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

On the collapsing of homogeneous bundles in arbitrary characteristic

András Cristian Lőrincz Humboldt–Universität zu Berlin, Institut für Mathematik, Berlin, Germany [email protected]
Abstract.

We study the geometry of equivariant, proper maps from homogeneous bundles G×PVG\times_{P}V over flag varieties G/PG/P to representations of GG, called collapsing maps. Kempf showed that, provided the bundle is completely reducible, the image GVG\cdot V of a collapsing map has rational singularities in characteristic zero. We extend this result to positive characteristic and show that for the analogous bundles the saturation GVG\cdot V is strongly FF-regular if its coordinate ring has a good filtration. We further show that in this case the images of collapsing maps of homogeneous bundles restricted to Schubert varieties are FF-rational in positive characteristic, and have rational singularities in characteristic zero. We provide results on the singularities and defining equations of saturations GXG\cdot X for PP-stable closed subvarieties XVX\subset V. We give criteria for the existence of good filtrations for the coordinate ring of GXG\cdot X.

Our results give a uniform, characteristic-free approach for the study of the geometry of a number of important varieties: multicones over Schubert varieties, determinantal varieties in the space of matrices, symmetric matrices, skew-symmetric matrices, and certain matrix Schubert varieties therein, representation varieties of radical square zero algebras (e.g. varieties of complexes), subspace varieties, higher rank varieties, etc.

SUR L’EFFONDREMENT DES FIBRéS HOMOGèNES EN CARACTéRISTIQUE ARBITRAIRE

Résumé. On étudie la géométrie des applications propres équivariantes de fibrés homogènes G×PVG\times_{P}V sur les variétés de drapeaux G/PG/P dans les représentations de GG, appelées applications d’effondrement. Kempf a montré que lorsque le fibré est complètement réductible, l’image GVG\cdot V d’une application d’effondrement a des singularités rationelles en caractéristique zéro. On étend ce résultat à la caractéristique positive et on montre que pour les fibrés analogues la saturation GVG\cdot V est fortement FF-régulière si son anneau des coordonnées a une bonne filtration. De plus, on montre que dans ce cas les images des applications d’effondrement de fibrés homogènes restreintes aux variétés de Schubert sont FF-rationelles en caractéristique positive, et ont des singularités rationelles en caractéristique zéro. On obtient des résultats sur les singularités et les équations qui définissent les saturations GXG\cdot X pour les sous-variétés XVX\subset V fermés PP-stables. On donne un critère pour l’existence de bonnes filtrations pour l’anneau des coordonnées de GXG\cdot X.

Nos résultats fournissent une approche uniforme et indépendante de la caractéristique, à l’étude de la géométrie de nombreuses variétés importantes: multicônes sur les variétés de Schubert, variétés déterminantales dans l’espace de matrices, matrices symétriques, matrices antisymétriques et certaines variétés de Schubert de matrices, variétés de représentations des algèbres dont le carré du radical est zéro (par ex. variétés de complexes), variétés de sous-espaces, variétés de rang supérieur, etc.

Key words and phrases:
Collapsing of bundles, Schubert varieties, FF-regularity, FF-rationality, rational singularities, good filtrations
2020 Mathematics Subject Classification:
14M15, 14L30, 13A35, 14B05, 14M05, 20G05, 14M12

1. Introduction

Let GG be a connected reductive group over an algebraically closed field 𝕜\Bbbk. Consider a parabolic subgroup PP of GG, and let WW be a GG-module and VWV\subset W a PP-stable submodule. The saturation GVWG\cdot V\subset W is the image of the homogeneous vector bundle G×PVG\times_{P}V under the proper “collapsing map”   G×PVWG\times_{P}V\to W  induced by the action of GG on WW.

Many remarkable varieties can be realized through such collapsing of bundles for various choices of GG, PP, WW, VV (cf. Section 4; for more such examples, see [Wey03]). Generally, the study of their geometry has been undertaken on case-by-case basis. An exception is the seminal work [Kem76], where it is shown that in characteristic zero GVG\cdot V has rational singularities whenever the unipotent radical U(P)U(P) of PP acts trivially on VV (see also [Kem86]). Further, in this case the singularities of GXG\cdot X are shown to be well-behaved for a closed PP-stable subvariety XVX\subset V [Kem76, Proposition 1 and Theorem 3].

In this paper, we generalize and extend the scope of Kempf’s results along several directions. In particular, we give characteristic-free strengthenings of the statements above, under the presence of good filtrations as initiated by Donkin [Don85], [Don90]. We say that a GG-variety ZZ is good, if 𝕜[Z]\Bbbk[Z] has a good filtration (see Section 2.4). We point out to the reader that all good-related properties hold automatically when char𝕜=0\operatorname{char}\Bbbk=0, and our results below are new in this case as well (with the exception of Theorem 1.3).

Let BPB\subset P a Borel subgroup of GG and TBT\subset B a maximal torus. We denote the set of dominant weights of GG by X(T)+X(T)_{+}. For λX(T)+\lambda\in X(T)_{+} we let ΔG(λ)\Delta_{G}(\lambda) denote the corresponding Weyl module (see Section 2.2). We consider the Levi decomposition P=LU(P)P=L\ltimes U(P) with LL reductive. Pick any λ1,λ2,,λnX(T)+\lambda_{1},\lambda_{2},\dots,\lambda_{n}\in X(T)_{+}, and for the rest of the introduction fix

(1.1) W=i=1nΔG(λi)andV=i=1nΔL(λi).W=\bigoplus_{i=1}^{n}\Delta_{G}(\lambda_{i})\quad\mbox{and}\quad V=\bigoplus_{i=1}^{n}\Delta_{L}(\lambda_{i}).

We have a natural inclusion VWU(P)V\subseteq W^{U(P)}, with equality if char𝕜=0\operatorname{char}\Bbbk=0 (when the bundle is completely reducible [Kem76]). While the examples in Section 4 fit into the setup (1.1), we note that in Section 3 we develop the results in a more general setting (see (3.1)).

Theorem 1.2.

Let XVX\subset V be an LL-submodule such that GXG\cdot X is good. Then GXG\cdot X is strongly FF-regular when char𝕜>0\operatorname{char}\Bbbk>0 (resp. is of strongly FF-regular type when char𝕜=0\operatorname{char}\Bbbk=0).

This illustrates that good filtrations are responsible for the geometric behavior of saturations in positive characteristic, a phenomenon that is apparent in invariant theory as well [Has01], [Has12]. Example 4.4 demonstrates that this assumption cannot be dropped.

The following is our main criterion for the existence of good filtrations (for the definition of good pairs, see Section 2.4).

Theorem 1.3.

Assume that WW is good, and that (V,X)(V,X) is a good pair for some closed LL-variety XVX\subset V. Then (W,GX)(\,W\,,\,\,G\cdot X) is a good pair of GG-varieties.

In particular, this implies that GVG\cdot V is good whenever char𝕜>max{dimΔG(λi)| 1in}\operatorname{char}\Bbbk>\max\{\dim\Delta_{G}(\lambda_{i})\,|\,1\leq i\leq n\}. However, in concrete situations the bound on char𝕜\operatorname{char}\Bbbk can be further improved significantly (cf. Sections 4.1, 4.2). See Theorem 3.6 for other criteria in this direction.

We extend the collapsing method to various relative settings, thus greatly increasing its versatility. These include restrictions to Schubert varieties or multiplicity-free subvarieties of flag varieties (for the latter, see Corollary 3.14). Below 𝒲\mathcal{W} denotes the Weyl group of GG.

Theorem 1.4.

Consider a closed LL-variety XVX\subset V and assume that GXG\cdot X is good. For any w𝒲w\in\mathcal{W}, we have:

  1. (1)

    BwX¯\overline{BwX} is normal if and only if XX is so.

  2. (2)

    If char𝕜=0\operatorname{char}\Bbbk=0, then BwX¯\overline{BwX} has rational singularities if and only if so does XX.

  3. (3)

    If char𝕜>0\operatorname{char}\Bbbk>0 and XX is an LL-submodule of VV, then BwX¯\overline{BwX} is FF-rational.

Note that when ww is the longest element in 𝒲\mathcal{W}, we have BwX¯=GX\overline{BwX}=G\cdot X.

Frequently (e.g. when GXG\cdot X is a spherical variety), the varieties BwX¯\overline{BwX} are orbit closures under the action of the Borel subgroup BB (see Section 4). The singularities of such varieties have been investigated mostly in the spherical case (e.g. [RR85], [Bri01], [BT06]), but they are not well understood [Per14, Comments 4.4.4]. Theorem 1.4 is one of the first of its kind at this level of generality, applicable equally in non-spherical situations as well.

When PP is itself a Borel subgroup, we sharpen some results on singularities (see Corollary 3.13), extending the case of multicones over Schubert varieties [KR87], [Has06].

Next, we provide a relative result on the defining ideals of saturations GXG\cdot X. For this, we introduce the notion of good generators of an ideal, see Definition 2.13.

Theorem 1.5.

Let (V,X)(V,X) be a good pair with GVG\cdot V good, and denote by IX𝕜[V]I_{X}\subset\Bbbk[V] the defining ideal of XVX\subset V. Let MM be the span of a set of good generators of IXI_{X} and take a basis 𝒫\mathcal{P}^{\prime} of the GG-module H0(G/P,𝒱(M))𝕜[GV]H^{0}(G/P,\mathcal{V}(M))\,\subset\Bbbk[G\cdot V]. Consider:

  1. (1)

    A set of generators 𝒫GV\mathcal{P}_{G\cdot V} of the defining ideal IGV𝕜[W]I_{G\cdot V}\subset\Bbbk[W] of GVG\cdot V;

  2. (2)

    A lift 𝒫~𝕜[W]\tilde{\mathcal{P}^{\prime}}\subset\Bbbk[W] of the set 𝒫𝕜[W]/IGV\mathcal{P}^{\prime}\subset\,\Bbbk[W]/I_{G\cdot V}.

Then the defining ideal of GXG\cdot X in 𝕜[W]\Bbbk[W] is generated by the set 𝒫GV𝒫~\mathcal{P}_{G\cdot V}\,\cup\,\tilde{\mathcal{P}^{\prime}}.

In Theorem 3.15 we give a version of the above that yields good defining equations, which we use to readily find (good) defining equations for the examples in Sections 4.1 and 4.2.

Saturations of the type GVG\cdot V appear in various forms throughout the existing literature, and a range of techniques have been developed to better understand their geometry. Applying the results above in the special case of radical square zero algebras (see Section 4.2), we simultaneously sharpen and generalize the main results in [Kem75], [DCS81], [Str82], [Bri85], [Str87], [MT99a], [MT99b] that concern the singularities and defining equations of the Buchsbaum–Eisenbud varieties of complexes as well as varieties of complexes of other type. In addition, we obtain that certain BB-orbit closures in varieties of complexes are FF-rational when char𝕜>0\operatorname{char}\Bbbk>0 (resp. have rational singularities when char𝕜=0\operatorname{char}\Bbbk=0).

Our results provide a general method for the investigation of the geometry of parabolically induced orbit closures in a representation WW of a reductive group GG. Namely, for any choice of a parabolic PGP\subset G, we can take the representation VV of the smaller reductive group LL as in (1.1) with trivial U(P)U(P)-action; choosing an LL-orbit closure X=Lx¯X=\overline{Lx} (for any xVx\in V), saturation gives a GG-orbit closure GX=Gx¯GVWG\cdot X=\overline{Gx}\subset G\cdot V\subset W. By considering all such possible choices, we obtain a large set of GG-orbit closures in WW whose singularities and defining equations are inherited from the smaller ones according to the results above.

The Cohen–Macaulay property for collapsing of bundles in positive characteristic is a consequence of the study of their FF-singularities. This relies on techniques from tight closure theory that was developed by Hochster and Huneke [HH90], [HH94a]. In Section 4.4, we translate this property into Griffiths-type vanishing results for the cohomology of such bundles on Schubert varieties in positive characteristic, extending the classical Kodaira-type vanishing results for line bundles [MR85], [LRPT06], [Smi00].

Acknowledgments

The author would like to express his gratitude to Ryan Kinser for his valuable comments and suggestions on this work.

2. Preliminaries

We work over an algebraically closed field 𝕜\Bbbk of arbitrary characteristic (see Remark 3.11).

An action of an algebraic group GG on an algebraic variety XX is always assumed to be algebraic, so that the map G×XXG\times X\to X is a morphism of algebraic varieties. We call a (possibly infinite-dimensional) vector space VV a rational GG-module, if VV is equipped with a linear action of GG, such that every vVv\in V is contained in a finite-dimensional GG-stable subspace on which GG acts algebraically. All modules considered are assumed to be rational of countable dimension.

Unless otherwise stated, throughout a ring or algebra is commutative, finitely generated over 𝕜\Bbbk with a multiplicative identity.

2.1. Reductive groups

Let GG be a connected reductive group over 𝕜\Bbbk, BB a Borel subgroup and UU its unipotent radical. We fix a maximal torus TBT\subset B, and denote by X(T)X(T) its group of characters. We denote by ,\langle\cdot,\cdot\rangle the standard pairing between X(T)X(T) and the group of cocharacters. Let ΦX(T)\Phi\subset X(T) denote the set of roots and Φ+Φ\Phi_{+}\subset\Phi the set of positive roots with respect to the choice of BB. We denote by ρ\rho the half sum of all the positive roots. The set of simple roots in Φ+\Phi_{+} is denoted by SS. We let 𝒲=N(T)/T\mathcal{W}=N(T)/T be the Weyl group of GG, and w0Ww_{0}\in W its longest element.

For ISI\subset S, consider the standard parabolic subgroup P:=PIGP:=P_{I}\subset G. We have a Levi decomposition PI=LIUIP_{I}=L_{I}\ltimes U_{I}, where UIU_{I} is the unipotent radical of PP and L:=LIL:=L_{I} is reductive. Let 𝒲I\mathcal{W}_{I} be the subgroup generated by the reflections sαs_{\alpha} with αI\alpha\in I, and wIw_{I} the longest element in 𝒲I\mathcal{W}_{I}. We choose the set 𝒲I\mathcal{W}^{I} of representatives of the cosets of 𝒲/𝒲I\mathcal{W}/\mathcal{W}_{I} as

(2.1) 𝒲I={w𝒲|w(α)Φ+, for all αI}.\mathcal{W}^{I}=\{w\in\mathcal{W}\,|\,w(\alpha)\in\Phi_{+},\mbox{ for all }\alpha\in I\}.

We have the Bruhat decomposition of GG into B×PB\times P-orbits (see [Jan03, Section II.13]):

G=w𝒲IBwP.G\,=\,\bigcup_{w\in\mathcal{W}^{I}}BwP.

For w𝒲Iw\in\mathcal{W}^{I}, we put U(w):=UwUw1U(w):=U\cap wU^{-}w^{-1}, where UU^{-} is the opposite unipotent radical. The multiplication map induces an isomorphism of U(w)U(w)-varieties (see [Jan03, Section II.13.8])

(2.2) U(w)×PBwP,(u,p)uwp.U(w)\times P\xrightarrow{\cong}BwP,\qquad(u,p)\,\mapsto\,uwp.

