Hostname: page-component-76fb5796d-2lccl Total loading time: 0 Render date: 2024-04-26T02:35:56.173Z Has data issue: false hasContentIssue false

Local duality for representations of finite group schemes

Published online by Cambridge University Press:  18 February 2019

Dave Benson
Affiliation:
Institute of Mathematics, University of Aberdeen, King’s College, Aberdeen AB24 3UE, UK
Srikanth B. Iyengar
Affiliation:
Department of Mathematics, University of Utah, Salt Lake City, UT 84112, USA email iyengar@math.utah.edu
Henning Krause
Affiliation:
Fakultät für Mathematik, Universität Bielefeld, 33501 Bielefeld, Germany
Julia Pevtsova
Affiliation:
Department of Mathematics, University of Washington, Seattle, WA 98195, USA
Rights & Permissions [Opens in a new window]

Abstract

A duality theorem for the stable module category of representations of a finite group scheme is proved. One of its consequences is an analogue of Serre duality, and the existence of Auslander–Reiten triangles for the $\mathfrak{p}$-local and $\mathfrak{p}$-torsion subcategories of the stable category, for each homogeneous prime ideal $\mathfrak{p}$ in the cohomology ring of the group scheme.

Type
Research Article
Copyright
© The Authors 2019 

1 Introduction

This work concerns the modular representation theory of finite groups and group schemes. A starting point for it is a duality theorem for finite groups due to Tate, that appears already in Cartan and Eilenberg [Reference Cartan and EilenbergCE56]. For our purposes it is useful to recast this theorem in terms of stable module categories. The stable module category of a finite group scheme $G$ over a field $k$ is the category obtained from the (abelian) category of finite-dimensional $G$ -modules by annihilating morphisms that factor through a projective module; we denote it $\mathsf{stmod}\,G$ , and write $\text{}\underline{\operatorname{Hom}}_{G}(-,-)$ for the morphisms in this category. The category $\mathsf{stmod}\,G$ is triangulated with suspension $\unicode[STIX]{x1D6FA}^{-1}$ , and Tate duality translates to the statement that for all finite-dimensional $G$ -modules $M$ and $N$ there are natural isomorphisms

$$\begin{eqnarray}\operatorname{Hom}_{k}(\text{}\underline{\operatorname{Hom}}_{G}(M,N),k)\cong \text{}\underline{\operatorname{Hom}}_{G}(N,\unicode[STIX]{x1D6FA}\unicode[STIX]{x1D6FF}_{G}\otimes _{k}M).\end{eqnarray}$$

Here $\unicode[STIX]{x1D6FF}_{G}$ is the modular character of $G$ , described in Jantzen [Reference JantzenJan03, § I.8.8]; it is isomorphic to the trivial representation $k$ when $G$ is a finite group. Tate duality can be deduced from a formula of Auslander and Reiten [Reference AuslanderAus78] that applies to general associative algebras; see Theorem 4.2.

In the language introduced by Bondal and Kapranov [Reference Bondal and KapranovBK89] the isomorphism above says that $\mathsf{stmod}\,G$ has Serre duality with Serre functor $M\mapsto \unicode[STIX]{x1D6FA}\unicode[STIX]{x1D6FF}_{G}\otimes _{k}M$ . One of the main results of our work is that such a duality also holds locally.

The precise statement involves a natural action of the cohomology ring $H^{\ast }(G,k)$ of $G$ on the graded abelian group

$$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M,N)=\bigoplus _{n\in \mathbb{Z}}\text{}\underline{\operatorname{Hom}}_{G}(M,\unicode[STIX]{x1D6FA}^{-n}N).\end{eqnarray}$$

The ring $H^{\ast }(G,k)$ is graded commutative, and also finitely generated as a $k$ -algebra, by a result of Friedlander and Suslin [Reference Friedlander and SuslinFS97]. Fix a homogeneous prime ideal $\mathfrak{p}$ not containing $H^{{\geqslant}1}(G,k)$ and consider the triangulated category $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ that is obtained from $\mathsf{stmod}\,G$ by localising the graded morphisms at $\mathfrak{p}$ and then taking the full subcategory of objects such that the graded endomorphisms are $\mathfrak{p}$ -torsion; see § 7 for details. Our interest in the subcategories $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ stems from the fact that they are the building blocks of $\mathsf{stmod}\,G$ and play a key role in the classification of its tensor ideal thick subcategories; see [Reference Benson, Iyengar and KrauseBIK15]. These subcategories may thus be viewed as analogues of the $K(n)$ -local spectra in stable homotopy theory that give the chromatic filtration of a spectrum; see [Reference RavenelRav92].

The following is our version of local Serre duality.

Theorem 1.1. Let $\mathsf{C}:=\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ and $d$ the Krull dimension of $H^{\ast }(G,k)/\mathfrak{p}$ . For all $M,N$ in $\mathsf{C}$ there are natural isomorphisms

$$\begin{eqnarray}\operatorname{Hom}_{H^{\ast }(G,k)}(\operatorname{Hom}_{\mathsf{C}}^{\ast }(M,N),I(\mathfrak{p}))\cong \operatorname{Hom}_{\mathsf{C}}(N,\unicode[STIX]{x1D6FA}^{d}\unicode[STIX]{x1D6FF}_{G}\otimes _{k}M),\end{eqnarray}$$

where $I(\mathfrak{p})$ is the injective hull of the graded $H^{\ast }(G,k)$ -module $H^{\ast }(G,k)/\mathfrak{p}$ .

One corollary is that $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ has Auslander–Reiten (AR) triangles, so one can bring to bear the techniques of AR theory to the study of $G$ -modules. These results are contained in Theorem 7.10.

We deduce Theorem 1.1 from a more general result concerning $\mathsf{StMod}\,G$ , the stable category of all (including infinite dimensional) $G$ -modules. Consider its subcategory $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ consisting of the $\mathfrak{p}$ -local $\mathfrak{p}$ -torsion modules; in other words, the $G$ -modules whose support is contained in $\{\mathfrak{p}\}$ . This is a compactly generated triangulated category and the full subcategory of compact objects is equivalent, up to direct summands, to $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ ; this is explained in Remark 7.2. There is an idempotent functor $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}:\mathsf{StMod}\,G\rightarrow \mathsf{StMod}\,G$ with image the $\mathfrak{p}$ -local $\mathfrak{p}$ -torsion modules; see § 2 for details. The central result of this work is that $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D6FF}_{G})$ is a dualising object for $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ , in the following sense.

Theorem 1.2. For any $G$ -module $M$ and $i\in \mathbb{Z}$ there is a natural isomorphism

$$\begin{eqnarray}\widehat{\text{Ext}}_{G}^{i}(M,\unicode[STIX]{x1D6E4}_{\mathfrak{ p}}(\unicode[STIX]{x1D6FF}_{G}))\cong \operatorname{Hom}_{H^{\ast }(G,k)}(H^{\ast -d-i}(G,M),I(\mathfrak{p})).\end{eqnarray}$$

This result is proved in § 5. In the isomorphism, the vector space on the left is $\text{}\underline{\operatorname{Hom}}_{G}(M,\unicode[STIX]{x1D6FA}^{-i}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D6FF}_{G}))$ . The statement is in terms of $\widehat{\text{Ext}}$ to underscore its similarity to Serre duality on a nonsingular projective variety $X$ of dimension  $n$ :

$$\begin{eqnarray}\operatorname{Ext}_{X}^{i}({\mathcal{F}},\unicode[STIX]{x1D714}_{X})\cong \operatorname{Hom}_{k}(H^{n-i}(X,{\mathcal{F}}),k),\end{eqnarray}$$

for any coherent sheave ${\mathcal{F}}$ on $X$ ; see, for example, Hartshorne [Reference HartshorneHar77].

When $G$ is a finite group $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k)$ is the Rickard idempotent module $\unicode[STIX]{x1D705}_{V}$ , introduced by Benson, Carlson and Rickard [Reference Benson, Carlson and RickardBCR96], that is associated to the irreducible subvariety $V$ of $\operatorname{Proj}H^{\ast }(G,k)$ defined by  $\mathfrak{p}$ . In this context, Theorem 1.2 was proved by Benson and Greenlees [Reference Benson and GreenleesBG08]; see the paragraph following Theorem 5.1 below for a detailed comparison with their work and that of Benson [Reference BensonBen08].

Concerning $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k)$ , the following consequences of Theorem 1.2 have been anticipated in [Reference BensonBen01] when $G$ is a finite group.

Theorem 1.3. Assume $\unicode[STIX]{x1D6FF}_{G}\cong k$ . The $H^{\ast }(G,k)$ -module $\widehat{\text{Ext}}_{G}^{\ast }(k,\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k))$ is injective and isomorphic to a twist of $I(\mathfrak{p})$ . Also, there is an isomorphism of $k$ -algebras

$$\begin{eqnarray}\widehat{\text{Ext}}_{G}^{\ast }(\unicode[STIX]{x1D6E4}_{\mathfrak{ p}}(k),\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k))\cong (H^{\ast }(G,k)_{\mathfrak{ p}})^{\wedge },\end{eqnarray}$$

where $(-)^{\wedge }$ denotes completion with respect to the $\mathfrak{p}$ -adic topology, and the $G$ -module $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k)$ is pure injective.

Theorem 1.2 can be interpreted to mean that the category $\mathsf{StMod}\,G$ is Gorenstein, for it is analogous to Grothendieck’s result that a commutative noetherian ring $A$ is Gorenstein if, and only if, $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}A$ is the injective hull of $A/\mathfrak{p}$ , up to suspension, for each $\mathfrak{p}$ in $\operatorname{Spec}A$ . In § 6 we propose a general notion of a Gorenstein triangulated category, in an attempt to place these results in a common framework.

To prove Theorem 1.2 we use a technique from algebraic geometry in the tradition of Zariski and Weil; namely, the construction of generic points for algebraic varieties. Given a point $\mathfrak{p}\subseteq H^{\ast }(G,k)$ , there is a purely transcendental extension $K$ of $k$ and a closed point $\mathfrak{m}$ of $\operatorname{Proj}H^{\ast }(G_{K},K)$ lying above the point $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ . Here, $G_{K}$ denotes the group scheme that is obtained from $G$ by extending the field to  $K$ . The crux is that one can choose $\mathfrak{m}$ such that the following statement holds.

Theorem 1.4. Restriction of scalars induces an exact functor

$$\begin{eqnarray}\mathsf{stmod}\,G_{K}\supseteq \unicode[STIX]{x1D6FE}_{\mathfrak{m}}(\mathsf{stmod}\,G_{K})\longrightarrow \unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)\end{eqnarray}$$

that is surjective on objects, up to isomorphism.

This result is proved in § 3, building on our work in [Reference Benson, Iyengar, Krause and PevtsovaBIKP18]. It gives a remarkable description of the compact objects in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ : they are obtained from the finite-dimensional objects in $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})$ by restriction of scalars. This allows one to reduce the proof of Theorem 1.2 to the case of a closed point, where it is essentially equivalent to classical Tate duality. The theorem above has other consequences; for example, it implies that the compact objects in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ are endofinite $G$ -modules in the sense of Crawley-Boevey [Reference Crawley-BoeveyCra92]; see § 3.

2 Cohomology and localisation

In this section we recall basic notions concerning certain localisation functors on triangulated categories with ring actions. The material is needed to state and prove the results in this work. The main triangulated category of interest is the stable module category of a finite group scheme, but the general framework is needed in §§ 6 and 7. Primary references for the material presented here are [Reference Benson, Iyengar and KrauseBIK08, Reference Benson, Iyengar and KrauseBIK11a]; see [Reference Benson, Iyengar, Krause and PevtsovaBIKP18] for the special case of the stable module category.

Triangulated categories with central action

Let $\mathsf{T}$ be a triangulated category with suspension $\unicode[STIX]{x1D6F4}$ . For objects $X$ and $Y$ in $\mathsf{T}$ set

$$\begin{eqnarray}\operatorname{Hom}_{\mathsf{T}}^{\ast }(X,Y):=\bigoplus _{i\in \mathbb{Z}}\operatorname{Hom}_{\mathsf{T}}(X,\unicode[STIX]{x1D6F4}^{i}Y)\quad \text{and}\quad \operatorname{End}_{\mathsf{T}}^{\ast }(X):=\operatorname{Hom}_{\mathsf{T}}^{\ast }(X,X).\end{eqnarray}$$

Composition makes $\operatorname{End}_{\mathsf{T}}^{\ast }(X)$ a graded ring and $\operatorname{Hom}_{\mathsf{T}}^{\ast }(X,Y)$ a left- $\operatorname{End}_{\mathsf{T}}^{\ast }(Y)$ right- $\operatorname{End}_{\mathsf{T}}^{\ast }(X)$ module.

Let $R$ be a graded commutative noetherian ring. In what follows we will only be concerned with homogeneous elements and ideals in  $R$ . In this spirit, ‘localisation’ will mean homogeneous localisation, and $\operatorname{Spec}R$ will denote the set of homogeneous prime ideals in  $R$ .

We say that a triangulated category $\mathsf{T}$ is $R$ -linear if for each $X$ in $\mathsf{T}$ there is a homomorphism of graded rings $\unicode[STIX]{x1D719}_{X}:R\rightarrow \operatorname{End}_{\mathsf{T}}^{\ast }(X)$ such that the induced left and right actions of $R$ on $\operatorname{Hom}_{\mathsf{T}}^{\ast }(X,Y)$ are compatible in the following sense: for any $r\in R$ and $\unicode[STIX]{x1D6FC}\in \operatorname{Hom}_{\mathsf{T}}^{\ast }(X,Y)$ , one has

$$\begin{eqnarray}\unicode[STIX]{x1D719}_{Y}(r)\unicode[STIX]{x1D6FC}=(-1)^{|r|\,|\unicode[STIX]{x1D6FC}|}\unicode[STIX]{x1D6FC}\unicode[STIX]{x1D719}_{X}(r).\end{eqnarray}$$

An exact functor $F:\mathsf{T}\rightarrow \mathsf{U}$ between $R$ -linear triangulated categories is $R$ -linear if the induced map

$$\begin{eqnarray}\operatorname{Hom}_{\mathsf{T}}^{\ast }(X,Y)\longrightarrow \operatorname{Hom}_{\mathsf{U}}^{\ast }(FX,FY)\end{eqnarray}$$

of graded abelian groups is $R$ -linear for all objects $X,Y$ in  $\mathsf{T}$ .

In what follows, we fix a compactly generated $R$ -linear triangulated category $\mathsf{T}$ and write $\mathsf{T}^{\mathsf{c}}$ for its full subcategory of compact objects.

Localisation

Fix an ideal $\mathfrak{a}$ in $R$ . An $R$ -module $M$ is $\mathfrak{a}$ -torsion if $M_{\mathfrak{q}}=0$ for all $\mathfrak{q}$ in $\operatorname{Spec}R$ with $\mathfrak{a}\not \subseteq \mathfrak{q}$ . Analogously, an object $X$ in $\mathsf{T}$ is $\mathfrak{a}$ -torsion if the $R$ -module $\operatorname{Hom}_{\mathsf{T}}^{\ast }(C,X)$ is $\mathfrak{a}$ -torsion for all $C\in \mathsf{T}^{\mathsf{c}}$ . The full subcategory of $\mathfrak{a}$ -torsion objects

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a})}\mathsf{T}:=\{X\in \mathsf{T}\mid X~\text{is}~\mathfrak{a}\text{-}\text{torsion}\}\end{eqnarray}$$

is localising and the inclusion $\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a})}\mathsf{T}\subseteq \mathsf{T}$ admits a right adjoint, denoted $\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a})}$ .

Fix a $\mathfrak{p}$ in $\operatorname{Spec}R$ . An $R$ -module $M$ is $\mathfrak{p}$ -local if the localisation map $M\rightarrow M_{\mathfrak{p}}$ is invertible, and an object $X$ in $\mathsf{T}$ is $\mathfrak{p}$ -local if the $R$ -module $\operatorname{Hom}_{\mathsf{T}}^{\ast }(C,X)$ is $\mathfrak{p}$ -local for all $C\in \mathsf{T}^{\mathsf{c}}$ . Consider the full subcategory of $\mathsf{T}$ of $\mathfrak{p}$ -local objects

$$\begin{eqnarray}\mathsf{T}_{\mathfrak{p}}:=\{X\in \mathsf{T}\mid X~\text{is}~\mathfrak{p}\text{-}\text{local}\}\end{eqnarray}$$

and the full subcategory of $\mathfrak{p}$ -local and $\mathfrak{p}$ -torsion objects

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T}:=\{X\in \mathsf{T}\mid X~\text{is}~\mathfrak{p}\text{-}\text{local and}~\mathfrak{p}\text{-}\text{torsion}\}.\end{eqnarray}$$

Note that $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T}\subseteq \mathsf{T}_{\mathfrak{p}}\subseteq \mathsf{T}$ are localising subcategories. The inclusion $\mathsf{T}_{\mathfrak{p}}\rightarrow \mathsf{T}$ admits a left adjoint $X\mapsto X_{\mathfrak{p}}$ while the inclusion $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T}\rightarrow \mathsf{T}_{\mathfrak{p}}$ admits a right adjoint. We denote by $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}:\mathsf{T}\rightarrow \unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T}$ the composition of those adjoints; it is the local cohomology functor with respect to  $\mathfrak{p}$ ; see [Reference Benson, Iyengar and KrauseBIK08, Reference Benson, Iyengar and KrauseBIK11a] for the construction of this functor.

The following observation is clear.

Lemma 2.1. For any element $r$ in $R\setminus \mathfrak{p}$ , say of degree  $n$ , and $\mathfrak{p}$ -local object  $X$ , the natural map $X\xrightarrow[{}]{r}\unicode[STIX]{x1D6F4}^{n}X$ is an isomorphism.

The functor $\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a})}$ commutes with exact functors preserving coproducts.

Lemma 2.2. Let $F:\mathsf{T}\rightarrow \mathsf{U}$ be an exact functor between $R$ -linear compactly generated triangulated categories such that $F$ is $R$ -linear and preserves coproducts. Suppose that the action of $R$ on $\mathsf{U}$ factors through a homomorphism $f:R\rightarrow S$ of graded commutative rings. For any ideal $\mathfrak{a}$ of $R$ there is a natural isomorphism

$$\begin{eqnarray}F\circ \unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a})}\cong \unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a}S)}\circ F\end{eqnarray}$$

of functors $\mathsf{T}\rightarrow \mathsf{U}$ , where $\mathfrak{a}S$ denotes the ideal of $S$ that is generated by  $f(\mathfrak{a})$ .

