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

Equivariant Hodge modules and rational singularities

Donu Arapura and Scott Hiatt Department of Mathematics
Purdue University
West Lafayette, IN 47907
U.S.A.
Department of Mathematics
DePauw University
Greencastle, IN 46135
U.S.A.
(Date: September 17, 2025)
Abstract.

We define a notion of Hodge modules with rational singularities. A variety has rational singularities in the usual sense, if it is normal and the Hodge module related to intersection cohomology has rational singularities in the present sense. Our main result is a generalization of Boutot’s theorem that if a reductive group acts on a smooth affine variety with a stable point, and {\mathcal{M}} is an equivariant Hodge module with rational singularities, then the induced module on the GIT quotient also has rational singularities.

2020 Mathematics Subject Classification:
14B05, 14F10, 14L30
First author supported by a grant from the Simon’s foundation

The goal of this paper is to develop a theory of rational singularities with coefficients, and to give an extension of Boutot’s theorem [Bou87] to this setting. Before we state the results precisely, we recall that a Hodge module {\mathcal{M}} on a smooth variety XX consist of a 𝒟\mathcal{D}-module MM with a good filtration FF_{\bullet} and a compatible perverse sheaf of {\mathbb{Q}}-vector spaces satisfying appropriate conditions [Sai88]. A simple example is XH[dimX]{\mathbb{Q}}_{X}^{H}[\dim X], which consists of the right 𝒟\mathcal{D}-module ΩXdimX\Omega^{\dim X}_{X} with trivial filtration, and perverse sheaf X[dimX]{\mathbb{Q}}_{X}[\dim X]. Hodge modules are also defined when XX is singular. If XsmXX_{\text{sm}}\subseteq X is the smooth locus, the module XsmH[dimX]{\mathbb{Q}}_{X_{\text{sm}}}^{H}[\dim X] has a unique extension ICXHIC_{X}^{H} to a Hodge module on XX, such that the underlying perverse sheaf is the intersection cohomology complex. Associated to a Hodge module {\mathcal{M}} is the 𝒪X{\mathcal{O}}_{X}-module SX()=Fp(M)0S_{X}({\mathcal{M}})=F_{p}(M)\neq 0, where pp is chosen minimal. We also have a complex of 𝒪X{\mathcal{O}}_{X}-modules QX()Q_{X}({\mathcal{M}}), which is roughly dual to SX()S_{X}({\mathcal{M}}); more precisely, QX()=RomX(SX(),ωX)Q_{X}({\mathcal{M}})=\textbf{R}{\mathcal{H}}om_{X}(S_{X}({\mathcal{M}}),\omega^{\bullet}_{X}). Suppose that {\mathcal{M}} is a Hodge module strictly supported on XX, which means that it supported on XX, and it has no nontrivial subquotient supported on a proper subvariety of XX. We define {\mathcal{M}} to have rational singularities if QX()Q_{X}({\mathcal{M}}) is quasi-isomorphic to a sheaf placed in degree dimX-\dim X, or equivalently if SX()S_{X}({\mathcal{M}}) is a maximal Cohen-Macaulay module. We can see that XH[dimX]{\mathbb{Q}}_{X}^{H}[\dim X] has rational singularities when XX is smooth, since QX(XH[dimX])=𝒪X[dimX]Q_{X}({\mathbb{Q}}_{X}^{H}[\dim X])={\mathcal{O}}_{X}[\dim X]. More generally, XX has rational singularities in the usual sense if and only if XX is normal and ICXHIC_{X}^{H} has rational singularities in the present sense (see Corollary 1.19). This gives a characterization of rational singularities which does not depend on a resolution of singularities.

Given a regular {\mathbb{C}}-algebra RR of finite type with an action of a reductive group GG, Hochster and Roberts [HR74] proved that the ring of invariants RGR^{G} is Cohen-Macaulay. The theorem was refined by Boutot [Bou87] who proved that SpecRG\operatorname{Spec}R^{G} has rational singularities. It seems natural to consider analogues of these results for modules. Suppose that GG is a reductive group that acts effectively on a smooth affine variety X=SpecRX=\operatorname{Spec}R such that XX has a stable point, then we show that there is an equivalence of categories between the category of equivariant Hodge modules strictly supported on XX, and Hodge modules strictly supported on Y=SpecRGY=\operatorname{Spec}R^{G}.

Proposition 0.1.

Suppose that GG is a reductive group that acts effectively on XX such that there exists a stable point. Let Y=X//GY=X//G. Then there is an equivalence of categories between HMY(Y,w)HM_{Y}(Y,w) and HMX,G(X,w+d)HM_{X,G}(X,w+d) .

Suppose that {\mathcal{M}} is an equivariant Hodge module strictly support on XX, and let π+G\pi^{G}_{+}{\mathcal{M}} be the corresponding Hodge module on YY under this equivalence. Our generalization of Boutot’s theorem is as follows:

Theorem 0.2.

Let GG be a dd-dimensional reductive algebraic group with an effective action on a smooth affine variety XX such that there exists a stable point. Let π:XY=X//G\pi:X\to Y=X//G be the quotient map. Suppose that HMX,G(X,w+d){\mathcal{M}}\in HM_{X,G}(X,w+d) has rational singularities, then 𝒩=π+GHMY(Y,w)\mathcal{N}=\pi^{G}_{+}{\mathcal{M}}\in HM_{Y}(Y,w) has rational singularities, and

QY(𝒩)[d](πQX())GQ_{Y}(\mathcal{N})[d]\cong(\pi_{*}Q_{X}({\mathcal{M}}))^{G}
SY(𝒩)(πSX())G.S_{Y}(\mathcal{N})\cong(\pi_{*}S_{X}({\mathcal{M}}))^{G}.

We recover (a special case of) Boutot’s theorem by applying our theorem to =XH[dimX]{\mathcal{M}}={\mathbb{Q}}_{X}^{H}[\dim X]. It implies that the corresponding module π+G=ICYH\pi^{G}_{+}{\mathcal{M}}=IC_{Y}^{H} has rational singularities. Since YY is easily seen to be normal, it must therefore have rational singularities by the result stated in the previous paragraph.

The analogue of Hochster-Roberts for modules was considered by Stanley [Sta79] and Van den Bergh [Van89]. When AA is a finitely generated free RR-module on which GG acts compatibly, they give some criteria when AGA^{G} is Cohen-Macaulay, and they give simple examples where this can fail. As a corollary of our main theorem, we deduce a different criterion:

Corollary 0.3.

Let X=SpecRX=\operatorname{Spec}R be a smooth variety admitting an action by a reductive group GG such that the stable locus XsX^{s}\not=\emptyset. Let VV be a rational representation of GG. Then (VR)G(V\otimes_{\mathbb{C}}R)^{G} is Cohen-Macaulay provided that VRV\otimes R is a direct summand of QX()[dimX]Q_{X}({\mathcal{M}})[-\dim X] for some HMX,G(X){\mathcal{M}}\in HM_{X,G}(X) with RS.

We give an explicit special case when RR is a polynomial ring in Example 3.21.

We work with complex algebraic varieties with their analytic topologies throughout this paper. The material from the first section of this paper is drawn from the second author’s thesis [Hia23]. Our thanks to the referee for careful reading, and for many helpful comments.

1. Rational Singularities for Pure Hodge Modules

Let XX be an irreducible complex variety. Assume UXU\subset X is a smooth open subset, and let H be a variation of Hodge structure (always assumed polarizable) on U.U. Kollár [Kol86, §5] conjectured that there were objects S(X,H)S(X,\textbf{H}) and Q(X,H)Q(X,\textbf{H}) that should behave similar to ωX\omega_{X} and 𝒪X{\mathcal{O}}_{X} when XX is smooth. Specifically, S(X,H)S(X,\textbf{H}) and Q(X,H)Q(X,\textbf{H}) should have the following properties:

  • S(X,H)S(X,\textbf{H}) is a coherent sheaf and Q(X,H)Q(X,\textbf{H}) is in the bounded coherent derived category Dcohb(X)D^{b}_{coh}(X)

  • Assume f:XXf:X^{\prime}\rightarrow X is a birational projective map. Since ff is birational, XX and XX^{\prime} contain isomorphic open subsets. If H\textbf{H}^{\prime} is the pullback of the variation of Hodge structure H, then

    RfS(X,H)=S(X,H)andRfQ(X,H)=Q(X,H).\textbf{R}f_{*}S(X^{\prime},\textbf{H}^{\prime})=S(X,\textbf{H})\quad\text{and}\quad\textbf{R}f_{*}Q(X^{\prime},\textbf{H}^{\prime})=Q(X,\textbf{H}).
  • If XX is smooth and H gives a variation of Hodge structure over an open subset UXU\subset X such that X\UX\backslash U is a divisor with normal crossings only, then

    S(X,H)=ωXub(H)andQ(X,H)=lGr0(H).S(X,\textbf{H})=\omega_{X}\otimes\hfill^{u}{\mathcal{F}}^{b}(\textbf{H})\quad\text{and}\quad Q(X,\textbf{H})=\hfill^{l}Gr^{0}(\textbf{H}).

    See [Kol86, Def. 2.3] for the definition of bu(H){}^{u}{\mathcal{F}}^{b}(\textbf{H}) and Glr0(H){}^{l}Gr^{0}(\textbf{H}).

  • If H\textbf{H}^{*} is the dual of H, then S(X,H)=𝒟(Q(X,H))S(X,\textbf{H}^{*})=\mathcal{D}(Q(X,\textbf{H})) and Q(X,H)=𝒟(S(X,H))Q(X,\textbf{H}^{*})=\mathcal{D}(S(X,\textbf{H})), where 𝒟=Rom(,ωX).\mathcal{D}=\textbf{R}{\mathcal{H}}om(\bullet,\omega^{\bullet}_{X}).

Notation 1.1.

In this paper, ωXDcohb(𝒪X))\omega^{\bullet}_{X}\in D^{b}_{coh}({\mathcal{O}}_{X})) will denote the dualizing complex of XX. Note that when XX is smooth, ωXΩXdimX[dimX]\omega^{\bullet}_{X}\simeq\Omega^{\dim X}_{X}[\dim X].

This conjecture was proven by M. Saito [Sai91] using properties of Hodge modules. This section will be devoted to reviewing Saito’s results [Sai91], and applying these results to define rational singularities for Hodge modules.

Although we will mostly treat Hodge module theory as a black box, we will need to recall a few details. We refer to Saito’s foundational papers [Sai88, Sai90], and to Schnell’s overview [Sch19] for a more precise account. Let XX denote an algebraic variety of dimension nn. For simplicity, in this section, we will always assume XX embeds into a smooth variety YY of dimension mm, and let

i:X{i:X}Y{Y}

denote the embedding. Saito associates to XX a semisimple {\mathbb{Q}}-linear abelian category HM(X)HM(X) of Hodge modules, such that every simple object is generically a variation of Hodge structure on a closed irreducible subvariety of XX. For the semisimplicity statement, we need to assume that Hodge modules are polarizable. To be a bit more explicit, a Hodge module consists of a right regular holonomic 𝒟Y\mathcal{D}_{Y}-module MM supported on XX, a good filtration FF_{\bullet} on MM, and a {\mathbb{Q}}-structure provided by a perverse sheaf LL of {\mathbb{Q}}-vector spaces and an isomorphism DR(M)LDR(M)\cong{\mathbb{C}}\otimes L subject to additional conditions. Although MM is a 𝒟Y\mathcal{D}_{Y}-module, the associated graded of the de Rham complex with respect to the induced filtration

FpDR(M)=[FpmMmTYFpm+1Mm1TYFpM][m]F_{p}DR(M)=[F_{p-m}M\otimes\wedge^{m}T_{Y}\to F_{p-m+1}M\otimes\wedge^{m-1}T_{Y}\to\ldots F_{p}M][m]

is a complex of 𝒪X{\mathcal{O}}_{X}-modules. We refer the reader to [HTT08] for more details into 𝒟\mathcal{D}-modules. Below, we find it convenient to refer to 𝒟Y\mathcal{D}_{Y}-modules supported on XX simply as 𝒟X\mathcal{D}_{X}-modules. When XX is smooth, these are indeed the same thing by a theorem of Kashiwara [HTT08, Theorem 1.6.1].

The category HM(X)HM(X) has a decomposition

HM(X)=w,ZXHMZ(X,w)HM(X)=\bigoplus_{w\in{\mathbb{Z}},Z\subseteq X}HM_{Z}(X,w)

where HMZ(X,w)HM_{Z}(X,w) is the category of Hodge modules of weight ww and strict support along a closed irreducible subvariety ZXZ\subset X. A Hodge module is strictly supported on a closed subvariety ZZ, when it is supported on ZZ and it has no nontrivial subquotient supported on a proper subvariety of ZZ. One of Saito’s key results [Sai90, theorem 3.21] is that HMZ(X,w)HM_{Z}(X,w) is equivalent to the category of variations of Hodge structure of weight wdimZw-\dim Z generically defined on ZZ. If H is a variation of Hodge structure defined on a Zariski open j:UZj:U\to Z, it extends to a Hodge module with underlying perverse sheaf given by intersection cohomology complex j!L[dimZ]j_{!*}L[\dim Z]. Saito [Sai90] constructs a bigger abelian category of mixed Hodge modules MHM(X)HM(X)MHM(X)\supset HM(X). A mixed Hodge module 𝒩MHM(X)\mathcal{N}\in MHM(X) contains a finite, increasing weight filtration W𝒩W_{\bullet}\mathcal{N}, where GriW𝒩Gr^{W}_{i}\mathcal{N} is a pure Hodge module of weight ii on XX. For a mixed Hodge module 𝒩MHM(X)\mathcal{N}\in MHM(X), we will denote the above data as

𝒩=((N,FN),W𝒩,L),\mathcal{N}=((N,F_{\bullet}N),W_{\bullet}\mathcal{N},L),

where NN is the underlying 𝒟X\mathcal{D}_{X}-module and LL is the perverse sheaf. The weight filtration will be omitted if 𝒩\mathcal{N} is pure of weight ww. It is worth noting that the derived category of mixed Hodge modules has standard operations such as direct images, denoted here by f+f_{+}, and these are compatible with the corresponding operations on the constructible derived category.

