectic variety, we will think of σ\sigma as a holomorphic form on the regular locus or a class in (intersection) cohomology without distinction.

Acknowledgements

This work is part of the author’s Ph.D. thesis at the University of Illinois at Chicago. I want to thank my advisor Benjamin Bakker for his insight into singular symplectic varieties, as well as the countless meetings devoted to this project. I also want to thank Christian Lehn for comments on a preliminary draft of this paper.

2.  Preliminaries and Auxiliary Results

The main objects of study in this work are primitive symplectic varieties, and so we recall both the local and the global properties which will be used further on. We also review the basic properties of intersection cohomology and study their Hodge theory for primitive symplectic varieties. Finally, we give some auxiliary results regarding bimeromorphic morphisms of primitive symplectic varieties, which will be useful for certain reductions later in the paper.

2.1.  Symplectic Varieties

2.1.1.  MMP singularities

Let XX be a normal complex variety. A log-resolution of singularities is a projective birational morphism π:X~X\pi\colon\widetilde{X}\to X from a smooth complex variety X~\widetilde{X} which is an isomorphism over the regular locus UU and for which π1()=E=Ei\pi^{-1}(\Sigma)=E=\sumop\displaylimits E_{i} is a simple normal crossing divisor, where is the singular locus of XX and the EiE_{i} are the smooth components of EE.

We say XX is Q\mathbb Q-Gorenstein if there is an integer mm such that mKXmK_{X} is Cartier, where KXK_{X} is the canonical divisor; the smallest such integer is called the index of XX. If XX has index 1, we say that XX is Gorenstein.

If XX is Q\mathbb Q-Gorenstein of index mm and π:X~X\pi\colon\widetilde{X}\to X a log-resolution of singularities, there are integers aia_{i} such that

mKX~=π(mKX)+aiEi.mK_{\widetilde{X}}=\pi^{*}(mK_{X})+\sumop\displaylimits a_{i}E_{i}.

We say that XX has canonical (resp. terminal ) singularities if we have ai0a_{i}\geq 0 (resp. ai>0a_{i}>0) for every ii. We call the aia_{i} the discrepancies of the exceptional divisors EiE_{i}.

If XX is just a normal variety, we say that XX has rational singularities if for some resolution of singularities π:X~X\pi\colon\widetilde{X}\to X, the higher direct image sheaves satisfy Riπ𝒪X~=0R^{i}\pi_{*}\mathscr{O}_{\widetilde{X}}=0 for every i>0i>0.

Canonical singularities are rational; see [KM98, Theorem 5.22]. Conversely, a normal variety with Gorenstein rational singularities has at worst canonical singularities. One way to see this is to consider the holomorphic extension problem for differentials: For which pp is the inclusion

(2.1) π⸦⟶X~pjUp\pi_{*}{}_{\widetilde{X}}^{p}\lhook\joinrel\longrightarrow j_{*}{}_{U}^{p}

is an isomorphism, where j:UXj\colon U\hookrightarrow X is the inclusion of the regular locus U:=XregU:=X_{\mathrm{reg}}? Kebekus–Schnell showed if XX has rational singularities, then πX~pjUp\pi_{*}{}_{\widetilde{X}}^{p}\hookrightarrow j_{*}{}_{U}^{p} is an isomorphism for each pp, see [KS21, Corollary 1.8], using the fact that the canonical sheaf ωX\omega_{X} is reflexive by Kempf’s criterion. In particular, the discrepancies aia_{i} are non-negative, as holomorphic nn-forms on UU extend with at worst zeros.

Since all our varieties are assumed normal, the sheaf jUpj_{*}{}_{U}^{p} is reflexive and isomorphic to the sheaf of reflexive pp-forms

:=X[p]()Xpj.Up{}_{X}^{[p]}:=({}_{X}^{p})^{**}\cong j_{*}{}_{U}^{p}.

Here 1X{}_{X}^{1} is the sheaf of Kähler differentials and pX{}_{X}^{p} is the pthp^{\mathrm{th}} exterior power.

One application of the work of Kebekus–Schnell is the existence of a functorial pullback morphism for reflexive differentials; see [KS21, Theorem 1.11]. Given a morphism f:YXf\colon Y\to X of reduced complex spaces with rational singularities, there is a pullback

(2.2) df:fX[p]Y[p]df\colon f^{*}{}_{X}^{[p]}\longrightarrow{}_{Y}^{[p]}

which satisfies natural universal properties; see [KS21, Section 14]. Specifically, it agrees with the pullback of Kähler differentials on smooth varieties, whenever this makes sense.

2.1.2.  Symplectic singularities

Definition 2.1.

Let XX be a normal variety. We say that XX is a symplectic variety if there is a holomorphic symplectic form σH0(U,)U2\sigma\in H^{0}(U,{}_{U}^{2}) on the regular locus UU which extends to a holomorphic 2-form σ~H0(X~,)X~2\tilde{\sigma}\in H^{0}(\widetilde{X},{}_{\widetilde{X}}^{2}) for any resolution of singularities π:X~X\pi\colon\widetilde{X}\to X.

Originally studied by Beauville [Bea00], symplectic varieties were defined as an attempt to extend results on KK-trivial manifolds to the singular setting. Symplectic varieties have well-behaved singularities: They are rational Gorenstein, see [Bea00, Proposition 1.3], and therefore have at worst canonical singularities. In fact, the strictly canonical locus is contained entirely in codimension 2 by [Nam01c]. Said differently, we have the following.

Proposition 2.2 (cf. [Nam01c, Corollary 1], [BL22, Theorem 3.4]).

A symplectic variety XX has terminal singularities if and only if  codimX()4\operatorname{codim}_{X}(\Sigma)\geq 4, where is the singular locus of  XX.

One of the key features of symplectic varieties is that they are stratified by varieties which, up to normalization, are again symplectic varieties. We will use the following structure theorem throughout the paper.

Proposition 2.3 (cf. [Kal06, Theorem 2.3,2.4], [BL22, Theorem 3.4]).

Let XX be a symplectic variety.

  1. (1)

    There is a stratification

    X=X0X1X2X=X_{0}\supset X_{1}\supset X_{2}\supset\cdots

    of  XX given by the singular locus of  XX, so that Xi=(Xi1)singX_{i}=(X_{i-1})_{\mathrm{sing}} for each ii. The normalization of each XiX_{i} is a symplectic variety, and Xi:=(Xi)regX_{i}^{\circ}:=(X_{i})_{\mathrm{reg}} admits a global holomorphic symplectic form.

  2. (2)

    Suppose that xXx\in X is such that xXix\in X_{i}^{\circ}. Let X^x\widehat{X}_{x} and Xi^x\widehat{X_{i}^{\circ}}_{x} be the completions of  XX and XiX_{i}^{\circ} at xx, respectively. Then there is a decomposition

    X^xYx×Xi^x,\widehat{X}_{x}\cong Y_{x}\times\widehat{X_{i}^{\circ}}_{x},

    where YxY_{x} is a symplectic variety.

In [Kal09], a symplectic variety YxY_{x} is a formal scheme rather than the completion of some symplectic variety, as its existence is predicted by solutions to differential equations derived from the Poisson structure (see Section 3 of op. cit.). By considering [KS24, Proposition 2.3, Appendix A], we can assume YxY_{x} is defined (and symplectic) in an analytic neighborhood of xx.

It would be interesting to understand how the holomorphic symplectic geometry of the regular strata XiX_{i}^{\circ} determines the geometry of a symplectic variety XX. For instance, if σH0(X,)X[2]\sigma\in H^{0}(X,{}_{X}^{[2]}) is the class of the symplectic form on the regular locus of XX, there is class jiσH0(Xinm,)Xi[2]j_{i}^{*}\sigma\in H^{0}(X_{i}^{\mathrm{nm}},{}_{X_{i}}^{[2]}), where ji:XinmXj_{i}\colon X_{i}^{\mathrm{nm}}\to X is the natural map from the normalization of XiX_{i}, defined by reflexive pullback for rational singularities; see [KS21, Theorem 14.1]. We claim this class is non-zero whenever dimXi>0\dim X_{i}>0. Since the problem is local, we can consider the product decomposition X^xYx×Xi^x\widehat{X}_{x}\cong Y_{x}\times\widehat{X_{i}^{\circ}}_{x} by Proposition 2.3. If jiσ=0j_{i}^{*}\sigma=0, then σ=p1σYx\sigma=p_{1}^{*}\sigma_{Y_{x}}, where p1:XYxp_{1}\colon X\to Y_{x} is the projection morphism and σYx\sigma_{Y_{x}} is the symplectic form on YxY_{x}. If dimYxdimX\dim Y_{x}\neq\dim X, then σdimX:=dimXσYx=0\sigma^{\dim X}:=\wedge^{\dim X}\sigma_{Y_{x}}=0. This is absurd if XX is a symplectic variety, and so we see that jiσj_{i}^{*}\sigma defines a non-zero global section of [2]Xi{}_{X_{i}}^{[2]}. Since XiX_{i}^{\circ} is symplectic, jiσj_{i}^{*}\sigma is a symplectic form.

2.1.3.  Primitive symplectic varieties

We now transition to global properties of symplectic varieties. The Hodge theory of singular symplectic varieties has been studied in [Nam01a, Nam06, Mat01, Mat15, Sch20, BL21, BL22] at varying levels of generality; primitive symplectic varieties, which were studied in [BL22], is the most general framework for studying the global properties of symplectic singularities.

The general framework of the global moduli theory of primitive symplectic varieties works in the category of complex Kähler varieties. A Kähler form on a reduced complex analytic space XX is given by an open covering X=iUiX=\bigcupop\displaylimits_{i}U_{i} and smooth strictly plurisubharmonic functions fi:UiRf_{i}\colon U_{i}\to\mathbb R such that on each intersection UijU_{ij}, the function fij=fi|Uijfj|Uijf_{ij}=f_{i}|_{U_{ij}}-f_{j}|_{U_{ij}} is locally the real part of a harmonic function. If XX admits a Kähler form, we say that XX is a Kähler variety. The most important property for this paper is that if XX is a Kähler variety, then XX admits a resolution by a Kähler manifold. If XX is a compact Kähler variety with at worst rational singularities and π:X~X\pi\colon\widetilde{X}\to X is a resolution of singularities, then there is an injection Hk(X,Z)Hk(X~,Z)H^{k}(X,\mathbb Z)\hookrightarrow H^{k}(\widetilde{X},\mathbb Z) for k2k\leq 2, see [BL21, Lemma 2.1], and Hk(X,Z)H^{k}(X,\mathbb Z) inherits a pure Hodge structure from Hk(X~,Z)H^{k}(\widetilde{X},\mathbb Z), which is described as follows. There is a spectral sequence

E1p,q:=Hq(X,j)UpHp+q(X,j)UE_{1}^{p,q}:=H^{q}(X,j_{*}{}_{U}^{p})\Longrightarrow\mathbb H^{p+q}(X,j_{*}{}_{U}^{\bullet})

which degenerates at E1E_{1} for p+q2p+q\leq 2; see [BL21, Lemma 2.2]. For k2k\leq 2, we have isomorphisms

(2.3) Hk(X,C)Hk(X,π)X~Hk(X,j)U,H^{k}(X,\mathbb C)\overset{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}{\longrightarrow}\mathbb H^{k}(X,\pi_{*}{}_{\widetilde{X}}^{\bullet})\cong\mathbb H^{k}(X,j_{*}{}_{U}^{\bullet}),

and the Hodge filtration is obtained by these isomorphisms and the degeneration of the reflexive Hodge-to-de Rham spectral sequence.

Definition 2.4.

A primitive symplectic variety is a compact Kähler symplectic variety (X,σ)(X,\sigma) such that H1(𝒪X)=0H^{1}(\mathscr{O}_{X})=0 and H0(X,)X[2]=CσH^{0}(X,{}_{X}^{[2]})=\mathbb C\cdot\sigma.

The second cohomology of a primitive symplectic variety is therefore a pure Hodge structure. As in the case of irreducible holomorphic symplectic manifolds, the geometry is controlled by the Hodge theory on H2H^{2}. To see this, we need the following quadratic form.

Definition 2.5.

Let XX be a primitive symplectic variety of dimension 2n2n with symplectic form σ\sigma. We define a quadratic form qX,σ:H2(X,C)Cq_{X,\sigma}\colon H^{2}(X,\mathbb C)\to\mathbb C given by the formula

qX,σ(α):=n2X(σσ¯)n1α2+(1n)Xσn1σ¯nαXσnσ¯n1α,q_{X,\sigma}(\alpha):=\frac{n}{2}\intop\nolimits_{X}(\sigma\overline{\sigma})^{n-1}\alpha^{2}+(1-n)\intop\nolimits_{X}\sigma^{n-1}\overline{\sigma}^{n}\alpha\intop\nolimits_{X}\sigma^{n}{\overline{\sigma}}^{n-1}\alpha,

where X\intop\nolimits_{X} is the cap product with the fundamental class.

This form qX,σq_{X,\sigma} defines a quadratic form over H2(X,R)H^{2}(X,\mathbb R) by definition. It is non-degenerate, and the signature of qX,σq_{X,\sigma} is (3,b23)(3,b_{2}-3), where b2=dimH2(X,R)b_{2}=\dim H^{2}(X,\mathbb R) is the second Betti number; see [Sch20, Theorem 2]. If we rescale σ\sigma so that X(σσ¯)n=1\intop\nolimits_{X}(\sigma\overline{\sigma})^{n}=1, then qX,σq_{X,\sigma} is independent of σ\sigma. In this case, we call qX:=qX,σq_{X}:=q_{X,\sigma} the Beauville–Bogomolov–Fujiki form.

We define the period domain of XX to be

:={[σ]P(H2(X,C))qX(σ)=0,qX(σ,σ¯)>0}.\Omega:=\left\{[\sigma]\in\mathbb P(H^{2}(X,\mathbb C))\mid q_{X}(\sigma)=0,\leavevmode\nobreak\ q_{X}(\sigma,\overline{\sigma})>0\right\}.

There is an associated period map

ρ:Deflt(X)\rho\colon\operatorname{Def}^{\,\mathrm{lt}}(X)\longrightarrow\Omega

from the universal family of locally trivial deformations of XX which sends tt to H2,0(Xt)H^{2,0}(X_{t}), where XtX_{t} is a locally trivial deformation of XX corresponding to a point tDeflt(X)t\in\operatorname{Def}^{\,\mathrm{lt}}(X). The period map ρ\rho is a local isomorphism by the local Torelli theorem, see [BL22, Proposition 5.5], which leads to two immediate consequences. First, qXq_{X} satisfies the Fujiki relation: There is a positive constant cc such that

(2.4) qX(α)n=cXα2n;q_{X}(\alpha)^{n}=c\intop\nolimits_{X}\alpha^{2n};

see [BL22, Proposition 5.15]. Second, we may rescale the BBF form to get an integral quadratic form on H2(X,Z)H^{2}(X,\mathbb Z); see [BL22, Lemma 5.7]. We denote the corresponding integral lattice by :=(H2(X,Z),qX)\Gamma:=(H^{2}(X,\mathbb Z),q_{X}).

The period map also satisfies many global properties. We can view the period domain as the moduli space of all weight 2 Hodge structures on which admit a quadratic form qq of signature (3,b23)(3,b_{2}-3) which is positive-definite on the real space underlying H2,0H0,2H^{2,0}\otimes H^{0,2}. We then write for the period domain. If XX^{\prime} is a locally trivial deformation of XX with (H2(X,Z),qX)(H^{2}(X^{\prime},\mathbb Z),q_{X^{\prime}})\cong\Gamma, then we get a period map p:p\colon\mathscr{M}^{\prime}\to from the moduli space of -marked locally trivial deformations of XX^{\prime}. The orthogonal group O()O(\Gamma) acts on \mathscr{M}^{\prime} and by changing the marking. For any connected component \mathscr{M} of \mathscr{M}^{\prime}, we denote by Mon()O()\operatorname{Mon}(\mathscr{M})\subset O(\Gamma) the image of the monodromy representation on second cohomology.

Theorem 2.6 (cf. [BL22, Theorem 1.1(1)]).

The monodromy group Mon()O()\operatorname{Mon}(\mathscr{M})\subset O(\Gamma) is of finite index.

Remark 2.7.

We will use the following interpretation of Theorem 2.6. Let Mon(X)\operatorname{Mon}(X) be the monodromy group associated to the connected component associated to a primitive symplectic variety XX. Since Mon(X)O()\operatorname{Mon}(X)\subset O(\Gamma) is of finite index, the restriction GX:=Mon(X)SO()SO()G_{X}:=\operatorname{Mon}(X)\cap\operatorname{SO}(\Gamma)\subset\operatorname{SO}(\Gamma) is also of finite index. By the Borel density theorem, GXG_{X} is therefore Zariski dense in SO()C\operatorname{SO}({}_{\mathbb C}).

2.2.  Intersection Cohomology

Our proof of the LLV structure theorem for primitive symplectic varieties uses the Hodge theory of intersection cohomology. We review the “Hodge–Kähler” package for the intersection cohomology of a compact complex space, which follows generally from Saito’s theory of mixed Hodge modules; see [Sai90]. We refer the reader to [dCM05] for an excellent treatment in the algebraic case.

2.2.1.  Hodge theory of intersection cohomology

The intersection cohomology is defined as the hypercohomology groups with respect to the intersection complex. The intersection cohomology complex is the perverse sheaf underlying the unique Hodge module determined by the constant variation of pure Hodge structures on QXreg[dimX]\mathbb Q_{X_{\mathrm{reg}}}[\dim X] over the regular locus of XX; see [Sai88, Section 5.3]. Writing 𝒞X\mathcal{IC}_{X} for this complex, we have

IHk(X,Q):=HkdimX(X,𝒞X).\operatorname{IH}^{k}(X,\mathbb Q):=\mathbb H^{k-\dim X}(X,\mathcal{IC}_{X}).
Proposition 2.8.

Suppose that XX is a compact Kähler variety.

  1. (1)

    ((Decomposition theorem)). If f:YXf\colon Y\to X is a projective morphism, there is a non-canonical isomorphism

    IHk(Y,Q)IHk(X,Q)λIH(Xλ,Lλ),\operatorname{IH}^{k}(Y,\mathbb Q)\cong\operatorname{IH}^{k}(X,\mathbb Q)\oplus\bigoplusop\displaylimits_{\lambda}\operatorname{IH}(X_{\lambda},L_{\lambda}),

    where the pairs (Xλ,Lλ)(X_{\lambda},L_{\lambda}) consist of closed subvarieties with semisimple local systems LλL_{\lambda}. In particular, IHk(X,Q)IHk(Y,Q)\operatorname{IH}^{k}(X,\mathbb Q)\subset\operatorname{IH}^{k}(Y,\mathbb Q).

  2. (2)

    ((Hodge decomposition)). The cohomology groups IHk(X,Q)\operatorname{IH}^{k}(X,\mathbb Q) all carry pure Hodge structures of weight kk.

  3. (3)

    ((hard Lefschetz)). Given an ample class αH2(X,Q)\alpha\in H^{2}(X,\mathbb Q) ((or, more generally, a Kähler class in H2(X,R))H^{2}(X,\mathbb R)), there is a cup product map Lα:IHk(X,Q)IHk+2(X,Q)L_{\alpha}\colon\operatorname{IH}^{k}(X,\mathbb Q)\to\operatorname{IH}^{k+2}(X,\mathbb Q) which produces isomorphisms

    Lαj:IHdimXj(X,Q)IHdimX+j(X,Q)L_{\alpha}^{j}\colon\operatorname{IH}^{\dim X-j}(X,\mathbb Q)\overset{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}{\longrightarrow}\operatorname{IH}^{\dim X+j}(X,\mathbb Q)

    for every jj.

  4. (4)

    For each kk, there is a natural morphism Hk(X,Q)IHk(X,Q)H^{k}(X,\mathbb Q)\to\operatorname{IH}^{k}(X,\mathbb Q) whose kernel is contained in Wk1Hk(X,Q)W_{k-1}H^{k}(X,\mathbb Q), the (k1)st(k-1)^{\mathrm{st}} piece of the weight filtration.

Remark 2.9.

If XX has rational singularities—such as a primitive symplectic variety—then H2(X,C)IH2(X,C)H^{2}(X,\mathbb C)\subset\operatorname{IH}^{2}(X,\mathbb C) since H2(X,C)H^{2}(X,\mathbb C) carries a pure Hodge structure by Proposition 2.8(4).

We will use the following description of intersection cohomology throughout the paper.

Lemma 2.10.

Suppose that XX is a projective variety with singular locus . If  codimX()3\operatorname{codim}_{X}(\Sigma)\geq 3, then IH2(X,C)H2(X,C)\operatorname{IH}^{2}(X,\mathbb C)\cong H^{2}(X\setminus\Sigma,\mathbb C).

Proof.

This is [Dur95, Lemma 1]. More generally, if codimX()X=d\operatorname{codim}_{X}({}_{X})=d, then IHk(X,C)Hk(X,C)\operatorname{IH}^{k}(X,\mathbb C)\cong H^{k}(X\setminus\Sigma,\mathbb C) for k<dk<d and k<dimXk<\dim X. ∎

Since the analytic topology of a singularity of a complex variety is preserved under locally trivial deformations, Lemma 2.10 holds for any primitive symplectic variety, as a general locally trivial deformation is projective; see [BL22, Corollary 6.10].

2.3.  Intersection Cohomology for Primitive Symplectic Varieties

In order to prove the structure theorem for the total Lie algebra, we need to adapt the hard Lefschetz theorem for non-ample classes. This can be done using the monodromy density theorem.

Proposition 2.11.

Let XX be a primitive symplectic variety of dimension 2n2n. Let αH2(X,Q)\alpha\in H^{2}(X,\mathbb Q) be any non-isotropic class with respect to the BBF form qXq_{X}. Then α\alpha satisfies hard Lefschetz: There is a cupping morphism Lα:IHk(X,Q)IHk+2(X,Q)L_{\alpha}\colon\operatorname{IH}^{k}(X,\mathbb Q)\to\operatorname{IH}^{k+2}(X,\mathbb Q) which induces isomorphisms

Lαk:IH2nk(X,Q)IH2n+k(X,Q).L_{\alpha}^{k}\colon\operatorname{IH}^{2n-k}(X,\mathbb Q)\longrightarrow\operatorname{IH}^{2n+k}(X,\mathbb Q).

Conversely, any class α\alpha which satisfies hard Lefschetz is non-isotropic.

Proof.

First, we note that there is a cupping morphism Lα:IHk(X,Q)IHk+2(X,Q)L_{\alpha}\colon\operatorname{IH}^{k}(X,\mathbb Q)\to\operatorname{IH}^{k+2}(X,\mathbb Q) for any αIH2(X,Q)\alpha\in\operatorname{IH}^{2}(X,\mathbb Q), which agrees with the usual cup product of Proposition 2.8(3) when α\alpha is a Kähler form. We follow [dCM05, Section 4.4]. By the decomposition theorem Proposition 2.8(1), any class αIH2(X,Q)\alpha\in\operatorname{IH}^{2}(X,\mathbb Q) can be lifted to a class α~H2(X~,Q)\tilde{\alpha}\in H^{2}(\tilde{X},\mathbb Q) for a resolution of singularities π:X~X\pi\colon\tilde{X}\to X. Consider the isomorphism

H2(X~,Q)HomD(X)(QX~,QX~[2]),H^{2}(\widetilde{X},\mathbb Q)\cong\operatorname{Hom}_{D(X)}\left(\mathbb Q_{\widetilde{X}},\mathbb Q_{\widetilde{X}}[2]\right),

where D(X)D(X) is the full subcategory of the bounded derived category Db(X)D^{b}(X) of constructible sheaves which are cohomologically constructible; see [dCM05, Definition 3.3.1]. In particular, we obtain a map

𝐑πα~:𝐑πQX~𝐑πQX~[2].\mathbf{R}\pi_{*}\tilde{\alpha}\colon\mathbf{R}\pi_{*}\mathbb Q_{\widetilde{X}}\longrightarrow\mathbf{R}\pi_{*}\mathbb Q_{\widetilde{X}}[2].

Now this construction is consistent with any pp-splitting of 𝐑πQX~\mathbf{R}\pi_{*}\mathbb Q_{\widetilde{X}}, in the sense of [dCM05, Definition 4.3.1]. In particular, this induces a map α:𝒞X𝒞X[2]\alpha\colon\mathcal{IC}_{X}\to\mathcal{IC}_{X}[2] via the decomposition theorem. This gives the desired morphism LαL_{\alpha} upon taking hypercohomology.