Proof. The statement follows from an explicit description of $\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a})}$ in terms of homotopy colimits; see [Reference Benson, Iyengar and KrauseBIK11a, Proposition 2.9].◻

Injective cohomology objects

Given an object $C$ in $\mathsf{T}^{\mathsf{c}}$ and an injective $R$ -module  $I$ , Brown representability yields an object $T(C,I)$ in $\mathsf{T}$ such that

(2.1) $$\begin{eqnarray}\operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{T}}^{\ast }(C,-),I)\cong \operatorname{Hom}_{\mathsf{T}}(-,T(C,I)).\end{eqnarray}$$

This yields a functor

$$\begin{eqnarray}T:\mathsf{T}^{\mathsf{c}}\times \mathsf{Inj}\,R\longrightarrow \mathsf{T}.\end{eqnarray}$$

For each $\mathfrak{p}$ in $\operatorname{Spec}R$ , we write $I(\mathfrak{p})$ for the injective hull of $R/\mathfrak{p}$ and set

$$\begin{eqnarray}T_{\mathfrak{p}}:=T(-,I(\mathfrak{p})),\end{eqnarray}$$

viewed as a functor $\mathsf{T}^{\mathsf{c}}\rightarrow \mathsf{T}$ .

Tensor triangulated categories

Let $\mathsf{T}=(\mathsf{T},\otimes ,\unicode[STIX]{x1D7D9})$ be a tensor triangulated category such that $R$ acts on $\mathsf{T}$ via a homomorphism of graded rings $R\rightarrow \operatorname{End}_{\mathsf{T}}^{\ast }(\unicode[STIX]{x1D7D9})$ . Brown representability yields functions objects ${\mathcal{H}}om(X,Y)$ satisfying an adjunction isomorphism

$$\begin{eqnarray}\operatorname{Hom}_{\mathsf{T}}(X\otimes Y,Z)\cong \operatorname{Hom}_{\mathsf{T}}(X,{\mathcal{H}}om(Y,Z))\quad \text{for all}~X,Y,Z~\text{in}~\mathsf{T}.\end{eqnarray}$$

Set $X^{\vee }:={\mathcal{H}}om(X,\unicode[STIX]{x1D7D9})$ for each $X$ in $\mathsf{T}$ . It is part of our definition of a tensor triangulated category that the unit, $\unicode[STIX]{x1D7D9}$ , is compact, and that compact objects are rigid: for all $C,X$ in $\mathsf{T}$ with $C$ compact the natural map

$$\begin{eqnarray}C^{\vee }\otimes X\longrightarrow {\mathcal{H}}om(C,X)\end{eqnarray}$$

is an isomorphism; see [Reference Benson, Iyengar and KrauseBIK08, § 8] for details.

The functors $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}$ and $T_{\mathfrak{p}}$ can be computed as follows:

(2.2) $$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\cong \unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D7D9})\otimes -\quad \text{and}\quad T_{\mathfrak{p}}\cong T_{\mathfrak{p}}(\unicode[STIX]{x1D7D9})\otimes -.\end{eqnarray}$$

Indeed, the first isomorphism is from [Reference Benson, Iyengar and KrauseBIK08, Corollary 8.3], while the second one holds because for each $X\in \mathsf{T}$ and compact object $C$ there are isomorphisms

$$\begin{eqnarray}\displaystyle \operatorname{Hom}_{\mathsf{T}}(X,T_{\mathfrak{p}}(\unicode[STIX]{x1D7D9})\otimes C) & \cong & \displaystyle \operatorname{Hom}_{\mathsf{T}}(X,{\mathcal{H}}om(C^{\vee },T_{\mathfrak{p}}(\unicode[STIX]{x1D7D9})))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{\mathsf{T}}(X\otimes C^{\vee },T_{\mathfrak{p}}(\unicode[STIX]{x1D7D9}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{T}}^{\ast }(\unicode[STIX]{x1D7D9},X\otimes C^{\vee }),I(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{T}}^{\ast }(\unicode[STIX]{x1D7D9},{\mathcal{H}}om(C,X),I(\mathfrak{p})))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{T}}^{\ast }(C,X),I(\mathfrak{p})).\nonumber\end{eqnarray}$$

The first and the fourth isomorphisms above hold because $C$ is rigid; the second and the last one are adjunction isomorphisms; the third one is by the defining isomorphism (2.1).

We turn now to modules over finite group schemes, following the notation and terminology from [Reference Benson, Iyengar, Krause and PevtsovaBIKP18].

The stable module category

Let $G$ be a finite group scheme over a field $k$ of positive characteristic. The coordinate ring and the group algebra of $G$ are denoted $k[G]$ and $kG$ , respectively. These are finite-dimensional Hopf algebras over $k$ that are dual to each other. We write $\mathsf{Mod}\,G$ for the category of $G$ -modules and $\mathsf{mod}\,G$ for its full subcategory consisting of finite-dimensional $G$ -modules. We often identity $\mathsf{Mod}\,G$ with the category of $kG$ -modules, which is justified by [Reference JantzenJan03, I.8.6].

We write $H^{\ast }(G,k)$ for the cohomology algebra, $\operatorname{Ext}_{G}^{\ast }(k,k)$ , of  $G$ . This is a graded commutative $k$ -algebra, because $kG$ is a Hopf algebra, and acts on $\operatorname{Ext}_{G}^{\ast }(M,N)$ , for any $G$ -modules $M,N$ . Moreover, the $k$ -algebra $H^{\ast }(G,k)$ is finitely generated, and, when $M,N$ are finite dimensional, $\operatorname{Ext}_{G}^{\ast }(M,N)$ is finitely generated over it; this is by a theorem of Friedlander and Suslin [Reference Friedlander and SuslinFS97].

The stable module category $\mathsf{StMod}\,G$ is obtained from $\mathsf{Mod}\,G$ by identifying two morphisms between $G$ -modules when they factor through a projective $G$ -module. An isomorphism in $\mathsf{StMod}\,G$ will be called a stable isomorphism, to distinguish it from an isomorphism in $\mathsf{Mod}\,G$ . In the same vein, $G$ -modules $M$ and $N$ are said to be stably isomorphic if they are isomorphic in $\mathsf{StMod}\,G$ ; this is equivalent to the condition that $M$ are $N$ are isomorphic in $\mathsf{Mod}\,G$ , up to projective summands.

The tensor product over $k$ of $G$ -modules passes to $\mathsf{StMod}\,G$ and yields a tensor triangulated category with unit $k$ and suspension $\unicode[STIX]{x1D6FA}^{-1}$ , the inverse of the syzygy functor. The category $\mathsf{StMod}\,G$ is compactly generated and the subcategory of compact objects identifies with $\mathsf{stmod}\,G$ , the stable module category of finite-dimensional $G$ -modules. See Carlson [Reference CarlsonCar96, § 5] and Happel [Reference HappelHap88, ch. I] for details.

We use the notation $\text{}\underline{\operatorname{Hom}}_{G}(M,N)$ for the $\operatorname{Hom}$ -sets in $\mathsf{StMod}\,G$ . The cohomology algebra $H^{\ast }(G,k)$ acts on $\mathsf{StMod}\,G$ via a homomorphism of $k$ -algebras

$$\begin{eqnarray}-\otimes _{k}M:H^{\ast }(G,k)=\operatorname{Ext}_{G}^{\ast }(k,k)\longrightarrow \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M,M).\end{eqnarray}$$

Thus, the preceding discussion on localisation functors on triangulated categories applies to the $H^{\ast }(G,k)$ -linear category $\mathsf{StMod}\,G$ .

Koszul objects

Each $b$ in $H^{d}(G,k)$ corresponds to a morphism $k\rightarrow \unicode[STIX]{x1D6FA}^{-d}k$ in $\mathsf{StMod}\,G$ ; let $k/\!\!/b$ denote its mapping cone. This gives a morphism $k\rightarrow \unicode[STIX]{x1D6FA}^{d}(k/\!\!/b)$ . For a sequence of elements $\boldsymbol{b}:=b_{1},\ldots ,b_{n}$ in $H^{\ast }(G,k)$ and a $G$ -module  $M$ , we set

$$\begin{eqnarray}k/\!\!/\boldsymbol{b}:=(k/\!\!/b_{1})\otimes _{k}\cdots \otimes _{k}(k/\!\!/b_{n})\quad \text{and}\quad M/\!\!/\boldsymbol{b}:=M\otimes _{k}k/\!\!/\boldsymbol{b}.\end{eqnarray}$$

It is easy to check that for a $G$ -module $N$ and $s=\sum _{i}|b_{i}|$ , there is an isomorphism

(2.3) $$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{G}(M,N/\!\!/\boldsymbol{b})\cong \text{}\underline{\operatorname{Hom}}_{G}(\unicode[STIX]{x1D6FA}^{n+s}M/\!\!/\boldsymbol{b},N).\end{eqnarray}$$

Let $\mathfrak{b}=(\boldsymbol{b})$ be the ideal of $H^{\ast }(G,k)$ generated by  $\boldsymbol{b}$ . By abuse of notation we set $M/\!\!/\mathfrak{b}:=M/\!\!/\boldsymbol{b}$ . If $\boldsymbol{b}^{\prime }$ is a finite set of elements in $H^{\ast }(G,k)$ such that $\sqrt{(\boldsymbol{b}^{\prime })}=\sqrt{(\boldsymbol{b})}$ , then, by [Reference Benson, Iyengar and KrauseBIK15, Proposition 3.10], for any $M$ in $\mathsf{StMod}\,G$ there is an equality

(2.4) $$\begin{eqnarray}\mathsf{Thick}(M/\!\!/\boldsymbol{b})=\mathsf{Thick}(M/\!\!/\boldsymbol{b}^{\prime }).\end{eqnarray}$$

Fix $\mathfrak{p}$ in $\operatorname{Spec}H^{\ast }(G,k)$ . We will repeatedly use the fact that $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)^{\mathsf{c}}$ is generated as a triangulated category by the family of objects $(M/\!\!/\mathfrak{p})_{\mathfrak{p}}$ with $M$ in $\mathsf{stmod}\,G$ ; see [Reference Benson, Iyengar and KrauseBIK11a, Proposition 3.9]. In fact, if $S$ denotes the direct sum of a representative set of simple $G$ -modules, then there is an equality

(2.5) $$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)^{\mathsf{c}}=\mathsf{Thick}((S/\!\!/\mathfrak{p})_{\mathfrak{p}}).\end{eqnarray}$$

It turns out that one has $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G)=\{0\}$ where $\mathfrak{m}$ denotes $H^{{\geqslant}1}(G,k)$ , the ideal of elements of positive degree; see Lemma 2.5 below. For this reason, it is customary to focus on $\operatorname{Proj}H^{\ast }(G,k)$ , the set of homogeneous prime ideals not containing $\mathfrak{m}$ , when dealing with $\mathsf{StMod}\,G$ .

Tate cohomology

By construction, the action of $H^{\ast }(G,k)$ on $\mathsf{StMod}\,G$ factors through $\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,k)$ , the graded ring of endomorphisms of the identity. The latter ring is not noetherian in general, which is one reason to work with $H^{\ast }(G,k)$ . In any case, there is little difference, vis-à-vis their action on $\mathsf{StMod}\,G$ , as the next remarks should make clear.

Remark 2.3. Let $M$ and $N$ be $G$ -modules. The map $\operatorname{Hom}_{G}(M,N)\rightarrow \text{}\underline{\operatorname{Hom}}_{G}(M,N)$ induces a map $\operatorname{Ext}_{G}^{\ast }(M,N)\rightarrow \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M,N)$ of $H^{\ast }(G,k)$ -modules. This map is surjective in degree zero, with kernel $\operatorname{PHom}_{G}(M,N)$ , the maps from $M$ to $N$ that factor through a projective $G$ -module. It is bijective in positive degrees and hence one gets an exact sequence of graded $H^{\ast }(G,k)$ -modules

(2.6) $$\begin{eqnarray}0\longrightarrow \operatorname{PHom}_{G}(M,N)\longrightarrow \operatorname{Ext}_{G}^{\ast }(M,N)\longrightarrow \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M,N)\longrightarrow X\longrightarrow 0\end{eqnarray}$$

with $X^{i}=0$ for $i\geqslant 0$ . For degree reasons, the $H^{\ast }(G,k)$ -modules $\operatorname{PHom}_{G}(M,N)$ and $X$ are $\mathfrak{m}$ -torsion. Consequently, for $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ the induced localised map is an isomorphism:

(2.7) $$\begin{eqnarray}\operatorname{Ext}_{G}^{\ast }(M,N)_{\mathfrak{ p}}\xrightarrow[{}]{~\cong ~}\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M,N)_{\mathfrak{ p}}.\end{eqnarray}$$

More generally, for each $r$ in $\mathfrak{m}$ localisation induces an isomorphism

$$\begin{eqnarray}\operatorname{Ext}_{G}^{\ast }(M,N)_{r}\xrightarrow[{}]{~\cong ~}\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M,N)_{r}\end{eqnarray}$$

of $H^{\ast }(G,k)_{r}$ -modules. This means that $\operatorname{Proj}H^{\ast }(G,k)$ has a finite cover by affine open sets on which ordinary cohomology and stable cohomology agree.

Given the finite generation result due to Friedlander and Suslin mentioned earlier, the next remark can be deduced from the exact sequence (2.6).

Remark 2.4. When $M,N$ are finite-dimensional $G$ -modules, $\text{}\underline{\operatorname{Hom}}_{G}^{{\geqslant}s}(M,N)$ is a finitely generated $H^{\ast }(G,k)$ -module for any $s\in \mathbb{Z}$ . Moreover the $H^{\ast }(G,k)_{\mathfrak{p}}$ -module

$$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M_{\mathfrak{ p}},N_{\mathfrak{p}})\cong \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M,N)_{\mathfrak{ p}}\end{eqnarray}$$

is finitely generated for each $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ .

Lemma 2.5. One has $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G)=\{0\}$ , where $\mathfrak{m}=H^{{\geqslant}1}(G,k)$ .

Proof. Given (2.5) it suffices to check that $S/\!\!/\mathfrak{m}=0$ in $\mathsf{StMod}\,G$ , where $S$ is the direct sum of representative set of simple $G$ -modules. For any $G$ -module  $M$ , the $H^{\ast }(G,k)$ -module $\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(M,S/\!\!/\mathfrak{m})$ is $\mathfrak{m}$ -torsion; see [Reference Benson, Iyengar and KrauseBIK08, Lemma 5.11(1)]. Thus, when $M$ is finite dimensional, the $H^{\ast }(G,k)$ -module $\text{}\underline{\operatorname{Hom}}_{G}^{{\geqslant}0}(M,S/\!\!/\mathfrak{m})$ is $\mathfrak{m}$ -torsion and finitely generated, so it follows that $\text{}\underline{\operatorname{Hom}}_{G}^{i}(M,S/\!\!/\mathfrak{m})=0$ for $i\gg 0$ , by Remark 2.4. This implies that $S/\!\!/\mathfrak{m}$ is projective, since $kG$ is self-injective.◻

To gain a better understanding of the discussion above, it helps to consider the homotopy category of $\mathsf{Inj}\,G$ , the injective $G$ -modules.

The homotopy category of injectives

Let $\mathsf{K}(\mathsf{Inj}\,G)$ and $\mathsf{D}(\mathsf{Mod}\,G)$ denote the homotopy category of $\mathsf{Inj}\,G$ and the derived category of $\mathsf{Mod}\,G$ , respectively. These are also $H^{\ast }(G,k)$ -linear compactly generated tensor triangulated categories, with the tensor product over  $k$ . The unit of the tensor product on $\mathsf{K}(\mathsf{Inj}\,G)$ is an injective resolution of the trivial $G$ -module  $k$ , while that of $\mathsf{D}(\mathsf{Mod}\,G)$ is  $k$ . The canonical quotient functor $\mathsf{K}(\mathsf{Inj}\,G)\rightarrow \mathsf{D}(\mathsf{Mod}\,G)$ induces an equivalence of triangulated category $\mathsf{K}(\mathsf{Inj}\,G)^{\mathsf{c}}\xrightarrow[{}]{{\sim}}\mathsf{D}^{b}(\mathsf{mod}\,G)$ , where the target is the bounded derived category of $\mathsf{mod}\,G$ ; see [Reference KrauseKra05, Proposition 2.3].

Taking Tate resolutions identifies $\mathsf{StMod}\,G$ with $\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)$ , the full subcategory of acyclic complexes in $\mathsf{K}(\mathsf{Inj}\,G)$ . In detail, let $\mathsf{p}k$ and $\mathsf{i}k$ be a projective and an injective resolution of the trivial $G$ -module  $k$ , respectively, and let $\mathsf{t}k$ be the mapping cone of the composed morphism $\mathsf{p}k\rightarrow k\rightarrow \mathsf{i}k$ ; this is a Tate resolution of  $k$ . Since projective and injective $G$ -modules coincide, one gets the exact triangle

(2.8) $$\begin{eqnarray}\mathsf{p}k\longrightarrow \mathsf{i}k\longrightarrow \mathsf{t}k\longrightarrow\end{eqnarray}$$

in $\mathsf{K}(\mathsf{Inj}\,G)$ . For a $G$ -module $M$ , the complex $M\otimes _{k}\mathsf{t}k$ is a Tate resolution of $M$ and the assignment $M\mapsto M\otimes _{k}\mathsf{t}k$ induces an equivalence of categories

$$\begin{eqnarray}\mathsf{StMod}\,G\xrightarrow[{}]{{\sim}}\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G),\end{eqnarray}$$

with quasi-inverse $X\mapsto Z^{0}(X)$ , the submodule of cycles in degree 0. Assigning $X$ in $\mathsf{K}(\mathsf{Inj}\,G)$ to $X\otimes _{k}\mathsf{t}k$ is a left adjoint of the inclusion $\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)\rightarrow \mathsf{K}(\mathsf{Inj}\,G)$ . These results are contained in [Reference KrauseKra05, Theorem 8.2]. Consider the composed functor

$$\begin{eqnarray}\unicode[STIX]{x1D70B}:\mathsf{K}(\mathsf{Inj}\,G)\xrightarrow[{}]{-\otimes _{k}\mathsf{t}k}\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)\xrightarrow[{}]{~\sim ~}\mathsf{StMod}\,G.\end{eqnarray}$$

A straightforward verification yields that these functors are $H^{\ast }(G,k)$ -linear. The result below is the categorical underpinning of Remark 2.3 and Lemma 2.5.

