Hostname: page-component-cd9895bd7-dk4vv Total loading time: 0 Render date: 2024-12-22T16:09:29.518Z Has data issue: false hasContentIssue false

A McKay Correspondence in Positive Characteristic

Published online by Cambridge University Press:  10 December 2024

Christian Liedtke*
Affiliation:
TU München, Department of Mathematics, Boltzmannstr. 3, D-85748 Garching bei München, Germany

Abstract

We establish a McKay correspondence for finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_2$ in positive characteristic. As an application, we obtain a McKay correspondence for all rational double point singularities in characteristic $p\geq 7$. We discuss linearly reductive quotient singularities and canonical lifts over the ring of Witt vectors. In dimension 2, we establish simultaneous resolutions of singularities of these canonical lifts via G-Hilbert schemes. In the appendix, we discuss several approaches towards the notion of conjugacy classes for finite group schemes: This is an ingredient in McKay correspondences, but also of independent interest.

Type
Algebraic and Complex Geometry
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2024. Published by Cambridge University Press

1 Introduction

1.1 Klein’s classification and McKay’s correspondence

Felix Klein [Reference KleinKl84] classified finite subgroups G of ${\mathbf {SL}}_2({\mathbb {C}})$ : up to conjugation, there are two infinite series and three isolated cases.

  1. 1. The associated quotient singularity ${\mathbb {C}}^2/G$ is called a Klein singularity and the singularities arising this way are precisely the rational double point singularities. Its minimal resolution of singularities is a union of ${\mathbb {P}}^1$ ’s, whose dual intersection graph $\Gamma $ is a simply-laced Dynkin diagram of finite type, that is, of type $A_n$ , $D_n$ , $E_6$ , $E_7$ or $E_8$ .

  2. 2. John McKay [Reference McKayMc80] associated a finite graph $\widehat {\Gamma }$ to $G \subset {\mathbf {SL}}_2({\mathbb {C}})$ , whose vertices correspond to the isomorphism classes of the simple representations of G. This graph is a Dynkin diagram of affine type $\widehat {A}_n$ , $\widehat {D}_n$ , $\widehat {E}_6$ , $\widehat {E}_7$ or $\widehat {E}_8$ .

  3. 3. After these preparations, the classical McKay correspondence consists of the following observations:

    1. (a) The graph $\Gamma $ is obtained from $\widehat {\Gamma }$ by removing the vertex corresponding to the trivial representation.

    2. (b) There exists a bijection between conjugacy classes of G, vertices of $\widehat {\Gamma }$ and isomorphism classes of simple representations of G.

    3. (c) There exists a bijection between finite subgroups of ${\mathbf {SL}}_2({\mathbb {C}})$ up to conjugacy, the above Dynkin diagrams of affine type, Klein singularities and the above Dynkin diagrams of finite type.

By now, there are various approaches to and a vast literature on this subject, such as [Reference KnörrerKn85, Reference KostantKo85, Reference McKayMc80, Reference McKayMc81, Reference SteinbergSt85] and many more. Also, there are now generalisations into very different directions: higher dimensional algebraic geometry [Reference ReidRe02], K-theory [Reference Gonzalez-Sprinberg and VerdierGV83], derived categories of coherent sheaves [Reference Bridgeland, King and ReidBKR01, Reference Kapranov and VasserotKV00], representations of quivers [Reference KirillovKi16], non-commutative geometry and Hopf algebras [Reference Chan, Kirkman, Walton and ZhangCKWZ16] and string theory [Reference Aspinwall, Bridgeland, Craw, Douglas, Gross, Kapustin, Moore, Segal, Szendrői and WilsonAetal09] - just to mention a few.

1.2 Positive characteristic

Now, let k be an algebraically closed field of characteristic $p>0$ .

1.2.1 Wild McKay correspondence

The classical McKay correspondence as sketched above is still partially available over k if G is assumed to be a finite subgroup of ${\mathbf {SL}}_2(k)$ of order prime to p - this is the tame case. If p divides the order of G - this is the modular or wild case - then this correspondence breaks down. We refer to Yasuda’s surveys [Reference YasudaYa22, Reference YasudaYa23] about conjectures and partial results concerning such wild McKay correspondences.

1.2.2 Linearly reductive McKay correspondence

In this article, we show that if G is a finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ , then there is a reasonable version of the classical McKay correspondence. For example, we obtain a McKay correspondence for all rational double point singularities if $p\geq 7$ . Instead of considering groups, we allow non-reduced group schemes over k, but we require their categories of k-linear and finite-dimensional representations to be semi-simple. We refer to [Reference Tonini and YasudaTY20] about conjectures and partial results concerning a McKay correspondence for the group scheme $\boldsymbol {\alpha }_p$ , which is not linearly reductive.

Thus, let G be a finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ with $p\geq 7$ . (In this introduction, we will exclude small characteristics whenever this is makes our discussion easier.) Let $x\in X:=U/G$ with $U={\mathbb {A}}_k^2$ or $U=\widehat {{\mathbb {A}}}_k^2$ be the associated Klein singularity, which is a rational double point. One goal of this article is to define a notion of conjugacy class for G, to construct graphs $\Gamma $ and $\widehat {\Gamma }$ and to establish bijections as above.

Let us make three comments:

  1. 1. It is interesting in its own that a McKay correspondence can be extended from finite group schemes of length prime to p, that is, the tame case, to linearly reductive group schemes.

  2. 2. What makes this linearly reductive McKay correspondence really interesting is that the bijection in Theorem 1.1 is not true when considering finite groups of order prime to p only, see Example 1.2.

  3. 3. Probably, many more aspects of the classical McKay correspondence can be carried over to the linearly reductive setting, but rather than writing a whole monograph, we decided to establish only some basic bijections.

1.3 Linearly reductive group schemes

Let G be a finite and linearly reductive group scheme over an algebraically closed field k of characteristic $p\geq 0$ . By definition, this means that the category $\mathrm {Rep}_k(G)$ of k-linear and finite-dimensional representations of G is semi-simple.

If $p=0$ , then every finite group scheme over k is étale and linearly reductive. and in fact, it is the constant group scheme associated to a finite group. In fact, the functor $G\mapsto G_{\mathrm {abs}}:=G(k)$ induces an equivalence of categories between finite group schemes over k and finite groups.

If $p>0$ , then every linearly reductive group scheme admits a canonical semi-direct product decomposition

$$ \begin{align*}G\,\cong\,G^\circ\rtimes G^{\mathrm{\acute{e}t}}, \end{align*} $$

where $G^{\mathrm {\acute {e}t}}$ is a group scheme of length prime to p (and thus, the constant group scheme associated to a finite group of order prime to p) and where $G^\circ $ is infinitesimal and diagonalisable. The latter implies that $G^\circ $ is a product of group schemes of the form $\boldsymbol {\mu }_{p^n}$ . Conversely, every such semi-direct product of a diagonalisable group scheme with the constant group scheme associated to a finite group of order prime to p is linearly reductive. This structure result is usually attributed to Nagata [Reference NagataNa61], but see also [Reference Abramovich, Olsson and VistoliAOV08], [Reference ChinCh92] and [Reference HashimotoHa15]. Linearly reductive group schemes over k strictly contains the class of constant group schemes associated to finite groups of order prime to p.

1.3.1 Abstract groups

Associated to G, there is an abstract finite group $G_{\mathrm {abs}}$ and we refer for the slightly technical definition to Section 2.2. The order of $G_{\mathrm {abs}}$ is equal to the length of G. For example, if G is étale over k, then we have $G_{\mathrm {abs}}\cong G(k)$ . The assignment $G\mapsto G_{\mathrm {abs}}$ establishes an equivalence of categories

(1.1) $$ \begin{align} \left\{ \begin{array}{l} \mbox{finite linearly reductive}\\ \mbox{group schemes over }k \end{array}\right\} \quad\leftrightarrow\quad \left\{ \begin{array}{l} \mbox{finite groups with a}\\ \mbox{unique }p\mbox{-Sylow subgroup} \end{array}\right\}, \end{align} $$

see [Reference Liedtke, Martin and MatsumotoLMM21] and Lemma 2.2.

1.3.2 Canonical lifts

In [Reference Liedtke, Martin and MatsumotoLMM21], we showed that if G is a finite and linearly reductive group scheme, then there exists a lift of G over the ring of Witt vectors $W(k)$ . We note that $G^\circ $ and $G^{\mathrm {\acute {e}t}}$ even lift uniquely to $W(k)$ and we define the canonical lift $G_{\mathrm {can}}\to \mathrm {Spec}\: K$ of G to be the unique lift that is a semi-direct product of the lifts of $G^\circ $ and $G^{\mathrm {\acute {e}t}}$ . Any lift of G to some extension field of K becomes isomorphic to $G_{\mathrm {can}}$ after possibly passing to some further field extension. Moreover, the finite group $G_{\mathrm {can}}(\overline {K})$ is isomorphic to $G_{\mathrm {abs}}$ .

1.3.3 Representation theory

By [Reference Liedtke, Martin and MatsumotoLMM21] and Proposition 2.6, there exist canonical equivalences of representation categories

(1.2) $$ \begin{align} \mathrm{Rep}_{k}(G) \,\to\, \mathrm{Rep}_{\overline{K}}(G_{\mathrm{can},\overline{K}}) \,\to\, \mathrm{Rep}_{\mathbb{C}}(G_{\mathrm{abs}}) \end{align} $$

that are compatible with degrees, direct sums, tensor products, duals and simplicity. These equivalences induce isomorphisms of rings

$$ \begin{align*}K_k(G) \,\to\, K_{\overline{K}}(G_{\mathrm{can},\overline{K}}) \,\to\, K_{{\mathbb{C}}}(G_{\mathrm{abs}}), \end{align*} $$

see Corollary 2.8. Here, $K_F(G)$ denotes the K-group associated to F-linear and finite-dimensional G-representations.

1.3.4 Hopf algebras

If G is a finite group scheme over k, then the multiplication map turns $H^0(G,{\mathcal O}_G)$ into a finite-dimensional Hopf algebra over k. We discuss finite group schemes and among them the linearly reductive ones from the point of Hopf algebras in Appendix A.

1.4 Linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$

Let k be an algebraically closed field of characteristic $p\geq 0$ . Hashimoto [Reference HashimotoHa15] extended Klein’s classification [Reference KleinKl84] of finite subgroups of ${\mathbf {SL}}_2({\mathbb {C}})$ up to conjugation to the setting of finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ . If $p\geq 7$ , then one obtains a list analogous to Klein’s classical list. If $p\in \{2,3,5\}$ , then some classical cases are missing, but there are no new cases.

1.5 McKay graph and McKay correspondence

Let G be a finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ . As in McKay’s original construction [Reference McKayMc80], we associate an affine Dynkin diagram $\widehat {\Gamma }$ to G, its embedding into ${\mathbf {SL}}_{2,k}$ and the set of isomorphism classes of simple representations of G. This is the McKay graph associated to this data. In fact, we will see that it is compatible with the equivalences induced by (1.1) and (1.2) and we refer to Section 3.1 for details. We establish the following version of McKay’s theorem [Reference McKayMc80] in positive characteristic.

Theorem 1.1 (Theorem 3.6)

Let k be an algebraically closed field of characteristic $p\geq 0$ . There exists a bijection between non-trivial, finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ up to conjugation and affine Dynkin graphs of type

$$ \begin{align*}\begin{array}{lcl} \widehat{A}_n, \widehat{D}_n, \widehat{E}_6, \widehat{E}_7, \widehat{E}_8 &\quad \mbox{if} \quad & p=0\quad \mbox{or} \quad p\geq7, \\ \widehat{A}_n, \widehat{D}_n, \widehat{E}_6, \widehat{E}_7 &\quad \mbox{if} \quad & p=5,\\ \widehat{A}_n, \widehat{D}_n &\quad \mbox{if} \quad & p=3, \\ \widehat{A}_n &\quad \mbox{if} \quad &p=2. \end{array} \end{align*} $$

By construction, this bijection is compatible with the classical McKay correspondence via the lifting results and the equivalences (1.1), (1.2).

Example 1.2. The linearly reductive group scheme corresponding to $\widehat {A}_n$ is $\boldsymbol {\mu }_{n+1}$ . This group scheme is reduced, that is, étale, if and only if p does not divide $n+1$ . In particular, it is crucial to allow non-reduced group schemes in order to obtain a bijection as in characteristic zero.

1.6 Linearly reductive quotient singularities

Consider ${\mathbf {GL}}_{2,k}$ with its usual linear action on $U={\mathbb {A}}^2_k$ or $U=\widehat {{\mathbb {A}}}_k^2$ . If G is a finite, linearly reductive and very small (see Definition 4.1) subgroup scheme of ${\mathbf {GL}}_{2,k}$ , then the associated quotient singularity $x\in X:=U/G$ is a two-dimensional linearly reductive quotient singularity in the sense of [Reference Liedtke, Martin and MatsumotoLMM21]. By loc.cit., such a singularity determines G together with its embedding $G\to {\mathbf {GL}}_{2,k}$ up to isomorphism and conjugation, respectively.

In Section 4.3, we will see that a minimal resolution of singularities of a two-dimensional linearly reductive quotient singularity $x\in X=U/G$ is provided by the G-Hilbert scheme

(1.3) $$ \begin{align} \pi \,:\, G\text{-}\mathrm{Hilb}(U) \,\to\, U/G, \end{align} $$

which generalises work of Ishii, Ito and Nakamura [Reference Ishii and NakamuraIN19, Reference Ito and NakamuraIN99].

Remark 1.3. In dimension two, a linearly reductive quotient singularity is the same as an F-regular singularity [Reference Liedtke, Martin and MatsumotoLMM21]. Thus, every two-dimensional F-regular singularity can be resolved by a suitable G-Hilbert scheme, see also Remark 4.6.

If moreover G is a subgroup scheme of ${\mathbf {SL}}_{2,k}$ , then $x\in X=U/G$ is called a Klein singularity. Klein singularities are rational double point singularities. If $p=0$ or $p\geq 7$ , then conversely every rational double point is a Klein singularity by Hashimoto [Reference HashimotoHa15] and, independently, by [Reference Liedtke and SatrianoLS14]. If $p\in \{2,3,5\}$ , then not every rational double point is a Klein singularity.

1.7 Canonical lifts and simultaneous resolutions

Let G be a very small, finite and linearly reductive subgroup scheme of ${\mathbf {GL}}_{2,k}$ and let $x\in X=U/G$ be the associated linearly reductive quotient singularity. Assume $p>0$ and let $W(k)$ be the ring of Witt vectors of k. In Section 4.4, we will establish the existence of a canonical lift

$$ \begin{align*}\mathcal{X}_{\mathrm{can}}\,\to\,\mathrm{Spec}\: W(k) \end{align*} $$

of $x\in X=U/G$ . Using G-Hilbert schemes in families, we will see in Section 4.5 that it admits a simultaneous and minimal resolution of singularities

$$ \begin{align*}\widetilde{\pi} \,:\, \mathcal{Y}\,\to\,\mathcal{X}_{\mathrm{can}}\,\to\,\mathrm{Spec}\: W(k). \end{align*} $$

We will prove this resolution to be unique, see Theorem 4.10.

1.8 McKay correspondence for Klein singularities

Let G be a finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ and let $x\in X=U/G$ be the associated Klein singularity.

As discussed in Section 1.5, we have the McKay graph $\widehat {\Gamma }$ associated to G, its embedding into ${\mathbf {SL}}_{2,k}$ and the set of isomorphism classes of simple representations of G.

Let $\pi :Y\to X$ be a minimal resolution of singularities. Since $x\in X$ is a rational double point, the exceptional divisor $\mathrm {Exc}(\pi )$ of $\pi $ is a configuration of ${\mathbb {P}}^1$ ’s, whose dual intersection graph $\Gamma $ is a simply-laced Dynkin diagram.

Theorem 1.4 (Theorem 5.3)

Let k be an algebraically closed field of characteristic $p\geq 0$ . Let G be a finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ and let $x\in X=U/G$ be the associated Klein singularity. Then, there exists a natural bijection of the graph $\Gamma $ with the graph obtained from $\widehat {\Gamma }$ by removing the vertex corresponding to the trivial representation.

Since every rational double point in characteristic $p\geq 7$ is a Klein singularity, we obtain the following.

Corollary 1.5. There exists a linearly reductive McKay correspondence for rational double point singularities in every characteristic $p\geq 7$ .

To establish this theorem, we use the Ishii-Ito-Nakamura resolution of singularities (1.3), as well as Hecke correspondences as in the work of Ito and Nakamura [Reference Ito and NakamuraIN99] and Nakajima [Reference NakajimaNa96, Reference NakajimaNa01].

1.9 Generalisations and variants

Let G be a very small, finite and linearly reductive subgroup scheme of ${\mathbf {GL}}_{2,k}$ and let $x\in X=U/G$ be the associated two-dimensional linearly reductive quotient singularity. Let $\pi :Y\to X$ be the minimal resolution of singularities and let $\mathrm {Exc}(\pi )$ be the exceptional divisor of $\pi $ .

  1. 1. In Theorem 5.4, we associate a representation of G to each point of $\mathrm {Exc}(\pi )$ . This generalises results of Ishii and Nakamura [Reference IshiiIs02, Reference Ishii and NakamuraIN19].

  2. 2. In Theorem 5.6, we establish a bijection between the components of $\textrm {Exc}(\pi )$ and reflexive ${\mathcal O}_X$ -modules. Probably, this result should be viewed as a theorem on two-dimensional F-regular singularities, see Remark 5.7. It generalises work of Artin and Verdier [Reference Artin and VerdierAV85], Wunram [Reference WunramWu88] and Ishii and Nakamura [Reference Ishii and NakamuraIN19].

  3. 3. In Theorem 7.1, we establish an equivalence of derived categories of coherent sheaves ${\mathcal {D}}(Y)$ and ${\mathcal {D}}^G(U)$ (G-equivariant sheaves on U). This generalises work of Kapranov and Vasserot [Reference Kapranov and VasserotKV00], Bridgeland, King and Reid [Reference Bridgeland, King and ReidBKR01] and Ishii, Nakamura and Ueda [Reference IshiiIs02, Reference Ishii and NakamuraIN19, Reference Ishii and UedaIU15]. Of course, these articles themselves generalise work of Gonzalez-Sprinberg and Verdier [Reference Gonzalez-Sprinberg and VerdierGV83] from K-theory to derived categories of coherent sheaves.

1.10 Ito-Reid correspondence

If G is a finite subgroup of ${\mathbf {SL}}_2({\mathbb {C}})$ , then Ito and Reid [Reference Ito and ReidIR96] found a natural bijection, the Ito-Reid correspondence, between the conjugacy classes of G and the vertices of the McKay graph $\widehat {\Gamma }$ .

Theorem 1.6. Let k be an algebraically closed field of characteristic $p\geq 0$ . There exists an Ito-Reid correspondence for finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ .

This can be proven using lifting results and the Ito-Reid correspondence over ${\mathbb {C}}$ , see Section 6.3 for details.