Let =(H2(X,Z),qX)\Gamma=(H^{2}(X,\mathbb Z),q_{X}), and consider the monodromy group Mon(X)\operatorname{Mon}(X) of -marked primitive symplectic varieties deformation-equivalent to XX. Then Mon(X)\operatorname{Mon}(X) is a finite-index subgroup of O()O(\Gamma) by Theorem 2.6. Then consider the subgroup

GX=Mon(X)SO(),G_{X}=\operatorname{Mon}(X)\cap\operatorname{SO}(\Gamma),

which is a Zariski-dense subset of SO()C\operatorname{SO}({}_{\mathbb C}) by Remark 2.7.

Let kerαk=ker(Lαk:IH2nk(X,Q)IH2n+k(X,Q))\ker\alpha^{k}=\ker(L_{\alpha}^{k}\colon\operatorname{IH}^{2n-k}(X,\mathbb Q)\to\operatorname{IH}^{2n+k}(X,\mathbb Q)). By Poincaré duality, it is enough to show that dimkerαk=0\dim\ker\alpha^{k}=0. Now for any gGXg\in G_{X}, it follows that dimkerαk=dimker(gα)k\dim\ker\alpha^{k}=\dim\ker(g\cdot\alpha)^{k}, as monodromy preserves cup product. If GX,αG_{X,\alpha} is the GX,αG_{X,\alpha}-orbit of α\alpha, we get the constant map GX,αZG_{X,\alpha}^{\circ}\to\mathbb Z given by gdimker(gα)kg\mapsto\dim\ker(g\cdot\alpha)^{k}. But GXG_{X} is Zariski dense, and therefore the constant map must extend to a constant map over the full SO()C\operatorname{SO}({}_{\mathbb C})-orbit of α\alpha; but this orbit necessarily contains the class of an ample divisor on some locally trivial deformation, which satisfies hard Lefschetz by Proposition 2.8(3). In particular, dimkerαk=0\dim\ker\alpha^{k}=0 for every α\alpha.

Conversely, if α\alpha satisfies hard Lefschetz, then the Fujiki relation (2.4) implies qX(α)0q_{X}(\alpha)\neq 0. ∎

We also need to move the intersection cohomology around in locally trivial deformations.

Proposition 2.12.

The intersection cohomology groups IHk(Xt,C)\operatorname{IH}^{k}(X_{t},\mathbb C) for tDeflt(X)t\in\operatorname{Def}^{\,\mathrm{lt}}(X) form a local system.

Proof.

As vector spaces, intersection cohomology is completely determined by the structure XtX_{t} admits as a stratified pseudomanifold by Deligne’s construction; see [GM83, Theorem 3.5]. The claim follows since locally trivial deformation are real analytically trivial; see [AV21, Proposition 5.1]. ∎

2.4.  Some Mixed Hodge Structures

The Hodge theory of the intersection cohomology of primitive symplectic varieties with isolated singularities can be completely described by the Hodge theory of its regular locus, see Section 3.3, and so we review the relevant parts of Deligne’s mixed Hodge structure on the (compactly supported) cohomology of smooth varieties. Deligne’s original treatment holds for algebraic varieties, but the same results hold in the Kähler setting by [Fuj80].

2.4.1.  Cohomology of the regular locus

Let UU be a smooth Kähler variety, and let X~\widetilde{X} be a smooth compactification of UU such that the complement X~UE\widetilde{X}\setminus U\cong E is a simple normal crossing (snc) divisor. A fundamental result of Deligne [Del71, Proposition 3.1.8] states that the cohomology of UU can be identified with the hypercohomology of the complex of logarithmic forms on the pair (X~,E)(\widetilde{X},E). Specifically, let (logE)X~p{}_{\widetilde{X}}^{p}(\log E) be the sheaf of logarithmic pp-forms, and let (logE)X~{}_{\widetilde{X}}^{\bullet}(\log E) the complex of logarithmic forms. Then there is an isomorphism

Hk(U,C)Hk(X~,(logE)X~),H^{k}(U,\mathbb C)\cong\mathbb H^{k}\left(\widetilde{X},{}_{\widetilde{X}}^{\bullet}(\log E)\right),

which induces two filtrations on Hk(U,C)H^{k}(U,\mathbb C). The first is the naive filtration associated to the complex (logE)X~{}_{\widetilde{X}}^{\bullet}(\log E):

FpHk(U,C)=im(Hk(X~,τp(logE)X~)Hk(X~,(logE)X~)),F^{p}H^{k}(U,\mathbb C)=\operatorname{im}\left(\mathbb H^{k}\left(\widetilde{X},\tau_{\geq p}{}_{\widetilde{X}}^{\bullet}(\log E)\right)\longrightarrow\mathbb H^{k}\left(\widetilde{X},{}_{\widetilde{X}}^{\bullet}(\log E)\right)\right),

where τp(logE)X~\tau_{\geq p}{}_{\widetilde{X}}^{\bullet}(\log E) is the complex (logE)X~{}_{\widetilde{X}}^{\bullet}(\log E) truncated in degree greater than or equal to pp. The second filtration is induced at the level of sheaves: There is an increasing filtration on logarithmic pp-forms given by

Wl(logE)X~p=(logE)X~lX~plW_{l}{}_{\widetilde{X}}^{p}(\log E)={}_{\widetilde{X}}^{l}(\log E)\wedge{}_{\widetilde{X}}^{p-l}

descending to Hk(X~,(logE)X~\mathbb H^{k}(\widetilde{X},{}_{\widetilde{X}}^{\bullet}(\log E). These filtrations correspond to the canonical mixed Hodge structure on Hk(U,Q)H^{k}(U,\mathbb Q)WW_{\bullet} is the complexification of the rational weight filtration, and FF^{\bullet} is the Hodge filtration.

In particular, there is a non-canonical decomposition

(2.5) Hk(U,C)p+q=kHq(X~,(logE)X~p).H^{k}(U,\mathbb C)\cong\bigoplusop\displaylimits_{p+q=k}H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{p}(\log E)\right).

We note that the above construction applies to the regular locus UU of any proper Kähler variety XX, where we may choose the smooth compactification to be a log-resolution X~\widetilde{X} of XX.

2.4.2.  Compactly supported cohomology of the regular locus

Deligne [Del71, Del74] also shows that the compactly supported cohomology of an algebraic variety carries a pure Hodge structure.

The compactly supported cohomology also carries a mixed Hodge structure. The easiest way to see this is to use Poincaré duality. Indeed, the Poincaré isomorphisms

Hck(U,C)(H2dimUk(U,C))C(dimU)H^{k}_{c}(U,\mathbb C)\cong\left(H^{2\dim U-k}(U,\mathbb C)\right)^{*}\otimes\mathbb C(-\dim U)

are isomorphisms of mixed Hodge structures. Noting that

Hq((logE)X~p)HdimX~q(X~,(logE)X~dimX~p(E))H^{q}\left({}_{\widetilde{X}}^{p}(\log E)\right)\cong H^{\dim\widetilde{X}-q}\left(\widetilde{X},{}_{\widetilde{X}}^{\dim\widetilde{X}-p}(\log E)(-E)\right)^{*}

by Serre duality, this completely describes the mixed Hodge structure on Hck(U,Q)H^{k}_{c}(U,\mathbb Q). In particular, we have another non-canonical splitting

(2.6) Hck(U,C)p+q=kHq(X~,(logE)X~p(E)).H_{c}^{k}(U,\mathbb C)\cong\bigoplusop\displaylimits_{p+q=k}H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{p}(\log E)(-E)\right).

2.4.3.  Cohomology of a simple normal crossing divisor

Finally, let us consider an snc divisor E=i=1rEiE=\sumop\displaylimits_{i=1}^{r}E_{i} with irreducible components EiE_{i}. For any subset J{1,,r}J\subset\left\{1,\ldots,r\right\}, write EJ=jJEjE_{J}=\bigcapop\displaylimits_{j\in J}E_{j} and let

(2.7) E(p):=|J|=pEJE_{(p)}:=\coprodop\displaylimits_{|J|=p}E_{J}

be the pp-fold intersections of the components. For each pp, the various inclusion maps E(p)E(p1)E_{(p)}\hookrightarrow E_{(p-1)} induce a simplicial map Hk(E(p1),Q)Hk(E(p),Q)H^{k}(E_{(p-1)},\mathbb Q)\to H^{k}(E_{(p)},\mathbb Q). From this we obtain a complex

0Hk(E(1),Q)δ1Hk(E(2),Q)δ2δp1Hk(E(p),Q)δp0\longrightarrow H^{k}(E_{(1)},\mathbb Q)\overset{\delta_{1}}{\longrightarrow}H^{k}(E_{(2)},\mathbb Q)\overset{\delta_{2}}{\longrightarrow}\cdots\xrightarrow{\delta_{p-1}}H^{k}(E_{(p)},\mathbb Q)\overset{\delta_{p}}{\longrightarrow}\cdots

which, for each kk, computes graded pieces of the weight filtration on Hk(E,C)H^{k}(E,\mathbb C):

grkWHk+r(E,C)=kerδr+1/imδr.\operatorname{gr}_{k}^{W}H^{k+r}(E,\mathbb C)=\ker\delta_{r+1}/\operatorname{im}\delta_{r}.

The (p,q)(p,q)-Hodge pieces are obtained by applying the functor Hq()pH^{q}({}^{p}) to the complex and taking cohomology. Precisely, there is an induced complex with morphisms

δrp,q:Hq(E(r),)E(r)pHq(E(r+1),)E(r+1)p,\delta^{p,q}_{r}\colon H^{q}\left(E_{(r)},{}_{E_{(r)}}^{p}\right)\longrightarrow H^{q}\left(E_{(r+1)},{}_{E_{(r+1)}}^{p}\right),

and

(2.8) grkWHk+r(E,C)p,q=kerδr+1p,q/imδrp,q.\operatorname{gr}_{k}^{W}H^{k+r}(E,\mathbb C)^{p,q}=\ker\delta_{r+1}^{p,q}/\operatorname{im}\leavevmode\nobreak\ \delta_{r}^{p,q}.

2.5.  A Result on Bimeromorphic Morphisms of Symplectic Varieties

Recall that a morphism ϕ:ZX\phi\colon Z\to X is semismall if dimZ×XZdimX\dim Z\times_{X}Z\leq\dim X. A result of Kaledin states that symplectic resolutions are semismall; see [Kal06, Lemma 2.11]. We want to extend this result to bimeromorphic morphisms of singular symplectic varieties, which will control the geometry of such morphisms.

Definition 2.13.

Let XX and ZZ be normal and Q\mathbb Q-Gorenstein varieties. A birational morphism ϕ:ZX\phi\colon Z\to X is crepant if ϕKX=KZ\phi^{*}K_{X}=K_{Z}.

Lemma 2.14.

If  XX is a ((primitive)) symplectic variety and ϕ:ZX\phi\colon Z\to X is a crepant birational morphism from a normal complex variety ZZ, then ZZ is also a ((primitive)) symplectic variety.

Proof.

By taking a common resolution of singularities of ZZ and XX, we can see that the symplectic form on XregX_{\mathrm{reg}} extends to a global reflexive 2-form σZ\sigma_{Z} on ZZ. Since KZ=ϕKX=𝒪ZK_{Z}=\phi^{*}K_{X}=\mathscr{O}_{Z}, then σZ\sigma_{Z} must define a holomorphic symplectic form on ZregZ_{\mathrm{reg}}, which therefore extends to a holomorphic 2-form on the common resolution (and therefore any resolution) of singularities. If XX is primitive symplectic, it is clear that ZZ is also a compact Kähler variety, H1(𝒪Z)=H1(𝒪X)=0H^{1}(\mathscr{O}_{Z})=H^{1}(\mathscr{O}_{X})=0, and H0(Z,)Z[2]=CσZH^{0}(Z,{}_{Z}^{[2]})=\mathbb C\cdot\sigma_{Z}. ∎

Lemma 2.15.

Let ϕ:ZX\phi\colon Z\to X be a crepant morphism of primitive symplectic varieties. Then the restriction of the class σZ\sigma_{Z} to ϕ1(x)\phi^{-1}(x) is 0 for every xx.

Proof.

Note that if XX is primitive symplectic, then so is ZZ. The result then follows from Hodge theory: By assumption, the class of the symplectic form ϕσX\phi^{*}\sigma_{X} extends to a symplectic form σZ\sigma_{Z} on ZZ, as both must extend to any common resolution of singularities. Therefore, its image under the morphism H2(Z,C)H0(X,R2ϕC)H^{2}(Z,\mathbb C)\to H^{0}(X,R^{2}\phi_{*}\mathbb C) is zero. The result then follows by proper base change. ∎

Proposition 2.16.

If XX is a primitive symplectic variety and ϕ:ZX\phi\colon Z\to X is a crepant morphism, then ϕ\phi is semismall.

Proof.

Let {Xi}\left\{X_{i}\right\} and {Zj}\left\{Z_{j}\right\} be the Kaledin stratifications of XX and ZZ (see Theorem 2.3). We will show that

(2.9) 2dimϕ1(x)+dimXidimX.2\dim\phi^{-1}(x)+\dim X_{i}\leq\dim X.

To prove this, we follow [Kal06, Lemma 2.11] in the case of symplectic resolutions. Kaledin’s proof, built upon work of Wierzba [Wie03] and Namikawa [Nam01b], uses the symplectic structure of a smooth symplectic variety ZZ to show that the tangent spaces Tzϕ1(x)T_{z}\phi^{-1}(x) and Tzϕ1(Xi)T_{z}\phi^{-1}(X_{i}) are mutually orthogonal, which immediately gives (2.9). We will therefore consider how the fibers intersect with the various smooth and symplectic strata ZjZ_{j}^{\circ}.

Let xXx\in X, and let XiX_{i} be the stratum for which xXix\in X_{i}^{\circ}. Write Ex=ϕ1(x)E_{x}=\phi^{-1}(x) and E=ϕ1(Xi)E=\phi^{-1}(X_{i}). For every zEregz\in E_{\mathrm{reg}}, there is a jj such that zZjz\in Z_{j}^{\circ}. There is then a map ϕ^x:Zj^zXi^x\widehat{\phi}_{x}\colon\widehat{Z_{j}^{\circ}}_{z}\to\widehat{X_{i}^{\circ}}_{x} induced from the map ϕ^:Z^zX^x\widehat{\phi}\colon\widehat{Z}_{z}\to\widehat{X}_{x} via the product decomposition. Consider the commutative diagram

Ereg^z{\widehat{E_{\mathrm{reg}}}_{z}}Zj^z{\widehat{Z_{j}^{\circ}}_{z}}Z^z{\widehat{Z}_{z}}Xi^x{\widehat{X_{i}^{\circ}}_{x}}X^x,{\widehat{X}_{x}\hbox to0.0pt{,\hss}}ϕ^x\scriptstyle{\widehat{\phi}_{x}}ϕ^\scriptstyle{\widehat{\phi}}

where Ereg^z\widehat{E_{\mathrm{reg}}}_{z} is the completion of EregE_{\mathrm{reg}} at zz. The varieties in the diagram all have at worst rational singularities, and the Xi^x\widehat{X_{i}^{\circ}}_{x}, Zj^z\widehat{Z_{j}^{\circ}}_{z} are symplectic varieties. By reflexive pullback, we see that the symplectic form σ^i\widehat{\sigma}_{i} on Xi^x\widehat{X_{i}^{\circ}}_{x} pulls back to the restriction σ^j|Ereg^z\widehat{\sigma}_{j}|_{\widehat{E_{\mathrm{reg}}}_{z}}, where σ^j\widehat{\sigma}_{j} is the symplectic form on Zj^z\widehat{Z_{j}^{\circ}}_{z}, which is just the restriction of the symplectic form σ^Z\widehat{\sigma}_{Z} on Z^z\widehat{Z}_{z}. Since σZ\sigma_{Z} vanishes on the fibers ϕ1(x)\phi^{-1}(x) (as a cohomology class), then σ^Z\widehat{\sigma}_{Z} vanishes on the fibers of ϕ^\widehat{\phi}. Since Zj^z\widehat{Z_{j}^{\circ}}_{z} is smooth and symplectic, we may assume the tangent spaces TzExT_{z}E_{x} and Tz(Ereg)T_{z}(E_{\mathrm{reg}}) are mutually orthogonal with respect to the symplectic form σj\sigma_{j} on ZjZ_{j}^{\circ} after passing to a small open neighborhood of zz. Therefore,

dimExdimXdimEdimXdimExdimXi.\dim E_{x}\leq\dim X-\dim E\leq\dim X-\dim E_{x}-\dim X_{i}.

The second inequality above follows from local product structure of XX along the smooth stratum XiX_{i}^{\circ}. This clearly agrees with (2.9), and so ϕ\phi is semismall. ∎

2.6.  A Q\mathbb Q-Factoriality Criterion

Let XX be a projective variety. We say that XX is Q\mathbb Q-factorial if every Weil divisor is Q\mathbb Q-Cartier. Equivalently, see [KM92, Proposition 12.1.6], the variety XX is Q\mathbb Q-Factorial if and only if for a resolution of singularities π:X~X\pi\colon\widetilde{X}\to X with exceptional divisor E=EiE=\sumop\displaylimits E_{i},

(2.10) im(H2(X~,Q)H0(X,R2πQ))=im(iQ[Ei]H0(X,R2πQ)).\operatorname{im}\left(H^{2}(\widetilde{X},\mathbb Q)\longrightarrow H^{0}(X,R^{2}\pi_{*}\mathbb Q)\right)=\operatorname{im}\left(\bigoplusop\displaylimits_{i}\mathbb Q[E_{i}]\longrightarrow H^{0}(X,R^{2}\pi_{*}\mathbb Q)\right).
Proposition 2.17.

If  XX is a terminal primitive symplectic variety, then XX is Q\mathbb Q-factorial if and only if the natural inclusion H2(X,C)IH2(X,C)H^{2}(X,\mathbb C)\hookrightarrow\operatorname{IH}^{2}(X,\mathbb C) is an isomorphism.

Proof.

Since XX is terminal, we have an isomorphism IH2(X,C)H2(U,C)\operatorname{IH}^{2}(X,\mathbb C)\xrightarrow{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}H^{2}(U,\mathbb C) by Lemma 2.10 and Proposition 2.2. The mixed Hodge structure on H2(U,Q)H^{2}(U,\mathbb Q) is then pure, and the logarithmic Hodge-to-de Rham spectral sequence computes the Hodge filtration; see [Del74, Section (9.2.3)]. In particular,

H2(U,C)H0(X~,(logE)X~2)H1(X~,(logE)X~1)H2(X~,𝒪X~).H^{2}(U,\mathbb C)\cong H^{0}\left(\widetilde{X},{}_{\widetilde{X}}^{2}(\log E)\right)\oplus H^{1}\left(\widetilde{X},{}_{\widetilde{X}}^{1}(\log E)\right)\oplus H^{2}\left(\widetilde{X},\mathscr{O}_{\widetilde{X}}\right).

Note that the inclusion H2(U,C)H2(X~,C)H^{2}(U,\mathbb C)\subset H^{2}(\widetilde{X},\mathbb C) differs only on the (1,1)(1,1)-pieces, as XX has rational singularities and πX~2\pi_{*}{}_{\widetilde{X}}^{2} is reflexive. Consider the long exact sequence

0H0(X~,)X~1H0(X~,(logE)X~1)H0(𝒪Ei)H1(X~,)X~1H1(X~,(logE)X~1)\begin{split}0\longrightarrow H^{0}\left(\widetilde{X},{}_{\widetilde{X}}^{1}\right)\longrightarrow H^{0}\left(\widetilde{X},{}_{\widetilde{X}}^{1}(\log E)\right)&\longrightarrow\bigoplusop\displaylimits H^{0}\left(\mathscr{O}_{E_{i}}\right)\\ &\longrightarrow H^{1}\left(\widetilde{X},{}_{\widetilde{X}}^{1}\right)\longrightarrow H^{1}\left(\widetilde{X},{}_{\widetilde{X}}^{1}(\log E)\right)\end{split}

By [KS21, Corollary 1.8], the first morphism is an isomorphism. The last morphism is surjective, as H1(X~,(logE)X~1)IH1,1(X)H^{1}(\widetilde{X},{}_{\widetilde{X}}^{1}(\log E))\xrightarrow{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}\operatorname{IH}^{1,1}(X), and so we have a surjection by the decomposition theorem. Therefore,

dimH1,1(X~)dimH1,1(U)=dimH0(Ei).\dim H^{1,1}\left(\widetilde{X}\right)-\dim H^{1,1}(U)=\sumop\displaylimits\dim H^{0}(E_{i}).

Now consider the inclusion H2(X,C)H2(X~,C)H^{2}(X,\mathbb C)\subset H^{2}(\widetilde{X},\mathbb C). Again, these vector spaces differ only in the (1,1)(1,1)-classes. The Leray spectral sequence induces the exact sequence

0H1(X,π)X~1H1(X~,)X~1H0(X,R1π)X~1.0\longrightarrow H^{1}\left(X,\pi_{*}{}_{\widetilde{X}}^{1}\right)\longrightarrow H^{1}\left(\widetilde{X},{}_{\widetilde{X}}^{1}\right)\longrightarrow H^{0}\left(X,R^{1}\pi_{*}{}_{\widetilde{X}}^{1}\right).

Thus the cokernel of the inclusion is exactly im(H2(X~,)X~1H0(R1π)X~1)\operatorname{im}(H^{2}(\widetilde{X},{}_{\widetilde{X}}^{1})\to H^{0}(R^{1}\pi_{*}{}_{\widetilde{X}}^{1})). But the symplectic form clearly gets killed in the morphism H2(X,C)H0(R2πC)H^{2}(X,\mathbb C)\to H^{0}(R^{2}\pi_{*}\mathbb C), so by Hodge theory

im(H1(X~,)X~1H0(R1π)X~1)=im(H2(X~,C)H0(R2πC)).\operatorname{im}\left(H^{1}\left(\widetilde{X},{}_{\widetilde{X}}^{1}\right)\longrightarrow H^{0}\left(R^{1}\pi_{*}{}_{\widetilde{X}}^{1}\right)\right)=\operatorname{im}\left(H^{2}\left(\widetilde{X},\mathbb C\right)\longrightarrow H^{0}\left(R^{2}\pi_{*}\mathbb C\right)\right).

We now see by (2.10) that XX is then Q\mathbb Q-factorial if and only if H2(X,C)IH2(X,C)H^{2}(X,\mathbb C)\cong\operatorname{IH}^{2}(X,\mathbb C). ∎

The results of the last two sections will be used to allow us to pass from a primitive symplectic variety to a bimeromorphic model with at worst Q\mathbb Q-factorial singularities to prove the results of this paper. Recall that a Q\mathbb Q-factorial terminalization of XX is a crepant morphism ϕ:ZX\phi\colon Z\to X from a Q\mathbb Q-factorial terminal variety ZZ. Such morphisms exist for projective varieties, see [BCHM10], and very general members of a locally trivial deformation of a primitive symplectic variety with b25b_{2}\geq 5; see [BL22, Corollary 9.2].

Corollary 2.18.

Suppose that XX is a primitive symplectic variety and ϕ:ZX\phi\colon Z\to X a Q\mathbb Q-factorial terminalization. Then ϕ\phi is semismall, and there is a canonical injection IH2(X,C)H2(Z,C)\operatorname{IH}^{2}(X,\mathbb C)\hookrightarrow H^{2}(Z,\mathbb C).

2.7.  Background on the LLV Algebra

To end this section, we review the construction of the LLV algebra and the LLV structure theorem for hyperkähler manifolds. This will also allow us to indicate the necessary pieces for an algebraic proof of the LLV structure theorem for intersection cohomology.

The total Lie algebra of a compact complex variety YY is the Lie algebra generated by all Lefschetz operators corresponding to hard Lefschetz, or HL, classes α\alpha:

(2.11) gtot(Y):=Lα,:ααisHL.\mathfrak g_{\mathrm{tot}}(Y):=\langle L_{\alpha},{}_{\alpha}:\leavevmode\nobreak\ \alpha\leavevmode\nobreak\ \mathrm{is\leavevmode\nobreak\ HL}\rangle.

Here, LαL_{\alpha} is the cupping operator with respect to the (1,1)(1,1)-class α\alpha, and α is its dual Lefschetz operator.

When XX is a hyperkähler manifold, we define the LLV algebra to be its total Lie algebra, and we write g=gtot(X)\mathfrak g=\mathfrak g_{\mathrm{tot}}(X).