We begin with a slightly generalized version of S(X,H)S(X,\textbf{H}) and Q(X,H)Q(X,\textbf{H}) given above.

Definition 1.2.

Let 𝒩MHM(X)\mathcal{N}\in MHM(X) be a mixed Hodge module. If p(𝒩)=min{p:FpN0}p(\mathcal{N})=\text{min}\{p:F_{p}N\neq 0\}, where NN is the underlying 𝒟X\mathcal{D}_{X}-module of 𝒩\mathcal{N}, then set

(1) SX(𝒩):=Grp(𝒩)FDR(𝒩)=Fp(𝒩)NS_{X}(\mathcal{N}):=Gr^{F}_{p(\mathcal{N})}DR(\mathcal{N})=F_{p(\mathcal{N})}N
(2) QX(𝒩):=𝒟(SX(𝔻(𝒩)))Dcohb(X),Q_{X}(\mathcal{N}):=\mathcal{D}(S_{X}(\mathbb{D}(\mathcal{N})))\in D^{b}_{coh}(X),

where 𝒟()=RomX(,ωX)\mathcal{D}(\bullet)=\textbf{R}{\mathcal{H}}om_{X}(\bullet,\omega^{\bullet}_{X}) and 𝔻(𝒩)\mathbb{D}(\mathcal{N}) is the dual Hodge Module.

Proposition 1.3.

[Sai88] If 𝒩MHM(X)\mathcal{N}\in MHM(X), then we have the quasi-isomorphism

RomX(GrpFDR(𝒩),ωX)GrpFDR(𝔻(𝒩))for every p,\textbf{R}{\mathcal{H}}om_{X}(Gr^{F}_{p}DR(\mathcal{N}),\omega^{\bullet}_{X})\simeq Gr^{F}_{-p}DR(\mathbb{D}(\mathcal{N}))\quad\text{for every $p\in{\mathbb{Z}}$},

where 𝔻(𝒩)\mathbb{D}(\mathcal{N}) is the dual Hodge module.

Remark 1.4.

When 𝒩HM(X,w)\mathcal{N}\in HM(X,w) is a pure Hodge module of weight ww, then a polarization gives an isomorphism 𝔻(𝒩)𝒩(w),\mathbb{D}(\mathcal{N})\cong\mathcal{N}(w), where

𝒩(w)=((N,FwN),W2w𝒩,L(w))\mathcal{N}(w)=((N,F_{\bullet-w}N),W_{\bullet-2w}\mathcal{N},L\otimes_{{\mathbb{Q}}}{\mathbb{Q}}(w))

is the Tate twist of 𝒩\mathcal{N} by the weight ww. Therefore, QX(𝒩)=𝒟(SX(𝒩)).Q_{X}(\mathcal{N})=\mathcal{D}(S_{X}(\mathcal{N})).

Lemma 1.5.

If 𝒩HM(X,w)\mathcal{N}\in HM(X,w) is a pure Hodge module of weight ww, then QX(𝒩)Grp(𝒩)wFDR(𝒩)Q_{X}(\mathcal{N})\simeq Gr^{F}_{-p(\mathcal{N})-w}DR(\mathcal{N}).

Proof.

By the previous remark, 𝔻(𝒩)𝒩(w)\mathbb{D}(\mathcal{N})\cong\mathcal{N}(w). Therefore, p(𝔻(𝒩))=p(𝒩)+wp(\mathbb{D}(\mathcal{N}))=p(\mathcal{N})+w. Applying Proposition 1.3 we obtain

QX(𝒩)=𝒟(Grp(𝒩)FDR(𝒩))=𝒟(Grp(𝒩)+wFDR(𝒩(w)))Grp(𝒩)wFDR(𝔻(𝒩(w)))Q_{X}(\mathcal{N})=\mathcal{D}(Gr^{F}_{p(\mathcal{N})}DR(\mathcal{N}))=\mathcal{D}(Gr^{F}_{p(\mathcal{N})+w}DR(\mathcal{N}(w)))\simeq Gr^{F}_{-p(\mathcal{N})-w}DR(\mathbb{D}(\mathcal{N}(w)))
=Grp(𝒩)wFDR(𝒩).=Gr^{F}_{-p(\mathcal{N})-w}DR(\mathcal{N}).

Remark 1.6.

If XX is irreducible, and 𝒩HMX(X,w)\mathcal{N}\in HM_{X}(X,w) is a Hodge module strictly supported on XX, then there exists a variation of Hodge structure H on a smooth open subset UXU\subset X such that 𝒩\mathcal{N} restricted to UU is isomorphic to H [Sai90, Thm. 3.21]. In this situation, SX(𝒩)=S(X,H)S_{X}(\mathcal{N})=S(X,\textbf{H}) and QX(𝒩)=Q(X,H)Q_{X}(\mathcal{N})=Q(X,\textbf{H}) satisfy the conditions conjectured above.

Example 1.7.

Assume XX is a smooth irreducible variety. If 𝒩=XH[n],\mathcal{N}={\mathbb{Q}}_{X}^{H}[n], then 𝒩\mathcal{N} is a pure Hodge module of weight n,n, and

p(XH[n])=nandSX(XH[n])=ΩXn.p({\mathbb{Q}}^{H}_{X}[n])=-n\quad\text{and}\quad S_{X}({\mathbb{Q}}^{H}_{X}[n])=\Omega^{n}_{X}.

By Lemma 1.5, we have

QX(XH[n])=Gr0F(DR(XH[n]))=𝒪X[n].Q_{X}({\mathbb{Q}}^{H}_{X}[n])=Gr^{F}_{0}(DR({\mathbb{Q}}^{H}_{X}[n]))=\mathcal{O}_{X}[n].
Proposition 1.8.

[KS21, Prop. 4.7] If XX is smooth, 𝒩MHM(X)\mathcal{N}\in MHM(X), and α=min{w:Ww𝒩0}\alpha=min\{w:W_{w}\mathcal{N}\neq 0\}, then

GrpFDR(𝒩)0unless p(𝒩)pp(𝒩)α.Gr^{F}_{p}DR(\mathcal{N})\simeq 0\quad\text{unless $p(\mathcal{N})\leq p\leq-p(\mathcal{N})-\alpha$}.
Proof.

By definition of p(𝒩)p(\mathcal{N}), we know GrpFDR(𝒩)=0Gr^{F}_{p}DR(\mathcal{N})=0 if p<p(𝒩)p<p(\mathcal{N}). So, we just have to show the other inequality.

When 𝒩\mathcal{N} is a pure Hodge module of weight α\alpha, there is a quasi-isomorphism

RomX(GrpFDR(𝒩),ωX)GrpαFDR(𝒩).\textbf{R}{\mathcal{H}}om_{X}(Gr^{F}_{p}DR(\mathcal{N}),\omega^{\bullet}_{X})\simeq Gr^{F}_{-p-\alpha}DR(\mathcal{N}).

If p>p(𝒩)αp>-p(\mathcal{N})-\alpha, then pα<p(𝒩)-p-\alpha<p(\mathcal{N}) and GrpαFDR(𝒩)=0Gr^{F}_{-p-\alpha}DR(\mathcal{N})=0. The quasi-isomorphism above implies GrpFDR(𝒩)0Gr^{F}_{p}DR(\mathcal{N})\simeq 0 for p>p(𝒩)αp>-p(\mathcal{N})-\alpha.

Now, we have the exact sequence of mixed Hodge modules

0GrαW𝒩𝒩𝒩/Wα𝒩0.0\rightarrow Gr^{W}_{\alpha}\mathcal{N}\rightarrow\mathcal{N}\rightarrow\mathcal{N}/W_{\alpha}\mathcal{N}\rightarrow 0.

There is an exact triangle

GrpFDR(GrαW𝒩){Gr^{F}_{p}DR(Gr^{W}_{\alpha}\mathcal{N})}GrpFDR(𝒩){Gr^{F}_{p}DR(\mathcal{N})}GrpFDR(𝒩/Wα𝒩){Gr^{F}_{p}DR(\mathcal{N}/W_{\alpha}\mathcal{N})}.{\hfill.}+1\scriptstyle{+1}

If p>p(𝒩)αp>-p(\mathcal{N})-\alpha, then GrpFDR(GrαW𝒩)0Gr^{F}_{p}DR(Gr^{W}_{\alpha}\mathcal{N})\simeq 0. So it suffices to show GrpFDR(𝒩/Wα𝒩)=0Gr^{F}_{p}DR(\mathcal{N}/W_{\alpha}\mathcal{N})=0 for p>p(𝒩)αp>-p(\mathcal{N})-\alpha. There is an exact triangle

GrpFDR(Grα+1W𝒩){Gr^{F}_{p}DR(Gr^{W}_{\alpha+1}\mathcal{N})}GrpFDR(𝒩/Wα𝒩){Gr^{F}_{p}DR(\mathcal{N}/W_{\alpha}\mathcal{N})}GrpFDR(𝒩/Wα+1𝒩){Gr^{F}_{p}DR(\mathcal{N}/W_{\alpha+1}\mathcal{N})}.{\hfill.}+1\scriptstyle{+1}

For the mixed Hodge module 𝒩/Wα𝒩\mathcal{N}/W_{\alpha}\mathcal{N}, we must have p(𝒩/Wα𝒩)p(𝒩)p(\mathcal{N}/W_{\alpha}\mathcal{N})\geq p(\mathcal{N}). So we must have p(Grα+1W𝒩)p(𝒩)p(Gr^{W}_{\alpha+1}\mathcal{N})\geq p(\mathcal{N}) as well. This forces GrpFDR(Grα+1W𝒩)0Gr^{F}_{p}DR(Gr^{W}_{\alpha+1}\mathcal{N})\simeq 0 for p>p(Grα+1W𝒩)α1.p>-p(Gr^{W}_{\alpha+1}\mathcal{N})-\alpha-1. In particular GrpFDR(Grα+1W𝒩)0Gr^{F}_{p}DR(Gr^{W}_{\alpha+1}\mathcal{N})\simeq 0 for p>p(𝒩)α.p>-p(\mathcal{N})-\alpha. So now it suffices to show GrpFDR(𝒩/Wα+1𝒩)=0Gr^{F}_{p}DR(\mathcal{N}/W_{\alpha+1}\mathcal{N})=0 for p>p(𝒩)αp>-p(\mathcal{N})-\alpha. Continuing in this same fashion, and using the fact that GrpFDR(Grα+iW𝒩)0Gr^{F}_{p}DR(Gr^{W}_{\alpha+i}\mathcal{N})\simeq 0 whenever i>0i>0 and p>p(𝒩)αp>-p(\mathcal{N})-\alpha, we must have GrpFDR(𝒩)0Gr^{F}_{p}DR(\mathcal{N})\simeq 0 for p>p(𝒩)αp>-p(\mathcal{N})-\alpha because the weight filtration is finite.

Assume XX is a smooth irreducible variety. Let 𝒩HMX(X,w)\mathcal{N}\in HM_{X}(X,w) and assume the singularities of 𝒩\mathcal{N} are along a simple normal crossing divisor DD, i.e. suppose that the restriction of 𝒩\mathcal{N} to X\DX\backslash D is a variation of Hodge structure H=((,F),L)\textbf{H}=(({\mathcal{L}},F^{\bullet}),L) of weight wnw-n on the open set U=X\D.U=X\backslash D. Let ι:DX\iota:D\hookrightarrow X denote the natural map. If j:UXj:U\hookrightarrow X is the natural map, recall that the Hodge filtration on 𝒩\mathcal{N} is given by

(3) Fp𝒩=i0(ΩXn(~>1jFipn))Fi𝒟X,\displaystyle F_{p}\mathcal{N}=\sum_{i\geq 0}\bigg{(}\Omega^{n}_{X}\otimes(\tilde{{\mathcal{L}}}^{>-1}\cap j_{\ast}F^{i-p-n}{\mathcal{L}})\bigg{)}F_{i}\mathcal{D}_{X},

where ~>α\tilde{{\mathcal{L}}}^{>\alpha} is Deligne’s extension of {\mathcal{L}} with eigenvalues of residue of connection in (α,α+1](\alpha,\alpha+1] [Sai17].

Proposition 1.9.

For the Hodge module 𝒩\mathcal{N} above, we have

(4) SX(𝒩)=ΩXn(~>1jFp(𝒩)n).S_{X}(\mathcal{N})=\Omega^{n}_{X}\otimes(\tilde{{\mathcal{L}}}^{>-1}\cap j_{\ast}F^{-p(\mathcal{N})-n}{\mathcal{L}}).
(5) QX(𝒩)~0/(~0jFw+p(𝒩)+1)[n].Q_{X}(\mathcal{N})\simeq\tilde{{\mathcal{L}}}^{\geq 0}/(\tilde{{\mathcal{L}}}^{\geq 0}\cap j_{\ast}F^{w+p(\mathcal{N})+1}{\mathcal{L}})[n].
Proof.

This is discussed in [Sai91]. For convenience to the reader, a proof will be given.

Equation (4) is clear by the definition of the Hodge filtration for 𝒩\mathcal{N}. To show equation (5), consider the exact triangle

ι+ι!𝒩{\iota_{+}\iota^{!}\mathcal{N}}𝒩{\mathcal{N}}j+j𝒩{j_{+}j^{*}\mathcal{N}}.{\hfill.}+\scriptstyle{+}

Since 𝒩\mathcal{N} has no non-trivial sub-object with support on DD, there is an exact sequence

0𝒩𝒩(D)ι+1ι!𝒩0.0\rightarrow\mathcal{N}\rightarrow\mathcal{N}(*D)\rightarrow\iota_{+}{\mathcal{H}}^{1}\iota^{!}\mathcal{N}\rightarrow 0.