We denote by X(w)PX(w)_{P} the Schubert variety that is the image of BwP¯\overline{BwP} under the locally trivial projection GG/PG\to G/P. For P=BP=B, we write X(w):=X(w)BX(w):=X(w)_{B}.

2.2. Cohomology of homogeneous bundles

For any representation MM of PP, we denote by 𝒱(M)\mathcal{V}(M) the sheaf of sections of the homogeneous vector bundle G×PMG\times_{P}M. For λX(T)\lambda\in X(T), we put (λ):=𝒱(𝕜λ)\mathcal{L}(\lambda):=\mathcal{V}(\Bbbk_{-\lambda}), where 𝕜λ\Bbbk_{-\lambda} is the 11-dimensional representation of BB.

A weight λX(T)\lambda\in X(T) is dominant if λ,α0\langle\lambda,\alpha^{\vee}\rangle\geq 0, for all simple roots αS\alpha\in S. The set of dominant weights is denoted by X(T)+X(T)_{+}. For λX(T)+\lambda\in X(T)_{+}, we call the space of sections

G(λ):=H0(G/B,(λ)),\nabla_{G}(\lambda):=H^{0}(G/B,\mathcal{L}(\lambda)),

a dual Weyl module. It has lowest weight λ-\lambda and highest weight w0λ-w_{0}\cdot\lambda. The module ΔG(λ)=G(λ)\Delta_{G}(\lambda)=\nabla_{G}(\lambda)^{*} is called a Weyl module, that has a non-zero highest weight vector of weight λ\lambda, and this generates ΔG(λ)\Delta_{G}(\lambda) as a GG-module. It is known that G(λ)\nabla_{G}(\lambda) has a unique simple submodule, of highest weight w0λ-w_{0}\cdot\lambda.

When char𝕜=p>0\operatorname{char}\Bbbk=p>0 and e1e\geq 1 is an integer such that (pe1)ρ(p^{e}-1)\rho is a weight of GG, we denote by Ste=G((pe1)ρ)\operatorname{St}_{e}=\nabla_{G}((p^{e}-1)\rho) the ethe^{\operatorname{th}} Steinberg module, and put St:=St1\operatorname{St}:=\operatorname{St}_{1}. The assumption is superficial as we can always replace GG by radG×G~\operatorname{rad}G\times\tilde{G}, where radG\operatorname{rad}G denotes the radical of GG and G~\tilde{G} the universal cover of [G,G][G,G], and (pe1)ρ(p^{e}-1)\rho is a weight of radG×G~\operatorname{rad}G\times\tilde{G} for all e1e\geq 1.

Let P=PIP=P_{I} be a parabolic subgroup. For λX(T)+\lambda\in X(T)_{+}, put 𝒱(λ):=𝒱(L(λ))\mathcal{V}(\lambda):=\mathcal{V}(\nabla_{L}(\lambda)) (here UIU_{I} acts trivially on L(λ)\nabla_{L}(\lambda)). The quotient map π:G/BG/P\pi:G/B\to G/P induces a quasi-isomorphism

(2.3) 𝐑π(λ)𝒱(λ).\mathbf{R}\pi_{*}\mathcal{L}(\lambda)\cong\mathcal{V}(\lambda).

By abuse of notation, we use the same notation for the respective bundles on Schubert varieties that are obtained by restriction. We record the following result.

Lemma 2.4.

Let λX(T)+\lambda\in X(T)_{+}, and w𝒲Iw\in\mathcal{W}^{I}. For all i0i\geq 0 we have Hi(X(w)P,𝒱(λ))Hi(X(wwI),(λ))H^{i}(X(w)_{P},\mathcal{V}(\lambda))\cong H^{i}(X(w\cdot w_{I}),\,\mathcal{L}(\lambda)), and the map induced by restriction is surjective:

Hi(G/P,𝒱(λ))Hi(X(w)P,𝒱(λ)).H^{i}(G/P,\,\mathcal{V}(\lambda))\to H^{i}(X(w)_{P},\mathcal{V}(\lambda)).

Moreover, Hi(G/P,𝒱(λ))=0H^{i}(G/P,\,\mathcal{V}(\lambda))=0 for i>0i>0.

Proof.

We have π1(X(w)P)=X(wwI)\pi^{-1}(X(w)_{P})=X(ww_{I}) [Jan03, Section 13.8] and a Cartesian square

G/B\textstyle{G/B\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}G/P\textstyle{G/P}X(wwI)\textstyle{X(ww_{I})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πw\scriptstyle{\,\,\pi^{w}}X(w)\textstyle{X(w)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

where the vertical maps are inclusions. As π\pi is proper and flat, by a base change argument (see [Har77, Corollary 12.9]) and (2.3) we get 𝐑πw(λ)𝒱(λ)\mathbf{R}\pi_{*}^{w}\,\mathcal{L}(\lambda)\cong\mathcal{V}(\lambda). This shows that Hi(X(w)P,𝒱(λ))Hi(X(wwI),(λ))H^{i}(X(w)_{P},\mathcal{V}(\lambda))\cong H^{i}(X(ww_{I}),\mathcal{L}(\lambda)) for all i0i\geq 0. The rest of the claims now follows from the diagram above using [RR85, Theorem 2]. ∎

2.3. Classes of singularities

When char𝕜=p>0\operatorname{char}\Bbbk=p>0, for a 𝕜\Bbbk-space VV and e0e\in\mathbb{Z}_{\geq 0} we denote by V(e)V^{(e)} the abelian group VV with the new 𝕜\Bbbk-space structure cv:=c1/pevc\cdot v:=c^{1/p^{e}}\cdot v. When VV is a module over an algebraic group GG, then V(e)V^{(e)} also has a GG-module structure [Jan03, Section I.9.10]. If AA is a 𝕜\Bbbk-algebra, then so is A(e)A^{(e)} by using the same multiplicative structure.

We call a domain AA strongly FF-regular if for every non-zero cAc\in A there exists e>0e>0 such that the A(e)A^{(e)}-map cFe:A(e)AcF^{e}:A^{(e)}\to A given by xcxpex\mapsto cx^{p^{e}} is A(e)A^{(e)}-split.

As we do not need it for our purposes, we refer the reader to [HH94a] for the definition of FF-rational rings (see (2.5) below for some of its important properties).

When char𝕜=0\operatorname{char}\Bbbk=0, an algebraic variety XX has rational singularities, if for some (hence, any) resolution of singularities f:ZXf:Z\to X (i.e. ZZ is smooth, and ff proper and birational), the natural map 𝒪X𝐑f𝒪Z\mathcal{O}_{X}\to\mathbf{R}f_{*}\mathcal{O}_{Z} is a (quasi-)isomorphism. Further, we say a ring AA is of strongly FF-regular type if there exist some subring RR of 𝕜\Bbbk which is of finite type over \mathbb{Z}, and some RR-algebra ARA_{R} which is flat of finite type over RR, such that ARR𝕜AA_{R}\otimes_{R}\Bbbk\cong A and for the closed points 𝔪\mathfrak{m} in a dense open subset of SpecR\operatorname{Spec}R, the ring ARRR/𝔪A_{R}\otimes_{R}R/\mathfrak{m} is strongly FF-regular.

An affine variety XX is FF-rational (resp. strongly FF-regular or of strongly FF-regular type) if 𝕜[X]\Bbbk[X] is so. We have the following implications (where CM stands for Cohen–Macaulay):

(2.5) char𝕜=0:\displaystyle\operatorname{char}\Bbbk=0: regularstrongly F-regular typerational sing.normal, CM;\displaystyle\mbox{ regular}\Rightarrow\mbox{strongly }F\mbox{-regular type}\!\Rightarrow\mbox{rational sing.}\Rightarrow\mbox{normal, CM};
char𝕜>0:\displaystyle\operatorname{char}\Bbbk>0: regularstrongly F-regularF-rationalnormal, CM.\displaystyle\mbox{ regular}\Rightarrow\,\mbox{strongly }F\mbox{-regular}\,\,\Longrightarrow\,F\mbox{-rational}\Rightarrow\mbox{normal, CM.}

Furthermore, FF-rationality implies pseudo-rationality [Smi97] and rational singularities in positive characteristic as defined in [Kov20]. When char𝕜=0\operatorname{char}\Bbbk=0, a ring has log terminal singularities if and only if it is of strongly FF-regular type and \mathbb{Q}-Gorenstein (see [HW02]).

Now let AA be a GG-algebra and char𝕜=p>0\operatorname{char}\Bbbk=p>0. We can assume that (pe1)ρ(p^{e}-1)\rho is a weight of GG for e1e\geq 1 (otherwise replace GG by radG×G~\operatorname{rad}G\times\tilde{G}). Following [Has12, Section 4], we say that AA is GG-FF-pure if there exists some e1e\geq 1 such that the map idFe:SteA(e)SteA\operatorname{id}\otimes F^{e}:\operatorname{St}_{e}\otimes A^{(e)}\to\operatorname{St}_{e}\otimes A splits as a (G,A(e))(G,A^{(e)})-linear map.

Now we study the coordinate ring of BwB¯G\overline{BwB}\subset G, where w𝒲w\in\mathcal{W}. Consider the rational B×TB\times T-subalgebra of 𝕜[BwB¯]U\Bbbk[\overline{BwB}]^{U} consisting of dominant TT-weight spaces

𝕜[BwB¯]+U:=λX(T)+𝕜[BwB¯]λU.\Bbbk[\overline{BwB}]^{U}_{+}:=\bigoplus_{\lambda\in X(T)_{+}}\Bbbk[\overline{BwB}]^{U}_{\lambda}.

Consider the section ring C(X(w)):=λX(T)+H0(X(w),(λ))C(X(w)):=\bigoplus_{\lambda\in X(T)_{+}}\!H^{0}(X(w),\mathcal{L}(\lambda)).

Lemma 2.6.

For any w𝒲w\in\mathcal{W}, we have an isomorphism of B×TB\times T-algebras

𝕜[BwB¯]+UC(X(w)).\Bbbk[\overline{BwB}]^{U}_{+}\cong C(X(w)).

As a consequence, the algebra 𝕜[BwB¯]+U\Bbbk[\overline{BwB}]^{U}_{+} is finitely generated, and strongly FF-regular when char𝕜>0\operatorname{char}\Bbbk>0 (resp. of strongly FF-regular type when char𝕜=0\operatorname{char}\Bbbk=0).

Proof.

Let Γ:=X(T)+\Gamma:=-X(T)_{+} and consider the semigroup ring 𝕜[Γ]\Bbbk[\Gamma], which is naturally a subalgebra of 𝕜[T]\Bbbk[T]. We have an isomorphism of B×TB\times T-algebras

(𝕜[BwB¯]U𝕜[Γ])TC(X(w)).(\Bbbk[\overline{BwB}]^{U}\otimes\Bbbk[\Gamma])^{T}\cong C(X(w)).

On the other hand, we have

(𝕜[BwB¯]U𝕜[Γ])T=(𝕜[BwB¯]+U𝕜[Γ])T=(𝕜[BwB¯]+U𝕜[T])T𝕜[BwB¯]+U,(\Bbbk[\overline{BwB}]^{U}\otimes\Bbbk[\Gamma])^{T}=(\Bbbk[\overline{BwB}]^{U}_{+}\otimes\Bbbk[\Gamma])^{T}=(\Bbbk[\overline{BwB}]^{U}_{+}\otimes\Bbbk[T])^{T}\cong\Bbbk[\overline{BwB}]^{U}_{+},

where the second equality follows from the decomposition 𝕜[T]λX(T)𝕜λ\Bbbk[T]\cong\bigoplus_{\lambda\in X(T)}\Bbbk_{\lambda} as TT-modules, and the last isomorphism from Lemma 2.17.

We now show that C(X(w))C(X(w)) is finitely generated. By [Gro97, Theorem 16.2], 𝕜[G]U\Bbbk[G]^{U} is finitely generated, and therefore so is C(X(w0))=(𝕜[G]U𝕜[Γ])TC(X(w_{0}))=(\Bbbk[G]^{U}\otimes\Bbbk[\Gamma])^{T}. By Lemma 2.4, we see that the map C(X(w0))C(X(w))C(X(w_{0}))\to C(X(w)) induced by restriction is onto, hence C(X(w))C(X(w)) is finitely generated (alternatively, this follows also from [RR85, Theorem 2]).

Let char𝕜>0\operatorname{char}\Bbbk>0. The (not necessarily noetherian) algebra λX(T)H0(X(w),(λ))\bigoplus_{\lambda\in X(T)}H^{0}(X(w),\mathcal{L}(\lambda)) is quasi-FF-regular, by [Has03, Theorem 2.6 (4)] and the global FF-regularity of Schubert varieties in the sense of [Smi00], see [LRPT06], [Has06]. Therefore, the algebra C(X(w))C(X(w)) is also quasi-FF-regular by [Has03, Lemma 2.4]. The latter is finitely generated, so strongly FF-regular (see [Has03, Section 2.1]).

For ring RR, consider the RR-algebra C(X(w)R)=λX(T)+H0(X(w)R,(λ)R)C(X(w)_{R})=\bigoplus_{\lambda\in X(T)_{+}}\!H^{0}(X(w)_{R},\mathcal{L}(\lambda)_{R}). We have C(X(w)𝕜)=C(X(w))kC(X(w)_{\Bbbk^{\prime}})=C(X(w)_{\mathbb{Z}})\otimes_{\mathbb{Z}}k^{\prime} (see [Jan03, Section II.14.15]), for any field 𝕜\Bbbk^{\prime}, and C(X(w))C(X(w)_{\mathbb{Z}}) is flat and finitely generated over \mathbb{Z} (e.g. from [Jan03, Sections II.14.1 and II.14.21]). By [HH94a, Theorem 5.5], C(X(w)𝕜)C(X(w)_{\Bbbk^{\prime}}) is strongly FF-regular for a perfect field 𝕜𝕜\Bbbk^{\prime}\subset\Bbbk. This shows that when char𝕜=0\operatorname{char}\Bbbk=0, C(X(w))C(X(w)) is of strongly FF-regular type. ∎

For the remainder of the subsection, we assume that char𝕜>0\operatorname{char}\Bbbk>0.

Lemma 2.7.

Let ΓX(T)+\Gamma\subset X(T)_{+} be a finitely generated semigroup, and A=λΓAλA=\bigoplus_{\lambda\in\Gamma}A_{\lambda} a Γ\Gamma-graded integral domain with a GG-action such that AλG(λ)A_{\lambda}\cong\nabla_{G}(\lambda). Then AA is GG-FF-pure.

Proof.

