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

A tensor-triangular property for categories of representations of restricted Lie algebras

Justin Bloom
(December 2024)
Abstract

We define a property for restricted Lie algebras in terms of cohomological support and tensor-triangular geometry of their categories of representations. By Tannakian reconstruction, the different symmetric tensor category structures on the underlying linear category of representations of a restricted Lie algebra correspond to different cocommutative Hopf algebra structures on the restricted enveloping algebra. In turn this equates together the linear categories of representations for various group scheme structures. The tensor triangular spectrum, for representations of a restricted Lie algebra, is known to be isomorphic to the scheme of 1-parameter subgroups of the infinitesimal group scheme structure associated to the Lie algebra. Points in the spectrum of a tensor triangulated category correspond to minimal radical thick tensor-ideals, provided the spectrum is noetherian, as is known in our case of finite group schemes. When the group scheme structure changes from the Lie algebra structure, a set of subgroups can still yield points of the spectrum, but there may not be enough to cover the spectrum. Restricted Lie algebras satisfy our property if, for each group scheme structure, the remaining set of subgroups correspond to minimal radical thick tensor-ideals having identical Green-ring structure to that of the original Lie algebra. Some small examples of algebras of finite and tame representation type satisfying our property are given. We show that no abelian restricted Lie algebra of wild representation type may have our property. We conjecture that satisfying our property is equivalent to having finite or tame representation type.

1 Introduction

1.1 Overview

In this paper we define Property PC for Jacobson’s restricted Lie algebras [Jac41] in terms of cohomological support for their category of restricted representations. The purpose of finding Lie algebras with this property is to illustrate how tensor-triangular geometry can be used to find constraints imposed on Green ring calculations for families of symmetric tensor categories.

We adopt the convention throughout that all representations of a restricted Lie algebra are restricted, and that kk is always an algebraically closed field of characteristic p>0p>0. We also adopt the convention that all finite group schemes over kk are affine. For a finite group scheme GG over kk, we denote by kGkG the group algebra, also called the measure algebra, a cocommutative Hopf algebra which is dual to the coordinate algebra k[G]=𝒪(G)k[G]={\mathcal{O}}(G) of the affine scheme G.G. We denote AA^{*} the kk-linear dual of a kk-space AA. We see any finite dimensional cocommutative Hopf algebra AA is kGkG for some finite group scheme G=SpecAG=\operatorname{Spec}A^{*}.

We will review the notions of cohomological support for finite group schemes in Section 2. The work of Benson, Carlson, and Rickard [BCR96] on finite groups, and later the works of Friedlander and Pevtsova [FrPev07], [FrPev05], and of Benson, Iyengar, Krause, and Pevtsova [BIKP15] on finite group schemes, establishes an equivalence between the data of cohomological support and tensor-triangular support (Balmer [Balmer05]) for the stable module category. Thus, for a finite group scheme GG, a closed point 𝔭ProjH(G,k){\mathfrak{p}}\in\operatorname{Proj}H^{*}(G,k) gives a subcategory 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) of finite modules supported only at the singleton {𝔭}\{{\mathfrak{p}}\}, and 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) is a minimal thick \otimes-ideal. What is interesting for our purposes is that when two finite group schemes both have a group algebra isomorphic to a given associative algebra AA, there is an equivalence of kk-linear categories of representations, but not of tensor categories. Yet the subcategories 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) (or any subcategory with support in a fixed subset of ProjH(A,k)\operatorname{Proj}H^{*}(A,k)) remain \otimes-ideal independent of which \otimes is chosen, so long as the one dimensional module kk comes from a fixed augmentation AkA\to k. This may be seen by noticing how the ring structure on H(A,k)=ExtA(k,k)H^{*}(A,k)=\operatorname{Ext}^{*}_{A}(k,k) is not dependent on a choice of Hopf algebra structure, and similarly the module action on cohomology H(A,M)=ExtA(k,M)H^{*}(A,M)=\operatorname{Ext}^{*}_{A}(k,M) for any AA-module MM. We investigate for which group schemes and which points 𝔭{\mathfrak{p}} the Green ring structure on 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) may be known to not change between group schemes.

When 𝔤{\mathfrak{g}} is a restricted Lie algebra of dimension rr, recall that restricted representations of 𝔤{\mathfrak{g}} coincide with modules over the restricted enveloping algebra A=u(𝔤)A=u({\mathfrak{g}}), which has dimension pr.p^{r}. The algebra AA is canonically a cocommutative Hopf algebra by taking elements of 𝔤{\mathfrak{g}} to be primitive. Thus there is a canonical group scheme G~\widetilde{G} with group algebra A,A, which we call the infinitesimal group scheme corresponding to 𝔤{\mathfrak{g}}. Given any cocommutative Hopf algebra structure Δ:AAA,\Delta:A\to A\otimes A, we have a corresponding group scheme GG with group algebra AA, and a corresponding tensor product \otimes for modules over AA, which are now also representations of G.G. We will always denote by Δ~,~\widetilde{\Delta},\widetilde{\otimes} the canonical Hopf algebra comultiplication and tensor product for the Lie algebra 𝔤{\mathfrak{g}}, or equivalently the infinitesimal group scheme G~.\widetilde{G}. We will only consider Hopf algebra structures sharing a counit Ak,A\to k, so that kk is a fixed AA-module acting as the monoidal unit with respect to any .\otimes.

Definition 1.1.1.

Let 𝔤{\mathfrak{g}} be a restricted Lie algebra over kk, and A=u(𝔤)A=u({\mathfrak{g}}). For finite group schemes GG coming from a Hopf algebra structure on AA, let \otimes denote the tensor product of GG-representations.

The algebra 𝔤{\mathfrak{g}} is said to satisfy Property PC if, for any such finite group scheme GG, and any 𝔭ProjH(A,k),{\mathfrak{p}}\in\operatorname{Proj}H^{*}(A,k), the tensor products ,~\otimes,\widetilde{\otimes} induce identical Green ring structures on the ideal 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) iff the prime 𝔭{\mathfrak{p}} is noble (3.1.1) for GG.

We include, as Proposition 2.2.1, a consequence of the work of Bendel, Friedlander, Parshall, and Suslin [FrPar87], [FrPar86], [SFB97], which tells us that all homogeneous primes of ProjH(𝔤,k)\operatorname{Proj}H^{*}({\mathfrak{g}},k) arise from 11-dimensional Lie subalgebras of 𝔤{\mathfrak{g}}. The idea of homogeneous primes arising from subgroups of a group scheme is generalized to our notion of a noble prime (3.1.1). In this way, Proposition 2.2.1 may be restated to say that every prime is noble for the infinitesimal group scheme corresponding to the restricted Lie algebra.

In context of Property PC, we see that deforming the comultiplication structure on the restricted enveloping algebra may change the group scheme in such a way that a given homogeneous prime of cohomology is no longer noble. It is at these ignoble primes where we expect the Green ring structure on the subcategory 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) to always change from its original structure from the Lie algebra. This we call Property PB, which in conjunction with its converse Property PA, becomes Property PC (see 3.1.4).

Theorem 1.1.2 below is proven in Section 3.3 and shows there exists an algebra of tame representation type satisfying Property PC.

Theorem 1.1.2.

Let 𝔤{\mathfrak{g}} be the 2-dimensional abelian Lie algebra with trivial restriction, i.e 𝔤[p]=0{\mathfrak{g}}^{[p]}=0, over kk with characteristic p=2p=2. Then 𝔤{\mathfrak{g}} satisfies Property PC.

Theorem 1.1.3 below is proven in Section 3.2 and provides an algebra of wild representation type in each odd characteristic, all failing to satisfy Property PC.

Theorem 1.1.3.

Let 𝔤{\mathfrak{g}} be the 2-dimensional abelian Lie algebra over kk with trivial restriction, i.e. 𝔤[p]=0{\mathfrak{g}}^{[p]}=0. If p>2p>2 then 𝔤{\mathfrak{g}} does not satisfy Property PA. In other words, there exists a group scheme GG as in Definition 1.1.1 and a point 𝔭{\mathfrak{p}} which is noble for GG, and modules V,W𝒞(𝔭)V,W\in{\mathcal{C}}({\mathfrak{p}}) such that VWV\otimes W is not isomorphic to V~W.V\widetilde{\otimes}W. So 𝔤{\mathfrak{g}} does not satisfy Property PC.

The restricted Lie algebra 𝔤{\mathfrak{g}} of 1.1.2 and 1.1.3 corresponds to the infinitesimal group scheme 𝔾a(1)2.{\mathbb{G}}_{a(1)}^{2}. The enveloping algebra u(𝔤)u({\mathfrak{g}}) is isomorphic as an associative algebra also to the group algebra k(/p)2k({\mathbb{Z}}/p)^{2} and to the group algebra for two other distinct group schemes over kk up to isomorphism. The classification of Hopf algebra structures on u(𝔤)=k[x,y]/(xp,yp)u({\mathfrak{g}})=k[x,y]/(x^{p},y^{p}) was first given as a corollary of the classification of all connected Hopf algebras of dimension p2p^{2}, by X. Wang [XWang13].

We state Conjecture 3.4.3, that having wild representation type is equivalent to failing to satisfy Property PC, and provide more examples of restriced Lie algebras having wild representation type, while not satisfying Property PC, in Section 4. We have found no restricted Lie algebra failing to satisfy Property PB, leading us to Conjecture 3.4.4.

It is fundamental to our use of noble points that the following algebra, of finite representation type, also satisfies Property PC. This is discussed further in 3.1.10.

Example 1.1.4.

Let 𝔤{\mathfrak{g}} be the one dimensional Lie algebra over kk with restriction 𝔤[p]=0.{\mathfrak{g}}^{[p]}=0. Then 𝔤{\mathfrak{g}} satisfies Property PC. In fact, there is only one homogeneous prime in ProjH(𝔤,k)={𝔭}\operatorname{Proj}H^{*}({\mathfrak{g}},k)=\{{\mathfrak{p}}\} and we find that all group schemes GG with kGu(𝔤)kG\cong u({\mathfrak{g}}) have that 𝔭{\mathfrak{p}} is noble for GG, and that they all define identical Green rings.

In order to affirm that a given restricted Lie algebra 𝔤{\mathfrak{g}} of tame or finite representation type satisfies Property PC, we resort to using an explicit classification of all cocommutative Hopf algebra structures on the universal enveloping algebra u(𝔤).u({\mathfrak{g}}). The work of Nguyen, Ng, L. Wang, and X. Wang, [NgW23], [NWW15], [NWW16], [WW14], [XWang13], [XWang15] has classified connected and pointed Hopf algebras of small dimension. As a corollary there are many small dimensional Lie algebras 𝔤{\mathfrak{g}} for which u(𝔤)u({\mathfrak{g}}) is a local ring (such 𝔤{\mathfrak{g}} are called unipotent), with all Hopf algebra structures on u(𝔤)u({\mathfrak{g}}) known to be dual to one of the connected Hopf algebras of the above mentioned authors. All local Hopf algebras of order p2p^{2} in characteristic pp are given explicitly as a corollary in [XWang13] and we use this directly in Section 3.3 to affirm Property PC for a tame algebra over a field of characteristic p=2p=2.

Our Property PC pertains to cocommutative Hopf algebra structures on a given finite augmented algebra, which we may think of as points of an affine algebraic set ,{\mathscr{H}}, whereas the classifications of the above mentioned authors yield only a complete set of representatives of the orbit space /{\mathscr{H}}/\sim under the action of twisting by augmented algebra automorphisms (3.1.5). Our technique for proving Theorem 1.1.2 consists of first showing the expected behavior for modules supported at noble and ignoble points for a complete set of representatives of /{\mathscr{H}}/\sim, as first classified by X. Wang in [XWang13]. In doing so it is shown that both the tame algebra of Theorem 1.1.2 and the wild algebra of Theorem 1.1.3 satisfy a weaker version of Property PB, quantified only over a complete set of representatives of /.{\mathscr{H}}/\sim. This is extended in the tame case p=2p=2 by using that if MM is any finite representation supported at a point 𝔭{\mathfrak{p}}, we find MφMM^{\varphi}\cong M whenever φAut(u(𝔤))\varphi\in\operatorname{Aut}(u({\mathfrak{g}})) is an algebra automorphism fixing the point 𝔭ProjH(𝔤,k).{\mathfrak{p}}\in\operatorname{Proj}H^{*}({\mathfrak{g}},k).

Our paper concludes with Section 4, an exposé on augmented algebra automorphisms, and on induced modules, for the restricted enveloping algebra of some Lie algebras having wild representation type. The group of (augmented) automorphisms Aut(u(𝔤))\operatorname{Aut}(u({\mathfrak{g}})) acts on {\mathscr{H}}, on the isomorphism classes π0(𝗋𝖾𝗉𝔤),\pi_{0}({\sf rep\ }{\mathfrak{g}}), and on the projective scheme ProjH(𝔤,k).\operatorname{Proj}H^{*}({\mathfrak{g}},k). Our technique, for proving that a Lie algebra 𝔤{\mathfrak{g}} fails to satisfy Property PC, is to provide nontrivial elements of the quotient

Aut(u(𝔤))𝔭/Aut(u(𝔤))π0(𝒞(𝔭))\operatorname{Aut}(u({\mathfrak{g}}))^{\mathfrak{p}}/\operatorname{Aut}(u({\mathfrak{g}}))^{\pi_{0}({\mathcal{C}}({\mathfrak{p}}))}

of isotropy subgroups Aut(u(𝔤))𝔭,Aut(u(𝔤))π0(𝒞(𝔭))\operatorname{Aut}(u({\mathfrak{g}}))^{{\mathfrak{p}}},\operatorname{Aut}(u({\mathfrak{g}}))^{\pi_{0}({\mathcal{C}}({\mathfrak{p}}))} of Aut(u(𝔤)),\operatorname{Aut}(u({\mathfrak{g}})), for a choice of point 𝔭ProjH(𝔤,k).{\mathfrak{p}}\in\operatorname{Proj}H^{*}({\mathfrak{g}},k). Given that an automorphism fixes each isomorphism class in π0(𝒞(𝔭)),\pi_{0}({\mathcal{C}}({\mathfrak{p}})), it must also fix the point 𝔭{\mathfrak{p}}. It is conjectured that the converse only fails for algebras 𝔤{\mathfrak{g}} of wild representation type; this is in turn informed by the conjecture that the continuous parameter for indecomposables of any tame Lie algebra is always realizable as support.

For the right choice of isotropy φAut(u(𝔤))𝔭\varphi\in\operatorname{Aut}(u({\mathfrak{g}}))^{\mathfrak{p}}, we find that twisting the Lie Hopf algebra structure gives a tensor product ~φ\widetilde{\otimes}^{\varphi} such that V~φWV~W,V\widetilde{\otimes}^{\varphi}W\not\cong V\widetilde{\otimes}W, where V,WV,W are a choice of modules having support {𝔭},\{{\mathfrak{p}}\}, induced from a subalgebra of 𝔤{\mathfrak{g}}. We show how to produce such an isotropy for any restricted Lie algebra with a wild abelian Lie algebra as a direct summand. We also produce such an isotropy in odd characteristic for the Heisenberg Lie algebra, which contains a wild elementary abelian Lie subalgebra, but not as a direct summand.

1.2 Why Lie algebras?

For our study of certain families of tensor-categories, we will review some tools from the tensor-triangular geometry of the stable module category of a group scheme in section 3.1. For one, we establish that subcategories supported at a subset of homogeneous primes of cohomology are closed under tensor product no matter which tensor product is chosen! Further, two modules will have tensor product a projective module if and only if they have disjoint support, and again their support and hence this property is independent of which tensor product is chosen. So by our account, comparing Green rings structures on the same underlying abelian category leads directly to tensor-triangular geometry. Many computations of the product in a tame category follow from abstract impositions from the theory of support before a product is even chosen.

Now we direct the reader to Proposition 2.2.1. This proposition states how every prime is noble for G~\widetilde{G}, whenever G~\widetilde{G} is the infinitesimal group scheme corresponding to a restricted Lie algebra 𝔤{\mathfrak{g}}. This is a necessary condition for the given restricted Lie algebra 𝔤{\mathfrak{g}} to have our Property PC. So why do we want to study deformations of finite group schemes GG such that every prime is noble for GG? The answer is a kind of local-to-global problem for tensor product structures on modules supported at only one point.

In place of localization we consider the pullback of a module along a π\pi-point for GG, as defined by Friedlander and Pevtsova [FrPev07], [FrPev05], which is a flat map

α:k[t]/tpkG\alpha:k[t]/t^{p}\to kG

that factors through a unipotent subgroup scheme of G.G. When the pullback of a module MM to k[t]/tpk[t]/t^{p} modules along α\alpha is not projective, we say MM is supported by α\alpha. It is then shown that the cohomological support for a module MM over a finite cocommutative Hopf algebra AA is equivalent (see Sections 2.2, 2.3) to the locus of π\pi-points α:k[t]/tpA\alpha:k[t]/t^{p}\to A such that the pullback is not projective. Our methods for proving Property PC are partly an investigation into whether isomorphisms

(MN)α(MN)α,(M\otimes N)\downarrow_{\alpha}\cong(M\otimes^{\prime}N)\downarrow_{\alpha},

known to hold ‘locally’ for each π\pi-point α\alpha, are enough to conclude the ‘global’ isomorphism MNMNM\otimes N\cong M\otimes^{\prime}N for two different tensor products ,\otimes,\otimes^{\prime} of modules over the algebra A.A. Further, the definition of noble (3.1.1) gives us a representing π\pi-point α:k[t]/tpA\alpha:k[t]/t^{p}\to A such that the restriction property holds

(MN)αMαNα,(M\otimes N)\downarrow_{\alpha}\cong M\downarrow_{\alpha}\otimes N\downarrow_{\alpha},

since \otimes may be defined on k[t]/tpk[t]/t^{p}-modules according to a Hopf-subalgebra of AA. Suppose M,NM,N are AA-modules supported only at the π\pi-point α\alpha, which is noble for both finite group schemes G,GG,G^{\prime} having an isomorphism kGkGkG\cong kG^{\prime} as associative algebras. Then assuming further that kGkG is a local ring we have an automatic local-isomorphism of products because for the point α\alpha in the support of both M,NM,N, we get

(MN)α\displaystyle(M\otimes N)\downarrow_{\alpha} =(Mα)(Nα)\displaystyle=(M\downarrow_{\alpha})\otimes(N\downarrow_{\alpha})
(Mα)(Nα)\displaystyle\cong(M\downarrow_{\alpha})\otimes(N\downarrow_{\alpha})
(Mα)(Nα)\displaystyle\cong(M\downarrow_{\alpha})\otimes^{\prime}(N\downarrow_{\alpha}) (1.2.1)
=(MM)α,\displaystyle=(M\otimes^{\prime}M)\downarrow_{\alpha},

and for the points not supported by M,NM,N the tensor product is projective on both sides, and hence free of the same rank. The third identity 1.2.1 is a direct application of our fundamental example 1.1.4. Without automatic local-isomorphisms as a starting point, it is significantly more difficult to address how the tensor categories compare between two arbitrary group schemes G,GG,G^{\prime} sharing a group algebra. A finite group scheme for which every prime, or π\pi-point, is noble, is as such a good starting point for this kind of investigation, so Proposition 2.2.1 offers an especially convenient start towards investigating Lie algebras.

2 Background

2.1 Tensor product of representations and reconstruction

If V,WV,W are two representations of a Lie algebra 𝔤{\mathfrak{g}} over a field kk, the tensor product VkWV\otimes_{k}W is given the structure of a 𝔤{\mathfrak{g}}-representation, we’ll call V~WV\widetilde{\otimes}W, according to the Leibniz rule on simple tensors

x(vw)=xvw+vxw,x(v\otimes w)=xv\otimes w+v\otimes xw,

for x𝔤,vV,wW.x\in{\mathfrak{g}},v\in V,w\in W. If kk is of characteristic p>0p>0 and 𝔤{\mathfrak{g}} is a restricted Lie algebra of dimension rr in the sense of Jacobson [Jac41], then the restricted universal enveloping algebra A=u(𝔤)A=u({\mathfrak{g}}) is defined, and is an associative algebra over kk of dimension prp^{r}, such that the restricted representations of 𝔤{\mathfrak{g}} are equivalent to modules over AA. In general, the products of representations, which endow the category of finite representations 𝗋𝖾𝗉𝔤{\sf rep\ }{\mathfrak{g}} with the structure of a tensor category, fibred via the usual forgetful functor

:𝗋𝖾𝗉𝔤𝖵𝖾𝖼k,{\mathcal{F}}:{\sf rep\ }{\mathfrak{g}}\to{\sf Vec}\ k,

are all induced from a Hopf algebra structure AAAA\to A\otimes A, a map of kk-algebras. The product ~\widetilde{\otimes} comes from the canonical Hopf algebra Δ~:AAA,\widetilde{\Delta}:A\to A\otimes A, which makes 𝔤{\mathfrak{g}} into the primitive subspace of P(A)AP(A)\subset A, i.e. each element xx of the generating set 𝔤A{\mathfrak{g}}\subset A is mapped to x1+1xAA.x\otimes 1+1\otimes x\in A\otimes A. The monoidal unit kk is given module structure by a counit map of kk-algebras AkA\to k.