For notation, let D1(𝒩)=ι+1ι!(𝒩).{\mathcal{H}}^{1}_{D}(\mathcal{N})=\iota_{+}{\mathcal{H}}^{1}\iota^{!}(\mathcal{N}). Since 𝒩\mathcal{N} has pure weight ww, we have GriW(D1(𝒩))=0Gr^{W}_{i}({\mathcal{H}}^{1}_{D}(\mathcal{N}))=0 for i<w+1i<w+1 [Sai90, Prop. 2.26]. By Proposition 1.8, we obtain

Grp(𝒩)wFDR(D1(𝒩))0.Gr^{F}_{-p(\mathcal{N})-w}DR({\mathcal{H}}^{1}_{D}(\mathcal{N}))\simeq 0.

So, from the exact sequence above,

QX(𝒩)=Grp(𝒩)wFDR(𝒩)Grp(𝒩)wFDR(𝒩(D)).Q_{X}(\mathcal{N})=Gr^{F}_{-p(\mathcal{N})-w}DR(\mathcal{N})\simeq Gr^{F}_{-p(\mathcal{N})-w}DR(\mathcal{N}(\ast D)).

By [Sai90, Prop. 3.11], there is a quasi-isomorphism

Grp(𝒩)wFDR(𝒩(D))Grp(𝒩)wF~0Ω1(logD)Grp(𝒩)w+1F~0.Gr^{F}_{-p(\mathcal{N})-w}DR(\mathcal{N}(\ast D))\simeq Gr^{F}_{-p(\mathcal{N})-w}\tilde{{\mathcal{L}}}^{\geq 0}\rightarrow\Omega^{1}(\log D)\otimes Gr^{F}_{-p(\mathcal{N})-w+1}\tilde{{\mathcal{L}}}^{\geq 0}\rightarrow\cdots.

On the variation of Hodge structure H,\textbf{H}, we have p(𝒩)+w=min{p:GrFp0}p(\mathcal{N})+w=\text{min}\{p:Gr^{p}_{F}{\mathcal{L}}\neq 0\}. So, when i>0i>0, we must have

Grp(𝒩)w+iF~0=(~0jFp(𝒩)+wi)/(~0jFp(𝒩)+wi+1)=0.Gr^{F}_{-p(\mathcal{N})-w+i}\tilde{{\mathcal{L}}}^{\geq 0}=(\tilde{{\mathcal{L}}}^{\geq 0}\cap j_{\ast}F^{p(\mathcal{N})+w-i}{\mathcal{L}})/(\tilde{{\mathcal{L}}}^{\geq 0}\cap j_{\ast}F^{p(\mathcal{N})+w-i+1}{\mathcal{L}})=0.

Hence,

QX(𝒩)Grp(𝒩)wF~0[n]=~0/(~0jFw+p(𝒩)+1)[n].Q_{X}(\mathcal{N})\simeq Gr^{F}_{-p(\mathcal{N})-w}\tilde{{\mathcal{L}}}^{\geq 0}[n]=\tilde{{\mathcal{L}}}^{\geq 0}/(\tilde{{\mathcal{L}}}^{\geq 0}\cap j_{\ast}F^{w+p(\mathcal{N})+1}{\mathcal{L}})[n].

Assume XX is an irreducible variety and 𝒩HMX(X,w)\mathcal{N}\in HM_{X}(X,w). If XX is smooth, and 𝒩\mathcal{N} is the trivial Hodge module, then we see that QX(𝒩)Q_{X}(\mathcal{N}) is given by a coherent sheaf (up to a shift). Similarly, if XX is smooth, and 𝒩\mathcal{N} is a variation of Hodge structure on XX, then QX(𝒩)Q_{X}(\mathcal{N}) is a coherent sheaf. This continues to hold when the singularity of 𝒩\mathcal{N} is a simple normal crossing divisor. In general, the object QX(𝒩)Q_{X}(\mathcal{N}) is not a coherent sheaf, but QX(𝒩)Dcohb(X)Q_{X}(\mathcal{N})\in D^{b}_{coh}(X). We will investigate Hodge modules on XX that have the property that QX(𝒩)Q_{X}(\mathcal{N}) is just a coherent sheaf.

Definition 1.10.

If XX is an irreducible variety, a Hodge module 𝒩HMX(X,w)\mathcal{N}\in HM_{X}(X,w) has rational singularities (or has RS) if

QX(𝒩)n(QX(𝒩))[n].Q_{X}(\mathcal{N})\simeq{\mathcal{H}}^{-n}(Q_{X}(\mathcal{N}))[n].
Lemma 1.11.

If XX is a smooth irreducible variety, then 𝒩HMX(X,w)\mathcal{N}\in HM_{X}(X,w) has RS if it has singularities along a divisor with normal crossings.

Proof.

This follows from Proposition 1.9. ∎

Lemma 1.12.

𝒩HMX(X,w)\mathcal{N}\in HM_{X}(X,w) has RS if and only if SX(𝒩)S_{X}(\mathcal{N}) is Cohen-Macaulay of pure dimension nn.

Proof.

Given the embedding i:XYi:X\hookrightarrow Y with YY smooth, by Remark 1.4 there is a quasi-isomorphism

iQX(𝒩)RomY(iSX(𝒩),ωY).i_{\ast}Q_{X}(\mathcal{N})\simeq\textbf{R}{\mathcal{H}}om_{Y}(i_{\ast}S_{X}(\mathcal{N}),\omega^{\bullet}_{Y}).

We see that QX(𝒩)n(QX(𝒩))[n]Q_{X}(\mathcal{N})\simeq{\mathcal{H}}^{-n}(Q_{X}(\mathcal{N}))[n] if and only if

xtYi(iSX(𝒩),ωY)=0unless i=n.{\mathcal{E}}xt^{i}_{Y}(i_{\ast}S_{X}(\mathcal{N}),\omega^{\bullet}_{Y})=0\quad\text{unless $i=-n$}.

Which is equivalent to SX(𝒩)S_{X}(\mathcal{N}) being Cohen-Macaulay of pure dimension nn [Kem+73, Chap. I, §3]. ∎

Corollary 1.13.

Assume XX is smooth. If 𝒩HMX(X,w)\mathcal{N}\in HM_{X}(X,w) has RS, then SX(𝒩)S_{X}(\mathcal{N}) and QX(𝒩)Q_{X}(\mathcal{N}) are (translates of) locally free sheaves.

Proof.

Locally, SX(𝒩)S_{X}(\mathcal{N}) is a maximal Cohen-Macaulay module. It is well known that a maximal Cohen-Macaulay module on a regular local ring is free. ∎

Next, we will explain a characterization of rational singularities for Hodge modules which is closer to the original definition. Specifically, we will show we can define rational singularities for Hodge modules by a resolution of singularities. To do this, we will need Saito’s main result from [Sai91]. Following Saito [Sai91], we define

q(𝒩):=np(𝒩)andq(𝒩):=p(𝒩)+w.q(\mathcal{N}):=-n-p(\mathcal{N})\quad\text{and}\quad q^{\prime}(\mathcal{N}):=p(\mathcal{N})+w.
Theorem 1.14.

[Sai91, Thm. 3.2] Let f:XXf:X\rightarrow X^{\prime} be a proper surjective morphism of irreducible varieties with d=dim(X)dim(X)d=dim(X)-dim(X^{\prime}), and HMX(X,w){\mathcal{M}}\in HM_{X}(X,w). If XiHMX(X,w+i){\mathcal{M}}^{i}_{X^{\prime}}\in HM_{X^{\prime}}(X^{\prime},w+i) is the direct factor of Hif+H^{i}f_{+}{\mathcal{M}} from the decomposition of strict support, then we have the (noncanonical) isomorphisms in Dcohb(X)D^{b}_{coh}(X^{\prime}):

RfSX()q(Xi)=q()+dSX(Xi)[i],\textbf{R}f_{\ast}S_{X}({\mathcal{M}})\simeq\displaystyle\bigoplus_{q({\mathcal{M}}^{i}_{X^{\prime}})=q({\mathcal{M}})+d}S_{X^{\prime}}({\mathcal{M}}_{X^{\prime}}^{i})[-i],
RfQX()q(Xi)=q()QX(Xi)[i],\textbf{R}f_{\ast}Q_{X}({\mathcal{M}})\simeq\displaystyle\bigoplus_{q^{\prime}({\mathcal{M}}^{i}_{X^{\prime}})=q^{\prime}({\mathcal{M}})}Q_{X^{\prime}}({\mathcal{M}}_{X^{\prime}}^{i})[-i],

and canonical isomorphisms

RifSX()=SX(Xi)for q(Xi)=q()+d R^{i}f_{\ast}S_{X}({\mathcal{M}})=S_{X^{\prime}}({\mathcal{M}}_{X^{\prime}}^{i})\quad\text{for $q({\mathcal{M}}^{i}_{X^{\prime}})=q({\mathcal{M}})+d$ }
HidRfQX()=QX(Xi)for q(Xi)=q(), {}^{d}H^{i}\textbf{R}f_{\ast}Q_{X}({\mathcal{M}})=Q_{X^{\prime}}({\mathcal{M}}_{X^{\prime}}^{i})\quad\text{for $q^{\prime}({\mathcal{M}}^{i}_{X^{\prime}})=q^{\prime}({\mathcal{M}}),$ }

where HidRfQX()=𝒟(Hi(𝒟(RfQX()))).{}^{d}H^{i}\textbf{R}f_{\ast}Q_{X}({\mathcal{M}})=\mathcal{D}(H^{-i}(\mathcal{D}(\textbf{R}f_{\ast}Q_{X}({\mathcal{M}})))).

Setting 1.15.

Let XX be an irreducible variety. If H=((,F),L)\textbf{H}=(({\mathcal{L}},F),L) is a variation of Hodge structure of weight wnw-n on a smooth open Zariski dense subset UU of XX, then there exists a unique Hodge module =((M,F),K)HMX(X,w){\mathcal{M}}=((M,F),K)\in HM_{X}(X,w) such that |UH{\mathcal{M}}|_{U}\cong\textbf{H}. If π:X~X\pi:\tilde{X}\rightarrow X is a resolution of singularities, then there exists a smooth open Zariski dense subset VV such that π|π1(V):π1(V)V\pi|_{\pi^{-1}(V)}:\pi^{-1}(V)\rightarrow V is an isomorphism. If we restrict the variation of Hodge structure H to the open subset UVU\cap V, then there is a variation of Hodge structure H~\tilde{\textbf{H}} on the open set π1(VU)\pi^{-1}(V\cap U). Hence there also exists a unique Hodge module ~=((M~,F),K~)HMX~(X~,w)\tilde{{\mathcal{M}}}=((\tilde{M},F),\tilde{K})\in HM_{\tilde{X}}(\tilde{X},w) such that ~|π1(VU)H~\tilde{{\mathcal{M}}}|_{\pi^{-1}(V\cap U)}\cong\tilde{\textbf{H}}.

Lemma 1.16.

In the setting 1.15, {\mathcal{M}} is the only direct factor of the Hodge module H0π+~H^{0}\pi_{+}\tilde{{\mathcal{M}}} with strict support on XX.

Proof.

By the condition of strict support, we have H0π+~=ZH^{0}\pi_{+}\tilde{{\mathcal{M}}}=\bigoplus{\mathcal{M}}_{Z}, where ZHMZ(X,w){\mathcal{M}}_{Z}\in HM_{Z}(X,w). There is also an isomorphism H0π+~|VUH|VUH^{0}\pi_{+}\tilde{{\mathcal{M}}}|_{V\cap U}\cong\textbf{H}|_{V\cap U}. Therefore, if HMX(X,w){\mathcal{M}}^{\prime}\in HM_{X}(X,w) is the unique Hodge module that extends H|VU\textbf{H}|_{V\cap U}, then {\mathcal{M}}^{{}^{\prime}} is the only direct factor of H0π+~H^{0}\pi_{+}\tilde{{\mathcal{M}}} with strict support on XX. So, we need to show ={\mathcal{M}}={\mathcal{M}}^{{}^{\prime}}. By uniqueness of {\mathcal{M}}, it suffices to show |UH{\mathcal{M}}^{\prime}|_{U}\cong\textbf{H}. By the Riemann-Hilbert correspondence, it suffices to show the underlying perverse sheaf of |U{\mathcal{M}}^{\prime}|_{U} is LL. If jVU:VUXj_{V\cap U}:V\cap U\hookrightarrow X is the natural inclusion, then the perverse sheaf of {\mathcal{M}}^{\prime} is given by the intermediate extension j(VU)!L|VUj_{(V\cap U)!\ast}L|_{V\cap U}. If jU:UXj_{U}:U\hookrightarrow X and j:VUUj:V\cap U\hookrightarrow U are the natural maps, then by the properties of the intermediary extension, there are isomorphisms

Lj!L|VUjU1(j(VU)!L|VU).L\cong j_{!\ast}L|_{V\cap U}\cong j_{U}^{-1}(j_{(V\cap U)!\ast}L|_{V\cap U}).

See [Dim92, Remark 5.2.7] for more details.

Proposition 1.17.

In the setting 1.15, there are quasi-isomorphisms

RπSX~(~)πSX~(~)SX()andRπQX~(~)QX().\textbf{R}\pi_{\ast}S_{\tilde{X}}(\tilde{{\mathcal{M}}})\simeq\pi_{\ast}S_{\tilde{X}}(\tilde{{\mathcal{M}}})\simeq S_{X}({\mathcal{M}})\quad\text{and}\quad\textbf{R}\pi_{\ast}Q_{\tilde{X}}(\tilde{{\mathcal{M}}})\simeq Q_{X}({\mathcal{M}}).
Proof.