Let gg be a hyperkähler metric on XX. Associated to (X,g)(X,g) are three differential forms ω1,ω2,ω3\omega_{1},\omega_{2},\omega_{3} which are Kähler forms with respect to gg. Let Wg:=ω1,ω2,ω3W_{g}:=\langle\omega_{1},\omega_{2},\omega_{3}\rangle be the three-space associated to this metric, and in fact a positive three-space with respect to the BBF form qXq_{X}, which means that qXq_{X} is positive-definite on WgW_{g}. Consider the algebra gg=Lωi,:ωiωiWg\mathfrak g_{g}=\langle L_{\omega_{i}},{}_{\omega_{i}}:\leavevmode\nobreak\ \omega_{i}\in W_{g}\rangle. Verbitsky [Ver90] showed that this algebra already has a structure generalizing the hard Lefschetz theorem for Kähler manifolds.

Theorem 2.19.

We have ggso(4,1)\mathfrak g_{g}\cong\mathfrak{so}(4,1).

The proof is as follows. Let 2n2n be the dimension of XX. The dual Lefschetz operators ωi{}_{\omega_{i}} are uniquely determined by the property that [Lωi,]ωi(α)=H(α)=(k2n)α[L_{\omega_{i}},{}_{\omega_{i}}](\alpha)=H(\alpha)=(k-2n)\alpha, where αHk(X,C)\alpha\in H^{k}(X,\mathbb C). Therefore, one can show that each ωi{}_{\omega_{i}} is the adjoint of LωiL_{\omega_{i}} with respect to the Hodge star operator, which depends only on the Kähler structure (X,g)(X,g). This then implies that the dual Lefschetz operators commute:

[,ωi]ωj=0.\left[{}_{\omega_{i}},{}_{\omega_{j}}\right]=0.

This is the main geometric input. The fact that ggso(4,1)\mathfrak g_{g}\cong\mathfrak{so}(4,1) follows from this geometric input, the hard Lefschetz theory for the ωi\omega_{i}, and the following additional commutator relations:

(2.12) Kij=Kji,[Kij,Kjk]=2Kik,[Kij,H]=0,[Kij,Lωj]=2Lωi,[Kij,]ωj=2,ωi[Kij,Lωk]=[Kij,]ωk=0,i,jk,\begin{split}&K_{ij}=-K_{ji},\quad\left[K_{ij},K_{jk}\right]=2K_{ik},\quad\left[K_{ij},H\right]=0,\\ &\left[K_{ij},L_{\omega_{j}}\right]=2L_{\omega_{i}},\quad\left[K_{ij},{}_{\omega_{j}}\right]=2{}_{\omega_{i}},\quad\left[K_{ij},L_{\omega_{k}}\right]=\left[K_{ij},{}_{\omega_{k}}\right]=0,\ i,j\neq k,\end{split}

where Kij:=[Lωi,]ωjK_{ij}:=[L_{\omega_{i}},{}_{\omega_{j}}]. Verbitsky observed in [Ver90] that KijK_{ij} acts as the Weil operator with respect to the complex structure induced by ωk\omega_{k} (and similarly for KjkK_{jk} and KkiK_{ki}). The commutator relations (2.12) follow from this key observation.

Verbitsky’s approach to the LLV structure theorem in his thesis [Ver96] was to then look at g\mathfrak g as the algebra generated by the gg\mathfrak g_{g} by varying the hyperkähler metric via the period map. The main technical input, observed by Verbitsky and also Looijenga–Lunts in [LL97], is that all the dual Lefschetz operators commute whenever they are defined. The following are the necessary pieces for obtaining the LLV structure theorem.

Theorem 2.20 (cf. [LL97, Theorem 4.5], [Ver96, Proposition 1.6]).

Let XX be a hyperkähler manifold.

  1. (1)

    For any two classes α,βH2(X,R)\alpha,\beta\in H^{2}(X,\mathbb R) satisfying hard Lefschetz, we have [,α]β=0[{}_{\alpha},{}_{\beta}]=0. Thus, the LLV algebra only exists in degrees 22, 0, and 2-2, and we get an eigenspace decomposition

    gg2g0g2\mathfrak g\cong\mathfrak g_{2}\oplus\mathfrak g_{0}\oplus\mathfrak g_{-2}

    with respect to the weight operator HH acting as the adjoint.

  2. (2)

    There are canonical isomorphisms g±2H2(X,R)\mathfrak g_{\pm 2}\cong H^{2}(X,\mathbb R) of  g\mathfrak g-modules.

  3. (3)

    There is a decomposition g0=g¯RH\mathfrak g_{0}=\overline{\mathfrak g}\oplus\mathbb R\cdot H, where g¯\overline{\mathfrak g} is the semisimple part. Moreover, g¯so(H2(X,R),qX)\overline{\mathfrak g}\cong\mathfrak{so}(H^{2}(X,\mathbb R),q_{X}), and g¯\overline{\mathfrak g} acts on H(X,R)H^{*}(X,\mathbb R) by derivations.

Theorem 2.20(1) follows from the fact that the collection of positive three-spaces WgW_{g} forms a dense open subset of the Grassmanian of three-spaces in H2(X,R)H^{2}(X,\mathbb R), whence the commutativity of the dual Lefschetz operators follows from the commutativity over the various WgW_{g}, and local Torelli. The decomposition holds since the direct sum g2g0g2\mathfrak g_{2}\oplus\mathfrak g_{0}\oplus\mathfrak g_{-2} is a Lie subalgebra of g\mathfrak g, which follows from Theorem 2.20(23). Indeed, the openness of the space of positive three-spaces in H2(X,R)H^{2}(X,\mathbb R) implies that the semisimple part is generated by the commutators [Lα,]β[L_{\alpha},{}_{\beta}] for HL classes α,β\alpha,\beta. If α,β\alpha,\beta come from a positive three-space WgW_{g}, then Verbitsky [Ver90, Lemma 2.2] shows that

(2.13) [Lα,]β(x)=i(pq)x,\left[L_{\alpha},{}_{\beta}\right](x)=i(p-q)x,

where xx is a (p,q)(p,q)-form with respect to the metric gg. But this certainly acts on H(X,C)H^{*}(X,\mathbb C) by derivations, and so every commutator [Lα,]β[L_{\alpha},{}_{\beta}] acts via derivations on the cohomology ring. This fact implies that [g¯,g±2]g±2[\overline{\mathfrak g},\mathfrak g_{\pm 2}]\subset\mathfrak g_{\pm 2}; since g2\mathfrak g_{2} and g2\mathfrak g_{-2} are abelian, this gives the eigenvalue decomposition. In order to prove that g¯so(H2(X,R),qX)\overline{\mathfrak g}\cong\mathfrak{so}(H^{2}(X,\mathbb R),q_{X}), we note that the semisimple part preserves cup product via derivation; the Fujiki relation (2.4) then implies that g¯\overline{\mathfrak g} preserves qXq_{X}, and so g¯so(H2(X,R),qX)\overline{\mathfrak g}\subset\mathfrak{so}(H^{2}(X,\mathbb R),q_{X}). The surjectivity follows by varying the so(4,1)\mathfrak{so}(4,1)-actions in the period domain since these generate the full so(H2(X,C),qX)\mathfrak{so}(H^{2}(X,\mathbb C),q_{X}).

We remark that our theorems for the LLV algebra are stated with rational coefficients, while the works of Looijenga–Lunts and Verbitsky work over real or complex coefficients. In the smooth case, it was observed in [GKLR22] that the LLV structure theorem holds over Q\mathbb Q, as the operators are all rationally defined. The same will hold for the singular version of the LLV algebra for the intersection cohomology of primitive symplectic varieties.

3.  Symplectic Symmetry on Intersection Cohomology

We prove that the canonical Hodge structure on IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q) inherited from the symplectic form σ\sigma satisfies the symplectic hard Lefschetz theorem, one of the main inputs in our algebraic proof of the LLV structure theorem.

3.1.  Degeneration of Hodge-to-de Rham on the Regular Locus

The first piece we need in proving the symplectic hard Lefschetz theorem is the degeneration of Hodge-to-de Rham on the regular locus. What is surprising here is that the degeneration holds with no restriction on the singularities of XX. We adapt a well-known trick to identify the Hodge-to-de Rham spectral sequence with the logarithmic Hodge-to-de Rham spectral sequence associated to a log-resolution of singularities.

Theorem 3.1.

Suppose XX is a proper symplectic variety of dimension 2n2n with regular locus UU and smooth singular locus . If the singular locus of  XX is smooth, then the Hodge-to-de Rham spectral sequence

E1p,q=Hq(U,)UpHp+q(U,C)E_{1}^{p,q}=H^{q}\left(U,{}_{U}^{p}\right)\Longrightarrow H^{p+q}(U,\mathbb C)

degenerates at E1E_{1} for p+q<2n1p+q<2n-1.

Proof.

Consider the logarithmic Hodge-to-de Rham spectral sequence

E1p,q=Hq(X~,(logE)X~p)Hp+q(U,C)E_{1}^{p,q}=H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{p}(\log E)\right)\Longrightarrow H^{p+q}(U,\mathbb C)

corresponding to a log-resolution of singularities π:X~X\pi\colon\widetilde{X}\to X with exceptional divisor EE, which degenerates at E1E_{1} for all p,qp,q; see [Del74, Section (9.2.3)]. It is enough to show that the restriction morphisms Hq(X~,(logE)X~p)Hq(U,)UpH^{q}(\widetilde{X},{}_{\widetilde{X}}^{p}(\log E))\to H^{q}(U,{}_{U}^{p}) are isomorphisms for p+q<2n1p+q<2n-1.

The restriction morphisms fit inside a long exact sequence

HEq(X~,(logE)X~p)Hq(X~,(logE)X~p)Hq(U,)Up.\cdots\longrightarrow H^{q}_{E}\left(\widetilde{X},{}_{\widetilde{X}}^{p}(\log E)\right)\longrightarrow H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{p}(\log E)\right)\longrightarrow H^{q}(U,{}_{U}^{p})\longrightarrow\cdots.

By local duality, there is an isomorphism

HEq(X~,(logE)X~p)H2nq(X~E,(logE)X~2np(E)),H_{E}^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{p}(\log E)\right)^{*}\cong H^{2n-q}\left(\widetilde{X}_{E},{}_{\widetilde{X}}^{2n-p}(\log E)(-E)\right),

where X~E\widetilde{X}_{E} is the completion of X~\widetilde{X} along EE. Now consider the Leray spectral sequence

E2r,s=Hr(X,Rsπ(logE)X~2np(E))Hr+s(X~E,(logE)X~2np(E)).E_{2}^{r,s}=H^{r}\left(X,R^{s}\pi_{*}{}_{\widetilde{X}}^{2n-p}(\log E)(-E)\right)\Longrightarrow H^{r+s}\left(\widetilde{X}_{E},{}_{\widetilde{X}}^{2n-p}(\log E)(-E)\right).

It is enough to show that E2r,sE_{2}^{r,s} vanishes in the range r+s>2n+1r+s>2n+1, which we note clearly holds when dim=0\dim\Sigma=0. To prove this, we claim that

Rqπ(logE)X~p(E)x=0R^{q}\pi_{*}{}_{\widetilde{X}}^{p}(\log E)(-E)_{x}=0

for p+q>2ndim,xp+q>2n-\dim{}_{x}, where x is the connected component of which contains xx. By Proposition 2.3, there is a product decomposition X^xYx×^x\widehat{X}_{x}\cong Y_{x}\times\widehat{\Sigma}_{x}, where YxY_{x} is a symplectic variety with isolated singularities. Since the claimed vanishing is local, we may assume that the log-resolution of singularities is π=πx×id\pi=\pi_{x}\times\operatorname{id},(2)(2)(2)Note that the complex 𝐑π(logE)X~p(E)\mathbf{R}\pi_{*}{}_{\tilde{X}}^{p}(\log E)(-E) is, up to quasi-isomorphism, independent of the choice of log-resolution π\pi. where πx:Yx~Yx\pi_{x}\colon\widetilde{Y_{x}}\to Y_{x} is a log-resolution of singularities of YxY_{x}. It follows that Rqπ(logE)X~p(E)=0R^{q}\pi_{*}{}_{\widetilde{X}}^{p}(\log E)(-E)=0 if Rq(πx)(logEx)Yx~p(Ex)=0R^{q}(\pi_{x})_{*}{}_{\widetilde{Y_{x}}}^{p}(\log E_{x})(-E_{x})=0, where ExE_{x} is the exceptional divisor of πx\pi_{x}.

To conclude, consider the terms E2r,sE_{2}^{r,s} for r+s>2n+1r+s>2n+1. On the one hand, E2r,s=0E_{2}^{r,s}=0 for r>dimr>\dim\Sigma, so we may assume otherwise. On the other hand, if we write =jj\Sigma=\coprodop\displaylimits_{j}{}_{j}, where j{}_{j}\subset\Sigma consists of the smooth components of of dimension jj, the preceding argument implies that SuppRsπ(logE)X~2np(E))j=\operatorname{Supp}\leavevmode\nobreak\ R^{s}\pi_{*}{}_{\widetilde{X}}^{2n-p}(\log E)(-E))\cap{}_{j}=\emptyset if s+(2np)>2njs+(2n-p)>2n-j. Putting this together, we can see that E2r,s=0E_{2}^{r,s}=0 implies that r+s2nr+s\leq 2n. ∎

An early draft claimed that this degeneration held for general primitive symplectic varieties—the following example indicates the gap in the previous argument when is not smooth; a new idea will be needed to extend this degeneration.

Example 3.2.

Let XX be a primitive symplectic variety, let X^x\widehat{X}_{x} and Xi^x\widehat{X_{i}^{\circ}}_{x} be the completions of XX and XiX_{i}^{\circ}, respectively, at xx, and let YxY_{x} be a symplectic variety. Then

Rqπ(logE)X~p(E)x=0,p+q>max{2n,2ndimXi+p}.R^{q}\pi_{*}{}_{\tilde{X}}^{p}(\log E)(-E)_{x}=0,\quad p+q>\max\left\{2n,2n-\dim X_{i}^{\circ}+p\right\}.

Consider the product decomposition X^xYx×Xi^x\widehat{X}_{x}\cong Y_{x}\times\widehat{X_{i}^{\circ}}_{x}. We note the claim is local and independent of the choice of π\pi, whence we may assume that π\pi is the resolution

(3.1) π^x×id:Yx~×Xi^xYx×Xi^x,\widehat{\pi}_{x}\times\operatorname{id}\colon\tilde{Y_{x}}\times\widehat{X_{i}^{\circ}}_{x}\longrightarrow Y_{x}\times\widehat{X_{i}^{\circ}}_{x},

where π^x:Y~xYx\widehat{\pi}_{x}\colon\tilde{Y}_{x}\to Y_{x} is a log-resolution of singularities with exceptional divisor ExE_{x}. On the one hand, there is a Künneth-type decomposition

(logE)Y~x×Xi^xp(E)p1+p2=p(logEx)Y~xp1(Ex),Xi^xp2{}_{\tilde{Y}_{x}\times\widehat{X_{i}^{\circ}}_{x}}^{p}(\log E)(-E)\cong\bigoplusop\displaylimits_{p_{1}+p_{2}=p}{}_{\tilde{Y}_{x}}^{p_{1}}(\log E_{x})(-E_{x})\otimes{}_{\widehat{X_{i}^{\circ}}_{x}}^{p_{2}},

where ExE_{x} is the π^x\widehat{\pi}_{x}-exceptional divisor. On the other hand, the qthq^{\mathrm{th}} higher direct image sheaf of (logEx)Y~xp1(Ex)Xi^xp2{}_{\tilde{Y}_{x}}^{p_{1}}(\log E_{x})(-E_{x})\otimes{}_{\widehat{X_{i}^{\circ}}_{x}}^{p_{2}} is just Rq(π^x)(logEx)Y~xp1(Ex)R^{q}(\widehat{\pi}_{x})_{*}{}_{\tilde{Y}_{x}}^{p_{1}}(\log E_{x})(-E_{x}) since we are taking the identity on the second factor. Therefore,

Rqπ(logE)X~p(E)xp1Rq(π^x)(logEx)Y~xp1(Ex).R^{q}\pi_{*}{}_{\tilde{X}}^{p}(\log E)(-E)_{x}\cong\bigoplusop\displaylimits_{p_{1}}R^{q}(\widehat{\pi}_{x})_{*}{}_{\tilde{Y}_{x}}^{p_{1}}(\log E_{x})(-E_{x}).

The vanishing then follows by Steenbrink vanishing applied to the lowest (logEx)Y~xp1(Ex){}_{\tilde{Y}_{x}}^{p_{1}}(\log E_{x})(-E_{x}).

3.2.  Hodge Theory of the Regular Locus

Next, we need to understand how the symplectic form of a primitive symplectic variety interacts with the (compactly supported) cohomology of the regular locus. The following lemma indicates how the symplectic form extends across the singularities of XX.

Lemma 3.3.

If  XX is a symplectic variety with smooth singular locus, the sheaves π(logE)X~p(E)\pi_{*}{}_{\widetilde{X}}^{p}(\log E)(-E) are reflexive for every 1p2n1\leq p\leq 2n.

Proof.

First assume that XX has isolated singularities. Since XX has rational singularities, it is enough to show that π(logE)X~p(E)πX~p\pi_{*}{}_{\widetilde{X}}^{p}(\log E)(-E)\hookrightarrow\pi_{*}{}_{\widetilde{X}}^{p} is an isomorphism for each 1p2n1\leq p\leq 2n. For p=2np=2n, this is immediate, and for p=2n1p=2n-1, this holds by [KS21, Theorem 1.6], where we note that Rn1π𝒪X~=0R^{n-1}\pi_{*}\mathscr{O}_{\widetilde{X}}=0. We may therefore assume that 1p2n21\leq p\leq 2n-2.

There is an exact complex

0(logE)X~p(E)X~pE(1)pE(2)pE(k)p,0\longrightarrow{}_{\widetilde{X}}^{p}(\log E)(-E)\longrightarrow{}_{\widetilde{X}}^{p}\longrightarrow{}_{E_{(1)}}^{p}\longrightarrow{}_{E_{(2)}}^{p}\longrightarrow\cdots\longrightarrow{}_{E_{(k)}}^{p}\longrightarrow\cdots,

where E(k)=|J|=kJEjE_{(k)}=\coprodop\displaylimits_{|J|=k}\bigcapop\displaylimits_{J}E_{j} is the pairwise union of the kk-fold intersections of the irreducible components of EE; see for example [MOP20, Lemma 4.1] and [MP22, Proof of Corollary 14.9]. In particular, there is a short exact sequence

0(logE)X~p(E)X~pp0,0\longrightarrow{}_{\widetilde{X}}^{p}(\log E)(-E)\longrightarrow{}_{\widetilde{X}}^{p}\longrightarrow\mathscr{M}_{p}\longrightarrow 0,

where p=ker(E(1)p)E(2)p\mathscr{M}_{p}=\ker({}_{E_{(1)}}^{p}\to{}_{E_{(2)}}^{p}). The lemma will follow if we can show that H0(p)=0H^{0}(\mathscr{M}_{p})=0. But this will follow from the Hodge theory of the exceptional divisor EE. Indeed, we have a complex

0Hq()E(1)pδ1p,qHq()E(2)pδ2p,q.0\longrightarrow H^{q}({}_{E_{(1)}}^{p})\xrightarrow{\delta_{1}^{p,q}}H^{q}({}_{E_{(2)}}^{p})\xrightarrow{\delta_{2}^{p,q}}\cdots.

As we have seen in Section 2.4, the mixed Hodge pieces of the cohomology of EE are then computed in terms of this complex:

grWp+qHp+q+r(E,C)p,q=kerδr+1p,q/imδrp,q.\operatorname{gr}_{W}^{p+q}H^{p+q+r}(E,\mathbb C)^{p,q}=\ker\delta_{r+1}^{p,q}/\operatorname{im}\delta_{r}^{p,q}.

In particular, grWpHp(E,C)p,0H0(p)\operatorname{gr}_{W}^{p}H^{p}(E,\mathbb C)^{p,0}\cong H^{0}(\mathscr{M}_{p}). But by Hodge symmetry, this is isomorphic to

grWpHp(E,C)0,pHp(E,𝒪E),\operatorname{gr}_{W}^{p}H^{p}(E,\mathbb C)^{0,p}\cong H^{p}(E,\mathscr{O}_{E}),

which vanishes by either [Nam01b, Lemma 1.2] or [MP22, Corollary 14.9] for 1p2n21\leq p\leq 2n-2.

More generally, suppose that the singular locus of XX is smooth. The problem is local and independent of the chosen resolution of singularities; we can then assume that π\pi is the resolution of singularities corresponding to a product decomposition given in Proposition 2.3. In particular, the sheaves π(logE)X~p(E)\pi_{*}{}_{\widetilde{X}}^{p}(\log E)(-E) are reflexive if and only if the sheaves (π^x)(logEx)Yx~p(Ex)(\widehat{\pi}_{x})_{*}{}_{\tilde{Y_{x}}}^{p}(\log E_{x})(-E_{x}) are reflexive for a transversal slice YxY_{x}. By assumption, YxY_{x} has at worst isolated singularities, and so the claim follows from the previous argument. ∎

In particular, the extension of the symplectic form σ\sigma gives a well-defined global section

σ~H0(X~,(logE)X~2(E))H0(X~,(logE)X~2),\tilde{\sigma}\in H^{0}\left(\widetilde{X},{}_{\widetilde{X}}^{2}(\log E)(-E)\right)\subset H^{0}\left(\widetilde{X},{}_{\widetilde{X}}^{2}(\log E)\right),

and so we get morphisms

(3.2) σ~p:(logE)X~np(logE)X~n+p,σ~p:(logE)X~np(E)(logE)X~n+p(E)\tilde{\sigma}^{p}\colon{}_{\widetilde{X}}^{n-p}(\log E)\longrightarrow{}_{\widetilde{X}}^{n+p}(\log E),\quad\tilde{\sigma}^{p}\colon{}_{\widetilde{X}}^{n-p}(\log E)(-E)\longrightarrow{}_{\widetilde{X}}^{n+p}(\log E)(-E)

which are induced by wedging. We emphasize that, unlike the corresponding morphisms σp:UnpUn+p\sigma^{p}\colon{}_{U}^{n-p}\to{}_{U}^{n+p}, the maps (3.2) are almost never isomorphisms. However, we note there is an important interaction between the sheaves (logE)X~p,(logE)X~2np(E){}_{\tilde{X}}^{p}(\log E),{}_{\tilde{X}}^{2n-p}(\log E)(-E) and the holomorphic extension σ~\tilde{\sigma}. The following lemma, which is a local computation, is a consequence of Lemma 3.3 and the canonical representation of the symplectic form on the regular locus.

Lemma 3.4.

Let XX be a symplectic variety of dimension 2n42n\geq 4 with isolated singularities and holomorphic symplectic form σH0(X,)X[2]\sigma\in H^{0}(X,{}_{X}^{[2]}). For 0pn0\leq p\leq n, there is a morphism

𝐑πσ~p:𝐑π(logE)X~np𝐑π(logE)X~n+p(E)\mathbf{R}\pi_{*}\widetilde{\sigma}^{p}\colon\mathbf{R}\pi_{*}{}_{\widetilde{X}}^{n-p}(\log E)\longrightarrow\mathbf{R}\pi_{*}{}_{\widetilde{X}}^{n+p}(\log E)(-E)

for any log-resolution of singularities π:X~X\pi\colon\widetilde{X}\to X, where σ~\widetilde{\sigma} is the unique extension of σ\sigma to H0(X~,)X~2H^{0}(\widetilde{X},{}_{\widetilde{X}}^{2}).

Proof.

For any xU:=Xregx\in U:=X_{\mathrm{reg}}, there exist an (analytic) neighborhood UxU_{x} of xx and local coordinates z1,,z2nz_{1},\ldots,z_{2n} such that σ\sigma, considered as a holomorphic symplectic form on UU, can be written as

σ=dz1dz2++dz2n1dz2n\sigma=dz_{1}\wedge dz_{2}+\cdots+dz_{2n-1}\wedge dz_{2n}

on UxU_{x}. By Lemma 3.3, there is a unique extension of σ\sigma to a global section σ~\widetilde{\sigma} of (logE)X~2(E){}_{\widetilde{X}}^{2}(\log E)(-E) for any log-resolution of singularities π:X~X\pi\colon\widetilde{X}\to X, which we describe over the exceptional divisor EE. For each point x0Ex_{0}\in E, choose local coordinates z1,,z2nz_{1}^{\prime},\ldots,z_{2n}^{\prime} of X~\widetilde{X} around x0x_{0} such that E=V(z1zk)E=V(z_{1}^{\prime}\cdots z_{k}^{\prime}) for some k2nk\leq 2n. Then

σ~=h(dz1dz2++dz2n1dz2n)\widetilde{\sigma}=h(dz_{1}^{\prime}\wedge dz_{2}^{\prime}+\cdots+dz_{2n-1}^{\prime}\wedge dz_{2n}^{\prime})

in a neighborhood VxV_{x} of x0x_{0}, where hH0(Vx,𝒪Vx)h\in H^{0}(V_{x},\mathscr{O}_{V_{x}}).