We recall from Etingof, Gelaki, Nikshych, and Ostrik [EGNO15] some definitions and the reconstruction theorem 2.1.4 to justify how tensor category structures on 𝗋𝖾𝗉𝔤{\sf rep\ }{\mathfrak{g}} are induced from Hopf algebra structures on A.A.

Definition 2.1.1.

[EGNO15]

  1. 1.

    A kk-linear abelian category is locally finite if

    1. (i)

      𝒞{\mathcal{C}} has finite dimensional spaces of morphisms;

    2. (ii)

      every object of 𝒞{\mathcal{C}} has finite length,

    and 𝒞{\mathcal{C}} is finite if in addition

    1. (c)

      𝒞{\mathcal{C}} has enough projectives; and

    2. (d)

      there are finitely many isomorphism classes of simple objects.

  2. 2.

    An object in a monoidal category is called rigid if it has left and right duals. A monoidal category 𝒞{\mathcal{C}} is called rigid if every object of 𝒞{\mathcal{C}} is rigid.

  3. 3.

    Let 𝒞{\mathcal{C}} be a locally finite kk-linear abelian rigid monoidal category. The category 𝒞{\mathcal{C}} is a tensor category if the bifunctor :𝒞×𝒞𝒞\otimes:{\mathcal{C}}\times{\mathcal{C}}\to{\mathcal{C}} is bilinear on morphisms, and End𝒞(𝟏)k\operatorname{End}_{\mathcal{C}}({\bf 1})\cong k.

Definition 2.1.2.

[EGNO15] Let 𝒞,𝒟{\mathcal{C}},{\mathcal{D}} be two locally finite abelian categories over kk. Deligne’s tensor product 𝒞×𝒟{\mathcal{C}}\hbox{$\Box$}\kern-7.7778pt\hbox{$\times$}{\mathcal{D}} is an abelian kk-linear category which is universal for the functor assigning, to every kk-linear abelian category 𝒜{\mathcal{A}}, the category of right exact in both variables bilinear bifunctors 𝒞×𝒟𝒜.{\mathcal{C}}\times{\mathcal{D}}\to{\mathcal{A}}.

The tensor product exists, is locally finite, and is unique up to unique equivalence, see [EGNO15, Proposition 1.11.2], or Deligne [Del90].

2.1.3.

A finite kk-linear abelian category 𝒞{\mathcal{C}} is equivalent to the category of modules over a finite algebra, or rather a Morita equivalence class of algebras [EGNO15, Section 1.8]. Fixing an exact faithful functor :𝒞𝖵𝖾𝖼k{\mathcal{F}}:{\mathcal{C}}\to{\sf Vec}\ k to finite dimensional vector spaces allows a finite algebra to be constructed as End()\operatorname{End}({\mathcal{F}}). Given an algebra AA such that 𝒞=𝗆𝗈𝖽A,{\mathcal{C}}={\sf mod}\ A, the forgetful functor {\mathcal{F}} is representable by the free module AA. Hence, by the Yoneda lemma End()=(A)=A\operatorname{End}({\mathcal{F}})={\mathcal{F}}(A)=A as a vector space, and indeed as an algebra. In fact any such exact faithful {\mathcal{F}} has that 𝒞{\mathcal{C}} is equivalent to modules over A=End()A=\operatorname{End}({\mathcal{F}}), and when modeled as such :𝗆𝗈𝖽A𝖵𝖾𝖼k{\mathcal{F}}:{\sf mod}\ A\to{\sf Vec}\ k is isomorphic to the forgetful functor.

Let :𝒞𝖵𝖾𝖼k{\mathcal{F}}:{\mathcal{C}}\to{\sf Vec}\ k and 𝒢:𝒟𝖵𝖾𝖼k{\mathcal{G}}:{\mathcal{D}}\to{\sf Vec}\ k be exact faithful functors with finite kk-linear abelian sources 𝒞,𝒟,{\mathcal{C}},{\mathcal{D}}, equivalent respectively to modules over A=End()A=\operatorname{End}({\mathcal{F}}) and B=End(𝒢)B=\operatorname{End}({\mathcal{G}}). Then the exact functors 𝒞𝒟{\mathcal{C}}\to{\mathcal{D}} relative to 𝖵𝖾𝖼k{\sf Vec}\ k correspond to homomorphisms of algebras BA.B\to A. In other words, given a commutative diagram of exact functors

𝒞{{\mathcal{C}}}𝒟{{\mathcal{D}}}𝖵𝖾𝖼k{{{\sf Vec}\ k}}Φ\scriptstyle{\Phi}\scriptstyle{{\mathcal{F}}}𝒢\scriptstyle{{\mathcal{G}}}

the homomorphism of algebras Ψ:BA\Psi:B\to A, defined by precomposition with Φ\Phi, is such that Φ\Phi is the pullback of modules along Ψ.\Psi.

2.1.4.

Now we review the relevant version of Tannakian reconstruction of a Hopf algebra from a finite tensor category. Let AA be a finite associative algebra over kk, the category of finite modules is a finite kk-linear abelian category we’ll call 𝒞.{\mathcal{C}}. It is known that 𝒞×𝒞{\mathcal{C}}\hbox{$\Box$}\kern-7.7778pt\hbox{$\times$}{\mathcal{C}} is equivalent to the category of (AA)(A\otimes A)-modules. By the previous discussion we see the tensor product of finite kk-linear abelian categories is again finite.

Denote by

:𝒞𝖵𝖾𝖼k{\mathcal{F}}:{\mathcal{C}}\to{\sf Vec}\ k

the forgetful functor. Then the composition

(×):𝒞×𝒞𝖵𝖾𝖼k×𝖵𝖾𝖼k𝖵𝖾𝖼k\otimes\circ({\mathcal{F}}\hbox{$\Box$}\kern-7.7778pt\hbox{$\times$}{\mathcal{F}}):{\mathcal{C}}\hbox{$\Box$}\kern-7.7778pt\hbox{$\times$}{\mathcal{C}}\to{\sf Vec}\ k\hbox{$\Box$}\kern-7.7778pt\hbox{$\times$}{\sf Vec}\ k\to{\sf Vec}\ k

is equivalent, when 𝒞×𝒞{\mathcal{C}}\hbox{$\Box$}\kern-7.7778pt\hbox{$\times$}{\mathcal{C}} is modeled as 𝗆𝗈𝖽(AA),{\sf mod}\ (A\otimes A), to the forgetful functor. A tensor category structure on 𝒞{\mathcal{C}} making {\mathcal{F}} into a functor of tensor categories is essentially one giving an AA-module structure to the kk-space VkWV\otimes_{k}W for each pair of modules V,WV,W. In general a tensor category structure on 𝒞{\mathcal{C}} is realizable as an exact functor :𝒞×𝒞𝒞,\otimes:{\mathcal{C}}\hbox{$\Box$}\kern-7.7778pt\hbox{$\times$}{\mathcal{C}}\to{\mathcal{C}}, relative to 𝖵𝖾𝖼k{\sf Vec}\ k, together with coherence laws of associators, etc. By the previous discussion, these are homomorphisms AAAA\to A\otimes A of algebras with coherence laws of associators corresponding to being a bialgebra. Tensor categories being rigid by definition will in turn see that these bialgebras are indeed Hopf algebras. Thus, we have proved the reconstruction theorem for tensor category structures on a fixed category of modules. That is, the classification of Hopf algebra structures on a given finite associative algebra is equivalent to the classification of tensor category structures on its category of modules. It also follows from reconstruction that Hopf algebra structures on a fixed finite augmented algebra are equivalent to tensor category structures with a fixed choice of unit object.

2.2 π\pi-points and a plausibility proposition

We will first review the machinery of π\pi-points, defined by Friedlander and Pevtsova [FrPev07], [FrPev05], which are used in Definition 3.1.1 to define when a homogeneous prime of cohomology is noble, a fundamental notion in Definition 1.1.1 of Property PC. From this machinery we conclude a basic proof for Proposition 2.2.1 stated below, which is otherwise a deeper consequence of earlier work of Friedlander and Parshall [FrPar87], [FrPar86], as well as Suslin, Friedlander, and Bendel [SFB97]. This shows plausibility for Property PC for any given restricted Lie algebra. Recall for a finite group scheme GG over kk, that the total cohomology H(G,k)=ExtG(k,k)=i0ExtiG(k,k)H^{*}(G,k)=\operatorname{Ext}^{*}_{G}(k,k)=\bigoplus_{i\geq 0}\operatorname{Ext}^{i}_{G}(k,k), for kk the trivial representation, is a graded commutative algebra with the cup product (see e.g. Benson [bensonI]).

Proposition 2.2.1.

Let 𝔤{\mathfrak{g}} be a finite dimensional restricted Lie algebra over kk, and H(𝔤,k)=Ext𝔤(k,k)H^{*}({\mathfrak{g}},k)=\operatorname{Ext}^{*}_{{\mathfrak{g}}}(k,k) the cohomology ring. Then every homogeneous prime 𝔭ProjH(𝔤,k){\mathfrak{p}}\in\operatorname{Proj}H^{*}({\mathfrak{g}},k) is the radical ideal ker(α)\sqrt{\operatorname{ker}(\alpha^{*})} for the composition

α:H(𝔤,k)kKH(𝔤K,K)H(ι)H(𝔥,K),\alpha^{*}:H^{*}({\mathfrak{g}},k)\xrightarrow{\otimes_{k}K}H^{*}({\mathfrak{g}}_{K},K)\xrightarrow{H^{*}(\iota)}H^{*}({\mathfrak{h}},K),

induced by the inclusion ι:𝔥𝔤K\iota:{\mathfrak{h}}\to{\mathfrak{g}}_{K} of some 1-dimensional Lie subalgebra 𝔥{\mathfrak{h}}, restricted by 𝔥[p]=0,{\mathfrak{h}}^{[p]}=0, of the base change 𝔤K=𝔤kK{\mathfrak{g}}_{K}={\mathfrak{g}}\otimes_{k}K to a field extension K/k.K/k.

If AA is any algebra over kk and MM is any module over AA, we write AK=AkKA_{K}=A\otimes_{k}K to be the base changed algebra over KK and MK=MkKM_{K}=M\otimes_{k}K to be the base changed module over AKA_{K}.

Definition 2.2.2.

[FrPev07] Let GG be a finite group scheme over kk. A π\pi-point of GG (defined over a field extension K/kK/k) is a (left) flat map of KK-algebras

αK:K[t]/tpKG\alpha_{K}:K[t]/t^{p}\to KG

which factors through the group algebra KCKKGK=KGKC_{K}\subset KG_{K}=KG of some unipotent abelian subgroup scheme CKC_{K} of GKG_{K}.

If βL:L[t]/tpLG\beta_{L}:L[t]/t^{p}\to LG is another π\pi-point of GG, then αK\alpha_{K} is said to be a specialization of βL\beta_{L}, written βLαK\beta_{L}\downarrow\alpha_{K}, provided that for any finite dimensional kGkG-module MM, αK(MK)\alpha^{*}_{K}(M_{K}) being free implies βL(ML)\beta^{*}_{L}(M_{L}) is free.

Two π\pi-points αK,βL,\alpha_{K},\beta_{L}, are said to be equivalent, written αKβL\alpha_{K}\sim\beta_{L}, if αKβL\alpha_{K}\downarrow\beta_{L} and βLαK.\beta_{L}\downarrow\alpha_{K}.

The points αK,βL\alpha_{K},\beta_{L} are said to be strongly equivalent if for any module (not necessarily finite dimensional) MM, αK(MK)\alpha^{*}_{K}(M_{K}) is projective if and only if βL(ML)\beta^{*}_{L}(M_{L}) is projective. It is shown that equivalence implies strong equivalence, and hence the notions coincide.

2.2.3.

Denote by 𝔾a(r){\mathbb{G}}_{a(r)} the rrth Frobenius kernel for the additive group scheme 𝔾a{\mathbb{G}}_{a} over kk. A quasi-elementary group scheme is one in the form

E𝔾a(r)×(/p)s.E\cong{\mathbb{G}}_{a(r)}\times({\mathbb{Z}}/p)^{s}.

When EE is quasi-elementary, we have isomorphism of the group algebra

kEk[t1,,tr+s]/(t1p,,tr+sp),kE\cong k[t_{1},\dots,t_{r+s}]/(t_{1}^{p},\dots,t_{r+s}^{p}),

with the first rr variables tit_{i} dual to the basis elements tpit^{p^{i}} in the coordinate algebra k[𝔾a(r)]=k[t]/tpr.k[{\mathbb{G}}_{a(r)}]=k[t]/t^{p^{r}}. It is shown in [FrPev05] that each π\pi-point defined over the base kk (originally called a pp-point with the same equivalence relation) is equivalent to some α:k[t]/tpkG\alpha:k[t]/t^{p}\to kG which factors through the group algebra kEkGkE\subset kG for some quasi-elementary subgroup scheme EGE\subset G. In fact the base changed statement can be shown for a π\pi-point defined over any field extension.

Notice, for any finite Hopf algebra AA with comultiplication Δ:AAA\Delta:A\to A\otimes A we may define a restricted Lie subalgebra P(A)Lie(A)P(A)\subset\operatorname{Lie}(A) the primitive subspace of AA, i.e. xAx\in A such that Δ(x)=x1+1x.\Delta(x)=x\otimes 1+1\otimes x. The universal enveloping algebra u(P(A))u(P(A)) is isomorphic to the Hopf subalgebra of AA generated as an associative algebra by the subspace P(A).P(A).

If EE is a quasi-elementary group scheme and the group algebra kEkE is given coordinates tit_{i} as above, direct computation shows that P(kE)P(kE) is one-dimensional, generated by t1.t_{1}. Thus, 𝔾a(1){\mathbb{G}}_{a(1)} is the only quasi-elementary group scheme with group algebra isomorphic as a Hopf algebra to a universal enveloping algebra for a restricted Lie algebra.

2.2.4.

Now we review the relationship with cohomology. The algebra H(G,k)H^{*}(G,k) is graded-commutative, meaning not necessarily commutative. However we do have that every homogeneous element is either central or nilpotent. So we write ProjH(G,k)\operatorname{Proj}H^{*}(G,k) to mean the space of homogeneous primes for the reduction of H(G,k)H^{*}(G,k), a commutative, graded algebra. In characteristic p=2p=2 we have that homogeneous elements of any degree may survive, but for characteristic p>2,p>2, this means only the even degree elements may survive.

Denote the algebra D=K[t]/tpD=K[t]/t^{p} over KK, an extension field of kk. There exists a Hopf algebra structure on DD (in fact there are two up to isomorphism, by a theorem of Oort and Tate [OT1970]) showing that the Hopf algebra cohomology H(D,K)H^{*}(D,K) is also a graded-commutative algebra. Given a π\pi-point αK:DKGK\alpha_{K}:D\to KG_{K} defined over K/kK/k, we define the ideal 𝔭(αK){\mathfrak{p}}(\alpha_{K}) as the radical of the kernel for the composition

H(G,k)kKH(GK,K)H(αK)H(D,K).H^{*}(G,k)\xrightarrow{\otimes_{k}K}H^{*}(G_{K},K)\xrightarrow{H^{*}(\alpha_{K})}H^{*}(D,K).

It is shown in [FrPev07] that 𝔭(αK){\mathfrak{p}}(\alpha_{K}) is always a homogeneous prime in ProjH(G,k),\operatorname{Proj}H^{*}(G,k), that every homogeneous prime of ProjH(G,k)\operatorname{Proj}H^{*}(G,k) is of the form 𝔭(α){\mathfrak{p}}(\alpha) for some π\pi-point α\alpha of GG, and further, for two π\pi-points αK,βL,\alpha_{K},\beta_{L}, that αKβL\alpha_{K}\sim\beta_{L} coincides with the equivalence relation 𝔭(αK)=𝔭(βL).{\mathfrak{p}}(\alpha_{K})={\mathfrak{p}}(\beta_{L}).

The last thing we need to prove Proposition 2.2.1 is the following theorem of Milnor and Moore, found as [MM65, Theorem 6.11]

Theorem 2.2.5.

Let 𝔤{\mathfrak{g}} be a finite dimensional restricted Lie algebra over kk, and A=u(𝔤)A=u({\mathfrak{g}}) the restricted enveloping algebra, a cocommutative Hopf algebra with 𝔤A{\mathfrak{g}}\subset A the subspace of primitive elements. Then if AAA^{\prime}\subset A is any Hopf subalgebra, there exists a restricted Lie subalgebra 𝔤𝔤{\mathfrak{g}}^{\prime}\subset{\mathfrak{g}} such that A=u(𝔤)A^{\prime}=u({\mathfrak{g}}^{\prime}) and AAA^{\prime}\subset A is the induced inclusion.

Contrast now Proposition 2.2.1 with the cohomological structure of π\pi-points reviewed in 2.2.4. Identifying R=H(u(𝔤),k)=H(𝔤,k)R=H^{*}(u({\mathfrak{g}}),k)=H^{*}({\mathfrak{g}},k), we claim that each prime in ProjR\operatorname{Proj}R comes from a π\pi-point K[t]/tpu(𝔤K)K[t]/t^{p}\to u({\mathfrak{g}}_{K}) which is the inclusion of a Hopf subalgebra isomorphic to K𝔾a(1).K{\mathbb{G}}_{a(1)}. It is not true in general that π\pi-points are equivalent to some Hopf subalgebra inclusion, and in fact the failure of this property for general finite group schemes is what we are studying with Property PC (Definition 1.1.1), at the level of tensor categories.

Proof of Proposition 2.2.1

Let GG be the finite group scheme with group algebra kG=u(𝔤)kG=u({\mathfrak{g}}) as cocommutative Hopf algebras. Let EKE_{K} be a quasi-elementary subgroup scheme of GKG_{K}. By theorem 2.2.5, the map KEKKGKKE_{K}\to KG_{K} is the induced map of enveloping algebras for a subalgebra of 𝔤.{\mathfrak{g}}. In particular, KEKKE_{K} is generated by its space of primitive elements. Since EE is quasi-elementary, by our discussion 2.2.3 we have E𝔾a(1)E\cong{\mathbb{G}}_{a(1)} with KEKK[t]/tp.KE_{K}\cong K[t]/t^{p}. Every π\pi-point is equivalent to some α:K[t]/tpKGK\alpha:K[t]/t^{p}\to KG_{K} factoring through a quasi-elementary subgroup scheme of EE of GG, and in this case we have shown K[t]/tpKEKK[t]/t^{p}\to KE_{K} is an isomorphism, and α\alpha is a map of Hopf algebras assuming tt is primitive. ∎

2.2.6.

The reason we say Proposition 2.2.1 shows plausibility for Property PC is as follows. In Definition 1.1.1, we have quantified Property PC over all cocommutative Hopf algebra structures on A=u(𝔤)A=u({\mathfrak{g}}) for a restricted Lie algebra 𝔤.{\mathfrak{g}}. One such structure is the Lie comultiplication and associated tensor product Δ=Δ~,=~,\Delta=\widetilde{\Delta},\otimes=\widetilde{\otimes}, defining a finite group scheme we’ll call GG. Without reviewing definitions of π\pi-support and the categories 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) yet, it is tautological that the products ,~\otimes,\widetilde{\otimes} define the same Green ring structure on 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}). So for 𝔤{\mathfrak{g}} to satisfy PC, it is necessary for each π\pi-point to be noble for GG i.e. equivalent to the inclusion of a Lie subalgebra 𝔥𝔤{\mathfrak{h}}\subset{\mathfrak{g}} by Theorem 2.2.5. This is what is guaranteed by Proposition 2.2.1.

2.2.7.