The main difficulty is to define a notion of conjugacy class for finite and linearly reductive group schemes that makes such a correspondence work: Let us recall the ring $K_k(G)$ from Section 1.3.3. If G is a finite and linearly reductive group scheme over k, then we define the set of conjugacy classes of G to be

$$ \begin{align*}\mathrm{Spec}\: \left({\mathbb{C}}\otimes K_k(G)\right). \end{align*} $$

At first sight, this might look rather artificial. One should think of it as defining conjugacy classes to be “dual” to simple representations. Moreover, our definition is compatible with lifting over $W(k)$ and it induces a bijection of the conjugacy classes of G with the conjugacy classes of $G_{\mathrm {abs}}$ . We refer to Section 6.1 and Appendix B.2 for details.

1.11 Conjugacy classes

In Theorem 1.6, we had to find a definition of conjugacy classes that makes this theorem true. This begs for the question whether there are other definitions or approaches, which is a question that is interesting in its own.

Let G be a finite group scheme (not necessarily linearly reductive) over an algebraically closed field k. In Appendix B, we study the following approaches to the notion of the set of conjugacy classes of G:

  1. 1. The set of conjugacy classes $G(k)/\sim $ of the group of k-rational points of G.

  2. 2. The spectrum of $F\otimes K_k(G)$ , where F is a field of characteristic zero that contains “sufficiently many” roots of unity.

  3. 3. The scheme that represents the functor of conjugacy classes from schemes over k to sets defined by $S\mapsto G(S)/\sim $ ,

  4. 4. the isotypical component of the trivial representation of the adjoint representation of G.

  5. 5. The simple subrepresentations of the extended adjoint representations $^{\mathrm {Ad}}A$ and $^{\mathrm {Ad}}(A^*)$ , which are defined using the quantum doubles of the Hopf algebra $A:=H^0(G,{\mathcal O}_G)$ and its dual $A^*$ .

All approaches lead essentially to the “same answer” in characteristic zero. However, they usually lead to very different notions in positive characteristic. On the other hand, all approaches have their merits and drawbacks. For linearly reductive group schemes, the approach (2) leads to a definition that is compatible with lifting and that leads to an Ito-Reid correspondence.

1.12 Organisation of this article

  1. - In Section 2, we recall basic facts about finite and linearly reductive group schemes over algebraically closed fields.

  2. - In Section 3, we construct the McKay graph $\widehat {\Gamma }$ . The main result is Theorem 3.6, a McKay correspondence.

  3. - In Section 4, we recall basic facts about linearly reductive quotient singularities. We establish an Ishii-Ito-Nakamura-type resolution of singularities, the canonical lift of such a singularity and a unique simultaneous resolution of singularities of the canonical lift.

  4. - In Section 5, we revisit the Ishii-Ito-Nakamura-type resolution $\pi $ of singularities of $x\in X=U/G$ and we introduce Hecke correspondences, which leads to Theorem 5.3, a bijection between simple and non-trivial representations of G and components of $\pi $ . We also study generalisations of this result to two-dimensional linearly reductive quotient singularities.

  5. - In Section 6, we establish an Ito-Reid correspondence between conjugacy classes of G and exceptional divisors in the minimal resolution of singularities of $x\in X=U/G$ .

  6. - In Section 7, we study derived categories of G-equivariant sheaves on U and on the minimal resolution $\pi :Y\to X=U/G$ .

  7. - In Appendix A, we recall results on finite group schemes from the point of view of Hopf algebras. We recall the adjoint representation, quantum doubles and the extended adjoint representation.

  8. - In Appendix B, we study several approaches toward the notion of a conjugacy class for finite group schemes.

2 Linearly reductive group schemes

In this section, we recall a couple of general facts about finite and linearly reductive group schemes over algebraically closed fields. We discuss the close relationship between such a group scheme G and a certain abstracted group $G_{\mathrm {abs}}$ associated to it. For the relationship to Hopf algebras, we refer to Appendix A.

2.1 Group schemes

Let G be a finite group scheme over an algebraically closed field k of characteristic $p\geq 0$ . Since k is perfect, there is a short exact sequence of finite group schemes over k

(2.1) $$ \begin{align} 1\,\to\,G^\circ\,\to\,G\,\to\, G^{\mathrm{\acute{e}t}} \,\to\,1, \end{align} $$

where $G^\circ $ is the connected component of the identity and where $G^{\mathrm {\acute {e}t}}$ is an étale group scheme over k. The reduction $G_{\mathrm {red}}\to G$ provides a canonical splitting of (2.1) and we obtain a canonical semi-direct product decomposition $G\cong G^\circ \rtimes G^{\mathrm {\acute {e}t}}$ . Since k is algebraically closed, $G^{\mathrm {\acute {e}t}}$ is the constant group scheme associated to the finite group $G(k)=G^{\mathrm {\acute {e}t}}(k)$ of k-rational points. Moreover, $G^\circ $ is an infinitesimal group scheme of length equal to some power of p. In particular, if $p=0$ or if the length of G is prime to p, then $G^\circ $ is trivial and then, G is étale.

If M is a finitely generated abelian group, then the group algebra $k[M]$ carries a Hopf algebra structure and the associated commutative group scheme is denoted $D(M):=\mathrm {Spec}\: k[M]$ . By definition, such group schemes are called diagonalisable. For example, we have $D({\mathbf {C}}_n)\cong \boldsymbol {\mu }_n$ , where ${\mathbf {C}}_n$ denotes the cyclic group of order n. We have that $\boldsymbol {\mu }_n$ is étale over k if and only if $p\nmid n$ .

A finite group scheme G over k is said to be linearly reductive if every k-linear and finite-dimensional representation of G is semi-simple. If $p=0$ , then all finite group schemes over k are étale and linearly reductive. If $p>0$ , then, by a theorem that is often attributed to Nagata [Reference NagataNa61, Theorem 2] (but see also [Reference Abramovich, Olsson and VistoliAOV08, Proposition 2.10], [Reference ChinCh92] and [Reference HashimotoHa15, Section 2]), a finite group scheme over k is linearly reductive if and only if it is an extension of a finite and étale group scheme, whose length is prime to p, by a diagonalisable group scheme.

2.2 Abstract groups and canonical lifts

Let G be a finite and linearly reductive group scheme over an algebraically closed field k of characteristic $p>0$ . Following [Reference Liedtke, Martin and MatsumotoLMM21, Section 2], we study the finite group

$$ \begin{align*}G_{\mathrm{abs}} \,:=\, \left(\underline{((G^\circ)^{D}(k))}_{{\mathbb{C}}}\right)^{D}({\mathbb{C}}) \,\rtimes\, G^{\mathrm{\acute{e}t}}(k). \end{align*} $$

Here, $-^D$ denotes $\underline {\mbox {Hom}}(-,{\mathbb {G}}_m)$ , the Cartier dual, of a commutative and finite group scheme. If G is étale, then $G_{\mathrm {abs}}=G(k)$ and if $G=\boldsymbol {\mu }_{p^n}$ , then $G_{\mathrm {abs}}={\mathbf {C}}_{p^n}$ . In any case, the order of $G_{\mathrm {abs}}$ is equal to the length of G.

Definition 2.1. The finite group $G_{\mathrm {abs}}$ is called the abstract finite group associated to G.

Lemma 2.2. The functor

$$ \begin{align*}G \,\mapsto\, G_{\mathrm{abs}} \end{align*} $$

establishes an equivalence of categories between the category of finite and linearly reductive group schemes over k and the category of finite groups with a normal and abelian p-Sylow subgroup.

Remark 2.3. A finite group G with a normal and abelian p-Sylow subgroup P is the same as a finite group with a unique and abelian p-Sylow subgroup. In this case, the Schur-Zassenhaus theorem implies that G is isomorphic to a semi-direct product $P\rtimes G/P$ .

Next, we study lifts to characteristic zero: let $W(k)$ be the ring of Witt vectors of k, let K be its field of fractions and let $\overline {K}$ be an algebraic closure of K. By [Reference Liedtke, Martin and MatsumotoLMM21, Proposition 2.4], there exist lifts of G as finite and flat group scheme over $W(k)$ . More precisely, $G^\circ $ and $G^{\mathrm {\acute {e}t}}$ even lift uniquely to $W(k)$ , but their extension class usually does not, see also [Reference Liedtke, Martin and MatsumotoLMM21, Example 2.6]. However, there is a unique lift ${\mathcal G}_{\mathrm {can}}$ of G to $W(k)$ that is characterised by being a semi-direct product of the unique lift of $G^\circ $ with the unique lift of $G^{\mathrm {\acute {e}t}}$ .

Definition 2.4. The lift $\mathcal {G}_{\mathrm {can}}\to \mathrm {Spec}\: W(k)$ is called the canonical lift of G. We set $G_{\mathrm {can}}:=\mathcal {G}_K\to \mathrm {Spec}\: K$ and also call it canonical lift.

Every other lift of G to some extension of $R\supseteq W(k)$ differs from ${\mathcal G}_{\mathrm {can},R}$ by a twist and thus, there is only one geometric lift of G to $\overline {K}$ up to isomorphism, namely $G_{\mathrm {can},\overline {K}}$ , see [Reference Liedtke, Martin and MatsumotoLMM21, Section 2.2].

Since $\overline {K}$ is algebraically closed and of characteristic zero, $G_{\mathrm {can},\overline {K}}$ is the constant group scheme associated to the finite group $G_{\mathrm {can}}(\overline {K})$ . In fact, we have $G^\circ \cong \prod _i \boldsymbol {\mu }_{p^{n_i}}$ for some $n_i$ ’s and if we set $N:=\max \{n_i\}_i$ , fix a primitive $p^N$ .th root of unity $\zeta _{p^N}$ and set $K_N:=K(\zeta _{p^N})$ , then $G_{\mathrm {can},K_N}$ is the constant group scheme associated to the finite group $G_{\mathrm {can}}(K_N)$ and we have $G_{\mathrm {can}}(K_N)=G_{\mathrm {can}}(\overline {K})$ .

Lemma 2.5. There exist isomorphisms of finite groups

$$ \begin{align*}G_{\mathrm{abs}} \,\cong\, G_{\mathrm{can}}(K_N) \,=\,G_{\mathrm{can}}(\overline{K}). \end{align*} $$

In particular, there exist isomorphisms

$$ \begin{align*}G_{\mathrm{can},K_N} \,\cong\, \left(\underline{G_{\mathrm{abs}}}\right)_{K_N} \quad \mbox{and} \quad G_{\mathrm{can},\overline{K}} \,\cong\, \left(\underline{G_{\mathrm{abs}}}\right)_{\overline{K}} \end{align*} $$

of finite group schemes over $K_N$ and $\overline {K}$ , respectively.

2.3 Representation theory

Let k be an algebraically closed field of characteristic $p>0$ , and let $W(k)$ , K and $\overline {K}$ be as in the previous section. The equivalence of Lemma 2.2 can be extended to representations and K-theory. If G is a finite group or a finite group scheme over k, we denote by $\mathrm {Rep}_k(G)$ the category of its k-linear and finite-dimensional representations. Let us first recall [Reference Liedtke, Martin and MatsumotoLMM21, Corollary 2.11].

Proposition 2.6. Let G be a finite and linearly reductive group scheme over k. Then, there exist canonical equivalences of categories

$$ \begin{align*}\mathrm{Rep}_{k}(G) \,\to\, \mathrm{Rep}_{K_N}(G_{\mathrm{can},K_N}) \,\to\, \mathrm{Rep}_{\overline{K}}(G_{\mathrm{can},\overline{K}}) \,\to\, \mathrm{Rep}_{\mathbb{C}}(G_{\mathrm{abs}}), \end{align*} $$

which are compatible with degrees, direct sums, tensor products, duals and simplicity.

Remark 2.7. In particular, this allows us to define characters or even a character table of G via Proposition 2.6 and $G_{\mathrm {abs}}$ .

Let $K_k(G)$ be the K-group associated to $\mathrm {Rep}_k(G)$ . In fact, $K_k(G)$ has a natural structure of a commutative ring with one, where the sum (resp. product) structure comes from direct sums (resp. tensor products) of representations. A straightforward application of Proposition 2.6 is the following.

Corollary 2.8. There exist isomorphisms of rings

$$ \begin{align*}K_k(G) \,\to\, K_{K_N}(G_{\mathrm{can},K_N}) \,\to\, K_{\overline{K}}(G_{\mathrm{can},\overline{K}}) \,\to\, K_{{\mathbb{C}}}(G_{\mathrm{abs}}). \end{align*} $$

3 McKay graph and McKay correspondence

In this section, we introduce the McKay graph associated to a finite and linearly reductive subgroup scheme G over an algebraically closed field k of characteristic $p\geq 0$ and a representation $\rho :G\to {\mathbf {GL}}_{n,k}$ . This induces a bijection between certain affine Dynkin diagrams and finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ . As an application, we establish a linearly reductive McKay correspondence.

3.1 McKay graph

Let G be a finite and linearly reductive group scheme over an algebraically closed field k of characteristic $p\geq 0$ . Let $\{\rho _i\}_i$ be the finite set of isomorphism classes of k-linear and simple representations of G. Following tradition, we assume that $\rho _0$ is the trivial representation. We fix a representation $\rho :G\to {\mathbf {GL}}(V)$ . If G is a subgroup scheme of ${\mathbf {SL}}_{n,k}$ or ${\mathbf {GL}}_{n,k}$ , then $\rho $ is usually the linear representation corresponding to the embedding of G into this linear algebraic group. By assumption, $\mathrm {Rep}_k(G)$ is semi-simple. Therefore, there exist unique integers $a_{ij}\in {\mathbb {Z}}_{\geq 0}$ for each i, such that we have isomorphisms of k-linear representations

$$ \begin{align*}\rho\otimes\rho_i \,\cong\, \bigoplus_{j}\, \rho_j^{\oplus a_{ij}}. \end{align*} $$

Associated to this data, we define the McKay graph, denoted $\Gamma (G,\{\rho _i\},\rho )$ :

  1. - The vertices are the $\{\rho _{i}\}_i$ . (Some sources exclude the the trivial representation $\rho _0$ .)

  2. - There are $a_{ij}$ edges from the vertex corresponding to $\rho _i$ to the vertex corresponding to $\rho _j$ .

We now establish a couple of elementary properties of this graph, which are well-known in the classical case and which immediately carry over to the linearly reductive situation. We leave the proof of the first lemma to the reader.

Lemma 3.1. We have

$$ \begin{align*}a_{ij} \,=\, \dim_k\, \mathrm{Hom}(\rho_i,\rho_j\otimes\rho). \end{align*} $$

In particular, if $\rho $ is self-dual, that is, $\rho \cong \rho ^\vee $ , then $a_{ij}=a_{ji}$ for all $i,j$ . In this case, we can consider $\Gamma (G,\{\rho _i\},\rho )$ as an undirected graph.

Lemma 3.2. Let $\rho :G\to {\mathbf {SL}}_{2,k}$ be a homomorphism of group schemes over k, considered as a 2-dimensional representation. Then, $\rho $ is self-dual.

Proof. Being a 2-dimensional representation, $\rho ^\vee $ is isomorphic to $\rho \otimes \det (\rho )$ and the lemma follows.

Lemma 3.3. Let $\rho :G\to {\mathbf {GL}}_{n,k}$ be a faithful representation. Then, every irreducible representation of G occurs as subrepresentation of $\rho ^{\otimes m}$ for some suitable m.

Proof. This is well-known for finite groups. Using Proposition 2.6, it carries over to finite and linearly reductive group schemes.

Corollary 3.4. If $\rho $ is a faithful representation, then the graph $\Gamma (G,\{\rho _i\},\rho )$ is connected.

Proof. It suffices to note that the number of paths in $\Gamma $ of length m that connect the vertices corresponding to $\rho _i$ and $\rho _j$ is equal to the multiplicity of $\rho _i$ in $\rho _j\otimes \rho ^{\otimes m}$ , see, for example, the proof of [Reference KirillovKi16, Theorem 8.13].

Let k be an algebraically closed field of characteristic $p>0$ , and let $W(k)$ , K and $\overline {K}$ be as in Section 2.2. There, we also discussed the canonical lift $G_{\mathrm {can}}$ of G over K and we saw that there exists an isomorphism of finite groups $G_{\mathrm {abs}}\cong G_{\mathrm {can}}(\overline {K})$ . Proposition 2.6 implies the following result.

Proposition 3.5. Let G be a finite and linearly reductive group scheme over k. Let $\{\rho _i\}_i$ be the set of isomorphism classes of simple representations of G. Let $\rho $ be a finite-dimensional representation of G.

  1. 1. Proposition 2.6 yields sets of representations $\{\rho _{\mathrm {can},i}\}_i$ and $\{\rho _{\mathrm {abs},i}\}_i$ of $G_{\mathrm {can}}$ and $G_{\mathrm {abs}}$ , respectively, which are the sets of isomorphism classes of simple representations of $G_{\mathrm {can}}$ and $G_{\mathrm {abs}}$ , respectively.

  2. 2. Proposition 2.6 yields representations $\rho _{\mathrm {can}}$ and $\rho _{\mathrm {abs}}$ of $G_{\mathrm {can}}$ and $G_{\mathrm {abs}}$ , respectively.

This data leads to a bijection of the three McKay graphs

$$ \begin{align*}\Gamma(G,\{\rho_i\}_i,\rho),\quad \Gamma(G_{\mathrm{can}},\{\rho_{\mathrm{can},i}\}_i,\rho_{\mathrm{can}}),\quad \Gamma(G_{\mathrm{abs}},\{\rho_{\mathrm{abs},i}\}_i,\rho_{\mathrm{abs}}). \end{align*} $$

3.2 McKay correspondence

We now run through the classical McKay correspondence [Reference McKayMc80] in our setting, where we follow [Reference KirillovKi16, Section 8.3]. Let k be an algebraically closed field of characteristic $p\geq 0$ and let G be a finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ .

  1. 1. The K-group $K_k(G)$ carries a symmetric bilinear form

    $$ \begin{align*}([V],[W])_0 \,:=\, \dim_k\,\mathrm{Hom}_G(V,W). \end{align*} $$
  2. 2. We consider the closed embedding $G\to {\mathbf {SL}}_{2,k}$ as a 2-dimensional representation $\rho :G\to {\mathbf {SL}}_{2,k}\to {\mathbf {GL}}_{2,k}$ and define an operator

    $$ \begin{align*}A\,:\,K_k(G)\,\to\, K_k(G),\qquad [V]\mapsto [V]\otimes\rho\,. \end{align*} $$

    Since $\rho $ is self-dual by Lemma 3.2, it follows that A is symmetric with respect to $(-,-)_0$ , see [Reference KirillovKi16, Lemma 8.12].

  3. 3. Using $\rho $ , we define a symmetric bilinear form on $K_k(G)\otimes _{\mathbb {Z}}{\mathbb {R}}$ by

    $$ \begin{align*}([V],[W]) \,:=\, \left([V],\, (2-A)[W]\right)_0, \end{align*} $$
    which is positive semi-definite. The class $\delta \in K_k(G)$ of the regular representation of G generates the radical of $(-,-)$ .
  4. 4. Let $\{\rho _i\}$ be the set of isomorphism classes of simple representations of G.