From Theorem 1.14,

RπSX~(~)q(Hiπ+(~))=q(~)SX(Hiπ+(~))[i].\textbf{R}\pi_{\ast}S_{\tilde{X}}(\tilde{{\mathcal{M}}})\simeq\displaystyle\bigoplus_{q(H^{i}\pi_{+}(\tilde{{\mathcal{M}}}))=q(\tilde{{\mathcal{M}}})}S_{X}(H^{i}\pi_{+}(\tilde{{\mathcal{M}}}))[-i].

But, for i0i\neq 0, the support of Hiπ+(~)H^{i}\pi_{+}(\tilde{{\mathcal{M}}}) is contained in a proper closed subset of XX. Therefore, by combining Theorem 1.14 and Lemma 1.16,

RπSX~(~)πSX~(~)SX(H0π+(~))SX().\textbf{R}\pi_{\ast}S_{\tilde{X}}(\tilde{{\mathcal{M}}})\simeq\pi_{\ast}S_{\tilde{X}}(\tilde{{\mathcal{M}}})\cong S_{X}(H^{0}\pi_{+}(\tilde{{\mathcal{M}}}))\cong S_{X}({\mathcal{M}}).

The second quasi-isomorphism follows from duality (Remark 1.4). ∎

Corollary 1.18.

Given a Hodge module HMX(X,w){\mathcal{M}}\in HM_{X}(X,w), choose a smooth open set UXU\subset X, such that |U{\mathcal{M}}|_{U} is a variation of Hodge structure. Fix a resolution of singularities π:X~X\pi:\tilde{X}\to X, such that π1(XU)\pi^{-1}(X-U) is a divisor with normal crossings. Then,

  1. (1)

    {\mathcal{M}} has RS if and only if RiπQX~(~)=0R^{i}\pi_{\ast}Q_{\tilde{X}}(\tilde{{\mathcal{M}}})=0 for i>ni>-n;

  2. (2)

    SX()S_{X}({\mathcal{M}}) is torsion free.

Proof.

The first item follows from the definition of RS for Hodge modules and Proposition 1.17. For the second item, SX~()S_{\tilde{X}}({\mathcal{M}}) is torsion free because it is locally free (Corollary 1.13). Hence πSX~(~)SX()\pi_{*}S_{\tilde{X}}(\tilde{{\mathcal{M}}})\cong S_{X}({\mathcal{M}}) is torsion free. ∎

Corollary 1.19.

Let UU be the nonsingular set of XX and ICXHHMX(X,n)IC^{H}_{X}\in HM_{X}(X,n) the unique Hodge module that extends the constant variation of Hodge structure on UU. The variety XX has rational singularities if and only if XX is normal and ICXHIC^{H}_{X} has rational singularities.

Proof.

Let π:X~X\pi:\tilde{X}\rightarrow X be a resolution of singularities. The Hodge module ICXHIC^{H}_{X} is a direct factor of H0π+X~H[n]H^{0}\pi_{+}{\mathbb{Q}}^{H}_{\tilde{X}}[n], and

Rπ𝒪X~[n]=RπQX~(X~H[n])QX(ICXH).\textbf{R}\pi_{\ast}{\mathcal{O}}_{\tilde{X}}[n]=\textbf{R}\pi_{\ast}Q_{\tilde{X}}({\mathbb{Q}}^{H}_{\tilde{X}}[n])\simeq Q_{X}(IC^{H}_{X}).

Hence ICXHIC^{H}_{X} has RS if and only if the higher direct images of 𝒪X~[n]{\mathcal{O}}_{\tilde{X}}[n] vanish. So, XX has rational singularities if and only if XX is normal and ICXHIC^{H}_{X} has RS. ∎

Using the definition of rational singularities for Hodge modules with Theorem 1.14, we also have a theorem analogous to Kollár’s result [Kol86a, Prop. 3.12].

Theorem 1.20.

Let f:XXf:X\rightarrow X^{\prime} be a surjective projective map between irreducible algebraic varieties of dimension nn and mm, respectively, and let d=nmd=n-m. If 𝒩HMX(X,w)\mathcal{N}\in HM_{X}(X,w), then the following are equivalent:

  1. (1)

    RifQX(𝒩)R^{i}f_{\ast}Q_{X}(\mathcal{N}) is torsion free for all ii.

  2. (2)

    RfQX(𝒩)iRifQX(𝒩)[i]\textbf{R}f_{\ast}Q_{X}(\mathcal{N})\simeq\displaystyle\bigoplus_{i}R^{i}f_{\ast}Q_{X}(\mathcal{N})[-i].

  3. (3)

    If 𝒩XjHMX(X,w+j)\mathcal{N}^{j}_{X^{\prime}}\in HM_{X^{\prime}}(X^{\prime},w+j) is the largest direct factor of Hjf+𝒩H^{j}f_{+}\mathcal{N} with strict support XX^{\prime}, then 𝒩Xj\mathcal{N}^{j}_{X^{\prime}} has RS.

Proof.

(1)(3):(1)\Rightarrow(3): By duality, and Theorem 1.14, we have the quasi-isomorphisms

RfQX(𝒩)RomX(RfSX(𝒩),ωX)0jdRomX(RjfSX(𝒩)[j],ωX)\textbf{R}f_{\ast}Q_{X}(\mathcal{N})\simeq\textbf{R}{\mathcal{H}}om_{X^{\prime}}(\textbf{R}f_{\ast}S_{X}(\mathcal{N}),\omega^{\bullet}_{X^{\prime}})\simeq\displaystyle\bigoplus_{0\leq j\leq d}\textbf{R}{\mathcal{H}}om_{X^{\prime}}(R^{j}f_{\ast}S_{X}(\mathcal{N})[-j],\omega^{\bullet}_{X^{\prime}})
0jdRomX(SX(𝒩Xj)[j],ωX).\simeq\displaystyle\bigoplus_{0\leq j\leq d}\textbf{R}{\mathcal{H}}om_{X^{\prime}}(S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}})[-j],\omega^{\bullet}_{X^{\prime}}).

To show 𝒩Xj\mathcal{N}^{j}_{X^{\prime}} has RS, it suffices to show SX(𝒩Xj)S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}}) is Cohen-Macaulay of pure dimension mm by Lemma 1.12. From the above formula, we obtain

RifQX(𝒩)0jdxtXi+j(SX(𝒩Xj),ωX).R^{i}f_{\ast}Q_{X}(\mathcal{N})\simeq\displaystyle\bigoplus_{0\leq j\leq d}{\mathcal{E}}xt^{i+j}_{X^{\prime}}(S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}}),\omega^{\bullet}_{X^{\prime}}).

By [Stacks, Tag 0ECM] (or by locally embedding XX^{\prime} into a smooth variety and applying [Gro67, Lemma 6.5]), we see that the dimension of support of xtXi+j(SX(𝒩Xj),ωX){\mathcal{E}}xt^{i+j}_{X^{\prime}}(S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}}),\omega^{\bullet}_{X^{\prime}}) is less than mm if i+j>mi+j>-m. Hence if RifQX(𝒩)R^{i}f_{\ast}Q_{X}(\mathcal{N}) is torsion free for every ii, then

xtXi+jSX(𝒩Xj),ωX)=0unless i+j=m.{\mathcal{E}}xt^{i+j}_{X^{\prime}}S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}}),\omega^{\bullet}_{X^{\prime}})=0\quad\text{unless $i+j=-m$}.

This is equivalent to SX(𝒩Xj)pS_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}})_{p} being Cohen-Macaulay of pure dimension mm.

(3)(2):(3)\Rightarrow(2): If 𝒩Xj\mathcal{N}^{j}_{X^{\prime}} has RS for every jj, then

RfQX(𝒩)0jdRomX(SX(𝒩Xj)[j],ωX)0jdRomX(SX(𝒩Xj),ωX)[j]\textbf{R}f_{\ast}Q_{X}(\mathcal{N})\simeq\displaystyle\bigoplus_{0\leq j\leq d}\textbf{R}{\mathcal{H}}om_{X^{\prime}}(S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}})[-j],\omega^{\bullet}_{X^{\prime}})\simeq\displaystyle\bigoplus_{0\leq j\leq d}\textbf{R}{\mathcal{H}}om_{X^{\prime}}(S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}}),\omega^{\bullet}_{X^{\prime}})[j]
0jdQX(𝒩Xj)[j]0jdHm(QX(𝒩Xj))[j+m].\simeq\displaystyle\bigoplus_{0\leq j\leq d}Q_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}})[j]\simeq\displaystyle\bigoplus_{0\leq j\leq d}H^{-m}(Q_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}}))[j+m].

Therefore, RmjfQX(𝒩)Hm(QX(𝒩Xj))R^{-m-j}f_{\ast}Q_{X}(\mathcal{N})\cong H^{-m}(Q_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}})) and

RfQX(𝒩)0jdHm(QX(𝒩Xj))[j+m]0jdRmjfQX(𝒩)[m+j].\textbf{R}f_{\ast}Q_{X}(\mathcal{N})\simeq\bigoplus_{0\leq j\leq d}H^{-m}(Q_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}}))[j+m]\simeq\bigoplus_{0\leq j\leq d}R^{-m-j}f_{\ast}Q_{X}(\mathcal{N})[m+j].

(2)(1)(2)\Rightarrow(1): By duality, and Theorem 1.14, we have the quasi-isomorphisms

0jdSX(𝒩Xj)[j]RfSX(𝒩)𝒟(RfQX(𝒩))i𝒟(RifQX(𝒩)[i]).\bigoplus_{0\leq j\leq d}S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}})[-j]\simeq\textbf{R}f_{\ast}S_{X}(\mathcal{N})\simeq\mathcal{D}(\textbf{R}f_{\ast}Q_{X}(\mathcal{N}))\simeq\bigoplus_{i}\mathcal{D}(R^{i}f_{\ast}Q_{X}(\mathcal{N})[-i]).

When we take the jthj^{th}-cohomology on each side, we have the isomorphism

SX(𝒩Xj)ixtXi+j(RifQX(𝒩),ωX).S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}})\cong\bigoplus_{i}{\mathcal{E}}xt^{i+j}_{X^{\prime}}(R^{i}f_{\ast}Q_{X}(\mathcal{N}),\omega^{\bullet}_{X^{\prime}}).

Since SX(𝒩Xj)S_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}}) is torsion free, we may use the same argument from (1)(3)(1)\Rightarrow(3) to prove

xtXi+j(RifQX(𝒩),ωX)=0unless i+j=m.{\mathcal{E}}xt^{i+j}_{X^{\prime}}(R^{i}f_{\ast}Q_{X}(\mathcal{N}),\omega^{\bullet}_{X^{\prime}})=0\quad\text{unless $i+j=-m$}.

Which implies RifQX(𝒩)R^{i}f_{\ast}Q_{X}(\mathcal{N}) is Cohen-Macaulay of pure dimension mm, and therefore torsion-free.

Remark 1.21.

In the proof of (3)(2)(3)\Rightarrow(2) it was shown there is a quasi-isomorphism

RfQX(𝒩)0jdQX(𝒩Xj)[j].\textbf{R}f_{\ast}Q_{X}(\mathcal{N})\simeq\displaystyle\bigoplus_{0\leq j\leq d}Q_{X^{\prime}}(\mathcal{N}^{j}_{X^{\prime}})[j].

But, in the statement of Theorem 1.14, we have

RfQX(𝒩)q(𝒩Xj)=q(𝒩)QX(𝒩Xj)[j].\textbf{R}f_{\ast}Q_{X}(\mathcal{N})\simeq\displaystyle\bigoplus_{q^{\prime}(\mathcal{N}^{j}_{X^{\prime}})=q^{\prime}(\mathcal{N})}Q_{X^{\prime}}(\mathcal{N}_{X^{\prime}}^{j})[-j].

It can be shown that q(𝒩Xj)=q(𝒩)q^{\prime}(\mathcal{N}^{j}_{X^{\prime}})=q^{\prime}(\mathcal{N}) whenever dj0-d\leq j\leq 0. Therefore,

RfQX(𝒩)0jdQX(𝒩Xj)[j].\textbf{R}f_{\ast}Q_{X}(\mathcal{N})\simeq\displaystyle\bigoplus_{0\leq j\leq d}Q_{X^{\prime}}(\mathcal{N}_{X^{\prime}}^{-j})[j].

So, these two statements appear to be contradictory to each other. However, f:XXf:X\rightarrow X^{\prime} is a projective morphism and there is an isomorpism

j:Hjf+𝒩Hjf+𝒩(j),\ell^{j}:H^{-j}f_{+}\mathcal{N}\rightarrow H^{j}f_{+}\mathcal{N}(j),

where \ell is the first Chern class of an ff-ample line bundle [Sai88]. Hence there is an isomorphism 𝒩Xj𝒩Xj(j)\mathcal{N}^{-j}_{X^{\prime}}\cong\mathcal{N}^{j}_{X^{\prime}}(j). Which then induces a quasi-isomorphism

QX(𝒩Xj)QX(𝒩Xj).Q_{X^{\prime}}(\mathcal{N}_{X^{\prime}}^{j})\simeq Q_{X^{\prime}}(\mathcal{N}_{X^{\prime}}^{-j}).

So the two statements above agree.

Corollary 1.22.

Let f:XXf:X\rightarrow X^{\prime} be a surjective projective map between irreducible algebraic varieties. Suppose that XX is smooth of dimension nn and ωX=fL\omega_{X}=f^{*}L, where LL is a line bundle on XX^{\prime}. If 𝒩Xj\mathcal{N}^{j}_{X^{\prime}} is the largest direct factor of Hjf+XHH^{j}f_{+}{\mathbb{Q}}_{X}^{H} with strict support XX^{\prime}, then 𝒩Xj\mathcal{N}^{j}_{X^{\prime}} has RS for all jj.

Proof.