(Generalized π\pi-points) Let αK:K[t]/tpKGK\alpha_{K}:K[t]/t^{p}\to KG_{K} be a flat map. The π\pi-point condition, that αK\alpha_{K} factors through a unipotent subgroup scheme of GKG_{K} turns out to be too strong for some of our purposes. We want to argue when the radical of the kernel for the composition H(αK)(kK)H^{*}(\alpha_{K})\circ(\otimes_{k}K) is a homogeneous prime in ProjH(G,k)\operatorname{Proj}H^{*}(G,k). Taking D=K[t]/tp,D=K[t]/t^{p}, the cohomology ring is calculated as

H(D,K)=ExtD(K,K)={K[ξ]Λ(η)p>2K[ζ]p=2H^{*}(D,K)=\operatorname{Ext}_{D}(K,K)=\begin{cases}K[\xi]\otimes\Lambda(\eta)&p>2\\ K[\zeta]&p=2\end{cases}

where |ξ|=2,|η|=1|\xi|=2,|\eta|=1, and |ζ|=1|\zeta|=1. In particular, regardless of characteristic, the reduced algebra H(D,K)redH^{*}(D,K)_{\mathrm{red}} is an integral domain. The radical of the ideal ker(H(αK)(kK))\operatorname{ker}(H^{*}(\alpha_{K})\circ(-\otimes_{k}K)) in H(G,k)H^{*}(G,k) agrees with the kernel of the composition redH(αK)(kK),\mathrm{red}\circ H^{*}(\alpha_{K})\circ(\otimes_{k}K), i.e. changing the target from H(D,K)H^{*}(D,K) to H(D,K)redH^{*}(D,K)_{\mathrm{red}}. This ideal is always a homogeneous prime, and so being in ProjH(G,k)\operatorname{Proj}H^{*}(G,k) is equivalent to the induced map redH(αK)(kK)\mathrm{red}\circ H^{*}(\alpha_{K})\circ(\otimes_{k}K) being nonzero in some positive degree. Such nondegenerate flat maps αK\alpha_{K} we may call generalized π\pi-points, and we denote by 𝔭(αK){\mathfrak{p}}(\alpha_{K}) the homogeneous prime of ProjH(G,k)\operatorname{Proj}H^{*}(G,k). Friedlander and Pevtsova [FrPev07] show that π\pi-points are nondegenerate in this way, i.e. that factoring through a unipotent subgroup scheme of GKG_{K} is sufficient for knowing the induced map of cohomology is nonzero in some degree. It remains to be shown, for generalized π\pi-points αK,βL\alpha_{K},\beta_{L}, that 𝔭(αK)=𝔭(βL){\mathfrak{p}}(\alpha_{K})={\mathfrak{p}}(\beta_{L}) if and only if αKβL\alpha_{K}\sim\beta_{L}, for an equivalence relation \sim defined similarly in terms of detecting projectivity. However, this can be shown by repeating the methods of [FrPev07] in a straightforward way, so we take it for granted. Note how it immediately follows that generalized π\pi-points are always equivalent to some π\pi-point.

Lemma 2.2.8.

Let GG be a finite group scheme over kk, and let φAut(kG)\varphi\in\operatorname{Aut}(kG) be an augmented automorphism of the augmented algebra kGkG. Let α:K[t]/tpKGK\alpha:K[t]/t^{p}\to KG_{K} be a π\pi-point of GG over KK. Then the composition φKα\varphi_{K}\circ\alpha is a generalized π\pi-point of GG over KK, where φKAut(KGK)\varphi_{K}\in\operatorname{Aut}(KG_{K}) is the base change. If βL\beta_{L} is another π\pi-point and αKβL,\alpha_{K}\sim\beta_{L}, then φKαKφLβL\varphi_{K}\circ\alpha_{K}\sim\varphi_{L}\circ\beta_{L} as generalized π\pi-points.

Proof.

Let A=KGKA=KG_{K} be the group algebra over KK, with Δ:AAA\Delta:A\to A\otimes A its cocommutative Hopf algebra structure. Suppose αK:K[t]/tpA\alpha_{K}:K[t]/t^{p}\to A factors through the inclusion KUAKU\hookrightarrow A of a unipotent subgroup scheme U<GK.U<G_{K}. The composition φKαK\varphi_{K}\circ\alpha_{K} is automatically a flat map (since αK\alpha_{K} and φK\varphi_{K} are both flat). So, to show that φKαK\varphi_{K}\circ\alpha_{K} is a generalized π\pi-point, it suffices to show that there is some cocommutative Hopf algebra structure Δ\Delta^{\prime} on AA such that φKαK\varphi_{K}\circ\alpha_{K} factors through the inclusion BAB\hookrightarrow A of a local Hopf-subalgebra BB. Letting G=(SpecA,Δ)G^{\prime}=(\operatorname{Spec}A^{*},\Delta^{\prime}) and U=SpecBU^{\prime}=\operatorname{Spec}B^{*}, we see UU^{\prime} is a unipotent subgroup scheme of G.G^{\prime}. Since the cohomology rings and their induced maps are invariant between Hopf algebra structures, we see any such choice of Δ\Delta^{\prime} and unipotent subgroup scheme U<GU^{\prime}<G^{\prime} will have that the induced map of cohomology from φKαK\varphi_{K}\circ\alpha_{K} is nondegenerate, making φKαK\varphi_{K}\circ\alpha_{K} a generalized π\pi-point of GG, as shown in [FrPev07].

Define Δ=(φKφK)ΔφK1\Delta^{\prime}=(\varphi_{K}\otimes\varphi_{K})\circ\Delta\circ\varphi_{K}^{-1} (c.f. 3.1.7). We see Δ\Delta^{\prime} is indeed a cocommutative comultiplication for a Hopf algebra structure on AA. Further, if we define B=φK(KU)B=\varphi_{K}(KU) the image of the group algebra, we see BB is a Hopf-subalgebra of A,ΔA,\Delta^{\prime}. Now BB is local as it is isomorphic to KUKU, and φKαK\varphi_{K}\circ\alpha_{K} factors through the inclusion BA.B\hookrightarrow A. We conclude φKαK\varphi_{K}\circ\alpha_{K} is a π\pi-point for GG^{\prime} and hence a generalized π\pi-point for GG over KK.

Now let βL\beta_{L} be a π\pi-point over L/kL/k such that αKβL\alpha_{K}\sim\beta_{L}. Then for any kGkG module MM we have αK(MK)\alpha_{K}^{*}(M_{K}) is projective if and only if βL(ML)\beta_{L}^{*}(M_{L}) is projective. There is natural isomorphism of K[t]/tpK[t]/t^{p} modules (φKαK)(M)αK(φ(M)K),(\varphi_{K}\circ\alpha_{K})^{*}(M)\cong\alpha_{K}^{*}(\varphi^{*}(M)_{K}), and similarly for βL.\beta_{L}. Since φ(M)\varphi^{*}(M) is a kGkG module, we have αK(φ(M)K)\alpha^{*}_{K}(\varphi^{*}(M)_{K}) is projective if and only if βL(φ(M)L)\beta_{L}^{*}(\varphi^{*}(M)_{L}) is projective, and hence φKαKφLβL.\varphi_{K}\circ\alpha_{K}\sim\varphi_{L}\circ\beta_{L}.

There is a hence a well defined Aut(kG)\operatorname{Aut}(kG)-action on ProjH(G,k)\operatorname{Proj}H^{*}(G,k) by

φ𝔭=𝔭(φα) for a π-point α such that 𝔭=𝔭(α).\varphi\cdot{\mathfrak{p}}={\mathfrak{p}}(\varphi\circ\alpha)\text{ for a }\pi\text{-point }\alpha\text{ such that }{\mathfrak{p}}={\mathfrak{p}}(\alpha).

2.3 Cohomological support and tt-geometry

2.3.1.

In 2.2 we reviewed how the space of π\pi-points for a finite group scheme GG over kk is equivalent to 𝒳(G)=ProjH(G,k){\mathscr{X}}(G)=\operatorname{Proj}H^{*}(G,k). Let R=H(G,k)R=H^{*}(G,k), a graded-commutative algebra. For each representation MM of GG, the cohomology

H(G,M)=ExtG(k,M)H^{*}(G,M)=\operatorname{Ext}^{*}_{G}(k,M)

has a canonical graded right-module structure over RR by the Yoneda splice product. By the theorem of Friedlander and Suslin [FS97], 𝒳(G){\mathscr{X}}(G) is a (possibly reducible) projective variety for finite group schemes GG, and H(G,M)H^{*}(G,M) gives a coherent sheaf over 𝒳(G){\mathscr{X}}(G) for finite dimensional MM. To MM we therefore associate a closed subvariety of cohomological support

𝒳(G,M)=V(AnnR(H(G,M))).{\mathscr{X}}(G,M)=V(\operatorname{Ann}_{R}(H^{*}(G,M))).

It is further shown by Friedlander and Pevtsova [FrPev07] that this subvariety is equivalent to π\pi-support, i.e.

𝒳(G,M)\displaystyle{\mathscr{X}}(G,M) =suppG(M)\displaystyle=\operatorname{supp}_{G}(M)
={[αK]αK(MK) is not projective},\displaystyle=\left\{[\alpha_{K}]\ \mid\ \alpha^{*}_{K}(M_{K})\text{ is not projective}\right\},

defined to range over equivalence classes of π\pi-points αK.\alpha_{K}.

2.3.2.

Now that we have a definition for support, we can elaborate on the equivalence relation for π\pi-points given in Definition 2.2.2. We continue our assumption that kk is algebraically closed, and so the closed points of 𝒳(G){\mathscr{X}}(G) are all of the form 𝔭(α){\mathfrak{p}}(\alpha) for a π\pi-point α\alpha of GG, defined over the ground field kk. The equivalence relation αβ\alpha\sim\beta between π\pi-points α,β\alpha,\beta of GG over kk is, by definition, that for any finite module MM, α(M)\alpha^{*}(M) is projective if and only if β(M)\beta^{*}(M) is projective. In practice, we may be given a finite group algebra kGkG with generators and relations, and the π\pi-points of GG over kk make an affine-algebraic subset Π\Pi of 𝔸(kG)=SpecS(𝒪(G)){\mathbb{A}}(kG)=\operatorname{Spec}S({\mathcal{O}}(G)). So it is preferable to characterize the equivalence relation in coordinates. It turns out fixing a closed point 𝔭𝒳(G){\mathfrak{p}}\in{\mathscr{X}}(G), in many important cases, there are standard techniques for producing a finite module MM such that 𝒳(G,M)={𝔭}.{\mathscr{X}}(G,M)=\{{\mathfrak{p}}\}. When such MM is known, the equivalence class of π\pi-points over kk {αΠ𝔭=𝔭(α)}\{\alpha\in\Pi\mid{\mathfrak{p}}={\mathfrak{p}}(\alpha)\} is the same as

{αΠα(M) is not projective}.\{\alpha\in\Pi\mid\alpha^{*}(M)\text{ is not projective}\}.

By computing Jordan canonical forms, fixing αΠ\alpha\in\Pi such that 𝔭=𝔭(α),{\mathfrak{p}}={\mathfrak{p}}(\alpha), it becomes straightforward to characterize βΠ\beta\in\Pi such that αβ.\alpha\sim\beta. The computational advantage here is that we only need to consider a single finite module MM rather than range over all MM.

2.3.3.

A universal approach toward support is given for tensor-triangulated categories in Balmer’s tensor-triangular geometry [Balmer05]. For representations of a finite group scheme GG, we look at the tensor triangulated category given by finite stable representations, denoted stmodkG.\operatorname{stmod}\nolimits kG. That is, with kGkG the cocommutative group algebra associated to GG, we look at the category of finite dimensional modules, with Hom spaces between objects X,YX,Y given by

Hom¯G(X,Y):=HomG(X,Y)/𝒫(X,Y),\operatorname{\underline{Hom}}_{G}(X,Y):=\operatorname{Hom}_{G}(X,Y)/{\mathcal{P}}(X,Y),

where 𝒫(X,Y){\mathcal{P}}(X,Y) is the subspace of maps factoring through a projective module. This category is tensor-triangulated, a general fact for the stable category of a Frobenius category with exact monoidal product (see e.g. Keller, [Keller94]). The work of Benson, Carlson, Rickard [BCR96], and Friedlander and Pevstova [BIKP15], [BIKP18], classifies the thick \otimes-ideals of stmodkG.\operatorname{stmod}\nolimits kG. In Balmer’s tt-geometric terms, what this means is that the projective variety 𝒳(G){\mathscr{X}}(G) defined above is the spectrum of the tt-category stmodkG\operatorname{stmod}\nolimits kG.

Recall the basic elements for tt-structure on stmodkG\operatorname{stmod}\nolimits kG: The algebra kGkG is a cocommutative Hopf algebra with counit kGkkG\to k defining kk to be the trivial module, the monoidal unit with respect to the product \otimes of representations (see Section 2.1). A finite representation PP is projective if and only if it is isomorphic to 0 as an object of stmodkG\operatorname{stmod}\nolimits kG, and the projective modules form an \otimes-ideal, meaning the monoidal structure of 𝗆𝗈𝖽kG{\sf mod}\ kG descends to the quotient stmodkG.\operatorname{stmod}\nolimits kG.

The triangulated structure has the suspension autoequivalence Σ,\Sigma, defined on objects as (co)syzygies, i.e. ΣM=coker(ι)\Sigma M=\operatorname{coker}(\iota) where ι:MI\iota:M\hookrightarrow I is a minimal injective embedding of MM, and the inverse supspension defined by syzygies Σ1M=ΩM=ker(ϵ)\Sigma^{-1}M=\Omega M=\operatorname{ker}(\epsilon) where ϵ:PM\epsilon:P\twoheadrightarrow M is a minimal projective cover of MM. The exact triangles come from exact sequences of modules; see, e.g., Happel [Happel87] or Keller [Keller94].

Definition 2.3.4.

A full subcategory 𝒞{\mathcal{C}} of 𝗆𝗈𝖽kG{\sf mod}\ kG is triangulated if

  1. (T1)

    Every finite projective kGkG-module is contained in 𝒞{\mathcal{C}}, and

  2. (T2)

    For every short exact sequence of finite modules

    0MMM0,0\to M^{\prime}\to M\to M^{\prime\prime}\to 0,

    if two of the modules in {M,M,M}\{M^{\prime},M,M^{\prime\prime}\} belong to 𝒞{\mathcal{C}} then so does the third.

Notice a triangulated subcategory is closed under isomorphism, and also under the autoequivalences Σ,Ω.\Sigma,\Omega.

The subcategory 𝒞{\mathcal{C}} is called thick if in addition

  1. 3.

    Whenever MMM\oplus M^{\prime} belongs to 𝒞{\mathcal{C}}, so do the summands M,MM,M^{\prime}.

The subcategory 𝒞{\mathcal{C}} is called \otimes-ideal if in addition

  1. 4.

    Whenever MM belongs to 𝒞{\mathcal{C}}, NMN\otimes M belongs to MM for any module NN.

The subcategory 𝒞{\mathcal{C}} is called radical if in addition

  1. 5.

    If the nn-fold product MnM^{\otimes n} belongs to 𝒞{\mathcal{C}}, so does the module MM.

The properties (T1)-(T5) for the tensor category 𝗆𝗈𝖽kG{\sf mod}\ kG all descend to the quotient stmodkG\operatorname{stmod}\nolimits kG to define the corresponding notions [Balmer05] for subcategories of the tt-category. Now we recall what it means to be a classifying support data on the tt-category stmodkG.\operatorname{stmod}\nolimits kG.

Definition 2.3.5.

[Balmer05] A support data on the tt-category stmodkG\operatorname{stmod}\nolimits kG is a pair (𝒳,σ),({\mathscr{X}},\sigma), where 𝒳{\mathscr{X}} is a topological space and σ\sigma is an assignment which associates to any object MstmodkGM\in\operatorname{stmod}\nolimits kG a closed subset σ(M)𝒳\sigma(M)\subset{\mathscr{X}} subject to the following rules:

  1. (S1)

    σ(0)=\sigma(0)=\emptyset,

  2. (S2)

    σ(MM)=σ(M)σ(M)\sigma(M\oplus M^{\prime})=\sigma(M)\cup\sigma(M^{\prime}),

  3. (S3)

    σ(ΣM)=σ(ΩM)=σ(M)\sigma(\Sigma M)=\sigma(\Omega M)=\sigma(M),

  4. (S4)

    σ(M)σ(M)σ(M)\sigma(M)\subset\sigma(M^{\prime})\cup\sigma(M^{\prime\prime}) for any short exact sequence

    0MMM0,0\to M^{\prime}\to M\to M^{\prime\prime}\to 0,
  5. (S5)

    σ(MM)=σ(M)σ(M).\sigma(M\otimes M^{\prime})=\sigma(M)\cap\sigma(M^{\prime}).

A support data (𝒳,σ)({\mathscr{X}},\sigma) for stmodkG\operatorname{stmod}\nolimits kG is a classifying support data if the following two conditions hold:

  1. (C1)

    The topological space 𝒳{\mathscr{X}} is noetherian and any non-empty irreducible closed subset 𝒵𝒳{\mathscr{Z}}\subset{\mathscr{X}} has a unique generic point: !x𝒵\exists!\ x\in{\mathscr{Z}} with {x}¯=𝒵\overline{\{x\}}={\mathscr{Z}},

  2. (C2)

    We have a bijection

    {Thomason subsets\displaystyle\{\text{Thomason subsets } 𝒴𝒳}\displaystyle{\mathscr{Y}}\subset{\mathscr{X}}\}\xrightarrow{\sim}
    {thick -ideals 𝒥stmodkG}\displaystyle\{\text{thick }\otimes\text{-ideals }{\mathcal{J}}\subset\operatorname{stmod}\nolimits kG\}

    defined by 𝒴{MstmodkGσ(M)𝒴}{\mathscr{Y}}\mapsto\{M\in\operatorname{stmod}\nolimits kG\mid\sigma(M)\subset{\mathscr{Y}}\}, with inverse 𝒥σ(𝒥):=M𝒥σ(M).{\mathcal{J}}\mapsto\sigma({\mathcal{J}}):=\bigcup_{M\in{\mathcal{J}}}\sigma(M).

The theorem of Friedlander and Suslin [FS97] shows that the purely topological condition 1 holds for 𝒳=𝒳(G){\mathscr{X}}={\mathscr{X}}(G). The work of Friedlander and Pevtsova [FrPev07], [FrPev05], includes that taking 𝒳=𝒳(G){\mathscr{X}}={\mathscr{X}}(G) and defining cohomological support σ(M)=𝒳(G,M)\sigma(M)={\mathscr{X}}(G,M) makes (𝒳,σ)({\mathscr{X}},\sigma) into a support data for stmodG\operatorname{stmod}\nolimits G. With Benson, Iyengar, Krause, and Pevtsova [BIKP15], [BIKP18], we have indeed that cohomology gives a classifying support data for stmodG.\operatorname{stmod}\nolimits G. As a consequence, 𝒳(G){\mathscr{X}}(G) has a certain universal property for support data (see Balmer’s [Balmer05, Theorem 5.2]) making it the spectrum of the tensor triangulated category stmodkG.\operatorname{stmod}\nolimits kG.

Some elementary considerations show that in fact the bijection from condition 2 preserves inclusion. Thus, we know how to characterize minimal (radical, \otimes-ideal) thick subcategories; they are the subcategories 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) supported at a singleton closed point 𝔭𝒳(G).{\mathfrak{p}}\in{\mathscr{X}}(G). Any finite module has support a closed set, so those with a singleton support in particular have support a closed point.

2.3.6.

On the constancy of classifying support between different Hopf algebra structures: If AkA\to k is an augmentation map, we can define a graded algebra structure on the cohomology H(A,k)=Ext(k,k)H^{*}(A,k)=\operatorname{Ext}^{*}(k,k) by splicing Yoneda extensions. If AA is given a Hopf algebra structure such that AkA\to k is the counit, one shows the splicing of Yoneda extensions is equivalent to the cup product, which is known to be graded-commutative (see e.g. Benson, [bensonI] in the case of cocommutative Hopf algebras). The same is true of the graded right-module structure on H(A,M)=Ext(k,M)H^{*}(A,M)=\operatorname{Ext}^{*}(k,M).

What we have now, is that even though applying the tt-geometry methods discussed for AA-modules depends on the existence of a cocommutative Hopf algebra structure on AA giving a symmetric monoidal product to begin with, the variety 𝒳(A)=ProjH(A,k){\mathscr{X}}(A)=\operatorname{Proj}H^{*}(A,k) does not depend on which Hopf algebra structure is chosen, and the supports 𝒳(A,M){\mathscr{X}}(A,M) of a finite module MM are in this way also independent. They all satisfy the tensor product property 5, and in particular the subcategories 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) supported at a point 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A) are closed under any symmetric tensor product chosen.

We have now concluded our review of terminology used in Definition 1.1.1.

3 A property of some Lie algebras

3.1 Definitions

Definition 3.1.1.

Let GG be a finite group scheme over a field kk. We say a π\pi-point is noble for GG if it is equivalent to a map

α:K[t]/tpKGK\alpha:K[t]/t^{p}\to KG_{K}

such that the image α(K[t]/tp)\alpha(K[t]/t^{p}) is a Hopf subalgebra of KGK,KG_{K}, and the point is ignoble for GG otherwise. We may also refer to an equivalence class of a π\pi-point being noble or ignoble, as well as its realization as a point 𝔭ProjH(G,k).{\mathfrak{p}}\in\operatorname{Proj}H^{*}(G,k).

Example 3.1.2.