    1. (a) The McKay graph $\widehat {\Gamma }=\Gamma (G,\{\rho _i\}_i,\rho )$ is an affine Dynkin diagram of type $\widehat {A}_n$ , $\widehat {D}_n$ with $n\geq 4$ , $\widehat {E}_6$ , $\widehat {E}_7$ or $\widehat {E}_8$ .

    2. (b) After removing the vertex corresponding to the trivial representation $\rho _0$ from $\widehat {\Gamma }$ , we obtain a finite Dynkin diagram of type $A_n$ , $D_n$ with $n\geq 4$ , $E_6$ , $E_7$ or $E_8$ , respectively.

If G is a finite subgroup of ${\mathbf {SL}}_2({\mathbb {C}})$ , then these statements are part of the classical McKay correspondence, see [Reference McKayMc80] or [Reference KirillovKi16, Section 8.3]. In our setting of finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ , the above claims immediately follow from the classical McKay correspondence together with the lifting results Proposition 2.6 and Proposition 3.5. We then obtain the following analog of McKay’s theorem [Reference McKayMc80].

Theorem 3.6. Let k be an algebraically closed field of characteristic $p\geq 0$ . There exists a bijection between finite, non-trivial and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ up to conjugation and affine Dynkin diagrams of type

$$ \begin{align*}\begin{array}{lcl} \widehat{A}_n, \widehat{D}_n, \widehat{E}_6, \widehat{E}_7, \widehat{E}_8 &\quad \mbox{if} \quad & p=0 \quad \mbox{or} \quad p\geq7, \\ \widehat{A}_n, \widehat{D}_n, \widehat{E}_6, \widehat{E}_7 &\quad \mbox{if} \quad & p=5,\\ \widehat{A}_n, \widehat{D}_n &\quad \mbox{if} \quad & p=3 \mbox{ and} \\ \widehat{A}_n &\quad \mbox{if} \quad &p=2. \end{array} \end{align*} $$

Proof. If $p=0$ , then this is part of the classical McKay correspondence. If $p>0$ , then this follows from the linearly reductive McKay correspondence just discussed together with Hashimoto’s classification [Reference HashimotoHa15, Theorem 3.8] of finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ .

Remark 3.7. The linearly reductive group schemes corresponding to $\widehat {E}_6$ , $\widehat {E}_7$ and $\widehat {E}_8$ are étale and correspond to finite groups of order prime to p. The linearly reductive group scheme corresponding to $\widehat {A}_n$ (resp. $\widehat {D}_n$ ) is étale if and only if $p\nmid (n+1)$ (resp. $p\nmid (n-2)$ ). We refer to [Reference Liedtke and SatrianoLS14, Proposition 4.2] for details. Thus, even if p is sufficiently large, then one does not obtain a bijection in Theorem 3.6 with finite groups of order to prime to p only.

Remark 3.8. Steinberg [Reference SteinbergSt85] established many properties of the McKay graph and the McKay correspondence for finite subgroups of ${\mathbf {SU}}_2({\mathbb {C}})$ without using classification lists. It seems plausible to obtain a proof of Theorem 3.6 along these lines without using lifting results or classification lists.

4 Linearly reductive quotient singularities

In this section, we recall linearly reductive quotient singularities in the sense of [Reference Liedtke, Martin and MatsumotoLMM21] and some general results, including the canonical lift of such a singularity over the ring of Witt vectors. In dimension two, we establish a minimal resolution of singularities using G-Hilbert schemes as in the work of Ishii, Ito and Nakamura [Reference IshiiIs02, Reference Ishii and NakamuraIN19, Reference Ito and NakamuraIN99]. We also show that the canonical lift admits a unique minimal and simultaneous resolution of singularities. Finally, we discuss Klein singularities and their relation to rational double point singularities.

4.1 Quotient singularities

Let k be an algebraically closed field of characteristic $p\geq 0$ . If G is a finite and linearly reductive group scheme over k, if V is a finite-dimensional k-vector space and if $\rho :G\to {\mathbf {GL}}(V)$ is a linear representation, then we define the $\lambda $ -invariant of $\rho $ as in [Reference Liedtke, Martin and MatsumotoLMM21, Definition 2.7] to be

$$ \begin{align*}\lambda(\rho) \,:=\, \max_{\{{\mathrm{id}}\}\neq\boldsymbol{\mu}_n\subseteq G} \dim V^{\boldsymbol{\mu}_n}, \end{align*} $$

where $V^{\boldsymbol {\mu }_n}$ denotes the subspace of $\boldsymbol {\mu }_n$ -invariants. As explained in [Reference Liedtke, Martin and MatsumotoLMM21, Remark 2.8], the representation $\rho $ is faithful if and only if $\lambda (\rho )\neq \dim V$ . Moreover, $\rho $ contains no pseudo-reflections if and only if $\lambda (\rho )\leq \dim V-2$ and in this case, the representation $\rho $ is said to be small.

Definition 4.1. The representation $\rho $ is called very small if $\lambda (\rho )=0$ .

Note that in dimension two the notions of small and very small coincide. We refer to [Reference Liedtke, Martin and MatsumotoLMM21, Section 3] for the classification of finite and linearly reductive group schemes that admit a very small representation. By [Reference Liedtke, Martin and MatsumotoLMM21, Proposition 6.5], $\lambda (\rho )$ is equal to the dimension of the non-free locus of the induced G-action on the spectrum of the formal power series ring $k[[V^\vee ]]=(k[V^\vee ])^\wedge $ . Following [Reference Liedtke, Martin and MatsumotoLMM21, Definition 6.4] and using the linearisation result [Reference Liedtke, Martin and MatsumotoLMM21, Proposition 6.3], we define:

Definition 4.2. A linearly reductive quotient singularity over k is an isolated singularity that is analytically isomorphic to $\mathrm {Spec}\: k[[V^\vee ]]^G$ , where G is a finite and linearly reductive group scheme over k and where $\rho :G\to {\mathbf {GL}}(V)$ is a very small representation.

Remark 4.3. We set $U:=\mathrm {Spec}\: k[V^\vee ]$ or $U:=\mathrm {Spec}\: k[[V^\vee ]]$ and simply write $x\in X=U/G$ with the assumptions of Definition 4.2 implicitly understood.

Properties of these singularities have been studied in [Reference Liedtke, Martin and MatsumotoLMM21]: for their local étale fundamental groups, class groups and F-signatures, we refer to [Reference Liedtke, Martin and MatsumotoLMM21, Section 7].

By [Reference Liedtke, Martin and MatsumotoLMM21, Theorem 8.1], a linearly reductive quotient singularity determines the finite and linearly reductive group scheme G together with the very small representation $\rho :G\to {\mathbf {GL}}(V)$ uniquely up to isomorphism and conjugacy, respectively. It thus makes sense to refer to $\rho (G)\subset {\mathbf {GL}}(V)$ as the finite and linearly reductive subgroup scheme of ${\mathbf {GL}}(V)$ associated to the linearly reductive quotient singularity $x\in X=U/G$ . In particular, the classification of linearly reductive quotient singularities in dimension d is “the same” as the classification of very small, finite and linearly reductive subgroup schemes of ${\mathbf {GL}}_{d,k}$ up to conjugacy. We refer to [Reference Liedtke, Martin and MatsumotoLMM21, Section 3] for details and this classification.

4.2 Minimal resolution of singularities

If $x\in X$ is a normal surface singularity, then it admits a unique minimal resolution of singularities

$$ \begin{align*}\pi\,:\, Y\,\to\,X. \end{align*} $$

If $x\in X$ is moreover a rational singularity, then the exceptional locus of $\pi $ is a union of ${\mathbb {P}}^1$ ’s, whose dual intersection graph $\Gamma $ contains no cycles. The graph $\Gamma $ is called the type of $x\in X$ . If the type determines the singularity up to analytic isomorphism, then the singularity is said to be taut.

Over ${\mathbb {C}}$ , taut singularities have been classified by Laufer [Reference LauferLa73]. For example, two-dimensional finite quotient singularities over ${\mathbb {C}}$ are taut. Since two-dimensional linearly reductive quotient singularities over algebraically closed fields of positive characteristic are F-regular (see [Reference Liedtke, Martin and MatsumotoLMM21, Proposition 7.1]), they are taut by Tanaka’s theorem [Reference TanakaTa15].

Remark 4.4. Very small, finite and linearly reductive subgroup schemes of ${\mathbf {GL}}_{2,k}$ have been classified in [Reference Liedtke, Martin and MatsumotoLMM21, Theorem 3.4], extending Brieskorn’s classification [Reference BrieskornBr67, Satz 2.9] of small subgroups of ${\mathbf {GL}}_2({\mathbb {C}})$ . The types of the associated quotient singularities in terms of this classification are given by [Reference BrieskornBr67, Satz 2.11].

Finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ are automatically very small and they have been classified by Hashimoto [Reference HashimotoHa15, Theorem 3.8], extending Klein’s classification [Reference KleinKl84] of finite subgroups of ${\mathbf {SL}}_2({\mathbb {C}})$ and we refer to [Reference HashimotoHa15, Reference Liedtke and SatrianoLS14] for the types of the associated quotient singularities (see also Theorem 4.12 below).

4.3 The Ishii-Ito-Nakamura resolution

In [Reference Ito and NakamuraIN99], Ito and Nakamura showed that if G is a finite subgroup of ${\mathbf {SL}}_2({\mathbb {C}})$ , then a minimal resolution of singularities of ${\mathbb {C}}^2/G$ is provided by the G-Hilbert scheme $G\text{-}\mathrm {Hilb}({\mathbb {C}}^2)$ . Ishii [Reference IshiiIs02] extended this to quotient singularities ${\mathbb {C}}^2/G$ for G a finite and very small subgroup of ${\mathbf {GL}}_2({\mathbb {C}})$ and Ishii and Nakamura [Reference Ishii and NakamuraIN19] extended this to quotient singularities $U/G$ for G a finite and very small subgroup of ${\mathbf {GL}}_2(k)$ of order prime to p.

Let k be an algebraically closed field of characteristic $p\geq 0$ . Let G be a very small, finite and linearly reductive subgroup scheme of ${\mathbf {GL}}_{2,k}$ . Set $U:={\mathbb {A}}^2_k$ or $\widehat {{\mathbb {A}}}_k^2$ and let $x\in X:=U/G$ be the associated linearly reductive quotient singularity. By [Reference BlumeBl11], there exists a G-Hilbert scheme $G\text{-}\mathrm {Hilb}(U)$ over k that parametrises zero-dimensional G-invariant subschemes $Z\subset U$ (so-called clusters), such that the G-representation on $H^0(Z,{\mathcal O}_Z)$ is the regular representation. Taking a cluster Z to its G-orbit (see, for example, [Reference BlumeBl11, Remark 3.3]) induces a morphism

$$ \begin{align*}\pi\,:\,Y\,:=\,G\text{-}\mathrm{Hilb}(U) \,\to\, U/G\,=\,X \end{align*} $$

over k.

Theorem 4.5. The morphism $\pi $ is a minimal resolution of singularities.

Proof. If $p=0$ , then this is shown in [Reference Ishii and NakamuraIN19], extending [Reference Ito and NakamuraIN99]. If $p>0$ and G is of length prime to p, then this is shown in [Reference Ishii and NakamuraIN19]. However, this proof also works if G is linearly reductive and p divides the length of G.

Remark 4.6. By [Reference Liedtke, Martin and MatsumotoLMM21, Theorem 11.5], every two-dimensional F-regular singularity $x\in X$ is a linearly reductive quotient singularity, say $x\in X=U/G$ for some finite and linearly reductive subgroup scheme G of ${\mathbf {GL}}_{2,k}$ . By [Reference Liedtke, Martin and MatsumotoLMM21, Theorem 8.1], G and its embedding into ${\mathbf {GL}}_{2,k}$ are unique up to isomorphism and conjugacy, respectively. By Theorem 4.5, $G\text{-}\mathrm {Hilb}(U)\to U/G=X\ni x$ is the minimal resolution of singularities. In particular, every two-dimensional F-regular singularity can be resolved by a suitable G-Hilbert scheme.

4.4 Canonical lifts

Let k be an algebraically closed field of characteristic $p>0$ , let $W(k)$ be the ring of Witt vectors, let K be its field of fractions and let $\overline {K}$ be an algebraic closure. Let $x\in X=U/G$ be a d-dimensional linearly reductive singularity, where $U={\mathbb {A}}^d_k$ or $U=\widehat {{\mathbb {A}}}_k^d$ . Next, we set ${\mathcal U}:={\mathbb {A}}^d_{W(k)}$ or $\widehat {{\mathbb {A}}}^d_{W(k)}$ , respectively. In Section 2.2, we recalled the canonical lift ${\mathcal G}_{\mathrm {can}}$ of G over $W(k)$ . By [Reference Liedtke, Martin and MatsumotoLMM21, Proposition 2.9], there exists a unique lift of the G-action from U to ${\mathcal U}$ . From this, we obtain a flat family

(4.1) $$ \begin{align} \mathcal{X}_{\mathrm{can}}\,:=\,{\mathcal U}/{\mathcal G}_{\mathrm{can}} \,\to\, \mathrm{Spec}\: W(k)\,. \end{align} $$

of linearly reductive quotient singularities over $W(k)$ , whose special fibre over k is $x\in X=U/G$ .

Definition 4.7. The family (4.1) is called the canonical lift of the linearly reductive quotient singularity $x\in X=U/G$ .

By the Lefschetz principle, the geometric generic fibre of $\mathcal {X}_{\mathrm {can}}$ can be identified with a finite quotient singularity of the form ${\mathbb {C}}^d/G_{\mathrm {abs}}$ , where $G_{\mathrm {abs}}$ is the abstract group associated to G and the embedding of $G_{\mathrm {abs}}\to {\mathbf {GL}}_d({\mathbb {C}})$ corresponds to the embedding $G\to {\mathbf {GL}}_{d,k}$ provided by Proposition 2.6. The canonical lift is unique in the following sense:

Proposition 4.8. We keep the notations and assumptions. Let $W(k)\subseteq R$ be a finite extension of complete DVRs and let $\mathcal {X}\to \mathrm {Spec}\: R$ be a lift of $x\in X$ that is of the form ${\mathcal V}/{\mathcal G}\to \mathrm {Spec}\: R$ for some flat lift ${\mathcal G}$ of G to R and ${\mathcal V}\cong {\mathcal U}\times _RS$ . Then, there exists a finite extension $R\subseteq S$ of complete DVRs, such that

  1. 1. There exists an isomorphism

    $$ \begin{align*}{\mathcal G}_{\mathrm{can}} \times_{W(k)}S \,\cong\, {\mathcal G}\times_RS \end{align*} $$
    of group schemes over S.
  2. 2. The very small representation $\rho :G\to {\mathbf {GL}}_{d,k}$ corresponding to the G-action on U lifts uniquely to ${\mathcal G}_{\mathrm {can}}$ and ${\mathcal G}$ , respectively, and they become conjugate over S.

  3. 3. There is an isomorphism

    $$ \begin{align*}\mathcal{X}_{\mathrm{can}}\times_{W(k)}S \,\cong\, \mathcal{X}\times_RS \end{align*} $$
    of deformations of $x\in X$ over S.

Proof. Since ${\mathcal G}$ and ${\mathcal G}_{\mathrm {can},R}$ are lifts of G to R, they become isomorphic after passing to a finite extension $R\subseteq S$ , see also the discussion in Section 2.2. There, we also saw that the linear representation $\rho :G\to {\mathbf {GL}}_{d,k}$ lifts uniquely to ${\mathcal G}$ and ${\mathcal G}_{\mathrm {can}}$ , respectively, and that they become conjugate over S. From this, we deduce an isomorphism $\mathcal {X}_{\mathrm {can}}\times _{W(k)}S\cong \mathcal {X}\times _RS$ .

Remark 4.9. If $d\geq 3$ , then [Reference Liedtke, Martin and MatsumotoLMM21, Corollary 10.10] shows that a d-dimensional linearly reductive quotient singularity $x\in X$ admits precisely one lift over $W(k)$ , namely the canonical lift. If $d=2$ , then linearly reductive quotient singularities usually have positive dimensional deformation spaces and admit many non-isomorphic lifts to $W(k)$ , see [Reference Liedtke, Martin and MatsumotoLMM21, Section 12].

4.5 Simultaneous resolution of singularities

If $\mathcal {X}\to S$ is a deformation of a rational double point singularity $x\in X$ , then it admits a simultaneous resolution of singularities, but usually only after some finite base-change $S'\to S$ , see [Reference ArtinAr74]. In the case where $\mathcal {X}\to S$ is a family of rational surface singularities, then such a finite $S'\to S$ exists if S maps to the so-called Artin component inside the versal deformation space of $x\in X$ . Moreover, due to the existence of flops, these simultaneous resolutions (if they exist) are not unique in general.

In the special case where $x\in X$ is a two-dimensional linearly reductive quotient singularity over some algebraically closed field k of characteristic $p>0$ and where $\mathcal {X}_{\mathrm {can}}\to \mathrm {Spec}\: W(k)$ is the canonical lift of $x\in X$ , we will now show that there exists a simultaneous and minimal resolution of singularities over $W(k)$ and that it is unique. This simultaneous resolution can be most elegantly constructed using the Ishii-Ito-Nakamura resolution from Section 4.3 in families.

Let k be an algebraically closed field of characteristic $p>0$ . Let G be a finite and linearly reductive group scheme over k and let $\rho :G\to {\mathbf {GL}}_{2,k}$ be a very small representation. Let ${\mathcal G}_{\mathrm {can}}\to \mathrm {Spec}\: W(k)$ be the canonical lift of G and let $\widetilde {\rho }:{\mathcal G}_{\mathrm {can}}\to {\mathbf {GL}}_{2,W(k)}$ be the lift of $\rho $ to $W(k)$ . Let ${\mathcal U}:={\mathbb {A}}^2_{W(k)}$ and let

$$ \begin{align*}\mathcal{X}_{\mathrm{can}} \,:=\, {\mathcal U}/{\mathcal G}_{\mathrm{can}} \,\to\, \mathrm{Spec}\: W(k) \end{align*} $$

be the canonical lift of $x\in X$ . By [Reference BlumeBl11], there exists a ${\mathcal G}_{\mathrm {can}}$ -Hilbert scheme

$$ \begin{align*}{\mathcal G}_{\mathrm{can}}\text{-}\mathrm{Hilb}({\mathcal U}) \,\to\,\mathrm{Spec}\: W(k) \end{align*} $$

that parametrises ${\mathcal G}_{\mathrm {can}}$ -invariant subschemes $Z\subset {\mathcal U}$ that are finite and flat over $W(k)$ (so-called ${\mathcal G}_{\mathrm {can}}-$ clusters) and such that the ${\mathcal G}_{\mathrm {can}}$ -representation on $H^0(Z,{\mathcal O}_Z)$ is the regular representation. Taking such a cluster Z to its ${\mathcal G}_{\mathrm {can}}$ -orbit (see, for example, [Reference BlumeBl11, Remark 3.3]) induces a morphism

(4.2) $$ \begin{align} \widetilde{\pi}\,:\, \mathcal{Y}\,:=\,{\mathcal G}_{\mathrm{can}}\text{-}\mathrm{Hilb}({\mathcal U}) \,\to\, {\mathcal U}/{\mathcal G}_{\mathrm{can}}\,=\,\mathcal{X}_{\mathrm{can}} \end{align} $$

over $W(k)$ .