Consider the global morphism (3.2) σ~p:(logE)X~np(logE)X~n+p\widetilde{\sigma}^{p}\colon{}_{\widetilde{X}}^{n-p}(\log E)\to{}_{\widetilde{X}}^{n+p}(\log E) induced by wedging. Let α\alpha be a section of (logE)X~np(Vx){}_{\widetilde{X}}^{n-p}(\log E)(V_{x}), written as

(3.3) α=fdzi1zi1dzildzildzj1dzj1dzjm,1i1,,il2r, 2r+1j1,,jm2n,\alpha=f\frac{dz_{i_{1}}^{\prime}}{z_{i_{1}}^{\prime}}\wedge\cdots\wedge\frac{dz_{i_{l}}^{\prime}}{dz_{i_{l}}}\wedge dz_{j_{1}}^{\prime}\wedge dz_{j_{1}}^{\prime}\wedge\cdots\wedge dz_{j_{m}}^{\prime},\quad 1\leq i_{1},\ldots,i_{l}\leq 2r,\leavevmode\nobreak\ 2r+1\leq j_{1},\ldots,j_{m}\leq 2n,

where fH0(Vx,𝒪x)f\in H^{0}(V_{x},\mathscr{O}_{x}) and l+m=npl+m=n-p, and consider the logarithmic form ασ~p\alpha\wedge\widetilde{\sigma}^{p}. Note that if k>2n1k>2n-1, then σ~α\widetilde{\sigma}\wedge\alpha vanishes along EE. Assume without loss of generality that k=2rk=2r for rn1r\leq n-1. By Lemma 3.3, hh must vanish along z2r+1,,z2nz_{2r+1}^{\prime},\ldots,z_{2n}^{\prime}. It follows immediately from (3.3) that the logarithmic form

αhp|I|=pdzI,\alpha\wedge h^{p}\sumop\displaylimits_{|I|=p}dz_{I}^{\prime},

where I{1,,2n}I\subset\left\{1,\ldots,2n\right\} and dzIdz_{I}^{\prime} is the pp-fold wedge product of dzjdz_{j}^{\prime} for jIj\in I, must vanish along EE. The logarithmic wedging map (3.2) must factor as

(logE)X~np{{}_{\widetilde{X}}^{n-p}(\log E)}(logE)X~n+p{{}_{\widetilde{X}}^{n+p}(\log E)}(logE)X~n+p(E).{{}_{\widetilde{X}}^{n+p}(\log E)(-E)\hbox to0.0pt{.\hss}}σ~p\scriptstyle{\widetilde{\sigma}^{p}}

The claim follows by taking cohomology. ∎

3.3.  Symplectic Hard Lefschetz

Lemma 3.3 indicates what kind of zeros the powers σ~p\tilde{\sigma}^{p} of the extended symplectic form pick up across the exceptional divisor EE. Moreover, it says that the symplectic form on UU defines a class in the compactly supported cohomology, as

(3.4) σ~H0(X~,(logE)X~p(E))grW2Hc2(U,C)2,0,\tilde{\sigma}\in H^{0}\left(\widetilde{X},{}_{\widetilde{X}}^{p}(\log E)(-E)\right)\cong\operatorname{gr}_{W}^{2}H^{2}_{c}(U,\mathbb C)^{2,0},

by (2.6).

For convenience, we want to consider the case that XX has at worst terminal singularities. By passing to a Q\mathbb Q-factorial terminalization of some bimeromorphic model, we will see that this assumption is sufficient once we prove the LLV structure theorem.

We fix a log-resolution of singularities π:X~X\pi\colon\widetilde{X}\to X with exceptional divisor EE. If XX has isolated singularities, the intersection cohomology groups are given by the pure Hodge structures

IHk(X,C)={Hk(U,C),k<2n,im(Hck(U,C)Hk(U,C)),k=2n,Hck(U,C),k>2n;\operatorname{IH}^{k}(X,\mathbb C)=\begin{cases}H^{k}(U,\mathbb C),&k<2n,\\ \operatorname{im}(H^{k}_{c}(U,\mathbb C)\to H^{k}(U,\mathbb C)),&k=2n,\\ H^{k}_{c}(U,\mathbb C),&k>2n;\end{cases}

see [GM80, Section 6.1] and [Ste77, Corollary (1.14)]. For k<2nk<2n, the degeneration of the logarithmic Hodge-to-de Rham spectral sequence at E1E_{1} induces the Hodge filtration on the pure Hodge structure Hk(U,C)H^{k}(U,\mathbb C). As a consequence of Theorem 3.1, we see that the intersection cohomology of a primitive symplectic variety satisfies

IHp,q(X,C){Hq(U,)Up,p+q<2n1,Hcq(U,)Up,p+q>2n+1.\operatorname{IH}^{p,q}(X,\mathbb C)\cong\begin{cases}H^{q}(U,{}_{U}^{p}),&p+q<2n-1,\\ H^{q}_{c}(U,{}_{U}^{p}),&p+q>2n+1.\end{cases}

To see this, note that since Hk(U,Q)H^{k}(U,\mathbb Q) is a pure Hodge structure for k<2nk<2n, the natural morphism Hk(X~,Q)Hk(U,Q)H^{k}(\widetilde{X},\mathbb Q)\to H^{k}(U,\mathbb Q) is a surjective morphism of pure Hodge structures. The (p,q)(p,q)-part of the canonical Hodge structure on the cohomology of X~\widetilde{X} must factor through Hq(U,)UpH^{q}(U,{}_{U}^{p}), which injects by Theorem 3.1, and so we have Hq(U,)UpIHp,q(X)H^{q}(U,{}_{U}^{p})\xrightarrow{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}\operatorname{IH}^{p,q}(X) for p+q<2n1p+q<2n-1. By Poincaré duality, we get the statement concerning the compactly supported cohomology groups.

It turns out that this gives us enough of the intersection Hodge diamond to show that the symplectic symmetry holds. This is the key input to the construction of the LLV algebra.

Theorem 3.5.

Let XX be a primitive symplectic variety of dimension 2n2n with isolated singularities. For 0pn0\leq p\leq n and 0q2n0\leq q\leq 2n, the cupping map

Lσp:IHnp,q(X)IHn+p,q(X)L_{\sigma}^{p}\colon\operatorname{IH}^{n-p,q}(X)\longrightarrow\operatorname{IH}^{n+p,q}(X)

is an isomorphism.

Proof.

From the above discussion, the theorem holds for (np)+q<2n(n-p)+q<2n and (n+p)+q<2n(n+p)+q<2n since σp:Hq(U,)UnpHq(U,)Un+p\sigma^{p}\colon H^{q}(U,{}_{U}^{n-p})\to H^{q}(U,{}_{U}^{n+p}) is an isomorphism. Similarly, the theorem holds if (np)+q>2n(n-p)+q>2n and (n+p)+q>2n(n+p)+q>2n. We check the remaining cases.

First, suppose that (np)+q<2n1(n-p)+q<2n-1 and (n+p)+q>2n+1(n+p)+q>2n+1 (that is, we are mapping across middle cohomology). Since σ~\tilde{\sigma} defines a compactly supported global section on UX~U\subset\widetilde{X}, we have the factorization

Hq(U,)Unp{H^{q}(U,{}_{U}^{n-p})}Hcq(U,)Un+p{H^{q}_{c}(U,{}_{U}^{n+p})}Hq(U,)Un+q.{H^{q}(U,{}_{U}^{n+q})\hbox to0.0pt{.\hss}}σp\scriptstyle{\sigma^{p}}

The diagonal morphism is an isomorphism since UU is symplectic. Therefore, the cupping morphism will be an isomorphism if Hq(U,)UnpH^{q}(U,{}_{U}^{n-p}) and Hcq(U,)Un+pH_{c}^{q}(U,{}_{U}^{n+p}) have the same dimension. Writing

hp,q(U)=dimHq(U,)Up,hcp,q(U)=dimHcq(U,)Up,h^{p,q}(U)=\dim H^{q}(U,{}_{U}^{p}),\quad h_{c}^{p,q}(U)=\dim H^{q}_{c}(U,{}_{U}^{p}),

this follows since

(3.5) hcn+p,q(U)=hnp,2nq(U)=h2nq,np(U)=hq,np(U)=hnp,q(U).h^{n+p,q}_{c}(U)=h^{n-p,2n-q}(U)=h^{2n-q,n-p}(U)=h^{q,n-p}(U)=h^{n-p,q}(U).

We now need to consider the case that we map to or from IH2n1(X,C)H2n1(U,C)\operatorname{IH}^{2n-1}(X,\mathbb C)\cong H^{2n-1}(U,\mathbb C), where the proof of Theorem 3.1 breaks down. First assume that we map into H2n1(U,C)H^{2n-1}(U,\mathbb C). We need to show that the morphism

σ~p:Hq(X~,(logE)X~np)Hq(X~,(logE)X~n+p)\widetilde{\sigma}^{p}\colon H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{n-p}(\log E)\right)\longrightarrow H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{n+p}(\log E)\right)

is an isomorphism for (n+p)+q=2n1(n+p)+q=2n-1. Following the proof of the degeneration of Hodge-to-de Rham, we have shown that the restriction morphisms

Hq(X~,(logE)X~n+p)Hq(U,)Un+pH^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{n+p}(\log E)\right)\longrightarrow H^{q}\left(U,{}_{U}^{n+p}\right)

are at least injective in this range. Consider the commutative diagram

Hq(X~,(logE)X~np){H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{n-p}(\log E)\right)}Hq(U,)Unp{H^{q}\left(U,{}_{U}^{n-p}\right)}Hq(X~,(logE)X~n+p){H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{n+p}(\log E)\right)}Hq(U,)Un+p.{H^{q}\left(U,{}_{U}^{n+p}\right)\hbox to0.0pt{.\hss}}σ~p\scriptstyle{\tilde{\sigma}^{p}}σp\scriptstyle{\sigma^{p}}

The top morphism is an isomorphism by Theorem 3.1 unless p=0p=0, which we do not need to check, and the map σp\sigma^{p} is an isomorphism since UU is symplectic. This proves the restriction morphism is also an isomorphism for (n+p)+q=2n1(n+p)+q=2n-1, and therefore σ~p\widetilde{\sigma}^{p} is an isomorphism, too.

Now we consider the case that we map from IH2n1(X,C)\operatorname{IH}^{2n-1}(X,\mathbb C), where we necessarily map across middle cohomology. Then assume that (np)+q=2n1(n-p)+q=2n-1. Consider the commutative diagram

Hq(X~,(logE)X~np){H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{n-p}(\log E)\right)}Hq(U,)Unp{H^{q}\left(U,{}_{U}^{n-p}\right)}Hq(X~,(logE)X~n+p(E)){H^{q}\left(\widetilde{X},{}_{\widetilde{X}}^{n+p}(\log E)(-E)\right)}Hcq(U,)Un+p{H^{q}_{c}\left(U,{}_{U}^{n+p}\right)}Hq(U,)Un+p,{H^{q}\left(U,{}_{U}^{n+p}\right)\hbox to0.0pt{,\hss}}σ~p\scriptstyle{\tilde{\sigma}^{p}}σp\scriptstyle{\sigma^{p}}

where the left vertical morphism follows from Lemma 3.4. Again, we are using the fact that XX is terminal, so that Hq(U,)UnpHq(U,)Un+pH^{q}(U,{}_{U}^{n-p})\to H^{q}(U,{}_{U}^{n+p}) factors through Hcq(U,)Un+pH_{c}^{q}(U,{}_{U}^{n+p}). The top morphism is injective, and so the cupping map σ~p:Hq(X~,(logE)X~np)Hq(X~,(logE)X~n+p(E))\tilde{\sigma}^{p}\colon H^{q}(\widetilde{X},{}_{\widetilde{X}}^{n-p}(\log E))\to H^{q}(\widetilde{X},{}_{\widetilde{X}}^{n+p}(\log E)(-E)) is injective. A similar symmetry argument to (3.5) shows that these groups have the same dimension.

Finally, we are left to check the case that LσpL_{\sigma}^{p} maps into middle cohomology. These morphisms are of the form

Lσp:IHnp,np(X)IHn+p,np(X).L_{\sigma}^{p}\colon\operatorname{IH}^{n-p,n-p}(X)\longrightarrow\operatorname{IH}^{n+p,n-p}(X).

for p<np<n. By a similar commutative diagram argument to above, we can see that these maps must be injective. To prove surjectivity, consider a class aIHn+p,np(X)=(W2nH2n(U,C))n+p,npa\in\operatorname{IH}^{n+p,n-p}(X)=(W_{2n}H^{2n}(U,\mathbb C))^{n+p,n-p}. We note that the restriction map H2n(X~,C)W2nH2n(U,C)H^{2n}(\widetilde{X},\mathbb C)\to W_{2n}H^{2n}(U,\mathbb C) is surjective in this case; since UU is smooth, we may represent a=[α]a=[\alpha] as the class of an (n+p,np)(n+p,n-p)-form α\alpha which is ¯\overline{\partial}-closed. At the level of sheaves, we have isomorphisms

σp:𝒜Unp,np𝒜Un+p,np\sigma^{p}\colon\mathcal{A}^{n-p,n-p}_{U}\longrightarrow\mathcal{A}^{n+p,n-p}_{U}

sending a form β\beta to σpβ\sigma^{p}\wedge\beta. If α=σpβ\alpha=\sigma^{p}\wedge\beta, then 0=¯α=¯(σpβ)=σp¯β0=\overline{\partial}\alpha=\overline{\partial}(\sigma^{p}\wedge\beta)=\sigma^{p}\wedge\overline{\partial}\beta. But since σ\sigma is symplectic, ¯β=0\overline{\partial}\beta=0 and defines a class [β]Hnp(U,)Unp[\beta]\in H^{n-p}(U,{}_{U}^{n-p}). This proves the surjectivity.

To finish, note that the only remaining case is mapping from middle cohomology; but this follows from Poincaré duality. ∎

4.  Hard Lefschetz for Symplectic Varieties

In this section, we describe a hard Lefschetz theorem for the classes σ\sigma and σ¯\overline{\sigma} using Theorem 3.5, which inherits IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) with the structure of an sl2×sl2\mathfrak{sl}_{2}\times\mathfrak{sl}_{2}-representation. This data, along with the monodromy representation of the second cohomology, completely describes the g\mathfrak g-representation structure of the intersection cohomology.

4.1.  Symplectic Hard Lefschetz

Let LσL_{\sigma} be the cupping morphism with respect to the class of the symplectic form σ\sigma, and let sσ\mathfrak s_{\sigma} be the completed sl2\mathfrak{sl}_{2}-triple

sσ=Lσ,,σHσ,\mathfrak s_{\sigma}=\langle L_{\sigma},{}_{\sigma},H_{\sigma}\rangle,

where HσH_{\sigma} is the weight operator of the corresponding weight decomposition.

The isomorphisms Lσp:IHnp,q(X)IHn+p,q(X)L_{\sigma}^{p}\colon\operatorname{IH}^{n-p,q}(X)\xrightarrow{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}\operatorname{IH}^{n+p,q}(X) induce a primitive decomposition on the intersection cohomology classes. A class αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X) for pqp\leq q is said to be σ\sigma-primitive if Lσnp+1α=0L_{\sigma}^{n-p+1}\alpha=0. We denote the subspace of primitive intersection (p,q)(p,q)-classes by IHσp,q(X)IHp,q(X)\operatorname{IH}_{\sigma}^{p,q}(X)\subset\operatorname{IH}^{p,q}(X). As in the case of classic hard Lefschetz, there is a σ\sigma-primitive decomposition

IHp,q(X)=jmax{pn,0}LσjIHσp2j,q(X).\operatorname{IH}^{p,q}(X)=\bigoplusop\displaylimits_{j\geq\mathrm{max}\left\{p-n,0\right\}}L_{\sigma}^{j}\operatorname{IH}_{\sigma}^{p-2j,q}(X).

While the hard Lefschetz theorem is a result of Kähler geometry, the relationship between the primitive decomposition theorem and the weight decomposition is purely algebraic. In particular, we get the following.

Corollary 4.1.

Let XX be a primitive symplectic variety with isolated singularities. For αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X), we have Hσ(α)=(pn)αH_{\sigma}(\alpha)=(p-n)\alpha.

Proof.

The dual Lefschetz operator σ can be uniquely defined to satisfy [Lσ,]σ=(pn)α[L_{\sigma},{}_{\sigma}]=(p-n)\alpha via the primitive decomposition. In particular, if α=jLσjασp2j,k\alpha=\sumop\displaylimits_{j}L_{\sigma}^{j}\alpha_{\sigma}^{p-2j,k} is the σ\sigma-primitive decomposition of a class αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X), then

(α)σ:=jj(np+j+1)Lσj1ασp2j,q.{}_{\sigma}(\alpha):=\sumop\displaylimits_{j}j(n-p+j+1)L_{\sigma}^{j-1}\alpha_{\sigma}^{p-2j,q}.

One checks that this operator acts on the intersection cohomology module as a degree (2,0)(-2,0) and satisfies the desired commutativity relation. See [Kle68, Section 1.4] for more details. ∎

Now consider the cupping operator Lσ¯:IHp,q(X)IHp,q+2(X)L_{\overline{\sigma}}\colon\operatorname{IH}^{p,q}(X)\to\operatorname{IH}^{p,q+2}(X). By Hodge symmetry, it induces isomorphisms Lσ¯q:IHp,nq(X)IHp,n+q(X)L_{\overline{\sigma}}^{q}\colon\operatorname{IH}^{p,n-q}(X)\xrightarrow{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}\operatorname{IH}^{p,n+q}(X). Considering the induced sl2\mathfrak{sl}_{2}-triple

sσ¯=Lσ¯,,σ¯Hσ¯,\mathfrak s_{\overline{\sigma}}=\langle L_{\overline{\sigma}},{}_{\overline{\sigma}},H_{\overline{\sigma}}\rangle,

it then follows that the corresponding weight operator satisfies Hσ¯=(qn)αH_{\overline{\sigma}}=(q-n)\alpha for αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X). As in the holomorphic case, a class αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X) with qnq\leq n is σ¯\overline{\sigma}-primitive if Lσ¯nq+1α=0L_{\overline{\sigma}}^{n-q+1}\alpha=0. If IHσ¯p,q(X)IHp,q(X)\operatorname{IH}_{\overline{\sigma}}^{p,q}(X)\subset\operatorname{IH}^{p,q}(X) is the subspace of σ¯\overline{\sigma}-primitive (p,q)(p,q)-classes, there is also a decomposition

IHp,q(X)=kmax{qn,0}Lσ¯kIHσ¯p,q2k(X).\operatorname{IH}^{p,q}(X)=\bigoplusop\displaylimits_{k\geq\mathrm{max}\left\{q-n,0\right\}}L_{\overline{\sigma}}^{k}\,\operatorname{IH}_{\overline{\sigma}}^{p,q-2k}(X).

If α=kLσ¯kασ¯p,q2k\alpha=\sumop\displaylimits_{k}L_{\overline{\sigma}}^{k}\,\alpha_{\overline{\sigma}}^{p,q-2k} is a σ¯\overline{\sigma}-primitive decomposition, then we have

(α)σ¯=kk(nq+k+1)Lσ¯k1ασ¯p,q2k.{}_{\overline{\sigma}}(\alpha)=\sumop\displaylimits_{k}k(n-q+k+1)L_{\overline{\sigma}}^{k-1}\alpha_{\overline{\sigma}}^{p,q-2k}.

As in case of compact Kähler manifolds, a class αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X) is σ\sigma-primitive (resp. σ¯\overline{\sigma}-primitive) if (α)σ=0{}_{\sigma}(\alpha)=0 (resp. (α)σ¯=0{}_{\overline{\sigma}}(\alpha)=0).

Proposition 4.2.

We have [,σ]σ¯=0[{}_{\sigma},{}_{\overline{\sigma}}]=0.

Proof.

Let αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X). We get two primitive decompositions with respect to σ\sigma and σ¯\overline{\sigma}:

α=jLσjασp2j,q=kLσ¯kασ¯p,q2k,\alpha=\sumop\displaylimits_{j}L_{\sigma}^{j}\alpha_{\sigma}^{p-2j,q}=\sumop\displaylimits_{k}L_{\overline{\sigma}}^{k}\,\alpha_{\overline{\sigma}}^{p,q-2k},

where ασp2j,q\alpha_{\sigma}^{p-2j,q} is σ\sigma-primitive for each jj and ασ¯p,q2k\alpha_{\overline{\sigma}}^{p,q-2k} is σ¯\overline{\sigma}-primitive for each kk.

Suppose that α\alpha is σ\sigma-primitive, so that pnp\leq n and Lσnp+1α=0L_{\sigma}^{n-p+1}\alpha=0. Then

0=Lσnp+1α=jLσnp+1Lσ¯kασ¯p,q2k=kLσ¯k(Lσnp+1ασ¯p,q2k).0=L_{\sigma}^{n-p+1}\alpha=\sumop\displaylimits_{j}L_{\sigma}^{n-p+1}L_{\overline{\sigma}}^{k}\,\alpha_{\overline{\sigma}}^{p,q-2k}=\sumop\displaylimits_{k}L_{\overline{\sigma}}^{k}\left(L_{\sigma}^{n-p+1}\alpha_{\overline{\sigma}}^{p,q-2k}\right).

If ασ¯p,q2k\alpha_{\overline{\sigma}}^{p,q-2k} is σ¯\overline{\sigma}-primitive of degree q2kq-2k, then Lσnp+1ασ¯p,q2kL_{\sigma}^{n-p+1}\alpha_{\overline{\sigma}}^{p,q-2k} is also σ¯\overline{\sigma}-primitive of degree q2kq-2k, as the antiholomorphic degree is not changing and the cup products commute. Therefore,

kLσ¯k(Lσnp+1ασ¯p,q2k)\sumop\displaylimits_{k}L_{\overline{\sigma}}^{k}\left(L_{\sigma}^{n-p+1}\alpha_{\overline{\sigma}}^{p,q-2k}\right)

is a σ¯\overline{\sigma}-primitive decomposition of Lnp+1α=0L^{n-p+1}\alpha=0, and so Lσ¯kLσnp+1ασ¯p,q2k=0L_{\overline{\sigma}}^{k}\,L_{\sigma}^{n-p+1}\alpha_{\overline{\sigma}}^{p,q-2k}=0 for each kk. This implies that Lσnp+1ασ¯p,q2k=0L^{n-p+1}_{\sigma}\alpha_{\overline{\sigma}}^{p,q-2k}=0, as Lσ¯kL^{k}_{\overline{\sigma}} is injective over antiholomorphic degree q2kq-2k, and so ασ¯p,q2k\alpha_{\overline{\sigma}}^{p,q-2k} is both σ¯\overline{\sigma}- and σ\sigma-primitive for each kk. Similarly, ασp2j,q\alpha_{\sigma}^{p-2j,q} is both σ\sigma- and σ¯\overline{\sigma}-primitive for each jj if α\alpha is σ¯\overline{\sigma}-primitive.

We have shown that any (p,q)(p,q)-class αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X) admits a simultaneous σ\sigma- and σ¯\overline{\sigma}-primitive decomposition. This implies that the operators σ and σ¯{}_{\overline{\sigma}} commute. ∎

Remark 4.3.

We note that the symplectic hard Lefschetz theory also holds for even-dimensional complex tori, as they admit a symplectic form. Looijenga–Lunts observed that the total Lie algebra of a complex torus detects the Hodge theory, see [LL97, Section 3], but the LLV algebra has a different structure than a compact hyperkähler manifold. This is because the Torelli theorem for a complex torus does not differentiate between it and its dual.

4.2.  Hard Lefschetz for Non-Isotropic Classes

For the rest of the section, we will assume that XX is a primitive symplectic variety with isolated singularities and b25b_{2}\geq 5. In particular, the global moduli theory of Bakker–Lehn [BL22] applies. In [Tig23b], the author shows that IH1,1(X)\operatorname{IH}^{1,1}(X) parametrizes isomorphism classes of certain deformations of XX, where we allow smoothing of the singularities; these deformations are unobstructed, and we get a global Torelli theorem in terms of the full intersection cohomology. The results of Section 2 will allow us to avoid developing this theory here.