The Klein 4-group G=h,gh2=g2=(gh)2=1G=\langle h,g\mid h^{2}=g^{2}=(gh)^{2}=1\rangle, for p=2p=2, has 3 noble π\pi-points up to equivalence, corresponding to its 3 cyclic subgroups generated by j=h,g,hg,j=h,g,hg, and defining a π\pi-point tj1.t\mapsto j-1. We revisit this in Sections 3.2, 3.3. Every flat map k[t]/tpkGk[t]/t^{p}\to kG is defined by

ta(h1)+b(g1)+c(h1)(g1),t\mapsto a(h-1)+b(g-1)+c(h-1)(g-1),

for nonzero vector (a,b)k2(a,b)\in k^{2}, and equivalence of π\pi-points identifies the triples (a,b,c)(a,b,c)(a,b,c)\sim(a^{\prime},b^{\prime},c^{\prime}) if and only if abab=0ab^{\prime}-a^{\prime}b=0 (this can be seen explicitly after classifying all modules, GG being of tame representation type). Thus 𝒳(G)=1{\mathscr{X}}(G)={\mathbb{P}}^{1}, i.e. points are given homogeneous coordinates [a:b][a:b]. The maps corresponding to the three generators h,g,ghh,g,gh in these coordinates are [1:0],[0:1],[1:1][1:0],[0:1],[1:1] respectively. For GG we have now that [1:0][1:0] is noble and [a:1][a:1] is noble iff a𝔽2a\in{\mathbb{F}}_{2}.

Definition 3.1.3.

A full subcategory 𝒟{\mathcal{D}} of a finite tensor category (𝒞,)({\mathcal{C}},\otimes) (over a field kk) is a semiring subcategory if the set of isomorphism classes of objects 𝒟{\mathcal{D}} is closed under direct sum and tensor product. If Φ:𝒞1𝒞2\Phi:{\mathcal{C}}_{1}\to{\mathcal{C}}_{2} is an equivalence of (kk-linear abelian) categories between finite tensor categories (𝒞i,i),({\mathcal{C}}_{i},\otimes_{i}), and Φ\Phi restricts essentially to an equivalence of categories 𝒟1𝒟2{\mathcal{D}}_{1}\to{\mathcal{D}}_{2} between semiring subcategories 𝒟i𝒞i{\mathcal{D}}_{i}\subset{\mathcal{C}}_{i}, we write (𝒟1,1)(𝒟2,2)({\mathcal{D}}_{1},\otimes_{1})\equiv({\mathcal{D}}_{2},\otimes_{2}) to mean an Φ\Phi induces an isomorphism of Green rings, i.e. Φ(a1b)Φ(a)2Φ(b)\Phi(a\otimes_{1}b)\cong\Phi(a)\otimes_{2}\Phi(b) for each pair of objects a,b𝒟1.a,b\in{\mathcal{D}}_{1}.

Definition 3.1.4.

Let 𝔤{\mathfrak{g}} be a restricted Lie algebra over kk, and A=u(𝔤)A=u({\mathfrak{g}}) the restricted enveloping algebra. Let Δ~,~\widetilde{\Delta},\widetilde{\otimes} be the Lie comultiplication on AA and its associated tensor product of 𝔤{\mathfrak{g}} representations. When GG is a finite group scheme arising from a Hopf algebra structure on the augmented algebra AA, we let Δ,\Delta,\otimes denote the group comultiplication on AA and its associated tensor product of GG-representations. Denote 𝒞(𝔭)\mathcal{C}({\mathfrak{p}}) the minimal thick subcategory of finite AA-modules, with support a singleton 𝔭ProjH(A,k){\mathfrak{p}}\in\operatorname{Proj}H^{*}(A,k).

  1. A.

    The algebra 𝔤{\mathfrak{g}} is said to satisfy Property PA if for any finite group scheme GG as above, and any noble 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A) for GG, that (𝒞(𝔭),)(𝒞(𝔭),~)({\mathcal{C}}({\mathfrak{p}}),\otimes)\equiv({\mathcal{C}}({\mathfrak{p}}),\widetilde{\otimes}) as semiring subcategories (Definition 3.1.3).

  2. B.

    The algebra 𝔤{\mathfrak{g}} is said to satisfy Property PB if for any finite group scheme GG as above, and any ignoble 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A) for GG, there exists V,W𝒞(𝔭)V,W\in{\mathcal{C}}({\mathfrak{p}}) such that VWV\otimes W is not isomorphic to V~W.V\widetilde{\otimes}W.

It is clear from definitions that Properties PA, PB are converse to one another and that in conjunction they are Property PC of 1.1.1.

3.1.5.

We must emphasize the formal meaning of quantifying our properties PA and PB over all group schemes GG having a given group algebra AA. On one hand, kk being algebraically closed and A=u(𝔤)A=u({\mathfrak{g}}) being finite dimensional, we may naïvely define an affine variety {\mathscr{B}} of cocommutative bialgebra structures Δ:AAA\Delta:A\to A\otimes A with fixed counit AkA\to k, an affine variety 𝒮{\mathscr{S}} of linear maps S:AAS:A\to A, and a closed subvariety ×𝒮{\mathscr{H}}\subset{\mathscr{B}}\times{\mathscr{S}} of cocommutative Hopf algebras (Δ,S)(\Delta,S) with SS an antipode for the comutiplication Δ\Delta. In fact, antipodes being uniquely determined by comultiplications, the composition ×𝒮{\mathscr{H}}\to{\mathscr{B}}\times{\mathscr{S}}\to{\mathscr{B}} is injective on kk-points. On the other hand the work of X. Wang et. al. [NgW23], [NWW15], [NWW16], [WW14], [XWang13], [XWang15] classifies Hopf algebras in small dimension up to equivalence. For our purposes, these classifications can provide a computation of the orbit space /Aut(A){\mathscr{H}}/\operatorname{Aut}(A) (of kk-points) where Aut(A)\operatorname{Aut}(A) is the group of augmented kk-algebra automorphisms φ:AA\varphi:A\to A acting on ×𝒮{\mathscr{B}}\times{\mathscr{S}} by

(Δ,S)(Δφ,Sφ)=((φφ)Δφ1,φSφ1),(\Delta,\,S)\mapsto(\Delta^{\varphi},\,S^{\varphi})=((\varphi\otimes\varphi)\circ\Delta\circ\varphi^{-1},\,\varphi\circ S\circ\varphi^{-1}),

making {\mathscr{H}} an invariant subvariety.

In these terms, what we have is that for the properties P == PA, PB, PC, we defined implicitly an existential P(s){}^{\prime}(s) dependent on a set ss\in{\mathscr{H}} of kk-points such that Property P is in the form

P:=s,(P(s) holds),\text{P}:=\forall s\in{\mathscr{H}},\,\,\left(\text{P}^{\prime}(s)\text{ holds}\right),

per definition. But to make good use the work of X. Wang et. al. while avoiding the enormous computation of {\mathscr{H}}, we must confirm a curtailment of Property P to be valid for u(𝔤)u({\mathfrak{g}}). That is, we would like to range over the set t/Aut(A)t\in{\mathscr{H}}/\operatorname{Aut}(A), realized as a complete set of Aut(A)\operatorname{Aut}(A)-orbit representatives in ,{\mathscr{H}}, and confirm that

(t/Aut(A),(P(t) holds ))P.\left(\forall t\in{\mathscr{H}}/\operatorname{Aut}(A),\,\,\left(\text{P}^{\prime}(t)\text{ holds }\right)\right)\implies\text{P}. (3.1.6)

The curtailments 3.1.6 for Properties P == PA, PB, PC are not immediate for a given 𝔤{\mathfrak{g}}. The curtailments would follow if for example it is known that

P(s)(P(sφ)φAut(A))\text{P}^{\prime}(s)\implies\left(\text{P}^{\prime}(s^{\varphi})\quad\forall\varphi\in\operatorname{Aut}(A)\right)

for each s=(Δ,S),s=(\Delta,S)\in{\mathscr{H}}, where sφ=(Δφ,Sφ)s^{\varphi}=(\Delta^{\varphi},S^{\varphi}). To see why this is not immediate, consider the following lemmas.

Lemma 3.1.7.

(Twisting Hopf algebras) Let φAut(A)\varphi\in\operatorname{Aut}(A) be an augmented algebra automorphism, and φ:ProjH(A,k)ProjH(A,k)\varphi^{*}:\operatorname{Proj}H^{*}(A,k)\to\operatorname{Proj}H^{*}(A,k) the induced automorphism on varieties (2.2.8). If Δ\Delta is the comultiplication for a group scheme GG with kGAkG\cong A, denote the twisted group scheme GφG^{\varphi} by the comultiplication Δφ=(φφ)Δφ1\Delta^{\varphi}=(\varphi\otimes\varphi)\circ\Delta\circ\varphi^{-1}. Then 𝔭ProjH(A,k){\mathfrak{p}}\in\operatorname{Proj}H^{*}(A,k) is noble for GG iff φ(𝔭)\varphi^{*}({\mathfrak{p}}) is noble for Gφ.G^{\varphi}.

3.1.8.

Denote by Ω(A,𝔭)\Omega(A,{\mathfrak{p}}) the isotropy subgroup of Aut(A)\operatorname{Aut}(A) for π\pi-point 𝔭{\mathfrak{p}}, and

Ω(A)=𝔭Ω(A,𝔭)\Omega(A)=\bigcap_{{\mathfrak{p}}}\Omega(A,{\mathfrak{p}})

the kernel of Aut(A)Aut(ProjH(A,k))\operatorname{Aut}(A)\to\operatorname{Aut}(\operatorname{Proj}H^{*}(A,k)) which takes φ\varphi to φ\varphi^{*} as in Lemma 3.1.7.

Suppose that Hopf algebras I=/Aut(A)I={\mathscr{H}}/\operatorname{Aut}(A) on a given A=u(𝔤)A=u({\mathfrak{g}}) are classified up to equivalence, as Δi,\Delta_{i}, corresponding to the scheme GiG_{i} and product i\otimes_{i} for iIi\in I. Then for any Hopf algebra structure Δ:AAA,\Delta:A\to A\otimes A, there is some φAut(A)\varphi\in\operatorname{Aut}(A) and iIi\in I with Δ=Δiφ.\Delta=\Delta_{i}^{\varphi}. For modules MM with action π:AkMM\pi:A\otimes_{k}M\to M, denote MφM^{\varphi} the twisted module on the same kk-space MM, but with action the composition

π(φ1idM):AkMφMφ,\pi\circ(\varphi^{-1}\otimes\operatorname{id}_{M}):A\otimes_{k}M^{\varphi}\to M^{\varphi},

i.e. Mφ:=φ(A)AM,M^{\varphi}:=\varphi(A)\otimes_{A}M, the base change along φ.\varphi. Then a prime 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A) belongs to the support variety 𝒳(A,M){\mathscr{X}}(A,M) if and only if φ(𝔭)\varphi^{*}({\mathfrak{p}}) belongs to 𝒳(A,Mφ).{\mathscr{X}}(A,M^{\varphi}).

Lemma 3.1.9.

Suppose for each closed point 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A), each finite AA-module MM with 𝒳(A,M)={𝔭},{\mathscr{X}}(A,M)=\{{\mathfrak{p}}\}, and each isotropy φΩ(A,𝔭),\varphi\in\Omega(A,{\mathfrak{p}}), that MφM.M^{\varphi}\cong M. Then the curtailment 3.1.6 holds for Properties P == PA, PB, PC.

In words for e.g. P == PA: under the same isotropy hypothesis, if for all iIi\in I, and any noble point 𝔭{\mathfrak{p}} for GiG_{i}, we have (𝒞(𝔭),i)(𝒞(𝔭),~)({\mathcal{C}}({\mathfrak{p}}),\otimes_{i})\equiv({\mathcal{C}}({\mathfrak{p}}),\widetilde{\otimes}) as semiring subcategories, then indeed 𝔤{\mathfrak{g}} satisfies Property PA.

The proofs of Lemmas 3.1.7, 3.1.9 are straightforward. One shows with Lemma 3.1.7 that the isotropy hypothesis of Lemma 3.1.9 has, as a consequence, that P(s)P(sφ)\text{P}^{\prime}(s)\implies\text{P}^{\prime}(s^{\varphi}) for Hopf algebras ss\in{\mathscr{H}}, and Properties P == PA, PB, PC. But the isotropy hypothesis is not immediate for restricted enveloping algebras A=u(𝔤),A=u({\mathfrak{g}}), a counterexample is given in 3.2.4.

We conclude this section with two easy examples of restricted Lie algebras of finite representation type which satisfy Property PC.

Example 3.1.10.

Let 𝔤=x{\mathfrak{g}}=\langle x\rangle be the one dimensional Lie algebra, with trivial restriction 𝔤[p]=0.{\mathfrak{g}}^{[p]}=0. The restricted enveloping algebra is given by

u(𝔤)=A=k[x]/xp,u({\mathfrak{g}})=A=k[x]/x^{p},

and by a theorem of Oort and Tate [OT1970], there are only two Hopf algebra structures on AA up to isomorphism. They are given by

  1. 1.

    Δ~:xx1+1x\widetilde{\Delta}:x\mapsto x\otimes 1+1\otimes x,

  2. 2.

    Δ:xx1+1x+xx,\Delta:x\mapsto x\otimes 1+1\otimes x+x\otimes x,

with tensor products denoted ~,\widetilde{\otimes},\otimes respectively. The structure (A,Δ~)(A,\widetilde{\Delta}) is equivalent to the group algebra for 𝔾a(1){\mathbb{G}}_{a(1)}, and the structure (A,Δ)(A,\Delta) is equivalent to the group algebra for /p.{\mathbb{Z}}/p. The algebra AA is a quotient of a PID and it is easy to see how there is exactly one indecomposable module JiJ_{i} of dimension ii, for 1ip1\leq i\leq p, with JpJ_{p} the unique indecomposable projective module. It is known (see e.g. Benson [Benson17]) that Ji~JjJiJjJ_{i}\,\widetilde{\otimes}\,J_{j}\cong J_{i}\otimes J_{j} for any 1i,jp.1\leq i,j\leq p.

It follows that 𝔤{\mathfrak{g}} satisfies Property PC. To elaborate, Proposition 2.2.1 (or better yet, a direct computation of cohomology and cup product) tells us that ProjH(𝔤,k)\operatorname{Proj}H^{*}({\mathfrak{g}},k) consists of a single point 𝔭{\mathfrak{p}}, represented by the identity for 𝔤{\mathfrak{g}}, and with that, each JiJ_{i} belongs to the unique minimal thick subcategory 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}). Finally, we may apply Lemma 3.1.9: in this case Ω(A)=Aut(A)\Omega(A)=\operatorname{Aut}(A), and we see for φΩ(A)\varphi\in\Omega(A), that JiφJ_{i}^{\varphi} is indecomposable of dimension ii, hence isomorphic to JiJ_{i}.

Example 3.1.11.

Let 𝔤=x,y[x,y]=0,x[p]=y,y[p]=0{\mathfrak{g}}=\langle x,y\ \mid\ [x,y]=0,x^{[p]}=y,y^{[p]}=0\rangle be the two dimensional abelian Lie algebra with nontrivial restriction. The restricted enveloping algebra is given by

u(𝔤)=A=k[x]/xp2.u({\mathfrak{g}})=A=k[x]/x^{p^{2}}.

By a corollary of X. Wang [XWang13], there are only three Hopf algebra structures on AA up to isomorphism. They are given by

  1. 1.

    Δ~:xx1+1x,\widetilde{\Delta}:x\mapsto x\otimes 1+1\otimes x,

  2. 2.

    Δ1:xx1+1x+ω(xp),\Delta_{1}:x\mapsto x\otimes 1+1\otimes x+\omega(x^{p}),

  3. 3.

    Δ2:xx1+1x+xx,\Delta_{2}:x\mapsto x\otimes 1+1\otimes x+x\otimes x,

with tensor products ~,1,2\widetilde{\otimes},\otimes_{1},\otimes_{2} respectively. The structure (A,Δ~)(A,\widetilde{\Delta}) is equivalent to the group algebra for the Frobenius kernel 𝕎2(1),{\mathbb{W}}_{2(1)}, where 𝕎i{\mathbb{W}}_{i} is the algebraic group of length ii Witt vectors. The comultiplication Δ1\Delta_{1} depends on a term ω(xp)\omega(x^{p}), defined as

ω(y)=(y1+1y)p(yp1+1yp)p,\omega(y)=\frac{(y\otimes 1+1\otimes y)^{p}-(y^{p}\otimes 1+1\otimes y^{p})}{p},

a formal division by pp in characteristic pp. The structure (A,Δ1)(A,\Delta_{1}) is the group algebra for a certain degree pp subgroup G2G_{2} of the second Frobenius kernel 𝕎2(2){\mathbb{W}}_{2(2)}. Both G2G_{2} and 𝕎2(2){\mathbb{W}}_{2(2)} are equal to their own Cartier dual. The structure (A,Δ2)(A,\Delta_{2}) is equivalent to the group algebra for /(p2).{\mathbb{Z}}/(p^{2}).

As in the Oort-Tate example above, there is only one π\pi-point for these group schemes, this time represented by the map

k[t]/tpxpk[x]/xp2,k[t]/t^{p}\xrightarrow{x^{p}}k[x]/x^{p^{2}},

a subgroup inclusion making the point noble for all three group schemes. It is not hard to see that the products ~,1,2\widetilde{\otimes},\otimes_{1},\otimes_{2} all give the same Green ring, so that 𝔤{\mathfrak{g}} has Property PC, again making use of Lemma 3.1.9.

3.2 Abelian Lie algebras of dimension 2

Throughout this section we let 𝔤{\mathfrak{g}} be the abelian Lie algebra of dimension 2 with the trivial restriction 𝔤[p]=0,{\mathfrak{g}}^{[p]}=0, and A=u(𝔤)A=u({\mathfrak{g}}) the restricted enveloping algebra which we endow with coordinates

A=k[x,y]/(xp,yp)A=k[x,y]/(x^{p},y^{p})

and take Δ~\widetilde{\Delta} to be the Hopf algebra comultiplication making xx and yy as primitive, and ~\widetilde{\otimes} the corresponding tensor product of AA-modules.

For p>2p>2, AA is of wild representation type, and we will show that 𝔤{\mathfrak{g}} does not meet the hypothesis of Lemma 3.1.9, and in fact, 𝔤{\mathfrak{g}} does not satisfy Property PA. In the tame case p=2p=2, we show in Section 3.3 that Lemma 3.1.9 can be applied directly.

3.2.1.

The Hopf algebra structures on AA, up to equivalence, are classified by X. Wang in [XWang13], and the cocommutative structures are given as follows.

  1. 0.

    The Lie algebra k𝔾a(1)2k{\mathbb{G}}_{a(1)}^{2}

    Δ~:x\displaystyle\widetilde{\Delta}:x x1+1x\displaystyle\mapsto x\otimes 1+1\otimes x
    y\displaystyle y y1+1y,\displaystyle\mapsto y\otimes 1+1\otimes y,
  2. 1.

    The quasi-elementary group algebra k𝔾a(2)k{\mathbb{G}}_{a(2)}

    Δ1:x\displaystyle\Delta_{1}:x x1+1x\displaystyle\mapsto x\otimes 1+1\otimes x
    y\displaystyle y y1+1y+ω(x),\displaystyle\mapsto y\otimes 1+1\otimes y+\omega(x),
  3. 2.

    The group-Lie product k(𝔾a(1)×/p)k\left({\mathbb{G}}_{a(1)}\times{\mathbb{Z}}/p\right)

    Δ2:x\displaystyle\Delta_{2}:x x1+1x\displaystyle\mapsto x\otimes 1+1\otimes x
    y\displaystyle y y1+1y+yy,\displaystyle\mapsto y\otimes 1+1\otimes y+y\otimes y,
  4. 3.

    The discrete group algebra k(/p)2k({\mathbb{Z}}/p)^{2}

    Δ3:x\displaystyle\Delta_{3}:x x1+1x+xx\displaystyle\mapsto x\otimes 1+1\otimes x+x\otimes x
    y\displaystyle y y1+1y+yy.\displaystyle\mapsto y\otimes 1+1\otimes y+y\otimes y.

For Δ1\Delta_{1} we have used the notation

ω(x)=(x1+1x)p(xp1+1xp)p,\omega(x)=\frac{(x\otimes 1+1\otimes x)^{p}-(x^{p}\otimes 1+1\otimes x^{p})}{p},

a formal division of binomial coefficients by pp.

3.2.2.

We calculate the spectrum 𝒳(𝔾a(1)2)=ProjH(𝔾a(1)2,k){\mathscr{X}}({\mathbb{G}}_{a(1)}^{2})=\operatorname{Proj}H^{*}({\mathbb{G}}_{a(1)}^{2},k), applying Lemma 2.2.1, to be 1,{\mathbb{P}}^{1}, since each linear subspace of 𝔤K{\mathfrak{g}}_{K} over an extension of fields K/kK/k is a Lie subalgebra with trivial restriction. Each π\pi-point αK\alpha_{K} is of the form

K[t]/tpK[x,y]/(xp,yp)K[t]/t^{p}\to K[x,y]/(x^{p},y^{p})

with tax+by+ξ,t\mapsto ax+by+\xi, for a,bKa,b\in K not both 0, where ξ\xi is a polynomial in the ideal (x2,xy,y2),(x^{2},xy,y^{2}), i.e. a higher order term. From Friedlander and Pevtsova [FrPev07], if we let βL\beta_{L} be another π\pi-point with tax+by+ξt\mapsto a^{\prime}x+b^{\prime}y+\xi^{\prime} in the same form over LL, then αKβL\alpha_{K}\sim\beta_{L} if and only if there is a common extension FF of KK and LL such that [a:b]=[a:b][a:b]=[a^{\prime}:b^{\prime}] as FF-points of the projective scheme 1.{\mathbb{P}}^{1}.