Lemma 2.6. There is a natural isomorphism $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}X\cong X\otimes _{k}\mathsf{p}k$ for $X\in \mathsf{K}(\mathsf{Inj}\,G)$ . For each $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ , the functor $\unicode[STIX]{x1D70B}$ induces triangle equivalences

$$\begin{eqnarray}\mathsf{K}(\mathsf{Inj}\,G)_{\mathfrak{p}}\xrightarrow[{}]{{\sim}}(\mathsf{StMod}\,G)_{\mathfrak{p}}\quad \text{and}\quad \unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{K}(\mathsf{Inj}\,G))\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G).\end{eqnarray}$$

Proof. We identify $\mathsf{StMod}\,G$ with $\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)$ . This entails $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G))=\{0\}$ , by Lemma 2.5. It is easy to check that $kG$ is $\mathfrak{m}$ -torsion, and hence so is $\mathsf{p}k$ , for it is in the localising subcategory generated by $kG$ , and the class of $\mathfrak{m}$ -torsion objects in $\mathsf{K}(\mathsf{Inj}\,G)$ is a tensor ideal localising subcategory; see, for instance, [Reference Benson, Iyengar and KrauseBIK08, § 8]. Thus, applying $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(-)$ to the exact triangle (2.8) yields $\mathsf{p}k\cong \unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{i}k)$ . It then follows from (2.2) that $X\otimes _{k}\mathsf{p}k\cong \unicode[STIX]{x1D6E4}_{\mathfrak{m}}X$ for any $X$ in $\mathsf{K}(\mathsf{Inj}\,G)$ .

From the construction of $\unicode[STIX]{x1D70B}$ and (2.8), the kernel of $\unicode[STIX]{x1D70B}$ is the subcategory

$$\begin{eqnarray}\{X\in \mathsf{K}(\mathsf{Inj}\,G)\mid X\otimes _{k}\mathsf{p}k\cong X\}.\end{eqnarray}$$

These are precisely the $\mathfrak{m}$ -torsion objects in $\mathsf{K}(\mathsf{Inj}\,G)$ , by the already established part of the result. Said otherwise, $X\in \mathsf{K}(\mathsf{Inj}\,G)$ is acyclic if and only if $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}X=0$ . It follows that $\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)$ contains the subcategory $\mathsf{K}(\mathsf{Inj}\,G)_{\mathfrak{p}}$ of $\mathfrak{p}$ -local objects, for each $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ . On the other hand, the inclusion $\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)\subseteq \mathsf{K}(\mathsf{Inj}\,G)$ preserves coproducts, so its left adjoint $\unicode[STIX]{x1D70B}$ preserves compactness of objects and all compacts of $\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)$ are in the image of  $\unicode[STIX]{x1D70B}$ . Given this a simple calculation shows that $\mathsf{K}(\mathsf{Inj}\,G)_{\mathfrak{p}}$ contains ${\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)}_{\mathfrak{p}}$ . Thus ${\mathsf{K}_{\text{ac}}(\mathsf{Inj}\,G)}_{\mathfrak{p}}=\mathsf{K}(\mathsf{Inj}\,G)_{\mathfrak{p}}$ .◻

3 Passage to closed points

Let $G$ be a finite group scheme over a field $k$ of positive characteristic. In this section we describe a technique that relates the $\mathfrak{p}$ -local $\mathfrak{p}$ -torsion objects in $\mathsf{StMod}\,G$ , for a point $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ , to the corresponding modules at a closed point defined over a field extension of $k$ . Recall that a point $\mathfrak{m}$ is closed when it is maximal with respect to inclusion: $\mathfrak{m}\subseteq \mathfrak{q}$ implies $\mathfrak{m}=\mathfrak{q}$ for all $\mathfrak{q}$ in $\operatorname{Proj}H^{\ast }(G,k)$ . In what follows, $k(\mathfrak{p})$ denotes the graded residue field of $H^{\ast }(G,k)$ at  $\mathfrak{p}$ .

For a field extension $K/k$ extension of scalars and restriction give exact functors

$$\begin{eqnarray}K\otimes _{k}(-):\mathsf{StMod}\,G\longrightarrow \mathsf{StMod}\,G_{K}\quad \text{and}\quad (-)\downarrow _{G}:\mathsf{StMod}\,G_{K}\longrightarrow \mathsf{StMod}\,G.\end{eqnarray}$$

The functors form an adjoint pair, with the left adjoint $K\otimes _{k}(-)$ mapping $k$ to $K$ and respecting tensor products, so one has a well-known projection formula:

(3.1) $$\begin{eqnarray}M\otimes _{k}N\downarrow _{G}\cong (M_{K}\otimes _{K}N)\downarrow _{G}\end{eqnarray}$$

for a $G$ -module $M$ and $G_{K}$ -module $N$ ; see [Reference Balmer, Dell’Ambrogio and SandersBDS16, (2.16)] or [Reference Benson, Iyengar, Krause and PevtsovaBIKP17, Lemma 2.2].

The functor $K\otimes _{k}(-)$ yields a homomorphism $H^{\ast }(G,k)\rightarrow H^{\ast }(G_{K},K)$ of rings. There is a natural isomorphism $H^{\ast }(G_{K},K)\xrightarrow[{}]{{\sim}}K\otimes _{k}H^{\ast }(G,k)$ of $K$ -algebras, so the preceding map above is just extension of scalars. There is an induced map

$$\begin{eqnarray}\operatorname{Proj}H^{\ast }(G_{K},K)\longrightarrow \operatorname{Proj}H^{\ast }(G,k),\end{eqnarray}$$

with $\mathfrak{q}$ mapping to $\mathfrak{p}:=\mathfrak{q}\cap H^{\ast }(G,k)$ . We say that $\mathfrak{q}$ lies over $\mathfrak{p}$ to indicate this.

The main objective of this section is the proof of the following result. We say a functor is dense if it is surjective on objects, up to isomorphism.

Theorem 3.1. Fix $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ and $K/k$ a purely transcendental extension of degree $\dim (H^{\ast }(G,k)/\mathfrak{p})-1$ . There exists a closed point $\mathfrak{m}$ in $\operatorname{Proj}H^{\ast }(G_{K},K)$ lying over $\mathfrak{p}$ with $k(\mathfrak{m})\cong k(\mathfrak{p})$ such that the functor $(-)\downarrow _{G}$ restricts to functors

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})\rightarrow \unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)\quad \text{and}\quad \unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})^{\mathsf{c}}\rightarrow \unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)^{\mathsf{c}}\end{eqnarray}$$

that are dense.

The proof of the theorem yields more: there is a subcategory of $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})$ on which $(-)\downarrow _{G}$ is full and dense; ditto for the category of compact objects. However, the functor need not be full on all of $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})^{\mathsf{c}}$ ; see Example 3.7.

Here is one consequence of Theorem 3.1.

Corollary 3.2. The compact objects in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ are, up to isomorphism, the restrictions of finite-dimensional $G_{K}$ -modules in $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})$ .

Proof. By [Reference Benson, Iyengar and KrauseBIK08, Theorem 6.4], for any ideal $\mathfrak{a}$ in $H^{\ast }(G,k)$ , we have

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a})}(\mathsf{StMod}\,G)^{\mathsf{c}}=\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{a})}(\mathsf{StMod}\,G)\cap \mathsf{stmod}\,G.\end{eqnarray}$$

Applying this observation to the ideal $\mathfrak{m}$ of $H^{\ast }(G_{K},K)$ and noting that $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}=\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{m})}$ , since $\mathfrak{m}$ is a closed point, the desired result follows from Theorem 3.1.◻

The closed point in Theorem 3.1 depends on the choice of a Noether normalisation of $H^{\ast }(G,k)/\mathfrak{p}$ as is explained in the construction below, from [Reference Benson, Iyengar, Krause and PevtsovaBIKP18, § 7].

Construction 3.3. Fix $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ ; the following construction is relevant only when $\mathfrak{p}$ is not a closed point. Choose elements $\boldsymbol{a}:=a_{0},\ldots ,a_{d-1}$ in $H^{\ast }(G,k)$ of the same degree such that their image in $H^{\ast }(G,k)/\mathfrak{p}$ is algebraically independent over $k$ and $H^{\ast }(G,k)/\mathfrak{p}$ is finitely generated as a module over the subalgebra $k[\boldsymbol{a}]$ . Thus the Krull dimension of $H^{\ast }(G,k)/\mathfrak{p}$ is  $d$ . Set $K:=k(t_{1},\ldots ,t_{d-1})$ , the field of rational functions in indeterminates $t_{1},\ldots ,t_{d-1}$ and

$$\begin{eqnarray}b_{i}:=a_{i}-a_{0}t_{i}\quad \text{for}~i=1,\ldots ,d-1\end{eqnarray}$$

viewed as elements in $H^{\ast }(G_{K},K)$ . Let $\mathfrak{p}^{\prime }$ be the extension of $\mathfrak{p}$ to $H^{\ast }(G_{K},K)$ and

$$\begin{eqnarray}\mathfrak{q}:=\mathfrak{p}^{\prime }+(\boldsymbol{b})\quad \text{and}\quad \mathfrak{m}:=\sqrt{\mathfrak{q}}.\end{eqnarray}$$

It is proved as part of [Reference Benson, Iyengar, Krause and PevtsovaBIKP18, Theorem 7.7] that the ideal $\mathfrak{m}$ is a closed point in $\operatorname{Proj}H^{\ast }(G_{K},K)$ with the property that $\mathfrak{m}\cap H^{\ast }(G,k)=\mathfrak{p}$ . What is more, it follows from the construction (see in particular [Reference Benson, Iyengar, Krause and PevtsovaBIKP18, Lemmas 7.6, and (7.2)]) that the induced extension of fields is an isomorphism

$$\begin{eqnarray}k(\mathfrak{p})\xrightarrow[{}]{~\cong ~}k(\mathfrak{m}).\end{eqnarray}$$

The sequence of elements $\boldsymbol{b}$ in $H^{\ast }(G_{K},K)$ yields a morphism $K\rightarrow \unicode[STIX]{x1D6FA}^{s}(K/\!\!/\boldsymbol{b})$ , where $s=\sum _{i}|b_{i}|$ , and composing its restriction to $G$ with the canonical morphism $k\rightarrow K\downarrow _{G}$ gives in $\mathsf{StMod}\,G$ a morphism

$$\begin{eqnarray}f:k\longrightarrow \unicode[STIX]{x1D6FA}^{s}(K/\!\!/\boldsymbol{b})\downarrow _{G}.\end{eqnarray}$$

Since the $a_{i}$ are not in $\mathfrak{p}$ , Lemma 2.1 yields a natural stable isomorphism

(3.2) $$\begin{eqnarray}\unicode[STIX]{x1D6FA}^{s}M\cong M\end{eqnarray}$$

for any $\mathfrak{p}$ -local $G$ -module $M$ . This remark will be used often in the sequel.

The result below extends [Reference Benson, Iyengar, Krause and PevtsovaBIKP18, Theorem 8.8]; the latter is the case $M=k/\!\!/\mathfrak{p}$ .

Theorem 3.4. For any $G$ -module $M$ , the morphism $M\otimes _{k}f$ induces a natural stable isomorphism of $G$ -modules

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}M\cong M\otimes _{k}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(K/\!\!/\boldsymbol{b})\downarrow _{G}\cong (M_{K}\otimes _{K}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(K/\!\!/\boldsymbol{b}))\downarrow _{G}.\end{eqnarray}$$

When $M$ is $\mathfrak{p}$ -torsion, these induce natural stable isomorphisms

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}M\cong M_{\mathfrak{p}}\cong M\otimes _{k}(K/\!\!/\boldsymbol{b})\downarrow _{G}\cong (M_{K}\otimes _{K}K/\!\!/\boldsymbol{b})\downarrow _{G}.\end{eqnarray}$$

Proof. We begin by verifying the second set of isomorphisms. As $M$ is $\mathfrak{p}$ -torsion so is $M_{\mathfrak{p}}$ and then it is clear that the natural map $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}M=\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p})}M_{\mathfrak{p}}\rightarrow M_{\mathfrak{p}}$ is an isomorphism. The third of the desired isomorphisms follows from (3.1). It thus remains to check that $M\otimes _{k}f$ induces an isomorphism

$$\begin{eqnarray}M_{\mathfrak{p}}\cong M\otimes _{k}(K/\!\!/\boldsymbol{b})\downarrow _{G}.\end{eqnarray}$$

It is easy to verify that the modules $M$ having this property form a tensor ideal localising subcategory of $\mathsf{StMod}\,G$ . Keeping in mind (3.2), from [Reference Benson, Iyengar, Krause and PevtsovaBIKP18, Theorem 8.8] one obtains that this subcategory contains $k/\!\!/\mathfrak{p}$ . The desired assertion follows since the $\mathfrak{p}$ -torsion modules form a tensor ideal localising subcategory of $\mathsf{StMod}\,G$ that is generated by $k/\!\!/\mathfrak{p}$ ; see [Reference Benson, Iyengar and KrauseBIK11a, Proposition 2.7].

Now we turn to the first set of isomorphisms. There the second one holds by (3.1), so we focus on the first. Let $M$ be an arbitrary $G$ -module, and let $\mathfrak{p}^{\prime }$ be as in Construction 3.3. Since $\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p})}M$ is $\mathfrak{p}$ -torsion, the already established isomorphism yields the second one below:

$$\begin{eqnarray}\displaystyle \unicode[STIX]{x1D6E4}_{\mathfrak{p}}M & \cong & \displaystyle (\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p})}M)_{\mathfrak{p}}\nonumber\\ \displaystyle & \cong & \displaystyle ((\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p})}M)_{K}\otimes _{K}K/\!\!/\boldsymbol{b})\downarrow _{G}\nonumber\\ \displaystyle & \cong & \displaystyle (\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p}^{\prime })}(M_{K})\otimes _{K}K/\!\!/\boldsymbol{b})\downarrow _{G}\nonumber\\ \displaystyle & \cong & \displaystyle (M_{K}\otimes _{K}\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p}^{\prime })}(K/\!\!/\boldsymbol{b}))\downarrow _{G}\nonumber\\ \displaystyle & \cong & \displaystyle (M_{K}\otimes _{K}\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p}^{\prime }+(\boldsymbol{b}))}(K/\!\!/\boldsymbol{b}))\downarrow _{G}\nonumber\\ \displaystyle & \cong & \displaystyle (M_{K}\otimes _{K}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(K/\!\!/\boldsymbol{b}))\downarrow _{G}.\nonumber\end{eqnarray}$$

The third one holds by Lemma 2.2, applied to the functor $K\otimes _{k}(-)$ from $\mathsf{StMod}\,G$ to $\mathsf{StMod}\,G_{K}$ . The next one is standard, the penultimate one holds as $K/\!\!/\boldsymbol{b}$ is $(\boldsymbol{b})$ -torsion, and the last follows from $\sqrt{\mathfrak{p}^{\prime }+(\boldsymbol{b})}=\mathfrak{m}$ . This completes the proof.◻

In the next remark we recast part of Theorem 3.4.

Remark 3.5. Fix a point $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ , and let $K$ , $\boldsymbol{b}$ and $\mathfrak{m}$ be as in Construction 3.3. Consider the following adjoint pair of functors.

$$\begin{eqnarray}\begin{array}{@{}ccc@{}}\unicode[STIX]{x1D706}:\mathsf{StMod}\,G\longrightarrow \mathsf{StMod}\,G_{K} & \text{and} & \unicode[STIX]{x1D70C}:\mathsf{StMod}\,G_{K}\longrightarrow \mathsf{StMod}\,G\\ \unicode[STIX]{x1D706}(M)=M_{K}\otimes _{K}K/\!\!/\boldsymbol{b} & & \unicode[STIX]{x1D70C}(N)=\operatorname{Hom}_{K}(K/\!\!/\boldsymbol{b},N)\downarrow _{G}\end{array}\end{eqnarray}$$

It is easy to check that this induces an adjoint pair

Theorem 3.4 implies that $(\unicode[STIX]{x1D706}M)\downarrow _{G}\cong M$ for any $M$ in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ .

Proof of Theorem 3.1.

Let $\mathfrak{m}$ , $\mathfrak{q}$ , and $\boldsymbol{b}$ be as in Construction 3.3. As noted there, $\mathfrak{m}$ is a closed point in $\operatorname{Proj}H^{\ast }(G_{K},K)$ lying over $\mathfrak{p}$ and $k(\mathfrak{m})\cong k(\mathfrak{p})$ . The modules in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ are precisely those with support contained in $\{\mathfrak{p}\}$ . It then follows from [Reference Benson, Iyengar, Krause and PevtsovaBIKP18, Proposition 6.2] that $(-)\downarrow _{G}$ restricts to a functor

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})\longrightarrow \unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G).\end{eqnarray}$$

This functor is dense because for any $G$ -module $M$ that is $\mathfrak{p}$ -local and $\mathfrak{p}$ -torsion one has $M\cong (\unicode[STIX]{x1D706}M)\downarrow _{G}$ where $\unicode[STIX]{x1D706}$ is the functor from Remark 3.5.

Consider the restriction of $(-)\downarrow _{G}$ to compact objects in $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})$ . First we verify that its image is contained in the compact objects of $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ . To this end, it suffices to check that there exists a generator of $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})^{\mathsf{c}}$ , as a thick subcategory, whose restriction is in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)^{\mathsf{c}}$ .