The proof follows closely that of [Has11, Lemma 3]. We can assume that G=G~×radGG=\tilde{G}\times\operatorname{rad}G. Further, we can assume that the product G(λ)G(μ)G(λ+μ)\nabla_{G}(\lambda)\otimes\nabla_{G}(\mu)\to\nabla_{G}(\lambda+\mu) in AA is given by multiplication of sections of the corresponding line bundles on G/BG/B, as seen in the proof of [Has03, Lemma 5.6]. We denote by ϕ\phi the composition of GG-maps

ϕ:StAλΓStG(pλ)λΓG(p(λ+ρ)ρ)StA(1),\phi\colon\operatorname{St}\otimes A\twoheadrightarrow\bigoplus_{\lambda\in\Gamma}\operatorname{St}\otimes\nabla_{G}(p\lambda)\twoheadrightarrow\bigoplus_{\lambda\in\Gamma}\nabla_{G}(p(\lambda+\rho)-\rho)\xrightarrow{\cong}\operatorname{St}\otimes A^{(1)},

where the first map is given by projection, the second by multiplication (see [RR85, Theorem 1]), and the third by the inverse of GG-isomorphism StG(λ)(1)G(p(λ+ρ)ρ)\operatorname{St}\otimes\nabla_{G}(\lambda)^{(1)}\xrightarrow{\cong}\nabla_{G}(p(\lambda+\rho)-\rho) induced also by multiplication of sections (see [And80, Theorem 2.5]). Then ϕ\phi gives the required splitting, since it is A(1)A^{(1)}-linear. The latter can be checked on the graded components, where it follows from the commutative diagram (with the obvious maps induced by multiplication):

StG(pλ)G(μ)(1)\textstyle{\operatorname{St}\otimes\nabla_{G}(p\lambda)\otimes\nabla_{G}(\mu)^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G(p(λ+ρ)ρ)G(μ)(1)\textstyle{\nabla_{G}(p(\lambda+\rho)-\rho)\otimes\nabla_{G}(\mu)^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}StG(λ)(1)G(μ)(1)\textstyle{\operatorname{St}\otimes\nabla_{G}(\lambda)^{(1)}\otimes\nabla_{G}(\mu)^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}StG(p(λ+μ))\textstyle{\operatorname{St}\otimes\nabla_{G}(p(\lambda+\mu))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\quad}G(p(λ+μ+ρ)ρ)\textstyle{\quad\nabla_{G}(p(\lambda+\mu+\rho)-\rho)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\quad}\scriptstyle{\cong}StG(λ+μ)(1)\textstyle{\quad\operatorname{St}\otimes\nabla_{G}(\lambda+\mu)^{(1)}}

When Γ\Gamma is saturated, the algebra AA as above is strongly FF-regular [Has03, Lemma 5.6].

Corollary 2.8.

The algebra 𝕜[G]UI×U\Bbbk[G]^{U_{I}\times U} is strongly FF-regular and LL-FF-pure.

Proof.

The algebra A=𝕜[G]UI×UA=\Bbbk[G]^{U_{I}\times U} has an L×TL\times T-action so that we have a decomposition A=λΓL(wIw0λ)A=\bigoplus_{\lambda\in\Gamma}\nabla_{L}(w_{I}w_{0}\lambda) as LL-modules (e.g. see [Don88, Theorem 3]). Clearly, the set {wIw0λ}λX(T)+\{w_{I}w_{0}\lambda\}_{\lambda\in X(T)_{+}} forms a saturated subsemigroup in the semigroup of dominant weights of LL. Hence, the claims follow by [Has03, Lemma 5.6] and Lemma 2.7, respectively. ∎

2.4. Good filtrations

Take a (possibly infinite-dimensional) GG-module VV. Following Donkin [Don85], an ascending exhaustive filtration

0=V0V1V20=V_{0}\subset V_{1}\subset V_{2}\subset\dots

of GG-submodules of VV is a good filtration (resp. Weyl filtration) of VV, if each Vi/Vi1V_{i}/V_{i-1} is isomorphic to a dual Weyl module (resp. to a Weyl module). If VV has both good and Weyl filtrations, then we call VV tilting.

Now let w𝒲w\in\mathcal{W}. We say that a BB-module VV has a ww-excellent filtration, if it has a BB-module filtration with successive quotients isomorphic to some H0(X(w),(λ))H^{0}(X(w),\mathcal{L}(\lambda)), with λX(T)+\lambda\in X(T)_{+}. This is a special type of excellent filtration, as defined in [vdK93, Definition 2.3.6]. Note that a good filtration of a GG-module is a w0w_{0}-excellent filtration.

A finite-dimensional GG-module WW good if SymdW\operatorname{Sym}_{d}W^{*} has a good filtration for all d0d\geq 0. In particular, in this case WW must have a Weyl filtration. Similarly, we call an affine GG-variety (resp. BB-variety) XX good (resp ww-excellent) if 𝕜[X]\Bbbk[X] has a good (resp. ww-excellent) filtration.

If XYX\subset Y is a closed GG-stable subvariety, then we say that (Y,X)(Y,X) is a good pair whenever YY is good and the defining ideal IX𝕜[Y]I_{X}\subset\Bbbk[Y] has a good filtration (see [Don90, Section 1.3]). In this case XX is automatically good.

If char𝕜=0\operatorname{char}\Bbbk=0, then all (pairs of) affine GG-varieties are good. An important feature of good filtrations is the following result of Donkin [Don85] and Mathieu [Mat90, Theorem 1].

Proposition 2.9.

If MM and NN are GG-modules with good filtrations, then M𝕜NM\otimes_{\Bbbk}N has a good filtration. In particular, if XX and YY are good affine GG-varieties, then so is X×YX\times Y.

We list some cases that imply the existence of good filtrations (see [AJ84, Section 4]).

Lemma 2.10.

Let V,WV,W be finite-dimensional GG-modules.

  1. (1)

    If χ+ρ,αchar𝕜\langle\chi+\rho,\alpha^{\vee}\rangle\leq\operatorname{char}\Bbbk for all weights χ\chi of VV and all αΦ+\alpha\in\Phi_{+}, then VV has a good filtration.

  2. (2)

    If VV has a good filtration and char𝕜>i\operatorname{char}\Bbbk>i, then iV\bigwedge^{i}V and SymiV\operatorname{Sym}_{i}V have good filtrations.

  3. (3)

    If V\bigwedge V and W\bigwedge W have good filtrations, then VWV\otimes W is good.

  4. (4)

    V\bigwedge V has a good filtration if and only if so does V\bigwedge V^{*} (i.e. V\bigwedge V is tilting).

We further need some basic results.

Lemma 2.11.

Let f:MNf:M\to N be a GG-module map. If MM has a good filtration and the induced map MUNUM^{U}\to N^{U} is onto, then NN and kerf\ker f have good filtrations and ff is onto.

Proof.

Put I=imagefI=\operatorname{image}f and K=kerfK=\ker f. Fix any λX(T)+\lambda\in X(T)_{+}. Since MM has a good filtration, we have an exact sequence (see [Jan03, Proposition II.4.16])

0HomG(ΔG(λ),K)HomG(ΔG(λ),M)HomG(ΔG(λ),I)ExtG1(ΔG(λ),K)0.0\to\operatorname{Hom}_{G}(\Delta_{G}(\lambda),K)\to\operatorname{Hom}_{G}(\Delta_{G}(\lambda),M)\to\operatorname{Hom}_{G}(\Delta_{G}(\lambda),I)\to\operatorname{Ext}^{1}_{G}(\Delta_{G}(\lambda),K)\to 0.

The assumption gives an exact sequence

0KUMUIU0.0\to K^{U}\to M^{U}\to I^{U}\to 0.

Taking λ\lambda-weights above we obtain that ExtG1(ΔG(λ),K)=0\operatorname{Ext}^{1}_{G}(\Delta_{G}(\lambda),K)=0 (see [Jan03, Lemma II.2.13]). Since λX(T)+\lambda\in X(T)_{+} was arbitrary, this shows that KK has a good filtration (see [Jan03, Proposition II.4.16]), and hence so does II. Let C=cokerfC=\operatorname{coker}f and consider an exact sequence 0INC00\to I\to N\to C\to 0. Since II has a good filtration, we see as above that the induced sequence 0IUNUCU00\to I^{U}\to N^{U}\to C^{U}\to 0 is also exact. By assumption CU=0C^{U}=0, hence C=0C=0. ∎

Corollary 2.12.

Let YY be a good affine GG-variety and XYX\subset Y a closed GG-stable subvariety. Then (Y,X)(Y,X) is a good pair if and only if the map 𝕜[Y]U𝕜[X]U\Bbbk[Y]^{U}\to\Bbbk[X]^{U} is surjective.

Proof.

If 𝕜[Y]U𝕜[X]U\Bbbk[Y]^{U}\to\Bbbk[X]^{U} is surjective, then it follows from Lemma 2.11 that (Y,X)(Y,X) is a good pair. The converse follows from [Don88, Proposition 1.4 and Proposition 2]. ∎

We introduce a notion for generators of ideals, that is again relevant only in positive characteristic.

Definition 2.13.

Let YY be a good affine GG-variety and XYX\subset Y a closed GG-stable subvariety with defining ideal IX𝕜[Y]I_{X}\subset\Bbbk[Y]. We say that a finite set of equations 𝒫IX\mathcal{P}\subset I_{X} are good defining equations (resp. good generators) of XX (resp. of IXI_{X}) if the following hold for M𝒫:=span𝕜𝒫IXM_{\mathcal{P}}:=\operatorname{span}_{\Bbbk}\mathcal{P}\,\subset I_{X}:

  1. (1)

    M𝒫M_{\mathcal{P}} is a GG-module with a good filtration;

  2. (2)

    The multiplication map m𝒫:k[Y]M𝒫IXm_{\mathcal{P}}:k[Y]\otimes M_{\mathcal{P}}\to I_{X} induces a surjective map on UU-invariants (k[Y]M𝒫)UIXU(k[Y]\otimes M_{\mathcal{P}})^{U}\to I_{X}^{U}.

Let us record some useful results regarding this notion. We continue with the notation in Definition 2.13.

Lemma 2.14.

There exist good defining equations for XYX\subset Y if and only if (Y,X)(Y,X) is a good pair.

Proof.

Assume that (Y,X)(Y,X) is a good pair. By [Gro97, Theorem 16.2], 𝕜[Y]U\Bbbk[Y]^{U} is noetherian, hence IXUI_{X}^{U} is finitely generated. Choose a finite set of generators. Taking a good filtration of IXI_{X}, there exists a finite dimensional piece MM that contains these generators. We can pick 𝒫\mathcal{P} to be a basis of MM.

Conversely, let 𝒫IX\mathcal{P}\subset I_{X} be a set of good generators. By Proposition 2.9, the domain of the multiplication map m𝒫m_{\mathcal{P}} has a good filtration. By Lemma 2.11, we obtain that m𝒫m_{\mathcal{P}} is surjective, and IXI_{X} has a good filtration. ∎

The proof above shows assumption (2) in Definition 2.13 can be replaced with the equivalent assumption that 𝒫\mathcal{P} generates IXI_{X} and kerm𝒫\ker m_{\mathcal{P}} has a good filtration. In particular, the notion does not depend on the choice of the Borel subgroup (see [Jan03, Remark II.4.16 (2)]). We record another convenient fact.

Lemma 2.15.

Assume that YY is good and let MIXM\subset I_{X} be GG-module such that a basis 𝒫\mathcal{P} of MM generates IXI_{X} and forms a regular sequence in 𝕜[Y]\Bbbk[Y]. Assume that M\bigwedge M has a good filtration. Then 𝒫\mathcal{P} are good defining equations of XYX\subset Y.

Proof.

This follows readily by considering the Koszul resolution, and using [Don85, Proposition 3.2.4] together with Proposition 2.9 repeatedly. ∎

Although we do not need it in this article, the assumption on M\bigwedge M in the lemma above can be weakened by requiring only that the good filtration dimension of iM\bigwedge^{i}M is at most i1i-1, for all i1i\geq 1 (see [Don90, Section 1.3]).

2.5. Deformation of algebras

We recall a filtration of algebras considered in [Pop86] and [Gro92]. There exists a homomorphism h:X(T)h:X(T)\to\mathbb{Z} satisfying the following properties:

  1. (1)

    h(λ)h(\lambda) is a non-negative integer for all λX(T)+\lambda\in X(T)_{+};

  2. (2)

    if χ,χX(T)\chi^{\prime},\chi\in X(T) with χ>χ\chi^{\prime}>\chi, then h(χ)>h(χ)h(\chi^{\prime})>h(\chi).

For a commutative GG-algebra AA over kk, we define the 0\mathbb{Z}_{\geq 0}-filtration

FiA:={aA|h(χ)i for all T-weights χ of span𝕜Ga}.F^{i}A:=\{a\in A\,|\,h(\chi)\leq i\mbox{ for all }T\mbox{-weights }\chi\mbox{ of }\operatorname{span}_{\Bbbk}G\cdot a\}.

Denote by grA\operatorname{gr}A the associated graded algebra. Then there is an injective map of GG-algebras

(2.16) grA(AU𝕜𝕜[G/U])T,\operatorname{gr}A\hookrightarrow(A^{U^{-}}\otimes_{\Bbbk}\Bbbk[G/U])^{T},

which is onto if and only if AA has a good filtration [Gro92, Theorem 16].

Consider LL a linear algebraic group, and HLH\subset L a closed subgroup. Let N:=NL(H)N:=N_{L}(H) be the normalizer of HH in LL. Let RR be an LL-algebra. The group NN acts naturally on RHR^{H} and on HH-invariants 𝕜[L]H=𝕜[L/H]\Bbbk[L]^{H}=\Bbbk[L/H] (by right multiplication). The following is a consequence of [Pop86, Theorem 4] (see also [Gro97, Theorem 9.1]).

Lemma 2.17.

There is an isomorphism of NN-algebras RH(R𝕜𝕜[L/H])LR^{H}\cong(R\otimes_{\Bbbk}\Bbbk[L/H])^{L}.

3. Main results

In this section we develop our general results on collapsing of bundles. We work over an algebraically closed field 𝕜\Bbbk of arbitrary characteristic (see Remark 3.11). In the special case when char𝕜=0\operatorname{char}\Bbbk=0 and the Schubert variety considered is the flag variety itself, the general framework agrees with that of completely reducible bundles as in [Kem76].

We fix the notation that is used throughout the section. Consider a parabolic subgroup PGP\subset G. Without loss of generality, we assume that PP is standard corresponding to a set of simple roots ISI\subset S. Let UIU_{I} be the unipotent radical of PP. Let P=LUIP=L\ltimes U_{I} be the Levi decomposition, with L:=LIL:=L_{I} reductive. We denote by PP^{-} the opposite parabolic subgroup, having decomposition P=LUIP^{-}=L\ltimes U_{I}^{-}.

Let WW be a finite-dimensional GG-module. We introduce the map of LL-modules

(3.1) ψ:WUI((W)UI),\psi\colon W^{U_{I}}\longrightarrow\left((W^{*})^{U^{-}_{I}}\right)^{*},

which is the dual of the composition (W)UIW(WUI)(W^{*})^{U^{-}_{I}}\hookrightarrow W^{*}\twoheadrightarrow(W^{U_{I}})^{*}.

Throughout we take an LL-submodule VWUIV\subset W^{U_{I}} such that the map ψ|V:V((W)UI)\left.\psi\right|_{V}:V\to((W^{*})^{U^{-}_{I}})^{*} is injective. The following shows that tracking the map ψ|V\left.\psi\right|_{V} is relevant only when char𝕜>0\operatorname{char}\Bbbk>0.

Lemma 3.2.