Theorem 4.10. Keeping the assumptions and notations

$$ \begin{align*}\widetilde{\pi} \,:\, \mathcal{Y} \,\to\, \mathcal{X}\,\to\,\mathrm{Spec}\: W(k) \end{align*} $$

is a simultaneous minimal resolution of singularities of the canonical lift $\mathcal {X}\to \mathrm {Spec}\: W(k)$ of the linearly reductive quotient singularity $x\in X=U/G$ .

  1. 1. The simultaneous resolution $\widetilde {\pi }$ is unique up to isomorphism.

  2. 2. The exceptional locus of $\widetilde {\pi }$ is a union of ${\mathbb {P}}^1_{W(k)}$ ’s meeting transversally and we denote by $\Gamma $ its dual intersection graph.

  3. 3. The special fibre and the generic fibre are linearly reductive quotient singularities of type $\Gamma $ and $\widetilde {\pi }$ identifies the components of the exceptional loci of $\widetilde {\pi }_k$ and $\widetilde {\pi }_K$ .

Proof. We recall that we defined $G_{\mathrm {can}}:={\mathcal G}_{\mathrm {can},K}$ , which will simplify the notation in the following. The generic and special fibre of (4.2) over K and k are isomorphic to

$$ \begin{align*}\begin{array}{cccccc} &G_{\mathrm{can}}\text{-}\mathrm{Hilb}({\mathcal U}_K) &\to& {\mathcal U}_K/G_{\mathrm{can}} &\to&\mathrm{Spec}\: K\\ \mbox{and} \quad& G\text{-}\mathrm{Hilb}(U) &\to& U/G &\to&\mathrm{Spec}\: k, \end{array} \end{align*} $$

respectively, by [Reference BlumeBl11, Remark 3.1]. Now, $G_{\mathrm {can},\overline {K}}\text{-}\mathrm {Hilb}({\mathcal U}_{\overline {K}}) \to {\mathcal U}_{\overline {K}}/G_{\mathrm {can},\overline {K}}$ and $G\text{-}\mathrm {Hilb}(U) \to U/G$ , are minimal resolutions of singularities by Theorem 4.5. Thus, $G_{\mathrm {can}}\text{-}\mathrm {Hilb}({\mathcal U}_K) \to {\mathcal U}_K/G_{\mathrm {can}}$ is a resolution of singularities, which is minimal since it is minimal over $\overline {K}$ . Thus, $\mathcal {Y}\to \mathcal {X}\to \mathrm {Spec}\: W(k)$ is a simultaneous minimal resolution of singularities.

The exceptional fibres $\widetilde {\pi }_K$ and $\widetilde {\pi }_k$ of $\widetilde {\pi }$ over K and k are unions of ${\mathbb {P}}^1$ ’s that intersect transversally. The types, that is dual resolution graphs, associated to $\widetilde {\pi }_K$ and $\widetilde {\pi }_k$ are the same (see also Remark 4.4) and we denote this graph by $\Gamma $ . In particular, the exceptional fibres of $\widetilde {\pi }_K$ and $\widetilde {\pi }_k$ have the same numbers of irreducible components. In particular, the specialisation maps of Néron-Severi lattices

$$ \begin{align*}\mathrm{NS}(\mathcal{Y}_K) \,\leftarrow\, \mathrm{NS}(\mathcal{Y}) \,\to\, \mathrm{NS}(\mathcal{Y}_k) \end{align*} $$

are isometries of lattices. These identify the components of the exceptional loci of $\widetilde {\pi }_K$ and $\widetilde {\pi }_k$ . Moreover, given such a component of the exceptional locus of $\widetilde {\pi }_k$ , it is isomorphic to ${\mathbb {P}}^1_k$ and it extends uniquely to a ${\mathbb {P}}^1_{W(k)}$ in the exceptional locus of $\widetilde {\pi }$ . This establishes claims (2) and (3).

It remains to prove claim (1): Let $\mathcal {Y}'\to \mathcal {X}\to \mathrm {Spec}\: W(k)$ be a simultaneous resolution of singularities that coincides with the minimal resolution on special and generic fibres, respectively. Let $\alpha :\mathcal {Y}^{\prime }_K\,\to \,\mathcal {Y}_K$ be an isomorphism over $\mathcal {X}_K$ and choose a relatively (to $\mathcal {X}$ ) ample invertible sheaf $\mathcal {L}$ on $\mathcal {Y}_K$ . Then, $\alpha ^*\mathcal {L}$ is relatively ample on $\mathcal {Y}_K$ . Since the types of the singularities of $\mathcal {X}_K$ and $\mathcal {X}_k$ are the same, the fibres of $\mathcal {Y}^{\prime }_K\to \mathcal {X}_K$ and $\mathcal {Y}^{\prime }_k\to \mathcal {Y}_k$ contain the same number of exceptional divisors. Thus, the specialisation map of Néron-Severi lattices $\mathrm {NS}(\mathcal {Y}^{\prime }_K)\to \mathrm {NS}(\mathcal {Y}^{\prime }_k)$ is an injective map between two lattices of the same rank and the same discriminant, whence an isometry. Similarly, the specialisation map of Néron-Severi lattices $\mathrm {NS}(\mathcal {Y}_K)\to \mathrm {NS}(\mathcal {Y}_k)$ is an isometry. In particular, we can identify the components of the exceptional loci of $\mathcal {Y}_K$ and $\mathcal {Y}_k$ . (If the types of $\mathcal {X}_k$ and $\mathcal {X}_K$ differ, then this specialisation map is usually only injective, there are usually more $(-2)$ -curves in the special fibre and this is also the place where the difficulties with flops begin.) In particular, $\mathcal {L}$ and $\alpha ^*\mathcal {L}$ extend to relative ample invertible sheaves on $\mathcal {Y}$ and $\mathcal {Y}'$ , respectively. By [Reference KovacsKo09, Theorem 5.14], the isomorphism $\alpha $ extends to an isomorphism $\mathcal {Y}'\to \mathcal {Y}$ over $\mathcal {X}$ and the claimed uniqueness follows.

4.6 Rational double point singularities and Klein singularities

We specialise Definition 4.2 to the following case.

Definition 4.11. A Klein singularity is a linearly reductive quotient singularity as in Definition 4.2 with $\dim V=2$ and $\det \rho =1$ , that is, $\rho $ is a homomorphism of G to ${\mathbf {SL}}_{2,k}$ .

In particular, a Klein singularity is a two-dimensional and linearly reductive quotient singularity, it is a rational surface singularity, and since $\det (\rho )=1$ , it is Gorenstein. Quite generally, rational and Gorenstein surface singularities are precisely the rational double point singularities [Reference ArtinAr66]. We refer to [Reference DurfeeDu79] or [Reference DemazureSSS] for more background on surface singularities and rational double point singularities.

If $x\in X$ is a rational double point, then its type $\Gamma $ is a simply-laced finite Dynkin graph. In characteristic zero, every rational double point singularity is taut. In positive characteristic, a Klein singularity is F-regular and thus, taut. In the case of rational double points, this also follows from Artin’s explicit classification [Reference ArtinAr77]. On the other hand, rational double point singularities that are not F-regular need not be taut. We have the following relation between rational double point singularities and Klein singularities in positive characteristic, see [Reference Liedtke, Martin and MatsumotoLMM21, Theorem 11.2].

Theorem 4.12. Let k be an algebraically closed field of characteristic $p\geq 0$ . Let $x\in X$ be a normal surface singularity over k.

  1. 1. If $p=0$ or $p\geq 7$ , then $x\in X$ is a Klein singularity if and only if $x\in X$ is a rational double point singularity.

  2. 2. If $p>0$ , then $x\in X$ is a Klein singularity if and only if $x\in X$ is a rational double point singularity and F-regular. The finite Dynkin graphs of these singularities are of type

    $$ \begin{align*}\begin{array}{lcl} A_n, D_n, E_6, E_7, E_8 &\quad \mbox{if} \quad & p\geq7, \\ A_n, D_n, E_6, E_7 &\quad \mbox{if} \quad & p=5,\\ A_n, D_n &\quad \mbox{if} \quad & p=3, \\ A_n &\quad \mbox{if} \quad &p=2. \end{array} \end{align*} $$

A Klein singularity $x\in X=U/G$ is a linearly reductive quotient singularity and thus, determines G and $\rho :G\to {\mathbf {SL}}_{2,k}$ up to isomorphism and conjugacy, respectively. Thus, the classification of Klein singularities boils down to the classification of finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ and we refer to Remark 4.4 for this classification.

Remark 4.13. Theorem 3.6 and Theorem 4.12 rely on the classification of finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ . It is therefore no surprise that the classification lists coincide.

5 Hecke correspondences

Let $x\in X=U/G$ be a Klein singularity over an algebraically closed field k of characteristic $p>0$ . In Section 2.2, we discussed the canonical lift $\mathcal {X}\to \mathrm {Spec}\: W(k)$ of $x\in X$ and in Section 4.5 we established a simultaneous minimal resolution of singularities of the canonical lift. More precisely, this simultaneous resolution was constructed using G-Hilbert schemes extending previous work of Ishii, Ito and Nakamura [Reference IshiiIs02, Reference Ishii and NakamuraIN19, Reference Ito and NakamuraIN99].

In this section, we refine this resolution of singularities as in the work of Ito and Nakamura [Reference Ito and NakamuraIN99] and Nakajima [Reference NakajimaNa96, Reference NakajimaNa01]: we eventually obtain a bijection between the components of the minimal resolution of singularities of the Klein singularity $x\in X=U/G$ and the simple and non-trivial representations of G using special Hecke correspondences.

5.1 The Ito-Nakamura resolution revisited

We first slightly extend the setup of Section 4.3. Let k be an algebraically closed field of characteristic $p>0$ and let $x\in X=U/G$ be a Klein singularity as in Definition 4.11, that is, we have $U={\mathbb {A}}^2_k$ and G a very small, finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ .

Let $\{\rho _i\}_{i\in I}$ be the set of isomorphism classes of simple representations of G. Given a finite-dimensional representation $\rho $ of G, there exist non-negative integers $\nu _i\in {\mathbb {Z}}_+$ , such that $\rho $ is isomorphic to $\bigoplus _i\rho _i^{\oplus \nu _i}$ and we combine these into a multi-index $\nu =(\{\nu _i\})\in {\mathbb {Z}}_{\geq 0}^I$ . We set $\dim (\nu ):=\sum _i\nu _i\dim \rho _i$ , which is the dimension of the representation associated to $\nu $ .

For any integer $n\geq 1$ , the G-action on U induces a G-action on the Hilbert scheme $\mathrm {Hilb}^n(U)$ , which parametrises zero-dimensional subschemes of length n of U. We consider the fixed point scheme

$$ \begin{align*}\mathrm{Hilb}^{n,G}(U) \,:=\, \left( \mathrm{Hilb}^n(U) \right)^G, \end{align*} $$

that is, the largest subscheme of $\mathrm {Hilb}^n(U)$ on which G acts trivially. It parametrises G-invariant and zero-dimensional subschemes of length n of U. Given $\nu =(\{\nu _i\})_i\in {\mathbb {Z}}_{\geq 0}^I$ with $\dim (\nu )=n$ , we define

$$ \begin{align*}H^\nu\,=\,\left\{ Z\,\in\,\mathrm{Hilb}^{n,G}(U)\,|\, H^0(Z,{\mathcal O}_Z)\,\cong\, \bigoplus \rho_i^{\oplus \nu_i}\mbox{ as }G\mbox{-representation} \right\}, \end{align*} $$

which defines a subscheme of $\mathrm {Hilb}^{n,G}(U)$ . Adapting [Reference KirillovKi16, Lemma 12.4], which follows [Reference Ito and NakamuraIN99, Section 9], to our situation, we have the following.

Lemma 5.1. We keep the assumptions and notations and let $n\geq 1$ be an integer.

  1. 1. The scheme $\mathrm {Hilb}^{n,G}(U)$ is smooth over k.

  2. 2. We have a decomposition

    $$ \begin{align*}\mathrm{Hilb}^{n,G}(U) \,=\, \bigsqcup_{\nu}\, H^\nu, \end{align*} $$
    where the disjoint union runs over all multi-indices $\nu \in {\mathbb {Z}}_{\geq 0}^I$ of dimension n. Each $H^\nu $ is a smooth subscheme of $\mathrm {Hilb}^{n,G}(U)$ .

Proof. Since U is two-dimensional, $\mathrm {Hilb}^n(U)$ is smooth over k by [Reference FogartyFo68, Theorem 2.4]. To show Claim (1), it remains to show that the fixed point scheme for the G-action is also smooth, which follows from Lemma 5.2 below. Then, Claim (2) is obvious.

Lemma 5.2. Let k be an algebraically closed field, let X be a scheme that is smooth over k, let G be a finite and linearly reductive group scheme over k and assume that G acts on X. Then, the fixed point scheme $X^G\subseteq X$ is smooth over k.

Proof. Let $x\in X^G$ . Passing to the completion of the local ring ${\mathcal O}_{X,x}$ , using that X is smooth over k and passing to coordinates such that the G-action is linear (this is always possible by [Reference SatrianoSa12, proof of Corollary 1.8]), we may assume that

$$ \begin{align*}\widehat{{\mathcal O}}_{X,x} \,\cong\, k[[u_1,...,u_d]] \end{align*} $$

and that the G-action is linear. In this description it is easy to see that the G-invariant subscheme of $\mathrm {Spec}\:\widehat {{\mathcal O}}_{X,x}$ is smooth (see also the proof of [Reference Ito and NakamuraIN99, Lemma 9.1]), which implies that $X^G$ is smooth near x.

An important special case is the regular representation of G, where we have $\nu _i=\dim \rho _i$ for all i and in this case, we will write $\delta $ for the corresponding multi-index. The dimension of $\delta $ is equal to the length of G. In this case, we have

$$ \begin{align*}\pi\,:\, H^\delta \,=\, G\text{-}\mathrm{Hilb}(U) \,\to\, U/G \,=\,X, \end{align*} $$

which we already studied in Section 4.3. There, we saw that $\pi $ is a minimal resolution of singularities of the Klein singularity $x\in X=U/G$ .

5.2 Hecke correspondences

We keep the assumptions and notations of the previous section. We let n be the length of G and let $\delta $ be the multi-index corresponding to the regular representation. For $i\in I$ , we set $\alpha _i:=(0,...,0,1,0...)$ with the non-zero entry in the i.th position. We define

$$ \begin{align*}B_i \,:=\, \left\{ J_1 \,\in\, H^{\delta-\alpha_i},\quad J_2\,\in\, H^\delta\quad |\quad J_2\subseteq J_1 \right\} \,\subseteq\, H^{\delta-\alpha_i}\times H^\delta, \end{align*} $$

which is a Hecke correspondence, see [Reference NakajimaNa96], [Reference NakajimaNa01, Section 6.1] or [Reference KirillovKi16, Section 12.4]. We let

$$ \begin{align*}E_i \,:=\, \mathrm{pr}_2(B_i)\,\subseteq\, H^\delta \,=\, G\text{-}\mathrm{Hilb}(U) \end{align*} $$

be the image under the projection $\mathrm {pr}_2$ onto the second factor.

In characteristic zero, the following result is due to Nakajima [Reference NakajimaNa96] and independently to Ito and Nakamura [Reference Ito and NakamuraIN99], see also [Reference KirillovKi16, Section 12.4].

Theorem 5.3. Let k be an algebraically closed field of characteristic $p\geq 0$ . Let G be a finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ , let $x\in X:=U/G$ be the associated Klein singularity, and let $\pi :H^\delta \to X$ be the Ito-Nakamura resolution.

  1. 1. The assignment $i\mapsto E_i$ defines a bijection between the set $\{\rho _i\}_i$ of isomorphism classes of non-trivial simple representations of G and the set of irreducible components of the exceptional divisor $\mathrm {Exc}(\pi )$ of $\pi $ .

  2. 2. If $i\neq j$ , then

    $$ \begin{align*}B_i\cdot B_j \,=\, A_{ij} \end{align*} $$
    where the numbering of the $\{\rho _i\}_i$ is as in [Reference Ito and NakamuraIN99, Section 10].

We thus obtain a bijection between the dual resolution graph of $\pi $ and the graph obtained by removing the vertex $\rho _0$ from the McKay graph of G with respect to the embedding $G\to {\mathbf {SL}}_{2,k}\to {\mathbf {GL}}_{2,k}$ .

Proof. The case-by-case analysis of [Reference Ito and NakamuraIN99, Section 12] carries directly over from finite subgroups of ${\mathbf {SL}}_2({\mathbb {C}})$ to finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ .

Here is a second proof of this theorem that uses our lifting results:

Proof. If $p=0$ , then the Lefschetz principle allows to reduce to $k={\mathbb {C}}$ , where the theorem is due to Ito, Nakajima and Nakamura [Reference Ito and NakamuraIN99, Reference NakajimaNa96].

We now assume $p>0$ . Let $\mathcal {X}_{\mathrm {can}}={\mathcal U}/{\mathcal G}_{\mathrm {can}}\to \mathrm {Spec}\: W(k)$ be the canonical lift of $x\in X=U/G$ (see Section 4.4) and let $\widetilde {\pi }:\mathcal {Y}\to \mathcal {X}_{\mathrm {can}}\to \mathrm {Spec}\: W(k)$ be the simultaneous resolution of singularities (see Section 4.5). Moreover, we have the abstract group $G_{\mathrm {abs}}$ and the canonical lift $G_{\mathrm {can}}$ associated to G (see Section 2.2) and we have a bijection of simple representations of $G_{\mathrm {abs}}$ , $G_{\mathrm {can}}$ and G by Proposition 2.6.

The generic fibre $\mathcal {X}_K$ is isomorphic to ${\mathcal U}_K/G_{\mathrm {can}}$ and using the Lefschetz principle, the geometric generic fibre can be identified with ${\mathbb {C}}^2/G_{\mathrm {abs}}$ . Over ${\mathbb {C}}$ , we have Theorem 5.3 for $G_{\mathrm {abs}}$ ${\mathbb {C}}^2/G_{\mathrm {abs}}$ and its minimal resolution of singularities by [Reference Ito and NakamuraIN99, Reference NakajimaNa96].

By the Lefschetz principle, we have it for $G_{\mathrm {can}}$ , $\mathcal {X}_K={\mathcal U}_{K}/G_{\mathrm {can}}$ and $\widetilde {\pi }_K$ . Using the comparison results Proposition 2.6 and Theorem 4.10, we obtain it for $\mathcal {X}_k={\mathcal U}_k/{\mathcal G}_{\mathrm {can},k}$ and $\widetilde {\pi }_k$ , that is, for $X=U/G$ , G and $\pi $ .