Let $S$ be the direct sum of a representative set of simple $G$ -modules. Each simple $G_{K}$ -module is (isomorphic to) a direct summand of  $S_{K}$ , so from (2.5) one gets the first equality below:

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})^{\mathsf{c}}=\mathsf{Thick}(S_{K}/\!\!/\mathfrak{m})=\mathsf{Thick}(S_{K}/\!\!/\mathfrak{q}).\end{eqnarray}$$

The second one holds by (2.4), since $\sqrt{\mathfrak{q}}=\mathfrak{m}$ . From Theorem 3.4 one gets isomorphisms of $G$ -modules

$$\begin{eqnarray}(S_{K}/\!\!/\mathfrak{q})\downarrow _{G}\cong ((S/\!\!/\mathfrak{p})_{K}\otimes _{K}K/\!\!/\boldsymbol{b})\downarrow _{G}\cong (S/\!\!/\mathfrak{p})_{\mathfrak{p}}.\end{eqnarray}$$

It remains to note that $(S/\!\!/\mathfrak{p})_{\mathfrak{p}}$ is in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)^{\mathsf{c}}$ , again by (2.5).

The last item to verify is that restriction is dense also on compacts. Since $K/\!\!/\boldsymbol{b}$ is compact, the functor $\unicode[STIX]{x1D70C}$ from Remark 3.5 preserves coproducts, and hence its left adjoint $\unicode[STIX]{x1D706}$ preserves compactness. Thus Theorem 3.4 gives the desired result.◻

Theorem 3.4 yields that $f\cong (f_{K})\downarrow _{G}$ for any morphism $f$ in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ ; in particular, the restriction functor is full and dense on the subcategory of $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,G_{K})$ consisting of objects of the form $\unicode[STIX]{x1D706}M$ , where $M$ is a $\mathfrak{p}$ -local $\mathfrak{p}$ -torsion $G$ -module. It need not be full on the entire category, or even on its subcategory of compact objects; see Example 3.7, modelled on the following example from commutative algebra.

Example 3.6. Let $k$ be a field and $k[a]$ the polynomial ring in an indeterminate  $a$ . Let $\mathsf{D}(k[a])$ denote its derived category; it is $k[a]$ -linear in an obvious way. For the prime $\mathfrak{p}:=(0)$ of $k[a]$ the $\mathfrak{p}$ -local $\mathfrak{p}$ -torsion subcategory $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{D}(k[a]))$ is naturally identified with the derived category of $k(a)$ , the field of rational functions in  $a$ .

With $k(t)$ denoting the field of rational functions in an indeterminate  $t$ , the maximal ideal $\mathfrak{m}:=(a-t)$ of $k(t)[a]$ lies over the prime ideal $\mathfrak{p}$ of $k[a]$ . The inclusion $k[a]\subset k(t)[a]$ induces an isomorphism $k(a)\cong k(t)[a]/\mathfrak{m}\cong k(t)$ . The analogue of Theorem 3.4 is that restriction of scalars along the inclusion $k[a]\subset k(t)[a]$ induces a dense functor

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{D}(k(t)[a]))\longrightarrow \unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{D}(k[a]))\simeq \mathsf{D}(k(a)).\end{eqnarray}$$

This property can be checked directly: the $\mathfrak{m}$ -torsion module $k(t)[a]/(a-t)$ restricts to $k(a)$ , and each object in $\mathsf{D}(k(a))$ is a direct sum of shifts of $k(a)$ . This functor is, however, not full: for $n\geqslant 1$ , the $k(t)[a]$ -module $L:=k(t)[a]/(a-t)^{n}$ is $\mathfrak{m}$ -torsion, and satisfies

$$\begin{eqnarray}\operatorname{rank}_{k(a)}\operatorname{End}_{\mathsf{D}}(L)=n\quad \text{and}\quad \operatorname{rank}_{k(a)}\operatorname{End}_{\mathsf{D}}(L\downarrow _{k[a]})=n^{2},\end{eqnarray}$$

where $\mathsf{D}$ stands for the appropriate derived category. In particular, if $n\geqslant 2$ , the canonical map $\operatorname{End}_{\mathsf{D}}(L)\rightarrow \operatorname{End}_{\mathsf{D}}(L\downarrow _{k[a]})$ is not surjective.

Indeed, the module of endomorphisms of $L$ in $\mathsf{D}(k(t)[a])$ is

$$\begin{eqnarray}\operatorname{End}_{\mathsf{D}}(L)=\operatorname{Hom}_{k(t)[a]}(L,L)\cong L.\end{eqnarray}$$

In particular, it has rank $n$ as an $k(a)$ -vector space. On the other hand, restricted to $k[a]$ , the $k(t)[a]$ -module $k(t)/(a-t)$ is isomorphic to $k(a)$ . It then follows from the exact sequences

$$\begin{eqnarray}0\longrightarrow \frac{k(t)[a]}{(a-t)}\xrightarrow[{}]{1\mapsto (a-t)^{i}}\frac{k(t)[a]}{(a-t)^{i+1}}\longrightarrow \frac{k(t)[a]}{(a-t)^{i}}\longrightarrow 0\end{eqnarray}$$

of $k(t)[a]$ -modules that $L$ restricts to a direct sum of $n$ copies of $k(a)$ , so that

$$\begin{eqnarray}\operatorname{End}_{\mathsf{D}}(L\downarrow _{k[a]})=\operatorname{Hom}_{k(a)}(k(a)^{n},k(a)^{n})\cong k(a)^{n^{2}}.\end{eqnarray}$$

In particular, this has rank $n^{2}$ as a $k(a)$ -vector space.

Example 3.7. Let $V=\mathbb{Z}/2\times \mathbb{Z}/2$ and $k$ a field of characteristic two. As $k$ -algebras, one has $H^{\ast }(V,k)\cong k[a,b]$ , where $a$ and $b$ are indeterminates of degree one. For the prime ideal $\mathfrak{p}=(0)$ of $k[a,b]$ , Construction 3.3 leads to the field extension $K:=k(t)$ of  $k$ , and the closed point $\mathfrak{m}=(b-at)$ of $\operatorname{Proj}H^{\ast }(V_{K},K)$ .

Set $F:=\text{}\underline{\operatorname{End}}_{V}(k_{\mathfrak{p}})$ ; this is the component in degree $0$ of the graded field $k[a,b]_{\mathfrak{p}}$ and can be identified with  $K$ ; see Construction 3.3.

Fix an integer $n\geqslant 1$ and set $N:=K/\!\!/(b-at)^{n}$ . This is a finite-dimensional $\mathfrak{m}$ -torsion $V_{K}$ -module and hence compact in $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\mathsf{StMod}\,V_{K})$ . We claim that

$$\begin{eqnarray}\operatorname{rank}_{F}\text{}\underline{\operatorname{End}}_{V_{K}}(N)=2n\quad \text{and}\quad \operatorname{rank}_{F}\text{}\underline{\operatorname{End}}_{V}(N\downarrow _{V})=n^{2},\end{eqnarray}$$

and hence that the map $\text{}\underline{\operatorname{End}}_{V_{K}}(N)\rightarrow \text{}\underline{\operatorname{End}}_{V}(N\downarrow _{V})$ is not surjective when $n\geqslant 3$ .

The claim can be checked as follows. Set $S:=\text{}\underline{\operatorname{End}}_{V_{K}}^{\ast }(K)_{\mathfrak{m}}\cong K[a,b]_{\mathfrak{m}}$ . Since $(b-at)^{n}$ is not a zero divisor on  $S$ , applying $\text{}\underline{\operatorname{Hom}}_{V_{K}}(K,-)$ to the exact triangle

$$\begin{eqnarray}K\xrightarrow[{}]{(b-at)^{n}}\unicode[STIX]{x1D6FA}^{-n}K\longrightarrow N\longrightarrow\end{eqnarray}$$

one gets that $\text{}\underline{\operatorname{Hom}}_{V_{K}}^{\ast }(K,N)$ is isomorphic to $S/(b-at)^{n}$ , as an $S$ -module; in particular $(b-at)^{n}$ annihilates it. Given this, applying $\text{}\underline{\operatorname{Hom}}_{V_{K}}(-,N)$ to the exact triangle above yields that the rank of $\text{}\underline{\operatorname{End}}_{V_{K}}(N)$ , as an $F$ -vector space, is  $2n$ .

As to the claim about $N\downarrow _{V}$ : the category $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,V)$ is semi-simple for its generator $k_{\mathfrak{p}}$ has the property that $\text{}\underline{\operatorname{End}}_{V}^{\ast }(k_{\mathfrak{p}})$ is a graded field. It thus suffices to verify that $N\downarrow _{V}\cong k_{\mathfrak{p}}^{n}$ ; equivalently, that $\operatorname{rank}_{F}\text{}\underline{\operatorname{Hom}}_{V}(k_{\mathfrak{p}},N\downarrow _{V})=n$ . This follows from the isomorphisms

$$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{V}(k_{\mathfrak{p}},N\downarrow _{V})\cong \text{}\underline{\operatorname{Hom}}_{V}(k,N\downarrow _{V})\cong \text{}\underline{\operatorname{Hom}}_{V_{K}}(K,N)\cong F^{n}.\end{eqnarray}$$

The first isomorphism holds because $N\downarrow _{V}$ is $\mathfrak{p}$ -local, the second one holds by adjunction.

There is a close connection between this example and Example 3.6. Namely, the Bernstein–Gelfand–Gelfand (BGG) correspondence sets up an equivalence between $\mathsf{StMod}\,V$ and the derived category of dg modules over $R:=k[a,b]$ , viewed as a dg algebra with zero differential, modulo the subcategory of $(a,b)$ -torsion dg modules; see [Reference Bernstein, Gelfand and GelfandBGG78] and also [Reference Benson, Iyengar and KrauseBIK12, § 5.2.2]. The BGG correspondence induces the equivalences in the following commutative diagram of categories:

where $\mathsf{D}(-)$ denotes the derived category of dg modules. The functor on the right is restriction of scalars along the homomorphism of rings $R_{\mathfrak{p}}\rightarrow S$ , which is induced by the inclusion $R=k[a,b]\subset K[a,b]$ . Under the BGG equivalence, the $V_{K}$ -module $N$ corresponds to $S/(b-at)^{n}$ , viewed as dg $S$ -module with zero differential. Since $R_{\mathfrak{p}}$ is a graded field, isomorphic to $K[a^{\pm 1}]$ , each dg $R_{\mathfrak{p}}$ -module is isomorphic to a direct sum of copies of $R_{\mathfrak{p}}$ . Arguing as in Example 3.6 one can verify that the dg $S$ -module $S/(b-at)^{n}$ restricts to a direct sum of $n$ copies of $R_{\mathfrak{p}}$ . This is another way to compute the endomorphism rings in question.

The remainder of this section is devoted to a further discussion of the compact objects in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ . This is not needed subsequently.

Endofiniteness

Following Crawley-Boevey [Reference Crawley-BoeveyCra91, Reference Crawley-BoeveyCra92], a module $X$ over an associative ring $A$ is endofinite if $X$ has finite length as a module over $\operatorname{End}_{A}(X)$ .

An object $X$ of a compactly generated triangulated category $\mathsf{T}$ is endofinite if the $\operatorname{End}_{\mathsf{T}}(X)$ -module $\operatorname{Hom}_{\mathsf{T}}(C,X)$ has finite length for all $C\in \mathsf{T}^{\mathsf{c}}$ ; see [Reference Krause and ReichenbachKR00].

Let $A$ be a self-injective algebra, finite dimensional over some field. Then an $A$ -module is endofinite if and only if it is endofinite as an object of $\mathsf{StMod}\,A$ . This follows from the fact $X$ is an endofinite $A$ -module if and only if the $\operatorname{End}_{A}(X)$ -module $\operatorname{Hom}_{A}(C,X)$ has finite length for every finite-dimensional $A$ -module  $C$ .

Lemma 3.8. Let $F:\mathsf{T}\rightarrow \mathsf{U}$ be a functor between compactly generated triangulated categories that preserves products and coproducts. Let $X$ be an object in  $\mathsf{T}$ . If $X$ is endofinite, then so is $FX$ and the converse holds when $F$ is fully faithful.

Proof. By Brown representability, $F$ has a left adjoint, say $F^{\prime }$ . It preserves compactness, as $F$ preserves coproducts. For $X\in \mathsf{T}$ and $C\in \mathsf{U}^{\mathsf{c}}$ , there is an isomorphism

$$\begin{eqnarray}\operatorname{Hom}_{\mathsf{U}}(C,FX)\cong \operatorname{Hom}_{\mathsf{T}}(F^{\prime }C,X)\end{eqnarray}$$

of $\operatorname{End}_{\mathsf{T}}(X)$ -modules. Thus if $X$ is endofinite, then $\operatorname{Hom}_{\mathsf{U}}(C,FX)$ is a module of finite length over $\operatorname{End}_{\mathsf{T}}(X)$ , and therefore also over $\operatorname{End}_{\mathsf{U}}(FX)$ . For the converse, observe that each compact object in $\mathsf{T}$ is isomorphic to a direct summand of an object of the form $F^{\prime }C$ for some $C\in \mathsf{U}^{\mathsf{c}}$ .◻

Proposition 3.9. Let $\mathfrak{p}$ be a point in $\operatorname{Proj}H^{\ast }(G,k)$ and $M$ a $G$ -module that is compact in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ . Then $M$ is endofinite both in $\mathsf{StMod}\,G$ and in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ .

Proof. By Corollary 3.2, the module $M$ is of the form $N\downarrow _{G}$ for a finite-dimensional $G_{K}$ -module  $N$ . Clearly, $N$ is endofinite in $\mathsf{StMod}\,G_{K}$ and $(-)\downarrow _{G}$ preserves products and coproducts, so it follows by Lemma 3.8 that $M$ is endofinite in $\mathsf{StMod}\,G$ . By the same token, as the inclusion $(\mathsf{StMod}\,G)_{\mathfrak{p}}\rightarrow \mathsf{StMod}\,G$ preserves products and coproducts, $M$ is endofinite in $(\mathsf{StMod}\,G)_{\mathfrak{p}}$ as well. Finally, the functor $\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p})}$ is a right adjoint to the inclusion $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)\rightarrow (\mathsf{StMod}\,G)_{\mathfrak{p}}$ . It preserves products, being a right adjoint, and also coproducts. Thus $M$ is endofinite in $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ , again by Lemma 3.8.◻

4 $G$ -modules and Tate duality

Now we turn to various dualities for modules over finite group schemes. We begin by recalling the construction of the transpose and the dual of a module over a finite group scheme, and certain functors associated with them. Our basic reference for this material is Skowroński and Yamagata [Reference Skowroński and YamagataSY11, ch. III].

Throughout $G$ will be a finite group scheme over $k$ . We write $(-)^{\vee }=\operatorname{Hom}_{k}(-,k)$ .

Transpose and dual

Let $G^{\text{op}}$ be the opposite group scheme of $G$ ; it can be realised as the group scheme associated to the cocommutative Hopf algebra $(kG)^{\text{op}}$ . Since $kG$ is a $G$ -bimodule, the assignment $M\mapsto \operatorname{Hom}_{G}(M,kG)$ defines a functor

$$\begin{eqnarray}(-)^{t}:\mathsf{Mod}\,G\longrightarrow \mathsf{Mod}\,G^{\text{op}}.\end{eqnarray}$$

Let $M$ be a finite-dimensional $G$ -module and $P_{1}\xrightarrow[{}]{f}P_{0}\rightarrow M$ a minimal projective presentation. The transpose of $M$ is the $G^{\text{op}}$ -module $\operatorname{Tr}M:=\operatorname{Coker}(f^{t})$ . By construction, there is an exact sequence of $G^{\text{op}}$ -modules:

$$\begin{eqnarray}0\longrightarrow M^{t}\longrightarrow P_{0}^{t}\xrightarrow[{}]{~f^{t}~}P_{1}^{t}\longrightarrow \operatorname{Tr}M\longrightarrow 0.\end{eqnarray}$$

The $P_{i}^{t}$ are projective $G^{\text{op}}$ -modules, so this yields an isomorphism of $G^{\text{op}}$ -modules

$$\begin{eqnarray}M^{t}\cong \unicode[STIX]{x1D6FA}^{2}\operatorname{Tr}M.\end{eqnarray}$$

Given a $G^{\text{op}}$ -module $N$ , the $k$ -vector space $\operatorname{Hom}_{k}(N,k)$ has a natural structure of a $G$ -module, and the assignment $N\mapsto \operatorname{Hom}_{k}(N,k)$ yields a functor

$$\begin{eqnarray}D:=\operatorname{Hom}_{k}(-,k):\mathsf{stmod}\,G^{\text{op}}\longrightarrow \mathsf{stmod}\,G.\end{eqnarray}$$

The Auslander–Reiten translate

In what follows we write $\unicode[STIX]{x1D70F}$ for the Auslander–Reiten translate of  $G$ :

$$\begin{eqnarray}\unicode[STIX]{x1D70F}:=D\circ \operatorname{Tr}:\mathsf{stmod}\,G\rightarrow \mathsf{stmod}\,G.\end{eqnarray}$$

Given an extension of fields $K/k$ , for any finite-dimensional $G$ -module $M$ there is a stable isomorphism of $G_{K}$ -modules

$$\begin{eqnarray}(\unicode[STIX]{x1D70F}M)_{K}\cong \unicode[STIX]{x1D70F}(M_{K}).\end{eqnarray}$$

Nakayama functor

The Nakayama functor

$$\begin{eqnarray}\unicode[STIX]{x1D708}:\mathsf{Mod}\,G\xrightarrow[{}]{~\sim ~}\mathsf{Mod}\,G\end{eqnarray}$$

is given by the assignment

$$\begin{eqnarray}M\mapsto D(kG)\otimes _{kG}M\cong \unicode[STIX]{x1D6FF}_{G}\otimes _{k}M,\end{eqnarray}$$

where $\unicode[STIX]{x1D6FF}_{G}=\unicode[STIX]{x1D708}(k)$ is the modular character of $G$ ; see [Reference JantzenJan03, I.8.8]. Since the group of characters of $G$ is finite, by [Reference WaterhouseWat79, §§ 2.1 and 2.2], there exists a positive integer $d$ such that $\unicode[STIX]{x1D6FF}_{G}^{\otimes d}\cong k$ and hence as functors on $\mathsf{Mod}\,G$ there is an equality

(4.1) $$\begin{eqnarray}\unicode[STIX]{x1D708}^{d}=\text{id}.\end{eqnarray}$$

When $M$ is a finite-dimensional $G$ -module, there are natural stable isomorphisms

$$\begin{eqnarray}\unicode[STIX]{x1D708}M\cong D(M^{t})\cong \unicode[STIX]{x1D6FA}^{-2}\unicode[STIX]{x1D70F}M.\end{eqnarray}$$

When in addition $M$ is projective, one has

$$\begin{eqnarray}\operatorname{Hom}_{G}(M,-)^{\vee }\cong (M^{t}\otimes _{kG}-)^{\vee }\cong \operatorname{Hom}_{G}(-,\unicode[STIX]{x1D708}M).\end{eqnarray}$$

Let $K/k$ be an extension of fields. For any $G$ -module $M$ there is a natural isomorphism of $G_{K}$ -modules

(4.2) $$\begin{eqnarray}(\unicode[STIX]{x1D708}M)_{K}\cong \unicode[STIX]{x1D708}(M_{K}).\end{eqnarray}$$

This is clear for $M=kG$ since

$$\begin{eqnarray}K\otimes _{k}\operatorname{Hom}_{k}(kG,k)\cong \operatorname{Hom}_{k}(kG,K)\cong \operatorname{Hom}_{K}(K\otimes _{k}kG,K),\end{eqnarray}$$

and the general case follows by taking a free presentation of  $M$ .