By Kollár’s theorem [Kol86, Theorem 2.1] and the projection formula, RifQ(XH[n])=Rif𝒪X[n]=Ri+nfωXL1R^{i}f_{\ast}Q({\mathbb{Q}}_{X}^{H}[n])=R^{i}f_{*}{\mathcal{O}}_{X}[n]=R^{i+n}f_{*}\omega_{X}\otimes L^{-1} is torsion free. So the corollary follows from the previous theorem. ∎

2. Invariant theory

In this section, we summarize some facts from geometric invariant theory needed below. The basic reference is [MFK94]. Let GG be a complex linear algebraic group acting algebraically on a complex algebraic variety XX in the sense that the action is defined by a morphism α:G×XX\alpha:G\times X\to X of varieties. We will say that XX is a GG-variety. A morphism of GG-varieties is simply an equivariant morphism. Given a GG-variety, the categorical quotient is a variety YY with a morphism π:XY\pi:X\to Y such that

G×X\textstyle{G\times X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}p2\scriptstyle{p_{2}}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}Y\textstyle{Y}

is coequalizer, where p2p_{2} is the second projection. If the categorical quotient exists, then it is characterized up to isomorphism by this property and we denote it by X//GX//G. The symbol X/GX/G is reserved for the set theoretic quotient, or orbit space, which might be different. The action is effective if the map GAut(X)G\to Aut(X) is injective. Effectivity can always be achieved by replacing GG by its image G¯\bar{G}. The quotient X//G=X//G¯X//G=X//\bar{G} is unaffected.

The categorical quotient exists when X=SpecRX=\operatorname{Spec}R is an affine variety and GG is reductive. In this case

X//G=SpecRGX//G=\operatorname{Spec}R^{G}

Suppose XX is a quasiprojective GG-variety. Then it possible to construct a categorical quotient of a large open set which depends on an equivariant locally closed embedding XNX\subset{\mathbb{P}}^{N}. The embedding is determined by the line bundle L=𝒪X(1)L={\mathcal{O}}_{X}(1), and GG will act on the space of sections of LL and its powers. A point x¯X\bar{x}\in X is semistable (with respect to LL) if the closure of the orbit of a point x𝔸N+1{0}x\in{\mathbb{A}}^{N+1}-\{0\}, lying over x¯\bar{x}, does not contain 0. The set of semistable points is denoted by X(L)ssX(L)^{ss}, or XssX^{ss} if LL is understood or unimportant. This forms a nonempty GG-invariant open set. Mumford [MFK94] shows that

Y=im[π:X(L)ssProj(i=0H0(X¯,Li))G]Y=\operatorname{im}\left[\pi:X(L)^{ss}\to\operatorname{Proj}\left(\bigoplus_{i=0}^{\infty}H^{0}(\overline{X},L^{\otimes i})\right)^{G}\right]

is the quotient X(L)ss//GX(L)^{ss}//G, where X¯N\overline{X}\subset{\mathbb{P}}^{N} is the closure and π\pi is induced by the inclusion of projective coordinate rings. The quotient π:XssY\pi:X^{ss}\to Y is a so called good quotient, which means that π\pi is affine, disjoint GG-invariant closed sets map to disjoint closed sets, and 𝒪Y=(π𝒪Xss)G{\mathcal{O}}_{Y}=(\pi_{*}{\mathcal{O}}_{X^{ss}})^{G}. We note that the affine case is subsumed by the second construction because:

Lemma 2.1.

When XX is affine, we can choose LL so that X(L)ss=XX(L)^{ss}=X.

Proof.

We can choose an equvariant embedding XVX\subset V into a rational representation. Let {\mathbb{C}} denote the trivial representation. Then the closure of XX in W=VW=V\oplus{\mathbb{C}} does not contain 0. Therefore the lemma holds for LL corresponding to the embedding X(W)X\subset{\mathbb{P}}(W^{*}). ∎

We say that YY is a geometric quotient of XX if it is a good quotient, and every point of XX has finite stabilizers and all orbits are closed. The last condition implies that Y=X/GY=X/G is the orbit space as a set. A point xXx\in X is stable if it has a finite stabilizer and a closed orbit. The set XsXssX^{s}\subset X^{ss} of stable points is an open, possibly empty, GG-invariant subset. The quotient Ys=Xs/GY^{s}=X^{s}/G is a geometric quotient. Luna’s theorem [Lun73, p 97] (see also [MFK94, appendix 1D]) gives a good local description of geometric quotients.

Theorem 2.2.

[Lun73] If GG is a reductive group and π:XX/G=Y\pi:X\to X/G=Y is a geometric quotient, then for each xXx\in X, there exists a finite group HGH\subset G, an HH-subvariety SXS\subset X, and a commutative diagram

G×S\textstyle{G\times S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(G×S)/H\textstyle{(G\times S)/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S/H\textstyle{S/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}Y\textstyle{Y}

such that maps labeled pp are étale and they map onto Zarski open neighbourhoods of xx and π(x)\pi(x). If XX is smooth, then SS can be chosen smooth.

3. Equivariant Hodge modules

Suppose that XX is a smooth GG-variety with quotient π:XY=X//G.\pi:X\to Y=X//G. Let {\mathcal{F}}^{\prime} be a sheaf on X//GX//G, and let =π{\mathcal{F}}=\pi^{*}{\mathcal{F}}^{\prime}. Then since πp2=πα\pi\circ p_{2}=\pi\circ\alpha, it follows that p2αp_{2}^{*}{\mathcal{F}}\cong\alpha^{*}{\mathcal{F}}. Furthermore, πid\pi^{*}id is an isomorphism ϕ:p2α\phi:p_{2}^{*}{\mathcal{F}}\cong\alpha^{*}{\mathcal{F}} which satisfies the cocycle condition:

sϕ=idandbϕqϕ=mϕ,s^{*}\phi=id\quad\text{and}\quad b^{*}\phi\circ q^{*}\phi=m^{*}\phi,

where the maps s,b,q,s,b,q, and mm are given below.

{s:XG×Xs(x)=(e,x)m:G×G×XG×Xm(g,h,x)=(gh,x)b:G×G×XG×Xb(g,h,x)=(g,α(h,x))q:G×G×XG×Xq(g,h,x)=(h,x)\begin{cases}s:X\rightarrow G\times X&s(x)=(e,x)\\ m:G\times G\times X\rightarrow G\times X&m(g,h,x)=(gh,x)\\ b:G\times G\times X\rightarrow G\times X&b(g,h,x)=(g,\alpha(h,x))\\ q:G\times G\times X\rightarrow G\times X&q(g,h,x)=(h,x)\\ \end{cases}

See also [HTT08, Defn 9.10.3] or [Ach]. A GG-equivariant sheaf is a pair (,ϕ)({\mathcal{F}},\phi), where ϕ\phi is an isomorphism as above satisfying the cocycle condition. Equivariance implies that gg^{*}{\mathcal{F}}\cong{\mathcal{F}} for all gGg\in G, but it is stronger than this. It implies that GG will act on the cohomology groups of {\mathcal{F}}. This fact was already used implicitly in the previous section. A morphism of equivariant sheaves is required to be compatible with the ϕ\phi’s. Let CohG(X)Coh_{G}(X) be the category of coherent equivariant sheaves.

Proposition 3.1.

When GG acts freely on XX in the sense that XX/GX\to X/G is a locally trivial principal bundle in the étale topology, π\pi^{*} gives an equivalence of categories Coh(X/G)CohG(X)Coh(X/G)\approx Coh_{G}(X). The inverse sends (π)G{\mathcal{F}}\mapsto(\pi_{*}{\mathcal{F}})^{G}.

We omit the details, but this is a standard consequence of faithfully flat descent.

Definition 3.2.

In general, for possibly nonfree actions, the functor πG:CohG(X)Coh(X//G)\pi_{*}^{G}:Coh_{G}(X)\to Coh(X//G) defined by πG:=(π)G{\mathcal{F}}\mapsto\pi_{*}^{G}{\mathcal{F}}:=(\pi_{*}{\mathcal{F}})^{G} is right adjoint to π\pi^{*}.

Lemma 3.3.

The functors π\pi_{*} and πG\pi^{G}_{*} are exact. When GG is reductive and CohG(X){\mathcal{F}}\in Coh_{G}(X), πG\pi^{G}_{*}{\mathcal{F}} is a direct summand of π\pi_{*}{\mathcal{F}}.

Proof.

The first statement follows from the fact that the map π\pi is affine. For any open set UYU\subset Y, reductivity of GG implies that (π1U)G(π1U){\mathcal{F}}(\pi^{-1}U)^{G}\subset{\mathcal{F}}(\pi^{-1}U) is a direct summand. ∎

If X=SpecRX=\operatorname{Spec}R, and VV is a rational representation of GG, then M=VRM=V\otimes_{\mathbb{C}}R, with diagonal action, is naturally an object of CohG(X)Coh_{G}(X). In fact, the set of objects of this form generates this category. We have that πGM~=MG~\pi_{*}^{G}\tilde{M}=\widetilde{M^{G}}.

Definition 3.4.

[Ach, Defn. 5.1] A GG-equivariant Hodge module on XX is a pair (,β),({\mathcal{M}},\beta), where MHM(X){\mathcal{M}}\in MHM(X), and β:p2α\beta:p_{2}^{*}{\mathcal{M}}\rightarrow\alpha^{*}{\mathcal{M}} is an isomorphism satisfying the cocycle condition.

Further information about equivariant (mixed) Hodge modules can be found in Achar’s notes [Ach]. Let HM?,G(X,??)HM_{?,G}(X,??) denote the category of GG-equivariant Hodge modules (with strict support ? and weight ??). This category can be described as follows:

Proposition 3.5.

If ZXZ\subseteq X is a GG-invariant closed subvariety, a variation of Hodge structure of weight ww on a smooth GG-invariant Zariski open set UXU\subseteq X with an isomorphism ϕ:p2VαV\phi:p_{2}^{*}V\cong\alpha^{*}V satisfying the cocycle identity extends to GG-equvariant Hodge module of weight w+dimZw+\dim Z. In fact, there is an equivalence of categories

HMZ,G(X,w)2-limUZVHSG(U,wdimZ)HM_{Z,G}(X,w)\approx 2\text{-}\varinjlim_{U\subseteq Z}VHS_{G}(U,w-\dim Z)
Proof.

This is an immediate consequence of [Sai90, Theorem 3.21]. ∎

Corollary 3.6.

If j:UXj:U\subset X is a nonempty GG-invariant open set, restriction jj^{*} gives an equivalence

HMX,G(X,w)HMU,G(U,w)HM_{X,G}(X,w)\approx HM_{U,G}(U,w)

with inverse j!j_{!*}.

Let d=dimGd=\dim G, and set πt𝒩:=π𝒩[d]\pi^{t}\mathcal{N}:=\pi^{*}\mathcal{N}[d]. Note that πt\pi^{t} is compatible with the usual pull back of variations of Hodge structure, i.e. when 𝒩\mathcal{N} corresponds to VVHS(U)V\in VHS(U), with UXU\subset X open, πt𝒩\pi^{t}\mathcal{N} corresponds to πV\pi^{*}V. If 𝒩HM(X//G)\mathcal{N}\in HM(X//G), then (πt𝒩,πtid)(\pi^{t}\mathcal{N},\pi^{t}id) (which we simply denote by πt𝒩\pi^{t}\mathcal{N}) gives an object of HMG(X)HM_{G}(X). Proposition 3.1 works for Hodge modules.

Proposition 3.7.

When GG acts freely on XX, πt\pi^{t} gives an equivalence of categories between HM(X/G)HM(X/G) and HMG(X)HM_{G}(X). More generally, if GG has a normal subgroup HH that acts freely on XX, then πt\pi^{t} gives an equivalence HMG/H(X/H)HMG(X)HM_{G/H}(X/H)\approx HM_{G}(X).

Proof.

See [Ach, §6]. Note that πt\pi^{t} is denoted by π\pi^{\dagger} in [loc. cit.]. ∎

More generally, we will prove the following proposition.

Proposition 3.8.

Suppose that GG is a reductive group that acts effectively on XX such that XsX^{s}\not=\emptyset. Let Y=X//GY=X//G. Then there is a functor

π:HMY(Y,w)HMX,G(X,w+d),\pi^{\dagger}:HM_{Y}(Y,w)\rightarrow HM_{X,G}(X,w+d),

which gives an equivalence of categories, and which agrees with πt\pi^{t} when GG acts freely.

To prove the proposition, we first need to prove two lemmas.

Lemma 3.9.

If f:STf:S\rightarrow T is a surjective finite map between irreducible algebraic varieties, then there is an equivalence of categories between HMS(S,w)HM_{S}(S,w) and HMT(T,w).HM_{T}(T,w). The equivalence are given by the functors

f#:HMT(T,w)HMS(S,w)f^{\#}:HM_{T}(T,w)\rightarrow HM_{S}(S,w)
𝒩=maximal summand of GrwW(0(f𝒩)) with strict support on S\mathcal{N}\mapsto{\mathcal{M}}=\text{maximal summand of $Gr^{W}_{w}({\mathcal{H}}^{0}(f^{*}\mathcal{N}))$ with strict support on $S$}

and

f#:HMS(S,w)HMT(T,w)f_{\#}:HM_{S}(S,w)\rightarrow HM_{T}(T,w)
𝒩=maximal summand of 0(f+) with strict support on T.{\mathcal{M}}^{\prime}\mapsto\mathcal{N}^{\prime}=\text{maximal summand of ${\mathcal{H}}^{0}(f_{+}{\mathcal{M}}^{\prime})$ with strict support on $T$.}
Proof.