We now turn to classic hard Lefschetz. Up to some locally trivial deformation, the classes σ,σ¯H2(X,C)\sigma,\overline{\sigma}\in H^{2}(X,\mathbb C) define real cohomology classes γ=R(σ)\gamma=\mathfrak{R}(\sigma) and γ=I(σ)\gamma^{\prime}=\mathfrak{I}(\sigma), corresponding to the real and imaginary parts of σ\sigma:

γ=σ+σ¯,γ=i(σσ¯).\gamma=\sigma+\overline{\sigma},\quad\gamma^{\prime}=-i(\sigma-\overline{\sigma}).

There are cupping maps Lγ,LγL_{\gamma},L_{\gamma^{\prime}} on the intersection cohomology module by the proof of Proposition 2.11; these morphisms, considered as nilpotent operators in gl(IH(X,Q))\mathfrak{gl}(\operatorname{IH}^{*}(X,\mathbb Q)), complete to sl2\mathfrak{sl}_{2}-triples

sγ=Lγ,,γHγ,sγ=Lγ,,γHγ.\mathfrak{s}_{\gamma}=\langle L_{\gamma},{}_{\gamma},H_{\gamma}\rangle,\quad\mathfrak{s}_{\gamma^{\prime}}=\langle L_{\gamma^{\prime}},{}_{\gamma^{\prime}},H_{\gamma^{\prime}}\rangle.
Proposition 4.4.

The cohomology classes γ\gamma and γ\gamma^{\prime} are HL.

Proof.

By assumption qX(γ),qX(γ)0q_{X}(\gamma),q_{X}(\gamma^{\prime})\neq 0. Therefore, this follows from Proposition 2.11. ∎

In particular, the weight operators satisfy Hγ=Hγ=HH_{\gamma}=H_{\gamma^{\prime}}=H, where HH acts as (k2n)id(k-2n)\operatorname{id} on IHk(X,R)\operatorname{IH}^{k}(X,\mathbb R).

Corollary 4.5.

We have [,γ]γ=0[{}_{\gamma},{}_{\gamma^{\prime}}]=0.

Proof.

Note that γ and γ{}_{\gamma^{\prime}} act as

=γ+σ,σ¯=γi(σ)σ¯,{}_{\gamma}={}_{\sigma}+{}_{\overline{\sigma}},\quad{}_{\gamma^{\prime}}=i({}_{\sigma}-{}_{\overline{\sigma}}),

as they are defined by the commutator relation H=[Lγ,]γ=[Lγ,]γH=[L_{\gamma},{}_{\gamma}]=[L_{\gamma^{\prime}},{}_{\gamma^{\prime}}]. Taking the commutator, the result follows since [,σ]σ¯=0[{}_{\sigma},{}_{\overline{\sigma}}]=0 by Proposition 4.2. ∎

We therefore have a pair of non-isotropic classes (γ,γ)(\gamma,\gamma^{\prime}) such that qX(γ,γ)=0q_{X}(\gamma,\gamma^{\prime})=0 whose dual Lefschetz operators commute. As we will see, it is important that this pair vanishes under the induced bilinear form, as this defines a Zariski-closed condition on the space of non-isotropic pairs.

Corollary 4.6.

We have [Lσ,]σ¯=[Lσ¯,]σ=0[L_{\sigma},{}_{\overline{\sigma}}]=-[L_{\overline{\sigma}},{}_{\sigma}]=0.

Proof.

Note that

[Lγ,]γ=[Lσ,]σ+[Lσ,]σ¯+[Lσ¯,]σ+[Lσ¯,]σ¯[L_{\gamma},{}_{\gamma}]=[L_{\sigma},{}_{\sigma}]+[L_{\sigma},{}_{\overline{\sigma}}]+[L_{\overline{\sigma}},{}_{\sigma}]+[L_{\overline{\sigma}},{}_{\overline{\sigma}}]

and

[Lγ,]γ=[Lσ,]σ[Lσ,]σ¯[Lσ¯,]σ+[Lσ¯,]σ¯.[L_{\gamma^{\prime}},{}_{\gamma^{\prime}}]=[L_{\sigma},{}_{\sigma}]-[L_{\sigma},{}_{\overline{\sigma}}]-[L_{\overline{\sigma}},{}_{\sigma}]+[L_{\overline{\sigma}},{}_{\overline{\sigma}}].

Either equation implies that [Lσ,]σ¯=[Lσ¯,]σ[L_{\sigma},{}_{\overline{\sigma}}]=-[L_{\overline{\sigma}},{}_{\sigma}] since ([Lσ,]σ+[Lσ¯,]σ¯)(α)=(p+q2n)α([L_{\sigma},{}_{\sigma}]+[L_{\overline{\sigma}},{}_{\overline{\sigma}}])(\alpha)=(p+q-2n)\alpha for αIHp,q\alpha\in\operatorname{IH}^{p,q} (see Corollary 4.1 and the following paragraph). Since [Lγ,]γ=[Lγ,]γ[L_{\gamma},{}_{\gamma}]=[L_{\gamma^{\prime}},{}_{\gamma^{\prime}}], subtracting the second equation from the first implies [Lσ,]σ¯=[Lσ¯,]σ=0[L_{\sigma},{}_{\overline{\sigma}}]=[L_{\overline{\sigma}},{}_{\sigma}]=0. ∎

Corollary 4.7.

For αIHp,q(X)\alpha\in\operatorname{IH}^{p,q}(X), we have [Lγ,]γ=i(pq)α[L_{\gamma},{}_{\gamma^{\prime}}]=i(p-q)\alpha.

Proof.

Note that i[Lγ,]γ=[Lσ,]σ[Lσ,]σ¯+[Lσ¯,]σ[Lσ¯,]σ¯-i[L_{\gamma},{}_{\gamma^{\prime}}]=[L_{\sigma},{}_{\sigma}]-[L_{\sigma},{}_{\overline{\sigma}}]+[L_{\overline{\sigma}},{}_{\sigma}]-[L_{\overline{\sigma}},{}_{\overline{\sigma}}]. The result follows since the two middle terms vanish by Corollary 4.6. ∎

Corollary 4.7 is a generalization of Verbitsky’s generalization of hard Lefschetz on compact hyperkähler manifolds, see [Ver90]: If (X,g,I,J,K)(X,g,I,J,K) is a compact hyperkähler manifold with hyperkähler metric gg and complex structures I,J,KI,J,K, the Weil operators Cσ=[Lωλ,]ωλ=i(pq)idC_{\sigma}=[L_{\omega_{\lambda}},{}_{\omega_{\lambda}}]=i(p-q)\operatorname{id} are contained in the algebras ggso(4,1)\mathfrak g_{g}\cong\mathfrak{so}(4,1) (see Theorem 2.19).

5.  The LLV Algebra for Intersection Cohomology

In this section, we prove the LLV structure theorem for intersection cohomology. We do this as follows: First, we prove the theorem for Q\mathbb Q-factorial singularities, which will allow us to consider all HL elements of H2(X,C)H^{2}(X,\mathbb C) and use the monodromy density theorem under the assumption b25b_{2}\geq 5. We then show that the theorem holds in general by passing to a Q\mathbb Q-factorial terminalization, using the semismallness of such crepant morphisms; see Corollary 2.18.

The key here is to observe that we have a pair (γ,γ)(\gamma,\gamma^{\prime}) of non-isotropic vectors such that qX(γ,γ)=0q_{X}(\gamma,\gamma^{\prime})=0. As we will see below, the second condition is necessary. In [Huy01, Theorem 2.2], it was shown that the commutativity of dual Lefschetz operators holds infinitesimally, and so we always have a pair of non-isotropic pairs which commute. The point though is that we get a variety LL, described by Zariski-closed conditions, for which the monodromy group acts.

5.1.  Q\mathbb Q-Factorial Terminal Case

As in the smooth case, we define the LLV algebra in terms of HL classes in intersection cohomology.

Definition 5.1.

If XX is a primitive symplectic variety, the (intersection) LLV algebra is

g:=gtot(X)=Lα,|ααIH2(X,Q)isHL.\mathfrak g:=\mathfrak g_{\mathrm{tot}}(X)=\langle L_{\alpha},{}_{\alpha}|\leavevmode\nobreak\ \alpha\in\operatorname{IH}^{2}(X,\mathbb Q)\leavevmode\nobreak\ \mathrm{is\leavevmode\nobreak\ HL}\rangle.

Now assume that XX is a primitive symplectic variety with at worst Q\mathbb Q-factorial terminal singularities. By Proposition 2.17, the LLV algebra is generated by the dual Lefschetz operators corresponding to HL classes in H2(X,Q)H^{2}(X,\mathbb Q).

5.1.1.  Commutativity of the dual Lefschetz operators

Recall that by Corollary 4.5, there exists a point (γ,γ)H2(X,R)×2(\gamma,\gamma^{\prime})\in H^{2}(X,\mathbb R)^{\times 2} of non-isotropic vectors such that [,γ]γ=0[{}_{\gamma},{}_{\gamma^{\prime}}]=0. Moreover, we have qX(γ,γ)=0q_{X}(\gamma,\gamma^{\prime})=0, where qX(,)q_{X}(-,-) is the induced bilinear form with respect to qXq_{X}. The upshot is that the monodromy group acts on the space of such non-isotropic pairs.

Theorem 5.2.

Let XX be a Q\mathbb Q-factorial primitive symplectic variety with isolated singularities and b25b_{2}\geq 5. For any pair of non-isotropic classes α,βH2(X,Q)\alpha,\beta\in H^{2}(X,\mathbb Q), we have [,α]β=0[{}_{\alpha},{}_{\beta}]=0.

Proof.

Let XX be a primitive symplectic variety with Q\mathbb Q-factorial terminal isolated singularities, so that Theorem 3.5 and the Lefschetz theory of Section 4.1 hold. If σ\sigma is the class of the symplectic form, let γ=R(σ)\gamma=\mathfrak R(\sigma) and γ=I(σ)\gamma^{\prime}=\mathfrak I(\sigma). We have seen that [,γ]γ=0[{}_{\gamma},{}_{\gamma^{\prime}}]=0, which is a Zariski-closed condition. Consider the space

L={(α,β)H2(X,C)×H2(X,C)qX(α),qX(β)0,qX(α,β)=0}.L=\left\{(\alpha,\beta)\in H^{2}(X,\mathbb C)\times H^{2}(X,\mathbb C)\mid q_{X}(\alpha),q_{X}(\beta)\neq 0,\leavevmode\nobreak\ \leavevmode\nobreak\ q_{X}(\alpha,\beta)=0\right\}.

The group SO()C\operatorname{SO}({}_{\mathbb C}) acts on LL diagonally. If Mon(X)O()\operatorname{Mon}(X)\subset O(\Gamma) is the monodromy group associated to XX (see Remark 2.7), let GX:=SO()Mon(X)SO()CG_{X}:=\operatorname{SO}(\Gamma)\cap\operatorname{Mon}(X)\subset\operatorname{SO}({}_{\mathbb C}). Note that the (γ,γ)(\gamma,\gamma^{\prime})-orbit of GXG_{X} preserves the commutator relation: If gGXg\in G_{X}, then [,gγ]gγ=0[{}_{g\cdot\gamma},{}_{g\cdot\gamma^{\prime}}]=0. Indeed, the dual Lefschetz operators satisfy the property

=gγgg1γ{}_{g\cdot\gamma}=g{}_{\gamma}g^{-1}

as they are uniquely determined by [Lgγ,]gγ=(k2n)id[L_{g\cdot\gamma},{}_{g\cdot\gamma}]=(k-2n)\operatorname{id}, and the monodromy group is invariant under conjugation. By the Borel density theorem (see [Fur76]), GXG_{X} is Zariski dense in SO()C\operatorname{SO}({}_{\mathbb C}), and [,gγ]gγ=0[{}_{g\cdot\gamma},{}_{g\cdot\gamma^{\prime}}]=0 for every gSO()Cg\in\operatorname{SO}({}_{\mathbb C}). The commutativity then holds for every (α,β)L(\alpha,\beta)\in L. ∎

Thus, the LLV algebra g\mathfrak g only contains elements of degrees 2,0,22,0,-2. As in Section 2.7, we define g2\mathfrak g_{2} as the subalgebra generated by the LαL_{\alpha}, g2\mathfrak g_{-2} as the subalgebra generated by the α, and g0\mathfrak g_{0} as the degree 0 piece which is (necessarily) of the form

g0g¯×QH.\mathfrak g_{0}\cong\overline{\mathfrak g}\times\mathbb Q\cdot H.

5.1.2.  g¯\overline{\mathfrak g} acts via derivations

We want to show that gg2g0g2\mathfrak g\cong\mathfrak g_{2}\oplus\mathfrak g_{0}\oplus\mathfrak g_{-2}. As in the smooth case, this will be done if we can show that g¯\overline{\mathfrak g} acts via derivations on intersection cohomology.

Proposition 5.3.

The semisimple part g¯\overline{\mathfrak g} acts via derivations.

Proof.

As g¯\overline{\mathfrak g} is generated by the commutators [Lα,]β[L_{\alpha},{}_{\beta}], it is enough to show that these elements act on the intersection cohomology module via derivations. Consider again the point (γ,γ)L(\gamma,\gamma^{\prime})\in L as in the proof of Theorem 5.2. The commutator satisfies [Lγ,]γ(x)=i(pq)x[L_{\gamma},{}_{\gamma^{\prime}}](x)=i(p-q)x for a (p,q)(p,q)-class xx. While this identity is not preserved by the restricted monodromy group GXG_{X}, we note that [Lγ,]γ[L_{\gamma},{}_{\gamma^{\prime}}] acts on IH(X,R)\operatorname{IH}^{*}(X,\mathbb R) via derivations, and this property is preserved via GXG_{X}. Specifically, for any gGXg\in G_{X}, [Lgγ,]gγ[L_{g\cdot\gamma},{}_{g\cdot\gamma^{\prime}}] also acts via derivations. But by the monodromy density theorem, this must be true for any [Lα,]β[L_{\alpha},{}_{\beta}], too. ∎

Thus, we get the eigenvalue decomposition

(5.1) gg2g0g2.\mathfrak g\cong\mathfrak g_{2}\oplus\mathfrak g_{0}\oplus\mathfrak g_{-2}.

5.1.3.  LLV structure theorem

We can now prove the LLV structure theorem for Q\mathbb Q-factorial singularities.

Theorem 5.4.

Let XX be a Q\mathbb Q-factorial primitive symplectic variety with isolated singularities with LLV algebra g\mathfrak g and b25b_{2}\geq 5. Then

gso((H2(X,Q),qX)h),\mathfrak g\cong\mathfrak{so}((H^{2}(X,\mathbb Q),q_{X})\oplus\mathfrak h),

where qXq_{X} is the BBF form and h\mathfrak h is a hyperbolic plane.

Another way of stating Theorem 5.4 is in terms of the Mukai completion. If we write

(H2(X,Q),qX)h=:(H~2(X,Q),q~X),(H^{2}(X,\mathbb Q),q_{X})\oplus\mathfrak h=:\left(\widetilde{H}^{2}(X,\mathbb Q),\widetilde{q}_{X}\right),

the Mukai completion (H~2(X,Q),q~X)(\widetilde{H}^{2}(X,\mathbb Q),\widetilde{q}_{X}) inherits a bilinear structure and Hodge structure which is compatible with (H2(X,Q),qX)(H^{2}(X,\mathbb Q),q_{X}). Theorem 5.4 then gives an isomorphism

gso(H~2(X,Q),q~X).\mathfrak g\cong\mathfrak{so}\left(\widetilde{H}^{2}(X,\mathbb Q),\widetilde{q}_{X}\right).

In the smooth case, the subalgebra g¯:=so(H2(X,Q),qX)\overline{\mathfrak g}:=\mathfrak{so}(H^{2}(X,\mathbb Q),q_{X}) corresponds to the semisimple part of the decomposition (5.1). Moreover, the completion (H2(X,Q),qX)(H~2(X,Q),q~X)(H^{2}(X,\mathbb Q),q_{X})\hookrightarrow(\widetilde{H}^{2}(X,\mathbb Q),\widetilde{q}_{X}) is compatible with the extension g¯g\overline{\mathfrak g}\hookrightarrow\mathfrak g by branching (see [GKLR22, Section B.2.1]). It is then sufficient to prove the structure theorem for the semisimple part.

Proof.

We will show that

(5.2) g¯so(H2(X,Q),qX).\overline{\mathfrak g}\cong\mathfrak{so}(H^{2}(X,\mathbb Q),q_{X}).

Since g¯\overline{\mathfrak g} acts by derivations and preserves cup product, we see that it preserves the BBF form qXq_{X}. Therefore, there is an injective map g¯so(H2(X,C),qX)\overline{\mathfrak g}\hookrightarrow\mathfrak{so}(H^{2}(X,\mathbb C),q_{X}). The surjectivity is a reproduction of Verbitsky’s result. Consider once more the point (γ,γ)L(\gamma,\gamma^{\prime})\in L. By global (or local) Torelli, this point completes to a positive three-space Wσ=α,γ,γW_{\sigma}=\langle\alpha,\gamma,\gamma^{\prime}\rangle for which qXq_{X} is positive-definite. Now Theorem 5.2 shows that the dual Lefschetz operators corresponding to WσW_{\sigma} all commute. Moreover, we can extend Corollary 4.7 to the entire three-space and show that the commutators [Lα,]γ[L_{\alpha},{}_{\gamma}] and [Lα,]γ[L_{\alpha},{}_{\gamma^{\prime}}] acts as the Weil operators with respect to the induced complex structure, after possibly passing to a locally trivial deformation. It then follows that

gσ:=Lω,:ωωWσ\mathfrak g_{\sigma}:=\langle L_{\omega},{}_{\omega}:\leavevmode\nobreak\ \omega\in W_{\sigma}\rangle

satisfies Verbitsky’s commutator relations (2.12). The upshot is that gσso(4,1)\mathfrak g_{\sigma}\cong\mathfrak{so}(4,1). Using monodromy density once more, we then see that any HL class xH2(X,R)x\in H^{2}(X,\mathbb R) completes to a positive three-space WxW_{x}, and the corresponding algebra satisfies gxso(4,1)\mathfrak g_{x}\cong\mathfrak{so}(4,1). The surjectivity then follows as the Lie algebra so(H2(X,C),qX)\mathfrak{so}(H^{2}(X,\mathbb C),q_{X}) is generated by the simultaneous so(4,1)\mathfrak{so}(4,1)-actions. ∎

5.2.  General Case

We now move to a general primitive symplectic variety. In order for the LLV structure theorem to make sense, we need to define a BBF form for intersection cohomology. The result will then follow from the representation theory of so(m)\mathfrak{so}(m).

5.2.1.  A BBF form on intersection cohomology

Definition 5.5.

Let XX be a primitive symplectic variety of dimension 2n2n and π:X~X\pi\colon\widetilde{X}\to X a resolution of singularities. For αIH2(X,C)\alpha\in\operatorname{IH}^{2}(X,\mathbb C), let α~\tilde{\alpha} be the class in H2(X~,C)H^{2}(\widetilde{X},\mathbb C) under the injection IH2(X,C)H2(X~,C)\operatorname{IH}^{2}(X,\mathbb C)\hookrightarrow H^{2}(\widetilde{X},\mathbb C). For σIH2,0(X)\sigma\in\operatorname{IH}^{2,0}(X), we define a quadratic form on IH2(X,C)\operatorname{IH}^{2}(X,\mathbb C) by

QX,σ(α):=n2X~(σσ¯)n1α~2+(1n)X~σnσ¯n1α~X~σn1σ¯nα~.Q_{X,\sigma}(\alpha):=\frac{n}{2}\intop\nolimits_{\widetilde{X}}(\sigma\overline{\sigma})^{n-1}\tilde{\alpha}^{2}+(1-n)\intop\nolimits_{\widetilde{X}}\sigma^{n}\overline{\sigma}^{n-1}\tilde{\alpha}\intop\nolimits_{\widetilde{X}}\sigma^{n-1}\overline{\sigma}^{n}\tilde{\alpha}.

The definition extends the quadratic form qX,σq_{X,\sigma} (see Definition 2.5) and is compatible with the decomposition theorem, Theorem 2.8. In particular, we have the following.

Lemma 5.6.

Let XX be a primitive symplectic variety of dimension 2n2n. If  π:X~X\pi\colon\widetilde{X}\to X is a resolution of singularities and ϕ:ZX\phi\colon Z\to X is a Q\mathbb Q-factorial terminalization, then

QX,σ|H2(X,C)=qX,σ,qZ,σZ|IH2(X,C)=QX,σ.Q_{X,\sigma}|_{H^{2}(X,\mathbb C)}=q_{X,\sigma},\quad q_{Z,\sigma_{Z}}|_{\operatorname{IH}^{2}(X,\mathbb C)}=Q_{X,\sigma}.
Proof.

This follows from Schwald’s description of the form qX,σq_{X,\sigma}, see [Sch20], where he showed that qX,σ(α)q_{X,\sigma}(\alpha) agrees with

n2X~(σσ¯)n1α~2+(1n)X~σnσ¯n1α~X~σn1σ¯nα~\frac{n}{2}\intop\nolimits_{\widetilde{X}}(\sigma\overline{\sigma})^{n-1}\tilde{\alpha}^{2}+(1-n)\intop\nolimits_{\widetilde{X}}\sigma^{n}\overline{\sigma}^{n-1}\tilde{\alpha}\intop\nolimits_{\widetilde{X}}\sigma^{n-1}\overline{\sigma}^{n}\tilde{\alpha}

for αH2(X,C)\alpha\in H^{2}(X,\mathbb C). The claim follows since we have inclusions H2(X,C)IH2(X,C)H2(Z,C)H2(X~,C)H^{2}(X,\mathbb C)\subset\operatorname{IH}^{2}(X,\mathbb C)\subset H^{2}(Z,\mathbb C)\subset H^{2}(\widetilde{X},\mathbb C). ∎

Then note that QX,σQ_{X,\sigma} is independent of σ\sigma whenever qX,σq_{X,\sigma} is. This leads to the following definition.

Definition 5.7.

The intersection Beauville–Bogomolov–Fujiki form on IH2(X,C)\operatorname{IH}^{2}(X,\mathbb C) of a primitive symplectic variety XX with X(σσ¯)n=1\intop\nolimits_{X}(\sigma\overline{\sigma})^{n}=1 is the quadratic form QX:=QX,σQ_{X}:=Q_{X,\sigma}.

Proposition 5.8.

Let XX be a primitive symplectic variety with a resolution of singularities π:X~X\pi\colon\widetilde{X}\to X.

  1. (1)

    There is a constant cR>0c\in\mathbb R_{>0} such that for every αH2(X,C)\alpha\in H^{2}(X,\mathbb C),

    cQX(α)n=X~α~n,c\cdot Q_{X}(\alpha)^{n}=\intop\nolimits_{\widetilde{X}}\widetilde{\alpha}^{n},

    where α~\widetilde{\alpha} is the extension of α\alpha under the inclusion H2(X,C)H2(X~,C)H^{2}(X,\mathbb C)\subset H^{2}(\widetilde{X},\mathbb C).

  2. (2)

    The quadratic form QXQ_{X} defines a real quadratic form on IH2(X,R)\operatorname{IH}^{2}(X,\mathbb R) of signature (3,B23)(3,B_{2}-3), where B2=dimIH2(X,R)B_{2}=\dim\operatorname{IH}^{2}(X,\mathbb R).

Proof.

We may assume without loss of generality that XX is projective, so let ϕ:ZX\phi\colon Z\to X be a Q\mathbb Q-factorial terminalization. The first claim follows from Lemma 5.6 and the fact that the statement holds for qZq_{Z} by [Sch20, Theorem 2(1)].