In either of the following cases, ψ|V\left.\psi\right|_{V} is an isomorphism:

  • (a)

    WW is a semi-simple GG-module and V=WUIV=W^{U_{I}}.

  • (b)

    W=i=1nΔG(λi)W=\bigoplus_{i=1}^{n}\Delta_{G}(\lambda_{i}) for some λiX(T)+\lambda_{i}\in X(T)_{+}, and VWUIV\subset W^{U_{I}} is V=i=1nΔL(λi)V=\bigoplus_{i=1}^{n}\Delta_{L}(\lambda_{i}).

Proof.

For part (a), we can assume that WW is a simple GG-module. Both WUIW^{U_{I}} and ((W)UI)((W^{*})^{U^{-}_{I}})^{*} are simple LL-modules [Jan03, Proposition II.2.11], and ψ\psi gives a non-trivial map between their respective highest weight vectors. Therefore, ψ\psi is an isomorphism.

For part (b), we can assume that W=ΔG(λ)W=\Delta_{G}(\lambda) is a Weyl module. The restriction map G(λ)L(λ)\nabla_{G}(\lambda)\to\nabla_{L}(\lambda) induced by P/BG/BP/B\subset G/B is surjective (see Lemma 2.4). Therefore, the LL-submodule of WW generated by its highest weight vector (of weight λ\lambda) is VΔL(λ)V\cong\Delta_{L}(\lambda). On the other hand, we have ((W)UI)ΔL(λ)((W^{*})^{U^{-}_{I}})^{*}\cong\Delta_{L}(\lambda) as LL-modules (see [Don88, Section 1.2]), generated as an LL-module by the highest weight vector. Since on the weight space of λ\lambda the map ψ|V\left.\psi\right|_{V} is easily seen to be non-zero, it is also surjective, hence an isomorphism. ∎

Let XX be a closed LL-stable subvariety of VV. As UIU_{I} acts on VV trivially, XX is PP-stable closed subvariety of WW. We have the following proper collapsing map

(3.3) q:G×PXW,q\colon G\times_{P}X\longrightarrow W,

with imq=GX\operatorname{im}q=G\cdot X a closed subvariety of WW. Let π:G×PXG/P\pi:G\times_{P}X\to G/P be the bundle map. For any closed subset YG/PY\subset G/P, the subvariety q(π1(Y))Wq(\pi^{-1}(Y))\subset W is closed. In the case when Y=X(w)PY=X(w)_{P} is a Schubert variety, then q(π1(Y))=BwX¯q(\pi^{-1}(Y))=\overline{BwX} is a BB-stable subvariety in XX.

Proposition 3.4.

For any w𝒲Iw\in\mathcal{W}^{I}, the restriction map 𝕜[BwX¯]𝕜[wX]\Bbbk[\overline{BwX}]\to\Bbbk[wX] induces an isomorphism of algebras

𝕜[BwX¯]U(w)𝕜[wX].\Bbbk[\overline{BwX}]^{U(w)}\,\xrightarrow{\,\,\cong\,\,}\,\Bbbk[wX].

Thus, the algebra 𝕜[X]\Bbbk[X] is a direct summand of 𝕜[BwX¯]\Bbbk[\overline{BwX}] as a 𝕜[X]\Bbbk[X]-module via 𝕜[X]𝕜[wX]\Bbbk[X]\cong\Bbbk[wX].

Proof.

The inclusions wXBwX¯WwX\subset\overline{BwX}\subset W give rise to a commutative diagram

𝕜[BwX¯]U(w)\textstyle{\Bbbk[\overline{BwX}]^{U(w)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{\quad f}𝕜[wX]\textstyle{\Bbbk[wX]}𝕜[W]U(w)\textstyle{\Bbbk[W]^{U(w)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{\,\,g}𝕜[wV]\textstyle{\Bbbk[wV]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

To show that ff is onto, it is enough to show that gg is so. For this, we show that the map (W)U(w)wV(W^{*})^{U(w)}\to w\cdot V^{*} is onto. As w1(W)U(w)=(W)w1UwUw^{-1}\cdot(W^{*})^{U(w)}=(W^{*})^{w^{-1}Uw\,\cap\,U^{-}} and w1UwUUIw^{-1}Uw\,\cap\,U^{-}\subset U_{I}^{-} (cf. (2.1)), this follows since the LL-module map ψ|V:(W)UIV\left.\psi\right|_{V}^{*}:(W^{*})^{U^{-}_{I}}\to V^{*} is onto.

The morphism (3.3) induces an injective map of algebras

𝕜[BwX¯](𝕜[BwP¯]𝕜[X])P.\Bbbk[\overline{BwX}]\hookrightarrow(\Bbbk[\overline{BwP}]\otimes\Bbbk[X])^{P}.

The multiplication map (2.2) gives an open immersion into BwP¯\overline{BwP}, inducing an injective map 𝕜[BwP¯]U(w)𝕜[wP]\Bbbk[\overline{BwP}]^{U(w)}\hookrightarrow\Bbbk[wP]. The previous maps give

𝕜[BwX¯]U(w)(𝕜[BwP¯]U(w)𝕜[X])P(𝕜[wP]𝕜[X])P𝕜[wX],\Bbbk[\overline{BwX}]^{U(w)}\hookrightarrow(\Bbbk[\overline{BwP}]^{U(w)}\!\!\otimes\Bbbk[X])^{P}\hookrightarrow(\Bbbk[wP]\otimes\Bbbk[X])^{P}\cong\Bbbk[wX],

thus proving the injectivity of ff. ∎

Remark 3.5.

Putting w=w0wI1w=w_{0}w_{I}^{-1} in Proposition 3.4, and twisting by ww we obtain an isomorphism of LL-algebras 𝕜[GX]UI𝕜[X]\Bbbk[G\cdot X]^{U_{I}^{-}}\xrightarrow{\cong}\Bbbk[X]. ∎

3.1. Good saturations

The following is our main tool for inducing the property of being good via saturations.

Theorem 3.6.
  • (a)

    The GG-variety GXG\cdot X is good if and only if the LL-variety XX is good and the induced map k[W]q𝒪G×PXk[W]\to q_{*}\mathcal{O}_{G\times_{P}X} is onto.

  • (b)

    Assume that (V,X)(V,X) is a good pair of LL-varieties and ψ|V\left.\psi\right|_{V} is a split map of LL-modules. If there is a good closed GG-subvariety ZWZ\subset W with GXZG\cdot X\subset Z, then (Z,GX)(\,Z\,,\,\,G\cdot X) is a good pair.

  • (c)

    Let YVY\subset V be a closed LL-stable subvariety such that (Y,X)(Y,X) is a good pair and GYG\cdot Y is good. Then (GY,GX)(G\cdot Y\,,\,G\cdot X) is a good pair.

Proof.

Assume that GXG\cdot X is good. By Remark 3.5 and [Don88], we obtain that XX is good. From the proof of Proposition 3.4 we have 𝕜[GX]UI(q𝒪G×PX)UI\Bbbk[G\cdot X]^{U^{-}_{I}}\xrightarrow{\,\cong\,}(q_{*}\mathcal{O}_{G\times_{P}X})^{U^{-}_{I}}, which yields isomorphisms

(3.7) 𝕜[GX]U(q𝒪G×PX)U𝕜[X]UL.\Bbbk[G\cdot X]^{U^{-}}\xrightarrow{\,\,\cong\,\,}(q_{*}\mathcal{O}_{G\times_{P}X})^{U^{-}}\xrightarrow{\,\,\cong\,\,}\Bbbk[X]^{U_{L}^{-}}.

Therefore, the map 𝕜[GX]q𝒪G×PX\Bbbk[G\cdot X]\to q_{*}\mathcal{O}_{G\times_{P}X} is onto by Lemma 2.11.

Now assume that XX is good and 𝕜[GX]q𝒪G×PX\Bbbk[G\cdot X]\to q_{*}\mathcal{O}_{G\times_{P}X} is onto (hence, an isomorphism). By [Don88, Theorem 3], Proposition 2.9 and [Don90, Proposition 1.2e (iii)] the GG-module q𝒪G×PX=(𝕜[G/UI]𝕜[X])Lq_{*}\mathcal{O}_{G\times_{P}X}=(\Bbbk[G/U_{I}]\otimes\Bbbk[X])^{L} has a good filtration, thus GXG\cdot X is good.

For part (b), by Corollary 2.12 the claim is equivalent to the map 𝕜[Z]U𝕜[GX]U\Bbbk[Z]^{U^{-}}\to\Bbbk[G\cdot X]^{U^{-}} being onto. By Proposition 3.4, it is enough to show that the map 𝕜[W]U𝕜[X]UL\Bbbk[W]^{U^{-}}\to\Bbbk[X]^{U_{L}^{-}} is onto. By Corollary 2.12, the map 𝕜[V]UL𝕜[X]UL\Bbbk[V]^{U_{L}^{-}}\to\Bbbk[X]^{U_{L}^{-}} is onto. Hence, the claim follows if we show that the map 𝕜[W]U𝕜[V]UL\Bbbk[W]^{U^{-}}\to\Bbbk[V]^{U_{L}^{-}} is onto. For this, we prove that the restriction of the latter map to the subalgebra (Sym((W)UI))UL(\operatorname{Sym}((W^{*})^{U_{I}^{-}}))^{U_{L}^{-}} is already onto.

Since the LL-map ψ|V\left.\psi\right|_{V} is split, then so is Sym((W)UI)Sym(V)\operatorname{Sym}((W^{*})^{U_{I}^{-}})\to\operatorname{Sym}(V^{*}). Therefore, taking ULU_{L}^{-}-invariants yields a surjective map.

Now we consider part (c). By Corollary 2.12 it is enough to see that the morphism 𝕜[GY]U𝕜[GX]U\Bbbk[G\cdot Y]^{U^{-}}\to\Bbbk[G\cdot X]^{U^{-}} is surjective. By Proposition 3.4, this is equivalent to showing that 𝕜[Y]UL𝕜[X]UL\Bbbk[Y]^{U^{-}_{L}}\to\Bbbk[X]^{U^{-}_{L}} is onto. This follows again by Corollary 2.12. ∎

Remark 3.8.

Assume VV is good and put η=𝒱(V)\eta=\mathcal{V}(V^{*}) and ξ=𝒱(W)/η\xi=\mathcal{V}(W^{*})/\eta. Then:

  • (a)

    GVG\cdot V is good if and only if Hi(G/P,iξ)=0H^{i}(G/P,\,\bigwedge^{i}\xi)=0, for all i>0i>0, by Theorem 3.6 (a), [Wey03, Theorem 5.1.2] and Remark 3.12 below.

  • (b)

    Assume further that WW has Weyl filtration and the LL-map ψ|V\left.\psi\right|_{V} is a split. Then using Lemma 2.11 we see as in the proof above that the induced map WH0(G/P,η)W^{*}\to H^{0}(G/P,\,\eta) is onto. Hence, by Theorem 3.6 (a), GVG\cdot V is good if and only if the algebra q𝒪G×PVH0(G/P,Symη)q_{*}\mathcal{O}_{G\times_{P}V}\cong H^{0}(G/P,\,\operatorname{Sym}\eta) is generated by H0(G/P,η)H^{0}(G/P,\,\eta).

Corollary 3.9.

If char𝕜>dimW\operatorname{char}\Bbbk>\dim W and char𝕜χ+ρ,α\operatorname{char}\Bbbk\geq\langle\chi+\rho,\alpha^{\vee}\rangle for all weights χ\chi of WW and all αΦ+\alpha\in\Phi_{+}, then (W,GV)(\,W\,,\,G\cdot V\,) is a good pair.

Proof.

By Lemma 2.10 parts (1)–(3), we see that both VV and WW are good. By [Jan03, Section 5.6], both VV and WW are semi-simple, therefore ψ|V\left.\psi\right|_{V} is split injective (see Lemma 3.2). The conclusion now follows from Theorem 3.6 (b). ∎

If WW is as in (1.1), then putting X=VX=V and Z=WZ=W in Theorem 3.6 (b), we see that (W,GV)(W,\,G\cdot V) is a good pair whenever char𝕜>max{dimΔG(λi)| 1in}\operatorname{char}\Bbbk>\max\{\dim\Delta_{G}(\lambda_{i})\,|\,1\leq i\leq n\} by Proposition 2.9 and Lemma 2.10. In particular, GVG\cdot V is then good as claimed in the Introduction.

3.2. Singularities via Schubert collapsing

Now we turn to Theorems 1.2 and 1.4. The following result describes the behavior of singularities under collapsing, and it strengthens [Kem76, Proposition 1 and Theorem 3] when w=w0wI1w=w_{0}w_{I}^{-1} (i.e. when BwX¯=GX\overline{BwX}=G\cdot X) in the characteristic zero case as well.

Theorem 3.10.

Assume that GXG\cdot X is good. For w𝒲Iw\in\mathcal{W}^{I}, the BB-variety BwX¯\overline{BwX} is wwIww_{I}-excellent. Furthermore, the following statements hold:

  1. (1)

    The map 𝒪BwX¯𝐑q𝒪BwP¯×PX\mathcal{O}_{\overline{BwX}}\,\xrightarrow{\,\cong\,}\,\mathbf{R}q_{*}\mathcal{O}_{\overline{BwP}\times_{P}X} is an isomorphism.

  2. (2)

    BwX¯\overline{BwX} is normal if and only if XX is so.

  3. (3)

    If char𝕜=0\operatorname{char}\Bbbk=0, then BwX¯\overline{BwX} has rational singularities if and only if so does XX.

  4. (4)

    If XX is an LL-submodule of VV, then GXG\cdot X is strongly FF-regular (resp. of strongly FF-regular type) when char𝕜>0\operatorname{char}\Bbbk>0 (resp. when char𝕜=0\operatorname{char}\Bbbk=0), and BwX¯\overline{BwX} is FF-rational when char𝕜>0\operatorname{char}\Bbbk>0.

Proof.

For part (1), observe that by (3.7) a good filtration of 𝕜[X]\Bbbk[X] has composition factors ΔL(λ)\Delta_{L}(\lambda) with such that λX(T)+\lambda\in X(T)_{+}. By Lemma 2.4, we obtain by induction on filtration that 𝐑iq𝒪BwP¯×PX=0\mathbf{R}^{i}q_{*}\mathcal{O}_{\overline{BwP}\times_{P}X}=0, for all i>0i>0. The map 𝒪BwX¯q𝒪BwP¯×PX\mathcal{O}_{\overline{BwX}}\to q_{*}\mathcal{O}_{\overline{BwP}\times_{P}X} is an isomorphism, since the composition 𝕜[W]q𝒪G×PXq𝒪BwP¯×PX\Bbbk[W]\to q_{*}\mathcal{O}_{G\times_{P}X}\to q_{*}\mathcal{O}_{\overline{BwP}\times_{P}X} is surjective by Theorem 3.6 (a) and Lemma 2.4.

For part (2), if BwX¯\overline{BwX} is normal, then by Proposition 3.4 so is XX. Conversely, if XX is normal, then so is BwX¯\overline{BwX} by the normality of X(w)PX(w)_{P} [RR85] and 𝒪BwX¯q𝒪BwP¯×PX\mathcal{O}_{\overline{BwX}}\cong q_{*}\mathcal{O}_{\overline{BwP}\times_{P}X}.