Remark 4.1. A finite group scheme is unimodular if the character $\unicode[STIX]{x1D6FF}_{G}$ is trivial; equivalently, when $kG$ is symmetric. Examples include finite groups, unipotent groups schemes, and Frobenius kernels of reductive groups; see [Reference JantzenJan03, I.8.9, II.3.4(a)]. Group schemes that are not unimodular also abound.

Frobenius kernels of Borel subgroups of reductive groups are not unimodular for $p\geqslant 3$ ; see [Reference JantzenJan03, II.3.4(c)]. The finite group scheme associated to a Lie algebra is unimodular if and only if $\text{tr}(\text{ad}\,x)=0$ for any $x$ in the Lie algebra; see [Reference JantzenJan03, I.9.7]. This condition fails for the upper triangular matrices inside $\text{sl}_{2}$ for $p\geqslant 3$ .

Tate duality

For finite groups, the duality theorem below is classical and due to Tate [Reference Cartan and EilenbergCE56, ch. XII, Theorem 6.4]. A proof of the extension to finite group schemes was sketched in [Reference Benson, Iyengar, Krause and PevtsovaBIKP17, § 2], and is reproduced here for readers convenience.

Theorem 4.2. Let $G$ be a finite group scheme over a field $k$ . For any $G$ -modules $M,N$ with $M$ finite dimensional, there are natural isomorphisms

$$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{G}(M,N)^{\vee }\cong \text{}\underline{\operatorname{Hom}}_{G}(N,\unicode[STIX]{x1D6FA}^{-1}\unicode[STIX]{x1D70F}M)\cong \text{}\underline{\operatorname{Hom}}_{G}(N,\unicode[STIX]{x1D6FA}\unicode[STIX]{x1D708}M).\end{eqnarray}$$

Proof. A formula of Auslander and Reiten [Reference AuslanderAus78, Proposition I.3.4], see also [Reference KrauseKra03, Corollary p. 269], yields the first isomorphism below

$$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{G}(M,N)^{\vee }\cong \operatorname{Ext}_{G}^{1}(N,\unicode[STIX]{x1D70F}M)\cong \text{}\underline{\operatorname{Hom}}_{G}(N,\unicode[STIX]{x1D6FA}^{-1}\unicode[STIX]{x1D70F}M).\end{eqnarray}$$

The second isomorphism is standard. It remains to recall that $\unicode[STIX]{x1D70F}M\cong \unicode[STIX]{x1D6FA}^{2}\unicode[STIX]{x1D708}M$ .◻

Restricted to finite-dimensional $G$ -modules, Tate duality is the statement that the $k$ -linear category $\mathsf{stmod}\,G$ has Serre duality, with Serre functor $\unicode[STIX]{x1D6FA}\unicode[STIX]{x1D708}$ . A refinement of this Serre duality will be proved in § 7.

5 Local cohomology versus injective cohomology

Let $k$ be a field and $G$ a finite group scheme over $k$ . In this section we establish the main result of this work; it identifies for a prime ideal $\mathfrak{p}$ in $H^{\ast }(G,k)$ , up to some twist and some suspension, the local cohomology object $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k)$ with the injective cohomology object $T_{\mathfrak{p}}(k)$ .

Theorem 5.1. Fix a point $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ and let $d$ be the Krull dimension of $H^{\ast }(G,k)/\mathfrak{p}$ . There is a stable isomorphism of $G$ -modules

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D6FF}_{G})\cong \unicode[STIX]{x1D6FA}^{-d}T_{\mathfrak{ p}}(k);\end{eqnarray}$$

equivalently, for any $G$ -module $M$ there is a natural isomorphism

$$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{G}(M,\unicode[STIX]{x1D6FA}^{d}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D6FF}_{G}))\cong \operatorname{Hom}_{H^{\ast }(G,k)}(H^{\ast }(G,M),I(\mathfrak{p})).\end{eqnarray}$$

When $G$ is the group scheme arising from a finite group the modular character $\unicode[STIX]{x1D6FF}_{G}$ is trivial, and the result above was proved by Benson and Greenlees [Reference Benson and GreenleesBG08, Theorem 2.4] using Gorenstein duality for cochains on the classifying space of  $G$ . Benson [Reference BensonBen08, Theorem 2] gave a different proof by embedding $G$ into a general linear group and exploiting the fact that its cohomology ring is a polynomial ring, as was proved by Quillen. These results have been extended to compact Lie groups; see [Reference Benson and GreenleesBG14, Theorem 6.10], and work of Barthel, Heard and Valenzuela [Reference Barthel, Heard and ValenzuelaBHV18, Proposition 4.33].

Theorem 5.1 is established using (by necessity) completely different arguments, thereby giving yet another proof in the case of finite groups that is, in a sense, more elementary than the other ones for it is based on classical Tate duality.

A caveat: in [Reference BensonBen08, Reference Benson and GreenleesBG08] it is asserted that $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k)\cong \unicode[STIX]{x1D6FA}^{d}T_{\mathfrak{p}}(k)$ . However, this is incorrect and the correct shift is the one in the preceding theorem. We illustrate this by computing these modules directly for the quaternions.

Example 5.2. Let $G:=Q_{8}$ , the quaternions, viewed as a group scheme over a field $k$ of characteristic $2$ . In this case $\unicode[STIX]{x1D6FF}_{G}=k$ , the trivial character. The cohomology algebra of $G$ is

$$\begin{eqnarray}H^{\ast }(G,k)=k[z]\otimes _{k}B\quad \text{where}~B=k[x,y]/(x^{2}+xy+y^{2},x^{2}y+xy^{2}),\end{eqnarray}$$

with $|x|=1=|y|$ and $|z|=4$ ; see, for instance, [Reference BensonBen84, p. 186]. Thus $\operatorname{Proj}H^{\ast }(G,k)$ consists of a single point, namely $\mathfrak{m}:=(x,y)$ . In particular, $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}k=k$ , in $\mathsf{StMod}\,G$ .

Next we compute $I(\mathfrak{m})$ as a module over $H^{\ast }(G,k)_{\mathfrak{m}}\cong k[z^{\pm 1}]\otimes _{k}B$ , using Lemma A.3. The extension $k[z^{\pm 1}]\subseteq k[z^{\pm 1}]\otimes _{k}B$ is evidently finite (and hence also residually finite). Since $\mathfrak{m}\cap k[z^{\pm 1}]=(0)$ and $k[z^{\pm 1}]$ is a graded field, from Lemma A.3 one gets an isomorphism of $H^{\ast }(G,k)_{\mathfrak{m}}$ -modules,

$$\begin{eqnarray}\displaystyle I(\mathfrak{m}) & \cong & \displaystyle \operatorname{Hom}_{k[z^{\pm 1}]}(k[z^{\pm 1}]\otimes _{k}B,k[z^{\pm 1}])\nonumber\\ \displaystyle & \cong & \displaystyle k[z^{\pm 1}]\otimes _{k}\operatorname{Hom}_{k}(B,k)\nonumber\\ \displaystyle & \cong & \displaystyle k[z^{\pm 1}]\otimes _{k}\unicode[STIX]{x1D6F4}^{3}B\nonumber\\ \displaystyle & \cong & \displaystyle \unicode[STIX]{x1D6F4}^{3}H^{\ast }(G,k)_{\mathfrak{m}}.\nonumber\end{eqnarray}$$

This yields the first isomorphism below of $G$ -modules

$$\begin{eqnarray}T_{\mathfrak{m}}(k)\cong \unicode[STIX]{x1D6FA}^{-3}k\cong \unicode[STIX]{x1D6FA}^{1}k,\end{eqnarray}$$

and the second one holds because $\unicode[STIX]{x1D6FA}^{4}k\cong k$ in $\mathsf{StMod}\,G$ .

In the proof of Theorem 5.1 the following simple observation will be used repeatedly; it is a direct consequence of Yoneda’s lemma.

Lemma 5.3. Let $X$ and $Y$ be $G$ -modules that are $\mathfrak{p}$ -local and $\mathfrak{p}$ -torsion. There is an isomorphism $X\cong Y$ in $\mathsf{StMod}\,G$ if and only if there is a natural isomorphism

$$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{G}(M,X)\cong \text{}\underline{\operatorname{Hom}}_{G}(M,Y)\end{eqnarray}$$

for all $\mathfrak{p}$ -local and $\mathfrak{p}$ -torsion (equivalently, for all) $G$ -modules  $M$ .

Proof of Theorem 5.1.

The $G$ -module $T_{\mathfrak{p}}(k)$ is $\mathfrak{p}$ -local and $\mathfrak{p}$ -torsion; see, for example, [Reference Benson, Iyengar and KrauseBIK11b, Lemma 11.10]. This fact will be used in the sequel, without comment.

The first isomorphism of the theorem is equivalent to the second: by (2.7) for any $\mathfrak{p}$ -local $H^{\ast }(G,k)$ -module $I$ there is an isomorphism

$$\begin{eqnarray}\operatorname{Hom}_{H^{\ast }(G,k)}(H^{\ast }(G,M),I)\cong \operatorname{Hom}_{H^{\ast }(G,k)}(\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,M),I).\end{eqnarray}$$

Consequently, one can rephrase the defining isomorphism (2.1) for the object $T_{\mathfrak{p}}(k)$ as a natural isomorphism

$$\begin{eqnarray}\text{}\underline{\operatorname{Hom}}_{G}(M,T_{\mathfrak{p}}(k))\cong \operatorname{Hom}_{H^{\ast }(G,k)}(H^{\ast }(G,M),I(\mathfrak{p})).\end{eqnarray}$$

The desired equivalence is then a consequence of Lemma 5.3.

The main task is to prove that there is a stable isomorphism:

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D708}k)\cong \unicode[STIX]{x1D6FA}^{-d}T_{\mathfrak{ p}}(k).\end{eqnarray}$$

Recall that $\unicode[STIX]{x1D708}k=\unicode[STIX]{x1D6FF}_{G}$ .

We first verify the isomorphism above for closed points of $\operatorname{Proj}H^{\ast }(G,k)$ and then use a reduction to closed points.

Claim.

The desired isomorphism holds when $\mathfrak{m}$ is a closed point of $\operatorname{Proj}H^{\ast }(G,k)$ .

Set $A:=H^{\ast }(G,k)$ and $R:=A_{\mathfrak{m}}$ . The injective hull, $I(\mathfrak{m})$ , of the $A$ -module $A/\mathfrak{m}$ is the same as that of the $R$ -module $k(\mathfrak{m})$ , viewed as an $A$ -module via restriction of scalars along the localisation map $A\rightarrow R$ . Thus $I(\mathfrak{m})$ is the module $I$ described in Lemma A.2; this is where the fact that $\mathfrak{m}$ is closed is used.

Let $M$ be a $G$ -module that is $\mathfrak{m}$ -local and $\mathfrak{m}$ -torsion. Given Lemma 5.3, the claim is a consequence of the following computation:

$$\begin{eqnarray}\displaystyle \text{}\underline{\operatorname{Hom}}_{G}(M,\unicode[STIX]{x1D6FA}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\unicode[STIX]{x1D708}k)) & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}(M,\unicode[STIX]{x1D6FA}\unicode[STIX]{x1D708}k)\nonumber\\ \displaystyle & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}(k,M)^{\vee }\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{R}(\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,M),I(\mathfrak{m}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{A}(\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,M),I(\mathfrak{m}))\nonumber\\ \displaystyle & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}(M,T_{\mathfrak{m}}(k)).\nonumber\end{eqnarray}$$

The first isomorphism holds because $M$ is $\mathfrak{m}$ -torsion; the second is Tate duality, Theorem 4.2, and the next one holds by Lemma A.2, which applies because $\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,M)$ is $\mathfrak{m}$ -local and $\mathfrak{m}$ -torsion as an $A$ -module. The penultimate one holds because the $A$ -module $\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,M)$ is $\mathfrak{m}$ -local, and the last one is by definition (2.1).

Let $\mathfrak{p}$ be a point in $\operatorname{Proj}H^{\ast }(G,k)$ that is not closed, and let  $K$ , $\boldsymbol{b}$ , and $\mathfrak{m}$ be as in Construction 3.3. Recall that $\mathfrak{m}$ is a closed point in $H^{\ast }(G_{K},K)$ lying over  $\mathfrak{p}$ .

Claim.

There is a stable isomorphism of $G$ -modules

(5.1) $$\begin{eqnarray}(T_{\mathfrak{m}}(K)/\!\!/\boldsymbol{b})\downarrow _{G}\cong \unicode[STIX]{x1D6FA}^{-d+1}T_{\mathfrak{ p}}(k),\end{eqnarray}$$

where $d$ is the Krull dimension of $H^{\ast }(G,k)/\mathfrak{p}$ .

Let $M$ be a $G$ -module that is $\mathfrak{p}$ -local and $\mathfrak{p}$ -torsion. Then Theorem 3.4 applies and yields isomorphisms of $G$ -modules:

$$\begin{eqnarray}(M_{K}/\!\!/\boldsymbol{b})\downarrow _{G}\cong (M_{K}\otimes _{K}K/\!\!/\boldsymbol{b})\downarrow _{G}\cong M.\end{eqnarray}$$

This gives the sixth isomorphism below:

$$\begin{eqnarray}\displaystyle \text{}\underline{\operatorname{Hom}}_{G}(M,\unicode[STIX]{x1D6FA}^{d-1}(T_{\mathfrak{m}}(K)/\!\!/\boldsymbol{b})\downarrow _{G}) & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G_{K}}(M_{K},\unicode[STIX]{x1D6FA}^{d-1}(T_{\mathfrak{m}}(K)/\!\!/\boldsymbol{b}))\nonumber\\ \displaystyle & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G_{K}}(M_{K}/\!\!/\boldsymbol{b},T_{\mathfrak{m}}(K))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{H^{\ast }(G_{K},K)}(\text{}\underline{\operatorname{Hom}}_{G_{K}}^{\ast }(K,M_{K}/\!\!/\boldsymbol{b}),I(\mathfrak{m}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{H^{\ast }(G,k)}(\text{}\underline{\operatorname{Hom}}_{G_{K}}^{\ast }(K,M_{K}/\!\!/\boldsymbol{b}),I(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{H^{\ast }(G,k)}(\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,(M_{K}/\!\!/\boldsymbol{b})\downarrow _{G}),I(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{H^{\ast }(G,k)}(\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,M),I(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}(M,T_{\mathfrak{p}}(k)).\nonumber\end{eqnarray}$$

The first and the fifth isomorphisms are true by adjunction. The second isomorphism is easily verified by a direct computation using (2.3) and (3.2). The next isomorphism is by definition and the fourth one holds by Lemma A.3 applied to the canonical homomorphism $H^{\ast }(G,k)\rightarrow H^{\ast }(G_{K},K)$ ; note that the $H^{\ast }(G_{K},K)$ -module $\text{}\underline{\operatorname{Hom}}_{G_{K}}^{\ast }(K,M_{K}/\!\!/\boldsymbol{b})$ is $\mathfrak{m}$ -torsion. The desired isomorphism (5.1) holds by Lemma 5.3, because both modules in question are $\mathfrak{p}$ -torsion; for the one on the left, see [Reference Benson, Iyengar, Krause and PevtsovaBIKP18, Proposition 6.2].

We are now ready to wrap up the proof of the theorem. Since $\mathfrak{m}$ is a closed point in $\operatorname{Proj}H^{\ast }(G_{K},K)$ , the first claim yields that the $G_{K}$ -modules $\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\unicode[STIX]{x1D708}K)$ and $\unicode[STIX]{x1D6FA}^{-1}T_{\mathfrak{m}}(K)$ are isomorphic. This then gives an isomorphism of $G_{K}$ -modules:

$$\begin{eqnarray}\unicode[STIX]{x1D708}K\otimes _{K}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(K/\!\!/\boldsymbol{b})\cong \unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\unicode[STIX]{x1D708}K)/\!\!/\boldsymbol{b}\cong \unicode[STIX]{x1D6FA}^{-1}T_{\mathfrak{ m}}(K)/\!\!/\boldsymbol{b}.\end{eqnarray}$$

Restricting to $G$ and applying (5.1) gives the last of the following isomorphisms of $G$ -modules:

$$\begin{eqnarray}\displaystyle \unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D708}k) & \cong & \displaystyle ((\unicode[STIX]{x1D708}k)_{K}\otimes _{K}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(K/\!\!/\boldsymbol{b}))\downarrow _{G}\nonumber\\ \displaystyle & \cong & \displaystyle (\unicode[STIX]{x1D708}K\otimes _{K}\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(K/\!\!/\boldsymbol{b}))\downarrow _{G}\nonumber\\ \displaystyle & \cong & \displaystyle \unicode[STIX]{x1D6FA}^{-d}T_{\mathfrak{p}}(k).\nonumber\end{eqnarray}$$

The first one holds by Theorem 3.4 and the second by (4.2). ◻

The following consequence of Theorem 5.1 was anticipated in [Reference BensonBen01, p. 203]. It concerns the Tate cohomology of a $G$ -module  $M$ , namely the $H^{\ast }(G,k)$ -module

$$\begin{eqnarray}\widehat{H}^{\ast }(G,M):=\widehat{\text{Ext}}_{G}^{\ast }(k,M)\cong \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,M).\end{eqnarray}$$

For a graded module $N=\bigoplus _{p\in \mathbb{Z}}N^{p}$ and $i\in \mathbb{Z}$ the twist $N(i)$ is the graded module with $N(i)^{p}=N^{p+i}$ .