5.3 Local McKay correspondence

We now extend work of Ishii and Nakamura [Reference IshiiIs02, Reference Ishii and NakamuraIN19] to our linearly reductive setting. Let G be a very small, finite and linearly reductive subgroup scheme of ${\mathbf {GL}}_{2,k}$ and let $x\in X=U/G$ be the associated two-dimensional linearly reductive quotient singularity. Until the end of this section, we do not require it to be a Klein singularity. Note that in dimension two, linearly reductive quotient singularities are the same as F-regular singularities, see Remark 4.6.

Let $\{\rho _i\}_i$ be the set of simple representations of G. Let $\rho _0$ be the trivial representation and we choose our numbering of the $\rho _i$ ’s to be the one of [Reference Ishii and NakamuraIN19, Theorem 3.6]. Let $\pi :Y\to X$ be its minimal resolution of singularities and let $\{E_i\}$ be the irreducible components of the exceptional divisor $\textrm {Exc}(\pi )$ of $\pi $ . Then, there is exists a connection between the $\{\rho _i\}$ and the exceptional divisors of $\pi $ as follows - this is yet another version of the McKay correspondence.

Theorem 5.4. Keeping assumptions and notations, let ${\mathfrak m}\subset {\mathcal O}_U$ be the maximal ideal corresponding to the origin, let $y\in Y$ be a closed point, let $Z_y$ be the G-invariant cluster of U corresponding to y and let ${\mathcal I}_{Z_y}\subset {\mathcal O}_U$ be its ideal sheaf. Then, the G-representation on ${\mathcal I}_{Z_y}/{\mathfrak m}{\mathcal I}_{Z_y}$ is given by

$$ \begin{align*}\left\{ \begin{array}{ll} \rho_i\,\oplus\,\rho_0 &\quad \mbox{if }y\in E_i\,\backslash\,\bigcup_{j\neq i}E_j\\ \rho_i\,\oplus\,\rho_j\,\oplus\,\rho_0 & \quad \mbox{if }y\in E_i\cap E_j \quad \mbox{with} \quad i\neq j. \end{array} \right. \end{align*} $$

Proof. For $k={\mathbb {C}}$ , this is [Reference IshiiIs02, Theorem 7.1]. For k algebraically closed of arbitrary characteristic and G a very small and finite subgroup of ${\mathbf {GL}}_{2}(k)$ of order prime to p, this is [Reference Ishii and NakamuraIN19, Theorem 3.6]. However, these proofs also work if G is a very small and finite and linearly reductive subgroup scheme of ${\mathbf {GL}}_{2,k}$ .

Remark 5.5. In [Reference Ishii and NakamuraIN19], the G-quiver structure of $G\text{-}\mathrm {Hilb}(U)$ was studied in the case where G is a very small and finite subgroup of ${\mathbf {GL}}_{2}(k)$ . We leave the extension of these results to the linearly reductive case to the reader.

5.4 Reflexive sheaves on the minimal resolution

We end this section by shortly digressing on work of Wunram [Reference WunramWu88], Ishii and Nakamura [Reference Ishii and NakamuraIN19], which generalises work of Artin and Verdier [Reference Artin and VerdierAV85]. We keep the assumptions and notations from Section 5.3. Let $F\subset Y$ be the fundamental divisor of $\pi $ , see [Reference ArtinAr66].

Theorem 5.6. Keeping assumptions and notations, there exists a bijection between

  1. 1. the set of irreducible components $\{E_i\}$ of $\mathrm {Exc}(\pi )$ and

  2. 2. the set of non-trivial indecomposable full ${\mathcal O}_Y$ -modules $\{{\mathcal M}_i\}$ , special in the sense that $H^1(Y,{\mathcal M}_i^{\vee })=0$

This correspondence ${\mathcal M}_i\mapsto E_i$ is defined by

$$ \begin{align*}c_1({\mathcal M}_i)\cdot E_j \,=\, \delta_{i,j}\,. \end{align*} $$

The rank of ${\mathcal M}_i$ is equal to $c_1({\mathcal M}_i)\cdot F$ , the multiplicity of $E_i$ in F.

Proof. For $k={\mathbb {C}}$ , this is [Reference WunramWu88]. For k algebraically closed of arbitrary characteristic and G a very small and finite subgroup of ${\mathbf {GL}}_{2,k}$ of order prime to p, this is [Reference Ishii and NakamuraIN19, Theorem 2.8]. However, these proofs also work if G is a very small, finite and linearly reductive subgroup scheme of ${\mathbf {GL}}_{2,k}$ .

Here, a full ${\mathcal O}_Y$ -module is as defined in [Reference Ishii and NakamuraIN19, Definition 2.4]. By [Reference Ishii and NakamuraIN19, Corollary 2.5], the assignment ${\mathcal M}\mapsto \pi _*{\mathcal M}$ sets up a bijection between the set of (indecomposable) full ${\mathcal O}_Y$ -modules with the set of (indecomposable) reflexive ${\mathcal O}_X$ -modules. In particular, the above theorem yields a bijection between the set of irreducible components of $\mathrm {Exc}(\pi )$ and the set of non-trivial, indecomposable and reflexive ${\mathcal O}_X$ -modules.

Remark 5.7. The group scheme G plays no rôle in this theorem and the discussion thereafter. Probably, these results should be viewed as results on two-dimensional F-regular singularities (see Remark 4.6).

6 Conjugacy classes and Ito-Reid correspondence

For a finite group, the number of isomorphism classes of complex simple representations is equal to the number of conjugacy classes. Thus, one can choose a bijection between these two sets. In particular, for a finite subgroup of ${\mathbf {SL}}_2({\mathbb {C}})$ one can choose a bijection between the vertices of its McKay graph and the conjugacy classes of G. In [Reference Ito and ReidIR96], Ito and Reid gave a canonical bijection between these two sets. In this section, we extend this Ito-Reid correspondence to finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{2,k}$ . The main difficulty is to define a suitable notion of conjugacy classes for finite and linearly reductive group schemes.

6.1 Conjugacy classes

Given a finite group $G_{\mathrm {abs}}$ , a representation $\rho :G_{\mathrm {abs}}\to {\mathbf {GL}}_n({\mathbb {C}})$ and an element $g\in G_{\mathrm {abs}}$ , we have the trace $\mathrm {tr}(\rho (g))\in {\mathbb {C}}$ . This pairing $(\rho ,g)\mapsto \mathrm {tr}(\rho (g))$ induces a non-degenerate pairing between isomorphism classes of simple representations of $G_{\mathrm {abs}}$ and conjugacy classes. Thus, conjugacy classes can be thought of as being “dual” to semi-simple representations, which can be used to give an unusual definition of conjugacy classes. This definition generalises to finite and linearly reductive group schemes. More precisely, we make the following definition, which is inspired by a result of Serre [Reference SerreSe77, Section 11.4] and which we discuss in detail in Appendix B.2.

Definition 6.1. Let G be a finite and linearly reductive group scheme over an algebraically closed field k of characteristic $p\geq 0$ . Then, the set of conjugacy classes is defined to be $\mathrm {Spec}\: {\mathbb {C}}\otimes K_k(G)$ .

We discuss several approaches to conjugacy classes for finite group schemes in Appendix B - there, we hope to convince the reader that Definition 6.1 is the best for the purposes of this article. For example, by Proposition B.2 it is compatible with canonical lifts and there is a natural bijection with the set of conjugacy classes of the abstract group $G_{\mathrm {abs}}$ associated to G. Concerning the choice of field ${\mathbb {C}}$ in Definition 6.1, we refer to Remarks B.3.

6.2 An explicit bijection

For a finite and linearly reductive subgroup scheme G of ${\mathbf {SL}}_{2,k}$ , we have an associated embedding of the finite group $G_{\mathrm {abs}}$ into ${\mathbf {SL}}_2({\mathbb {C}})$ . Let $\{\rho _i\}_i$ be the set of isomorphism classes of semi-simple representations of G and let $\{\rho _{\mathrm {abs},i}\}_i$ be the corresponding set for $G_{\mathrm {abs}}$ obtained by Proposition 2.6. The associated McKay graphs $\Gamma (G,\rho ,\{\rho _i\}_i)$ and $\Gamma (G_{\mathrm {abs}},\rho _{\mathrm {abs}}, \{\rho _{\mathrm {abs},i}\}_i)$ coincide by Proposition 3.5 and we denote both by $\widehat {\Gamma }$ . Next, the group $G_{\mathrm {abs}}$ admits a presentation of the form

$$ \begin{align*}G_{\mathrm{abs}} \,=\, \left\langle A,B,C \,|\, A^r=B^s=C^t=ABC \right\rangle \end{align*} $$

for suitable non-negative integers $r,s,t$ . The non-trivial conjugacy classes of $G_{\mathrm {abs}}$ can be uniquely represented by the following elements

$$ \begin{align*}ABC,\quad A^i \mbox{ with }1\leq i\leq r-1, \quad B^i \mbox{ with }1\leq i\leq s-1, \quad C^i \mbox{ with }1\leq i\leq t-1. \end{align*} $$

This allows to give an explicit bijection between the conjugacy classes of $G_{\mathrm {abs}}$ and the vertices of $\widehat {\Gamma }$ . We refer to [Reference KirillovKi16, Section 2] for details. Using Proposition B.2, we obtain an explicit bijection between the conjugacy classes of G and the vertices of $\widehat {\Gamma }$ .

6.3 The Ito-Reid correspondence

If G is a finite subgroup of ${\mathbf {SL}}_2({\mathbb {C}})$ , then Ito and Reid [Reference Ito and ReidIR96] (see also [Reference ReidRe02]) constructed a canonical bijection. Let us run through [Reference ReidRe02, Section 2] and show that this can be carried over to our linearly reductive setting: let k be an algebraically closed field of characteristic $p>0$ and let G be a finite and linearly reductive subgroup scheme of ${\mathbf {SL}}_{2,k}$ .

  1. 1. Associated to G, we have the abstract group $G_{\mathrm {abs}}$ . By Proposition 2.6, an embedding $G\to {\mathbf {SL}}_{n,k}$ yields an embedding $G_{\mathrm {abs}}\to {\mathbf {SL}}_n({\mathbb {C}})$ . By Proposition B.2, we can identify conjugacy classes of G (in the sense of Definition 6.1) and conjugacy classes of $G_{\mathrm {abs}}$ , which allows us to define the age of a conjugacy class of G via the corresponding notion for $G_{\mathrm {abs}}$ as, for example, in [Reference ReidRe02, Section 2]. A conjugacy class of age 1 is called junior and if $n=2$ , then all conjugacy classes are junior.

Remark 6.2. Using the “toric mechanism” mentioned in [Reference ReidRe02, Section 2], one can define the age directly and without referring to lifts, but we will not pursue this here.

  1. 2. Let $x\in X:=U/G$ be the associated Klein singularity. We have the canonical lift $\mathcal {X}_{\mathrm {can}}\to \mathrm {Spec}\: W(k)$ and the simultaneous resolution of singularities $\widetilde {\pi }:\mathcal {Y}\to \mathcal {X}_{\mathrm {can}}\to \mathrm {Spec}\: W(k)$ by Theorem 4.10. Passing to geometric generic fibres and using the Lefschetz principle, we obtain the minimal resolution of singularities of ${\mathbb {C}}^2/G_{\mathrm {abs}}$ . The special fibre of $\widetilde {\pi }$ , the geometric generic fibre of $\widetilde {\pi }$ and the minimal resolution of ${\mathbb {C}}^2/G_{\mathrm {abs}}$ are crepant in the sense of [Reference ReidRe87] and we can identify the exceptional divisors of these three resolutions with each other, see Theorem 4.10. This way, we obtain an Ito-Reid correspondence between junior conjugacy classes of G and crepant divisors of the resolution [Reference ReidRe02, Theorem 2.1].

Remark 6.3. It seems reasonable that one can extend this correspondence to finite and linearly reductive subgroup schemes of ${\mathbf {SL}}_{n,k}$ with $n\geq 3$ , but we will not pursue this here.

7 Derived categories

Let G be a very small, finite and linearly reductive subgroup scheme of ${\mathbf {GL}}_{2,k}$ , let $x\in X:=U/G$ be the associated linearly reductive quotient singularity and let $\pi :Y\to X$ be its minimal resolution of singularities. Gonzalez-Sprinberg and Verdier [Reference Gonzalez-Sprinberg and VerdierGV83] gave an interpretation of the McKay correspondence as an isomorphism between the K-groups $K^G(U)$ and $K(Y)$ . Kapranov and Vasserot [Reference Kapranov and VasserotKV00] and Bridgeland, King and Reid [Reference Bridgeland, King and ReidBKR01] generalised this to an equivalence of derived categories ${\mathcal {D}}^G(U)$ and ${\mathcal {D}}(Y)$ . In this section, we extend this to our setting, following Ishii, Ito, Nakamura and Ueda [Reference IshiiIs02, Reference Ishii and NakamuraIN19, Reference Ishii and UedaIU15].

We have a commutative diagram

By Theorem 4.5, the minimal resolution $\pi $ can be constructed by the Ishii-Ito-Nakamura resolution

$$ \begin{align*}G\text{-}\mathrm{Hilb}(U) \,\to\, U/G \end{align*} $$

We let ${\mathcal {Z}}$ be the universal cluster over $G\text{-}\mathrm {Hilb}(U)$ , we identify $G\text{-}\mathrm {Hilb}(U)$ with Y and then, we have a commutative diagram

Let ${\mathcal {D}}(Y)$ be the derived category of coherent sheaves on Y. Let ${\mathcal {D}}^G(U)$ be the derived category of G-equivariant coherent sheaves on U. Following [Reference IshiiIs02] and [Reference Ishii and NakamuraIN19], we define two functors

$$ \begin{align*}\begin{array}{ccccc} \Psi &:& {\mathcal{D}}^G(U) &\to& {\mathcal{D}}(Y)\\ \Phi &:& {\mathcal{D}}(Y) &\to& {\mathcal{D}}^G(U) \end{array} \end{align*} $$

by

$$ \begin{align*}\Psi(-) \,:=\, [p_*\, \mathbf{L}q^*\, (-)]^G \end{align*} $$

and

$$ \begin{align*}\Phi(-) \,:=\, \mathbf{R}\pi_{U,*} \left( {\mathcal O}_{{\mathcal{Z}}}^\vee \,\otimes^{\mathbf{L}}\,\pi_Y^*(-\otimes\rho_0)\,\otimes^{\mathbf{L}}\, \pi_U^*K_{U}\right) [2], \end{align*} $$

where ${\mathcal O}_{{\mathcal {Z}}}^\vee :=\mathbf {R}Hom({\mathcal O}_{{\mathcal {Z}}},{\mathcal O}_{Y\times U})$ denotes the dual of ${\mathcal O}_{{\mathcal {Z}}}$ , where $-\otimes \rho _0:{\mathcal {D}}(Y)\to {\mathcal {D}}^G(Y)$ denotes the functor that attaches the trivial G-action and where $K_U$ denotes the canonical sheaf of U. We refer to [Reference Ishii and NakamuraIN19, Section 3.1] for details, conventions and notations.

Theorem 7.1. Keeping assumptions and notations, $\Phi $ is fully faithful and $\Psi $ is a left adjoint of $\Phi $ .

Proof. For G is a very small subgroup of ${\mathbf {GL}}_2({\mathbb {C}})$ , this is [Reference IshiiIs02, Section 6] and [Reference Ishii and UedaIU15, Proposition 1.1 and Lemma 2.9]. For k algebraically closed of arbitrary characteristic and G a very small and finite subgroup of ${\mathbf {GL}}_{2}(k)$ of order prime to p, this is [Reference Ishii and NakamuraIN19, Theorem 3.2]. However, these proofs also work if G is a very small, finite and linearly reductive subgroup scheme of ${\mathbf {GL}}_{2,k}$ .

A Hopf algebras

In the first section of the appendix, we study finite group schemes from the point of view of finite-dimensional Hopf algebras. We also recall the quantum double of a Hopf algebra, as well as the adjoint and the extended adjoint representation. Many results of this section should be well-known to the experts, but are somewhat scattered over the literature - especially, since many sources (sometimes implicitly) assume characteristic zero or work even over the complex numbers.

A.1 Generalities

If G is a finite group scheme over a field k, then the k-algebra $A:=H^0(G,{\mathcal O}_G)$ is commutative and finite-dimensional as k-vector space. The multiplication $m:G\times G\to G$ the inverse $i:G\to G$ and the neutral element $e:\mathrm {Spec}\: k\to G$ induce k-algebra homomorphisms $m^*:A\to A\otimes _k A$ , $i^*:A\to A$ and $e^*:A\to k$ , which turn A into a co-algebra over k with co-multiplication $m^*$ and antipode $S:=i^*$ , and thus, into a Hopf algebra over k. Since i is the inverse of G, the antipode S satisfies $S^2=\mathrm {id}_A$ , that is, A is an involutive Hopf algebra.

Conversely, if A is a finite-dimensional and commutative Hopf algebra over k, then it is involutive and $\mathrm {Spec}\: A$ is a finite group scheme over k. Moreover, A is a co-commutative Hopf algebra if and only if G is a commutative group scheme.

Example A.1. Let $G_{\mathrm {abs}}$ be a finite group. The group algebra $k[G_{\mathrm {abs}}]$ becomes a Hopf algebra by defining the co-multiplication to be $\Delta (g)=g\otimes g$ and the antipode to be $S(g)=g^{-1}$ .

  1. 1. Clearly, $k[G_{\mathrm {abs}}]$ is an involutive and co-commutative Hopf algebra. Moreover, $k[G_{\mathrm {abs}}]$ is commutative if and only if $G_{\mathrm {abs}}$ is commutative.

  2. 2. There exist examples of non-isomorphic finite groups $H_{\mathrm {abs}}$ and $G_{\mathrm {abs}}$ , whose group rings $k[H_{\mathrm {abs}}]$ and $k[G_{\mathrm {abs}}]$ are isomorphic as k-algebras.

    However, they are not isomorphic as Hopf algebras: Recall that an element $x\in B$ in a Hopf algebra B is called group-like if $\Delta (x)=x\otimes x$ . The set of group-like elements of B form a group. If $B=k[G_{\mathrm {abs}}]$ , then the group of group-like elements of B is isomorphic to $G_{\mathrm {abs}}$ and thus, recovers the group.

  3. 3. There exists an isomorphism of finite group schemes over k

    $$ \begin{align*}G\,\cong\,\mathrm{Spec}\: k[G_{\mathrm{abs}}]^*, \end{align*} $$
    where $k[G_{\mathrm {abs}}]^*$ denotes the dual Hopf algebra and where G denotes the constant group scheme over k associated to $G_{\mathrm {abs}}$ .

A.2 Simplicity

A Hopf algebra A is co-semi-simple if its dual Hopf algebra $A^*$ is semi-simple and A is bi-semi-simple if both A and $A^*$ are semi-simple.

Let G be a finite group scheme over k and let $A:=H^0(G,{\mathcal O}_G)$ be the associated Hopf algebra. To give a finite-dimensional and k-linear representation $\rho :G\to \mathrm {GL}(V)$ is the same as to give an A-co-module $V\to V\otimes _k A$ , see [Reference WaterhouseWa79, Section 3.2] for details. Using this equivalence, we see that G is linearly reductive if and only if A is co-semi-simple. For the classification of linearly reductive group schemes (see Section 2) in the language of Hopf algebras we refer to [Reference ChinCh92].