First consider 𝒩HMT(T,w).\mathcal{N}\in HM_{T}(T,w). Since the fibers of ff are zero dimensional, i(f𝒩)=0{\mathcal{H}}^{i}(f^{*}\mathcal{N})=0 for i>0i>0 [HTT08, Prop. 8.1.40]. By Saito’s theory of weights [Sai90, Prop. 2.26], Grw+1W(0(f𝒩))=0Gr^{W}_{w+1}({\mathcal{H}}^{0}(f^{*}\mathcal{N}))=0. We have a functor

f#:HMT(T,w)HMS(S,w)f^{\#}:HM_{T}(T,w)\rightarrow HM_{S}(S,w)
𝒩=maximal summand of GrwW(0(f𝒩)) with strict support on S.\mathcal{N}\mapsto{\mathcal{M}}=\text{maximal summand of $Gr^{W}_{w}({\mathcal{H}}^{0}(f^{*}\mathcal{N}))$ with strict support on $S$.}

Now suppose that HMS(S,w).{\mathcal{M}}^{\prime}\in HM_{S}(S,w). Since the map ff is finite, f+0(f+).f_{+}{\mathcal{M}}^{\prime}\simeq{\mathcal{H}}^{0}(f_{+}{\mathcal{M}}^{\prime}). There is a functor

f#:HMS(S,w)HMT(T,w)f_{\#}:HM_{S}(S,w)\rightarrow HM_{T}(T,w)
𝒩=maximal summand of 0(f+) with strict support on T.{\mathcal{M}}^{\prime}\mapsto\mathcal{N}^{\prime}=\text{maximal summand of ${\mathcal{H}}^{0}(f_{+}{\mathcal{M}}^{\prime})$ with strict support on $T$.}

We will show that f#f#f_{\#}\circ f^{\#} is isomorphic to the identity functor on HMT(T,w)HM_{T}(T,w) and f#f#f^{\#}\circ f_{\#} is isomorphic to the identity functor on HMS(S,w).HM_{S}(S,w).

By construction, there is a natural map 𝒩(f#f#)(N)\mathcal{N}\rightarrow(f_{\#}\circ f^{\#})(N). Using [Sai90, Thm. 3.21] and possibly restricting to a smooth Zariski-open subset of T,T, we may assume TT is smooth, 𝒩\mathcal{N} is a variation of Hodge structure, and the map ff is étale of degree .\ell. Let {\mathcal{L}} denote the underlying locally constant sheaf of 𝒩\mathcal{N}. The adjunction map

adj:Rff1R0ff1adj_{*}:{\mathcal{L}}\rightarrow\textbf{R}f_{*}f^{-1}{\mathcal{L}}\simeq R^{0}f_{*}f^{-1}{\mathcal{L}}

realizes {\mathcal{L}} as a direct summand of R0ff1R^{0}f_{*}f^{-1}{\mathcal{L}} because the composition 1adj!adj\displaystyle\frac{1}{\ell}\cdot adj_{!}\circ adj_{*} is the identity map on ,{\mathcal{L}}, where adj!adj_{!} is given by the map

R0ff1Rf!f!.R^{0}f_{*}f^{-1}{\mathcal{L}}\simeq\textbf{R}f_{!}f^{!}{\mathcal{L}}\rightarrow{\mathcal{L}}.

Therefore, the natural map 𝒩f+f𝒩\mathcal{N}\rightarrow f_{+}f^{*}\mathcal{N} splits and 𝒩\mathcal{N} is a direct summand of f+f𝒩.f_{+}f^{*}\mathcal{N}. Hence the natural map 𝒩(f#f#)(𝒩)\mathcal{N}\rightarrow(f_{\#}\circ f^{\#})(\mathcal{N}) is an isomorphism.

For the other direction, there is a natural map (f#f#)().(f^{\#}\circ f_{\#})({\mathcal{M}}^{\prime})\rightarrow{\mathcal{M}}^{\prime}. Again using [Sai90, Thm. 3.21], after possibly restricting to smooth Zariski-open subsets of SS and T,T, we may assume {\mathcal{M}}^{\prime} and (f#f#)()(f^{\#}\circ f_{\#})({\mathcal{M}}^{\prime}) are variations of Hodge structure, and the map f:STf:S\rightarrow T is étale. But then by construction, the map (f#f#)()(f^{\#}\circ f_{\#})({\mathcal{M}}^{\prime})\rightarrow{\mathcal{M}}^{\prime} is a local isomorphism, and hence the map is an isomorphism.

Lemma 3.10.

Suppose f:STf:S\rightarrow T is a surjective, finite map between smooth, irreducible GG-varieties, and the map ff is equivariant. Then there is an equivalence of categories f#:HMS,G(S,w)HMT,G(T,w).f^{\#}:HM_{S,G}(S,w)\approx HM_{T,G}(T,w).

Proof.

By the previous lemma, we already know there is an equivalence of categories between HMS(S,w)HM_{S}(S,w) and HMT(T,w)HM_{T}(T,w). We have to show this equivalence of categories is compatible with the data defining equivariant Hodge modules. Using the notation of the previous proof, we will show that the compatibility for the functor f#:HMT(T,w)HMS(S,w)f^{\#}:HM_{T}(T,w)\rightarrow HM_{S}(S,w). The argument for the inverse functor f#f_{\#} is similar.

Since the map ff is equivariant, we have the following commutative diagrams

G×S{G\times S}S{S}G×T{G\times T}T{T}αS\scriptstyle{\alpha_{S}}(id;f)\scriptstyle{(id;f)}f\scriptstyle{f}αT\scriptstyle{\alpha_{T}}  G×S{G\times S}S{S}G×T{G\times T}T,{T,}pS\scriptstyle{p_{S}}(id;f)\scriptstyle{(id;f)}f\scriptstyle{f}pT\scriptstyle{p_{T}}

where pp is the projection map and α\alpha is the action map. Note that the map (id;f):G×SG×T(id;f):G\times S\rightarrow G\times T is finite, and hence there is an equivalence of categories between HMG×S(G×S)HM_{G\times S}(G\times S) and HMG×T(G×T)HM_{G\times T}(G\times T). If 𝒩HMT,G(T,w),\mathcal{N}\in HM_{T,G}(T,w), then there is an isomorphism βT:pT𝒩αT𝒩\beta_{T}:p_{T}^{*}\mathcal{N}\rightarrow\alpha^{*}_{T}\mathcal{N} satisfying the cocycle condition. Therefore, there is a quasi-isomorphism

pSf𝒩(id;f)pT𝒩(id;f)αT𝒩αSf𝒩.p_{S}^{*}f^{*}\mathcal{N}\simeq(id;f)^{*}p_{T}^{*}\mathcal{N}\rightarrow(id;f)^{*}\alpha_{T}^{*}\mathcal{N}\simeq\alpha_{S}^{*}f^{*}\mathcal{N}.

The projection and action maps are smooth. Hence there are quasi-isomorphisms

pS𝒩d(pS𝒩)[d]andαS𝒩d(αS𝒩)[d],p^{*}_{S}\mathcal{N}\simeq{\mathcal{H}}^{d}(p^{*}_{S}\mathcal{N})[-d]\quad\text{and}\quad\alpha^{*}_{S}\mathcal{N}\simeq{\mathcal{H}}^{d}(\alpha^{*}_{S}\mathcal{N})[-d],

where d=dimGd=\dim G [Sai90, Lemma 2.25]. Therefore, by the use of spectral sequences, we have an isomorphism

d(pS(0(f𝒩)))d(pSf𝒩)d(αSf𝒩)d(αS(0(f𝒩))).{\mathcal{H}}^{d}(p_{S}^{*}({\mathcal{H}}^{0}(f^{*}\mathcal{N})))\cong{\mathcal{H}}^{d}(p_{S}^{*}f^{*}\mathcal{N})\rightarrow{\mathcal{H}}^{d}(\alpha_{S}^{*}f^{*}\mathcal{N})\cong{\mathcal{H}}^{d}(\alpha_{S}^{*}({\mathcal{H}}^{0}(f^{*}\mathcal{N}))).

Also, by [Sai90, Lemma 2.25],

Grw+dWd(pS(0(f𝒩)))d(pS(GrwW0(f𝒩)))Gr^{W}_{w+d}{\mathcal{H}}^{d}(p_{S}^{*}({\mathcal{H}}^{0}(f^{*}\mathcal{N})))\cong{\mathcal{H}}^{d}(p_{S}^{*}(Gr^{W}_{w}{\mathcal{H}}^{0}(f^{*}\mathcal{N})))
Grw+dWd(αS(0(f𝒩)))d(αS(GrwW0(f𝒩))).Gr^{W}_{w+d}{\mathcal{H}}^{d}(\alpha_{S}^{*}({\mathcal{H}}^{0}(f^{*}\mathcal{N})))\cong{\mathcal{H}}^{d}(\alpha_{S}^{*}(Gr^{W}_{w}{\mathcal{H}}^{0}(f^{*}\mathcal{N}))).

Hence there is an isomorphism βS:pSαS.\beta_{S}:p_{S}^{*}{\mathcal{M}}\rightarrow\alpha_{S}^{*}{\mathcal{M}}. The cocycle condition can be shown in a similar matter. Therefore, {\mathcal{M}} is an equivariant Hodge module. ∎

Proof of Proposition 3.8.

Since Xs,X^{s}\neq\emptyset, by Theorem 2.2, we have a diagram

G×S\textstyle{G\times S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}r\scriptstyle{r}p\scriptstyle{p}(G×S)/H\textstyle{(G\times S)/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}Xs\textstyle{X^{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}S/H\textstyle{S/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h}Ys\textstyle{Y^{s}}

satisfying the conditions stated in that theorem. There is also an equivalence of categories

HMXs,G(Xs,w+d)HM(G×S)/H,G((G×S)/H,w+d)HM_{X^{s},G}(X^{s},w+d)\approx HM_{(G\times S)/H,G}((G\times S)/H,w+d)

by Lemma 3.10 because the map (G×S)/HXs(G\times S)/H\rightarrow X^{s} is GG-equivariant. Note that the map G×S(G×S)/HG\times S\rightarrow(G\times S)/H is GG-equivariant because the GG-action on (G×S)/H(G\times S)/H is the natural action

g[(g,s)]=[(gg,s)]for g,gG and sS.g\cdot[(g^{\prime},s)]=[(gg^{\prime},s)]\quad\text{for $g,g^{\prime}\in G$ and $s\in S.$}

So, by applying Lemma 3.10, we have an equivalence of categories

HMG×S,G(G×S,w+d)HM(G×S)/H,G((G×S)/H,w+d).HM_{G\times S,G}(G\times S,w+d)\approx HM_{(G\times S)/H,G}((G\times S)/H,w+d).

Using a similar argument with Lemma 3.9, there is an equivalence of categories

HMY(Y,w)HMS(S,w).HM_{Y}(Y,w)\approx HM_{S}(S,w).

With Proposition 3.7 we obtain the diagram

HMG×S,G(G×S,w+d)\textstyle{HM_{G\times S,G}(G\times S,w+d)}HMXs,G(Xs,w+d)\textstyle{HM_{X^{s},G}(X^{s},w+d)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\approx}HMS(S,w)\textstyle{HM_{S}(S,w)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\approx}HMS/H(S/H,w)\textstyle{HM_{S/H}(S/H,w)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\approx}HMYs(Ys,w).\textstyle{HM_{Y^{s}}(Y^{s},w).\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\approx}

By Corollary 3.6, there are equivalences of categories

j:HMX,G(X,w+d)HMXs,G(Xs,w+d),j^{*}:HM_{X,G}(X,w+d)\approx HM_{X^{s},G}(X^{s},w+d),
k:HMY(Y,w)HMYs(Ys,w),k^{*}:HM_{Y}(Y,w)\approx HM_{Y^{s}}(Y^{s},w),

where j:XsXj:X^{s}\to X and k:YsYk:Y^{s}\to Y are the inclusions. Then

(6) π=j!f#r#ptq#h#k\pi^{\dagger}=j_{!*}f_{\#}r_{\#}p^{t}q^{\#}h^{\#}k^{*}

gives the desired equivalence. One can check that π=πt\pi^{\dagger}=\pi^{t} when GG acts freely on XX and π=π#\pi=\pi^{\#} when GG is finite. ∎

Remark 3.11.

The formula (6) is not easy to use directly. However, it is not hard to see that the functor π:HMY(Y,w)HMX,G(X,w+d)\pi^{\dagger}:HM_{Y}(Y,w)\rightarrow HM_{X,G}(X,w+d) is given by:

𝒩= the maximal summand of Grw+dWd(π𝒩) with strict support on X.\mathcal{N}\mapsto{\mathcal{M}}=\text{ the maximal summand of $Gr^{W}_{w+d}{\mathcal{H}}^{d}(\pi^{*}\mathcal{N})$ with strict support on $X$.}

Here is a basic example.

Lemma 3.12.

With the same assumptions as in Proposition 3.8, πICYH=ICXH\pi^{\dagger}IC_{Y}^{H}=IC_{X}^{H}.

Proof.

Choosing smooth Zariski open sets UXU\subset X and VYV\subset Y such that π(U)V\pi(U)\subset V, the pullback of the constant variation of Hodge structure πVH=UH\pi^{*}{\mathbb{Q}}_{V}^{H}={\mathbb{Q}}_{U}^{H}. Since ICYHIC_{Y}^{H} and ICXHIC_{X}^{H} are the unique extensions of VH[dimY]{\mathbb{Q}}_{V}^{H}[\dim Y] and UH[dimX]{\mathbb{Q}}_{U}^{H}[\dim X] to Hodge modules with strict support on YY and XX, we must have πICYH=ICXH\pi^{\dagger}IC_{Y}^{H}=IC_{X}^{H}. ∎

We will use π+G:HMX,G(X,w+d)HMY(Y,w)\pi^{G}_{+}:HM_{X,G}(X,w+d)\rightarrow HM_{Y}(Y,w) to denote the inverse functor of π.\pi^{\dagger}. The next proposition gives an explicit description of π+G.\pi^{G}_{+}.

Proposition 3.13.