The group schemes corresponding to the four cocommutative Hopf algebras listed in 3.2.1 are as we have claimed in notation:

𝔾a(1)2,𝔾a(2),𝔾a(1)×/p,(/p)2.{\mathbb{G}}_{a(1)}^{2},\quad{\mathbb{G}}_{a(2)},\quad{\mathbb{G}}_{a(1)}\times{\mathbb{Z}}/p,\quad({\mathbb{Z}}/p)^{2}.

We already know that each π\pi-point is noble for the Lie algebra, corresponding to 𝔾a(1)2{\mathbb{G}}_{a(1)}^{2}. The noble points for the three Group schemes representing the remaining points of /Aut(A){\mathscr{H}}/\operatorname{Aut}(A) are calculated below. One checks that each of these group schemes has finitely many subgroup schemes and that base changing to any field extension does not change the number of subgroup schemes.

  1. 1.

    The quasi-elementary group scheme 𝔾a(2){\mathbb{G}}_{a(2)} has only one nontrivial proper subgroup, and it is isomorphic 𝔾a(1){\mathbb{G}}_{a(1)}. The inclusion of 𝔾a(1){\mathbb{G}}_{a(1)} gives a noble π\pi-point k[t]/tpAk[t]/t^{p}\to A which maps tx.t\mapsto x. Therefore [1:0]1[1:0]\in{\mathbb{P}}^{1} is the only noble point for 𝔾a(2){\mathbb{G}}_{a(2)}.

  2. 2.

    The group schemes 𝔾a(1){\mathbb{G}}_{a(1)} and /p{\mathbb{Z}}/p are disjunct in the sense that any subgroup of the product 𝔾a(1)×/p{\mathbb{G}}_{a(1)}\times{\mathbb{Z}}/p is the natural inclusion of a product H1×H2H_{1}\times H_{2} for H1𝔾a(1)H_{1}\leq{\mathbb{G}}_{a(1)} and H2/pH_{2}\leq{\mathbb{Z}}/p. Therefore there are two inclusions 𝔾a(1)×0<𝔾a(1)×/p{\mathbb{G}}_{a(1)}\times 0<{\mathbb{G}}_{a(1)}\times{\mathbb{Z}}/p and 0×/p<𝔾a(1)×/p0\times{\mathbb{Z}}/p<{\mathbb{G}}_{a(1)}\times{\mathbb{Z}}/p, which give noble π\pi-points k[t]/tpAk[t]/t^{p}\to A, mapping txt\mapsto x and tyt\mapsto y respectively. Therefore [1:0][1:0] and [0:1][0:1] are the only noble points for the product 𝔾a(1)×/p{\mathbb{G}}_{a(1)}\times{\mathbb{Z}}/p.

  3. 3.

    The discrete group scheme (/p)2({\mathbb{Z}}/p)^{2} is a 2-dimensional vector space over the prime field /p{\mathbb{Z}}/p. Therefore the nontrivial proper subgroups are all isomorphic to 1-dimensional subspaces, i.e. the cyclic subgroups. The inclusion of the cyclic subgroup generated by (i,j)(/p)2(i,j)\in({\mathbb{Z}}/p)^{2} gives a noble π\pi-point k[t]/tpAk[t]/t^{p}\to A mapping t(x+1)i(y+1)j1.t\mapsto(x+1)^{i}(y+1)^{j}-1. Computing the linear term then tells us that the of the noble points for the discrete group are precisely those in the form [i:j]1[i:j]\in{\mathbb{P}}^{1} for i,j/p,i,j\in{\mathbb{Z}}/p, and up to equivalence there are p+1p+1 of them.

Lemma 3.2.3.

Let I={Δ~,Δ1,Δ2,Δ3}=/Aut(A)I=\{\widetilde{\Delta},\Delta_{1},\Delta_{2},\Delta_{3}\}={\mathscr{H}}/\operatorname{Aut}(A), with corresponding schemes GiG_{i} and tensor products i\otimes_{i} for iIi\in I. Then for any GiG_{i}, and any ignoble point 𝔭{\mathfrak{p}} for GiG_{i}, there exists V,W𝒞(𝔭)V,W\in{\mathcal{C}}({\mathfrak{p}}) such that ViWV\otimes_{i}W is not isomorphic to V~W.V\widetilde{\otimes}W.

Proof.

Given 𝔭=[a:b]1,{\mathfrak{p}}=[a:b]\in{\mathbb{P}}^{1}, we let α(𝔭)\alpha({\mathfrak{p}}) be the canonical π\pi-point, a map

k[t]/tpA=k[x,y]/(xp,yp)k[t]/t^{p}\to A=k[x,y]/(x^{p},y^{p})

which takes tax+by.t\mapsto ax+by. We generate a module with support {𝔭}\{{\mathfrak{p}}\} by inducing up the trivial k[t]/tpk[t]/t^{p} module kk up to AA, i.e. take

V(𝔭)=Ak[t]/tpk,V({\mathfrak{p}})=A\otimes_{k[t]/t^{p}}k,

where AA is a k[t]/tpk[t]/t^{p} algebra via α(𝔭).\alpha({\mathfrak{p}}). Explicitly the AA-module structure for V(𝔭)V({\mathfrak{p}}) is given as a pp-dimensional space over kk such that s2=ax+bys_{2}=ax+by acts as the 0-matrix and, for s1=cx+dys_{1}=cx+dy with {s1,s2}\{s_{1},s_{2}\} linearly independent, s1s_{1} acts by the nilpotent p×pp\times p Jordan block

Jp=(010000100000100000),J_{p}=\begin{pmatrix}0&1&0&\dots&0\\ 0&0&1&\dots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&0&1\\ 0&0&0&0&0\end{pmatrix},

written in a fixed ordered basis s1p1v,,s1v,v,s_{1}^{p-1}v,\dots,s_{1}v,v, where v=11Ak[t]/tpk.v=1\otimes 1\in A\otimes_{k[t]/t^{p}}k. Then 𝔭𝒳(Gi,V(𝔭)){\mathfrak{p}}\in{\mathscr{X}}(G_{i},V({\mathfrak{p}})), since tt annihilates the restricted module α(𝔭)(V(𝔭))\alpha({\mathfrak{p}})^{*}(V({\mathfrak{p}})). If 𝔮=[c:d]1{\mathfrak{q}}=[c:d]\in{\mathbb{P}}^{1} is a closed point not equivalent to 𝔭{\mathfrak{p}}, then the restriction α(𝔮)(V(𝔭))\alpha({\mathfrak{q}})^{*}(V({\mathfrak{p}})) is free of rank 1, and hence 𝔮𝒳(Gi,V(𝔭)),{\mathfrak{q}}\not\in{\mathscr{X}}(G_{i},V({\mathfrak{p}})), and the generic point is not in 𝒳(Gi,V(𝔭)){\mathscr{X}}(G_{i},V({\mathfrak{p}})) either. Thus 𝒳(Gi,V(𝔭)){\mathscr{X}}(G_{i},V({\mathfrak{p}})) is the singleton {𝔭}.\{{\mathfrak{p}}\}.

Now we compute the products V(𝔭)iV(𝔭)V({\mathfrak{p}})\otimes_{i}V({\mathfrak{p}}). First, each canonical π\pi-point α=α(𝔮),𝔮1,\alpha=\alpha({\mathfrak{q}}),\,{\mathfrak{q}}\in{\mathbb{P}}^{1}, is the inclusion of a subgroup scheme of 𝔾a(1)2.{\mathbb{G}}_{a(1)}^{2}. Therefore for any AA-modules M,M,M,M^{\prime}, we have

α(M~M)α(M)~α(M),\alpha^{*}(M\widetilde{\otimes}M^{\prime})\cong\alpha^{*}(M)\,\widetilde{\otimes}\,\alpha^{*}(M^{\prime}),

where the right-hand ~\widetilde{\otimes} is the tensor product of 𝔾a(1){\mathbb{G}}_{a(1)} representations as in Example 3.1.10. Therefore V(𝔭)~V(𝔭)V({\mathfrak{p}})\widetilde{\otimes}V({\mathfrak{p}}) is annihilated by s2s_{2} and is free of rank pp when restricted along s1,s_{1}, so

V(𝔭)~V(𝔭)pV(𝔭).V({\mathfrak{p}})\widetilde{\otimes}V({\mathfrak{p}})\cong pV({\mathfrak{p}}).

By Proposition 2.2.1 there are no ignoble points for 𝔾a(1)2,{\mathbb{G}}_{a(1)}^{2}, so PB(Δ~){}^{\prime}(\widetilde{\Delta}) is vacuously true. To show PB(Δi){}^{\prime}(\Delta_{i}) for i=1,2,3,i=1,2,3, we show that V(𝔭)iV(𝔭)V({\mathfrak{p}})\otimes_{i}V({\mathfrak{p}}) is not annihilated by s2s_{2}, for each ignoble point 𝔭{\mathfrak{p}} for Gi,G_{i}, and is therefore not isomorphic to V(𝔭)~V(𝔭)V({\mathfrak{p}})\widetilde{\otimes}V({\mathfrak{p}}). Notice the point [1:0]1[1:0]\in{\mathbb{P}}^{1} is noble for each of G1,G2,G3.G_{1},G_{2},G_{3}. Therefore an ignoble point is always in the form [a:1][a:1] for ak,a\in k, so we may always assume s1=xs_{1}=x and s2=ax+y.s_{2}=ax+y.

The elements s21,1s2AAs_{2}\otimes 1,1\otimes s_{2}\in A\otimes A annihilate the kk-space V(𝔭)V(𝔭)V({\mathfrak{p}})\otimes V({\mathfrak{p}}). Therefore the action of s2s_{2} for the representation V(𝔭)iV(𝔭)V({\mathfrak{p}})\otimes_{i}V({\mathfrak{p}}) of GiG_{i}, given by Δi(s2),\Delta_{i}(s_{2}), is well defined as an element of AA/Σ,A\otimes A/\Sigma, where Σ\Sigma is the ideal (s21,1s2)AA.(s_{2}\otimes 1,1\otimes s_{2})\triangleleft A\otimes A. The elements of the quotient AA/ΣA\otimes A/\Sigma are cosets, which we write as z+Σz+\Sigma for zAA.z\in A\otimes A.

  1. 1.

    The ignoble points for G1=𝔾a(2)G_{1}={\mathbb{G}}_{a(2)} are 𝔭=[a:1]{\mathfrak{p}}=[a:1] for any ak.a\in k. We have

    Δ1:s2s21+1s2+ω(s1)ω(s1)+Σ.\Delta_{1}:s_{2}\mapsto s_{2}\otimes 1+1\otimes s_{2}+\omega(s_{1})\in\omega(s_{1})+\Sigma.

    We described ω(s1)\omega(s_{1}) in 3.2.1 using a formal division of binomial coefficients by pp. This makes ω(s1)\omega(s_{1}) a sum with coefficients on terms s1ns1ms_{1}^{n}\otimes s_{1}^{m} for n+m=pn+m=p, and n,m1.n,m\geq 1. Using the same basis elements s1pvs_{1}^{p-\ell}v from our description of the Jordan block JpJ_{p}, we see

    ω(s1)(s1p2vv)=(s1s1p1)(s1p2vv)=s1p1vs1p1v,\omega(s_{1})\cdot(s_{1}^{p-2}v\otimes v)=(s_{1}\otimes s_{1}^{p-1})\cdot(s_{1}^{p-2}v\otimes v)=s_{1}^{p-1}v\otimes s_{1}^{p-1}v,

    so ω(s1)\omega(s_{1}) does not annihilate V(𝔭)kV(𝔭).V({\mathfrak{p}})\otimes_{k}V({\mathfrak{p}}).

  2. 2.

    The ignoble points for G2=𝔾a(1)×/pG_{2}={\mathbb{G}}_{a(1)}\times{\mathbb{Z}}/p are [a:1][a:1] for a0a\neq 0. We have

    Δ2:s2s21+1s2+(s2as1)(s2as1)a2s1s1+Σ.\Delta_{2}:s_{2}\mapsto s_{2}\otimes 1+1\otimes s_{2}+(s_{2}-as_{1})\otimes(s_{2}-as_{1})\in a^{2}s_{1}\otimes s_{1}+\Sigma.

    Using the same s1pvs_{1}^{p-\ell}v we see

    (a2s1s1)(vv)=a2s1vs1v,(a^{2}s_{1}\otimes s_{1})\cdot(v\otimes v)=a^{2}s_{1}v\otimes s_{1}v,

    and in particular, supposing a0,a\neq 0, we have a2s1s1a^{2}s_{1}\otimes s_{1} does not annihilate V(𝔭)kV(𝔭).V({\mathfrak{p}})\otimes_{k}V({\mathfrak{p}}).

  3. 3.

    The ignoble points for G3=(/p)2G_{3}=({\mathbb{Z}}/p)^{2} are [a:1][a:1] for apa0.a^{p}-a\neq 0. We have

    Δ3:s2s21+1s2+as1s1\displaystyle\Delta_{3}:s_{2}\mapsto s_{2}\otimes 1+1\otimes s_{2}+as_{1}\otimes s_{1} +(s2as1)(s2as1)\displaystyle+(s_{2}-as_{1})\otimes(s_{2}-as_{1})
    (a+a2)s1s1+Σ.\displaystyle\in(a+a^{2})s_{1}\otimes s_{1}+\Sigma.

    The same considerations as for G2G_{2} above shows that, supposing a+a20,a+a^{2}\neq 0, we have (a+a2)s1s1(a+a^{2})s_{1}\otimes s_{1} does not annihilate V(𝔭)kV(𝔭).V({\mathfrak{p}})\otimes_{k}V({\mathfrak{p}}). Note that apaa^{p}-a is divisible by a2+aa^{2}+a in k[a]k[a] in any characteristic p>0p>0.

We conclude that if 𝔭{\mathfrak{p}} is ignoble for GiG_{i}, then V(𝔭)iV(𝔭)V(𝔭)~V(𝔭)V({\mathfrak{p}})\otimes_{i}V({\mathfrak{p}})\not\cong V({\mathfrak{p}})\,\widetilde{\otimes}\,V({\mathfrak{p}}), for each i=G1,G2,G3/Aut(A),i=G_{1},G_{2},G_{3}\in{\mathscr{H}}/\operatorname{Aut}(A), since V(𝔭)~V(𝔭)V({\mathfrak{p}})\,\widetilde{\otimes}\,V({\mathfrak{p}}) is annihilated by s1s_{1} and V(𝔭)iV(𝔭)V({\mathfrak{p}})\otimes_{i}V({\mathfrak{p}}) is not. Thus PB(s){}^{\prime}(s) holds for s/Aut(A).s\in{\mathscr{H}}/\operatorname{Aut}(A).

Theorem 3.2.4.

(c.f. Theorem 1.1.3) If p>2,p>2, then 𝔤{\mathfrak{g}} does not satisfy Property PA.

Proof.

Consider the automorphism φ:AA\varphi:A\to A of algebras, with inverse φ1\varphi^{-1}

φ:x\displaystyle\varphi:x x\displaystyle\mapsto x φ1:x\displaystyle\varphi^{-1}:x x\displaystyle\mapsto x
y\displaystyle y y+x2,\displaystyle\mapsto y+x^{2}, y\displaystyle y yx2.\displaystyle\mapsto y-x^{2}.

Now fix 𝔭=[0:1]{\mathfrak{p}}=[0:1]. By examining linear terms of φ(ax+by)\varphi(ax+by) we see that φΩ(A).\varphi\in\Omega(A). For this reason MφM^{\varphi} has the same support 𝔭{\mathfrak{p}}. We compute directly that the module MφM^{\varphi} is determined by the p×pp\times p-matrix below, representing the action of a generic ax+byax+by

(0ab0000ab00000ab00000a000000)\begin{pmatrix}0&a&-b&0&\dots&0\\ 0&0&a&-b&\dots&0\\ \vdots&\vdots&\vdots&\ddots&\ddots&\vdots\\ 0&0&0&0&a&-b\\ 0&0&0&0&0&a\\ 0&0&0&0&0&0\end{pmatrix}

Now letting α(𝔭)\alpha({\mathfrak{p}}) be the canonical π\pi-point with tyt\mapsto y, we see α(𝔭)(Mφ)\alpha({\mathfrak{p}})^{*}(M^{\varphi}) has a Jordan block of size N=p+12N=\frac{p+1}{2} (we have assumed pp is odd). In particular MφM^{\varphi} is not isomorphic to MM because it is not annihilated by yy, thus contradicting the hypothesis in Lemma 3.1.6. The Jordan block of size NN is also present in α(𝔭)(Mφ1).\alpha({\mathfrak{p}})^{*}(M^{\varphi^{-1}}). We shall see that M~MM\widetilde{\otimes}M is not isomorphic as an AA-module to M~φM,M\widetilde{\otimes}^{\varphi}M, despite 𝔭{\mathfrak{p}} being noble for both the group scheme GG asssociated to the Lie algebra Δ~\widetilde{\Delta}, and its twist GφG^{\varphi} associated to Δ~φ=(φφ)Δ~φ1\widetilde{\Delta}^{\varphi}=(\varphi\otimes\varphi)\circ\widetilde{\Delta}\circ\varphi^{-1}. This proves Theorem 1.1.3, as our claim shows that 𝔤{\mathfrak{g}} does not satisfy Property PA.

Our argument is as follows: if \otimes is induced from any Δ\Delta, and φ\otimes^{\varphi} induced from the twist Δφ\Delta^{\varphi}, then we have for AA-modules V,W,V,W, that there is equality of AA-modules

VφφWφ=(VW)φ.V^{\varphi}\otimes^{\varphi}W^{\varphi}=(V\otimes W)^{\varphi}.

Therefore M~MM\widetilde{\otimes}M is isomorphic to M~φM=(Mφ1~Mφ1)φM\widetilde{\otimes}^{\varphi}M=(M^{\varphi^{-1}}\widetilde{\otimes}M^{\varphi^{-1}})^{\varphi} if and only if (M~M)φ1(M\widetilde{\otimes}M)^{\varphi^{-1}} is isomorphic to Mφ1~Mφ1M^{\varphi^{-1}}\widetilde{\otimes}M^{\varphi^{-1}}. But we see that the latter is false by restricting along α(𝔭).\alpha({\mathfrak{p}}).

We know already that M~MM\widetilde{\otimes}M is a direct sum of pp copies of MM from the restriction property of ~\widetilde{\otimes} used in 3.2.3. In particular (M~M)φ1=pMφ1(M\widetilde{\otimes}M)^{\varphi^{-1}}=pM^{\varphi^{-1}} has no Jordan blocks of size pp when restricted along α(𝔭)\alpha({\mathfrak{p}}), the largest Jordan block is instead of size N=p+12.N=\frac{p+1}{2}.

But we also have the restriction property

α(𝔭)(Mφ1~Mφ1)α(𝔭)(Mφ1)~α(𝔭)(Mφ1),\alpha({\mathfrak{p}})^{*}(M^{\varphi^{-1}}\widetilde{\otimes}M^{\varphi^{-1}})\cong\alpha({\mathfrak{p}})^{*}(M^{\varphi^{-1}})\,\widetilde{\otimes}\,\alpha({\mathfrak{p}})^{*}(M^{\varphi^{-1}}),

where the right-hand ~\widetilde{\otimes} is of k[t]/tpk[t]/t^{p}-modules as in Example 3.1.10. The product ~\widetilde{\otimes} of k[t]/tpk[t]/t^{p}-modules is known to multiply a pair of Jordan blocks of size nn to a module containing a Jordan block of size max{2n1,p}\operatorname{max}\{2n-1,p\} (see e.g. Benson [Benson17]). We conclude α(𝔭)(Mφ1~Mφ1)\alpha({\mathfrak{p}})^{*}(M^{\varphi^{-1}}\widetilde{\otimes}M^{\varphi^{-1}}) has a Jordan block of size 2N1=p2N-1=p but α(𝔭)((M~M)φ1)\alpha({\mathfrak{p}})^{*}((M\widetilde{\otimes}M)^{\varphi^{-1}}) does not, and hence

M~MM~φM.M\widetilde{\otimes}M\not\cong M\widetilde{\otimes}^{\varphi}M.

3.3 A tame algebra

We continue the assumptions of Section 3.2, and we specialize to p=2,p=2, so that A=k[x,y]/(x2,y2)A=k[x,y]/(x^{2},y^{2}) is of tame representation type. Note that in X. Wang’s classification 3.2.1, we can now replace the term ω(t)=tt\omega(t)=t\otimes t for tAt\in A, per its definition.