The following equivalences are probably well-known to the experts.

Proposition A.2. Let G be a finite group scheme over an algebraically closed field k of characteristic $p\geq 0$ . Let $A:=H^0(G,{\mathcal O}_G)$ be the Hopf algebra associated to G.

  1. 1. A is semi-simple if and only if G is étale.

  2. 2. A is co-semi-simple if and only if G is linearly reductive.

  3. 3. A is bi-semi-simple if and only if G is of length prime to p.

Proof. We already established Claim (2) above.

If G is étale, then $A\cong k[G_{\mathrm {abs}}]^*$ , where $G_{\mathrm {abs}}:=G(k)$ is the abstract group associated to G. Since group rings are co-semi-simple, A is semi-simple. Conversely, if A is semi-simple, then $\langle \varepsilon ,\int _H^r\rangle \neq 0$ by Maschke’s theorem for Hopf algebras (see, for example, [Reference LorenzLo18, Section 12.3.1] for notation, statement and proof), which implies that A is a separable k-algebra, which implies that G is étale over k. This establishes Claim (1).

Of course, (3) follows immediately from (1) and (2), but we can also give an independent proof: Being an involutive Hopf algebra, A is bi-semi-simple if and only if p does not divide $\dim _k A$ , see [Reference Etingof and GelakiEG98, Corollary 3.2] or [Reference Larson and RadfordLR88, Corollary 2.6]. The latter is equivalent to G being of length prime to p. This establishes Claim (3).

A.3 Adjoint representation

Let $G_{\mathrm {abs}}$ be a finite group and let k be an algebraically closed field of characteristic $p\geq 0$ (p may or may not divide the order of $G_{\mathrm {abs}}$ ). Then, the action of $G_{\mathrm {abs}}$ on itself by conjugation is a permutation representation and we denote by

$$ \begin{align*}\rho_{\mathrm{ad}} \,:\, G_{\mathrm{abs}} \,\to\, {\mathbf{GL}}(V_{\mathrm{ad}}) \end{align*} $$

the associated k-linear representation.

Remark A.3. To be more precise: Depending on whether one considers $x\mapsto gxg^{-1}$ or $x\mapsto g^{-1}xg$ , one should speak about left or right adjoint actions and representations. For our discussion, this is not important, as long as one chooses one of them and stays with it.

Remark A.4. Let us recall a couple of general results about $\rho _{\mathrm {ad}}$ , which are well-known (I thank Frank Himstedt for explaining them to me):

  1. 1. Let C be the set of conjugacy classes of $G_{\mathrm {abs}}$ and let $\{g_c\}_{c\in C}$ be a system of representatives. Then, we have an isomorphism of k-linear representations

    where $C(g_c)$ denotes the centraliser of $g_c\in G_{\mathrm {abs}}$ , where denotes the one-dimensional trivial representation and where $\mathrm {Ind}$ denotes induction from a subgroup. See, for example, [Reference Nagao and TsushimaNT89, page 171 and Exercise 1.5].
  2. 2. Let $g\in G_{\mathrm {abs}}$ and let

    be a decomposition into indecomposable representations, which is unique up to isomorphism and numbering.

    There is precisely one summand, say $V_{i_1}$ , that has a subrepresentation that is isomorphic to . Moreover, this subrepresentation is unique. Also, there is precisely one summand, say $V_{i_2}$ , that has a quotient representation that is isomorphic to . Moreover, this quotient representation is unique. Then, we have $i_1=i_2$ and set

    $$ \begin{align*}S(V) \,:=\, V_{i_1}\,=\, V_{i_2}, \end{align*} $$
    which is called the Scott representation of V. See, for example, [Reference Nagao and TsushimaNT89, page 296 and Theorem 8.4].

In particular, the dimension of the largest trivial subrepresentation (resp. quotient representation) of $\rho _{\mathrm {ad}}$ is equal to the number of conjugacy classes of $G_{\mathrm {abs}}$ .

If A is a finite dimensional Hopf algebra over k, then there is an adjoint action of A on itself (in fact, a left adjoint action and a right adjoint action), see, for example, [Reference MontgomeryMo93, Definition 3.4.1]. We will write

$$ \begin{align*}\mathrm{ad}\,:\,A \,\to\, \mathrm{End}(A) \end{align*} $$

or simply ${}^{\mathrm {ad}} A$ for this representation.

Remarks A.5.

  1. 1. If $G_{\mathrm {abs}}$ is a finite group, then the adjoint representation of $H:=k[G_{\mathrm {abs}}]$ can be identified with the k-linear extension from $G_{\mathrm {abs}}$ to H of the conjugation action of $G_{\mathrm {abs}}$ on itself.

  2. 2. Let G be a finite group scheme over k and let $A:=H^0(G,{\mathcal O}_G)$ be the $A:=H^0(G,{\mathcal O}_G)$ . There is an adjoint representation

    $$ \begin{align*}\rho_{\mathrm{ad}} \,:\,G\,\to\,{\mathbf{GL}}(V_{\mathrm{ad}}), \end{align*} $$
    which corresponds to the adjoint representation of the dual Hopf algebra $A^*$ .
    1. (a) If $G_{\mathrm {abs}}$ is a finite group and G is the constant group scheme associated to it, then A is isomorphic to the dual of $k[G_{\mathrm {abs}}]$ equipped with its usual Hopf algebra structure. This shows that $\rho _{\mathrm {ad}}$ should be defined via the adjoint representation of $A^*$ rather than A.

    2. (b) The representation $\rho _{\mathrm {ad}}$ should not be confused with the adjoint representation of G on its Lie algebra, see for example, [Reference WaterhouseWa79, page 100, Exercise 13]. The latter can be identified with a subquotient of $\rho _{\mathrm {ad}}$ .

A.4 Quantum doubles

If A is a finite-dimensional Hopf algebra over a field k, then Drinfeld [Reference DrinfeldDr87] defined a Hopf algebra $D(A):=(A^{\mathrm {op}})^* \bowtie A$ , called the quantum double or Drinfeld double, where the bicrossed product structure is defined using the co-adjoint representation of A on $A^*$ and the co-adjoint representation of $A^*$ on A, see also [Reference MontgomeryMo93, Definition 10.3.1].

Remarks A.6. Let A be a finite-dimensional Hopf algebra over k and let $D(A):=(A^{\mathrm {op}})^* \bowtie A$ be its quantum double.

  1. 1. As k-vector space, $D(A)$ is of dimension $(\dim _kA)^2$ .

  2. 2. A is a Hopf subalgebra of $D(A)$ via $\varepsilon \bowtie A$ and $(A^{\mathrm {op}})^*$ is a Hopf subalgebra of $D(A)$ via $(A^{\mathrm {op}})^*\bowtie 1$ .

  3. 3. If A is commutative and co-commutative, then we have $A^{\mathrm {op}}=A$ and using the definition of the quantum double, we obtain isomorphisms

    $$ \begin{align*}D(A)\,\cong\,A^*\otimes_k A \,\cong\, D(A^*) \end{align*} $$
    of Hopf algebras over k, where the tensor product is the trivial tensor product of Hopf algebras. If we set $G:=\mathrm {Spec}\: A$ , then we have $G^D\cong \mathrm {Spec}\: A^*$ , where $-^D$ denotes the Cartier dual group scheme, and we have an isomorphism
    $$ \begin{align*}\mathrm{Spec}\: D(A) \,\cong\, G\times_{\mathrm{Spec}\: k} G^D \end{align*} $$
    of finite and commutative group schemes over k.
  4. 4. In particular, $D(A)$ is commutative if and only if A is commutative and co-commutative.

The assertions on simplicity in the next proposition extend results of Witherspoon [Reference WitherspoonWi96, Proposition 1.2] from groups to group schemes. They are also related to general semi-simplicity results of quasi-triangular Hopf algebras in positive characteristic due to Etingof and Gelaki [Reference Etingof and GelakiEG98] and in this form, they might be known to the experts.

Proposition A.7. Let G be a finite group scheme over an algebraically closed field k of characteristic $p\geq 0$ and let $A:=H^0(G,{\mathcal O}_G)$ be the associated Hopf algebra. Then, the following are equivalent

$$ \begin{align*} \begin{array}{llll} (1) & D(A)\ \text{is semi-simple} & (1') & D(A^*)\ \text{is semi-simple} \\ (2) & D(A)\ \text{is co-semi-simple} \qquad \qquad \ & (2') & D(A^*)\ \text{is co-semi-simple} \\ (3) & A\ \text{is bi-semi-simple} & (3') & A^*\ \text{is bi-semi-simple} \\ (4) & G\ \text{is of length prime to}\ p \end{array} \end{align*} $$

Proof. The equivalences (1) $\Leftrightarrow $ (2) $\Leftrightarrow $ (3) and (1’) $\Leftrightarrow $ (2’) $\Leftrightarrow $ (3’) are shown in [Reference MontgomeryMo93, Corollary 10.3.13]. The equivalence (3) $\Leftrightarrow $ (3’) is trivial and finally, the equivalence (3) $\Leftrightarrow $ (4) was shown in Proposition A.2.

A.5 Extended adjoint representation

Given a Hopf algebra A, we defined the adjoint representation ${}^{\mathrm {ad}}A$ in Section A.3 above. There is a way to extend this to a representation

$$ \begin{align*}\mathrm{Ad} \,:\, D(A) \,\to\, \mathrm{End}(A) \end{align*} $$

the extended adjoint representation of A, which is denoted by ${}^{\mathrm {Ad}} A$ . For its definition, we refer to Zhu’s article [Reference ZhuZhu97], as well as [Reference Cohen and WestreichCW11] for subsequent work on this representation. We note that it can be described as an induced representation

(A.1)

where denotes the trivial one-dimensional representation and where we consider A as a subalgebra of $D(A)$ via $\varepsilon \bowtie A$ , see [Reference BurciuBu06] or [Reference JacobyJa17, Section 4.3.4].

Example A.8. Let $G_{\mathrm {abs}}$ be a finite group and let $H:=k[G_{\mathrm {abs}}]$ be the group algebra equipped with its usual Hopf algebra structure. The elements $g\in G_{\mathrm {abs}}$ form a basis of H as k-vector space and we denote by $\rho _g\in H^*$ the dual basis elements.

  1. 1. The multiplication of $D(H)$ is given by

    $$ \begin{align*}(\rho_h\bowtie g)\cdot(\rho_k\bowtie \ell) \,=\, \delta_{h,gkg^{-1}}\cdot \rho_h\bowtie g\ell, \end{align*} $$
    see, for example, [Reference JacobyJa17, Example 2.4.2].
    1. (a) The extended adjoint representation ${}^{\mathrm {Ad}} H$ is given by

      $$ \begin{align*}(\rho_h\bowtie k)\,\mapsto\, (g\,\mapsto\, \delta_{h^{-1},kgk^{-1}}\,\cdot\, kgk^{-1} ) \end{align*} $$
      see, for example, [Reference JacobyJa17, Example 2.5.3].
    2. (b) The unit of $H^*$ is $\varepsilon :=\sum _{g\in G_{\mathrm {abs}}}\rho _g$ and the restriction of ${}^{\mathrm {Ad}}H$ to $\varepsilon \bowtie H$ is ${}^{\mathrm {ad}}H$ . Thus, the extended adjoint representation extends the adjoint representation from an A-representation to a $D(A)$ -representation, whence the name.

    3. (c) The extended adjoint representation ${}^{\mathrm {Ad}}H$ is semi-simple. More precisely, the k-subvector space of H generated by the elements of a conjugacy class of $G_{\mathrm {abs}}$ is a simple $D(H)$ -subrepresentation of ${}^{\mathrm {Ad}}H$ . This gives a bijection between conjugacy classes of $G_{\mathrm {abs}}$ and simple subrepresentations of ${}^{\mathrm {Ad}}H$ . By [Reference JacobyJa17, Lemma 4.3.2], this is also true if p divides the order of $G_{\mathrm {abs}}$ .

    We refer to [Reference BurciuBu06, Reference Dijkgraaf, Pasquier and RocheDPR91, Reference GouldGo93, Reference WitherspoonWi96] for more results about the representation theory of $D(H)$ and ${}^{\mathrm {Ad}}H$ .

  2. 2. If we identify $H^{**}$ with H, then the multiplication of $D(H^*)$ is given by

    $$ \begin{align*}(g\bowtie \rho_h)\cdot(\ell\bowtie\rho_k) \,=\, \delta_{h,k} \,\cdot\, (g\cdot \ell) \bowtie \rho_h \end{align*} $$
    and the extended adjoint representation ${}^{\mathrm {Ad}} (H^*)$ is given by
    $$ \begin{align*}(h\bowtie \rho_k)\,\mapsto\, (\rho_g\,\mapsto\, \rho_{gh^{-1}}), \end{align*} $$
    which can be identified with the dual of the regular representation of $G_{\mathrm {abs}}$ . The restriction of ${}^{\mathrm {Ad}}(H^*)$ to $1\bowtie H^*$ is trivial, which is equal to the adjoint representation $^{\mathrm {ad}}(H^*)$ , which is also trivial.

Example A.9. Let A be a finite-dimensional Hopf algebra over k that is commutative and co-commutative.

  1. 1. The adjoint representations ${}^{\mathrm {ad}}A$ and ${}^{\mathrm {ad}}(A^*)$ are trivial.

By Remark A.6, we have isomorphisms

$$ \begin{align*}D(A)\,\cong\,A^*\otimes_k A \,\cong\, D(A^*) \end{align*} $$

of Hopf algebras over k, where the tensor product is the trivial tensor product of Hopf algebras. Thus, to give a representation $\rho :D(A)\to \mathrm {End}(V)$ is equivalent to giving two representations $\rho _1:A\to \mathrm {End}(V)$ and $\rho _2:A^*\to \mathrm {End}(V)$ , whose actions on V commute.

  1. 2. Using the description (A.1) of ${}^{\mathrm {Ad}}A$ as induced representation, we obtain ${}^{\mathrm {Ad}}A$ from the the trivial A-action on A (this is $\rho _1$ ) and the dual of the regular representation of $A^*$ (this is $\rho _2$ ).

  2. 3. Similarly, we obtain ${}^{\mathrm {Ad}}(A^*)$ from the trivial $A^*$ -action on $A^*$ and the dual of the regular representation of A.

B Conjugacy classes for finite group schemes

Let G be a finite group scheme over an algebraically closed field k of characteristic $p\geq 0$ . In the second section of the appendix, we discuss several approaches towards the notion of a conjugacy class for G. If $p=0$ , then all of them lead to the same notion, namely, the familiar one. All approaches look reasonable at first sight if $p>0$ and they lead to essentially “the same” answer if G is of length prime to p. In general however, they lead to different notions, all of which have their merits and drawbacks. This appendix serves as a motivation for Definition 6.1, but the discussion and results may be interesting in themselves.

B.1 First approach: via rational points

Let $G(k)$ be the group of k-rational points. We obtain an equivalence relation $\sim $ on this set by defining $g_1\sim g_2$ if and only if there exists $h\in G(k)$ such that $g_1=hg_2h^{-1}$ . The quotient $G(k)/\sim $ is the first candidate for the set of conjugacy classes.

Example B.1. If G is étale over k, then it is the constant group scheme associated to the finite group $G_{\mathrm {abs}}:=G(k)$ . In this case, $G(k)/\sim $ coincides with the set of conjugacy classes of $G_{\mathrm {abs}}$ .

If G is étale, which is automatic if $p=0$ , then this approach is satisfactory. However, if $p>0$ , then the connected-étale sequence (2.1) induces a bijection

$$ \begin{align*}G(k) \,\cong\, G^{\mathrm{\acute{e}t}}(k), \end{align*} $$

which is an isomorphism of finite groups. In particular, $G(k)/\sim $ depends on the maximal étale quotient $G^{\mathrm {\acute {e}t}}$ of G only. For example, in the extremal case where G is connected, we have $G(k)=\{1\}$ and then, $G(k)/\sim $ consists of one element, and we do not gain much information about G.

Concerning functoriality: if $\varphi :G\to H$ is a homomorphism of finite group schemes over k, then we have induced morphisms $G(k)\to H(k)$ , $G^{\mathrm {\acute {e}t}}\to H^{\mathrm {\acute {e}t}}$ and $G(k)/\sim \to H(k)/\sim $ .

B.2 Second approach: via representations and K-theory

If $G_{\mathrm {abs}}$ is a finite group of order prime to p, then the category $\operatorname {\mathrm {Rep}}_k(G)$ of k-linear and finite-dimensional $G_{\mathrm {abs}}$ -representations is semi-simple. In this case, the number of isomorphism classes of simple representations is equal to the number of conjugacy classes. More precisely, if $\rho $ is a k-linear and finite-dimensional representation of $G_{\mathrm {abs}}$ and $g\in G_{\mathrm {abs}}$ is an element, then

$$ \begin{align*}(\rho,g) \,\mapsto\,\mathrm{Tr}\left(\rho(g)\right)\,\in\, k \end{align*} $$

induces a pairing between representations and conjugacy classes. Since the character table of a finite group is a quadratic and invertible matrix (see, for example, [Reference SerreSe77, Proposition I.7]), this pairing is non-degenerate and one can think of conjugacy classes as being “dual” to simple representations.

This idea can be made precise as follows: Let F be a field, let $\mathrm {Cl}(G_{\mathrm {abs}})$ be the set of conjugacy classes of $G_{\mathrm {abs}}$ and let $F^{\mathrm {Cl}(G_{\mathrm {abs}})}$ be the ring of class functions on $G_{\mathrm {abs}}$ with values in F. Let g be the order of $G_{\mathrm {abs}}$ , let $\zeta _g\in {\mathbb {C}}$ be a primitive g.th primitive root of unity and assume that F contains ${\mathbb {Q}}(\zeta _g)$ . Since the characters of $G_{\mathrm {abs}}$ take values in F (here, we use ${\mathbb {Q}}(\zeta _g)\subseteq F$ ), we have injective ring homomorphisms

$$ \begin{align*}F \,\to\, F\otimes_{\mathbb{Z}} K_k(G_{\mathrm{abs}}) \,\stackrel{\gamma}{\longrightarrow}\, F^{\mathrm{Cl}(G_{\mathrm{abs}})} \end{align*} $$

and $\gamma $ is an isomorphism, see [Reference SerreSe77, Section 11.4] and the first paragraph of [Reference SerreSe77, Section 9.1]. In particular,

$$ \begin{align*}\mathrm{Spec}\:\, F\otimes_{\mathbb{Z}} K_k(G_{\mathrm{abs}}) \end{align*} $$

is a finite set consisting of maximal ideals only and the topology is discrete. Since $\gamma $ is an isomorphism, the cardinality of this set is equal to that of $\mathrm {Cl}(G_{\mathrm {abs}})$ . This carries over to finite and linearly reductive schemes as follows.