Corollary 5.4. Fix $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ . With $d$ the Krull dimension of $H^{\ast }(G,k)/\mathfrak{p}$ there are isomorphisms of $H^{\ast }(G,k)$ -modules

$$\begin{eqnarray}\widehat{H}^{\ast }(G,\unicode[STIX]{x1D6E4}_{\mathfrak{ p}}(k))\cong \operatorname{Hom}_{H^{\ast }(G,k)}^{\ast }(H^{\ast }(G,\unicode[STIX]{x1D6FF}_{G}),I(\mathfrak{p}))(d)\end{eqnarray}$$

and

$$\begin{eqnarray}\widehat{H}^{\ast }(G,\operatorname{End}_{k}(\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k)))\cong (H^{\ast }(G,k)_{\mathfrak{ p}})^{\wedge },\end{eqnarray}$$

where $(-)^{\wedge }$ denotes completion with respect to the $\mathfrak{p}$ -adic topology.

Proof. Set $R=H^{\ast }(G,k)$ . The first of the stated isomorphisms is a composition of the following isomorphisms of $R$ -modules:

$$\begin{eqnarray}\displaystyle \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k)) & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(\unicode[STIX]{x1D6FF}_{G},\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D6FF}_{G}))\nonumber\\ \displaystyle & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(\unicode[STIX]{x1D6FF}_{G},\unicode[STIX]{x1D6FA}^{-d}T_{\mathfrak{p}}(k))\nonumber\\ \displaystyle & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(\unicode[STIX]{x1D6FF}_{G},T_{\mathfrak{p}}(k))(d)\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{R}^{\ast }(H^{\ast }(G,\unicode[STIX]{x1D6FF}_{G}),I(\mathfrak{p}))(d).\nonumber\end{eqnarray}$$

The second isomorphism holds by Theorem 5.1, the one after is standard, while the last one is by the definition of $T_{\mathfrak{p}}(k)$ .

In the same vein, one has the following chain of isomorphisms:

$$\begin{eqnarray}\displaystyle \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,\operatorname{End}_{k}(\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k))) & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k),\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(k))\nonumber\\ \displaystyle & \cong & \displaystyle \text{}\underline{\operatorname{Hom}}_{G}^{\ast }(T_{\mathfrak{p}}(k),T_{\mathfrak{p}}(k))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{R}^{\ast }(\text{}\underline{\operatorname{Hom}}_{G}^{\ast }(k,T_{\mathfrak{p}}(k)),I(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{R}^{\ast }(I(\mathfrak{p}),I(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle (R_{\mathfrak{p}})^{\wedge }.\nonumber\end{eqnarray}$$

The second isomorphism holds by Theorem 5.1 and the rest are standard. ◻

Remark 5.5. Another consequence of Theorem 5.1 is that $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\unicode[STIX]{x1D6FF}_{G})$ , and hence also $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}k$ , is an indecomposable pure injective object in $\mathsf{StMod}\,G$ ; see [Reference Benson and KrauseBK02, Theorem 5.1].

6 The Gorenstein property

In this section we introduce a notion of a Gorenstein triangulated category and reinterpret Theorem 5.1 to mean that $\mathsf{StMod}\,G$ has this property. The definition is justified by its consequences for modular representations; it yields duality results which will be discussed in the subsequent § 7.

Let $R$ be a graded commutative noetherian ring and $\mathsf{T}$ a compactly generated $R$ -linear triangulated category. The support of $\mathsf{T}$ is by definition the set

$$\begin{eqnarray}\operatorname{supp}_{R}(\mathsf{T})=\{\mathfrak{p}\in \operatorname{Spec}R\mid \unicode[STIX]{x1D6E4}_{\mathfrak{p}}\neq 0\}.\end{eqnarray}$$

Definition 6.1. We say that $\mathsf{T}$ is $R$ -Gorenstein (or simply Gorenstein when the action of $R$ is clear) if there is an $R$ -linear triangle equivalence

$$\begin{eqnarray}F:\mathsf{T}^{\mathsf{c}}\xrightarrow[{}]{{\sim}}\mathsf{T}^{\mathsf{c}}\end{eqnarray}$$

and for every $\mathfrak{p}$ in $\operatorname{supp}_{R}(\mathsf{T})$ there is an integer $d(\mathfrak{p})$ and a natural isomorphism

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\circ F\cong \unicode[STIX]{x1D6F4}^{d(\mathfrak{p})}\circ T_{\mathfrak{ p}}\end{eqnarray}$$

of functors $\mathsf{T}^{\mathsf{c}}\rightarrow \mathsf{T}$ .

In this context we call $F$ a global Serre functor, because in Proposition 7.3 we show that localising at $\mathfrak{p}$ induces a functor $\unicode[STIX]{x1D6F4}^{-d(\mathfrak{p})}F_{\mathfrak{p}}$ which is an analogue of a Serre functor in the sense of Bondal and Kapranov [Reference Bondal and KapranovBK89].

We have not as yet explored fully the dependence of the Gorenstein property of $\mathsf{T}$ on the ring  $R$ ; a beginning has been made in the work of Yuliawan [Reference YuliawanYul17].

Let $\mathsf{T}=(\mathsf{T},\otimes ,\unicode[STIX]{x1D7D9})$ be a tensor triangulated category such that $R$ acts on $\mathsf{T}$ via a homomorphism of graded rings $R\rightarrow \operatorname{End}_{\mathsf{T}}^{\ast }(\unicode[STIX]{x1D7D9})$ . We assume that $\unicode[STIX]{x1D7D9}$ is compact and that each compact object is rigid; see § 2 for details. The Gorenstein property is implied by the existence of a dualising object in $\mathsf{T}$ , as explained below.

Lemma 6.2. Let $\mathsf{T}$ be a tensor triangulated category with an $R$ -action. Suppose that there exists a compact object $W$ with the following properties.

  1. (i) There is a compact object $W^{-1}$ such that $W\otimes W^{-1}\cong \unicode[STIX]{x1D7D9}$ .

  2. (ii) For each $\mathfrak{p}$ in $\operatorname{supp}_{R}(\mathsf{T})$ there exists an integer $d(\mathfrak{p})$ and an isomorphism

    $$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}W\cong \unicode[STIX]{x1D6F4}^{d(\mathfrak{p})}T_{\mathfrak{ p}}(\unicode[STIX]{x1D7D9}).\end{eqnarray}$$

Then $\mathsf{T}$ is $R$ -Gorenstein, with global Serre functor $F:=W\otimes -$ .

Proof. Since $W$ is compact, so is $W\otimes C$ for any compact object $C$ of  $\mathsf{T}$ . Thus $F$ induces a functor on compact objects. It is an equivalence of categories with quasi-inverse $W^{-1}\otimes -$ , by condition (i). Moreover for any compact object $C$ and $\mathfrak{p}\in \operatorname{supp}_{R}(\mathsf{T})$ one has isomorphisms

$$\begin{eqnarray}\unicode[STIX]{x1D6F4}^{d(\mathfrak{p})}T_{\mathfrak{ p}}(C)\cong \unicode[STIX]{x1D6F4}^{d(\mathfrak{p})}T_{\mathfrak{ p}}(\unicode[STIX]{x1D7D9})\otimes C\cong \unicode[STIX]{x1D6E4}_{\mathfrak{p}}W\otimes C\cong \unicode[STIX]{x1D6E4}_{\mathfrak{p}}(W\otimes C),\end{eqnarray}$$

where the first and the last one are consequences of (2.2), and the middle one is by (ii). ◻

The example below justifies the language of Gorenstein triangulated categories.

Example 6.3. Let $A$ be a commutative noetherian ring and $\mathsf{D}$ the derived category of $A$ -modules. This is an $A$ -linear compactly generated tensor triangulated category, with compact objects the perfect complexes of $A$ -modules, that is to say, complexes quasi-isomorphic to bounded complexes of finitely generated projective $A$ -modules.

Recall that the ring $A$ is Gorenstein if for each $\mathfrak{p}\in \operatorname{Spec}A$ the injective dimension of  $A_{\mathfrak{p}}$ , as a module over itself, is finite; see [Reference Bruns and HerzogBH98, 3.1]. By Grothendieck’s local duality theorem [Reference Bruns and HerzogBH98, § 3.5], this is equivalent to an isomorphism of $A_{\mathfrak{p}}$ -modules

$$\begin{eqnarray}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}A\cong \unicode[STIX]{x1D6F4}^{-\text{dim}\,A_{\mathfrak{p}}}I(\mathfrak{p}).\end{eqnarray}$$

Thus $\mathsf{D}$ is Gorenstein with dualising object $A$ and $d(\mathfrak{p})=-\text{dim}\,A_{\mathfrak{p}}$ ; see Lemma 6.2. Conversely, it is not difficult to check that $\mathsf{D}$ is Gorenstein only if $A$ is Gorenstein.

For a finite group scheme $G$ over a field $k$ , the Gorenstein property for $\mathsf{StMod}\,G$ is basically a reformulation of Theorem 5.1.

Corollary 6.4. As an $H^{\ast }(G,k)$ -linear triangulated category, $\mathsf{StMod}\,G$ is Gorenstein, with $F=\unicode[STIX]{x1D6FF}_{G}\otimes _{k}-$ the Nakayama functor and $d(\mathfrak{p})=\dim H^{\ast }(G,k)/\mathfrak{p}$ for each $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ .

Next we discuss the Gorenstein property for $\mathsf{K}(\mathsf{Inj}\,G)$ for a finite group scheme  $G$ . To this end observe that the assignment $X\mapsto \unicode[STIX]{x1D708}X$ induces triangle equivalences

$$\begin{eqnarray}\mathsf{K}(\mathsf{Inj}\,G)\xrightarrow[{}]{{\sim}}\mathsf{K}(\mathsf{Inj}\,G)\quad \text{and}\quad \mathsf{D}^{b}(\mathsf{mod}\,G)\xrightarrow[{}]{{\sim}}\mathsf{D}^{b}(\mathsf{mod}\,G).\end{eqnarray}$$

We are ready to establish the Gorenstein property for $\mathsf{K}(\mathsf{Inj}\,G)$ .

Theorem 6.5. Let $G$ be a finite group scheme over a field $k$ . Then $\mathsf{K}(\mathsf{Inj}\,G)$ is Gorenstein as an $H^{\ast }(G,k)$ -linear triangulated, with $F$ induced by the Nakayama functor and $d(\mathfrak{p})=\dim H^{\ast }(G,k)/\mathfrak{p}$ for each $\mathfrak{p}\in \operatorname{Spec}H^{\ast }(G,k)$ .

Proof. Set $R:=H^{\ast }(G,k)$ and $\mathfrak{m}:=R^{{\geqslant}1}$ , the maximal ideal of $R$ . It is easy to verify that $I(\mathfrak{m}):=\operatorname{Hom}_{k}(R,k)$ is the injective hull of $R/\mathfrak{m}$ as an $R$ -module and hence that $\operatorname{Hom}_{R}(-,I(\mathfrak{m}))=\operatorname{Hom}_{k}(-,k)$ on the category of graded $R$ -modules; see [Reference Bruns and HerzogBH98, Proposition 3.6.16]. This observation is used in the first isomorphism below.

Let $\mathsf{p}k$ be a projective resolution of the trivial representation. For any $X$ in $\mathsf{K}(\mathsf{Inj}\,G)$ the complex $\mathsf{p}X:=\mathsf{p}k\otimes _{k}X$ is a projective resolution of  $X$ .

Suppose $X$ is compact in $\mathsf{K}(\mathsf{Inj}\,G)$ ; one may assume $X^{n}=0$ for $n\ll 0$ and that $H^{n}(X)$ is finitely generated for all $n$ and equal to $0$ for $n\gg 0$ ; see [Reference KrauseKra05, Proposition 2.3(2)]. Then, for each $Y$ in $\mathsf{K}(\mathsf{Inj}\,G)$ from [Reference Krause and LeKL06, Theorem 3.4] one gets the first of the following natural isomorphisms:

$$\begin{eqnarray}\displaystyle \operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{K}}^{\ast }(X,Y),I(\mathfrak{m})) & \cong & \displaystyle \operatorname{Hom}_{\mathsf{K}}(Y,D(kG)\otimes _{kG}\mathsf{p}X)\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{\mathsf{K}}(Y,\unicode[STIX]{x1D6FF}_{G}\otimes _{k}\mathsf{p}X)\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{\mathsf{K}}(Y,\mathsf{p}(\unicode[STIX]{x1D6FF}_{G}\otimes _{k}X)).\nonumber\end{eqnarray}$$

The second one holds by the definition of the modular character, and the last one is immediate from the definition of $\mathsf{p}$ and the commutativity of tensor products. Since $\mathsf{p}(\unicode[STIX]{x1D6FF}_{G}\otimes _{k}X)=\unicode[STIX]{x1D6E4}_{\mathfrak{m}}(\unicode[STIX]{x1D6FF}_{G}\otimes _{k}X)$ , by Lemma 2.6, it follows that $T_{\mathfrak{m}}\cong \unicode[STIX]{x1D6E4}_{\mathfrak{m}}\circ F$ .

For a prime ideal $\mathfrak{p}\neq \mathfrak{m}$ , the assertion follows from Theorem 5.1, for localisation at $\mathfrak{p}$ yields a triangle equivalence $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{K}(\mathsf{Inj}\,G))\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ by Lemma 2.6.◻

Remark 6.6. There are extensions of the Gorenstein property to differential graded algebras, see [Reference Frankild, Iyengar and JørgensenFIJ03] and  [Reference Dwyer, Greenlees and IyengarDGI06], for instance, and a natural question is how these relate to the Gorenstein property, in the sense of Definition 6.1, of the associated derived categories. We defer exploring these connections to another occasion.

Balmer, Dell’Ambrogio and Sanders [Reference Balmer, Dell’Ambrogio and SandersBDS16] have introduced a categorical framework extending the duality theory for schemes due to Grothendieck and Neeman. The relationship to our work might be explained thus: a commutative noetherian ring $R$ is Gorenstein precisely when it has an invertible dualising complex. The framework in [Reference Balmer, Dell’Ambrogio and SandersBDS16] captures the relative version (dealing with a morphism of rings, or schemes) of the Gorenstein property and its characterisation in terms of the relative dualising complex. We are interested in the characterisation of the Gorenstein property in terms of local cohomology, and in the fact that when $R$ is Gorenstein, so is $R_{\mathfrak{p}}$ for each prime $\mathfrak{p}$ in $\operatorname{Spec}R$ . Theorem 5.1 may be seen as an analogue of these results for modular representations.

7 Local Serre duality

In this section we introduce a notion of local Serre duality for an essentially small $R$ -linear triangulated category and link it to the Gorenstein property from § 6. We use the concept of a Serre functor for a triangulated category which is due to Bondal and Kapranov [Reference Bondal and KapranovBK89]; this provides a conceptual way to formulate classical Serre duality and Grothendieck’s local duality in a triangulated setting.

In the second part of this section we discuss the existence of AR-triangles. These were introduced by Happel for derived categories of finite-dimensional algebras [Reference HappelHap88], and in [Reference HappelHap91] he established the connection with the Gorenstein property, while Reiten and Van den Bergh [Reference Reiten and Van den BerghRV02] discovered the connection between AR-triangles and the existence of a Serre functor.

Small triangulated categories with central action

Let $\mathsf{C}$ be an essentially small $R$ -linear triangulated category. Fix $\mathfrak{p}\in \operatorname{Spec}R$ and let $\mathsf{C}_{\mathfrak{p}}$ denote the triangulated category obtained from $\mathsf{C}$ by keeping the objects of $\mathsf{C}$ and setting

$$\begin{eqnarray}\operatorname{Hom}_{\mathsf{C}_{\mathfrak{ p}}}^{\ast }(X,Y):=\operatorname{Hom}_{C}^{\ast }(X,Y)_{\mathfrak{ p}}.\end{eqnarray}$$

Then $\mathsf{C}_{\mathfrak{p}}$ is an $R_{\mathfrak{p}}$ -linear triangulated category and localising the morphisms induces an exact functor $\mathsf{C}\rightarrow \mathsf{C}_{\mathfrak{p}}$ ; see [Reference BalmerBal10, Theorem 3.6] or [Reference Benson, Iyengar and KrauseBIK15, Lemma 3.5].

Let $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}$ be the full subcategory of $\mathfrak{p}$ -torsion objects in $\mathsf{C}_{\mathfrak{p}}$ , namely

$$\begin{eqnarray}\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}:=\{X\in \mathsf{C}_{\mathfrak{p}}\mid \operatorname{End}_{\mathsf{C}_{\mathfrak{ p}}}^{\ast }(X)~\text{is}~\mathfrak{p}\text{-}\text{torsion}\}.\end{eqnarray}$$

This is a thick subcategory of $\mathsf{C}_{\mathfrak{p}}$ ; see [Reference Benson, Iyengar and KrauseBIK15, p. 458f]. In [Reference Benson, Iyengar and KrauseBIK15] this category is denoted $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{C}$ . The notation has been changed to avoid confusion.

Remark 7.1. Let $F:\mathsf{C}\rightarrow \mathsf{C}$ be an $R$ -linear equivalence. It is straightforward to check that this induces triangle equivalences $F_{\mathfrak{p}}:\mathsf{C}_{\mathfrak{p}}\xrightarrow[{}]{{\sim}}\mathsf{C}_{\mathfrak{p}}$ and $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}$ making the following diagram commutative.