The Hopf algebra Δ3\Delta_{3} of 3.2.1 corresponds to the discrete group G3=(/2)2G_{3}=({\mathbb{Z}}/2)^{2}. Finite indecomposable AA-modules were first classified by Bašev [Basev61], identifying A=k(/2)2A=k({\mathbb{Z}}/2)^{2}. The semirings relative to G3,3G_{3},\otimes_{3} for thick subcategories 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}), supported at noble points 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A) for G3G_{3}, were also successfully calculated by Bašev, and we will see that 𝔤{\mathfrak{g}} satisfies property PA. So for each V,WV,W supported only at a noble point for G3G_{3}, we will see that V3WV~WV\otimes_{3}W\cong V\widetilde{\otimes}W with the same methods as likely used by Bašev. For the 2-dimensional modules V(𝔭)=Ak[t]/t2kV({\mathfrak{p}})=A\otimes_{k[t]/t^{2}}k defined in 3.2.3, our computations of V(𝔭)3V(𝔭)V({\mathfrak{p}})\otimes_{3}V({\mathfrak{p}}) also agree with Bašev for 𝔭{\mathfrak{p}} both noble and ignoble points for G3G_{3}. Note that the complete semiring (𝒞(𝔭),3)({\mathcal{C}}({\mathfrak{p}}),\otimes_{3}) for the ignoble points 𝔭{\mathfrak{p}} for G3G_{3} was initially computed in error in [Basev61], and corrected first by Conlon in [Conlon65].

3.3.1.

Our argument for Theorem 1.1.2 is as follows. The algebra AA is local so there is up to isomorphism only one projective indecomposable we call PP, of dimension 44. For each closed point 𝔭𝒳(A,k)=1{\mathfrak{p}}\in{\mathscr{X}}(A,k)={\mathbb{P}}^{1}, the minimal thick subcategory 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) contains up to isomorphism only PP, and for each n=1,2,n=1,2,\dots, a single indecomposable module V2n(𝔭)V_{2n}({\mathfrak{p}}) of dimension 2n2n with support {𝔭}\{{\mathfrak{p}}\}. This is shown by Bašev [Basev61]. We will see for n=1n=1 that V2(𝔭)V_{2}({\mathfrak{p}}) agrees with our induced module V(𝔭)V({\mathfrak{p}}) defined in 3.2.3. Bašev’s classification shows for us that the algebra AA satisfies the isotropy hypothesis for applying the curtailment Lemma 3.1.9. By Lemma 3.2.3 we then know that 𝔤{\mathfrak{g}} satisfies Property PB. What remains is to repeat the methods of [Basev61] to conclude that

V2n(𝔭)V2m(𝔭)2V2min(n,m)+(nmmin(n,m))PV_{2n}({\mathfrak{p}})\otimes V_{2m}({\mathfrak{p}})\cong 2V_{2\operatorname{min}(n,m)}+(nm-\operatorname{min}(n,m))P (3.3.2)

for each point 𝔭{\mathfrak{p}} which is noble for the group scheme GG, having induced product =~,1,2,3\otimes=\widetilde{\otimes},\otimes_{1},\otimes_{2},\otimes_{3}, and each n,m.n,m. Then by the curtailment Lemma 3.1.9, as well as Propositon 2.2.1 and X. Wang’s classification 3.2.1, we know in particular that V2n(𝔭)V2m(𝔭)V2n(𝔭)~V2m(𝔭)V_{2n}({\mathfrak{p}})\otimes V_{2m}({\mathfrak{p}})\cong V_{2n}({\mathfrak{p}})\,\widetilde{\otimes}\,V_{2m}({\mathfrak{p}}) for each Hopf algebra structure of {\mathscr{H}}, with group scheme GG, product \otimes, and 𝔭{\mathfrak{p}} noble for GG. Since these are all possible pairs of nonprojective indecomposables, we conclude that 𝔤{\mathfrak{g}} satisfies Property PA and thus Property PC.

The next three results below extend the techniques of [Basev61] and reduce Bašev’s formula 3.3.2 to direct computation with matrices.

Proposition 3.3.3.

Let n,m>0n,m\in{\mathbb{Z}}_{>0} and 𝔭1.{\mathfrak{p}}\in{\mathbb{P}}^{1}. Then for =~,1,2,3,\otimes=\widetilde{\otimes},\otimes_{1},\otimes_{2},\otimes_{3}, we have

V2n(𝔭)V2m(𝔭)=V(nmmin(n,m))PV_{2n}({\mathfrak{p}})\otimes V_{2m}({\mathfrak{p}})=V\oplus(nm-\operatorname{min}(n,m))P

for some finite AA-module VV with no nontrivial projective submodule.

Lemma 3.3.4.

[Basev61] Let RR be a (nonunital) associative, commutative ring with a {\mathbb{Z}}-linear basis {xss=1,2,}\{x_{s}\mid s=1,2,\dots\} ; x0=0x_{0}=0. Assume that each product xmxnx_{m}x_{n} is a nonnegative integer combination of the xsx_{s} for s>0s>0, and further that x1x2=2x1x_{1}x_{2}=2x_{1}. If the {\mathbb{Z}}-linear functional f:Rf:R\to{\mathbb{Z}} defined by f(xs)=sf(x_{s})=s satisfies f(xsxt)=2min(s,t)f(x_{s}x_{t})=2\operatorname{min}(s,t), then xmxn=2xmin(m,n)x_{m}x_{n}=2x_{\operatorname{min}(m,n)} whenever mn.m\neq n.

Corollary 3.3.5.

Let n>m>0n>m\in{\mathbb{Z}}_{>0} and 𝔭1.{\mathfrak{p}}\in{\mathbb{P}}^{1}. Then for =~,1,2,3,\otimes=\widetilde{\otimes},\otimes_{1},\otimes_{2},\otimes_{3}, we have

V2n(𝔭)V2m(𝔭)=2V2m(𝔭)(nmm)P.V_{2n}({\mathfrak{p}})\otimes V_{2m}({\mathfrak{p}})=2V_{2m}({\mathfrak{p}})\oplus(nm-m)P.

The proof of Proposition 3.3.3 comes from the fact that P=AP=A is an injective module over AA, and has a linear basis 1,x,y,xy.1,x,y,xy. So the element xyxy annihilates any indecomposable module which is not free. It follows that for the module M=V2n(𝔭)V2m(𝔭)M=V_{2n}({\mathfrak{p}})\otimes V_{2m}({\mathfrak{p}}), the rank of the matrix in Endk(M)\operatorname{End}_{k}(M) representing xyAxy\in A is equal to the rank of the largest free submodule (a direct summand) of MM. Then since the ideal (xy1,1xy)AA(xy\otimes 1,1\otimes xy)\triangleleft A\otimes A is in the AAA\otimes A-annihilator for MM, we see all the matrices in Endk(M)\operatorname{End}_{k}(M) represented by the elements

Δ~(xy),Δ1(xy),Δ2(xy),Δ3(xy)AA\widetilde{\Delta}(xy),\quad\Delta_{1}(xy),\quad\Delta_{2}(xy),\quad\Delta_{3}(xy)\in A\otimes A

are the same as that of xy+yxAA.x\otimes y+y\otimes x\in A\otimes A. After properly defining the AA-modules V2n(𝔭)V_{2n}({\mathfrak{p}}) for n=1,2,,n=1,2,\dots, it is easily verified that the matrix representing xy+yxx\otimes y+y\otimes x indeed has rank (nmm)(nm-m). Lemma 3.3.4 is a tedious exercise in induction. Corollary 3.3.5 is immediate from taking RR to be the semiring generated by the indecomposables xs=V2sx_{s}=V_{2s} with products \otimes defined modulo PP, and applying Lemma 3.3.4.

All that is left to verify the formula 3.3.2 is the square in each dimension

V2n(𝔭)V2n(𝔭)=2V2n(𝔭)(n2n)PV_{2n}({\mathfrak{p}})\otimes V_{2n}({\mathfrak{p}})=2V_{2n}({\mathfrak{p}})\oplus(n^{2}-n)P (3.3.6)

when 𝔭{\mathfrak{p}} is noble for the group scheme GG having product =~,1,2,3.\otimes=\widetilde{\otimes},\otimes_{1},\otimes_{2},\otimes_{3}. For this, before defining the modules V2n(𝔭)V_{2n}({\mathfrak{p}}), we state two more tedious results pertaining to Lemma 3.3.4.

Proposition 3.3.7.

Let SS be any subset of the natural numbers such that no two consecutive numbers are elements of SS. Let R=xii=1R=\langle x_{i}\rangle_{i=1}^{\infty} be the free abelian group, denoting x0=0,x_{0}=0, with multiplication defined by

xsxt={2xmin(s,t)st2xss=tSxs1+xs+1s=tS,x_{s}x_{t}=\begin{cases}2x_{\operatorname{min}(s,t)}&s\neq t\\ 2x_{s}&s=t\not\in S\\ x_{s-1}+x_{s+1}&s=t\in S,\end{cases}

extended linearly. Then RR is an associative, commutative ring admitting a functional f:Rf:R\to{\mathbb{Z}} as in Lemma 3.3.4.

Lemma 3.3.8.

Let RR be a ring with functional f:Rf:R\to{\mathbb{Z}} as in Lemma 3.3.4. Then there is some subset SS of the positive integers such that RR is the ring defined in Proposition 3.3.7.

Now let us define the modules V2n(𝔭)V_{2n}({\mathfrak{p}}) in Bašev’s classification theorem below. We omit the classification of modules supported everywhere in 𝒳(A){\mathscr{X}}(A), which includes the structure of the syzygy modules Ωnk=Σnk\Omega^{n}k=\Sigma^{-n}k for nn\in{\mathbb{Z}}, and that these are all such everywhere-supported indecomposable modules up to isomorphism.

Theorem 3.3.9.

(Bašev, 1961, [Basev61]) Let 𝔭=[a:b]𝒳(A)=1{\mathfrak{p}}=[a:b]\in{\mathscr{X}}(A)={\mathbb{P}}^{1}, and define s2=ax+byAs_{2}=ax+by\in A and s1=cx+dys_{1}=cx+dy such that s1,s2s_{1},s_{2} forms a basis for the subspace x,yA.\langle x,y\rangle\subset A.

Let MM denote a vector space of dimension 2n2n, with kk-linear decomposition into lower and upper blocks M=MMuM=M_{\ell}\oplus M_{u}, with M,MuM_{\ell},M_{u} each of dimension nn. We define V2n(𝔭)=MV_{2n}({\mathfrak{p}})=M to be the AA-module defined by following matrix representations of the actions of s1,s2As_{1},s_{2}\in A

s1:(0In00),s2:(0𝔑n00),s_{1}:\begin{pmatrix}0&I_{n}\\ 0&0\end{pmatrix},\quad s_{2}:\begin{pmatrix}0&\mathfrak{N}_{n}\\ 0&0\end{pmatrix},

where InI_{n} is the diagonal ones matrix and 𝔑n\mathfrak{N}_{n} is an upper triangular nilpotent Jordan block of rank n1.n-1.

Then we have that

  1. I.

    The module V2n(𝔭)V_{2n}({\mathfrak{p}}) is, up to isomorphism, not dependent on choice of a,b,c,d,a,b,c,d, such that 𝔭=[a:b]1{\mathfrak{p}}=[a:b]\in{\mathbb{P}}^{1}, and such that s1=cx+dys_{1}=cx+dy and s2s_{2} are linearly independent,

  2. II.

    The module V2n(𝔭)V_{2n}({\mathfrak{p}}) is indecomposable,

  3. III.

    The support 𝒳(A,V2n(𝔭)){\mathscr{X}}(A,V_{2n}({\mathfrak{p}})) is {𝔭}1,\{{\mathfrak{p}}\}\subset{\mathbb{P}}^{1}, and

  4. IV.

    Any finite indecomposable module VV with support 𝒳(A,V)={𝔭}{\mathscr{X}}(A,V)=\{{\mathfrak{p}}\} is of even dimension 2n2n and is isomorphic to V2n(𝔭)V_{2n}({\mathfrak{p}}), for some nn.

The following Lemma is proven with the exact same method used for computing V(𝔭)~V(𝔭)2V(𝔭)V({\mathfrak{p}})\widetilde{\otimes}V({\mathfrak{p}})\cong 2V({\mathfrak{p}}) in 3.2.3, after noting V(𝔭)=V2(𝔭)V({\mathfrak{p}})=V_{2}({\mathfrak{p}}) and applying the curtailment Lemma 3.1.9.

Lemma 3.3.10.

Let 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A) be a noble π\pi-point for a group scheme GG corresponding to a Hopf algebra structure Δ\Delta\in{\mathscr{H}} on AA, with \otimes the product of GG-representations. Then V2(𝔭)V2(𝔭)2V2(𝔭).V_{2}({\mathfrak{p}})\otimes V_{2}({\mathfrak{p}})\cong 2V_{2}({\mathfrak{p}}).

Now we prove the formula 3.3.6, giving an original proof of the computations of tensor products, first described in the case G=G3G=G_{3} from 3.2.1, by Bašev [Basev61] and Conlon [Conlon65] without proof.

Theorem 3.3.11.

Let GG be the group scheme associated to one of the Hopf algebras /Aut(A)={Δ~,Δ1,Δ2,Δ3}{\mathscr{H}}/\operatorname{Aut}(A)=\{\widetilde{\Delta},\Delta_{1},\Delta_{2},\Delta_{3}\} from X. Wang’s classification 3.2.1. Let \otimes be the associated tensor product of representations of GG. Then for each noble point 𝔭{\mathfrak{p}} for GG and each nn, we have an isomorphism

V2n(𝔭)V2n(𝔭)2V2n(𝔭)(n2n)PV_{2n}({\mathfrak{p}})\otimes V_{2n}({\mathfrak{p}})\cong 2V_{2n}({\mathfrak{p}})\oplus(n^{2}-n)P
Proof.

We will make explicit computations using matrices in each case of 3.2.1 for representatives Gi/Aut(A)G_{i}\in{\mathscr{H}}/\operatorname{Aut}(A), and promote this to a general formula of GG with our curtailment Lemma 3.1.9. Throughout, we fix a noble π\pi-point 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A) for GG and denote V2n=V2n(𝔭)V_{2n}=V_{2n}({\mathfrak{p}}).

Given the basic matrix construction 3.3.9 of the indecomposable reps V2nV_{2n}, the choice of basis induces short exact sequences of modules we call the canonical monos/epis

0V2pV2(p+q)V2q0.0\to V_{2p}\to V_{2(p+q)}\to V_{2q}\to 0.

Suppose for n>1n>1 the contrary to formula 3.3.6, which by Lemma 3.3.8 lets us assume

V2nV2nV2(n1)V2(n+1)(n2n)P.V_{2n}\otimes V_{2n}\cong V_{2(n-1)}\oplus V_{2(n+1)}\oplus(n^{2}-n)P.

But also by Lemma 3.3.8 we have

V2(n+1)V2(n+1)2V2(n+1)(n2+n)P.V_{2(n+1)}\otimes V_{2(n+1)}\cong 2V_{2(n+1)}\oplus(n^{2}+n)P.

Consider the canonical mono ι:V2nV2(n+1).\iota:V_{2n}\to V_{2(n+1)}. Then its \otimes-square

ιι:V2nV2nV2(n+1)V2(n+1),\iota\otimes\iota:V_{2n}\otimes V_{2n}\to V_{2(n+1)}\otimes V_{2(n+1)},

is also a monomorphism, and by the injectivity of PP, this descends to

ι:V2(n1)V2(n+1)2V2(n+1)(2n)P.\iota^{\prime}:V_{2(n-1)}\oplus V_{2(n+1)}\to 2V_{2(n+1)}\oplus(2n)P.

We will show directly that the restriction ι\iota^{\prime} of ιι\iota\otimes\iota indeed has image contained in the nonprojective component of V2(n+1)V2(n+1)V_{2(n+1)}\otimes V_{2(n+1)}, and argue that ι\iota^{\prime} is a sum of canonical monos composed with the split embedding. From this it will follow that the cokernel of ιι\iota\otimes\iota is isomorphic to V4(2n)PV_{4}\oplus(2n)P. But this is a contradiction: by the restriction formula (see 1.2) there is no exact sequence

0V2nV2nV2(n+1)V2(n+1)V4(2n)P0,0\to V_{2n}\otimes V_{2n}\to V_{2(n+1)}\otimes V_{2(n+1)}\to V_{4}\oplus(2n)P\to 0,

for its restriction to the subgroup representing 𝔭{\mathfrak{p}} is an exact sequence of k[t]/t2k[t]/t^{2}-modules (see 3.1.10)

0{0}4J14(n1)J2{{4J_{1}\oplus 4(n-1)J_{2}}}4J14(n+1)J2{{4J_{1}\oplus 4(n+1)J_{2}}}2J1(4n+1)J2{{2J_{1}\oplus(4n+1)J_{2}}}0{0}

This is a contradiction when n>1n>1 since J2J_{2} is projective and injective, and 4(n+1)<(4n+1)+4(n1).4(n+1)<(4n+1)+4(n-1).

For direct computation we consider two cases, one in which 𝔭=[1:0]{\mathfrak{p}}=[1:0], hence we may assume s2=x,s_{2}=x, and s1=ys_{1}=y, and the other, in which 𝔭=[a:1]{\mathfrak{p}}=[a:1] for a=0,1ka=0,1\in k we take s2=ax+ys_{2}=ax+y and s1=x.s_{1}=x.

Suppose 𝔭=[1:0]{\mathfrak{p}}=[1:0], s2=x,s_{2}=x, s1=y.s_{1}=y. Now we have

Δ~:s1\displaystyle\widetilde{\Delta}:s_{1} s11+1s1\displaystyle\mapsto s_{1}\otimes 1+1\otimes s_{1}
s2\displaystyle s_{2} s21+1s2,\displaystyle\mapsto s_{2}\otimes 1+1\otimes s_{2},
Δ1:s1\displaystyle\Delta_{1}:s_{1} s11+1s1+s2s2\displaystyle\mapsto s_{1}\otimes 1+1\otimes s_{1}+s_{2}\otimes s_{2}
s2\displaystyle s_{2} s21+1s2,\displaystyle\mapsto s_{2}\otimes 1+1\otimes s_{2},
Δ2:s1\displaystyle\Delta_{2}:s_{1} s11+1s1+s1s1\displaystyle\mapsto s_{1}\otimes 1+1\otimes s_{1}+s_{1}\otimes s_{1}
s2\displaystyle s_{2} s21+1s2,\displaystyle\mapsto s_{2}\otimes 1+1\otimes s_{2},
Δ3:s1\displaystyle\Delta_{3}:s_{1} s11+1s1+s1s1\displaystyle\mapsto s_{1}\otimes 1+1\otimes s_{1}+s_{1}\otimes s_{1}
s2\displaystyle s_{2} s21+1s2+s2s2.\displaystyle\mapsto s_{2}\otimes 1+1\otimes s_{2}+s_{2}\otimes s_{2}.

Now we describe the action of the Δ(si)AA\Delta(s_{i})\in A\otimes A over the kk-linear decomposition

MM=(MM)(MMu)(MuM)(MuMu)M\otimes M=(M_{\ell}\otimes M_{\ell})\oplus(M_{\ell}\otimes M_{u})\oplus(M_{u}\otimes M_{\ell})\oplus(M_{u}\otimes M_{u})

into blocks of size n2,n^{2}, where V2n=M=MMuV_{2n}=M=M_{\ell}\oplus M_{u} as in Theorem 3.3.9. There are four actions to check:

  1. 0.

    Δ=Δ~\Delta=\widetilde{\Delta}

    Δ(s1):(01InIn10000In10001In0000),Δ(s2):(01𝔑n𝔑n10000𝔑n10001𝔑n0000),\Delta(s_{1}):\begin{pmatrix}0&1\otimes I_{n}&I_{n}\otimes 1&0\\ 0&0&0&I_{n}\otimes 1\\ 0&0&0&1\otimes I_{n}\\ 0&0&0&0\end{pmatrix},\quad\Delta(s_{2}):\begin{pmatrix}0&1\otimes\mathfrak{N}_{n}&\mathfrak{N}_{n}\otimes 1&0\\ 0&0&0&\mathfrak{N}_{n}\otimes 1\\ 0&0&0&1\otimes\mathfrak{N}_{n}\\ 0&0&0&0\end{pmatrix},
  2. 1.

    Δ=Δ1\Delta=\Delta_{1}

    Δ(s1):(01InIn1𝔑n𝔑n000In10001In0000),Δ(s2):(01𝔑n𝔑n10000𝔑n10001𝔑n0000),\Delta(s_{1}):\begin{pmatrix}0&1\otimes I_{n}&I_{n}\otimes 1&\mathfrak{N}_{n}\otimes\mathfrak{N}_{n}\\ 0&0&0&I_{n}\otimes 1\\ 0&0&0&1\otimes I_{n}\\ 0&0&0&0\end{pmatrix},\quad\Delta(s_{2}):\begin{pmatrix}0&1\otimes\mathfrak{N}_{n}&\mathfrak{N}_{n}\otimes 1&0\\ 0&0&0&\mathfrak{N}_{n}\otimes 1\\ 0&0&0&1\otimes\mathfrak{N}_{n}\\ 0&0&0&0\end{pmatrix},
  3. 2.