Proposition B.2. Let G be a finite and linearly reductive group scheme over k and let $G_{\mathrm {abs}}$ be the associated abstract group. Let g be the length of G, fix a g.th root of unity $\zeta _g\in {\mathbb {C}}$ and let F be a field that contains ${\mathbb {Q}}(\zeta _{g})$ . Then, there exista canonical bijection of sets

$$ \begin{align*}\left\{\begin{array}{l} \mbox{conjugacy classes of }G_{\mathrm{abs}} \end{array}\right\} \,\to\, \mathrm{Spec}\:\, F\otimes_{\mathbb{Z}} K_k(G_{\mathrm{abs}}) \end{align*} $$

and homeomorphisms

$$ \begin{align*}\mathrm{Spec}\:\, F\otimes_{\mathbb{Z}} K_k(G_{\mathrm{abs}}) \,\leftarrow\, \mathrm{Spec}\:\, F\otimes_{\mathbb{Z}} K_{\overline{K}}(G_{\mathrm{abs}}) \,\leftarrow\, \mathrm{Spec}\:\, F\otimes_{\mathbb{Z}} K_k(G). \end{align*} $$

Proof. If c is a conjugacy class of $G_{\mathrm {abs}}$ , then $P_{0,c}$ (notation as in [Reference SerreSe77, Section 11.4, Proposition 30]) is an element of $\mathrm {Spec}\: F\otimes K_k(G_{\mathrm {abs}})$ and by loc. cit. this defines a bijection. The isomorphisms of Corollary 2.8 induce homeomorphisms and thus, bijections of spectra as stated.

Remarks B.3.

  1. 1. The maximal abelian extension ${\mathbb {Q}}\subset {\mathbb {Q}}^{\mathrm {ab}}$ is generated by all roots of unity by the Kronecker-Weber theorem. Thus, if we have ${\mathbb {Q}}^{\mathrm {ab}}\subseteq F$ , then we have a field that works independent of the length of G. In Definition 6.1, we have chosen $F={\mathbb {C}}$ as this field may be more familiar than ${\mathbb {Q}}^{\mathrm {ab}}$ .

  2. 2. In [Reference SerreSe77, Section 11.4], Serre described $\mathrm {Spec}\: A\otimes _{\mathbb {Z}} K_k(G_{\mathrm {abs}})$ , where $A={\mathbb {Z}}[\zeta _g]$ . Using Corollary 2.8, we obtain a homeomorphism

    $$ \begin{align*}\mathrm{Spec}\: A\otimes_{\mathbb{Z}} K_k(G) \,\to\,\mathrm{Spec}\: A\otimes_{\mathbb{Z}} K_k(G_{\mathrm{abs}}). \end{align*} $$

    In particular, Serre’s results from loc.cit. carry over to $A\otimes _{\mathbb {Z}} K_k(G)$ . For the purposes of this article, we are only interested in the fibre over $0\in \mathrm {Spec}\: F$ with $F=\mathrm {Frac}(A)={\mathbb {Q}}(\zeta _g)$ , that is, the Zariski open subset $\mathrm {Spec}\: F\otimes K_k(G)\subset \mathrm {Spec}\: A\otimes K_k(G)$ .

Example B.4. Let G be the group scheme $\boldsymbol {\alpha }_p$ or ${\mathbf {C}}_p$ over the algebraically closed field k of characteristic $p>0$ . Then, $\operatorname {\mathrm {Rep}}_k(G)$ is not semi-simple: the only simple k-linear representation of G is the trivial one-dimensional representation . Thus, induces an isomorphism of rings $K_k(G)\cong {\mathbb {Z}}$ and $\mathrm {Spec}\: F\otimes _{\mathbb {Z}} K_k(G)$ consists of one point only. The approach to conjugacy classes in this subsection may therefore lead to somewhat unexpected results if G is not linearly reductive.

Proposition B.5. Let $\varphi :G\to H$ be a morphism of finite and linearly reductive group schemes over k. Let $\varphi _{\mathrm {abs}}:G_{\mathrm {abs}}\to H_{\mathrm {abs}}$ be the induced homomorphism of their associated abstract groups. Let F be a field that contains ${\mathbb {Q}}(\zeta _g,\zeta _h)$ , where g (resp. h) denotes the length of G (resp. H).

  1. 1. There maps $\varphi $ and $\varphi _{\mathrm {abs}}$ induce ring homomorphisms $K_k(H)\to K_k(G)$ and $K_{\overline {K}}(H_{\mathrm {abs}})\to K_{\overline {K}}(G_{\mathrm {abs}})$ , respectively. We obtain a commutative diagram of continuous maps

    whose vertical arrows are the homeomorphisms from Proposition B.2.
  2. 2. Let $G_{\mathrm {abs}}/\sim \to H_{\mathrm {abs}}/\sim $ be the map on conjugacy classes induced by $\varphi _{\mathrm {abs}}$ . We obtain a commutative diagram of maps of sets

    whose vertical maps are the bijections from Proposition B.2.

Proof. Clearly, $\varphi $ induces a morphism $K_k(H)\to K_k(G)$ of rings since every H-representation becomes a G-representation via $\varphi $ . Similarly, $\varphi _{\mathrm {abs}}$ induces a ring homomorphism $\varphi _{K,\mathrm {abs}}\,:\,K_{\overline {K}}(H_{\mathrm {abs}})\to K_{\overline {K}}(G_{\mathrm {abs}})$ . We leave it to the reader to check the compatibility of these maps with the homeomorphisms of Proposition B.2.

To check commutativity of the second diagram, let $g\in G_{\mathrm {abs}}$ . With the notations and definitions of [Reference SerreSe77, Section 11.4], it is easy to see that we have

$$ \begin{align*}P_{0,\varphi_{\mathrm{abs}}(g)} \,=\,\varphi_{K,\mathrm{abs}}^{-1}(P_{0,g}) \,=\, \varphi_{K,\mathrm{abs}}^\sharp(P_{0,g}), \end{align*} $$

where

$$ \begin{align*}\varphi_{K,\mathrm{abs}}^\sharp \,:\,\mathrm{Spec}\: F\otimes K_{\overline{K}}(G_{\mathrm{abs}}) \,\to\, \mathrm{Spec}\: F\otimes K_{\overline{K}}(H_{\mathrm{abs}}) \end{align*} $$

is the induced map on spectra. Since the image of the conjugacy class $[g]$ of $G_{\mathrm {abs}}$ is the conjugacy class $[\varphi _{\mathrm {abs}}(g)]$ of $H_{\mathrm {abs}}$ , the assertion follows.

B.3 Third approach: the scheme of conjugacy classes

Just as group schemes generalise the notion of a group, one could try to replace the set of conjugacy classes by a suitable notion of scheme of conjugacy classes. More precisely, let G be a finite group scheme over k and let $\underline {\mathrm {Aut}}(G)$ be the automorphism group scheme of G. For every scheme $T\to \mathrm {Spec}\: k$ , we have the set $G(T)$ of T-valued points of G and a conjugation action of $G(T)$ on $G(T)$ . This induces a morphism of schemes $G\to \underline {\mathrm {Aut}}(G)$ . We will say that two elements of $G(T)$ are equivalent if they differ by such an automorphism and we denote the resulting equivalence relation by $\sim $ .

We obtain a functor from the category of schemes over k to sets

$$ \begin{align*}\begin{array}{cccccc} \underline{\mathrm{Conj}}_G &:& (\mathrm{Schemes}/k) &\to& (\mathrm{Sets})\\ && T &\mapsto& G(T)/\sim &. \end{array} \end{align*} $$

This functor should somehow represent the conjugacy classes of G in the sense of schemes.

Proposition B.6. Let G be a finite group scheme over k. Then, the functor $\underline {\mathrm {Conj}}_G$ is representable by a scheme $\mathrm {Conj}_G$ , which is finite over $\mathrm {Spec}\: k$ .

Proof. Set $V:=H^0(G,{\mathcal O}_G)$ and let $\rho _{\mathrm {ad}}:G\to {\mathbf {GL}}(V)$ be the adjoint representation of G (see Appendix A.3). Then, the functor $\underline {\mathrm {Conj}}_G$ can be rephrased as the functor that associates to each $T\to \mathrm {Spec}\: k$ the quotient of $V\times {\mathcal O}_T$ modulo $G(T)$ . This amounts to representing the quotient $V/G$ by a scheme. Since V is a vector space and thus, can be identified with an affine scheme, and since G is a finite group scheme, this quotient is representable by a scheme. In fact, it is representable by the spectrum of the invariant ring $\mathrm {Spec}\: V^G$ .

Definition B.7. $\mathrm {Conj}_G$ is called the scheme of conjugacy classes of G.

Remark B.8. If $\rho _{\mathrm {ad}}:G\to {\mathbf {GL}}(H^0(G,{\mathcal O}_G))$ is the adjoint representation, then the previous proof shows that the length of $\mathrm {Conj}_G$ over $\mathrm {Spec}\: k$ is equal to the dimension of the maximal trivial subrepresentation of $\rho _{\mathrm {ad}}$ . If $G_{\mathrm {abs}}$ is a finite group or if G is a finite and étale group scheme over $\mathrm {Spec}\: k$ , then we gave an explicit description of this maximal trivial subrepresentation of $\rho _{\mathrm {ad}}$ in Remark A.4.

Examples B.9.

  1. 1. If G is étale over k, then G is isomorphic to the constant group scheme associated to $G_{\mathrm {abs}}:=G(k)$ . In this case, $\mathrm {Conj}_G$ is a disjoint union of copies of $\mathrm {Spec}\: k$ , one copy for each conjugacy class of the abstract group $G_{\mathrm {abs}}$ .

  2. 2. If G is commutative, then the adjoint representation is trivial and thus, $\mathrm {Conj}_G$ is isomorphic to the scheme underlying G. For example, if $G=\boldsymbol {\alpha }_p$ or $G=\boldsymbol {\mu }_p$ , then $\mathrm {Conj}_G$ is a non-reduced scheme of length p with reduction $(\mathrm {Conj}_G)_{\mathrm {red}}\cong \mathrm {Spec}\: k$ .

In characteristic zero, all finite group schemes are étale and Case (1) applies. However, in characteristic $p>0$ , the second example shows that for non-reduced group schemes $\mathrm {Conj}_G$ may also be non-reduced. The reduction $(\mathrm {Conj}_G)_{\mathrm {red}}$ is related to our first approach to conjugacy classes:

Proposition B.10.

  1. 1. A morphism $G\to H$ of finite group schemes over k induces a morphism $\mathrm {Conj}_G\to \mathrm {Conj}_H$ of schemes over k.

  2. 2. Assume $p>0$ and let $G\to G^{\mathrm {\acute {e}t}}$ be the maximal étale quotient of the finite group scheme G over k. Then, reduction identifies $G_{\mathrm {red}}$ with $G^{\mathrm {\acute {e}t}}$ and the induced natural inclusion $G^{\mathrm {\acute {e}t}}\to G$ induces an isomorphism

    $$ \begin{align*}\mathrm{Conj}_{G^{\mathrm{\acute{e}t}}} \,\to\, \left( \mathrm{Conj}_G \right)_{\mathrm{red}} \end{align*} $$
    of schemes over k. In particular, $(\mathrm {Conj}_G)_{\mathrm {red}}$ is a disjoint union of copies of $\mathrm {Spec}\: k$ with one copy for each conjugacy class of the abstract group $G_{\mathrm {abs}}:=G(k)=G^{\mathrm {\acute {e}t}}(k)$ .

Proof. We first prove Claim (1). If $G_{\mathrm {abs}}\to H_{\mathrm {abs}}$ is a homomorphism of groups, then we get a well-defined induced map of conjugacy classes. Thus, if $G\to H$ is as in (1) and if T is a scheme over k, then we get a well-defined map $(G(T)/\sim )\to (H(T)/\sim )$ . This induces a morphism of functors $\underline {\mathrm {Conj}}_G\to \underline {\mathrm {Conj}}_H$ and thus, a morphism of schemes $\mathrm {Conj}_G\to \mathrm {Conj}_H$ .

Let G be as in Claim (2). Let $G_{\mathrm {red}}\to G$ be the reduction, which is a morphism of group schemes. We thus obtain canonical homomorphisms $G^{\mathrm {\acute {e}t}}\to G\to G^{\mathrm {\acute {e}t}}$ of group schemes over k, which (by Claim (1)) induce morphisms of their associated schemes of conjugacy classes

$$ \begin{align*}\mathrm{Conj}_{G^{\mathrm{\acute{e}t}}} \,\to\, \mathrm{Conj}_G \,\to\, \mathrm{Conj}_{G^{\mathrm{\acute{e}t}}}. \end{align*} $$

The composition is the identity. Since $\mathrm {Conj}_{G^{\mathrm {\acute {e}t}}}$ is reduced, we obtain a factorisation

(B.1) $$ \begin{align} \mathrm{Conj}_{G^{\mathrm{\acute{e}t}}} \,\to\, \left(\mathrm{Conj}_G\right)_{\mathrm{red}} \,\to\, \mathrm{Conj}_{G^{\mathrm{\acute{e}t}}}. \end{align} $$

All these schemes are reduced and finite over the algebraically closed field k. Thus, all of them are finite disoint unions of copies of $\mathrm {Spec}\: k$ . To prove that the morphisms in (B.1) are isomorphisms, it suffices to check that the induced maps on k-rational points are bijections. This follows easily from the fact that the maps $G_{\mathrm {red}}(k)\to G(k)\to G^{\mathrm {\acute {e}t}}(k)$ are bijections.

If G is moreover linearly reductive, then the length of $\mathrm {Conj}_G$ is related to the approach to conjugacy classes from Section B.2.

Proposition B.11. Let G be a finite and linearly reductive group scheme over k, let $G_{\mathrm {abs}}$ be the abstract group associated to G and let F be a field as in Proposition B.2. Then,

$$ \begin{align*}\mathrm{length}_k\, \mathrm{Conj}_G \,=\, | \mathrm{Spec}\: F\otimes_{\mathbb{Z}} K_k(G) | \end{align*} $$

and this length agrees with the number of conjugacy classes of $G_{\mathrm {abs}}$ .

Proof. By Remark B.8, the length of $\mathrm {Conj}_k(G)$ is equal to the dimension of the largest trivial subrepresentation of $\rho _{\mathrm {ad}}$ . Since $\mathrm {Rep}_k(G)$ is semi-simple, this is the same as the multiplicity of in $\rho _{\mathrm {ad}}$ . Using the isomorphism from Prop 2.6, this multiplicity is the same as the multiplicity of in the adjoint representation of $G_{\mathrm {abs}}$ . This latter multiplicity is equal to the number of conjugacy classes of $G_{\mathrm {abs}}$ , see, for example, Remark A.4. By Proposition B.2, this number is equal to the cardinality of $\mathrm {Spec}\: F\otimes _{\mathbb {Z}} K_k(G)$ .

B.4 Fourth approach: via lifting to characteristic zero

Another idea might be to use lifting to characteristic zero: Let k be an algebraically closed field of characteristic $p>0$ , let $W(k)$ be the ring of Witt vectors of k and let K be field of fractions of $W(k)$ .

Example B.12 (Mumford–Oort)

Let $\varphi :\boldsymbol {\mu }_p\to \underline {\mathrm {Aut}}(\boldsymbol {\alpha }_p)\cong {\mathbb {G}}_m$ be a non-trivial homomorphism and let $G:=\boldsymbol {\alpha }_p\rtimes _\varphi \boldsymbol {\mu }_p$ be the corresponding semi-direct product group scheme. Then, G is a non-commutative group scheme of length $p^2$ over k. Oort and Mumford [Reference Mumford and OortMO68, Introduction, Example (-B)] (but see also [Reference OortOo71, page 266]) showed that there does not exist a lift of G to any extension of $W(k)$ .

In particular, one cannot define conjugacy classes by first lifting G over some possibly ramified extension of $W(k)$ , then passing to the geometric generic fibre of the lift, which would be the constant group associated to an abstract finite group $G_{\mathrm {abs}}$ , and finally use the conjugacy classes of $G_{\mathrm {abs}}$ as a replacement for the conjugacy classes of G. The reason is simply that lifts may not exist to start with.

In the following two cases, lifts do exist and we leave the straightforward proofs of our assertions to the reader:

Remark B.13. Assume that G is étale over k. Then, there exists a unique flat lift of G over $W(k)$ , which is the constant group scheme associated to $G_{\mathrm {abs}}:=G(k)$ over $W(k)$ . The constant scheme associated to the set of conjugacy classes of $G_{\mathrm {abs}}$ is the unique a flat lift of the scheme $\mathrm {Conj}_G$ over $W(k)$ , which is étale over $W(k)$ . More precisely, it is the constant scheme associated to the set of conjugacy classes of $G_{\mathrm {abs}}$ over $W(k)$ .

Remark B.14. Assume that G is linearly reductive. Let $G_{\mathrm {abs}}$ abstract group associated to G and recall that their representation categories are equivalent by Proposition 2.6. Proposition B.2 shows that Definition 6.1 is compatible with $G_{\mathrm {abs}}$ .

If ${\mathcal G}\to \mathrm {Spec}\: W(k)$ is a flat lift of G over $W(k)$ , for example, the canonical lift ${\mathcal G}_{\mathrm {can}}$ , then

$$ \begin{align*}\mathrm{Spec}\:\, H^0({\mathcal G},{\mathcal O}_{{\mathcal G}})^{{\mathcal G}}\,\to\,\mathrm{Spec}\: W(k), \end{align*} $$

where the invariants are taken with respect to the adjoint representation, is a flat lift of $\mathrm {Conj}_G$ over $W(k)$ , whose geometric generic fibre is the constant scheme associated to the set of conjugacy classes of $G_{\mathrm {abs}}$ . (Flatness follows from the fact that the special and the geometric generic fibre have the same length by Remark B.8.)

The length of $\mathrm {Conj}_G$ is equal to the number of conjugacy classes of $G_{\mathrm {abs}}$ . The length of $\mathrm {Conj}_G$ is at least the number of k-rational points $\mathrm {Conj}_G(k)$ and in general not equal, since $\mathrm {Conj}_G$ may not be étale over k.

If G is of length prime to p, then it is étale and linearly reductive and in this case, all previous approaches yield essentially the same notion of conjugacy class. On the other hand, the examples of the above subsections show that if G is étale of length divisible by p or if G is linearly reductive but not étale, then the various approaches of the above subsections usually lead to different notions of conjugacy classes.

B.5 Fifth approach: via adjoint representation

Let $G_{\mathrm {abs}}$ be a finite group and let $\rho _{\mathrm {ad}}:G_{\mathrm {abs}}\to {\mathbf {GL}}(V_{\mathrm {ad}})$ be its adjoint representation over k. By Remark A.4, the dimension of the largest trivial subrepresentation of $\rho _{\mathrm {ad}}$ is equal to the dimension of largest trivial quotient representation of $\rho _{\mathrm {ad}}$ (even if p divides the order of $G_{\mathrm {abs}}$ ) and these dimensions are equal to the number of conjugacy classes of $G_{\mathrm {abs}}$ . Unfortunately, these trivial sub- or quotient representations do not admit canonical decompositions into one-dimensional subspaces. Thus, there is no canonical bijection between the conjugacy classes of $G_{\mathrm {abs}}$ with trivial sub- or quotient representations of $\rho _{\mathrm {ad}}$ .