Next, we prove the statements regarding BwX¯\overline{BwX} in part (3) and (4). If BwX¯\overline{BwX} has rational singularities, then due to the direct summand property in Proposition 3.4 so does XX according to [Bou87, Théorème].

Consider the filtration Fi𝕜[X]F^{i}\Bbbk[X] as in Section 2.5. This gives an exhaustive filtration on A:=𝕜[BwX¯]A:=\Bbbk[\overline{BwX}] by FiA:=(𝕜[BwP¯]Fi𝕜[X])PF^{i}A:=(\Bbbk[\overline{BwP}]\otimes F^{i}\Bbbk[X])^{P}. The associated graded is

grA=(𝕜[BwP¯]UIgr𝕜[X])L(2.16)(𝕜[BwP¯]UI(𝕜[L/UL]𝕜[X]UL)T)L\operatorname{gr}A=(\Bbbk[\overline{BwP}]^{U_{I}}\otimes\operatorname{gr}\Bbbk[X])^{L}\,\stackrel{{\scriptstyle(\ref{eq:grA})}}{{\cong}}\,(\Bbbk[\overline{BwP}]^{U_{I}}\otimes(\Bbbk[L/U_{L}]\otimes\Bbbk[X]^{U^{-}_{L}})^{T})^{L}\cong
((𝕜[BwP¯]UI𝕜[L/UL])L𝕜[X]UL)T(𝕜[BwP¯]U𝕜[X]UL)T=(𝕜[BwwIB¯]+U𝕜[X]UL)T,\cong((\Bbbk[\overline{BwP}]^{U_{I}}\otimes\Bbbk[L/U_{L}])^{L}\otimes\Bbbk[X]^{U^{-}_{L}})^{T}\cong(\Bbbk[\overline{BwP}]^{U}\otimes\Bbbk[X]^{U^{-}_{L}})^{T}=(\Bbbk[\overline{Bww_{I}B}]^{U}_{+}\otimes\Bbbk[X]^{U^{-}_{L}})^{T},

where the last equality is a consequence of BwP¯=BwwIB¯\overline{BwP}=\overline{Bww_{I}B} and (3.7), and the isomorphism before it follows from Lemma 2.17.

Now assume that XX has rational singularities when char𝕜=0\operatorname{char}\Bbbk=0 (resp. XX is an LL-module when char𝕜>0\operatorname{char}\Bbbk>0). By [Pop86, Theorem 6] (resp. by [Has12, Corollary 4.14]), 𝕜[X]UL\Bbbk[X]^{U^{-}_{L}} has rational singularities (resp. is strongly FF-regular). By Lemma 2.6 and (2.5), 𝕜[BwwIB¯]+U\Bbbk[\overline{Bww_{I}B}]^{U}_{+} has rational singularities (resp. is strongly FF-regular). Hence, grA\operatorname{gr}A has rational singularities (resp. is strongly FF-regular) by [Bou87] (resp. [HH94a, Theorem 5.5]). As in [Pop86, Section 5], the algebra grA\operatorname{gr}A is a flat deformation of AA. Therefore, AA has rational singularities by [Elk78] (resp. is FF-rational by (2.5) and [HH94a, Theorem 4.2]).

Now we show that GXG\cdot X is strongly FF-regular in part (4). Let G=G~×ZG^{\prime}=\tilde{G}\times Z, with G~\tilde{G} a covering of [G,G][G,G] and ZTZ\subset T a torus so that GG is a quotient of GG^{\prime}. We can view WW as a GG^{\prime}-representation. Since TLT\subset L, we have GX=G~XG\cdot X=\tilde{G}\cdot X. Moreover, we can lift PP to a parabolic PP^{\prime} of G~\tilde{G} with unipotent radical UIU^{\prime}_{I} and Levi subgroup LL^{\prime}. We have WUI=WUIW^{U^{\prime}_{I}}=W^{U_{I}} and (W)UI=(W)UI(W^{*})^{U^{\prime-}_{I}}=(W^{*})^{U^{-}_{I}}. Furthermore, GXG\cdot X (resp. XX) is GG-good (resp. LL-good) if and only if it is G~\tilde{G}-good (resp. LL^{\prime}-good) [Don85, Section 3]. This shows that we can assume that GG is simply connected and semisimple.

Assume that char𝕜>0\operatorname{char}\Bbbk>0. Since XX and GG are good, using [Don88, Theorem 3] and Proposition 2.9 we have

q𝒪G×PX=(𝕜[G/UI]𝕜[X])L=((𝕜[G/UI]𝕜[X])UL)T.q_{*}\mathcal{O}_{G\times_{P}X}=(\Bbbk[G/U_{I}]\otimes\Bbbk[X])^{L}=\left((\Bbbk[G/U_{I}]\otimes\Bbbk[X])^{U_{L}}\right)^{T}.

As TT is linearly reductive, by [HH94a, Theorem 5.5] the claim follows once we show that R:=(𝕜[G/UI]𝕜[X])ULR:=(\Bbbk[G/U_{I}]\otimes\Bbbk[X])^{U_{L}} is strongly FF-regular. Since 𝕜[X]\Bbbk[X] and 𝕜[G]\Bbbk[G] are factorial rings (see [Pop74]), so is RR and 𝕜[G]U×UI\Bbbk[G]^{U\times U_{I}} (see [VP89, Theorem 3.17]). In particular, since 𝕜[G]U×UI\Bbbk[G]^{U\times U_{I}} is Cohen–Macaulay by Corollary 2.8 and (2.5), it is Gorenstein [Mur64].

We have an action of GG on RR induced from its left action on 𝕜[G]\Bbbk[G]. We have an isomorphism R(𝕜[L/UL]𝕜[G/UI]𝕜[X])LR\cong(\Bbbk[L/U_{L}]\otimes\Bbbk[G/U_{I}]\otimes\Bbbk[X])^{L}, which is easily seen to be GG-equivariant. The algebra 𝕜[L/UL]𝕜[G/UI]𝕜[X]\Bbbk[L/U_{L}]\otimes\Bbbk[G/U_{I}]\otimes\Bbbk[X] has a good filtration as a G×LG\times L-module, as seen using [Don88, Theorem 3] and Proposition 2.9. By [Don90, Proposition 1.2e (iii)], we obtain that RR has a good filtration as a GG-module. We consider the invariant ring RUR^{U}. By Corollary 2.8, [Has03, Theorem 5.2] and [Has12, Theorem 4.4 and Lemma 4.7], the 0\mathbb{Z}_{\geq 0}-graded ring 𝕜[G]U×UI𝕜[X]\Bbbk[G]^{U\times U_{I}}\otimes\Bbbk[X] is Gorenstein, strongly FF-regular, and LL-FF-pure. Then [Has12, Corollary 4.13] implies that RUR^{U} is strongly FF-regular. Using the filtration in Section 2.5, this implies that RR is FF-rational by (2.5) and [HH94a, Theorem 4.2] (see also [Has12, Corollary 3.9]). Since RR is factorial and Cohen–Macaulay, it is also Gorenstein [Mur64]. This shows that RR is strongly FF-regular (see [HH94a, Corollary 4.7] or [HH94b]).

Now let char𝕜=0\operatorname{char}\Bbbk=0. We can choose a suitable large set of primes SS such that for D=[S1]D=\mathbb{Z}[S^{-1}] we have: the map G×WWG\times W\to W (resp. G×XGXG\times X\to G\cdot X) is defined over DD; GX=(GDXD)×SpecDSpec(𝕜)G\cdot X=(G_{D}\cdot X_{D})\times_{\operatorname{Spec}D}\operatorname{Spec}(\Bbbk); the affine scheme (GX)D=GDXD(G\cdot X)_{D}=G_{D}\cdot X_{D} is flat over DD; both W𝔽¯pW_{\overline{\mathbb{F}}_{p}} and (GX)𝔽¯p(G\cdot X)_{\overline{\mathbb{F}}_{p}} are good for pSp\notin S (see Corollary 3.9); W𝔽¯pW_{\overline{\mathbb{F}}_{p}} (resp. X𝔽¯pX_{\overline{\mathbb{F}}_{p}}) is a semi-simple G𝔽¯pG_{\overline{\mathbb{F}}_{p}}-module (resp. L𝔽¯pL_{\overline{\mathbb{F}}_{p}}-module) (see [Jan03, Section II.5.6]). For such pSp\notin S, for V=X𝔽¯pV=X_{\overline{\mathbb{F}}_{p}} the map ψ|V\left.\psi\right|_{V} in (3.1) is injective (see Lemma 3.2). By the previous paragraph and [HH94a, Theorem 5.5], we obtain that (GX)𝔽p(G\cdot X)_{\mathbb{F}_{p}} is strongly FF-regular. Hence, GXG\cdot X is of strongly FF-regular type. ∎

Remark 3.11.

As seen in the proof above, the assumption on the field to be algebraically closed is not essential. The claims about rational singularities and strongly FF-regular type (resp. FF-rational singularities) hold over any field, e.g. by [Bou87] (resp. proof of [Smi97, Lemma 1.4]), as do claims (1) and (2). The claim on strong FF-regularity holds for any FF-finite (e.g. perfect) field [HH94a, Theorem 5.5]. ∎

Remark 3.12.

Even if XX is good, it may happen that GXG\cdot X is not, as can be seen in Example 4.4. Nevertheless, we still have 𝐑iq𝒪BwP¯×PX=0\mathbf{R}^{i}q_{*}\mathcal{O}_{\overline{BwP}\times_{P}X}=0 for i>0i>0. Further if XX is good, normal, and q:BwP¯×PXBwX¯q:\overline{BwP}\times_{P}X\to\overline{BwX} is birational (or, more generally, the generic fiber of qq is connected and qq is separable, as in [LW19, Theorem 2.1 (a)]), then the results in Theorem 3.10 carry over if we replace the variety BwX¯\overline{BwX} in each statement (besides part (2)) with its normalization, which is then in turn a wwIww_{I}-excellent variety. ∎

We further note that if one knows a good filtration of 𝕜[X]\Bbbk[X] explicitly, then by Theorem 3.10 one obtains readily a corresponding wwIww_{I}-excellent filtration for 𝕜[BwX¯]\Bbbk[\overline{BwX}]. It is then possible to compute the (TT-equivariant) Hilbert function for 𝕜[BwX¯]\Bbbk[\overline{BwX}] using Lemma 2.4 and the Demazure character formula (e.g. [BK05, Corollary 3.3.11]).

By Proposition 3.4 and [HH94a, Theorem 5.5] if BwX¯\overline{BwX} is strongly FF-regular (when char𝕜>0\operatorname{char}\Bbbk>0), then XX must also be strongly FF-regular. In the case of a Borel subgroup, we can strengthen Theorem 3.10 by giving the following converse to this statement.

Corollary 3.13.

Assume that P=BP=B is a Borel subgroup and WW has a Weyl filtration. Then GXG\cdot X is good. Moreover, for w𝒲w\in\mathcal{W}, the variety BwX¯\overline{BwX} is strongly FF-regular (resp. of strongly FF-regular type) when char𝕜>0\operatorname{char}\Bbbk>0 (resp. when char𝕜=0\operatorname{char}\Bbbk=0) if and only if so is XX.

Proof.

We can assume that P=BP=B. Since TT is linearly reductive, (V,X)(V,X) is a good pair. By Theorem 3.6 (c), in order to show that GXG\cdot X is good it is enough to show that GVG\cdot V is so. For this, we use Theorem 3.6(a). Since VWUV\subset W^{U}, we have a TT-decomposition V=i=1n𝕜λiV=\bigoplus_{i=1}^{n}\Bbbk_{\lambda_{i}}, where λiX(T)+\lambda_{i}\in X(T)_{+}. The section ring

q𝒪G×BV=(mi)nH0((i=1nmiλi))q_{*}\mathcal{O}_{G\times_{B}V}=\bigoplus_{(m_{i})\in\mathbb{N}^{n}}H^{0}(\mathcal{L}(\sum_{i=1}^{n}m_{i}\lambda_{i}))

is generated in the components of the unit tuples, i.e. by the sum i=1nG(λi)\bigoplus_{i=1}^{n}\nabla_{G}(\lambda_{i}), as it follows from [RR85] (see also [KR87]). By Remark 3.8 (b), GVG\cdot V is good.

Assume that XX is strongly FF-regular. Note that both 𝕜[BwB¯]+U\Bbbk[\overline{BwB}]^{U}_{+} and 𝕜[X]\Bbbk[X] are X(T)+X(T)_{+}-graded algebras, so also 0\mathbb{Z}_{\geq 0}-graded, using for instance the map hh in Section 2.5. Then the algebra q𝒪BwB¯×BX=(𝕜[BwB¯]U𝕜[X])T=(𝕜[BwB¯]+U𝕜[X])Tq_{*}\mathcal{O}_{\overline{BwB}\times_{B}X}=(\Bbbk[\overline{BwB}]^{U}\otimes\Bbbk[X])^{T}=(\Bbbk[\overline{BwB}]^{U}_{+}\otimes\Bbbk[X])^{T} is strongly FF-regular, as it follows by combining Lemma 2.6, [Has03, Theorem 5.2] and [HH94a, Theorem 5.5]. Since GXG\cdot X is good, the conclusion follows from Theorem 3.10 (1).

Now let char𝕜=0\operatorname{char}\Bbbk=0. Assume XX is of strongly FF-regular type, and consider a finitely generated \mathbb{Z}-algebra RkR\subset k as in the definition in Section 2.3 (enlarging, if necessary, so that the action of TRT_{R} is well-defined). Let (BwX¯)R=Spec((C(X(w)R)R[XR])TR)(\overline{BwX})_{R}=\operatorname{Spec}((C(X(w)_{R})\otimes R[X_{R}])^{T_{R}}). As in the proof of Lemma 2.6, (BwX¯)R(\overline{BwX})_{R} is flat of finite type over RR, and (C(X(w)R)R[XR])TRR𝕜(C(X(w)𝕜)𝕜[X𝕜])T𝕜(C(X(w)_{R})\otimes R[X_{R}])^{T_{R}}\otimes_{R}\Bbbk^{\prime}\cong(C(X(w)_{\Bbbk^{\prime}})\otimes\Bbbk^{\prime}[X_{\Bbbk^{\prime}}])^{T_{\Bbbk^{\prime}}}, for any field 𝕜\Bbbk^{\prime} over RR (see [Jan03, Section I.2.11]). By Theorem 3.10 (1), we have (BwX¯)R×Spec(R)Spec(𝕜)BwX¯(\overline{BwX})_{R}\times_{\operatorname{Spec}(R)}\operatorname{Spec}(\Bbbk)\cong\overline{BwX}. When 𝕜\Bbbk^{\prime} is a residue field of RR, it is finite, in which case C(X(w)𝕜)C(X(w)_{\Bbbk^{\prime}}) is strongly FF-regular, as seen in the proof of Lemma 2.6. As in the previous paragraph, we conclude that (BwX¯)R/𝔪(\overline{BwX})_{R/\mathfrak{m}} is strongly FF-regular for maximal ideals 𝔪\mathfrak{m} in a dense open subset of Spec(R)\operatorname{Spec}(R).