    Δ=Δ2\Delta=\Delta_{2}

    Δ(s1):(01InIn1InIn000In10001In0000),Δ(s2):(01𝔑n𝔑n10000𝔑n10001𝔑n0000),\Delta(s_{1}):\begin{pmatrix}0&1\otimes I_{n}&I_{n}\otimes 1&I_{n}\otimes I_{n}\\ 0&0&0&I_{n}\otimes 1\\ 0&0&0&1\otimes I_{n}\\ 0&0&0&0\end{pmatrix},\ \Delta(s_{2}):\begin{pmatrix}0&1\otimes\mathfrak{N}_{n}&\mathfrak{N}_{n}\otimes 1&0\\ 0&0&0&\mathfrak{N}_{n}\otimes 1\\ 0&0&0&1\otimes\mathfrak{N}_{n}\\ 0&0&0&0\end{pmatrix},
  4. 3.

    Δ=Δ3\Delta=\Delta_{3}

    Δ(s1):(01InIn1InIn000In10001In0000),Δ(s2):(01𝔑n𝔑n1𝔑n𝔑n000𝔑n10001𝔑n0000),\Delta(s_{1}):\begin{pmatrix}0&1\otimes I_{n}&I_{n}\otimes 1&I_{n}\otimes I_{n}\\ 0&0&0&I_{n}\otimes 1\\ 0&0&0&1\otimes I_{n}\\ 0&0&0&0\end{pmatrix},\ \Delta(s_{2}):\begin{pmatrix}0&1\otimes\mathfrak{N}_{n}&\mathfrak{N}_{n}\otimes 1&\mathfrak{N}_{n}\otimes\mathfrak{N}_{n}\\ 0&0&0&\mathfrak{N}_{n}\otimes 1\\ 0&0&0&1\otimes\mathfrak{N}_{n}\\ 0&0&0&0\end{pmatrix},

with each block entry of the matrices above representing an n2×n2n^{2}\times n^{2} matrix.

The projective component of MMM\otimes M is actually a subspace which is independent of choice of Δ,\Delta, and free of rank n2n.n^{2}-n. To see this, as in the discussion following Corollary 3.3.5, we see for each choice Δ=Δ~,Δ1,Δ2,Δ3,\Delta=\widetilde{\Delta},\Delta_{1},\Delta_{2},\Delta_{3}, that Δ(s1s2)\Delta(s_{1}s_{2}) has the same matrix representation as that of s1s2+s2s1,s_{1}\otimes s_{2}+s_{2}\otimes s_{1}, given by the block matrix

(000In𝔑n+𝔑nIn000000000000).\begin{pmatrix}0&0&0&I_{n}\otimes{\mathfrak{N}}_{n}+{\mathfrak{N}}_{n}\otimes I_{n}\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&0\end{pmatrix}.

We can check that this matrix has rank n2nn^{2}-n as follows: writing the linear map

In𝔑n+𝔑nInHomk(MuMu,MM)I_{n}\otimes{\mathfrak{N}}_{n}+{\mathfrak{N}}_{n}\otimes I_{n}\in\operatorname{Hom}_{k}(M_{u}\otimes M_{u},M_{\ell}\otimes M_{\ell})

with respect to the tensor basis, we get the n2×n2n^{2}\times n^{2} block matrix, with each entry an n×nn\times n matrix, as below

(𝔑nIn000𝔑nIn0000𝔑nIn0000𝔑n).\begin{pmatrix}{\mathfrak{N}}_{n}&I_{n}&0&\dots&0\\ 0&{\mathfrak{N}}_{n}&I_{n}&\dots&0\\ \vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&0&{\mathfrak{N}}_{n}&I_{n}\\ 0&0&0&0&{\mathfrak{N}}_{n}\end{pmatrix}.

Then of the columns numbered 0,,n21,0,\dots,n^{2}-1, the iith column is a pivot column if and only if ii is not divisible by nn (i.e. it is not the leftmost column of its respective block). The basis elements 𝔅{\mathfrak{B}} of MuMuM_{u}\otimes M_{u} corresponding to the pivot columns identified in turn generate a free submodule F𝔅MMF^{{\mathfrak{B}}}\subset M\otimes M of rank n2nn^{2}-n, with total dimension 4(n2n)4(n^{2}-n) over kk.

Now, the canonical mono ι:V2nV2(n+1)\iota:V_{2n}\to V_{2(n+1)} is induced from an inclusion of basis elements, between upper blocks V2n(u)V2(n+1)(u)V_{2n(u)}\to V_{2(n+1)(u)} and lower blocks V2n()V2(n+1)().V_{2n(\ell)}\to V_{2(n+1)(\ell).} Our assertion now follows, that ιι:V2nV2nV2(n+1)V2(n+1)\iota\otimes\iota:V_{2n}\otimes V_{2n}\to V_{2(n+1)}\otimes V_{2(n+1)} is the direct sum of monomorphisms between projective components and between nonprojective components, for each =~,1,2,3\otimes=\widetilde{\otimes},\otimes_{1},\otimes_{2},\otimes_{3}.

Further, the basic module structure of the indecomposables V2nV_{2n} in Theorem 3.3.9 tells us any embedding V2nVV_{2n}\to V, for VV a finite module with 𝒳(A,V)={𝔭}{\mathscr{X}}(A,V)=\{{\mathfrak{p}}\}, and such that VV is annihilated by s1s2s_{1}s_{2}, is a canonical mono into one indecomposable factor of VV. Thus the assertion follows that ι:V2(n1)V2(n+1)2V2(n+1)2V2(n+1)(2n)P\iota^{\prime}:V_{2(n-1)}\oplus V_{2(n+1)}\to 2V_{2(n+1)}\to 2V_{2(n+1)}\oplus(2n)P is a direct sum of canonical monos. Hence, a contradiction, as explained above.

We will be able to make the same contradiction by observing the same projective components in the case of 𝔭=[a:1]{\mathfrak{p}}=[a:1] for a=0,1ka=0,1\in k, but we omit the matrices that let us see this directly. ∎

Theorem 3.3.12.

(c.f. Theorem 1.1.2) The restricted Lie algebra 𝔤{\mathfrak{g}} satisfies Property PC.

Proof.

From Theorem 3.3.9 and Lemma 3.1.7, our curtailment Lemma 3.1.9 applies. Then with Lemma 3.2.3, we have that 𝔤{\mathfrak{g}} satisfies Property PB.

We know from Lemma 3.3.10, and the calculation of projective components in Proposition 3.3.3, that we may apply 3.3.4 to calculate that each non-square product of indecomposables in 𝒞(𝔭){\mathcal{C}}({\mathfrak{p}}) agrees with Bašev’s formula 3.3.2 regardless of GiG_{i} chosen from 3.2.1, and regardless of nobility of 𝔭{\mathfrak{p}} for GiG_{i}. Finally Theorem 3.3.11 tells us that the remaining products of indecomposables, the squares of those supported at a noble point, all follow the same formula 3.3.2. Then by the curtailment Lemma 3.1.9, we have that 𝔤{\mathfrak{g}} satisfies property PA. ∎

3.4 Conjectures for Lie algebras

Consider the Lie algebra 𝔰𝔩2\mathfrak{sl}_{2}, with presentation

e,f,h[e,f]=h,[h,e]=2e,[h,f]=2f.\langle e,f,h\mid[e,f]=h,[h,e]=2e,[h,f]=2f\rangle.

In characteristic p=2p=2, any restriction 𝔤{\mathfrak{g}} of 𝔰𝔩2\mathfrak{sl}_{2} which includes relations

e[2]=f[2]=0e^{[2]}=f^{[2]}=0

has in turn that u(𝔤)u({\mathfrak{g}}) is a quotient of the tame noncommutative algebra B=kx,y/(x2,y2).B=k\langle x,y\rangle/(x^{2},y^{2}). The modules of BB are classified by Bondarenko [Bon75], originally applied to determine how dihedral groups are tame. In turn the work applies directly to the case of the restriction 𝔤1{\mathfrak{g}}_{1} of 𝔰𝔩2\mathfrak{sl}_{2} which has h[2]=0h^{[2]}=0 (this restricted Lie algebra coincides with the canonical restriction for the Heisenberg algebra). This is because the dihedral group of order 88 has group algebra isomorphic as an associative algebra to u(𝔤1)u({\mathfrak{g}}_{1}). Thus, A=u(𝔤1)A=u({\mathfrak{g}}_{1}) has been shown so far to have two different cocommutative Hopf algebra structures, but it is an open problem to classify all of the cocommutative Hopf algebra structures on AA. There are at least ten, including those of the Lie algebra and dihedral group, each specializing under h0h\mapsto 0 to a Hopf algebra on the commutative algebra k[e,f]/(e2,f2)k[e,f]/(e^{2},f^{2}) covered in Sections 3.2 and 3.3. Similar local algebras of order p3p^{3} have Hopf algebra structures classified as a corollary of the work of Nguyen, L. Wang, and X. Wang [NWW15] for p>2.p>2.

The other restriction 𝔤2{\mathfrak{g}}_{2} on 𝔰𝔩2\mathfrak{sl}_{2} with u(𝔤2)u({\mathfrak{g}}_{2}) a quotient of B,B, is the canonical restriction derived from the trace-free 2×22\times 2 matrix representation of 𝔰𝔩2\mathfrak{sl}_{2}, which has that h[2]=h.h^{[2]}=h. We present three conjectures, each encompassed by the next.

Conjecture 3.4.1.

The restrictions 𝔤1{\mathfrak{g}}_{1} and 𝔤2{\mathfrak{g}}_{2} of the Lie algebra 𝔰𝔩2\mathfrak{sl}_{2} satisfy Property PC.

Conjecture 3.4.2.

Let 𝔤{\mathfrak{g}} be a restricted Lie algebra of tame representation type over kk. Then 𝔤{\mathfrak{g}} satisfies Property PC.

Conjecture 3.4.3.

Let 𝔤{\mathfrak{g}} be any restricted Lie algebra over kk. Then 𝔤{\mathfrak{g}} satisfies Property PC if and only if 𝔤{\mathfrak{g}} is of tame or finite representation type.

A fourth conjecture is also likely to hold, but we also suspect a proof would require classification of Hopf algebras far beyond what is known.

Conjecture 3.4.4.

All restricted Lie algebras satisfy Property PB regardless of representation type.

The first point to address the ‘if’ direction of Conjecture 3.4.3 is to show whether tame and finite representation type is equivalent to the isotropy hypothesis of Lemma 3.1.9. For converse, in the wild case, we can continue in Section 4 to give ad-hoc arguments for how the failure of the isotropy hypothesis leads to a failure of Property PA, as per our technique in 3.2.4, which proved Theorem 1.1.3. Tame restricted Lie algebras of dimension 3\leq 3 for odd characteristic may also be studied making direct use of Nguyen, L. Wang, and X. Wang’s classification [NWW15], towards Conjecture 3.4.2.

4 Lie algebras of wild representation type

In this section we will move toward one direction of Conjecture 3.4.3, that no Lie algebra of wild representation type satisfies Property PA.

We begin by adapting our proof of Theorem 3.2.4 into a more general situation. Recall the setting for disproving Property PA for a given restricted Lie algebra 𝔤{\mathfrak{g}}: we wish to produce a Hopf algebra structure Δ\Delta on the enveloping algebra A=u(𝔤)A=u({\mathfrak{g}}), differing from the Lie structure Δ~\widetilde{\Delta}, corresponding to a tensor structure \otimes on AA-modules differing from the Lie structure ~\widetilde{\otimes}. Then we produce AA-modules V,WV,W, with the support condition 𝒳(A,V)=𝒳(A,W)={𝔭},{\mathscr{X}}(A,V)={\mathscr{X}}(A,W)=\{{\mathfrak{p}}\}, where 𝔭{\mathfrak{p}} is noble for the group scheme GG corresponding to Δ\Delta, but there is non-isomorphism

VWV~W.V\otimes W\not\cong V\widetilde{\otimes}W.

Our technique proving Theorem 3.2.4 was to leverage the action of Aut(A)\operatorname{Aut}(A) on the space of Hopf algebra structures {\mathscr{H}}, on AA-modules, and on the variety 𝒳(A){\mathscr{X}}(A), with the natural isomorphism

(VW)φVφφWφ.(V\otimes W)^{\varphi}\cong V^{\varphi}\otimes^{\varphi}W^{\varphi}.

Specifically, we found one representation VV of G~=𝔾a(1)2\widetilde{G}={\mathbb{G}}_{a(1)}^{2} satisfying a polynomial identity ρ2=nρ,\rho^{2}=n\rho, i.e.

V~VnV,V\widetilde{\otimes}V\cong nV, (4.0.1)

where n=dimVn=\operatorname{dim}V, in the Green ring for the infinitesimal group scheme G~\widetilde{G}, which corresponds to a Lie Hopf algebra structure Δ~\widetilde{\Delta}. But there exists an augmented automorphism φAut(A)\varphi\in\operatorname{Aut}(A) such that the identity ρ2=nρ\rho^{2}=n\rho is not satisfied by Vφ1,V^{\varphi^{-1}}, provided p>2.p>2. Now, if we take =~φ,\otimes=\widetilde{\otimes}^{\varphi}, having VVV~VV\otimes V\cong V\widetilde{\otimes}V is equivalent to Vφ1~Vφ1nVφ1V^{\varphi^{-1}}\widetilde{\otimes}V^{\varphi^{-1}}\cong nV^{\varphi^{-1}}, which is known to be false. All that is left is to confirm the support condition that 𝒳(A,V)=𝒳(A,Vφ1)={𝔭}{\mathscr{X}}(A,V)={\mathscr{X}}(A,V^{\varphi^{-1}})=\{{\mathfrak{p}}\} for some point 𝔭𝒳(A).{\mathfrak{p}}\in{\mathscr{X}}(A). This follows from construction, that VV was chosen to meet this support condition and φ\varphi was chosen to be an isotropy in Ω(A,𝔭)<Aut(A)\Omega(A,{\mathfrak{p}})<\operatorname{Aut}(A) (3.1.8).

We will see that very little needs to be changed for a given abelian restricted Lie algebra 𝔤{\mathfrak{g}} of wild representation type. The polynomial identity 4.0.1 will always be satisfied by a given induced module V=k𝔤𝔥,V=k\uparrow^{\mathfrak{g}}_{\mathfrak{h}}, but not by some twist Vφ1,V^{\varphi^{-1}}, for a choice of Lie subalgebra 𝔥𝔤{\mathfrak{h}}\subset{\mathfrak{g}} with kk the trivial 𝔥{\mathfrak{h}}-module. In fact, in most cases 𝔥{\mathfrak{h}} can be taken to be a subalgebra isomorphic to tt[p]=0.\langle t\mid t^{[p]}=0\rangle. For p=2p=2, we will see in Proposition 4.3.2 that some wild abelian algebras require extra care.

For nonabelian Lie algebras, the polynomial identity 4.0.1 usually fails for similarly constructed induced modules. But Frobenius reciprocity can still be used to find other polynomial identities involving induced modules: whenever 𝔥𝔤{\mathfrak{h}}\subset{\mathfrak{g}} is a Lie subalgebra, denoting ~\widetilde{\otimes} the tensor product for both 𝔥{\mathfrak{h}} and 𝔤{\mathfrak{g}} representations, we have a natural isomorphism

M𝔥𝔤~N(M~(N𝔤𝔥))𝔤𝔥,M\uparrow_{{\mathfrak{h}}}^{\mathfrak{g}}\widetilde{\otimes}N\cong(M\widetilde{\otimes}(N\downarrow^{{\mathfrak{g}}}_{{\mathfrak{h}}}))\uparrow^{{\mathfrak{g}}}_{\mathfrak{h}}, (4.0.2)

whenever MM is an 𝔥{\mathfrak{h}}-representation and NN is a 𝔤{\mathfrak{g}}-representation. Choosing 𝔥{\mathfrak{h}} to be a subalgebra with an easily calculated Green ring can then let us derive ad-hoc polynomial identities (4.1.2, 4.1.3) for induced representations of 𝔤{\mathfrak{g}}, which fail after twisting by some isotropy φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}). We offer the Heisenberg Lie algebra of arbitrary dimension 2n+1,n12n+1,n\geq 1, as an example of a nonabelian Lie algebra, of wild representation type (p>2p>2), for which this generalized technique can be applied.

4.1 Induction from the nullcone

Let 𝔤{\mathfrak{g}} be a restricted Lie algebra over a field kk of characteristic pp. Define 𝒩r(𝔤){\mathscr{N}}_{r}({\mathfrak{g}}) to be the rrth restricted nullcone of 𝔤{\mathfrak{g}}, i.e. the subset of 𝔤{\mathfrak{g}} defined by

𝒩r(𝔤)={x𝔤x[p]r=0},{\mathscr{N}}_{r}({\mathfrak{g}})=\{x\in{\mathfrak{g}}\mid x^{[p]^{r}}=0\},

with 𝒩0(𝔤)=𝟎{\mathscr{N}}_{0}({\mathfrak{g}})={\bf 0} and 𝒩(𝔤)=r𝒩r(𝔤).{\mathscr{N}}({\mathfrak{g}})=\bigcup_{r}{\mathscr{N}}_{r}({\mathfrak{g}}).

Each 𝒩r(𝔤){\mathscr{N}}_{r}({\mathfrak{g}}) is a homogeneous subvariety (i.e. a cone) of the affine space 𝔸(𝔤)=SpecS(𝔤){\mathbb{A}}({\mathfrak{g}})=\operatorname{Spec}S({\mathfrak{g}}^{*}). A simple argument showed in Proposition 2.2.1 how the projective variety (𝒩1(𝔤)){\mathbb{P}}({\mathscr{N}}_{1}({\mathfrak{g}})) covers the support variety 𝒳(u(𝔤))=ProjH(𝔤,k),{\mathscr{X}}(u({\mathfrak{g}}))=\operatorname{Proj}H^{*}({\mathfrak{g}},k), using the machinery of π\pi-points. But the earlier work of Friedlander and Parshall [FrPar86] showed how the support variety 𝒳(u(𝔤)){\mathscr{X}}(u({\mathfrak{g}})) maps homeomorphically onto (𝒩1(𝔤)){\mathbb{P}}({\mathscr{N}}_{1}({\mathfrak{g}})), an inverse to our map using π\pi-points.

We will say the elements in the difference of sets x𝒩r(𝔤)𝒩r1(𝔤)x\in{\mathscr{N}}_{r}({\mathfrak{g}})\setminus{\mathscr{N}}_{r-1}({\mathfrak{g}}) have order rr. When x𝒩(𝔤)x\in{\mathscr{N}}({\mathfrak{g}}) has order rr, we denote x𝔤\langle x\rangle\subset{\mathfrak{g}} to be the subalgebra of dimension rr, with basis x,x[p],x[p]rx,x^{[p]},\dots x^{[p]^{r}}. Notice the restricted enveloping algebra u(x)u(\langle x\rangle) is of the form

k[x]/xpr,k[x]/x^{p^{r}},

a Hopf algebra with xx primitive. This Hopf algebra agrees with the group algebra for the infinitesimal group 𝕎r(1){\mathbb{W}}_{r(1)} of length rr Witt vectors. When r=1r=1, we have 𝕎1(1)𝔾a(1).{\mathbb{W}}_{1(1)}\cong{\mathbb{G}}_{a(1)}.

For nonzero x𝒩(𝔤)x\in{\mathscr{N}}({\mathfrak{g}}) having any order r1r\geq 1, we have that x[p]r1x^{[p]^{r-1}} has order 1. Thus the subalgebra x𝔤\langle x\rangle\subset{\mathfrak{g}} will always produce a π\pi-point over kk for the infinitesimal group scheme G~\widetilde{G} corresponding to 𝔤{\mathfrak{g}}, in the form of the composition

u(x[p]r1)u(x)u(𝔤).u(\langle x^{[p]^{r-1}}\rangle)\hookrightarrow u(\langle x\rangle)\hookrightarrow u({\mathfrak{g}}).

For x𝒩(𝔤)x\in{\mathscr{N}}({\mathfrak{g}}), we will denote 𝔭(x)𝒳(u(𝔤)){\mathfrak{p}}(x)\in{\mathscr{X}}(u({\mathfrak{g}})) the homogeneous prime arising from this π\pi-point.

4.1.1.

(Representations of Witt vectors) Consider the algebra A=k[x]/xprA=k[x]/x^{p^{r}} over a field kk of characteristic pp. Assuming xx is primitive, i.e. Δ~(x)=x1+1x\widetilde{\Delta}(x)=x\otimes 1+1\otimes x, defines AA to be a cocommutative Hopf algebra, and in fact a restricted enveloping algebra for the rr-dimensional P(A)=xP(A)=\langle x\rangle (xx being of order rr in 𝒩(P(A)){\mathscr{N}}(P(A))). The infinitesimal group scheme corresponding to P(A)P(A) is denoted 𝕎r(1){\mathbb{W}}_{r(1)}, the first Frobenius kernel for length rr Witt vectors.