For the second claim, note that by [Sch20, Proposition 25], there are decompositions H2(X,R)=VX+VXH^{2}(X,\mathbb R)=V_{X}^{+}\oplus V_{X}^{-} and H2(Z,R)=VZ+VZH^{2}(Z,\mathbb R)=V_{Z}^{+}\oplus V_{Z}^{-} which are qXq_{X}- and qZq_{Z}-orthogonal, respectively, such that the restrictions qX|VX+q_{X}|_{V_{X}^{+}} and qZ|VZ+q_{Z}|_{V_{Z}^{+}} (resp. qX|VXq_{X}|_{V_{X}^{-}} and qZ|VZq_{Z}|_{V_{Z}^{-}}) are positive-definite (resp. negative-definite) for qXq_{X} and qZq_{Z}. For σH2,0(X)H2,0(Z)\sigma\in H^{2,0}(X)\cong H^{2,0}(Z), the positive parts can be described by VX+=VZ+=Re(σ),im(σ),αV_{X}^{+}=V_{Z}^{+}=\langle\mathrm{Re}(\sigma),\operatorname{im}(\sigma),\alpha\rangle for some ample class αH2(X,R)H2(Z,R)\alpha\in H^{2}(X,\mathbb R)\subset H^{2}(Z,\mathbb R). We can extend VX+V_{X}^{+} to IH2(X,R)\operatorname{IH}^{2}(X,\mathbb R) to a 3-dimension real space (VX+)IH2(X)(V_{X}^{+})^{\prime}\subset\operatorname{IH}^{2}(X) for which QXQ_{X} is positive-definite. It is now immediate that the signature must be (3,B23)(3,B_{2}-3), as having an isotropic vector subspace in (IH2(X,R),QX)(\operatorname{IH}^{2}(X,\mathbb R),Q_{X}) would necessarily lower the rank of (VX+)(V_{X}^{+})^{\prime}. One can define a QXQ_{X}-orthogonal decomposition IH2(X,R)=(VX+)(VX)\operatorname{IH}^{2}(X,\mathbb R)=(V_{X}^{+})^{\prime}\oplus(V_{X}^{-})^{\prime} as in [Sch20]. The negative-definite part is given by

(VX)=αIH1,1(X)IH2(X,R),(V_{X}^{-})^{\prime}=\alpha^{\perp}\cap\operatorname{IH}^{1,1}(X)\cap\operatorname{IH}^{2}(X,\mathbb R),

and the proof, which is entirely linear algebraic on the Hodge structure, is the same as for the case of H2(X,C)H^{2}(X,\mathbb C). ∎

5.2.2.  Proof of Theorem 1.1

We now want to prove the LLV structure theorem for intersection cohomology. The LLV algebra is defined as in the smooth case: Recall that a class αIH2(X,Q)\alpha\in\operatorname{IH}^{2}(X,\mathbb Q) is HL if it satisfies hard Lefschetz.

By Proposition 5.7, the BBF form on IH2(X,R)\operatorname{IH}^{2}(X,\mathbb R) descends to IH2(X,Q)\operatorname{IH}^{2}(X,\mathbb Q), and we get a rational quadratic vector space (IH2(X,Q),QX)(\operatorname{IH}^{2}(X,\mathbb Q),Q_{X}).

Theorem 5.9.

If  XX is a primitive symplectic variety with isolated singularities and b25b_{2}\geq 5, the intersection LLV algebra satisfies

gso((IH2(X,Q),QX)h).\mathfrak{g}\cong\mathfrak{so}((\operatorname{IH}^{2}(X,\mathbb Q),Q_{X})\oplus\mathfrak h).
Proof.

First suppose that XX is projective. Since XX has canonical singularities, it admits a Q\mathbb Q-factorial terminalization ϕ:ZX\phi\colon Z\to X by [BCHM10]. We denote by gZ\mathfrak g_{Z} the LLV algebra on ZZ. Since ϕ\phi satisfies reflexive pullback of differentials, see (2.2), ZZ has at worst isolated singularities. By Theorem 5.4,

gZLβ,|βqZ(β)0so((H2(X,Q),qZ)h).\mathfrak g_{Z}\cong\langle L_{\beta},{}_{\beta}|\leavevmode\nobreak\ q_{Z}(\beta)\neq 0\rangle\cong\mathfrak{so}((H^{2}(X,\mathbb Q),q_{Z})\oplus\mathfrak h).

Suppose αIH2(X,C)\alpha\in\operatorname{IH}^{2}(X,\mathbb C) is any class which satisfies hard Lefschetz. Since ϕ\phi is semismall, then the pullback ϕα\phi^{*}\alpha, which exists since XX has rational singularities, must also satisfy hard Lefschetz. Note that this means qZ(ϕα)0q_{Z}(\phi^{*}\alpha)\neq 0 by Proposition 2.11, and so QX(α)0Q_{X}(\alpha)\neq 0 by Lemma 5.6. It follows that

g=Lα,|αQX(α)0.\mathfrak g=\langle L_{\alpha},{}_{\alpha}|\leavevmode\nobreak\ Q_{X}(\alpha)\neq 0\rangle.

By [dCM05, Remark 4.4.3], it follows that α is a direct summand of ϕα{}_{\phi^{*}\alpha} as α is uniquely determined by the commutator relation [Lα,]α=(k2n)id[L_{\alpha},{}_{\alpha}]=(k-2n)\operatorname{id}. By definition, then ggZ\mathfrak g\subset\mathfrak g_{Z}, which is a canonical injection, and the Lα,αL_{\alpha},{}_{\alpha} satisfy all the same commutator relations as the Lϕα,ϕαL_{\phi^{*}\alpha},{}_{\phi^{*}\alpha}. In particular, g\mathfrak g is just the restriction of gZso((H2(Z,Q),qZ)h)\mathfrak g_{Z}\cong\mathfrak{so}((H^{2}(Z,\mathbb Q),q_{Z})\oplus\mathfrak h) to the subspace generated by the hard Lefschetz operators corresponding to the elements of IH2(X,C)\operatorname{IH}^{2}(X,\mathbb C). Since qZ|IH2(X,Q)=QXq_{Z}|_{\operatorname{IH}^{2}(X,\mathbb Q)}=Q_{X} by Lemma 5.6, it follows that gso((IH2(X,Q),QX)h)\mathfrak g\cong\mathfrak{so}((\operatorname{IH}^{2}(X,\mathbb Q),Q_{X})\oplus\mathfrak h).

In general, the total Lie algebra is a locally trivial diffeomorphism invariant, and so the Lie algebra structure of g\mathfrak g is preserved under small deformations. Since a general locally trivial deformation of XX is projective, see [BL22, Corollary 6.11], the result follows. ∎

The proof indicates that any primitive symplectic variety which admits a Q\mathbb Q-factorial terminalization with isolated singularities also satisfies the LLV structure theorem. This, for example, holds in dimension 4 by Proposition 2.2.

Corollary 5.10.

If  XX is a primitive symplectic variety of dimension 4 and b25b_{2}\geq 5, then there is an isomorphism gso((IH2(X,Q),QX)h)\mathfrak g\cong\mathfrak{so}((\operatorname{IH}^{2}(X,\mathbb Q),Q_{X})\oplus\mathfrak h).

Another interesting consequence of the proof of Theorem 5.9 is in the case b2=4b_{2}=4, where [BL22, Theorem 1.1] does not apply. If such an XX is non-terminal, then the inclusion IH2(X,Q)H2(Z,Q)\operatorname{IH}^{2}(X,\mathbb Q)\hookrightarrow H^{2}(Z,\mathbb Q) of a Q\mathbb Q-factorialization must be strict, and so dimH2(Z,Q)5\dim H^{2}(Z,\mathbb Q)\geq 5. Therefore, gZso((H2(Z,Q),qZ)h)\mathfrak g_{Z}\cong\mathfrak{so}((H^{2}(Z,\mathbb Q),q_{Z})\oplus\mathfrak h), and αH2(Z,Q)\alpha\in H^{2}(Z,\mathbb Q) is HL if and only if qZ0q_{Z}\neq 0. We get the following.

Corollary 5.11.

If  XX is a non-terminal projective primitive symplectic variety and b2=4b_{2}=4, then

g=Lα,|αQX(α)0so((IH2(X,Q),QX)h).\mathfrak g=\langle L_{\alpha},{}_{\alpha}|\leavevmode\nobreak\ Q_{X}(\alpha)\neq 0\rangle\cong\mathfrak{so}((\operatorname{IH}^{2}(X,\mathbb Q),Q_{X})\oplus\mathfrak h).
Remark 5.12.

As indicated in the introduction, we note that our proof gives an algebraic proof of the LLV structure theorem for compact hyperkähler manifolds with b25b_{2}\geq 5. We note that by Verbitsky’s global Torelli theorem (which depends on the existence of twistor deformations), the monodromy groups Mon(X)O()\operatorname{Mon}(X)\subset O(\Gamma) are always finite-index subgroups, and so GXSO()CG_{X}\subset\operatorname{SO}({}_{\mathbb C}) is Zariski dense (see [BL22, Remark 8.12] for more details). The proof of Theorem 5.2 follows through for any b2b_{2} in the smooth case.

5.3.  Holomorphic Symplectic Orbifolds

The methods of this paper show that the LLV structure theorem can be extended to any primitive symplectic variety XX for which the Hodge filtration on (intersection) cohomology is generated by the symplectic form. One other case where this can be seen is when XX has at worst quotient singularities. If XX is a primitive symplectic orbifold, then we have the following:

  1. (1)

    IH(X,Q)H(X,Q)\operatorname{IH}^{*}(X,\mathbb Q)\cong H^{*}(X,\mathbb Q) since XX is a Q\mathbb Q-homology manifold; see [HTT08, Proposition 8.2.21]. In particular, H(X,Q)H^{*}(X,\mathbb Q) satisfies hard Lefschetz, and we may define the total Lie algebra g\mathfrak g.

  2. (2)

    The Hodge filtration on H(X,C)H^{*}(X,\mathbb C) is induced by a spectral sequence

    E1p,q:=Hq(X,)X[p]Hp+q(X,C),E_{1}^{p,q}:=H^{q}\left(X,{}_{X}^{[p]}\right)\Longrightarrow H^{p+q}(X,\mathbb C),

    which is known classically by Steenbrink [Ste77, Section 1.6], but see also [SVV23, Corollary 4.3] and [Tig23a, Theorem 7.2] for an argument using du Bois complexes.

We therefore get the symmetry

Hnp,q(X)Hn+p,q(X)H^{n-p,q}(X)\overset{\vbox to0.0pt{\vss\hbox{$\sim$}\vskip-3.0pt}}{\longrightarrow}H^{n+p,q}(X)

induced by the symplectic form, and the Hodge filtration is generated by the class of the symplectic form. We conclude the following from the methods of Section 5.

Proposition 5.13.

If  XX is a primitive symplectic orbifold, then there is an isomorphism

gso((H2(X,Q),qX)h).\mathfrak g\cong\mathfrak{so}((H^{2}(X,\mathbb Q),q_{X})\oplus\mathfrak h).
Proof.

From the discussion above, the only thing to remark is that we can drop the assumption that b25b_{2}\geq 5 (see also Section 5.4). In this case, we can use Menet’s global Torelli theorem for holomorphic symplectic orbifolds [Men20, Theorem 1.1] and note that the surjectivity of the period map is sufficient to prove Theorem 2.6 (see [BL22, Remark 8.12]), and so Theorem 5.2, Proposition 5.3, and Theorem 5.4 extend to this setting. ∎

5.4.  A Remark on b𝟐<𝟓\boldsymbol{b_{2}<5}

Our methods leave open the case b2=3b_{2}=3, and b2=4b_{2}=4 in the terminal case. It is not known if there exists a compact hyperkähler manifold with b2=3b_{2}=3 or 44, but Verbitsky’s global Torelli theorem predicts that the cohomology is completely described by its structure as an so(b22,4)\mathfrak{so}(b_{2}-2,4)-representation. When b2=3b_{2}=3, this is exactly the action of so(4,1)\mathfrak{so}(4,1) coming from the twistor deformation.

In the singular world, there is a primitive symplectic variety with b2=3b_{2}=3. It is obtained by taking the Fano variety of lines of a special cubic 4-fold admitting an automorphism of order 11; see [Mon13]. The induced automorphism is symplectic, and the resulting quotient has the desired Betti number; see [FM21, Section 5.2]. In this case, the action of so(4,1)\mathfrak{so}(4,1) exists since symplectic orbifolds admit twistor deformations; see [Men20, Theorem 5.4].

It is unknown if there exists a primitive symplectic variety with b2=4b_{2}=4.

Of course, IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) inherits the structure of an sl2\mathfrak{sl}_{2}-representation by hard Lefschetz. The symplectic hard Lefschetz theorem gives intersection cohomology the structure of a representation of a larger Lie algebra with more symmetries.

Proposition 5.14.

If  XX is a primitive symplectic variety with isolated singularities, then IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) inherits the structure of an so(4)\mathfrak{so}(4)-representation. Moreover, the Hodge filtration on IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) induced by σ\sigma is completely determined by so(4)\mathfrak{so}(4).

Proof.

The hard Lefschetz theory for the pair (σ,σ¯)(\sigma,\overline{\sigma}) (see Section 4.1) induces two structures on IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) via the sl2\mathfrak{sl}_{2}-triples gσ=Lσ,,σHσ\mathfrak g_{\sigma}=\langle L_{\sigma},{}_{\sigma},H_{\sigma}\rangle and gσ¯=Lσ¯,,σ¯Hσ¯\mathfrak g_{\overline{\sigma}}=\langle L_{\overline{\sigma}},{}_{\overline{\sigma}},H_{\overline{\sigma}}\rangle. Proposition 4.2 and Corollary 4.6 imply that the operators

Lσ,Lσ¯,,σ,σ¯Hσ,Hσ¯L_{\sigma},L_{\overline{\sigma}},{}_{\sigma},{}_{\overline{\sigma}},H_{\sigma},H_{\overline{\sigma}}

are linearly independent. The Lie algebra gσ,σ¯\mathfrak g_{\sigma,\overline{\sigma}} generated by these six operators is isomorphic to so(4)\mathfrak{so}(4). Moreover, gσ,σ¯\mathfrak{g}_{\sigma,\overline{\sigma}} contains the Weil operator

(5.3) Cσ=i(pq)id=i(HσHσ¯)=i([Lσ,]σ[Lσ¯,]σ¯),C_{\sigma}=i(p-q)\operatorname{id}=i(H_{\sigma}-H_{\overline{\sigma}})=i([L_{\sigma},{}_{\sigma}]-[L_{\overline{\sigma}},{}_{\overline{\sigma}}]),

and so the Hodge filtration on IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) is completely predicted by gσ,σ¯\mathfrak g_{\sigma,\overline{\sigma}}. ∎

Therefore, the Hodge structure on intersection cohomology on any primitive symplectic variety is detected by the symplectic hard Lefschetz theorem, with no restriction on b2b_{2}.

6.  Representation Theory and Hodge Theory of the LLV Algebra

6.1.  Verbitsky Component of 𝐈𝐇(X,Q)\boldsymbol{\operatorname{IH}^{*}(X,\mathbb Q)}

Let XX be a compact hyperkähler manifold. The LLV algebra gives the cohomology ring H(X,Q)H^{*}(X,\mathbb Q) the structure of a g\mathfrak g-representation. This structure has been extensively studied in [GKLR22] for the known cases of compact hyperkähler manifolds and has been used to produce bounds on b2b_{2} in low dimensions, see [Gua01, Saw22, Kur15], and conjecturally in all dimensions; see [KL20].

We wish to extend some well-known results on the representation theory of the LLV algebra action on the intersection cohomology module IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q) of a primitive symplectic variety XX. We believe that the g\mathfrak g-structure on IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q) can restrict both the number and types of singularities that primitive symplectic varieties can admit, which will be explored in future work. The first step is to understand the Verbitsky component, which is the submodule of IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q) generated by IH2(X,Q)\operatorname{IH}^{2}(X,\mathbb Q).

Theorem 6.1.

Let XX be a primitive symplectic variety of dimension 2n2n with isolated singularities and b25b_{2}\geq 5. Then the submodule SIH2(X,Q)IH(X,Q)\operatorname{SIH}^{2}(X,\mathbb Q)\subset\operatorname{IH}^{*}(X,\mathbb Q) generated by IH2(X,Q)\operatorname{IH}^{2}(X,\mathbb Q) is an irreducible g\mathfrak g-module of  SymH2(X,Q)\operatorname{Sym}^{*}H^{2}(X,\mathbb Q).

Proof.

The proof is nearly identical to the smooth case; see [Ver96, Theorem 1.7] and also [GKLR22, Theorem 2.15]. Consider the decomposition

g=Lα,|αQX(α)0=g2(g¯×QH)g2,\mathfrak g=\langle L_{\alpha},{}_{\alpha}|\leavevmode\nobreak\ Q_{X}(\alpha)\neq 0\rangle=\mathfrak g_{2}\oplus(\overline{\mathfrak g}\times\mathbb Q\cdot H)\oplus\mathfrak g_{-2},

which exists by restricting the decomposition (5.1) of a Q\mathbb Q-factorial terminalization, after possibly passing to a locally trivial deformation. The semisimple part g¯g0\overline{\mathfrak g}\subset\mathfrak g_{0} of the degree 0 part of the LLV algebra acts on SIH2(X,Q)\operatorname{SIH}^{2}(X,\mathbb Q) as it acts by derivations on cup product. Clearly, the weight operators HH and LαL_{\alpha} for αH2(X,Q)\alpha\in H^{2}(X,\mathbb Q) act on SIH2(X,C)\operatorname{SIH}^{2}(X,\mathbb C). To see that α acts on SIH2(X,Q)\operatorname{SIH}^{2}(X,\mathbb Q), let α1αkSH2(X,Q)\alpha_{1}\cdots\alpha_{k}\in\operatorname{SH}^{2}(X,\mathbb Q) be any element, and note that

(α1αk)α=[Lα1,]α(α2αk)Lα1(α2α1αk).{}_{\alpha}(\alpha_{1}\cdots\alpha_{k})=[L_{\alpha_{1}},{}_{\alpha}](\alpha_{2}\cdots\alpha_{k})-L_{\alpha_{1}}({}_{\alpha_{1}}\alpha_{2}\cdots\alpha_{k}).

The result follows by induction on kk and the fact that [Lα1,]αg0[L_{\alpha_{1}},{}_{\alpha}]\in\mathfrak g_{0}.

Now we may consider SIH2(X,Q)\operatorname{SIH}^{2}(X,\mathbb Q) as a representation of g¯so(IH2(X,C),QX)\overline{\mathfrak g}\cong\mathfrak{so}(\operatorname{IH}^{2}(X,\mathbb C),Q_{X}). Just as in [Ver96, Section 15], we see that

SIH2(X,Q)2k={SymkIH2(X,Q),kn,Sym2nkIH2(X,Q),k>n,\operatorname{SIH}^{2}(X,\mathbb Q)_{2k}=\begin{cases}\operatorname{Sym}^{k}\operatorname{IH}^{2}(X,\mathbb Q),&k\leq n,\\ \operatorname{Sym}^{2n-k}\operatorname{IH}^{2}(X,\mathbb Q),&k>n,\end{cases}

which depends only on the representation theory of so(IH2(X,C),QX)\mathfrak{so}(\operatorname{IH}^{2}(X,\mathbb C),Q_{X}). Therefore,

SIH2(X,Q)=SymnIH2(X,Q)k1(SymnkIH2(X,Q))2.\operatorname{SIH}^{2}(X,\mathbb Q)=\operatorname{Sym}^{n}\operatorname{IH}^{2}(X,\mathbb Q)\oplus\bigoplusop\displaylimits_{k\geq 1}\left(\operatorname{Sym}^{n-k}\operatorname{IH}^{2}(X,\mathbb Q)\right)^{\oplus 2}.

It then follows that SIH2(X,Q)\operatorname{SIH}^{2}(X,\mathbb Q) extends to g\mathfrak g as an irreducible g\mathfrak g-representation due to the “branching rules” for special orthogonal groups; see [GKLR22, Section B.2]. ∎

We also get the following description of SIH2(X,Q)\operatorname{SIH}^{2}(X,\mathbb Q), due to Bogomolov [Bog96] in the smooth case. A similar description was given for a general primitive symplectic variety in terms of H2(X,Q)H^{2}(X,\mathbb Q) in [BL22, Proposition 5.11].

Proposition 6.2.

Let XX be a primitive symplectic variety of dimension 2n2n with isolated singularities and b25b_{2}\geq 5. Then

SIH2(X,Q)=SymIH2(X,Q)/αn+1|QX(α)=0.\operatorname{SIH}^{2}(X,\mathbb Q)=\operatorname{Sym}^{*}\operatorname{IH}^{2}(X,\mathbb Q)/\langle\alpha^{n+1}|\leavevmode\nobreak\ Q_{X}(\alpha)=0\rangle.

The proof is completely algebraic, and the main input is the following standard lemma.

Lemma 6.3.

Let (H,Q)(H,Q) be a complex vector space with a non-degenerate quadratic form QQ, and let AA be a graded quotient of  SymH\operatorname{Sym}^{*}H by a graded ideal II^{*} such that

  1. (1)

    A2n0A^{2n}\neq 0,

  2. (2)

    Iαn+1|QX(α)=0I^{*}\supset\langle\alpha^{n+1}|\leavevmode\nobreak\ Q_{X}(\alpha)=0\rangle.

Then I=αn+1|QX(α)=0I^{*}=\langle\alpha^{n+1}|\leavevmode\nobreak\ Q_{X}(\alpha)=0\rangle.

Proof of Proposition 6.2. Lemma 6.3 applies to (IH2(X,C),QX)(\operatorname{IH}^{2}(X,\mathbb C),Q_{X}) and A=SIH2(X,C)A=\operatorname{SIH}^{2}(X,\mathbb C) by Theorem 6.1. Note that A2nIH2n,2n(X)A^{2n}\supset\operatorname{IH}^{2n,2n}(X) is non-zero. The problem is invariant under small deformations, so we may assume that XX admits a Q\mathbb Q-factorial terminalization ϕ:ZX\phi\colon Z\to X. There is a commutative diagram

SymH2(Z,Q){\operatorname{Sym}^{*}H^{2}(Z,\mathbb Q)}SymH2(Z,Q)/αn+1|qZ(α)=0{\operatorname{Sym}^{*}H^{2}(Z,\mathbb Q)/\langle\alpha^{n+1}|\leavevmode\nobreak\ q_{Z}(\alpha)=0\rangle}SH2(Z,Q){\operatorname{SH}^{2}(Z,\mathbb Q)}SymIH2(X,Q){\operatorname{Sym}^{*}\operatorname{IH}^{2}(X,\mathbb Q)}SIH2(X,Q).{\operatorname{SIH}^{2}(X,\mathbb Q)\hbox to0.0pt{.\hss}}

The natural map SymH2(Z,Q)SH2(Z,Q)\operatorname{Sym}^{*}H^{2}(Z,\mathbb Q)\to\operatorname{SH}^{2}(Z,\mathbb Q) factors through SymH2(Z,Q)/αn+1|qZ(α)=0\operatorname{Sym}^{*}H^{2}(Z,\mathbb Q)/\langle\alpha^{n+1}|\leavevmode\nobreak\ q_{Z}(\alpha)=0\rangle by [BL22, Proposition 5.11]. Therefore, the bottom map must factor as

SymIH2(X,Q)SymIH2(X,Q)/αn+1|QX(α)=0SIH2(X,Q)\operatorname{Sym}^{*}\operatorname{IH}^{2}(X,\mathbb Q)\longrightarrow\operatorname{Sym}^{*}\operatorname{IH}^{2}(X,\mathbb Q)/\langle\alpha^{n+1}|\leavevmode\nobreak\ Q_{X}(\alpha)=0\rangle\longrightarrow\operatorname{SIH}^{2}(X,\mathbb Q)

by Lemma 5.6. If this map were not injective, then Lemma 6.3 would imply that the kernel would contain the rational cohomology class corresponding to the generator of IH2n,2n(X)=(σ+σ¯)2n\operatorname{IH}^{2n,2n}(X)=(\sigma+\overline{\sigma})^{2n}, which does not vanish. This finishes the proof. ∎

Corollary 6.4.

Let XX be a primitive symplectic variety of dimension 2n2n with isolated singularities and b25b_{2}\geq 5. For every knk\leq n, there is an injection

SymkIH2(X,Q)⸦⟶IH2k(X,Q).\operatorname{Sym}^{k}\operatorname{IH}^{2}(X,\mathbb Q)\lhook\joinrel\longrightarrow\operatorname{IH}^{2k}(X,\mathbb Q).

6.2.  Kuga–Satake Construction on the Cohomology of Primitive Symplectic Varieties

The Kuga–Satake construction, see [KS67], associates to a polarized Hodge structure H=HZH=H_{\mathbb Z} of K3 type a complex torus TT for which HH is a sub-Hodge structure of Hom(H1(T),H1(T))(1)\operatorname{Hom}(H_{1}(T),H_{1}(T))(1). The construction associates to the weight 2 Hodge structure HH (with its bilinear form) its Clifford algebra C(H)C(H). There is an induced complex structure on C(H)RC(H)\otimes\mathbb R, and one can show that the quotient C(H)R/C(H)C(H)\otimes\mathbb R/C(H) is a complex torus, which is in fact an abelian variety in the case that HH is the second cohomology of a projective K3 surface.