Finally, if BwX¯\overline{BwX} is of strongly FF-regular type, using Proposition 3.4 we see by an argument similar to the above that XX is also of strongly FF-regular type. ∎

Further, we provide a result that can lead to more general varieties outside the equivariant setting. Following [Bri03], we call a closed subvariety YG/PY\subset G/P multiplicity-free if it is rationally equivalent to a multiplicity-free linear combination of Schubert cycles.

Corollary 3.14.

Let YY be a multiplicity-free subvariety of G/PG/P, and assume that GXG\cdot X is good. Then 𝒪q(π1(Y))𝐑q𝒪π1(Y)\mathcal{O}_{q(\pi^{-1}(Y))}\,\xrightarrow{\,\cong\,}\,\mathbf{R}q_{*}\mathcal{O}_{\pi^{-1}(Y)} is an isomorphism. Moreover, if XX is normal (resp. has rational singularities when char𝕜=0\operatorname{char}\Bbbk=0), then q(π1(Y))q(\pi^{-1}(Y)) is normal (resp. has rational singularities).

Proof.

The proof of the isomorphism 𝒪q(π1(Y))𝐑q𝒪π1(Y)\mathcal{O}_{q(\pi^{-1}(Y))}\,\xrightarrow{\,\cong\,}\,\mathbf{R}q_{*}\mathcal{O}_{\pi^{-1}(Y)} follows as in Theorem 3.10 (a) using [Bri03, Theorem 0.1] and Lemma 2.4. The claim on normality follows from this, as YY itself is normal [Bri03, Theorem 0.1]. Moreover, YY has rational singularities when char𝕜=0\operatorname{char}\Bbbk=0 [Bri03, Theorem 0.1 and Remark 3.3], hence we conclude that so does q(π1(Y))q(\pi^{-1}(Y)) by [Kov00, Theorem 1]. ∎

3.3. Defining equations of saturations

In this section we give a result on the defining equations of GXG\cdot X in WW. Assume that GVG\cdot V is good. Let M𝕜[V]M\subset\Bbbk[V] be an LL-stable module with a good filtration. We can associate to it a GG-module M𝕜[GV]M^{\prime}\subset\Bbbk[G\cdot V] in the following way. Consider the inclusion of sheaves 𝒱(M)𝒱(SymV)\mathcal{V}(M)\subset\mathcal{V}(\operatorname{Sym}V^{*}) on G/PG/P. Then we put M=H0(G/P,𝒱(M))M^{\prime}=H^{0}(G/P,\mathcal{V}(M)). As in the proof of Theorem 3.10 (1), we see that MM^{\prime} has a good filtration as a GG-module. Note that MM^{\prime} contains span𝕜GM\operatorname{span}_{\Bbbk}G\cdot M via the inclusion given by Remark 3.5, and this containment is an equality when MM^{\prime} is a semi-simple GG-module.

Theorem 3.15.

Let (V,X)(V,X) be a good pair with GVG\cdot V good, and denote by IX𝕜[V]I_{X}\subset\Bbbk[V] the defining ideal of XVX\subset V. Let MM be the span of a set of good generators of IXI_{X} and take a basis 𝒫\mathcal{P}^{\prime} of the GG-module M𝕜[GV]M^{\prime}\subset\Bbbk[G\cdot V] associated to MM as above. Consider the following:

  1. (1)

    A set of generators 𝒫GV\mathcal{P}_{G\cdot V} of the defining ideal IGV𝕜[W]I_{G\cdot V}\subset\Bbbk[W] of GVG\cdot V;

  2. (2)

    A lift 𝒫~𝕜[W]\tilde{\mathcal{P}^{\prime}}\subset\Bbbk[W] of the set 𝒫𝕜[W]/IGV\mathcal{P}^{\prime}\subset\,\Bbbk[W]/I_{G\cdot V}.

Then the defining ideal of GXG\cdot X in 𝕜[W]\Bbbk[W] is generated by 𝒫:=𝒫GV𝒫~\mathcal{P}:=\,\mathcal{P}_{G\cdot V}\,\cup\,\tilde{\mathcal{P}^{\prime}}.

Furthermore, assume that (W,GV)(W,\,G\cdot V) is a good pair. If either MM^{\prime} is a tilting module, or there are no dominant weights λ>μ\lambda>\mu such that (M)λU0(IGV)μU(M^{\prime})^{U}_{\lambda}\neq 0\neq(I_{G\cdot V})^{U}_{\mu} , then the lift 𝒫~\tilde{\mathcal{P}^{\prime}} can be chosen such that span𝕜𝒫~𝕜[W]\operatorname{span}_{\Bbbk}\tilde{\mathcal{P}^{\prime}}\,\subset\Bbbk[W] is GG-stable; with such lift, if 𝒫GV\mathcal{P}_{G\cdot V} are good generators of IGVI_{G\cdot V} then 𝒫\mathcal{P} is a set of good defining equations of GXWG\cdot X\,\subset W.

Proof.

Let J𝕜[GV]J\subset\Bbbk[G\cdot V] denote the defining ideal of GXG\cdot X in GVG\cdot V. We have an exact sequence

0J𝕜[GV]𝕜[GX]0.0\to J\to\Bbbk[G\cdot V]\to\Bbbk[G\cdot X]\to 0.

By Remark 3.5, taking UIU^{-}_{I}-invariants in the sequence above we get that JUIIXJ^{U^{-}_{I}}\cong I_{X}. Furthermore, by construction MJM^{\prime}\subset J and MMUIM\subset M^{\prime U^{-}_{I}}. Consider the multiplication map

m𝒫:𝕜[GV]MJ.m_{\mathcal{P}^{\prime}}:\Bbbk[G\cdot V]\otimes M^{\prime}\to J.

By Lemma 2.11 and Proposition 2.9, to see that m𝒫m_{\mathcal{P}^{\prime}} is surjective, it is enough to show that the induced map on UU^{-}-invariants is so. This is a consequence of the fact that the following composition of maps is surjective by the assumption on good generators of IXI_{X}:

(3.16) (𝕜[V]M)UL(𝕜[GV]UIMUI)UL(𝕜[GV]M)UJU=IXUL.(\Bbbk[V]\otimes M)^{U_{L}^{-}}\hookrightarrow(\Bbbk[G\cdot V]^{U^{-}_{I}}\otimes M^{\prime U^{-}_{I}})^{U_{L}^{-}}\hookrightarrow(\Bbbk[G\cdot V]\otimes M^{\prime})^{U^{-}}\to J^{U^{-}}=I_{X}^{U^{-}_{L}}.

As 𝒫\mathcal{P}^{\prime} generates J=IGX/IGVJ=I_{G\cdot X}/I_{G\cdot V}, it is clear that 𝒫\mathcal{P} generates IGXI_{G\cdot X}.

Let NN be the GG-submodule NIGXN\subset I_{G\cdot X} corresponding to MJM^{\prime}\subset J. We have an exact sequence

0IGVNM0.0\to I_{G\cdot V}\to N\to M^{\prime}\to 0.

To show that 𝒫~\tilde{\mathcal{P}^{\prime}} can be chosen in the required way, we show that the sequence splits as ExtG1(M,IGV)=0\operatorname{Ext}^{1}_{G}(M^{\prime},I_{G\cdot V})=0. When MM^{\prime} is tilting, this is a consequence of [Jan03, Proposition II.4.13], as IGVI_{G\cdot V} has a good filtration and MM^{\prime} has a Weyl filtration. The other case is a consequence of [Fri85, Proposition 2].

By the splitting above, we have M𝒫~=span𝒫~MM_{\tilde{\mathcal{P}^{\prime}}}=\operatorname{span}\tilde{\mathcal{P}^{\prime}}\cong M^{\prime} as GG-modules. It has a good filtration, as the module M𝒫GVM_{\mathcal{P}_{G\cdot V}}, since 𝒫GV\mathcal{P}_{G\cdot V} is a set of good generators. Therefore, M𝒫=M𝒫~M𝒫GVM_{\mathcal{P}}=M_{\tilde{\mathcal{P}^{\prime}}}\oplus M_{\mathcal{P}_{G\cdot V}} has a good filtration [Don85, Corollary 3.2.5]. Consider the commutative diagram

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(k[W]M𝒫GV)U\textstyle{(k[W]\otimes M_{\mathcal{P}_{G\cdot V}})^{U}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(k[W]M𝒫)U\textstyle{(k[W]\otimes M_{\mathcal{P}})^{U}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(k[W]M𝒫~)U\textstyle{(k[W]\otimes M_{\tilde{\mathcal{P}^{\prime}}})^{U}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(IGV)U\textstyle{\quad(I_{G\cdot V})^{U}\quad\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(IGX)U\textstyle{\quad(I_{G\cdot X})^{U}\quad\ignorespaces\ignorespaces\ignorespaces\ignorespaces}JU\textstyle{\quad J^{U}\quad\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

Due to the respective modules having good filtrations by Proposition 2.9, the rows of the diagrams are exact [Don88, Proposition 1.4 and Proposition 2]. Since 𝒫GV\mathcal{P}_{G\cdot V} is a set of good generators, the first vertical map is onto. We are left to show that the third vertical map is onto, or equivalently, that the following composition is surjective (see comment after Lemma 2.14):

(𝕜[W]M𝒫~)U(𝕜[GV]M)UJU.(\Bbbk[W]\otimes M_{\tilde{\mathcal{P}^{\prime}}})^{U^{-}}\to(\Bbbk[G\cdot V]\otimes M^{\prime})^{U^{-}}\to J^{U^{-}}.

The first map is onto since M𝒫~MM_{\tilde{\mathcal{P}^{\prime}}}\xrightarrow{\cong}M^{\prime} and (W,GV)(W,\,G\cdot V) is a good pair. The second map is onto as seen in (3.16). Thus, 𝒫\mathcal{P} is a good generating set of IGXI_{G\cdot X}. ∎

Remark 3.17.

With the assumptions above, one can similarly give defining equations of BwX¯\overline{BwX}, provided we have defining equations of BwV¯\overline{BwV} in 𝕜[GV]\Bbbk[G\cdot V]. ∎

When GVG\cdot V is good, by Theorem 3.10 one can in principle apply [Wey03, Theorem 5.1.3] to obtain a (minimal) set of generators 𝒫GV\mathcal{P}_{G\cdot V} (as seen in Remark 3.8), or even its minimal free resolution. We note that the minimal free resolution of GVG\cdot V given by loc. cit. has length equal to codimGVW\operatorname{codim}_{G\cdot V}W, since GVG\cdot V is Cohen–Macaulay (2.5). For variations of this technique, see for example [Wey03, Section 6] or [KL21, Proposition 4.4].

4. Special cases and applications

This section is devoted to demonstrate the strength of our results through some important applications, both classical and new. The examples in the next three subsections fit into the situation described in the Introduction (1.1).

4.1. Varieties of determinantal type

Let mn0m\geq n\geq 0, and consider the case when WW is the space of m×nm\times n matrices, n×nn\times n skew-symmetric matrices, or n×nn\times n symmetric matrices – the latter can be also identified with the 2nd divided power of 𝕜n\Bbbk^{n}. Then we choose GG to be GL(m)×GL(n)\operatorname{GL}(m)\times\operatorname{GL}(n), GL(n)\operatorname{GL}(n) or GL(n)\operatorname{GL}(n), and W=ΔG(λ)W=\Delta_{G}(\lambda) to be 𝕜m𝕜n\Bbbk^{m}\otimes\Bbbk^{n}, 2𝕜n\bigwedge^{2}\Bbbk^{n}, or ΔG(2ω1)\Delta_{G}(2\omega_{1}), respectively. For 0rn0\leq r\leq n, we put LL to be GL(r)×GL(r)\operatorname{GL}(r)\times\operatorname{GL}(r), GL(r)\operatorname{GL}(r) or GL(r)\operatorname{GL}(r), respectively (and V=ΔL(λ)V=\Delta_{L}(\lambda)). Then GVG\cdot V is precisely the closed subvariety in WW of matrices of rank at most rr (see [Wey03, Section 6]).

The variety WW (resp. VV) is good in arbitrary characteristic (see Lemma 2.10 and [Bof91]). Thus, by Theorem 1.3 (with X=VX=V) the GG-variety GVG\cdot V is good as well. Therefore, by Theorem 3.10 GVG\cdot V is strongly FF-regular when char𝕜>0\operatorname{char}\Bbbk>0 (resp. is of strongly FF-regular type when char𝕜=0\operatorname{char}\Bbbk=0) and BwV¯\overline{BwV} is FF-rational (resp. has rational singularities if char𝕜=0\operatorname{char}\Bbbk=0). This yields all GG-orbit closures GVG\cdot V and many BB-orbit closures BwV¯\overline{BwV} in WW.

For GG-orbit closures in the case of symmetric matrices, this answers [KMN19, Question 5.10]. For GG-orbit closures in 𝕜m𝕜n\Bbbk^{m}\otimes\Bbbk^{n} and 2𝕜n\bigwedge^{2}\Bbbk^{n}, we recover the results [HH94b], [Băe01, Theorem 1.3] (see also [Băe06, Chapter 7]).

The BB-orbit closures are called matrix Schubert varieties in the literature. As far as we are aware, in this case the results are new even in characteristic 0, except in the space of m×nm\times n matrices, when it is known that all matrix Schubert varieties are strongly FF-regular, as this can be reduced to the corresponding statement on Schubert varieties [LRPT06] (see Corollary 3.13) by an identification as done in [Ful92].

Let us show that the (r+1)×(r+1)(r+1)\times(r+1) minors of a generic symmetric matrix give good defining equations for the space of symmetric matrices of rank r\leq r in WW using Theorem 3.15 (the other cases are analogous and slightly easier). We work on downwards induction on rr, the case r=nr=n being trivial. Let VV be the space of r×rr\times r symmetric matrices as above, and consider XVX\subset V the matrices of rank <r<r. Clearly, the symmetric determinant is a good defining equation for XVX\subset V (e.g. Lemma 2.15). The associated GG-module M𝕜[GV]M^{\prime}\subset\Bbbk[G\cdot V] in Theorem 3.15 is M=G(2ωr)M^{\prime}=\nabla_{G}(2\omega_{r}), and it is easy to see that it satisfies the condition that there are no dominant weights μ<2ωr\mu<2\omega_{r} with (IGV)μU0(I_{G\cdot V})^{U}_{\mu}\neq 0. The lift 𝒫~\tilde{\mathcal{P}}^{\prime} can be chosen to be the r×rr\times r minors of a generic symmetric matrix, while 𝒫GV\mathcal{P}_{G\cdot V} are the (r+1)×(r+1)(r+1)\times(r+1) minors, by the induction hypothesis. By Theorem 3.15, we conclude that 𝒫~\tilde{\mathcal{P}}^{\prime} is a good set of defining equations for GXG\cdot X in WW.