Remark 7.2. Let $\mathsf{T}$ be a compactly generated $R$ -linear triangulated category. Set $\mathsf{C}:=\mathsf{T}^{\mathsf{c}}$ and fix $\mathfrak{p}\in \operatorname{Spec}R$ . The triangulated categories $\mathsf{T}_{\mathfrak{p}}$ and $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T}$ are compactly generated. The left adjoint of the inclusion $\mathsf{T}_{\mathfrak{p}}\rightarrow \mathsf{T}$ induces (up to direct summands) a triangle equivalence $\mathsf{C}_{\mathfrak{p}}\xrightarrow[{}]{{\sim}}(\mathsf{T}_{\mathfrak{p}})^{\mathsf{c}}$ and restricts to a triangle equivalence

$$\begin{eqnarray}\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}\xrightarrow[{}]{{\sim}}(\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T})^{\mathsf{c}}.\end{eqnarray}$$

This follows from the fact that the localisation functor $\mathsf{T}\rightarrow \mathsf{T}_{\mathfrak{p}}$ taking $X$ to $X_{\mathfrak{p}}$ preserves compactness and that for compact objects $X,Y$ in $\mathsf{T}$

$$\begin{eqnarray}\operatorname{Hom}_{\mathsf{T}}^{\ast }(X,Y)_{\mathfrak{ p}}\xrightarrow[{}]{{\sim}}\operatorname{Hom}_{\mathsf{T}_{\mathfrak{p}}}^{\ast }(X_{\mathfrak{ p}},Y_{\mathfrak{p}}).\end{eqnarray}$$

For details we refer to [Reference Benson, Iyengar and KrauseBIK08].

Local Serre duality

Let $R$ be a graded commutative ring that is local; thus there is a unique homogeneous maximal ideal, say  $\mathfrak{m}$ . Extrapolating from Bondal and Kapranov [Reference Bondal and KapranovBK89, § 3], we call an $R$ -linear triangle equivalence $F:\mathsf{C}\xrightarrow[{}]{{\sim}}\mathsf{C}$ a Serre functor if for all objects $X,Y$ in $\mathsf{C}$ there is a natural isomorphism

(7.1) $$\begin{eqnarray}\operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{C}}^{\ast }(X,Y),I(\mathfrak{m}))\xrightarrow[{}]{{\sim}}\operatorname{Hom}_{\mathsf{C}}(Y,FX).\end{eqnarray}$$

The situation when $R$ is a field was the one considered in [Reference Bondal and KapranovBK89]. For a general ring  $R$ , the appearance of $\operatorname{Hom}_{R}^{\ast }(-,I(\mathfrak{m}))$ , which is the Matlis duality functor, is natural for it is an extension of vector-space duality; see also Lemma A.2. The definition proposed above is not the only possible extension to the general context, but it is well suited for our purposes.

For an arbitrary graded commutative ring $R$ , we say that an $R$ -linear triangulated category $\mathsf{C}$ satisfies local Serre duality if there exists an $R$ -linear triangle equivalence $F:\mathsf{C}\xrightarrow[{}]{{\sim}}\mathsf{C}$ such that for every $\mathfrak{p}\in \operatorname{Spec}R$ and some integer $d(\mathfrak{p})$ the induced functor $\unicode[STIX]{x1D6F4}^{-d(\mathfrak{p})}F_{\mathfrak{p}}:\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}$ is a Serre functor for the $R_{\mathfrak{p}}$ -linear category $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}$ . Thus for all objects $X,Y$ in $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}$ there is a natural isomorphism

$$\begin{eqnarray}\operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{C}_{\mathfrak{ p}}}^{\ast }(X,Y),I(\mathfrak{p}))\xrightarrow[{}]{{\sim}}\operatorname{Hom}_{\mathsf{C}_{\mathfrak{p}}}(Y,\unicode[STIX]{x1D6F4}^{-d(\mathfrak{p})}F_{\mathfrak{ p}}X).\end{eqnarray}$$

For a compactly generated triangulated category, the Gorenstein property 6.1 is linked to local Serre duality for the subcategory of compact objects.

Proposition 7.3. Let $R$ be a graded commutative noetherian ring and $\mathsf{T}$ a compactly generated $R$ -linear triangulated category. Suppose that $\mathsf{T}$ is Gorenstein, with global Serre functor $F$ and shifts $\{d(\mathfrak{p})\}$ . Then for each $\mathfrak{p}\in \operatorname{supp}_{R}(\mathsf{T})$ , object $X\in (\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T})^{\mathsf{c}}$ and $Y\in \mathsf{T}_{\mathfrak{p}}$ there is a natural isomorphism

$$\begin{eqnarray}\operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{T}}^{\ast }(X,Y),I(\mathfrak{p}))\cong \operatorname{Hom}_{\mathsf{T}}(Y,\unicode[STIX]{x1D6F4}^{-d(\mathfrak{p})}F_{\mathfrak{ p}}(X)).\end{eqnarray}$$

Proof. Given Remark 7.2 we can assume $X=C_{\mathfrak{p}}$ for a $\mathfrak{p}$ -torsion compact object $C$ in  $\mathsf{T}$ . The desired isomorphism is a concatenation of the following natural ones:

$$\begin{eqnarray}\displaystyle \operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{T}}^{\ast }(C_{\mathfrak{p}},Y),I(\mathfrak{p})) & \cong & \displaystyle \operatorname{Hom}_{R}(\operatorname{Hom}_{\mathsf{T}}^{\ast }(C,Y),I(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{\mathsf{T}}(Y,T_{\mathfrak{p}}(C))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{\mathsf{T}}(Y,\unicode[STIX]{x1D6F4}^{-d(\mathfrak{p})}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}F(C))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{\mathsf{T}}(Y,\unicode[STIX]{x1D6F4}^{-d(\mathfrak{p})}\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p})}F(C)_{\mathfrak{p}})\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{\mathsf{T}}(Y,\unicode[STIX]{x1D6F4}^{-d(\mathfrak{p})}\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{p})}F_{\mathfrak{p}}(C_{\mathfrak{p}}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{\mathsf{T}}(Y,\unicode[STIX]{x1D6F4}^{-d(\mathfrak{p})}F_{\mathfrak{p}}(C_{\mathfrak{p}})).\nonumber\end{eqnarray}$$

In this chain, the first map is induced by the localisation $C\mapsto C_{\mathfrak{p}}$ and is an isomorphism because $Y$ is $\mathfrak{p}$ -local. The second one is by the definition of $T_{\mathfrak{p}}(C)$ ; the third is by the Gorenstein property of  $\mathsf{T}$ ; the fourth is by the definition of $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}$ ; the last two are explained by Remark 7.2, where for the last one uses also the fact that $C_{\mathfrak{p}}$ , and hence also $F_{\mathfrak{p}}(C_{\mathfrak{p}})$ , is $\mathfrak{p}$ -torsion.◻

Corollary 7.4. Let $R$ be a graded commutative noetherian ring and $\mathsf{T}$ a compactly generated $R$ -linear triangulated category. If $\mathsf{T}$ is Gorenstein, then $\mathsf{T}^{\mathsf{c}}$ satisfies local Serre duality.

Proof. Given Remark 7.2, the assertion follows from Proposition 7.3. ◻

Example 7.5. In the notation of Example 6.3, when $A$ is a (commutative noetherian) Gorenstein ring, local Serre duality reads: For each $\mathfrak{p}\in \operatorname{Spec}A$ and integer $n$ there are natural isomorphisms

$$\begin{eqnarray}\operatorname{Hom}_{A_{\mathfrak{p}}}(\operatorname{Ext}_{A_{\mathfrak{p}}}^{n}(X,Y),I(\mathfrak{p}))\cong \operatorname{Ext}_{A_{\mathfrak{p}}}^{n+\dim A_{\mathfrak{p}}}(Y,X),\end{eqnarray}$$

where $X$ is a perfect complexes of $A_{\mathfrak{p}}$ -modules with finite length cohomology, and $Y$ is a complex of $A_{\mathfrak{p}}$ -modules.

Auslander–Reiten triangles

Let $\mathsf{C}$ be an essentially small triangulated category. Following Happel [Reference HappelHap88], an exact triangle $X\xrightarrow[{}]{\unicode[STIX]{x1D6FC}}Y\xrightarrow[{}]{\unicode[STIX]{x1D6FD}}Z\xrightarrow[{}]{\unicode[STIX]{x1D6FE}}\unicode[STIX]{x1D6F4}X$ in $\mathsf{C}$ is an Auslander–Reiten triangle if:

  1. (i) any morphism $X\rightarrow X^{\prime }$ that is not a split monomorphism factors through $\unicode[STIX]{x1D6FC}$ ;

  2. (ii) any morphism $Z^{\prime }\rightarrow Z$ that is not a split epimorphism factors through $\unicode[STIX]{x1D6FD}$ ;

  3. (iii) $\unicode[STIX]{x1D6FE}\neq 0$ .

In this case, the endomorphism rings of $X$ and $Z$ are local, and in particular the objects are indecomposable. Moreover, each of $X$ and $Z$ determines the AR-triangle up to isomorphism. Assuming conditions (ii) and (iii), the condition (i) is equivalent to the following.

  1. (i′) The endomorphism ring of $X$ is local.

See [Reference KrauseKra00, § 2] for details.

Let $\mathsf{C}$ be a Krull–Schmidt category, that is, each object decomposes into a finite direct sum of objects with local endomorphism rings. We say that $\mathsf{C}$ has AR-triangles if for every indecomposable object $X$ there are AR-triangles

$$\begin{eqnarray}V\rightarrow W\rightarrow X\rightarrow \unicode[STIX]{x1D6F4}V\quad \text{and}\quad X\rightarrow Y\rightarrow Z\rightarrow \unicode[STIX]{x1D6F4}X.\end{eqnarray}$$

The next proposition establishes the existence of AR-triangles; it is the analogue of a result of Reiten and Van den Bergh [Reference Reiten and Van den BerghRV02, I.2] for triangulated categories that are Hom-finite over a field.

Proposition 7.6. Let $R$ be a graded commutative ring that is local, and let $\mathsf{C}$ be an essentially small $R$ -linear triangulated category that is Krull–Schmidt. If $\mathsf{C}$ has a Serre functor, then it has AR-triangles.

Proof. Let $F$ be a Serre functor for $\mathsf{C}$ and $X$ an indecomposable object in  $\mathsf{C}$ . The ring $\operatorname{End}_{\mathsf{C}}(X)$ is thus local; let $J$ be its maximal ideal and $I$ the right ideal of $\operatorname{End}_{\mathsf{C}}^{\ast }(X)$ that it generates. One has $I^{0}=J$ , by Remark 7.7, so $\operatorname{End}_{\mathsf{C}}^{\ast }(X)/I$ is nonzero. Choose a nonzero morphism $\operatorname{End}_{\mathsf{C}}^{\ast }(X)/I\rightarrow I(\mathfrak{m})$ and let $\unicode[STIX]{x1D6FE}:X\rightarrow FX$ be the corresponding morphism in $\mathsf{C}$ provided by Serre duality (7.1). We claim that the induced exact triangle

$$\begin{eqnarray}\unicode[STIX]{x1D6F4}^{-1}FX\rightarrow W\rightarrow X\xrightarrow[{}]{\unicode[STIX]{x1D6FE}}FX\end{eqnarray}$$

is an AR-triangle. Indeed, by construction, if $X^{\prime }\rightarrow X$ is not a split epimorphism, then the composition $\operatorname{Hom}_{\mathsf{C}}^{\ast }(X,X^{\prime })\rightarrow \operatorname{End}_{\mathsf{C}}^{\ast }(X)/I\rightarrow I(\mathfrak{m})$ is zero, and therefore the naturality of (7.1) yields that the composition $X^{\prime }\rightarrow X\xrightarrow[{}]{\unicode[STIX]{x1D6FE}}FX$ is zero. Moreover as $X$ is indecomposable so is  $FX$ .

Applying this construction to $F^{-1}\unicode[STIX]{x1D6F4}X$ yields an AR-triangle starting at  $X$ .◻

The following observations about graded rings has been, and will again be, used.

Remark 7.7. Let $E$ be a graded ring. For any $J$ a right ideal in $E^{0}$ , the right ideal $JE$ of $E$ it generates satisfies $JE\cap E^{0}=J$ ; this can be verified directly, or by noting that $E^{0}$ is a direct summand of  $E$ , as right $E^{0}$ -modules. It follows that if $E$ is artinian than so is  $E^{0}$ : any descending chain of ideals in $E^{0}$ stabilises, because the chain of ideals in $E$ that they generate stabilises; cf. the proof of [Reference Bruns and HerzogBH98, Theorem 1.5.5].

Let $R$ be a graded commutative ring. An $R$ -linear triangulated category $\mathsf{C}$ is noetherian if the $R$ -module $\operatorname{Hom}_{\mathsf{C}}^{\ast }(X,Y)$ is noetherian for all $X,Y$ in  $\mathsf{C}$ .

Lemma 7.8. Let $R$ be a graded commutative ring and $\mathsf{C}$ an essentially small, noetherian, $R$ -linear triangulated category. For each $\mathfrak{p}\in \operatorname{Spec}R$ , the idempotent completion of $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}$ is a Krull–Schmidt category.

Proof. The noetherian property implies that for any object $X$ in $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}$ the $R_{\mathfrak{p}}$ -module $E:=\operatorname{End}_{\unicode[STIX]{x1D6FE}_{\mathfrak{ p}}\mathsf{C}}^{\ast }(X)$ is of finite length. The graded ring $E$ is thus artinian, and then so is the ring  $E^{0}$ , by Remark 7.7. Artinian rings are semi-perfect, so the idempotent completion of $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{C}$ is a Krull–Schmidt category; see [Reference KrauseKra15, Corollary 4.4].◻

Corollary 7.9. Let $\mathsf{T}$ be a compactly generated $R$ -linear triangulated category with $\mathsf{T}^{\mathsf{c}}$ noetherian. If $\mathsf{T}$ is Gorenstein, then $(\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T})^{\mathsf{c}}$ has AR-triangles for $\mathfrak{p}\in \operatorname{Spec}R$ .

Proof. Set $\mathsf{S}:=(\unicode[STIX]{x1D6E4}_{\mathfrak{p}}\mathsf{T})^{\mathsf{c}}$ ; this is an idempotent complete, essentially small, $R_{\mathfrak{p}}$ -linear triangulated category. Since $\mathsf{T}^{\mathsf{c}}$ is noetherian, it follows from Remark 7.2 and Lemma 7.8, that $\mathsf{S}$ is a Krull–Schmidt category. The Gorenstein hypothesis implies that the $R_{\mathfrak{p}}$ -linear category $\mathsf{S}$ has a Serre functor, by Proposition 7.3, and then Proposition 7.6 yields the existence of AR-triangles.◻

Next we consider local Serre duality for $\mathsf{D}^{b}(\mathsf{mod}\,G)$ and $\mathsf{stmod}\,G$ . Recall from Lemma 2.6 that localisation at $\mathfrak{p}\in \operatorname{Proj}H^{\ast }(G,k)$ induces a triangle equivalence $\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{K}(\mathsf{Inj}\,G))\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6E4}_{\mathfrak{p}}(\mathsf{StMod}\,G)$ . Using Remark 7.2, this yields (up to direct summands) triangle equivalences

$$\begin{eqnarray}\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{D}^{b}(\mathsf{mod}\,G))\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6E4}_{\mathfrak{ p}}(\mathsf{K}(\mathsf{Inj}\,G))^{\mathsf{c}}\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6E4}_{\mathfrak{ p}}(\mathsf{StMod}\,G)^{\mathsf{c}}\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6FE}_{\mathfrak{ p}}(\mathsf{stmod}\,G).\end{eqnarray}$$

The result below contains Theorem 1.1 in the introduction.

Theorem 7.10. Let $G$ be a finite group scheme over a field $k$ . Then the $H^{\ast }(G,k)$ -linear triangulated category $\mathsf{D}:=\mathsf{D}^{b}(\mathsf{mod}\,G)$ satisfies local Serre duality. Said otherwise, given $\mathfrak{p}$ in $\operatorname{Spec}R$ and with $d$ the Krull dimension of $H^{\ast }(G,k)/\mathfrak{p}$ , for each $M$ in $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{D}$ and $N$ in $\mathsf{D}_{\mathfrak{p}}$ , there are natural isomorphisms

$$\begin{eqnarray}\operatorname{Hom}_{H^{\ast }(G,k)}(\operatorname{Hom}_{\mathsf{D}_{\mathfrak{ p}}}^{\ast }(M,N),I(\mathfrak{p}))\cong \operatorname{Hom}_{\mathsf{D}_{\mathfrak{p}}}(N,\unicode[STIX]{x1D6FA}^{d}\unicode[STIX]{x1D6FF}_{G}\otimes _{k}M).\end{eqnarray}$$

In particular, $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}\mathsf{D}$ has AR-triangles.

Proof. The first assertion follows from Theorem 6.5 and Proposition 7.3. The existence of AR-triangles then follows from Corollary 7.9, as $\mathsf{D}$ is noetherian.◻

AR-components and periodicity

The existence of AR-triangles for a triangulated category $\mathsf{C}$ gives rise to an AR-quiver; see for example [Reference Happel, Preiser and RingelHPR82, Reference LiuLiu10]. The vertices are given by the isomorphism classes of indecomposable objects in $\mathsf{C}$ and an arrow $[X]\rightarrow [Y]$ exists if there is an irreducible morphism $X\rightarrow Y$ .

In the context of $\mathsf{stmod}\,G$ , one can describe part of the structure of the AR-quiver of the $\mathfrak{p}$ -local $\mathfrak{p}$ -torsion objects as the Serre functor is periodic.

Proposition 7.11. Let $G$ be a finite group scheme over a field $k$ . Fix a point $\mathfrak{p}$ in $\operatorname{Proj}H^{\ast }(G,k)$ and set $d=\dim H^{\ast }(G,k)/\mathfrak{p}$ . Then the Serre functor

$$\begin{eqnarray}\unicode[STIX]{x1D6FA}^{d}\unicode[STIX]{x1D708}:\unicode[STIX]{x1D6FE}_{\mathfrak{ p}}(\mathsf{stmod}\,G)\xrightarrow[{}]{{\sim}}\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)\end{eqnarray}$$

is periodic, that is, $(\unicode[STIX]{x1D6FA}^{d}\unicode[STIX]{x1D708})^{r}=\operatorname{id}$ for some positive integer  $r$ .