Understanding the geometric connection between varieties admitting Hodge structures of K3 type and the Kuga–Satake construction is a difficult problem, as the Mumford–Tate group of

Hom(H1(T),H1(T))(1)H1(T)H1(T)\operatorname{Hom}(H_{1}(T),H_{1}(T))(1)\cong H^{1}(T)\otimes H^{1}(T)

is highly restricted, while the Mumford–Tate group associated to a Hodge structure of K3 type can be quite large (see Section 6.3). In special cases, it has been observed that the Kuga–Satake construction for K3 surfaces is related to the Hodge conjecture; see [vGe00]. It is therefore interesting to understand the geometry of the Kuga–Satake construction, as well as generalizations to Hodge structures of higher weights.

In [KSV19], Kurnosov–Soldatenkov–Verbitsky observe that there is a multidimensional Kuga–Satake construction on the cohomology ring of a compact hyperkähler manifold XX. Namely, associated to XX are a complex torus TT, a non-negative integer ll, and embeddings

g⸦⟶gtot(T),:H(X,C)⸦⟶H+l(T,C),\mathfrak g\lhook\joinrel\longrightarrow\mathfrak g_{\mathrm{tot}}(T),\quad\Psi\colon H^{*}(X,\mathbb C)\lhook\joinrel\longrightarrow H^{*+l}(T,\mathbb C),

where gtot(T)\mathfrak g_{\mathrm{tot}}(T) is the total Lie algebra of TT. Here, the embedding is a morphism with respect to the induced structures as g\mathfrak g-representations (resp. gtot(T)\mathfrak g_{\mathrm{tot}}(T)-representations). Fixing a complex structure on XX makes a morphism of Hodge structures.

Using the existence of the LLV algebra, we outline how the Kuga–Satake construction holds for intersection cohomology, further demonstrating how unique the geometry of primitive symplectic varieties is.

6.2.1.  LLV embedding

Consider a finite-dimensional complex vector space HH and a non-degenerate symmetric bilinear form QQ. Let THT^{*}H denote the tensor algebra, and let aTH\mathfrak a\subset T^{*}H be the ideal generated by elements of the form

vvQ(v,v),vH.v\otimes v-Q(v,v),\quad v\in H.

The Clifford algebra of (H,Q)(H,Q) is C=C(H,Q):=TH/aC=C(H,Q):=T^{*}H/\mathfrak a.

The main technical result of [KSV19] is the following.

Theorem 6.5 (cf. [KSV19, Theorem 3.14]).

If  (H,Q)(H,Q) is any quadratic vector space and WW is any representation of  g:=so(H~,Q~)\mathfrak g:=\mathfrak{so}(\tilde{H},\tilde{Q}), where (H~,Q~)(\tilde{H},\tilde{Q}) is the Mukai completion of  (H,Q)(H,Q), then there is a C(H,Q)C(H,Q)-module VV with an invariant symmetric bilinear form τ\tau such that V\bigwedgeop\displaylimits^{\bullet}V^{*} contains WW as a g\mathfrak g-submodule.

The construction is seen by applying Theorem 6.5 to (IH2(X,C),QX)(\operatorname{IH}^{2}(X,\mathbb C),Q_{X}), where XX is a primitive symplectic variety with isolated singularities and b25b_{2}\geq 5. We let W=IH(X,C)W=\operatorname{IH}^{*}(X,\mathbb C), which is a g\mathfrak g-representation by Theorem 5.9. Then there exists a C=C(IH2(X,C),QX)C=C(\operatorname{IH}^{2}(X,\mathbb C),Q_{X})-module VV with an embedding

IH2(X,C)⸦⟶2V\operatorname{IH}^{2}(X,\mathbb C)\lhook\joinrel\longrightarrow\displaystyle\mathop{\bigwedgeop\displaylimits\nolimits^{\!2}}V^{*}

and an embedding of g\mathfrak g-modules

IH(X,C)⸦⟶V.\operatorname{IH}^{*}(X,\mathbb C)\lhook\joinrel\longrightarrow\displaystyle\mathop{\bigwedgeop\displaylimits\nolimits^{\!\bullet}}V^{*}.

Now g\mathfrak g induces a grading on IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) and V\displaystyle\mathop{\bigwedgeop\displaylimits\nolimits^{\!\bullet}}V^{*}, whence we have a degree ll morphism

(6.1) ψ:IH(X,C)⸦⟶+lV\psi\colon\operatorname{IH}^{*}(X,\mathbb C)\lhook\joinrel\longrightarrow\displaystyle\mathop{\bigwedgeop\displaylimits\nolimits^{\!\bullet+l}}V^{*}

of graded vector spaces for some ll. If we take T=V/T=V/\Gamma for some lattice V\Gamma\subset V, we get an embedding :H(X,C)H+l(T,C)\Psi\colon H^{*}(X,\mathbb C)\hookrightarrow H^{*+l}(T,\mathbb C) by (6.1).

As we have seen, the Hodge structure on IH(X,R)\operatorname{IH}^{*}(X,\mathbb R) is detected by the Weil operators Cσg¯gC_{\sigma}\in\overline{\mathfrak g}\subset\mathfrak g, which are determined once we fix a point [σ][\sigma]\in\Omega in the period domain. Let μg¯so(IH2(X,R),QX)\mu\in\overline{\mathfrak g}\cong\mathfrak{so}(\operatorname{IH}^{2}(X,\mathbb R),Q_{X}) be the corresponding skew-symmetric matrix of rank 2. Let Wσ=γ,γW_{\sigma}^{\prime}=\langle\gamma,\gamma^{\prime}\rangle be the positive two-space corresponding to σ\sigma, where γ=R(σ),γ=I(σ)\gamma=\mathfrak R(\sigma),\gamma^{\prime}=\mathfrak I(\sigma) as in Section 4.2. Then μ=γγg¯C\mu=\gamma\gamma^{\prime}\in\overline{\mathfrak g}\subset C. It acts trivially on the orthogonal complement to WσW_{\sigma}, and μ2=1\mu^{2}=-1 in the Clifford algebra. Thus μ\mu defines a complex structure on H1(T,R)H^{1}(T,\mathbb R), noting that TT is smooth.

Finally, since TT is a complex torus, the Hodge structures on IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q) and H(T,Q)H^{*}(T,\mathbb Q) are both determined by the symplectic hard Lefschetz theorem (see Section 3), which is compatible with the choice CσC_{\sigma}. Therefore, the morphism is compatible with the Hodge structures.

6.2.2.  Polarized Kuga–Satake construction

Just as in the case of compact hyperkähler manifolds, the existence of a polarization in (intersection) cohomology induces a polarization on the complex torus TT described above.

Let XX be a projective primitive symplectic variety and b25b_{2}\geq 5. Suppose that hIH2(X,Q)h\in\operatorname{IH}^{2}(X,\mathbb Q) is an ample class on XX, and let hIH2(X,R)h^{\perp}\subset\operatorname{IH}^{2}(X,\mathbb R) be the orthogonal complement of hh with respect to the intersection BBF form QXQ_{X}. The torus TT is a quotient of some VV, where (as in [KSV19]), V=V1NV=V_{1}^{\bigoplusop\displaylimits N} with V1C=C(IH2(X,R),QX)V_{1}\cong C=C(\operatorname{IH}^{2}(X,\mathbb R),Q_{X}).

Constructing the polarization on TT is done as follows; see [KSV19, Section 4.2]. It is enough to construct a polarization on Ch:=C(h,q|h)C_{h}:=C(h^{\perp},q|_{h^{\perp}}). Note that (h,q)(h^{\perp},q_{\perp}) is of signature (2,k)(2,k) for some kk. Let Wh=γ,γhW_{h}=\langle\gamma,\gamma^{\prime}\rangle\subset h^{\perp} be the subspace where q|hq|_{h^{\perp}} is positive. By local Torelli, we may assume that q|h(γ,γ)=0q|_{h^{\perp}}(\gamma,\gamma^{\prime})=0. Consider the product a=γ1γ2C(h,q|h)a=\gamma_{1}\gamma_{2}\in C(h^{\perp},q|_{h^{\perp}}). For any x,yC(h,q|h)x,y\in C(h^{\perp},q|_{h^{\perp}}), we define

σa(x,y):=Tr(xay¯),\sigma_{a}(x,y):=\operatorname{Tr}(xa\overline{y}),

where Tr\operatorname{Tr} is the trace map corresponding to the algebra CC and y¯\overline{y} is the operator

y¯=αβ(y),\overline{y}=\alpha\beta(y),

where α\alpha is the natural parity involution on the Clifford algebra and β\beta is the anti-automorphism which sends a tensor y1yky_{1}\otimes\cdots\otimes y_{k} to yky1y_{k}\otimes\cdots\otimes y_{1}.

By [KSV19, Proposition 4.2], either σa\sigma_{a} or σa-\sigma_{a} is a polarization on C(h,q|h)C(h^{\perp},q|_{h^{\perp}}). The proof holds here, as the statement holds for any quadratic vector space (H,q)(H,q).

6.3.  The Mumford–Tate Algebra

There is a connection between the Mumford–Tate group of the intersection cohomology of a primitive symplectic variety and the LLV algebra. In the compact hyperkähler case, this was studied in [GKLR22, Section 2]. Given the LLV structure theorem, similar results follow through with only minor adjustments. For convenience, we reference loc. cit. to indicate the corresponding statement in the hyperkähler setting.

Definition 6.6.

Let VV be a Q\mathbb Q-Hodge structure. The special Mumford–Tate algebra mt¯(V)\overline{\mathfrak{mt}}(V) of VV is the smallest Q\mathbb Q-algebraic Lie subalgebra of gl(V)\mathfrak{gl}(V) such that mt¯(V)R\overline{\mathfrak{mt}}(V)_{\mathbb R} contains the Weil operator C=i(pq)idC=i(p-q)\operatorname{id}. The Mumford–Tate algebra is defined as

mt0(V)=mt¯(V)QH,\mathfrak{mt}_{0}(V)=\overline{\mathfrak{mt}}(V)\oplus\mathbb Q\cdot H,

where HH is the weight operator.

Proposition 6.7.

Let XX be a primitive symplectic variety with at worst Q\mathbb Q-factorial isolated singularities and b25b_{2}\geq 5. Let m¯=mt¯(IH(X,Q))\overline{\mathfrak m}=\overline{\mathfrak{mt}}(\operatorname{IH}^{*}(X,\mathbb Q)) be the special Mumford–Tate algebra of the pure Hodge structure IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q). Then m¯g¯\overline{\mathfrak m}\subset\overline{\mathfrak g}, with equality if  XX is very general.

Proof.

The proof is similar to that of [GKLR22, Proposition 2.38], although the main input is that the Weil operator CσC_{\sigma} with respect to a (fixed) Hodge structure on IH(X,Q)\operatorname{IH}^{*}(X,\mathbb Q) is contained in the semisimple part g¯\overline{\mathfrak g} of g0\mathfrak g_{0}. The proof of loc. cit. uses the hyperkähler structure, so we indicate how this works algebraically.

Recall that if we fix the point [σ][\sigma]\in\Omega in the period domain, then there is a pair (γ,γ)(\gamma,\gamma^{\prime}) of non-isotropic classes in H2(X,R)H^{2}(X,\mathbb R) such that [Lγ,]γ=Cσ[L_{\gamma},{}_{\gamma^{\prime}}]=C_{\sigma}; see Corollary 4.7. We also saw that by the surjectivity of the period map with respect to H2H^{2}, this pair completes to a positive three-space

Wσ=α,γ,γso(4,1).W_{\sigma}=\langle\alpha,\gamma,\gamma^{\prime}\rangle\cong\mathfrak{so}(4,1).

It follows that the operators satisfy [Lα,]γ=[Lα,]γ=Cσ[L_{\alpha},{}_{\gamma}]=[L_{\alpha},{}_{\gamma^{\prime}}]=C_{\sigma} as well. Using the commutativity of the dual Lefschetz operators, one can show that

Cσ=[Lγ,]γ=12[[Lα,]γ,[Lα,]γ]g¯;C_{\sigma}=[L_{\gamma},{}_{\gamma^{\prime}}]=-\frac{1}{2}\left[[L_{\alpha},{}_{\gamma}],[L_{\alpha},{}_{\gamma^{\prime}}]\right]\in\overline{\mathfrak g};

see [GKLR22, Proposition 2.24] for the computation. This shows that mt¯(IH(X,Q))g¯Q\overline{\mathfrak{mt}}(\operatorname{IH}^{*}(X,\mathbb Q))\subset\overline{\mathfrak g}_{\mathbb Q} since mt¯(IH(X,Q))R\overline{\mathfrak{mt}}(\operatorname{IH}^{*}(X,\mathbb Q))_{\mathbb R} is the smallest subalgebra to contain the Weil operator CσC_{\sigma}.

For the statement regarding a general primitive symplectic variety, the proof follows as in [GKLR22, Proposition 2.38], and we sketch the main details. The key observation is to notice that the special Mumford–Tate group of any IHk(X,Q)\operatorname{IH}^{k}(X,\mathbb Q) is

(6.2) mt¯(IHk(X,Q))=m¯.\overline{\mathfrak{mt}}(\operatorname{IH}^{k}(X,\mathbb Q))=\overline{\mathfrak m}.

This follows as the g¯\overline{\mathfrak g}-module structure on IHk(X,Q)\operatorname{IH}^{k}(X,\mathbb Q) is determined by the composition

ρk:g¯gl(IH(X,Q))gl(IHk(X,Q)).\rho_{k}\colon\overline{\mathfrak g}\subset\mathfrak{gl}(\operatorname{IH}^{*}(X,\mathbb Q))\longrightarrow\mathfrak{gl}(\operatorname{IH}^{k}(X,\mathbb Q)).

In the smooth case, this map is shown to be injective; see [GKLR22, Corollary 2.36]. This follows in the singular case, however, since the proof only depends on the representation theory of the Verbitsky component V(n):=SH2(X,𝐂)V_{(n)}:=\operatorname{SH}^{2}(X,\mathbf{C}) (this is [GKLR22, Proposition 2.35]), which is identical to the smooth case by Theorem 6.1 and Proposition 6.2. It follows that g¯\overline{\mathfrak g} and ρk(g¯)\rho_{k}(\overline{\mathfrak g}) are isomorphic. Since mt¯(IHk(X,Q))\overline{\mathfrak{mt}}(\operatorname{IH}^{k}(X,\mathbb Q)) is the smallest Q\mathbb Q-algebraic subgroup such that ρk(Cσ)mt¯(IHk(X,Q))\rho_{k}(C_{\sigma})\in\overline{\mathfrak{mt}}(\operatorname{IH}^{k}(X,\mathbb Q)), we see that (6.2) holds.

By the Noether–Lefschetz theory of period domains of Hodge structures of hyperkähler type (see [GGK12]), it follows that a very general primitive symplectic variety with b25b_{2}\geq 5 must satisfy

mt¯(H2(X,Q))so(H2(X,Q),qX),\overline{\mathfrak{mt}}(H^{2}(X,\mathbb Q))\cong\mathfrak{so}(H^{2}(X,\mathbb Q),q_{X}),

noting that H2(X,Q)H^{2}(X,\mathbb Q) satisfies the local Torelli theorem; see [BL22, Proposition 5.5]. ∎

7.  Weak P=W\boldsymbol{P=W} for Primitive Symplectic Varieties

One of the more interesting applications of the LLV algebra for compact hyperkähler manifolds involves the P=WP=W conjecture. Given a degeneration 𝒳\mathscr{X}\to\Delta of a compact hyperkähler manifold XX, the cohomology groups Hk(X,C)H^{k}(X,\mathbb C) inherit a weight filtration from the limit mixed Hodge structure on the unique singular fiber. There is an induced (logarithmic) monodromy operator Nso(H2(X,Q),q)g¯N\in\mathfrak{so}(H^{2}(X,\mathbb Q),q)\cong\overline{\mathfrak g}, which is nilpotent of index either 2 or 3. We say a degeneration is of type III if NN has index 3.

Any compact hyperkähler manifold admits a type III degeneration; see [Sol20]. The P=WP=W conjecture for Lagrangian fibrations states that the induced weight filtration from the limit mixed Hodge structure agrees with the perverse filtration when XX admits a Lagrangian fibration. This was answered positively in [HLSY21] by showing that the data of these filtrations agree with the Hodge filtration induced from a positive three-space WgW_{g} corresponding to a hyperkähler metric gg.

We can form an analog of the P=WP=W conjecture for primitive symplectic varieties admitting a Lagrangian fibration, relating the data of a Lagrangian fibration to the filtration induced by the logarithmic monodromy operator NN of a type III degeneration. We expect that the filtration induced by NN agrees with a limit mixed Hodge structure for intersection cohomology, although there is a subtle issue describing the intersection cohomology in terms of a variation of pure Hodge structures rather than the underlying pure Hodge module. This will be explored in future work.

7.1.  Perverse =\boldsymbol{=} Hodge

The LLV algebra for a compact hyperkähler manifold detects the information of a Lagrangian fibration. We outline how some of these results hold in the case of primitive symplectic varieties, which is based purely on the work of Shen–Yin [SY22]; see also [HM22] for a survey and [FSY22] in the case of primitive symplectic varieties admitting a symplectic resolution.

7.1.1.  Perverse filtration on the cohomology of a Lagrangian fibration

Throughout this section, we assume that XX is a primitive symplectic variety of dimension 2n2n, admitting a Lagrangian fibration f:XBf\colon X\to B to a projective base BB of dimension nn.

Given a projective morphism f:XBf\colon X\to B, there is a natural filtration on the cohomology of XX induced from the images of the truncated complexes of the perverse tt-structure associated to the morphism ff:

PmHk(X,C)=im(Hk2n(B,τmp(𝐑f(𝒞XC)[2n]))Hk(X,C)).P_{m}H^{k}(X,\mathbb C)=\operatorname{im}\left(\mathbb H^{k-2n}(B,{}^{\mathfrak p}\tau_{\leq m}(\mathbf{R}f_{*}(\mathcal{IC}_{X}\otimes\mathbb C)[-2n]))\longrightarrow H^{k}(X,\mathbb C)\right).

The filtration is completely determined by an ample class on the base BB. Indeed, if αH2(B,R)\alpha\in H^{2}(B,\mathbb R) is ample and β=fα\beta=f^{*}\alpha, then

PmHk(X,C)=i(ker(Lβ2n+m+ik)im(Lβi1))IHk(X,C),P_{m}H^{k}(X,\mathbb C)=\sumop\displaylimits_{i}\left(\ker\left(L_{\beta}^{2n+m+i-k}\right)\cap\operatorname{im}\left(L_{\beta}^{i-1}\right)\right)\cap\operatorname{IH}^{k}(X,\mathbb C),

where LβL_{\beta} is the cupping operator, see [dCM05, Proposition 5.2.4].

By the Fujiki relations on the BBF form qXq_{X}, the pullback β\beta is qXq_{X}-isotropic and therefore QXQ_{X}-isotropic. For any isotropic class μIH2(X,C)\mu\in\operatorname{IH}^{2}(X,\mathbb C), we may define an analogous filtration

PmμHk(X,C)=i(ker(Lμ2n+m+ik)im(Lμi1))IHk(X,C).P_{m}^{\mu}H^{k}(X,\mathbb C)=\sumop\displaylimits_{i}\left(\ker\left(L_{\mu}^{2n+m+i-k}\right)\cap\operatorname{im}\left(L_{\mu}^{i-1}\right)\right)\cap\operatorname{IH}^{k}(X,\mathbb C).
Lemma 7.1.

For any QXQ_{X}-isotropic classes μ1,μ2\mu_{1},\mu_{2}, we have

dimPmμ1IHk(X,C)=dimPmμ2IHk(X,C).\dim P_{m}^{\mu_{1}}\operatorname{IH}^{k}(X,\mathbb C)=\dim P_{m}^{\mu_{2}}\operatorname{IH}^{k}(X,\mathbb C).
Proof.

For classes μ1,μ2H2(X,C)\mu_{1},\mu_{2}\in H^{2}(X,\mathbb C), this is [FSY22, Proposition 1.2], and so the lemma holds in the Q\mathbb Q-factorial case. By taking a Q\mathbb Q-factorial terminalization ϕ:ZX\phi\colon Z\to X (noting that XX is projective), we see that dimPmϕμ1Hk(Z,C)=dimPmϕμ2Hk(Z,C)\dim P_{m}^{\phi^{*}\mu_{1}}H^{k}(Z,\mathbb C)=\dim P_{m}^{\phi^{*}\mu_{2}}H^{k}(Z,\mathbb C). But this implies the statement of the theorem as ϕ:ZX\phi\colon Z\to X is semismall. Indeed, we must have that jp(ϕ𝒞Z[2n])=0{}^{\mathfrak p}\mathscr{H}^{j}(\phi_{*}\mathcal{IC}_{Z}[2n])=0 for every j0j\neq 0, so the induced perverse filtration with respect to the morphism ϕ:ZX\phi\colon Z\to X is trivial, and the dimension of the perverse filtration will be determined by the pullback. ∎

Now consider the filtration Pmσ¯IHk(X,C)P_{m}^{\overline{\sigma}}\operatorname{IH}^{k}(X,\mathbb C) associated to the antiholomorphic symplectic form σ¯\overline{\sigma}. It is an increasing filtration which detects the Hodge filtration by Theorem 3.5. Specifically,

Pmσ¯IHk(X,C)=pmIHp,mp(X).P_{m}^{\overline{\sigma}}\operatorname{IH}^{k}(X,\mathbb C)=\bigoplusop\displaylimits_{p\leq m}\operatorname{IH}^{p,m-p}(X).

We therefore see that the Hodge numbers equal the perverse Hodge numbers:

dimIHp,q(X)=hp,qp:=dimgrpPIHm(X,C).\dim\operatorname{IH}^{p,q}(X)={}^{\mathfrak p}h^{p,q}:=\dim\operatorname{gr}_{p}^{P}\operatorname{IH}^{m}(X,\mathbb C).

7.1.2.  A Lefschetz class corresponding to (β,η)\boldsymbol{(\beta,\eta)}

As in the smooth case, the data of a Q\mathbb Q-factorial terminal primitive symplectic variety admitting Lagrangian fibration is encoded in the LLV algebra; see [SY22, HLSY21]. Let β\beta be as above (which is qXq_{X}-isotropic by the Fujiki relations), and let ηH2(X,Q)\eta\in H^{2}(X,\mathbb Q) be an ff-relative ample class. By replacing η\eta with a Q\mathbb Q-linear combination of η\eta and β\beta as needed, we may assume qX(η)=0q_{X}(\eta)=0. By global Torelli, there is a class ρH2(X,Q)\rho\in H^{2}(X,\mathbb Q) such that

qX(ρ)>0,qX(η,ρ)=qX(β,ρ)=0.q_{X}(\rho)>0,\quad q_{X}(\eta,\rho)=q_{X}(\beta,\rho)=0.

The corresponding Lie algebra gρ\mathfrak g_{\rho} generated by the simultaneous sl2\mathfrak{sl}_{2}-triples induced by ρ,β,η\rho,\beta,\eta naturally sits inside g¯so(H2(X,C),qX)\overline{\mathfrak g}\cong\mathfrak{so}(H^{2}(X,\mathbb C),q_{X}); see [HLSY21, Equation (5)]. The Lie algebra gρ\mathfrak g_{\rho} inherits IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) with the structure of an so(5)\mathfrak{so}(5)-representation.

By [SY22, Proposition 1.1], there is a canonical splitting of the perverse filtration:

(7.1) PlβIHk(X,Q)=p+q=lPp,q.P_{l}^{\beta}\operatorname{IH}^{k}(X,\mathbb Q)=\bigoplusop\displaylimits_{p+q=l}P^{p,q}.

We note that the proof is stated for compact hyperkähler manifolds, but it is completely algebraic. The contents of Sections 4 and 5 immediately imply (7.1).

7.2.  Degenerations

Soldatenkov’s proof of the existence of maximally unipotent degenerations is based purely on lattice theory and knowledge of the period domain of compact hyperkähler manifolds. As primitive symplectic varieties satisfy global Torelli, the existence of degenerations for locally trivial families will follow exactly as in the smooth case.

Definition 7.2.

Let XX be a primitive symplectic variety. A degeneration of XX is a flat proper morphism g:𝒳g\colon\mathscr{X}\to\Delta of complex analytic spaces such that

  1. (1)

    for some tt\in{}^{*}, the fiber satisfies 𝒳tX\mathscr{X}_{t}\cong X;

  2. (2)

    the restriction g:𝒳g^{\prime}\colon\mathscr{X}^{*}\to{}^{*} is a locally trivial deformation; and

  3. (3)

    the monodromy action on the H2(𝒳t,Q)H^{2}(\mathscr{X}_{t},\mathbb Q) is unipotent and non-trivial.