4.2. Varieties of complexes on arbitrary quivers

The geometry of the Buchsbaum–Eisenbud varieties of complexes has been investigated thoroughly in a number of articles. In [Kem75] it has been shown that these varieties have rational singularities in characteristic zero, based on the method in [Kem76]. A characteristic-free approach has been pursued in [DCS81] using Hodge algebras, where defining equations are provided as well. In characteristic zero, this result has been proved also in [Bri85] by showing that their algebra of covariants is a polynomial ring. Frobenius splitting methods have been applied in [MT99b]. One can realize such varieties as certain open subsets in Schubert varieties [Zel85], [LM98]. Similar varieties have been studied in [Str82], [Str87], [MT99a] for other special quivers. These varieties can be considered for any quiver, and are particular cases of certain rank varieties of radical square zero algebras, as in explained in [KL21]. In ibid., it is shown that in characteristic zero all such varieties have rational singularities, and defining equations are provided. We explain now how to extend such results to arbitrary characteristic, as announced in Remark 4.16 of ibid. Additionally, we obtain analogous results for BB-varieties.

We follow closely the notation established in [KL21]. Consider the (associative, non-commutative) radical square zero algebra A=𝕜Q/𝕜Q2A=\Bbbk Q/\Bbbk Q_{\geq 2}, with QQ an arbitrary finite quiver with the set of vertices Q0Q_{0} and arrows Q1Q_{1}. For a dimension vector 𝐝:Q00\mathbf{d}:Q_{0}\to\mathbb{Z}_{\geq 0}, we consider the representation space

repQ(𝐝)=αQ1Hom𝕜(𝕜𝐝(tα),𝕜𝐝(hα))=αQ1(𝕜𝐝(tα))𝕜𝐝(hα),\operatorname{rep}_{Q}(\mathbf{d})=\prod_{\alpha\in Q_{1}}\operatorname{Hom}_{\Bbbk}(\Bbbk^{\mathbf{d}(t\alpha)},\Bbbk^{\mathbf{d}(h\alpha)})=\bigoplus_{\alpha\in Q_{1}}(\Bbbk^{\mathbf{d}(t\alpha)})^{*}\otimes\Bbbk^{\mathbf{d}(h\alpha)},

and within the representation variety of AA

repA(𝐝)={MrepQ(𝐝)MβMα=0, for all α,βQ1 with hα=tβ},\operatorname{rep}_{A}(\mathbf{d})=\{M\in\operatorname{rep}_{Q}(\mathbf{d})\,\mid\,M_{\beta}\circ M_{\alpha}=0,\mbox{ for all }\alpha,\beta\in Q_{1}\mbox{ with }h\alpha=t\beta\},

which has a natural action of the reductive group GL(𝐝)=xQ0GL(𝐝(x))\operatorname{GL}(\mathbf{d})=\prod_{x\in Q_{0}}\operatorname{GL}(\mathbf{d}(x)). For xQ0x\in Q_{0} and MrepQ(𝐝)M\in\operatorname{rep}_{Q}(\mathbf{d}), we put

hx(M)=hα=xMα:hα=xMtαMx.h_{x}(M)=\bigoplus_{h\alpha=x}M_{\alpha}\colon\,\bigoplus_{h\alpha=x}M_{t\alpha}\to M_{x}.

For a dimension vector 𝐫𝐝\mathbf{r}\leq\mathbf{d}, we denote by C𝐫C_{\mathbf{r}} the closure of the set of representations MrepA(𝐝)M\in\operatorname{rep}_{A}(\mathbf{d}) such that rankhx(M)=𝐫(x)\operatorname{rank}h_{x}(M)=\mathbf{r}(x), for all xQ0x\in Q_{0}. Let 𝐬=𝐝𝐫\mathbf{s}=\mathbf{d}-\mathbf{r}. By [KL21, Theorem 3.19] the variety C𝐫C_{\mathbf{r}} is irreducible, and it is non-empty if and only if

(4.1) hα=x𝐬(tα)𝐫(x), for all xQ0.\sum_{h\alpha=x}\mathbf{s}(t\alpha)\,\geq\mathbf{r}(x),\mbox{ for all }x\in Q_{0}.

Furthermore, each irreducible component of repA(𝐝)\operatorname{rep}_{A}(\mathbf{d}) is of the form C𝐫C_{\mathbf{r}}, for some 𝐫𝐝\mathbf{r}\leq\mathbf{d}.

Now fix 𝐫𝐝\mathbf{r}\leq\mathbf{d} as in (4.1). With the notation from Section 3, we let W=repQ(𝐝)W=\operatorname{rep}_{Q}(\mathbf{d}), V=αQ1(𝕜𝐬(tα))𝕜𝐫(hα)V=\bigoplus_{\alpha\in Q_{1}}(\Bbbk^{\mathbf{s}(t\alpha)})^{*}\otimes\Bbbk^{\mathbf{r}(h\alpha)}, G=GL(𝐝)G=\operatorname{GL}(\mathbf{d}), L=xQ0(GL(𝐬(x))×GL(𝐫(x)))L=\prod_{x\in Q_{0}}(\operatorname{GL}(\mathbf{s}(x))\times\operatorname{GL}(\mathbf{r}(x))) . It is implicit from the proof of [KL21, Theorem 3.19] that C𝐫=GVC_{\mathbf{r}}=G\cdot V (in fact, the collapsing map q:G×PVC𝐫q:G\times_{P}V\to C_{\mathbf{r}} is a resolution of singularities). The variety WW (resp. VV) is good in arbitrary characteristic by Lemma 2.10 and Proposition 2.9. Thus, by Theorem 1.3 the GG-variety C𝐫C_{\mathbf{r}} is good and Theorem 3.10 implies the following result.

Corollary 4.2.

The rank variety C𝐫C_{\mathbf{r}} is strongly FF-regular when char𝕜>0\operatorname{char}\Bbbk>0 (resp. of strongly FF-regular type when char𝕜=0\operatorname{char}\Bbbk=0).

Moreover, the varieties BwV¯C𝐫\overline{BwV}\subset C_{\mathbf{r}} are FF-rational when char𝕜>0\operatorname{char}\Bbbk>0 (resp. have rational singularities when char𝕜=0\operatorname{char}\Bbbk=0). Note that the Buchsbaum–Eisenbud varieties of complexes are spherical (e.g. [Bri85]), therefore such varieties are always BB-orbit closures in this case as there are only finitely many BB-orbits [Bri86], [Vin86]. We leave the details of the combinatorial characterization of such BB-orbit closures to the interested reader.

In [KL21, Corollary 4.13], explicit defining equations are provided for all C𝐫C_{\mathbf{r}} when char𝕜=0\operatorname{char}\Bbbk=0. We give a self-contained argument to show that, in the case when QQ has no loops, these equations are also defining equations when char𝕜>0\operatorname{char}\Bbbk>0.

For αQ1\alpha\in Q_{1}, we let XαX_{\alpha} be the 𝐝(tα)×𝐝(hα)\mathbf{d}(t\alpha)\times\mathbf{d}(h\alpha) generic matrix of variables. We identify the coordinate ring 𝕜[rep𝕜Q(𝐝)]\Bbbk[\operatorname{rep}_{\Bbbk Q}(\mathbf{d})] with a polynomial ring in the entries of the matrices {Xα}αQ1\{X_{\alpha}\}_{\alpha\in Q_{1}}. For xQ0x\in Q_{0}, we write HxH_{x} (resp. TxT_{x}) for the 𝐝(x)×(hα=x𝐝(tα))\mathbf{d}(x)\times\left(\displaystyle\sum_{h\alpha=x}\mathbf{d}(t\alpha)\right) matrix (resp. (tα=x𝐝(hα))×𝐝(x)\left(\displaystyle\sum_{t\alpha=x}\mathbf{d}(h\alpha)\right)\times\mathbf{d}(x) matrix) obtained by placing the matrices XαX_{\alpha} with hα=xh\alpha=x next to (resp. with tα=xt\alpha=x on top of) each other.

Corollary 4.3.

Assume QQ has no loops, and let C𝐫repA(𝐝)C_{\mathbf{r}}\subset\operatorname{rep}_{A}(\mathbf{d}) be non-empty. The following set of polynomials in 𝕜[rep𝕜Q(𝐝)]\Bbbk[\operatorname{rep}_{\Bbbk Q}(\mathbf{d})] form a good set of generators for the prime ideal of C𝐫C_{\mathbf{r}}, as xx runs through all the vertices in Q0Q_{0}:

  1. (1)

    The (𝐫(x)+1)×(𝐫(x)+1)(\mathbf{r}(x)+1)\times(\mathbf{r}(x)+1) minors of HxH_{x};

  2. (2)

    The (𝐬(x)+1)×(𝐬(x)+1)(\mathbf{s}(x)+1)\times(\mathbf{s}(x)+1) minors of TxT_{x};

  3. (3)

    The entries of   TxHxT_{x}\cdot H_{x};

Proof.

We work by splitting nodes one at a time, analogously to [KL21, Corollary 4.13]. We note that in Theorem 3.15, the module MM^{\prime} is tilting in this case (see Lemma 2.10 (4)). To conclude using Theorem 3.15 as in [KL21, Corollary 4.13], we are reduced to show that the equations (1)–(3) with x=2x=2 are good defining equations of C𝐫C_{\mathbf{r}} for the following quiver (compare with [KL21, Proposition 4.4])

1\textstyle{1\ignorespaces\ignorespaces\ignorespaces\ignorespaces}a\scriptstyle{a}2\textstyle{2\ignorespaces\ignorespaces\ignorespaces\ignorespaces}b\scriptstyle{b}3\textstyle{3}

As in the case of determinantal varieties in Section 4.1, we can further reduce using Theorem 3.15 (applied at vertices 11 and 33) to the case 𝐫=(0,d1,d3)\mathbf{r}=(0,d_{1},d_{3}) (when we have d2d1+d3d_{2}\geq d_{1}+d_{3}). In such case only the equations of type (3) appear, and they form a regular sequence. Using the Jacobian criterion, one readily obtains that the ideal generated by these polynomials is radical. Moreover, by Lemma 2.15 they give good defining equations for C𝐫rep𝕜Q(𝐝)C_{\mathbf{r}}\subset\operatorname{rep}_{\Bbbk Q}(\mathbf{d}), thus yielding the conclusion. ∎

The article [KL21] further demonstrates the usefulness of working in the relative situation XVX\subset V. By splitting nodes one at a time, the method is applied to a large number of other quiver varieties in characteristic zero. The main obstruction to extending such results to positive characteristics readily is that so far the good property of the corresponding LL-variety XX has been studied only in a handful of cases (e.g.  [Don90]).

4.3. Further examples

When G=GL(n)G=\operatorname{GL}(n), L=GL(r)L=\operatorname{GL}(r) (with rnr\leq n), W=ΔG(λ)W=\Delta_{G}(\lambda) and V=ΔL(λ)V=\Delta_{L}(\lambda), the variety GVG\cdot V is called higher rank variety [Wey03, Section 7]. Thus, Theorem 3.10 generalizes Proposition 7.1.2 in loc. cit. to characteristics that are not “too small”, and further gives new results for the varieties BwV¯\overline{BwV}. We note that the result does not hold in arbitrary characteristic, as the following example shows.

Example 4.4.

Let G=GL(3)G=\operatorname{GL}(3), W=3𝕜6W=\bigwedge^{3}\Bbbk^{6}, V=3𝕜5V=\bigwedge^{3}\Bbbk^{5} with char𝕜=2\operatorname{char}\Bbbk=2. Then VV is a good variety, but GVG\cdot V is not normal, as shown by Weyman [Wey03, Proposition 7.3.10]. Using Theorems 3.6 and 3.10 we see that WW is not good (nor is the hypersurface given by the discriminant of degree 44), a fact further observed in [vdK04, Example 3.3]. Nevertheless, by Remark 3.12 the normalization of GVG\cdot V is strongly FF-regular.

We can extrapolate this to higher dimensions as follows. Set X:=GVX:=G\cdot V from above. Let n6n\geq 6, and consider inclusions 3𝕜53𝕜63𝕜n\bigwedge^{3}\Bbbk^{5}\subset\bigwedge^{3}\Bbbk^{6}\subset\bigwedge^{3}\Bbbk^{n}. Then the saturation Y:=GL(n)3𝕜53𝕜nY:=\operatorname{GL}(n)\cdot\bigwedge^{3}\Bbbk^{5}\,\subset\,\bigwedge^{3}\Bbbk^{n} is the same as the saturation GL(n)XGL(n)3𝕜63𝕜n\operatorname{GL}(n)\cdot X\,\subset\operatorname{GL}(n)\cdot\bigwedge^{3}\Bbbk^{6}\,\subset\bigwedge^{3}\Bbbk^{n}. We have seen that XX is not strongly FF-regular, hence neither is YY by Proposition 3.4 and [HH94a, Theorem 5.5], but the normalization of YY is again strongly FF-regular by Remark 3.12. In particular, 3𝕜n\bigwedge^{3}\Bbbk^{n} is not good by Theorems 3.6 and 3.10. ∎

Other examples of saturations GVG\cdot V (and BwV¯\overline{BwV}) where our results can be readily applied include varieties considered in [Kem76, Section 2], [KR87], [SW15], [Fri10], [LW09], and the subspace varieties in [LW07] (including the relative setting for secant varieties, as in [LW07, Proposition 5.1]), thus strengthening the corresponding results therein.

As explained in the Introduction, the results can be effectively used in the study of the geometry of orbit closures for any representation WW (as in (1.1)) of a reductive group. Since such problems have been pursued intensively in numerous articles for various special representations, it would be difficult to list them all in relation with our results. We simply direct the reader to [Wey03] and the references therein for a large collection of such examples.

4.4. Vanishing results for bundles on Schubert varieties

First, we record the following positive characteristic version of the Grauert–Riemenschneider theorem for collapsing of bundles (cf. [Kem76, Section 3]). Such results are of interest (see [BK05, Theorem 1.3.14]), as in general they do not hold in positive characteristic. We continue with the notation from Section 3. We denote by ωY\omega_{Y} the canonical sheaf of a Cohen–Macaulay variety YY and put η=𝒱(V)\eta=\mathcal{V}(V^{*}) as in Remark 3.8.

Proposition 4.5.

Take w𝒲Iw\in\mathcal{W}^{I} and put c=dimX(w)+dimVdimBwV¯c=\dim X(w)+\dim V-\dim\overline{BwV}. If GVG\cdot V is good then 𝐑cqωBwP¯×PVωBwV¯\mathbf{R}^{c}q_{*}\,\omega_{\overline{BwP}\times_{P}V}\cong\omega_{\overline{BwV}} and

Hi(X(w)P,SymdηdetηωX(w)P)=0 for all ic,d0.H^{i}(X(w)_{P},\,\,\operatorname{Sym}_{d}\eta\otimes\det\eta\otimes\omega_{X(w)_{P}}\,)=0\mbox{ for all }i\neq c,d\geq 0.
Proof.

Put Y=BwP¯×PVY=\overline{BwP}\times_{P}V and Z=BwV¯Z=\overline{BwV}. By Theorem 3.10 we have 𝐑q𝒪Y𝒪Z\mathbf{R}q_{*}\mathcal{O}_{Y}\cong\mathcal{O}_{Z}. and ZZ is Cohen–Macaulay (2.5). As q!ωZωY[c]q^{!}\omega_{Z}\cong\omega_{Y}[c], we obtain using Grothendieck duality [Har66, Theorem III.11.1]

𝐑qωY𝐑qom𝒪Y(𝒪Y,ωY)om𝒪Z(𝒪Z,ωZ[c])ωZ[c].\mathbf{R}q_{*}\omega_{Y}\cong\mathbf{R}q_{*}\mathcal{H}\!om_{\mathcal{O}_{Y}}(\mathcal{O}_{Y},\omega_{Y})\cong\mathcal{H}\!om_{\mathcal{O}_{Z}}(\mathcal{O}_{Z},\omega_{Z}[-c])\cong\omega_{Z}[-c].