If HMX,G(X,w+d),{\mathcal{M}}\in HM_{X,G}(X,w+d), then

π+G()= the maximal summand of GrwWd(π+) with strict support on Y.\pi^{G}_{+}({\mathcal{M}})=\text{ the maximal summand of $Gr^{W}_{w}{\mathcal{H}}^{-d}(\pi_{+}{\mathcal{M}})$ with strict support on $Y$.}
Proof.

If GG is finite, then the proposition follows from Lemma 3.9. So, we may assume dimG1\dim G\geq 1 and GG is connected. After possibly restricting to open subsets of XX and YY, we may assume π:XY\pi:X\rightarrow Y is a geometric quotient. By Luna’s theorem 2.2, we have the diagram

G×S\textstyle{G\times S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}r\scriptstyle{r}(G×S)/H\textstyle{(G\times S)/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Xs\textstyle{X^{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}S/H\textstyle{S/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ys\textstyle{Y^{s}}

satisfying the assumptions stated in that theorem. After possibly restricting further to open subsets of XX and Y,Y, we may assume the maps

S/HYand(G×S)/HXS/H\rightarrow Y\quad\text{and}\quad(G\times S)/H\rightarrow X

are surjective. Hence we may asssume the maps G×SXG\times S\rightarrow X and SYS\rightarrow Y are surjective and finite. Therefore, by Lemma 3.9 and Lemma 3.10, we may reduce to the case when X=G×YX=G\times Y and π:XY\pi:X\rightarrow Y is the natural projection. Let s:YXs:Y\rightarrow X be the natural map given by

y(e,y) eG is the identity element.y\mapsto(e,y)\quad\text{ $e\in G$ is the identity element.}

We have πs=idY\pi\circ s=id_{Y}, and s:HMX,G(X,w+d)HMY(Y,w)s^{\dagger}:HM_{X,G}(X,w+d)\rightarrow HM_{Y}(Y,w) is the inverse functor for the equivalence of categories. For 𝒩HMY(Y,w)\mathcal{N}\in HM_{Y}(Y,w) it was shown in the proof of [Sai90, Lemma 2.27] that the map

d(π+(dπ𝒩))0sπ𝒩=𝒩{\mathcal{H}}^{-d}(\pi_{+}({\mathcal{H}}^{d}\pi^{*}\mathcal{N}))\rightarrow{\mathcal{H}}^{0}s^{\dagger}\pi^{\dagger}\mathcal{N}=\mathcal{N}

is an isomorphism. ∎

Corollary 3.14.

If HMX,G(X,w+d),{\mathcal{M}}\in HM_{X,G}(X,w+d), then

π+G()=the maximal summand of d(π¯+¯) with strict support on Y,\pi^{G}_{+}({\mathcal{M}})=\text{the maximal summand of ${\mathcal{H}}^{-d}(\bar{\pi}_{+}\bar{{\mathcal{M}}})$ with strict support on $Y$,}

where π¯:X¯Y\bar{\pi}:\bar{X}\rightarrow Y is any compactification of π,\pi, and ¯HMX¯(X¯,w+d)\bar{{\mathcal{M}}}\in HM_{\bar{X}}(\bar{X},w+d) is the unique extension of .{\mathcal{M}}.

Proof.

Compactify the morphism π\pi to a projective morphism π¯:X¯Y\bar{\pi}:\bar{X}\rightarrow Y such that the diagram commutes

X{X}X¯{\bar{X}}Y.{Y.}j\scriptstyle{j}π\scriptstyle{\pi}π¯\scriptstyle{\bar{\pi}}

There is a natural map

d(π¯+¯)GrwWd(π+).{\mathcal{H}}^{-d}(\bar{\pi}_{+}\bar{{\mathcal{M}}})\rightarrow Gr^{W}_{w}{\mathcal{H}}^{-d}(\pi_{+}{\mathcal{M}}).

Furthermore, this map is surjective by the weight spectral sequence [Sai90, Prop. 2.15]. Which implies π+G()\pi^{G}_{+}({\mathcal{M}}) is a direct summand of d(π¯+¯){\mathcal{H}}^{-d}(\bar{\pi}_{+}\bar{{\mathcal{M}}}). ∎

Lemma 3.15.

If f:STf:S\rightarrow T is a surjective finite map between irreducible algebraic varieties, then HMS(S,w){\mathcal{M}}\in HM_{S}(S,w) has RS if and only if 𝒩HMT(T,w)\mathcal{N}\in HM_{T}(T,w) has RS under the equivalence of categories.

Proof.

The morphism f:Sf:S\rightarrow is projective since it is finite. By Theorem 1.14, and using ff_{*} is exact,

RfQS()fQS()QT(𝒩).\textbf{R}f_{*}Q_{S}({\mathcal{M}})\simeq f_{*}Q_{S}({\mathcal{M}})\simeq Q_{T}(\mathcal{N}).

So, if {\mathcal{M}} has RS, then it is clear that 𝒩\mathcal{N} must also have RS. Similarly, since f:STf:S\rightarrow T is affine and ff_{*} is exact, if 𝒩\mathcal{N} has RS then {\mathcal{M}} must also have RS.

Lemma 3.16.

Let GG be a reductive group of dimension dd and π:XX/G=Y\pi:X\to X/G=Y a geometric quotient of a smooth variety. If HMX,G(X){\mathcal{M}}\in HM_{X,G}(X), then {\mathcal{M}} has rational singularities if and only if π+G\pi^{G}_{+}{\mathcal{M}} has rational singularities. In particular, if {\mathcal{M}} or π+G\pi^{G}_{+}{\mathcal{M}} has rational singularities, then Q(π+G)[d]πGQ()Q(\pi_{+}^{G}{\mathcal{M}})[d]\cong\pi_{*}^{G}Q({\mathcal{M}}) and S(π+G)πGS()S(\pi_{+}^{G}{\mathcal{M}})\cong\pi_{*}^{G}S({\mathcal{M}}).

Proof.

Recall, by Theorem 2.2, for xXx\in X and π(x)Y\pi(x)\in Y there exists open neighborhoods xUXx\in U\subset X and π(x)VY\pi(x)\in V\subset Y such that we have the commutative diagram

G×S\textstyle{G\times S\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Π\scriptstyle{\Pi}(G×S)/H\textstyle{(G\times S)/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}U\textstyle{U\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π|U\scriptstyle{\pi|_{U}}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}S\textstyle{S\ignorespaces\ignorespaces\ignorespaces\ignorespaces}S/H\textstyle{S/H\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}V\textstyle{V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Y.\textstyle{Y.}

The variety SS is smooth, and the maps pp are étale and surjective. Since the condition of rational singularities is local, we may replace XX with UU and YY with VV. Hence we may assume the induced maps

α:G×SX andβ:SY\alpha:G\times S\rightarrow X\quad\text{ and}\quad\beta:S\rightarrow Y

are finite and surjective. Let 𝒩=π+G.\mathcal{N}=\pi^{G}_{+}{\mathcal{M}}. Recall from Lemma 3.9 there exists an equivalence of categories between HMY(Y,w)HM_{Y}(Y,w) and HMS(S,w).HM_{S}(S,w). Let 𝒩HMS(S)\mathcal{N}^{\prime}\in HM_{S}(S) be the Hodge module that corresponds to 𝒩.\mathcal{N}. Now, because the map SYS\rightarrow Y is finite, 𝒩\mathcal{N} has RS if and only if 𝒩\mathcal{N}^{\prime} has RS. Similarly, let HMG×S,G(G×S){\mathcal{M}}^{\prime}\in HM_{G\times S,G}(G\times S) be the Hodge module that corresponds to {\mathcal{M}} from Lemma 3.10. Note that {\mathcal{M}} has RS if and only {\mathcal{M}}^{\prime} has RS. By construction, we have

=Π𝒩[d].{\mathcal{M}}^{\prime}=\Pi^{*}\mathcal{N}^{\prime}[d].

The explicit formulas in [Sch19, §30] show that

QG×S()=ΠQS(𝒩)[d]Q_{G\times S}({\mathcal{M}}^{\prime})=\Pi^{*}Q_{S}(\mathcal{N}^{\prime})[d]

and

SG×S()=ΠSS(𝒩).S_{G\times S}({\mathcal{M}}^{\prime})=\Pi^{*}S_{S}(\mathcal{N}^{\prime}).

Since Π:G×SS\Pi:G\times S\rightarrow S is faithfully flat, we see that {\mathcal{M}}^{\prime} has RS if and only if 𝒩\mathcal{N}^{\prime} has RS. Hence {\mathcal{M}} has RS if and only if 𝒩\mathcal{N} has RS.

When {\mathcal{M}}^{\prime} or 𝒩\mathcal{N}^{\prime} has RS, we have

(7) QS(𝒩)[d]RΠGQG×S()=ΠGQG×S()SS(𝒩)RΠGSG×S()=ΠGSG×S()\begin{split}Q_{S}(\mathcal{N}^{\prime})[d]&\cong\textbf{R}\Pi^{G}_{*}Q_{G\times S}({\mathcal{M}}^{\prime})=\Pi^{G}_{*}Q_{G\times S}({\mathcal{M}}^{\prime})\\ S_{S}(\mathcal{N}^{\prime})&\cong\textbf{R}\Pi^{G}_{*}S_{G\times S}({\mathcal{M}}^{\prime})=\Pi^{G}_{*}S_{G\times S}({\mathcal{M}}^{\prime})\end{split}

because GG acts freely on SS and we may apply Proposition 3.1. The second set of equalities, which involves a slight abuse of notation, stems from the fact that ΠG\Pi^{G}_{*} is exact on the category of coherent sheaves by Lemma 3.3. Now it is well known that if SS is an affine regular {\mathbb{C}}-domain, then the ring of GG-invariants SGS^{G} is a direct summand of SS as a SGS^{G}-module [HR74]. The sheaves (up to translation) QS(𝒩)Q_{S}(\mathcal{N}^{\prime}), QG×S()Q_{G\times S}({\mathcal{M}}^{\prime}), SS(𝒩)S_{S}(\mathcal{N}^{\prime}) and SG×S()S_{G\times S}({\mathcal{M}}^{\prime}) are locally free by Corollary 1.13, because SS and G×SG\times S are smooth. By (7), QS(𝒩)[d]Q_{S}(\mathcal{N}^{\prime})[d] is a direct summand of ΠQG×S().\Pi_{*}Q_{G\times S}({\mathcal{M}}^{\prime}). Let

ΠQG×S()=ΠGQG×S()A[dimX]=QS(𝒩)[d]A[dimX]\Pi_{*}Q_{G\times S}({\mathcal{M}}^{\prime})=\Pi^{G}_{*}Q_{G\times S}({\mathcal{M}}^{\prime})\oplus A[\dim X]=Q_{S}(\mathcal{N}^{\prime})[d]\oplus A[\dim X]

for some coherent 𝒪S{\mathcal{O}}_{S}-module AA by Lemma 3.3. Then

πQX()παQG×S()βΠQG×S()βΠGQG×S()βA[dimX]\pi_{*}Q_{X}({\mathcal{M}})\cong\pi_{*}\alpha_{*}Q_{G\times S}({\mathcal{M}}^{\prime})\cong\beta_{*}\Pi_{*}Q_{G\times S}({\mathcal{M}}^{\prime})\cong\beta_{*}\Pi^{G}_{*}Q_{G\times S}({\mathcal{M}}^{\prime})\oplus\beta_{*}A[\dim X]
βQS(𝒩)[d]βA[dimX]QY(𝒩)[d]βA[dimX].\cong\beta_{*}Q_{S}(\mathcal{N}^{\prime})[d]\oplus\beta_{*}A[\dim X]\cong Q_{Y}(\mathcal{N})[d]\oplus\beta_{*}A[\dim X].

Therefore, when taking GG-invariants, we obtain

QY(𝒩)[d]πGQX().Q_{Y}(\mathcal{N})[d]\cong\pi^{G}_{*}Q_{X}({\mathcal{M}}).

By a similar argument, S(π+G)πGS()S(\pi_{+}^{G}{\mathcal{M}})\cong\pi_{*}^{G}S({\mathcal{M}}). ∎

Theorem 3.17.

Let GG be a dd-dimensional reductive algebraic group with an effective action on a smooth affine variety XX such that XsX^{s}\not=\emptyset. Let π:XY=X//G\pi:X\to Y=X//G be the quotient map. Suppose that HMX,G(X,w+d){\mathcal{M}}\in HM_{X,G}(X,w+d) has rational singularities, then 𝒩=π+GHMY(Y,w)\mathcal{N}=\pi^{G}_{+}{\mathcal{M}}\in HM_{Y}(Y,w) has rational singularities, and

QY(𝒩)[d](πQX())GQ_{Y}(\mathcal{N})[d]\cong(\pi_{*}Q_{X}({\mathcal{M}}))^{G}
SY(𝒩)(πSX())G.S_{Y}(\mathcal{N})\cong(\pi_{*}S_{X}({\mathcal{M}}))^{G}.
Proof.