We say that a degeneration is projective if gg is a projective morphism.

If gg is a degeneration of XX, then every fiber 𝒳t\mathscr{X}_{t} is also primitive symplectic, and each fiber is Q\mathbb Q-factorial terminal if XX is. Moreover, R2gZR^{2}g^{\prime}_{*}\mathbb Z, where g:𝒳g^{\prime}\colon\mathscr{X}^{*}\to{}^{*} is the restriction, is a local system as we restrict ourselves to locally trivial deformations. In particular, we get a variation of pure Hodge structures 𝒱\mathscr{V}, where each fiber is isomorphic to H2Z(𝒳t,Z)H2(X,Z){}_{\mathbb Z}\cong H^{2}(\mathscr{X}_{t},\mathbb Z)\cong H^{2}(X,\mathbb Z). If hZh\in{}_{\mathbb Z} is the class of a polarization, let 𝒱h\mathscr{V}^{h} be the qXq_{X}-orthogonal complement of the bilinear pairing induced by the Beauville–Bogomolov–Fujiki form on 𝒱\mathscr{V}. It forms a local system with fiber equal to the qq-complement of hh, that is, hZ{}_{\mathbb Z}^{h}.

7.2.1.  Limit mixed Hodge structure for intersection cohomology

The weight filtration induced by a degeneration of primitive symplectic varieties is a consequence of Schmid’s work on the limit mixed Hodge structure; see [Sch73]. Let 𝒱\mathscr{V} be an (integral) variation of pure Hodge structures over admitting a maximally unipotent monodromy operator TT, and let V0=𝒱¯0V_{0}=\overline{\mathscr{V}}_{0} be the fiber of the unique extension of 𝒱\mathscr{V} over . For each kZ0k\in\mathbb Z_{\geq 0}, the corresponding log-monodromy operator NN defines a unique increasing weight filtration WkW^{\bullet}_{k} satisfying the following properties:

  1. (1)

    NWkjWkj2NW^{j}_{k}\subset W^{j-2}_{k} for j2j\geq 2, and

  2. (2)

    the induced map Nl:grj+lWkV0grjlWkV0N^{l}\colon\operatorname{gr}_{j+l}^{W_{k}}V_{0}\to\operatorname{gr}_{j-l}^{W_{k}}V_{0} is an isomorphism for all l0l\geq 0;

see [Sch73, Lemma 6.4]. In particular, the triple (V0,Wk,¯0)(V_{0},W_{k}^{\bullet},\overline{\mathcal{F}}^{\bullet}_{0}) gives rise to a mixed Hodge structure, where ¯\overline{\mathcal{F}}^{\bullet} is a holomorphic extension of the Hodge bundle \mathcal{F}^{\bullet} underlying 𝒱\mathscr{V}.

The main application of the nilpotent weight filtration is on the cohomology of XX induced by a degeneration of compact Kähler manifolds. Extending Proposition 2.12, we can describe a limit mixed Hodge structure on the intersection cohomology of the central fiber of a degeneration g:𝒳g\colon\mathscr{X}\to\Delta of a primitive symplectic variety XX. By definition, the locally trivial family g:𝒳g^{\prime}\colon\mathscr{X}^{*}\to{}^{*} admits a simultaneous resolution of singularities f:𝒴f^{\prime}\colon\mathscr{Y}^{*}\to{}^{*}; see [BL22, Lemma 4.9]. For each kk, let 𝒴k\mathscr{H}_{\mathscr{Y}}^{k} be the variation of pure Hodge structures with local system RkfQ𝒴R^{k}f^{\prime}_{*}\mathbb Q_{\mathscr{Y}}. If IHk\mathbb{IH}^{k} is the local system determined by the intersection cohomology of the fibers (g)1(t)(g^{\prime})^{-1}(t), the decomposition theorem (see Proposition 2.8) implies that IHk\mathbb{IH}^{k} underlies a sub-variation of pure Hodge structures 𝒳k\mathscr{H}_{\mathscr{X}}^{k} of 𝒴k\mathscr{H}_{\mathscr{Y}}^{k}. In particular, we have the following.

Definition-Theorem 7.3.

If g:𝒳g\colon\mathscr{X}\to\Delta is a degeneration of a primitive symplectic variety XX, there is a mixed Hodge structure, called the limit mixed Hodge structure, on the intersection cohomology IHk(X,Q)\operatorname{IH}^{k}(X_{\infty},\mathbb Q) of the canonical fiber XX_{\infty}.

7.2.2.  Existence of type III degenerations

There is an induced monodromy transformation λAut(𝒱h,q)O(,Zhq)\lambda\in\operatorname{Aut}(\mathscr{V}^{h},q)\cong O({}_{\mathbb Z}^{h},q) which, by definition, must be of the form λ=eN\lambda=e^{N}, where Nso(,Qhq)N\in\mathfrak{so}({}^{h}_{\mathbb Q},q), and, by [Sch73, Theorem 6.1], must be of index 2 or 3. We say that a degeneration is maximally unipotent, or has maximally unipotent monodromy, if NN is of index 3.

Proposition 7.4.

Let XX be a primitive symplectic variety with at worst Q\mathbb Q-factorial isolated singularities and b25b_{2}\geq 5. There exists a projective degeneration of XX with maximally unipotent monodromy.

Proof.

The proof follows as in [Sol20, Section 4] almost verbatim, and so we only briefly indicate the main details. Since b25b_{2}\geq 5 and (,Zq)({}_{\mathbb Z},q) is of signature (3,b23)(3,b_{2}-3) as in the smooth case, there exist a polarization hZh\in{}_{\mathbb Z} and an endomorphism Nso(,Qhq)N\in\mathfrak{so}({}_{\mathbb Q}^{h},q) of index 3; see [Sol20, Lemma 4.1]. Moreover, the restriction of qq to the image of NN is semi-positive with one-dimensional kernel.

Let be the period domain with respect to (,Zq)({}_{\mathbb Z},q), and let ^\widehat{\Omega} be the compact dual. The polarization defines a period domain h with compact dual ^h\widehat{\Omega}^{h}. For Nso(,Qhq)N\in\mathfrak{so}({}_{\mathbb Q}^{h},q) and x^hx\in\widehat{\Omega}^{h}, Soldatenkov defines the pair (N,x)(N,x) to be a nilpotent orbit if eitNhe^{itN}\in{}^{h} for every t>>0t>>0. Equivalently, see [Sol20, Lemma 4.4], (N,x)(N,x) is a nilpotent orbit if and only if q(Nx,Nx¯)>0q(Nx,N\overline{x})>0, and the image of such points in ^h\widehat{\Omega}^{h} therefore defines a non-empty open subset. For such a nilpotent monodromy operator NN of index 3, let 𝒩={xh(N,x)isanilpotentorbit}\mathscr{N}=\{x\in{}^{h}\mid(N,x)\leavevmode\nobreak\ \mathrm{is\leavevmode\nobreak\ a\leavevmode\nobreak\ nilpotent\leavevmode\nobreak\ orbit}\}. It is open and non-empty, whence we get an open subset of the period domain which corresponds to this nilpotent operator NN. As in [Sol20, Theorem 4.6], this open subset predicts a degeneration 𝒳\mathscr{X}\to\Delta of XX with logarithmic monodromy operator NN. ∎

7.3.  Singular P=W\boldsymbol{P=W} Theorem

We can now state a singular version of the Lagrangian P=WP=W conjecture, as follows.

Theorem 7.5.

Let XX be a Q\mathbb Q-factorial terminal primitive symplectic variety with isolated singularities and b25b_{2}\geq 5. If f:XBf\colon X\to B is a Lagrangian fibration, the perverse filtration PβP^{\beta} on IH(X,C)\operatorname{IH}^{*}(X,\mathbb C) with respect to the pullback of an ample class on BB agrees with the weight filtration WNW_{N} on IH(X,C)\operatorname{IH}^{*}(X_{\infty},\mathbb C) with respect to the logarithmic monodromy operator NN of a type III degeneration 𝒳\mathscr{X}\to\Delta of  XX.

Proof.

The proof follows [HLSY21, Sections 9–11]. Note that by Proposition 2.17 and Lemma 5.6,

(H2(X,Q),qX)=(IH2(X,Q),QX).(H^{2}(X,\mathbb Q),q_{X})=(\operatorname{IH}^{2}(X,\mathbb Q),Q_{X}).

Let ρ,β,η\rho,\beta,\eta be the triple associated to the Lagrangian fibration f:XBf\colon X\to B (see Section 7.1.2), and let gρso(5)\mathfrak g_{\rho}\cong\mathfrak{so}(5) be the Lie algebra associated to ρ\rho. By [Sol20, Lemma 4.1], there is a nilpotent operator(3)(3)(3)The proof uses Meyer’s theorem on the lattice (H2(X,Z),qX)(H^{2}(X,\mathbb Z),q_{X}), which also requires b25b_{2}\geq 5. Nβ,ρN_{\beta,\rho} of index 3 in so(H2(X,Q),qX)\mathfrak{so}(H^{2}(X,\mathbb Q),q_{X}) corresponding to the pair (β,ρ)(\beta,\rho). By (5.2) and [KSV19, Lemma 3.9], we have Nβ,ρ=[Lβ,]ρgρg¯N_{\beta,\rho}=[L_{\beta},{}_{\rho}]\in\mathfrak g_{\rho}\subset\overline{\mathfrak g}.

Let WNW_{N}^{\bullet} be the weight filtration corresponding to the completion of the nilpotent operator Nβ,ρN_{\beta,\rho} to an sl2\mathfrak{sl}_{2}-triple. By Proposition 7.4 and Section 7.2.1, WNW_{N}^{\bullet} restricts to the weight filtration of the limit mixed Hodge structure IHk(X,Q)\operatorname{IH}^{k}(X,\mathbb Q) of the corresponding degeneration 𝒳\mathscr{X}\to\Delta with logarithmic monodromy operator Nρ,βN_{\rho,\beta}. On the other hand, by local Torelli and Section 4, the isotropic pairs (β,η)(\beta,\eta) induce an sl2×sl2\mathfrak{sl}_{2}\times\mathfrak{sl}_{2}-action with weight decomposition

IH(X,C)=p,qHp,q\operatorname{IH}^{*}(X,\mathbb C)=\bigoplusop\displaylimits_{p,q}H^{p,q}

such that the corresponding weight operators Hβ,HηH_{\beta},H_{\eta} satisfy

Hβ|Hp,q=(qn)id,Hη|Hp,q=(pn)id.H_{\beta}|_{H^{p,q}}=(q-n)\operatorname{id},\quad H_{\eta}|_{H^{p,q}}=(p-n)\operatorname{id}.

From the proof of Lemma 7.1, local Torelli, and the symplectic hard Lefschetz theory of Section 4.1, it follows that Hp,q=Pp,qH^{p,q}=P^{p,q}, where the Pp,qP^{p,q} are the summands of the splitting (7.1). It follows that the perverse filtration restricted to IHk(X,C)\operatorname{IH}^{k}(X,\mathbb C) agrees with the weight filtration on IHk(X,C)\operatorname{IH}^{k}(X_{\infty},\mathbb C).

References

  • [AV21] E. Amerik and M. Verbitsky, Contraction centers in families of hyperkähler manifolds, Selecta Math. (N.S.) 27 (2021), no. 4, Paper No. 60, doi:10.1007/s00029-021-00677-8.
  • [Ara90] D. Arapura, Local cohomology of sheaves of differential forms and Hodge theory, J. reine angew. Math. 409 (1990), 172–179, doi:10.1515/crll.1990.409.172.
  • [BL21] B. Bakker and C. Lehn, A global Torelli theorem for singular symplectic varieties, J. Eur. Math. Soc. (JEMS) 23 (2021), no. 3, 949–994, doi:10.4171/jems/1026.
  • [BL22] by same author, The global moduli theory of symplectic varieties, J. reine angew. Math. 790 (2022), 223–265, doi:10.1515/crelle-2022-0033.
  • [Bea00] A. Beauville, Symplectic singularities, Invent. Math. 139 (2000), no. 3, 541–549, doi:10.1007/s002229900043.
  • [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, vol. 100, Soc. Math. France, 1982.
  • [BCHM10] C. Birkar, P. Cascini, C. D. Hacon, and J. McKernan, Existence of minimal models for varieties of log general type, J. Amer. Math. Soc. 23 (2010), no. 2, 405–468, doi:10.1090/S0894-0347-09-00649-3.
  • [Bog96] F. A. Bogomolov, On the cohomology ring of a simple hyperkähler manifold (on the results of Verbitsky), Geom. Funct. Anal. 6 (1996), no. 4, 612–618, doi:10.1007/BF02247113.
  • [dCM05] M. A. A. de Cataldo and L. Migliorini, The Hodge theory of algebraic maps, Ann. Sci. École Norm. Sup. (4) 38 (2005), no. 5, 693–750, doi:10.1016/j.ansens.2005.07.001.
  • [Del71] P. Deligne, Théorie de Hodge. II, Inst. Hautes Études Sci. Publ. Math. 40 (1971), 5–57, http://www.numdam.org/item?id=PMIHES_1971__40__5_0.
  • [Del74] by same author, Théorie de Hodge. III, Inst. Hautes Études Sci. Publ. Math. 44 (1974), 5–77, http://www.numdam.org/item?id=PMIHES_1974__44__5_0.
  • [Dur95] A. H. Durfee, Intersection homology Betti numbers, Proc. Amer. Math.  Soc. 123 (1995), no. 4, 989–993, doi:10.2307/2160693.
  • [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, doi:10.1090/tran/8592.
  • [FM21] L. Fu and G. Menet, On the Betti numbers of compact holomorphic symplectic orbifolds of dimension four, Math. Z. 299 (2021), no. 1-2, 203–231, doi:10.1007/s00209-020-02682-7.
  • [Fuj80] A. Fujiki, Duality of mixed Hodge structures of algebraic varieties, Publ. Res. Inst. Math. Sci. 16 (1980), no. 3, 635–667, doi:10.2977/prims/1195186924.
  • [Fur76] H. Furstenberg, A note on Borel’s density theorem, Proc. Amer. Math. Soc. 55 (1976), no. 1, 209–212, doi:10.2307/2041874.
  • [vGe00] B. van Geemen, Kuga-Satake varieties and the Hodge conjecture, In: The arithmetic and geometry of algebraic cycles (Banff, AB, 1998), pp. 51–82, NATO Sci.  Ser. C Math. Phys. Sci., vol. 548, Kluwer Acad. Publ., Dordrecht, 2000.
  • [GGK12] M. Green, P. Griffiths, and M. Kerr, Mumford-Tate groups and domains, Ann. of Math. Stud., vol. 183, Princeton Univ. Press, Princeton, NJ, 2012.
  • [GKLR22] 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, doi:10.1007/s00208-021-02238-y.
  • [GM80] M. Goresky and R. MacPherson, Intersection homology theory, Topology 19 (1980), no. 2, 135–162, doi:10.1016/0040-9383(80)90003-8.
  • [GM83] by same author, Intersection homology II, Invent. Math. 72 (1983), no. 1, 77–129, doi:10.1007/BF01389130.
  • [Gua01] D. Guan, On the Betti numbers of irreducible compact hyperkähler manifolds of complex dimension four, Math. Res. Lett. 8 (2001), no. 5-6, 663–669, doi:10.4310/MRL.2001.v8.n5.a8.
  • [HLSY21] A. Harder, Z. Li, J. Shen, and Q. Yin, P=WP=W for Lagrangian fibrations and degenerations of hyper-Kähler manifolds, Forum Math. Sigma 9 (2021), Paper No. e50, doi:10.1017/fms.2021.31.
  • [HTT08] R. Hotta, T. Takeuchi, and T. Tanisaki DD-modules, perverse sheaves, and representation theory, Progr. Math., vol. 236, Birkhäuser Boston, Inc., Boston, MA, 2008.
  • [Huy01] D. Huybrechts, Infinitesimal variation of harmonic forms and Lefschetz decomposition, preprint arXiv:math/0102116 (2001).
  • [Huy16] by same author, Lectures on K3 surfaces, Cambridge Stud. Adv. Math., vol. 158, Cambridge Univ. Press, Cambridge, 2016, doi:10.1017/CBO9781316594193.
  • [HM22] D. Huybrechts and M. Mauri, Lagrangian fibrations, Milan J. Math. 90 (2022), no. 2, 459–483, doi:10.1007/s00032-022-00349-y.
  • [Kal06] D. Kaledin, Symplectic singularities from the Poisson point of view, J. reine angew. Math. 600 (2006), 135–156, doi:10.1515/CRELLE.2006.089.
  • [Kal09] by same author, Geometry and topology of symplectic resolutions, In: Algebraic geometry, Part 2: Seattle 2005, pp. 595–628, Proc. Sympos. Pure Math., vol. 80, Part 2, Amer. Math. Soc., Providence, RI, 2009, doi:10.1090/pspum/080.2/2483948.
  • [KS24] D. Kaplan and T. Schedler, Crepant resolutions of stratified varieties via gluing, Int. Math. Res. Not. IMRN 2024, no. 17, 12161–12200, doi:10.1093/imrn/rnae135.
  • [KS21] S. Kebekus and C. Schnell, Extending holomorphic forms from the regular locus of a complex space to a resolution of singularities, J. Amer. Math. Soc. 34 (2021), no. 2, 315–368, doi:10.1090/jams/962.
  • [KL20] Y. J. Kim and R. Laza, A conjectural bound on the second Betti number for hyper-Kähler manifolds, Bull. Soc. Math. France 148 (2020), no. 3, 467–480, doi:10.24033/bsmf.2813.
  • [Kle68] S. L. Kleiman, Algebraic cycles and the Weil conjectures, In: Dix exposés sur la cohomologie des schémas, pp. 359–386, Adv. Stud. Pure Math., vol. 3, North-Holland Publ. Co., Amsterdam, 1968.
  • [KM92] J. Kollár and S. Mori, Classification of three-dimensional flips, J. Amer. Math. Soc. 5 (1992), no. 3, 533–703, doi:10.2307/2152704.
  • [KM98] by same author, Birational geometry of algebraic varieties, (with the collaboration of C. H. Clemens and A. Corti; translated from the 1998 Japanese original), Cambridge Tracts in Math., vol. 134, Cambridge Univ. Press, Cambridge, 1998, doi:10.1017/CBO9780511662560.
  • [KS67] M. Kuga and I. Satake, Abelian varieties attached to polarized K3K_{3}-surfaces, Math. Ann. 169 (1967), 239–242, doi:10.1007/BF01399540.
  • [Kur15] N. Kurnosov, Boundness of  b2b_{2} for hyperkähler manifolds with vanishing odd-Betti numbers, preprint arXiv:1511.02838 (2015).
  • [KSV19] N. Kurnosov, A. Soldatenkov, and M. Verbitsky, Kuga-Satake construction and cohomology of hyperkähler manifolds, Adv. Math. 351 (2019), 275–295, doi:10.1016/j.aim.2019.04.060.
  • [LL97] E. Looijenga and V. A. Lunts, A Lie algebra attached to a projective variety, Invent. Math. 129 (1997), no. 2, 361–412, doi:10.1007/s002220050166.
  • [Mat01] D. Matsushita, Fujiki relation on symplectic varieties, preprint arXiv:math/0109165 (2001).
  • [Mat15] by same author, On base manifolds of Lagrangian fibrations, Sci. China Math. 58 (2015), no. 3, 531–542, doi:10.1007/s11425-014-4927-7.
  • [Men20] G. Menet, Global Torelli theorem for irreducible symplectic orbifolds, J. Math. Pures Appl. (9) 137 (2020), 213–237, doi:10.1016/j.matpur.2020.03.010.
  • [Mon13] G. Mongardi, On symplectic automorphisms of hyper-Kähler fourfolds of  K3[2]\rm K3^{[2]} type, Michigan Math. J. 62 (2013), no. 3, 537–550, doi:10.1307/mmj/1378757887.
  • [MOP20] M. Mustaţă, S. Olano, and M. Popa, Local vanishing and Hodge filtration for rational singularities, J. Inst. Math. Jussieu 19 (2020), no. 3, 801–819, doi:10.1017/s1474748018000208.
  • [MP22] M. Mustaţă and M. Popa, Hodge filtration on local cohomology, Du Bois complex and local cohomological dimension, Forum Math. Pi 10 (2022), Paper No. e22, doi:10.1017/fmp.2022.15.
  • [Nam01a] Y. Namikawa, Deformation theory of singular symplectic nn-folds, Math. Ann. 319 (2001), no. 3, 597–623, doi:10.1007/PL00004451.
  • [Nam01b] by same author, Extension of  22-forms and symplectic varieties, J. reine angew. Math. 539 (2001), 123–147, doi:10.1515/crll.2001.070.
  • [Nam01c] by same author, A note on symplectic singularities, preprint arXiv:math/0101028 (2001).
  • [Nam06] by same author, On deformations of  Q\mathbb{Q}-factorial symplectic varieties, J. reine angew. Math. 599 (2006), 97–110, doi:10.1515/CRELLE.2006.079.
  • [Sai88] M. Saito, Modules de Hodge polarisables, Publ. Res. Inst. Math. Sci.  24 (1988), no. 6, 849–995, doi:10.2977/prims/1195173930.
  • [Sai90] by same author, Mixed Hodge modules, Publ. Res. Inst. Math. Sci. 26 (1990), no. 2, 221–333, doi:10.2977/prims/1195171082.
  • [Saw22] J. Sawon, A bound on the second Betti number of hyperkähler manifolds of complex dimension six, Eur. J. Math. 8 (2022), no. 3, 1196–1212, doi:10.1007/s40879-021-00526-0.
  • [Sch73] W. Schmid, Variation of Hodge structure: The singularities of the period mapping, Invent. Math. 22 (1973), no. 3, 211–319.
  • [Sch20] M. Schwald, Fujiki relations and fibrations of irreducible symplectic varieties, Épijournal Géom. Algébrique 4 (2020), Art. 7, doi:10.46298/epiga.2020.volume4.4557.
  • [SVV23] W. Shen, S. Venkatesh, and A. D. Vo, On kk-Du Bois and kk-rational singularities, preprint arXiv:2306.03977 (2023).
  • [SY22] J. Shen and Q. Yin, Topology of Lagrangian fibrations and Hodge theory of hyper-Kähler manifolds (with Appendix B by C. Voisin), Duke Math. J. 171 (2022), no. 1, 209–241, doi:10.1215/00127094-2021-0010.
  • [Sol20] A. Soldatenkov, Limit mixed Hodge structures of hyperkähler manifolds, Mosc. Math. J. 20 (2020), no. 2, 423–436.
  • [Ste77] J. H. M. Steenbrink, Mixed Hodge structure on the vanishing cohomology, In: Real and complex singularities (Proc. Ninth Nordic Summer School/NAVF Sympos. Math., Oslo, 1976), pp. 525–563, Sijthoff & Noordhoff Internat. Publ., Alphen aan den Rijn, 1977.
  • [Tig23a] B. Tighe, The holomorphic extension property for higher Du Bois singularities, preprint arXiv:2312.01245 (2023).
  • [Tig23b] by same author, The LLV algebra for primitive symplectic varieties and applications to the Du Bois complex, Ph.D. thesis, Univ. of Illinois at Chicago, 2023, available at https://indigo.uic.edu/articles/thesis/The_LLV_Algebra_for_Primitive_Symplectic_Varieties_and_Applications_to_the_du_Bois_Complex/24241552?file=42558205.
  • [Ver90] M. Verbitsky, Action of the Lie algebra of  SO(5){\rm SO}(5) on the cohomology of a hyperkähler manifold, Funktsional. Anal. i Prilozhen. 24 (1990), no. 3, 70–71, doi:10.1007/BF01077967.
  • [Ver96] by same author, Cohomology of compact hyperkähler manifolds and its applications, Geom. Funct. Anal. 6 (1996), no. 4, 601–611, doi:10.1007/BF02247112.
  • [Ver13] by same author, Mapping class group and a global Torelli theorem for hyperkähler manifolds (with Appendix A by E. Markman), Duke Math. J. 162 (2013), no. 15, 2929–2986, doi:10.1215/00127094-2382680.
  • [Wie03] J. Wierzba, Contractions of symplectic varieties, J. Algebraic Geom. 12 (2003), no. 3, 507–534, doi:10.1090/S1056-3911-03-00318-7.