The conclusion follows by the adjunction formula [Har77, Proposition II.8.20]. ∎

Remark 4.6.

When X(w)=G/PX(w)=G/P and char𝕜=0\operatorname{char}\Bbbk=0, the bundle SymdηdetηωX(w)P\operatorname{Sym}_{d}\eta\otimes\det\eta\otimes\omega_{X(w)_{P}} is semi-simple. Thus, using the Borel–Weil–Bott theorem (see [Wey03, Section 4] and [Jan03, Corollary 5.5]) and Serre duality [Har77, Corollary 7.7], in this case we can deduce from Proposition 4.5 that the LL-dominant weights that appear in SymVdetV\operatorname{Sym}V\otimes\det V are either singular or lie in a single Bott chamber (giving cohomology in degree dimGVdimV\dim G\cdot V-\dim V). ∎

If we only assume that VV is good, one can give an analogous result to Proposition 4.5 using normalization as in Remark 3.12. Along these lines, we give the following version of Griffiths’ vanishing theorem [Gri69] for Schubert varieties in positive characteristic.

Corollary 4.7.

Assume VV is a good and let λX(T)+\lambda\in X(T)_{+} with λ,αi=0\langle\lambda,\alpha_{i}^{\vee}\rangle=0 if and only if iIi\in I (i.e. (λ)\mathcal{L}(\lambda) is ample on G/PG/P). Then

Hi(X(w)P,Symdηdetη(λ)ωX(w)P)=0 for all i>0,d0,w𝒲I.H^{i}(X(w)_{P},\,\,\operatorname{Sym}_{d}\eta\otimes\det\eta\otimes\mathcal{L}(\lambda)\otimes\omega_{X(w)_{P}}\,)=0\mbox{ for all }i>0,d\geq 0,w\in\mathcal{W}^{I}.
Proof.

We put W=ΔG(λ)WW^{\prime}=\Delta_{G}(\lambda)\oplus W, V=𝕜λVV^{\prime}=\Bbbk_{\lambda}\oplus V and consider q:G×PVGVq:G\times_{P}V^{\prime}\to G\cdot V^{\prime}. To conclude by Proposition 4.5 in combination with Remark 3.12, it is enough to show that qq is an isomorphism on the open G×P((𝕜λ{0})×V)G\times_{P}((\Bbbk_{\lambda}\setminus\{0\})\times V) (so qq is birational). It is known that the map q1:G×P𝕜λG𝕜λq_{1}:G\times_{P}\Bbbk_{\lambda}\to G\cdot\Bbbk_{\lambda} is an isomorphism on the open G×P(𝕜λ{0})G\times_{P}(\Bbbk_{\lambda}\setminus\{0\}) (e.g. [Wey03, Exercise 5.8]). Further, we have an isomorphism G×P(𝕜λ×W)(G×P𝕜λ)×WG\times_{P}(\Bbbk_{\lambda}\times W)\cong(G\times_{P}\Bbbk_{\lambda})\times W given by (g,l,w)(g,l,gw)(g,l,w)\mapsto(g,l,gw). Composing the latter map with q1q_{1} we obtain the result. ∎

Note that when η\eta is a line bundle and d=0d=0 (or when V=0V=0), the result amounts to the classical Kodaira-type vanishing property for Schubert varieties that can be realized as a consequence of Frobenius splitting or global FF-regularity [MR85], [LRPT06], [Smi00].

References

  • [AJ84] H. H. Andersen and J. C. Jantzen. Cohomology of induced representations for algebraic groups. Math. Ann., 269(4):487–525, 1984.
  • [And80] H. H. Andersen. The Frobenius morphism on the cohomology of homogeneous vector bundles on G/BG/B. Ann. of Math. (2), 112(1):113–121, 1980.
  • [Băe01] C. Băeţică. FF-rationality of algebras defined by Pfaffians. Math. Rep. (Bucur.), 3(53):139–144, 2001. Memorial issue dedicated to Nicolae Radu.
  • [Băe06] C. Băeţică. Combinatorics of determinantal ideals. Nova Science Publishers, Inc., Hauppauge, NY, 2006.
  • [BK05] M. Brion and S. Kumar. Frobenius splitting methods in geometry and representation theory, volume 231 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 2005.
  • [Bof91] G. Boffi. On some plethysms. Adv. Math., 89(2):107–126, 1991.
  • [Bou87] J.-F. Boutot. Singularités rationalles et quotients par les groupes réductifs. Invent. Math., 88:65–68, 1987.
  • [Bri85] M. Brion. Représentations exceptionnelles des groupes semi-simples. Ann. Sci. École Norm. Sup. (4), 18(2):345–387, 1985.
  • [Bri86] M. Brion. Quelques propriétés des espaces homogènes sphériques. Manuscripta Math., 55(2):191–198, 1986.
  • [Bri01] M. Brion. On orbit closures of spherical subgroups in flag varieties. Comment. Math. Helv., 76:263–299, 2001.
  • [Bri03] M. Brion. Multiplicity-free subvarieties of flag varieties. In Commutative algebra (Grenoble/Lyon, 2001), volume 331 of Contemp. Math., pages 13–23. Amer. Math. Soc., Providence, RI, 2003.
  • [BT06] M. Brion and J. F. Thomsen. FF-regularity of large Schubert varieties. Amer. J. Math., 128(4):949–962, 2006.
  • [DCS81] C. De Concini and E. Strickland. On the variety of complexes. Adv. Math., 41:57–77, 1981.
  • [Don85] S. Donkin. Rational representations of algebraic groups. Tensor products and filtration, volume 1140 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1985.
  • [Don88] S. Donkin. Invariants of unipotent radicals. Math. Z., 198(1):117–125, 1988.
  • [Don90] S. Donkin. The normality of closures of conjugacy classes of matrices. Invent. Math., 101(3):717–736, 1990.
  • [Elk78] R. Elkik. Singularités rationnelles et déformations. Invent. Math., 47(2):139–147, 1978.
  • [Fri85] E. M. Friedlander. A canonical filtration for certain rational modules. Math. Z., 188(3):433–438, 1985.
  • [Fri10] M. L. Fries. Standard bases for coordinate rings of cotangent varieties. ProQuest LLC, Ann Arbor, MI, 2010. Thesis (Ph.D.)–Northeastern University.
  • [Ful92] W. Fulton. Flags, Schubert polynomials, degeneracy loci, and determinantal formulas. Duke Math. J., 65(3):381–420, 1992.
  • [Gri69] P. A. Griffiths. Hermitian differential geometry, Chern classes, and positive vector bundles. In Global Analysis (Papers in Honor of K. Kodaira), pages 185–251. Univ. Tokyo Press, Tokyo, 1969.
  • [Gro92] F. D. Grosshans. Contractions of actions of reductive algebraic groups in arbitrary characteristic. Invent. Math., 107:127–133, 1992.
  • [Gro97] F. D. Grosshans. Algebraic homogeneous spaces and invariant theory, volume 1673 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1997.
  • [Har66] R. Hartshorne. Residues and duality. Lecture notes of a seminar on the work of A. Grothendieck, given at Harvard 1963/64. With an appendix by P. Deligne. Lecture Notes in Mathematics, No. 20. Springer-Verlag, Berlin-New York, 1966.
  • [Har77] R. Hartshorne. Algebraic geometry. Springer-Verlag, New York, 1977. Graduate Texts in Mathematics, No. 52.
  • [Has01] M. Hashimoto. Good filtrations of symmetric algebras and strong F-regularity of invariant subrings. Math. Z., 236(3):605–623, 2001.
  • [Has03] M. Hashimoto. Surjectivity of multiplication and FF-regularity of multigraded rings. In Commutative algebra (Grenoble/Lyon, 2001), volume 331 of Contemp. Math., pages 153–170. Amer. Math. Soc., Providence, RI, 2003.
  • [Has06] M. Hashimoto. Another proof of the global FF-regularity of Schubert varieties. Tohoku Math. J. (2), 58(3):323–328, 2006.
  • [Has11] M. Hashimoto. Good filtrations and FF-purity of invariant subrings. J. Math. Soc. Japan, 63(3):815–818, 2011.
  • [Has12] M. Hashimoto. Good filtrations and strong FF-regularity of the ring of UPU_{P}-invariants. J. Algebra, 370:198–220, 2012.
  • [HH90] M. Hochster and C. Huneke. Tight closure, invariant theory, and the Briançon-Skoda theorem. J. Amer. Math. Soc., 3(1):31–116, 1990.
  • [HH94a] M. Hochster and C. Huneke. FF-regularity, test elements, and smooth base change. Trans. Amer. Math. Soc., 346(1):1–62, 1994.
  • [HH94b] M. Hochster and C. Huneke. Tight closure of parameter ideals and splitting in module-finite extensions. J. Algebraic Geom., 3(4):599–670, 1994.
  • [HW02] N. Hara and K.-I. Watanabe. F-regular and F-pure rings vs. log terminal and log canonical singularities. J. Algebraic Geom., 11(2):363–392, 2002.
  • [Jan03] J. C. Jantzen. Representations of algebraic groups, volume 107 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, second edition, 2003.
  • [Kem75] G. R. Kempf. Images of homogeneous vector bundles and varieties of complexes. Bull. Amer. Math. Soc., 81(5):900–901, 1975.
  • [Kem76] G. R. Kempf. On the collapsing of homogeneous bundles. Invent. Math., 37(3):229–239, 1976.
  • [Kem86] G. R. Kempf. Varieties with rational singularities. In The Lefschetz centennial conference, Part I (Mexico City, 1984), volume 58 of Contemp. Math., pages 179–182. Amer. Math. Soc., Providence, RI, 1986.
  • [KL21] R. Kinser and A. C. Lőrincz. Representation varieties of algebras with nodes. arXiv:1810.10997, to appear in J. Inst. Math. Jussieu, 2021.
  • [KMN19] M. Katzman and C. B. Miranda-Neto. Strong FF-regularity and generating morphisms of local cohomology modules. J. Algebra, 525:19–41, 2019.
  • [Kov00] S. J. Kovács. A characterization of rational singularities. Duke Math. J., 102(2):187–191, 2000.
  • [Kov20] S. J. Kovács. Rational singularities. arXiv:1703.02269, 2020.
  • [KR87] G. R. Kempf and A. Ramanathan. Multicones over Schubert varieties. Invent. Math., 87(2):353–363, 1987.
  • [LM98] V. Lakshmibai and P. Magyar. Degeneracy schemes, quiver schemes, and Schubert varieties. Internat. Math. Res. Notices, (12):627–640, 1998.
  • [LRPT06] N. Lauritzen, U. Raben-Pedersen, and J. F. Thomsen. Global FF-regularity of Schubert varieties with applications to DD-modules. J. Amer. Math. Soc., 19(2):345–355, 2006.
  • [LW07] J. M. Landsberg and J. Weyman. On the ideals and singularities of secant varieties of Segre varieties. Bull. Lond. Math. Soc., 39(4):685–697, 2007.
  • [LW09] J. M. Landsberg and J. Weyman. On secant varieties of compact Hermitian symmetric spaces. J. Pure Appl. Algebra, 213(11):2075–2086, 2009.
  • [LW19] A. C. Lőrincz and J. Weyman. Free resolutions of orbit closures of Dynkin quivers. Trans. Amer. Math. Soc., 372:2715–2734, 2019.
  • [Mat90] O. Mathieu. Filtrations of GG-modules. Ann. Sci. École Norm. Sup. (4), 23(4):625–644, 1990.
  • [MR85] V. B. Mehta and A. Ramanathan. Frobenius splitting and cohomology vanishing for Schubert varieties. Ann. of Math. (2), 122(1):27–40, 1985.
  • [MT99a] V. B. Mehta and V. Trivedi. The variety of circular complexes and FF-splitting. Invent. Math., 137:449–460, 1999.
  • [MT99b] V. B. Mehta and V. Trivedi. Variety of complexes and FF-splitting. J. Algebra, 215:352–365, 1999.
  • [Mur64] M. P. Murthy. A note on factorial rings. Arch. Math. (Basel), 15:418–420, 1964.
  • [Per14] N. Perrin. On the geometry of spherical varieties. Transform. Groups, 19(1):171–223, 2014.
  • [Pop74] V. L. Popov. Picard groups of homogeneous spaces of linear algebraic groups and one-dimensional homogeneous vector fiberings. Izv. Akad. Nauk SSSR Ser. Mat., 38:294–322, 1974.
  • [Pop86] V. L. Popov. Contractions of actions of reductive algebraic groups. Mat. Sb. (N.S.), 130(172)(3):310–334, 431, 1986.
  • [RR85] S. Ramanan and A. Ramanathan. Projective normality of flag varieties and Schubert varieties. Invent. Math., 79(2):217–224, 1985.
  • [Smi97] K. E. Smith. FF-rational rings have rational singularities. Amer. J. Math., 119(1):159–180, 1997.
  • [Smi00] K. E. Smith. Globally F-regular varieties: applications to vanishing theorems for quotients of Fano varieties. Michigan Math. J., 48:553–572, 2000. Dedicated to William Fulton on the occasion of his 60th birthday.
  • [Str82] E. Strickland. On the conormal bundle of the determinantal variety. J. Algebra, 75:523–537, 1982.
  • [Str87] E. Strickland. On the variety of projectors. J. Algebra, 106:135–147, 1987.
  • [SW15] S. V. Sam and J. Weyman. Littlewood complexes and analogues of determinantal varieties. Int. Math. Res. Not. IMRN, (13):4663–4707, 2015.
  • [vdK93] W. van der Kallen. Lectures on Frobenius splittings and BB-modules. Published for the Tata Institute of Fundamental Research, Bombay; by Springer-Verlag, Berlin, 1993. Notes by S. P. Inamdar.
  • [vdK04] W. van der Kallen. Cohomology with Grosshans graded coefficients. In Invariant theory in all characteristics, volume 35 of CRM Proc. Lecture Notes, pages 127–138. Amer. Math. Soc., Providence, RI, 2004.
  • [Vin86] E. B. Vinberg. Complexity of actions of reductive groups. Funktsional. Anal. i Prilozhen., 20(1):1–13, 1986.
  • [VP89] E. B. Vinberg and V. L. Popov. Invariant theory. In Algebraic geometry, 4 (Russian), Itogi Nauki i Tekhniki, pages 137–314, 315. Akad. Nauk SSSR, Vsesoyuz. Inst. Nauchn. i Tekhn. Inform., Moscow, 1989.
  • [Wey03] J. Weyman. Cohomology of vector bundles and syzygies, volume 149 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2003.
  • [Zel85] A. V. Zelevinskiĭ. Two remarks on graded nilpotent classes. Uspekhi Mat. Nauk, 40(1(241)):199–200, 1985.