Proof. Lemma 2.1 and (4.1) provide an integer $r\geqslant 0$ such that $\unicode[STIX]{x1D708}^{r}M\cong M$ and $\unicode[STIX]{x1D6FA}^{r}M\cong M$ for $M$ in $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ . Thus $(\unicode[STIX]{x1D6FA}^{d}\unicode[STIX]{x1D708})^{r}=\operatorname{id}$ , since $\unicode[STIX]{x1D708}$ and $\unicode[STIX]{x1D6FA}$ commute.◻

This has the following consequence.

Corollary 7.12. Every connected component of the AR-quiver of $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ is a stable tube in case it is infinite; and otherwise, it is of the form $\mathbb{Z}\unicode[STIX]{x1D6E5}/U$ , where $\unicode[STIX]{x1D6E5}$ is a quiver of Dynkin type and $U$ is a group of automorphisms of $\mathbb{Z}\unicode[STIX]{x1D6E5}$ .

Proof. Since the Serre functor on $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ is periodic, the desired result follows from [Reference LiuLiu10, Theorem 5.5]; see also [Reference Happel, Preiser and RingelHPR82].◻

The preceding result may be seen as a first step in the direction of extending the results of Farnsteiner’s [Reference FarnsteinerFar07, § 3] concerning $\mathsf{stmod}\,G$ to $\unicode[STIX]{x1D6FE}_{\mathfrak{p}}(\mathsf{stmod}\,G)$ for a general (meaning, not necessarily closed) point $\mathfrak{p}$ of $\operatorname{Proj}H^{\ast }(G,k)$ .

Acknowledgements

It is a pleasure to thank Tobias Barthel, and a referee, for detailed comments on an earlier version of this manuscript.

Appendix A Injective modules at closed points

In this section we collect some remarks concerning injective hulls over graded rings, for use in § 5. Throughout $k$ will be a field and $A:=\bigoplus _{i\geqslant 0}A^{i}$ a finitely generated graded commutative $k$ -algebra with $A^{0}=k$ ; we have in mind $H^{\ast }(G,k)$ , for a finite group scheme $G$ over  $k$ .

As usual $\operatorname{Proj}A$ denotes the homogeneous prime ideals in $A$ that do not contain the ideal $A^{{\geqslant}1}$ . Given a point $\mathfrak{p}$ in $\operatorname{Proj}A$ , we write $k(\mathfrak{p})$ for the graded residue field at  $\mathfrak{p}$ ; this is the homogeneous field of fractions of the graded domain $A/\mathfrak{p}$ . Observe that $k(\mathfrak{p})^{0}$ is a field extension of $k$ and $k(\mathfrak{p})$ is of the form $k(\mathfrak{p})^{0}[t^{\pm 1}]$ for some indeterminate $t$ over $k(\mathfrak{p})^{0}$ ; see, for example, [Reference Bruns and HerzogBH98, Lemma 1.5.7].

Lemma A.1. The degree of $k(\mathfrak{m})^{0}/k$ is finite for any closed point $\mathfrak{m}$ in  $\operatorname{Proj}A$ .

Proof. One way to verify the lemma is as follows. The Krull dimension of $A/\mathfrak{m}$ is one so, by noetherian normalisation, there exists a subalgebra $k[t]$ of $A/\mathfrak{m}$ where $t$ is an indeterminate over $k$ and the $A/\mathfrak{m}$ is finitely generated $k[t]$ -module. Thus, inverting  $t$ , one gets that $(A/\mathfrak{m})_{t}$ is a finitely generated module over the graded field $k[t^{\pm 1}]$ , and hence isomorphic to $k(\mathfrak{m})$ . The finiteness of the extension $k(\mathfrak{m})/k[t^{\pm 1}]$ implies that the extension $k(\mathfrak{m})^{0}/k$ of fields is finite.◻

The result below is familiar; cf. [Reference Bruns and HerzogBH98, Proposition 3.6.16].

Lemma A.2. Let $A$ be as above, let $\mathfrak{m}$ be a closed point in $\operatorname{Proj}A$ and set $R:=A_{\mathfrak{m}}$ . The $R$ -submodule $I:=\bigcup _{i\geqslant 0}\operatorname{Hom}_{k}^{\ast }(R/\mathfrak{m}^{i},k)$ of $\operatorname{Hom}_{k}^{\ast }(R,k)$ is the injective hull of $k(\mathfrak{m})$ , and for any $\mathfrak{m}$ -torsion $R$ -module  $N$ , there is a natural isomorphism

$$\begin{eqnarray}\operatorname{Hom}_{R}(N,I)\cong \operatorname{Hom}_{k}(N,k).\end{eqnarray}$$

Proof. Set $K=k(\mathfrak{m})^{0}$ and recall that $k(\mathfrak{m})=K[t^{\pm 1}]$ , for some indeterminate $t$ over  $K$ . Thus, one has isomorphisms of graded $k(\mathfrak{m})$ -modules

(A.1) $$\begin{eqnarray}\displaystyle \operatorname{Hom}_{k}^{\ast }(k(\mathfrak{m}),k) & \cong & \displaystyle \operatorname{Hom}_{K}^{\ast }(k(\mathfrak{m}),\operatorname{Hom}_{k}(K,k))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{K}^{\ast }(k(\mathfrak{m}),K)\nonumber\\ \displaystyle & \cong & \displaystyle k(\mathfrak{m}).\end{eqnarray}$$

The first isomorphism is adjunction, the second holds because $\operatorname{rank}_{k}K$ is finite, by Lemma A.1, and the last one is a direct verification.

The $R$ -module $\operatorname{Hom}_{k}^{\ast }(R,k)$ is injective and hence so is its $\mathfrak{m}$ -torsion submodule

$$\begin{eqnarray}\mathop{\bigcup }_{i\geqslant 0}\operatorname{Hom}_{R}^{\ast }(R/\mathfrak{m}^{i},\operatorname{Hom}_{k}^{\ast }(R,k)).\end{eqnarray}$$

This is precisely the $R$ -module $I$ , by standard adjunction. Thus $I$ must be a direct sum of shifts of injective hulls of $k(\mathfrak{m})$ . It remains to verify that $I$ is in fact just the injective hull of $k(\mathfrak{m})$ . To this end, note that for any $\mathfrak{m}$ -torsion $R$ -module  $N$ , one has isomorphisms of graded $k(\mathfrak{m})$ -modules

$$\begin{eqnarray}\displaystyle \operatorname{Hom}_{R}^{\ast }(N,I) & \cong & \displaystyle \operatorname{Hom}_{R}^{\ast }(N,\operatorname{Hom}_{k}^{\ast }(R,k))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{k}^{\ast }(N,k).\nonumber\end{eqnarray}$$

This settles the last assertion in the desired result and also yields the first isomorphism below of graded $k(\mathfrak{m})$ -modules:

$$\begin{eqnarray}\operatorname{Hom}_{R}^{\ast }(k(\mathfrak{m}),I)\cong \operatorname{Hom}_{k}^{\ast }(k(\mathfrak{m}),k)\cong k(\mathfrak{m}).\end{eqnarray}$$

The second one holds by (A.1). It follows that $I$ is the injective hull of $k(\mathfrak{m})$ .◻

The next result, whose proof is rather similar to the one above, gives yet another way to get to the injective hull at a closed point of  $\operatorname{Proj}$ .

Recall that $I(\mathfrak{p})$ denotes the injective hull of $A/\mathfrak{p}$ for any $\mathfrak{p}$ in $\operatorname{Spec}A$ .

Lemma A.3. Let $A\rightarrow B$ be a homomorphism of graded commutative algebras, let $\mathfrak{m}$ be a closed point in $\operatorname{Proj}B$ , and set $\mathfrak{p}:=\mathfrak{m}\cap A$ . If the extension of residue fields $k(\mathfrak{p})\subseteq k(\mathfrak{m})$ is finite, then the $B$ -module $\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{m})}\operatorname{Hom}_{A}^{\ast }(B,I(\mathfrak{p}))$ is the injective hull of $B/\mathfrak{m}$ , and, for any $\mathfrak{m}$ -torsion $B$ -module  $N$ , adjunction induces an isomorphism

$$\begin{eqnarray}\operatorname{Hom}_{B}(N,I(\mathfrak{m}))\cong \operatorname{Hom}_{A}(N,I(\mathfrak{p})).\end{eqnarray}$$

Proof. The $B$ -module $I:=\unicode[STIX]{x1D6E4}_{{\mathcal{V}}(\mathfrak{m})}\operatorname{Hom}_{A}^{\ast }(B,I(\mathfrak{p}))$ is injective, for it is the $\mathfrak{m}$ -torsion submodule of the injective $B$ -module $\operatorname{Hom}_{A}^{\ast }(B,I(\mathfrak{p}))$ . As $\mathfrak{m}$ is a closed point, $I$ is a direct sum of shifts of copies of $I(\mathfrak{m})$ . It remains to make the computation below:

$$\begin{eqnarray}\displaystyle \operatorname{Hom}_{B}^{\ast }(k(\mathfrak{m}),I) & \cong & \displaystyle \operatorname{Hom}_{B}^{\ast }(k(\mathfrak{m}),\operatorname{Hom}_{A}^{\ast }(B,I(\mathfrak{p})))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{A}^{\ast }(k(\mathfrak{m}),I(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{k(\mathfrak{p})}^{\ast }(k(\mathfrak{m}),\operatorname{Hom}_{A}^{\ast }(k(\mathfrak{p}),I(\mathfrak{p})))\nonumber\\ \displaystyle & \cong & \displaystyle \operatorname{Hom}_{k(\mathfrak{p})}^{\ast }(k(\mathfrak{m}),k(\mathfrak{p}))\nonumber\\ \displaystyle & \cong & \displaystyle k(\mathfrak{m}).\nonumber\end{eqnarray}$$

These are all isomorphisms of $k(\mathfrak{m})$ -modules. The last one is where the hypothesis that $k(\mathfrak{m})/k(\mathfrak{p})$ is finite is required. This implies that $I\cong I(\mathfrak{m})$ . Given this, the last isomorphism follows by standard adjunction.◻

Footnotes

SBI was partly supported by NSF grant DMS-1503044 and DMS-1700985 and JP was partly supported by NSF grants DMS-0953011 and DMS-1501146.

References

Auslander, M., Functors and morphisms determined by objects , in Representation theory of algebras, Philadelphia, PA, 1976, Lecture Notes in Pure and Applied Mathematics, vol. 37 (Dekker, New York, NY, 1978), 1244.Google Scholar
Balmer, P., Spectra, spectra, spectra—tensor triangular spectra versus Zariski spectra of endomorphism rings , Algebr. Geom. Topol. 10 (2010), 15211563.Google Scholar
Balmer, P., Dell’Ambrogio, I. and Sanders, B., Grothendieck–Neeman duality and the Wirthmüller isomorphism , Compos. Math. 152 (2016), 17401776.Google Scholar
Barthel, T., Heard, D. and Valenzuela, G., Local duality for structured ring spectra , J. Pure Appl. Algebra 222 (2018), 433463.Google Scholar
Benson, D., Modular representation theory: new trends and methods, Lecture Notes in Mathematics, vol. 1081 (Springer, Berlin, 1984).Google Scholar
Benson, D., Modules with injective cohomology, and local duality for a finite group , New York J. Math. 7 (2001), 201215.Google Scholar
Benson, D., Idempotent kG-modules with injective cohomology , J. Pure Appl. Algebra 212 (2008), 17441746.Google Scholar
Benson, D. J., Carlson, J. F. and Rickard, J., Complexity and varieties for infinitely generated modules. II , Math. Proc. Cambridge Philos. Soc. 120 (1996), 597615.Google Scholar
Benson, D. and Greenlees, J. P. C., Localization and duality in topology and modular representation theory , J. Pure Appl. Algebra 212 (2008), 17161743.Google Scholar
Benson, D. and Greenlees, J. P. C., Stratifying the derived category of cochains on BG for G a compact Lie group , J. Pure Appl. Algebra 218 (2014), 642650.Google Scholar
Benson, D. J., Iyengar, S. B. and Krause, H., Local cohomology and support for triangulated categories , Ann. Sci. Éc. Norm. Supér. (4) 41 (2008), 575621.Google Scholar
Benson, D. J., Iyengar, S. B. and Krause, H., Stratifying triangulated categories , J. Topology 4 (2011), 641666.Google Scholar
Benson, D. J., Iyengar, S. B. and Krause, H., Stratifying modular representations of finite groups , Ann. of Math. (2) 174 (2011), 16431684.Google Scholar
Benson, D. J., Iyengar, S. B. and Krause, H., Representations of finite groups: local cohomology and support, Oberwolfach Seminars, vol. 43 (Birkhäuser/Springer, Basel, 2012).Google Scholar
Benson, D. J., Iyengar, S. B. and Krause, H., A local-global principle for small triangulated categories , Math. Proc. Cambridge Philos. Soc. 158 (2015), 451476.Google Scholar
Benson, D. J., Iyengar, S. B., Krause, H. and Pevtsova, J., Colocalising subcategories of modules over finite group schemes , Ann. K-Theory 2 (2017), 387408.Google Scholar
Benson, D. J., Iyengar, S. B., Krause, H. and Pevtsova, J., Stratification for module categories of finite group schemes , J. Amer. Math. Soc. 31 (2018), 265302.Google Scholar
Benson, D. J. and Krause, H., Pure injectives and the spectrum of the cohomology ring of a finite group , J. Reine Angew. Math. 542 (2002), 2351.Google Scholar
Bernstein, I. N., Gelfand, I. M. and Gelfand, S. I., Algebraic vector bundles on P n and problems of linear algebra , Funct. Anal. Appl. 12 (1978), 212214.Google Scholar
Bondal, A. I. and Kapranov, M. M., Representable functors, Serre functors, and mutations , Izv. Akad. Nauk SSSR Ser. Mat. 53 (1989), 11831205; 1337; English translation in Math. USSR Izv. 35 (1990), 519–541.Google Scholar
Bruns, W. and Herzog, J., Cohen–Macaulay rings, Cambridge Studies in Advanced Mathematics, vol. 39, revised edition (Cambridge University Press, Cambridge, 1998).Google Scholar
Carlson, J. F., Modules and group algebras, Lecture Notes in Mathematics ETH Zürich (Birkhäuser, Basel, 1996).Google Scholar
Cartan, H. and Eilenberg, S., Homological algebra (Princeton University Press, Princeton, NJ, 1956).Google Scholar
Crawley-Boevey, W., Tame algebras and generic modules , Proc. Lond. Math. Soc. (3) 63 (1991), 241265.Google Scholar
Crawley-Boevey, W., Modules of finite length over their endomorphism rings , in Representations of algebras and related topics, Kyoto, 1990, London Mathematical Society Lecture Note Series, vol. 168 (Cambridge University Press, Cambridge, 1992), 127184.Google Scholar
Dwyer, W., Greenlees, J. P. C. and Iyengar, S., Duality in algebra and topology , Adv. Math. 200 (2006), 357402.Google Scholar
Farnsteiner, R., Support spaces and Auslander–Reiten components , in Lie algebras, vertex operator algebras and their applications, Contemporary Mathematics, vol. 442 (American Mathematical Society, Providence, RI, 2007), 6187.Google Scholar
Frankild, A., Iyengar, S. and Jørgensen, P., Dualizing DG modules and Gorenstein DG algebras , J. Lond. Math. Soc. (2) 68 (2003), 288306.Google Scholar
Friedlander, E. M. and Suslin, A., Cohomology of finite group schemes over a field , Invent. Math. 127 (1997), 209270.Google Scholar
Happel, D., Triangulated categories in the representation theory of finite dimensional algebras, London Mathematical Society Lecture Note Series, vol. 119 (Cambridge University Press, Cambridge, 1988).Google Scholar
Happel, D., On Gorenstein algebras , in Representation theory of finite groups and finite-dimensional algebras, Bielefeld, 1991, Progress in Mathematics, vol. 95 (Birkhäuser, Basel, 1991), 389404.Google Scholar
Happel, D., Preiser, U. and Ringel, C. M., Vinberg’s characterization of Dynkin diagrams using subadditive functions with application to D Tr-periodic modules , in Representation theory II, Proc. Second Internat. Conf., Ottawa, ON, 1979, Lecture Notes in Mathematics, vol. 832 (Springer, Berlin, 1982), 280294.Google Scholar
Hartshorne, R., Algebraic geometry, Graduate Texts in Mathematics, vol. 52 (Springer, New York, NY, 1977).Google Scholar
Jantzen, J. C., Representations of algebraic groups, second edition (American Mathematical Society, Providence, RI, 2003).Google Scholar
Krause, H., Auslander–Reiten theory via Brown representability , K-Theory 20 (2000), 331344.Google Scholar
Krause, H., A short proof of Auslander’s defect formula , Linear Algebra Appl. 365 (2003), 267270.Google Scholar
Krause, H., The stable derived category of a noetherian scheme , Compos. Math. 141 (2005), 11281162.Google Scholar
Krause, H., Krull–Schmidt categories and projective covers , Expo. Math. 33 (2015), 535549.Google Scholar
Krause, H. and Le, J., The Auslander–Reiten formula for complexes of modules , Adv. Math. 207 (2006), 133148.Google Scholar
Krause, H. and Reichenbach, U., Endofiniteness in stable homotopy theory , Trans. Amer. Math. Soc. 353 (2000), 157173.Google Scholar
Liu, S., Auslander–Reiten theory in a Krull–Schmidt category , São Paulo J. Math. Sci. 4 (2010), 425472.Google Scholar
Ravenel, D. C., Nilpotence and periodicity in stable homotopy theory, Annals of Mathematics Studies, vol. 128 (Princeton University Press, Princeton, NJ, 1992).Google Scholar
Reiten, I. and Van den Bergh, M., Noetherian hereditary abelian categories satisfying Serre duality , J. Amer. Math. Soc. 15 (2002), 295366.Google Scholar
Skowroński, A. and Yamagata, K., Frobenius algebras. I, EMS Textbooks in Mathematics (European Mathematical Society (EMS), Zürich, 2011).Google Scholar
Waterhouse, W. C., Introduction to affine group schemes, Graduate Texts in Mathematics, vol. 66 (Springer, New York, NY, 1979).Google Scholar
Yuliawan, F., Actions of Hochschild cohomology and local duality in modular representation theory, PhD thesis, Universität Bielefeld (2017).Google Scholar