By a theorem of Kirwan [Kir85, Prop 6.3], we can find a nonsingular GG-variety X~\tilde{X} with an ample GG-equivariant line bundle L~\tilde{L}, and an equivariant map p:X~Xp:\tilde{X}\to X, such that p(X~(L~)ss)=Xp(\tilde{X}(\tilde{L})^{ss})=X, X~(L~)ss=X~(L~)s\tilde{X}(\tilde{L})^{ss}=\tilde{X}(\tilde{L})^{s}. We also have that the induced map q:Y~=Xss//GYq:\tilde{Y}=X^{ss}//G\to Y is birational. Furthermore, the arguments of [Kir85, §3] show that X~\tilde{X} can be chosen to dominate any given equivariant blow up of XX. Therefore, the pull back of a given invariant closed set can be assumed to be a divisor with simple normal crossings. To simplify notation, let us replace X~\tilde{X} by X~ss\tilde{X}^{ss}. Then we have a commutative diagram

X~\textstyle{\tilde{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π~\scriptstyle{\tilde{\pi}}p\scriptstyle{p}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}Y~\textstyle{\tilde{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}Y.\textstyle{Y.}

The map p:X~Xp:\tilde{X}\rightarrow X is birational. So let ~HMX~,G(X~,w+d)\tilde{\mathcal{M}}\in HM_{\tilde{X},G}(\tilde{X},w+d) be the unique Hodge module that agrees with {\mathcal{M}} on some smooth open dense subset. We can assume that ~\tilde{\mathcal{M}} has singularities along a normal crossing divisor by blowing up X~\tilde{X} further. So, by Lemma 1.11, ~\tilde{\mathcal{M}} has RS. Since the map π~:X~Y~\tilde{\pi}:\tilde{X}\rightarrow\tilde{Y} is a geometric quotient, 𝒩~=π~+G~\tilde{\mathcal{N}}=\tilde{\pi}^{G}_{+}\tilde{\mathcal{M}} has RS by Lemma 3.16. Since the maps π\pi and π~\tilde{\pi} are affine, the functors π\pi_{*} and π~\tilde{\pi}_{*} are exact on the category of coherent sheaves, so we use the same symbols for their derived functors. Since

QY~(𝒩~)[d]π~GQX~(~)Q_{\tilde{Y}}(\tilde{\mathcal{N}})[d]\cong\tilde{\pi}^{G}_{*}Q_{\tilde{X}}(\tilde{{\mathcal{M}}})
SY~(𝒩~)π~GSX~(~)S_{\tilde{Y}}(\tilde{\mathcal{N}})\cong\tilde{\pi}^{G}_{*}S_{\tilde{X}}(\tilde{{\mathcal{M}}})

we deduce that

π~QX~(~)QY~(𝒩~)[d]A[dimX]\tilde{\pi}_{*}Q_{\tilde{X}}(\tilde{{\mathcal{M}}})\cong Q_{\tilde{Y}}(\tilde{\mathcal{N}})[d]\oplus A[\dim X]
π~SX~(~)SY~(𝒩~)B\tilde{\pi}_{*}S_{\tilde{X}}(\tilde{{\mathcal{M}}})\cong S_{\tilde{Y}}(\tilde{\mathcal{N}})\oplus B

for some 𝒪Y~{\mathcal{O}}_{\tilde{Y}}-modules AA and BB by Lemma 3.3. Therefore, we have

πQX()πRpQX~(~)Rqπ~QX~(~)\pi_{*}Q_{X}({\mathcal{M}})\simeq\pi_{*}\textbf{R}p_{*}Q_{\tilde{X}}(\tilde{{\mathcal{M}}})\simeq\textbf{R}q_{*}\tilde{\pi}_{*}Q_{\tilde{X}}(\tilde{{\mathcal{M}}})
RqQY~(𝒩~)[d]RqA[dimX]QY(𝒩)[d]RqA[dimX].\simeq\textbf{R}q_{*}Q_{\tilde{Y}}(\tilde{\mathcal{N}})[d]\oplus\textbf{R}q_{*}A[\dim X]\simeq Q_{Y}(\mathcal{N})[d]\oplus\textbf{R}q_{*}A[\dim X].

Similarly,

πSX()SY(𝒩)RqB.\pi_{*}S_{X}({\mathcal{M}})\simeq S_{Y}(\mathcal{N})\oplus\textbf{R}q_{*}B.

Therefore if {\mathcal{M}} has RS, then we see that 𝒩\mathcal{N} has RS and

QY(𝒩)[d]πGQX()Q_{Y}(\mathcal{N})[d]\cong\pi^{G}_{*}Q_{X}({\mathcal{M}})
SY(𝒩)πGSX().S_{Y}(\mathcal{N})\cong\pi^{G}_{*}S_{X}({\mathcal{M}}).

We recover a special case of Boutot’s theorem (needed below) by a method entirely different from the original proof.

Corollary 3.18.

Let X,GX,G and YY be as in Theorem 3.17. The variety YY has rational singularities.

Proof.

We note first of all that YY is normal because the normalization of YY would also be a categorical quotient of XX by GG. Since XX is smooth, XH[dimX]HMX,G(X,dimX){\mathbb{Q}}^{H}_{X}[\dim X]\in HM_{X,G}(X,\dim X) has rational singularities. By Lemma 3.12, πGXH[dimX]=ICYH\pi_{*}^{G}{\mathbb{Q}}^{H}_{X}[\dim X]=IC^{H}_{Y}, and this has rational singularities by Theorem 3.17. Therefore YY has rational singularities by Corollary 1.19.

The following should be standard, but we indicate the proof for lack of a suitable reference.

Lemma 3.19.

A direct summand of a Cohen-Macaulay module over a noetherian ring is again Cohen-Macaulay.

Proof.

It suffices to prove this when the ring is local. If NMN\subset M is a summand, then we have inequalities

dimMdimNdepthNdepthM.\dim M\geq\dim N\geq\operatorname{depth}N\geq\operatorname{depth}M.

Therefore if MM is CM, this forces NN to be CM as well. We only prove the last inequality, since the remaining inequalities are elementary. If kk denotes the residue field, then by [Mat89, Theorem 16.7]

depthN=inf{iExti(k,N)0}inf{iExti(k,M)0}=depthM\operatorname{depth}N=\inf\{i\mid Ext^{i}(k,N)\not=0\}\geq\inf\{i\mid Ext^{i}(k,M)\not=0\}=\operatorname{depth}M

because Exti(k,N)Ext^{i}(k,N) is a summand of Exti(k,M)Ext^{i}(k,M). ∎

Corollary 3.20.

Let X=SpecRX=\operatorname{Spec}R be a smooth variety admitting an action by a reductive group GG such that XsX^{s}\not=\emptyset. Let VV be a rational representation of GG. Then (VR)G(V\otimes_{\mathbb{C}}R)^{G} is Cohen-Macaulay provided that VRV\otimes R is a direct summand of the global sections of either QX()[dimX]Q_{X}({\mathcal{M}})[-\dim X] or SX()S_{X}({\mathcal{M}}) for some HMX,G(X){\mathcal{M}}\in HM_{X,G}(X) with RS.

Proof.

Set Y=SpecRG=X//GY=\operatorname{Spec}R^{G}=X//G. By the previous corollary YY has rational singularities, therefore it is CM [Kem+73, chap I §3]. Thus ωY=ωY[dimY]\omega_{Y}^{\bullet}=\omega_{Y}[\dim Y], where ωY\omega_{Y} is the dualizing sheaf. Let HMX,G(X){\mathcal{M}}\in HM_{X,G}(X) have RS and let 𝒩=π+G\mathcal{N}=\pi^{G}_{+}{\mathcal{M}}. Theorem 3.17 implies that 𝒩\mathcal{N} has RS. Therefore SY(𝒩)S_{Y}(\mathcal{N}) is maximal Cohen-Macaulay by Lemma 1.12. So by Theorem 3.17 and [HK71, Satz 6.1]

πG(QX()[dimX])=omY(SY(𝒩),ωY)\pi_{*}^{G}(Q_{X}({\mathcal{M}})[-\dim X])=\mathcal{H}om_{Y}(S_{Y}(\mathcal{N}),\omega_{Y})

is CM. A direct summand of QX()[dimX]Q_{X}({\mathcal{M}})[-\dim X] or SX()S_{X}({\mathcal{M}}) is also CM by Lemma 3.19. ∎

Following the referee’s request, we give a simple example where this corollary applies.

Example 3.21.

Let GGLn()G\subseteq GL_{n}({\mathbb{C}}) be a reductive subgroup acting in the standard way on R=[x1,,xn]R={\mathbb{C}}[x_{1},\ldots,x_{n}]. We suppose that XsX^{s}\not=\emptyset. In the special case, where G=G={\mathbb{C}}^{*}, with tGt\in G acting by xitaixix_{i}\mapsto t^{a_{i}}x_{i}, XsX^{s}\not=\emptyset if and only if there exists j,kj,k with aj>0a_{j}>0 and ak<0a_{k}<0. Let V=vV={\mathbb{C}}v, with AGA\in G acting by vdet(A)vv\mapsto\det(A)v. Then (VR)G(V\otimes R)^{G} is CM because we can identify VRΩRnΓ(QX(XH[n])[n])V\otimes R\cong\Omega^{n}_{R}\cong\Gamma(Q_{X}({\mathbb{Q}}_{X}^{H}[n])[-n]).

References

  • [Ach] Pramod Achar “Equivariant mixed Hodge modules” Lecture notes from the Clay Mathematics Institute workshop on Mixed Hodge Modules and Applications (2013), Preprint URL: https://www.math.lsu.edu/~pramod/docs/emhm.pdf
  • [Bou87] Jean-François Boutot “Singularités rationnelles et quotients par les groupes réductifs” In Invent. Math. 88.1, 1987, pp. 65–68 DOI: 10.1007/BF01405091
  • [Dim92] Alexandru Dimca “Singularities and topology of hypersurfaces”, Universitext Springer-Verlag, New York, 1992, pp. xvi+263 DOI: 10.1007/978-1-4612-4404-2
  • [Gro67] A. Grothendieck “Local cohomology. A seminar given by A. Grothendieck, Harvard University, Fall 1961. Notes by R. Hartshorne” 41, Lect. Notes Math. Springer, Cham, 1967 DOI: 10.1007/BFb0073971
  • [Hia23] Scott Hiatt “Duality and Local Cohomology in Hodge Theory”, 2023
  • [HK71] “Der kanonische Modul eines Cohen-Macaulay-Rings. (The canonical moduls of a Cohen-Macaulay-ring)” 238, Lect. Notes Math. Springer, Cham, 1971 DOI: 10.1007/BFb0059377
  • [HR74] Melvin Hochster and Joel L. Roberts “Rings of invariants of reductive groups acting on regular rings are Cohen-Macaulay” In Advances in Math. 13, 1974, pp. 115–175 DOI: 10.1016/0001-8708(74)90067-X
  • [HTT08] Ryoshi Hotta, Kiyoshi Takeuchi and Toshiyuki Tanisaki DD-modules, perverse sheaves, and representation theory” Translated from the 1995 Japanese edition by Takeuchi 236, Progress in Mathematics Birkhäuser Boston, Inc., Boston, MA, 2008, pp. xii+407 DOI: 10.1007/978-0-8176-4523-6
  • [Kem+73] G. Kempf, Finn Faye Knudsen, D. Mumford and B. Saint-Donat “Toroidal embeddings. I”, Lecture Notes in Mathematics, Vol. 339 Springer-Verlag, Berlin-New York, 1973, pp. viii+209
  • [Kir85] Frances Clare Kirwan “Partial desingularisations of quotients of nonsingular varieties and their Betti numbers” In Ann. of Math. (2) 122.1, 1985, pp. 41–85 DOI: 10.2307/1971369
  • [Kol86] János Kollár “Higher direct images of dualizing sheaves. I” In Ann. of Math. (2) 123.1, 1986, pp. 11–42 DOI: 10.2307/1971351
  • [Kol86a] János Kollár “Higher direct images of dualizing sheaves. II” In Ann. of Math. (2) 124.1, 1986, pp. 171–202 DOI: 10.2307/1971390
  • [KS21] Stefan Kebekus and Christian Schnell “Extending holomorphic forms from the regular locus of a complex space to a resolution of singularities” In J. Amer. Math. Soc. 34.2, 2021, pp. 315–368 DOI: 10.1090/jams/962
  • [Lun73] Domingo Luna “Slices étales” In Sur les groupes algébriques, Bull. Soc. Math. France, Mém. 33 Soc. Math. France, Paris, 1973, pp. 81–105 DOI: 10.24033/msmf.110
  • [Mat89] Hideyuki Matsumura “Commutative ring theory” Translated from the Japanese by M. Reid 8, Cambridge Studies in Advanced Mathematics Cambridge University Press, Cambridge, 1989, pp. xiv+320
  • [MFK94] D. Mumford, J. Fogarty and F. Kirwan “Geometric invariant theory” 34, Ergebnisse der Mathematik und ihrer Grenzgebiete (2) [Results in Mathematics and Related Areas (2)] Springer-Verlag, Berlin, 1994, pp. xiv+292
  • [Sai17] Morihiko Saito “A young person’s guide to mixed Hodge modules” In Hodge theory and L2L^{2}-analysis 39, Adv. Lect. Math. (ALM) Int. Press, Somerville, MA, 2017, pp. 517–553
  • [Sai88] Morihiko Saito “Modules de Hodge polarisables” In Publ. Res. Inst. Math. Sci. 24.6, 1988, pp. 849–995 (1989) DOI: 10.2977/prims/1195173930
  • [Sai90] Morihiko Saito “Mixed Hodge modules” In Publ. Res. Inst. Math. Sci. 26.2, 1990, pp. 221–333 DOI: 10.2977/prims/1195171082
  • [Sai91] Morihiko Saito “On Kollár’s conjecture” In Several complex variables and complex geometry, Part 2 (Santa Cruz, CA, 1989) 52, Proc. Sympos. Pure Math. Amer. Math. Soc., Providence, RI, 1991, pp. 509–517
  • [Sch19] Christian Schnell “An overview of Morihiko Saito’s theory of mixed Hodge modules” In Representation theory, automorphic forms & complex geometry Int. Press, Somerville, MA, [2019] ©2019, pp. 27–80
  • [Sta79] Richard P. Stanley “Combinatorics and invariant theory” In Relations between combinatorics and other parts of mathematics (Proc. Sympos. Pure Math., Ohio State Univ., Columbus, Ohio, 1978), Proc. Sympos. Pure Math., XXXIV Amer. Math. Soc., Providence, R.I., 1979, pp. 345–355
  • [Stacks] The Stacks Project Authors Stacks Project”, https://stacks.math.columbia.edu, 2018
  • [Van89] Michel Van den Bergh “Trace rings of generic matrices are Cohen-Macaulay” In J. Amer. Math. Soc. 2.4, 1989, pp. 775–799 DOI: 10.2307/1990894