Remark B.15. Let $H:=k[G_{\mathrm {abs}}]$ be the group algebra with its usual Hopf algebra structure. The set of group-like elements of H recovers the group $G_{\mathrm {abs}}$ , see Example A.1. This suggests to use the adjoint representation ${}^{\mathrm {ad}}H$ together with the Hopf algebra structure of H - in particular, its co-algebra structure - to define a useful notion of conjugacy class.

As we have seen (somewhat implicitly) in Section B.3 above, this works: Let $G:=\mathrm {Spec}\: H^*$ be the constant group scheme associated to $G_{\mathrm {abs}}$ . Then, the topological space underlying $\mathrm {Conj}_G$ is the set of group-like elements of the Hopf algebra H modulo the adjoint representation. This set can be identified with the set of conjugacy classes of $G_{\mathrm {abs}}$ .

Unfortunately, the previous remark does not carry over to finite group schemes that are not étale, as the following example shows.

Example B.16. Let $G=\boldsymbol {\mu }_p$ or $G=\boldsymbol {\alpha }_p$ over the algebraically closed field k of characteristic $p>0$ .

  1. 1. The adjoint representation $\rho _{\mathrm {ad}}$ of G is trivial of dimension p, but there is no canonical way to decompose it into one-dimensional subspaces. This would suggest to have p conjugacy classes, but without being able to distinguish them.

  2. 2. Moreover, there is only one group-like element in the Hopf algebra $H^0(G,{\mathcal O}_G)^*$ and the adjoint representation on this element is trivial as well. This would suggest to have one conjugacy class only, see also Example B.9.(2).

B.6 Sixth approach: via extended adjoint representation

Given a finite-dimensional Hopf algebra A over a field k, we have the adjoint representation ${}^{\mathrm {ad}}A$ , see Appendix A.3. In Appendix A.4, we recalled the quantum double $D(A)=(A^{\mathrm {op}})^*\bowtie A$ . In Appendix A.5, we recalled that ${}^{\mathrm {ad}}A$ can be extended to a representation of $D(A)$ on A, the extended adjoint representation ${}^{\mathrm {Ad}}A$ .

Generalising work of Witherspoon [Reference WitherspoonWi99], Cohen and Westreich [Reference Cohen and WestreichCW11] defined the set of conjugacy classes of A for a finite-dimensional semi-simple Hopf algebra over ${\mathbb {C}}$ to be the set of simple subrepresentations of ${}^{\mathrm {Ad}}A$ . In this context, we also refer to the work of Jacoby [Reference JacobyJa17] and Zhu [Reference ZhuZhu97].

If G is a finite group scheme over k with Hopf algebra $A:=H^0(G,{\mathcal O}_G)$ , then $D(A)$ is semi-simple if and only if $D(A^*)$ is semi-simple if and only if G is of length prime to p, see Proposition A.7. In particular, the (extended) adjoint representations of A or $A^*$ may not be semi-simple. We will now study these representations and their relation to conjugacy classes.

Example B.17. Let G be étale over k. Then it is the constant group scheme associated to $G_{\mathrm {abs}}:=G(k)$ . If $A:=H^0(G,{\mathcal O}_G)$ is the associated Hopf algebra, then we have $A^*\cong k[G_{\mathrm {abs}}]$ . As seen in Example A.8, the extended adjoint representation ${}^{\mathrm {Ad}}(A^*)$ is semi-simple and there is a natural bijection between conjugacy classes of $G_{\mathrm {abs}}$ and simple subrepresentations of ${}^{\mathrm {Ad}}(A^*)$ . We stress that this is also true if p divides the order of $G_{\mathrm {abs}}$ .

Example B.18. Let G be a finite and linearly reductive group scheme over k, let $G_{\mathrm {abs}}$ be the associated abstract finite group and let $A:=H^0(G,{\mathcal O}_G)$ be the associated Hopf algebra. Since A is commutative, ${}^{\mathrm {ad}}A$ is trivial. Using the description (A.1) of ${}^{\mathrm {Ad}}A$ as an induced representation, we can identify it with the dual of the regular representation

$$ \begin{align*}\rho_{\mathrm{reg}}^\vee\,:\, A^*\,\to\, \mathrm{End}(A) \end{align*} $$

together with the trivial representation of A. By Proposition 2.6, representations of $A^*$ can be identified with representations of ${\mathbb {C}}[G_{\mathrm {abs}}]$ and thus, representations of $G_{\mathrm {abs}}$ . Thus, we can decompose ${}^{\mathrm {Ad}}A$ like the dual of the regular representation of $G_{\mathrm {abs}}$ : the number of isotypical components of this representation is equal to the number of conjugacy classes of $G_{\mathrm {abs}}$ . Thus, one can think of the conjugacy classes of $G_{\mathrm {abs}}$ as being “dual” to these isotypical components similarly to Appendix B.2.

The upshot of this discussion is the following: Let G be a finite group scheme over k and let $A:=H^0(G,{\mathcal O}_G)$ be the associated Hopf algebra.

  1. 1. If G is étale over k, then simple subrepresentations of ${}^{\mathrm {Ad}}(A^*)$ are in bijection with conjugacy classes of $G_{\mathrm {abs}}:=G(k)$ .

  2. 2. If G is linearly reductive, then the isotypical components of ${}^{\mathrm {Ad}}A$ give a reasonably good definition for the “dual” of a conjugacy class.

The following example shows that in general, neither ${}^{\mathrm {Ad}}A$ nor ${}^{\mathrm {Ad}}(A^*)$ leads to a good approach toward the notion of a conjugacy class - at least none that is better than the ones already discussed.

Example B.19. Let G be a finite and commutative group scheme over k and let $A:=H^0(G,{\mathcal O}_G)$ be the associated Hopf algebra. As a consequence of Example A.9, we have the following.

  1. 1. If $G=\boldsymbol {\mu }_p$ , then we have $A=k[{\mathbf {C}}_p]$ . Thus, ${}^{\mathrm {Ad}}A$ splits into the direct sum of p pairwise non-isomorphic one-dimensional representations, which correspond to the characters of G. On the other hand, ${}^{\mathrm {Ad}}(A^*)$ is a non-trivial successive p-fold and non-split extension of the trivial representation , whose semi-simplification is trivial of dimension p.

  2. 2. If $G={\mathbf {C}}_p$ , then we obtain the same as before with the rôles of A and $A^*$ interchanged.

  3. 3. If $G=\boldsymbol {\alpha }_p$ , then $A\cong A^*$ and ${}^{\mathrm {Ad}}A\cong {}^{\mathrm {Ad}}(A^*)$ is a successive p-fold and non-split extension of the trivial representation .

Acknowledgements

I thank Frank Himstedt, Martin Lorenz, Yuya Matsumoto, Frans Oort, Matt Satriano, Takehiko Yasuda and the referees for discussions and comments.

References

Abramovich, D., Olsson, M. and Vistoli, A., ‘Tame stacks in positive characteristic’, Ann. Inst. Fourier (Grenoble) 58(4) (2008), 10571091.CrossRefGoogle Scholar
Artin, M., ‘On isolated rational singularities of surfaces’, Amer. J. Math. 88 (1966), 129136.CrossRefGoogle Scholar
Artin, M., ‘Algebraic construction of Brieskorn’s resolutions’, J. Algebra 29 (1974), 330348.CrossRefGoogle Scholar
Artin, M., ‘Coverings of the rational double points in characteristic p’, in Complex Analysis and Algebraic Geometry (Iwanami-Shoten, Tokyo, 1977), 1122.CrossRefGoogle Scholar
Artin, M. and Verdier, J.-L., ‘Reflexive modules over rational double points’, Math. Ann. 270(1) (1985), 7982.CrossRefGoogle Scholar
Aspinwall, P., Bridgeland, T., Craw, A., Douglas, M., Gross, M., Kapustin, A., Moore, G., Segal, G., Szendrői, B. and Wilson, P., Dirichlet Branes and Mirror Symmetry (Clay Mathematics Monographs 4) (American Mathematical Society, Providence, Clay Mathematics Institute, Cambridge, MA, 2009).Google Scholar
Auslander, M. and Reiten, I., ‘McKay quivers and extended Dynkin diagrams’, Trans. Amer. Math. Soc. 293(1) (1986), 293301.CrossRefGoogle Scholar
Blume, M., ‘Construction of G-Hilbert schemes’, Math. Nachr. 284(8–9) (2011), 953959.CrossRefGoogle Scholar
Bridgeland, T., King, A. and Reid, M., ‘The McKay correspondence as an equivalence of derived categories’, J. Amer. Math. Soc. 14(3) (2001), 535554.CrossRefGoogle Scholar
Brieskorn, E., ‘Rationale Singularitäten komplexer Flächen’, Invent. Math. 4 (1967/68), 336358.CrossRefGoogle Scholar
Burciu, S., ‘On some representations of the Drinfeld double’, J. Algebra 296(2) (2006), 480504.CrossRefGoogle Scholar
Chan, K., Kirkman, E., Walton, C. and Zhang, J. J., ‘Quantum binary polyhedral groups and their actions on quantum planes’, J. Reine Angew. Math. 719 (2016), 211252.CrossRefGoogle Scholar
Chin, W., ‘Crossed products of semisimple cocommutative Hopf algebras’, Proc. Amer. Math. Soc. 116(2) (1992), 321327.CrossRefGoogle Scholar
Cohen, M. and Westreich, S., ‘Conjugacy classes, class sums and character tables for Hopf algebras’, Comm. Algebra 39(12) (2011), 46184633.CrossRefGoogle Scholar
Coquereaux, R. and Zuber, J.-B., ‘Drinfeld doubles for finite subgroups of $\mathbf{SU}(2)$ and $\mathbf{SU}(3)$ Lie groups’, SIGMA Symmetry Integrability Geom. Methods Appl. 9 (2013), Paper 039, 36 pp.CrossRefGoogle Scholar
Dijkgraaf, R., Pasquier, V. and Roche, P., ‘Quasi Hopf algebras, group cohomology and orbifold models’, Nuclear Phys. B Proc. Suppl. 18B (1991), 6072.CrossRefGoogle Scholar
Drinfeld, V. G., Quantum Groups (Proceedings of the International Congress of Mathematicians) vol. 1, 2 (Berkeley, Calif., 1986) (Amer. Math. Soc., 1987), 798820.Google Scholar
Durfee, A. H., ‘Fifteen characterizations of rational double points and simple critical points’, Enseign. Math. 25 (1979), 131163.Google Scholar
Etingof, P. and Gelaki, S., ‘On finite-dimensional semisimple and cosemisimple Hopf algebras in positive characteristic’, Int. Math. Res. Not. 1998(16) (1998), 851864.Google Scholar
Fogarty, J., ‘Algebraic families on an algebraic surface’, Amer. J. Math. 90 (1968), 511521.CrossRefGoogle Scholar
Gonzalez-Sprinberg, G. and Verdier, J-L., ‘Construction géométrique de la correspondence de McKay’, Ann. Sc. ENS 16 (1983), 409449.Google Scholar
Gould, M. D., ‘Quantum double finite group algebras and their representations’, Bull. Austral. Math. Soc. 48(2) (1993), 275301.CrossRefGoogle Scholar
Hashimoto, M., ‘Classification of the linearly reductive finite subgroup schemes of SL2’, Acta Math. Vietnam. 40 (2015), 527534.CrossRefGoogle Scholar
Ishii, A., ‘On the McKay correspondence for a finite small subgroup of $\mathbf{GL}\left(2.\mathrm{C}\right)$ ’, J. Reine Angew. Math. 549 (2002), 221233.Google Scholar
Ishii, A. and Nakamura, I., ‘Extended McKay correspondence for quotient surface singularities’, Q. J. Math. 70(2) (2019), 395408.CrossRefGoogle Scholar
Ishii, A. and Ueda, K., ‘The special McKay correspondence and exceptional collections’, Tohoku Math. J. (2) 67(4) (2015), 585609.CrossRefGoogle Scholar
Ito, Y. and Nakamura, I., ‘Hilbert schemes and simple singularities’, in New Trends in Algebraic Geometry (Warwick, 1996) (London Math. Soc. Lecture Note Ser. 264) (Cambridge Univ. Press, Cambridge, 1999), 151233.CrossRefGoogle Scholar
Ito, Y. and Reid, M., ‘The McKay correspondence for finite subgroups of $\mathbf{SL}\left(3.\mathrm{C}\right)$ ’, in Higher-Dimensional Complex Varieties (Trento, 1994) (de Gruyter, Berlin, 1996), 221240.Google Scholar
Jacoby, A., ‘On representation theory of finite-dimensional Hopf algebras’, PhD Thesis, Temple University, 2017.Google Scholar
Kapranov, M. and Vasserot, E., ‘Kleinian singularities, derived categories and Hall algebras’, Math. Ann. 316(3) (2000), 565576.CrossRefGoogle Scholar
Kirillov, A., Quiver Representations and Quiver Varieties (Graduate Studies in Mathematics) vol. 174 (American Mathematical Society, Providence, RI, 2016).CrossRefGoogle Scholar
Klein, F., Vorlesungen über das Ikosaeder und die Auflösung der Gleichungen vom fünften Grade (B. G. Teubner, Leipzig, 1884).Google Scholar
Knörrer, H., ‘Group representations and the resolution of rational double points’, in Finite Groups – Coming of Age (Montreal, Que., 1982) (Contemp. Math.) vol. 45 (Amer. Math. Soc., 1985), 175222.CrossRefGoogle Scholar
Kostant, B., ‘On finite subgroups of $\mathbf{SU}(2)$ , simple Lie algebras and the McKay correspondence’, Astérisque S131 (1985).CrossRefGoogle Scholar
Kovacs, S., ‘Young person’s guide to moduli of higher dimensional varieties’, in Algebraic Geometry - Seattle 2005 (Proc. Symp. Pure Math.) Part 2 (AMS, 2009), 711743.fyCrossRefGoogle Scholar
Larson, R. G. and Radford, D. E., ‘Finite dimensional cosemisimple Hopf algebras in characteristic 0 are semisimple’, J. Algebra 117, 267289 (1988).CrossRefGoogle Scholar
Laufer, H. B., ‘Taut two-dimensional singularities’, Math. Ann. 205 (1973), 131164.CrossRefGoogle Scholar
Liedtke, C. and Satriano, M., ‘On the birational nature of lifting’, Adv. Math. 254 (2014), 118137.CrossRefGoogle Scholar
Liedtke, C., Martin, G. and Matsumoto, Y., ‘Linearly reductive quotient singularities’, Preprint, 2021, arXiv:2102.01067, to appear in Astérisque.Google Scholar
Lorenz, M., A Tour of Representation Theory (Graduate Studies in Mathematics) vol. 193 (American Mathematical Society, 2018).CrossRefGoogle Scholar
McKay, J., ‘Graphs, singularities, and finite groups’, in The Santa Cruz Conference on Finite Groups (Univ. California, Santa Cruz, Calif., 1979) (Proc. Sympos. Pure Math.) vol. 37 (Amer. Math. Soc., Providence, RI, 1980), 183186.CrossRefGoogle Scholar
McKay, J., ‘Cartan matrices, finite groups of quaternions, and kleinian singularities’, Proc. Amer. Math. Soc. (1981), 153154.Google Scholar
Montgomery, S., Hopf Algebras and Their Actions on Rings (CBMS Regional Conference Series in Mathematics) vol. 82 (American Mathematical Society, 1993).CrossRefGoogle Scholar
Mumford, D. and Oort, F., ‘Deformations and liftings of finite, commutative group schemes’, Invent. Math. 5 (1968), 317334.Google Scholar
Nagao, H. and Tsushima, Y., Representations of Finite Groups (Academic Press, 1989). Translated from Japanese.Google Scholar
Nagata, M., ‘Complete reducibility of rational representations of a matric group’, J. Math. Kyoto Univ. 1 (1961/1962), 8799.Google Scholar
Nakajima, H., ‘Varieties associated with quivers’, in Representation Theory of Algebras and Related Topics (Mexico City, 1994) (CMS Conf. Proc.) vol. 19 (Amer. Math. Soc., 1996), 139157.Google Scholar
Nakajima, H., Quiver Varieties and McKay Correspondence - Lectures at Hokkaido University, December 2001, available from the author’s webpage at http://www.kurims.kyoto-u.ac.jp/nakajima/TeX/hookaido.pdf Google Scholar
Oort, F., ‘Finite group schemes, local moduli for abelian varieties, and lifting problems’, Compos. Math. 23 (1971), 265296.Google Scholar
Reid, M., ‘Young person’s guide to canonical singularities’, in Algebraic Geometry , Bowdoin 1985 (Proc. Sympos. Pure Math.) vol. 46, part 1 (Amer. Math. Soc. 1987), 345414.Google Scholar
Reid, M., ‘La correspondance de McKay’, Séminaire Bourbaki, vol. 1999/2000, Astérisque 276 (2002), 5372.Google Scholar
Serre, J.-P., Linear Representations of Finite Groups (Graduate Texts in Mathematics) vol. 42 (Springer, 1977).CrossRefGoogle Scholar
Steinberg, R., ‘Finite subgroups of ${\mathbf{SU}}_2$ , Dynkin diagrams and affine Coxeter elements’, Pacific J. Math. 118(2) (1985), 587598.CrossRefGoogle Scholar
Satriano, M., ‘The Chevalley-Shephard-Todd theorem for finite linearly reductive group schemes’, Algebra Number Theory 6(1) (2012), 126.CrossRefGoogle Scholar
Demazure, M. et al., Séminaire sur les Singularités des Surfaces (Lecture Notes in Mathematics) vol. 777 (Springer, 1980).CrossRefGoogle Scholar
Tanaka, Y., ‘On tautness of two-dimensional F-regular and F-pure rational singularities’, Preprint, 2015, arXiv:1502.07236.Google Scholar
Tonini, F. and Yasuda, T., ‘Notes on the Motivic McKay Correspondence for the Group Scheme ${\alpha}_p$ ’, in Singularities - Kagoshima 2017 (World Scientific, 2020), 215233.CrossRefGoogle Scholar
Waterhouse, W. C., Introduction to Affine Group Schemes (Graduate Texts in Mathematics) vol. 66 (Springer, 1979).CrossRefGoogle Scholar
Witherspoon, S. J., ‘The representation ring of the quantum double of a finite group’, J. Algebra 179(1) (1996), 305329.CrossRefGoogle Scholar
Witherspoon, S. J., ‘The representation ring and the centre of a Hopf algebra’, Canad. J. Math. 51(4) (1999), 881896.CrossRefGoogle Scholar
Wunram, J., ‘Reflexive modules on quotient surface singularities’, Math. Ann. 279(4) (1988), 583598.CrossRefGoogle Scholar
Yasuda, T., ‘The wild McKay correspondence via motivic integration’, Sugaku Expositions 35(1) (2022), 128.CrossRefGoogle Scholar
Yasuda, T., ‘Open problems in the wild McKay correspondence and related fields’, (Adv. Stud. Pure Math.) vol. 88 (Mathematical Society of Japan, 2023), 279303.Google Scholar
Zhu, Y., ‘A commuting pair in Hopf algebras’, Proc. Amer. Math. Soc. 125 (1997), 28472851.CrossRefGoogle Scholar