The representations of 𝕎r(1){\mathbb{W}}_{r(1)} are modules over AA. The indecomposable representations are thus described by Jordan blocks JiJ_{i}, for 1ipr,1\leq i\leq p^{r}, each of dimension ii. The block JprJ_{p^{r}} is the unique indecomposable projective AA-module. The Green ring for 𝕎r(1){\mathbb{W}}_{r(1)} is easily calculated, and especially well known when r=1.r=1. For now we will only use the following calculation, pertaining to the polynomial identity 4.0.1.

Let G=𝕎r(1)G={\mathbb{W}}_{r(1)}. The blocks JpsJ_{p^{s}} for srs\leq r are isomorphic to the induced modules kHGk\uparrow_{H}^{G} of trivial modules for the infinitesimal subgroup HH corresponding to the subalgebra xprs\langle x^{p^{r-s}}\rangle. By Frobenius reciprocity 4.0.2, we get the identity

Jps~JiJiHGGH.J_{p^{s}}\widetilde{\otimes}J_{i}\cong J_{i}\downarrow_{H}^{G}\uparrow^{G}_{H}.

In particular, since JpsJ_{p^{s}} is annihilated by xprsx^{p^{r-s}}, we get that each JpsJ_{p^{s}} satisfies the polynomial identity 4.0.1, i.e.

Jps~JpsnJps,J_{p^{s}}\widetilde{\otimes}J_{p^{s}}\cong nJ_{p^{s}},

where n=ps=dimJps.n=p^{s}=\operatorname{dim}J_{p^{s}}. In fact, the only GG-modules MM (of dimension nn) such that M~MnMM\widetilde{\otimes}M\cong nM are of the form MnpsJpsM\cong\frac{n}{p^{s}}J_{p^{s}} for some ss. This follows from basic considerations using Jordan canonical forms.

Definition 4.1.2.

For a restricted Lie algebra 𝔤{\mathfrak{g}} and t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}) of order r>0r>0, denote Vi,𝔤(t)=Jit𝔤V_{i,{\mathfrak{g}}}(t)=J_{i}\uparrow_{\langle t\rangle}^{{\mathfrak{g}}} for i=1,,pri=1,\dots,p^{r}. Define the nthn^{th} Mackey coefficients at tt for 𝔤{\mathfrak{g}}, denoted ci,𝔤n(t)c_{i,{\mathfrak{g}}}^{n}(t), such that

Vn,𝔤(t)𝔤tci,𝔤n(t)JiV_{n,{\mathfrak{g}}}(t)\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}\cong\sum c_{i,{\mathfrak{g}}}^{n}(t)J_{i}

as t\langle t\rangle-representations. We define the polynomial identity

ρ12=ci,𝔤1(t)ρi,\rho_{1}^{2}=\sum c_{i,{\mathfrak{g}}}^{1}(t)\rho_{i},

we call the 1st1^{st} Clebsch-Gordon-Mackey (CGM) polynomial identity for 𝔤{\mathfrak{g}}-representations. Hence, by Frobenius reciprocity, the 1st1^{st} CGM polynomial identity is always satisfied by ρi=Vi,𝔤(t)\rho_{i}=V_{i,{\mathfrak{g}}}(t), i.e.

V1,𝔤(t)~V1,𝔤(t)ci,𝔤1(t)Vi,𝔤(t).V_{1,{\mathfrak{g}}}(t)\widetilde{\otimes}V_{1,{\mathfrak{g}}}(t)\cong\sum c_{i,{\mathfrak{g}}}^{1}(t)V_{i,{\mathfrak{g}}}(t).

Higher CGM polynomials are derivable from the Green ring for 𝕎r(1),{\mathbb{W}}_{r(1)}, but we don’t make use of these.

Definition 4.1.3.

Let 𝔤{\mathfrak{g}} be a restricted Lie algebra. Let

F=F(ρ1,,ρn)[ρ1,,ρn]F=F(\rho_{1},\dots,\rho_{n})\in{\mathbb{Z}}[\rho_{1},\dots,\rho_{n}]

be an integer polynomial, with F+,FF_{+},F_{-} the positive and negative components of FF respectively, so that F=F++FF=F_{+}+F_{-}. If V1,,VnV_{1},\dots,V_{n} are representations of 𝔤{\mathfrak{g}} and G[ρ1,,ρn]G\in{\mathbb{Z}}[\rho_{1},\dots,\rho_{n}] is a polynomial with positive coefficients, we write G(V1,,Vn)G(V_{1},\dots,V_{n}) to mean the representation of 𝔤{\mathfrak{g}} built from sums \oplus and products ~.\widetilde{\otimes}. If t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}) is of positive order and there is isomorphism

F+(V1,,Vn)𝔤tF(V1,,Vn)𝔤tF_{+}(V_{1},\dots,V_{n})\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}\cong-F_{-}(V_{1},\dots,V_{n})\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}

as representations of t\langle t\rangle, then we say the identity F=0F=0 (or equivalently F+=FF_{+}=-F_{-}) is witnessed by tt for ρi=Vi.\rho_{i}=V_{i}. Note that if there exists some tt such that a polynomial identity is not witnessed by tt for ρi=Vi,\rho_{i}=V_{i}, then the polynomial identity does not hold as representations of 𝔤{\mathfrak{g}}.

The following lemma may be used to extend the negation of Property PA to a larger restricted Lie algebra in a general, nonabelian setting, provided that the induced modules Vi,𝔤(t)V_{i,{\mathfrak{g}}}(t) are not projective.

Lemma 4.1.4.

Let 𝔤{\mathfrak{g}}^{\prime} be a restricted Lie algebra and B=u(𝔤)B=u({\mathfrak{g}}^{\prime}) the restricted enveloping algebra. Assume there exists t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}^{\prime}) with associated prime 𝔭=𝔭(t)𝒳(B){\mathfrak{p}}^{\prime}={\mathfrak{p}}(t)\in{\mathscr{X}}(B), and an isotropy ψΩ(B,𝔭)\psi\in\Omega(B,{\mathfrak{p}}^{\prime}) such that the 1st1^{st} CGM polynomial identity is not witnessed by tt for ρi=Vi,𝔤(t)ψ1\rho_{i}=V_{i,{\mathfrak{g}}^{\prime}}(t)^{\psi^{-1}}.

Now suppose that 𝔤{\mathfrak{g}} is a restricted Lie algebra with 𝔤𝔤{\mathfrak{g}}^{\prime}\subset{\mathfrak{g}} a central Lie subalgebra, i.e. [𝔤,𝔤]=0[{\mathfrak{g}}^{\prime},{\mathfrak{g}}]=0. Suppose that there exists an augmented automorphism φ\varphi of the algebra A=u(𝔤)A=u({\mathfrak{g}}) extending ψ,\psi, i.e. denoting A=u(𝔤),A=u({\mathfrak{g}}), that the following diagram commutes

B{B}A{A}B{B}A.{A.}ψ\scriptstyle{\psi}φ\scriptstyle{\varphi}

Then the 1st1^{st} CGM polynomial identity is not witnessed by tt for ρi=Vi,𝔤(t)φ1,\rho_{i}=V_{i,{\mathfrak{g}}}(t)^{\varphi^{-1}}, where t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}) by the inclusion 𝒩(𝔤)𝒩(𝔤).{\mathscr{N}}({\mathfrak{g}}^{\prime})\subset{\mathscr{N}}({\mathfrak{g}}).

Proof.

Let Wi=Vi,𝔤(t)=Jit𝔤W_{i}=V_{i,{\mathfrak{g}}^{\prime}}(t)=J_{i}\uparrow_{\langle t\rangle}^{{\mathfrak{g}}^{\prime}} and Vi=Vi,𝔤(t)=Wi𝔤𝔤.V_{i}=V_{i,{\mathfrak{g}}}(t)=W_{i}\uparrow_{{\mathfrak{g}}^{\prime}}^{{\mathfrak{g}}}. Let di=c1i,𝔤(t)d_{i}=c^{1}_{i,{\mathfrak{g}}^{\prime}}(t) and ci=c1i,𝔤c_{i}=c^{1}_{i,{\mathfrak{g}}}. Now we have

W1𝔤tdiJi,V1𝔤tciJi.W_{1}\downarrow^{{\mathfrak{g}}^{\prime}}_{\langle t\rangle}\cong\sum d_{i}J_{i},\qquad V_{1}\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}\cong\sum c_{i}J_{i}.

Since 𝔤𝔤{\mathfrak{g}}^{\prime}\subset{\mathfrak{g}} is central, we have W𝔤𝔤𝔤𝔤nWW\uparrow_{{\mathfrak{g}}^{\prime}}^{{\mathfrak{g}}}\downarrow_{{\mathfrak{g}}^{\prime}}^{{\mathfrak{g}}}\cong nW for all representations WW of 𝔤,{\mathfrak{g}}^{\prime}, and hence also ci=ndi,c_{i}=nd_{i}, where n=[𝔤:𝔤].n=[{\mathfrak{g}}:{\mathfrak{g}}^{\prime}]. This follows from the theorem of Nichols and Zoeller [NZ89], which states that u(𝔤)u({\mathfrak{g}}) is free of rank nn as a left module over the Hopf-subalgebra u(𝔤),u({\mathfrak{g}}^{\prime}), and the PBW theorem.

By assumption, there is non-isomorphism

(W1ψ1~W1ψ1)𝔤tdiWiψ1𝔤t,(W_{1}^{\psi^{-1}}\widetilde{\otimes}W_{1}^{\psi^{-1}})\downarrow^{{\mathfrak{g}}^{\prime}}_{\langle t\rangle}\not\cong\sum d_{i}W_{i}^{\psi^{-1}}\downarrow^{{\mathfrak{g}}^{\prime}}_{\langle t\rangle},

and we want to show

(V1φ1~V1φ1)𝔤tciViφ1𝔤t.(V_{1}^{\varphi^{-1}}\widetilde{\otimes}V_{1}^{\varphi^{-1}})\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}\not\cong\sum c_{i}V_{i}^{\varphi^{-1}}\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}.

Let CC denote the image ψ(u(t))B.\psi(u(\langle t\rangle))\subset B. Since φ\varphi extends ψ\psi, we have also C=φ(u(t))A.C=\varphi(u(\langle t\rangle))\subset A. For representations WW of 𝔤{\mathfrak{g}}^{\prime} and VV of 𝔤{\mathfrak{g}}, we have by definition Wψ1𝔤t=WBCW^{\psi^{-1}}\downarrow^{{\mathfrak{g}}^{\prime}}_{\langle t\rangle}=W\downarrow^{B}_{C} and Vφ1𝔤t=VACV^{\varphi^{-1}}\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}=V\downarrow^{A}_{C} as representations of 𝕎r(1){\mathbb{W}}_{r(1)}, where rr is the order of tt. Now on one side we have isomorphisms

(V1φ1~V1φ1)𝔤t\displaystyle(V_{1}^{\varphi^{-1}}\widetilde{\otimes}V_{1}^{\varphi^{-1}})\downarrow^{{\mathfrak{g}}}_{\langle t\rangle} V1AC~V1AC\displaystyle\cong V_{1}\downarrow^{A}_{C}\widetilde{\otimes}V_{1}\downarrow^{A}_{C}
kt𝔤𝔤𝔤ABBC~kt𝔤𝔤𝔤ABBC\displaystyle\cong k\uparrow_{\langle t\rangle}^{{\mathfrak{g}}^{\prime}}\uparrow^{{\mathfrak{g}}}_{{\mathfrak{g}}^{\prime}}\downarrow^{A}_{B}\downarrow^{B}_{C}\widetilde{\otimes}k\uparrow_{\langle t\rangle}^{{\mathfrak{g}}^{\prime}}\uparrow^{{\mathfrak{g}}}_{{\mathfrak{g}}^{\prime}}\downarrow^{A}_{B}\downarrow^{B}_{C}
n2W1BC~W1BC\displaystyle\cong n^{2}W_{1}\downarrow^{B}_{C}\widetilde{\otimes}W_{1}\downarrow^{B}_{C}
n2(W1ψ1~W1ψ1)𝔤t.\displaystyle\cong n^{2}(W_{1}^{\psi^{-1}}\widetilde{\otimes}W_{1}^{\psi^{-1}})\downarrow^{{\mathfrak{g}}^{\prime}}_{\langle t\rangle}.

On the other side we have similarly Viφ1𝔤tnWiψ1𝔤tV_{i}^{\varphi^{-1}}\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}\cong nW_{i}^{\psi^{-1}}\downarrow^{{\mathfrak{g}}^{\prime}}_{\langle t\rangle} for each ii, and hence

ciViφ1𝔤tn2(diWiψ1𝔤t).\sum c_{i}V_{i}^{\varphi^{-1}}\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}\cong n^{2}\left(\sum d_{i}W_{i}^{\psi^{-1}}\downarrow^{{\mathfrak{g}}^{\prime}}_{\langle t\rangle}\right).

The desired non-isomorphism follows immediately. ∎

For abelian restricted Lie algebras, or more generally whenever t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}) is central, the Mackey coefficients are quite simple. We state the next few results which are considerably specialized to this situation, and directly adapt the proof of Theorem 3.2.4.

Lemma 4.1.5.

Let 𝔤{\mathfrak{g}} be a restricted Lie algebra and A=u(𝔤)A=u({\mathfrak{g}}) its restricted enveloping algebra. Suppose there exists a central nilpotent element t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}) with associated prime 𝔭=𝔭(t)𝒳(A){\mathfrak{p}}={\mathfrak{p}}(t)\in{\mathscr{X}}(A), and an isotropy φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}) (3.1.8), such that Vφ1𝔤tV^{\varphi^{-1}}\!\!\!\downarrow^{\mathfrak{g}}_{\langle t\rangle} is not isomorphic to the t\langle t\rangle-module nJpsnJ_{p^{s}}, for any n,s0n,s\geq 0, where V=V1,𝔤(t)V=V_{1,{\mathfrak{g}}}(t). Then 𝔤{\mathfrak{g}} does not satisfy Property PA.

In particular, 𝒳(A,V)={𝔭},{\mathscr{X}}(A,V)=\{{\mathfrak{p}}\}, and 𝔭{\mathfrak{p}} is noble for both G~,G~φ\widetilde{G},\widetilde{G}^{\varphi} (the infinitesimal group scheme and its twist), but there is non-isomorphism

V~VV~φV.V\widetilde{\otimes}V\not\cong V\widetilde{\otimes}^{\varphi}V.
Proof.

The support condition 𝒳(A,V)={𝔭}{\mathscr{X}}(A,V)=\{{\mathfrak{p}}\} is automatic for V=V1,𝔤(t)V=V_{1,{\mathfrak{g}}}(t) with tt a central nilpotent element: For general tt, it follows from the PBW theorem that 𝒳(A,V1,𝔤(t)){𝔭(t)},{\mathscr{X}}(A,V_{1,{\mathfrak{g}}}(t))\subseteq\{{\mathfrak{p}}(t)\}, and for central tt we have V1,𝔤(t)𝔤t=[𝔤:t]J1.V_{1,{\mathfrak{g}}}(t)\downarrow^{{\mathfrak{g}}}_{\langle t\rangle}=[{\mathfrak{g}}:\langle t\rangle]J_{1}. The prime 𝔭=𝔭(t){\mathfrak{p}}={\mathfrak{p}}(t) is noble for G~\widetilde{G} by construction, and since we assumed φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}), we also have, by Lemma 3.1.7, that φ(𝔭)=𝔭\varphi^{*}({\mathfrak{p}})={\mathfrak{p}} is noble for G~φ.\widetilde{G}^{\varphi}.

Suppose t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}) is of order rr, and let 𝔥=t{\mathfrak{h}}=\langle t\rangle be the rr-dimensional subalgebra. Since 𝔤{\mathfrak{g}} is abelian, say of dimension nn, we get

V𝔤𝔥=k𝔥𝔤𝔤𝔥=[𝔤:𝔥]k,V\downarrow^{\mathfrak{g}}_{\mathfrak{h}}=k\uparrow_{{\mathfrak{h}}}^{\mathfrak{g}}\downarrow^{\mathfrak{g}}_{\mathfrak{h}}=[{\mathfrak{g}}:{\mathfrak{h}}]k,

where [𝔤:𝔥]=dimV=pnr[{\mathfrak{g}}:{\mathfrak{h}}]=\operatorname{dim}V=p^{n-r}. By 4.0.2, we have the polynomial identity 4.0.1 is satisfied by VV:

V~V[𝔤:𝔥]V.V\widetilde{\otimes}V\cong[{\mathfrak{g}}:{\mathfrak{h}}]V.

But we assumed that Vφ1𝔥𝔤V^{\varphi^{-1}}\!\!\!\downarrow_{{\mathfrak{h}}}^{\mathfrak{g}} is not isomorphic to any nJpsnJ_{p^{s}}. Supposing that Vφ1V^{\varphi^{-1}} satisfies the same polynomial identity 4.0.1, by restricting we get

[𝔤:𝔥]Vφ1𝔥𝔤\displaystyle[{\mathfrak{g}}:{\mathfrak{h}}]V^{\varphi^{-1}}\!\!\!\downarrow_{\mathfrak{h}}^{\mathfrak{g}} (Vφ1~Vφ1)𝔥𝔤\displaystyle\cong(V^{\varphi^{-1}}\widetilde{\otimes}V^{\varphi^{-1}})\downarrow_{\mathfrak{h}}^{\mathfrak{g}}
(Vφ1𝔥𝔤)~(Vφ1𝔥𝔤).\displaystyle\cong(V^{\varphi^{-1}}\!\!\!\downarrow_{\mathfrak{h}}^{\mathfrak{g}})\widetilde{\otimes}(V^{\varphi^{-1}}\!\!\!\downarrow_{\mathfrak{h}}^{\mathfrak{g}}).

This contradicts the calculation we gave at the end of 4.1.1. Now we know Vφ1~Vφ1[𝔤:𝔥]Vφ1V^{\varphi^{-1}}\widetilde{\otimes}V^{\varphi^{-1}}\not\cong[{\mathfrak{g}}:{\mathfrak{h}}]V^{\varphi^{-1}} and therefore, twisting both sides, we have

V~φV[𝔤:𝔥]VV~V.V\widetilde{\otimes}^{\varphi}V\not\cong[{\mathfrak{g}}:{\mathfrak{h}}]V\cong V\widetilde{\otimes}V.

Corollary 4.1.6.

Let 𝔤{\mathfrak{g}}^{\prime} be an abelian restricted Lie algebra, and let B=u(𝔤)B=u({\mathfrak{g}}^{\prime}) denote the restricted enveloping algebra. Assume there exists t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}^{\prime}) with associated prime 𝔭=𝔭(t)𝒳(B),{\mathfrak{p}}^{\prime}={\mathfrak{p}}(t)\in{\mathscr{X}}(B), and an isotropy ψΩ(B,𝔭)\psi\in\Omega(B,{\mathfrak{p}}^{\prime}) such that Wψ1t𝔤W^{\psi^{-1}}\!\!\!\downarrow_{\langle t\rangle}^{{\mathfrak{g}}^{\prime}} is not isomorphic to the t\langle t\rangle-module nJpsnJ_{p^{s}} for any n,s0n,s\geq 0, where W=V1,𝔤(t).W=V_{1,{\mathfrak{g}}^{\prime}}(t).

Now suppose that 𝔤{\mathfrak{g}} is a restricted Lie algebra with 𝔤𝔤{\mathfrak{g}}^{\prime}\subset{\mathfrak{g}} a central Lie subalgebra. Suppose that there exists an augmented automorphism φ\varphi of the algebra u(𝔤)u({\mathfrak{g}}) extending ψ\psi.

Let 𝔭=𝔭(t)𝒳(A){\mathfrak{p}}={\mathfrak{p}}(t)\in{\mathscr{X}}(A) be the prime associated to t𝒩(𝔤)𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}^{\prime})\subset{\mathscr{N}}({\mathfrak{g}}). Then we have that φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}) is an isotropy such that Vφ1𝔤tV^{\varphi^{-1}}\!\!\!\downarrow^{\mathfrak{g}}_{\langle t\rangle} is not isomorphic to any nJpsnJ_{p^{s}}, for any n,s0n,s\geq 0, where V=V1,𝔤(t)V=V_{1,{\mathfrak{g}}}(t). Further, we have that 𝒳(A,V)={𝔭}{\mathscr{X}}(A,V)=\{{\mathfrak{p}}\}, that 𝔭{\mathfrak{p}} is noble for both the infinitesimal group scheme G~\widetilde{G} associated to 𝔤{\mathfrak{g}} and for its twist G~φ\widetilde{G}^{\varphi}, and that

V~VV~φV,V\widetilde{\otimes}V\not\cong V\widetilde{\otimes}^{\varphi}V,

so we may conclude that 𝔤{\mathfrak{g}} does satisfy Property PA.

Proof.

Let CC denote the subalgebra ψ(u(t)B\psi(u(\langle t\rangle)\subset B. That ψΩ(B,𝔭)\psi\in\Omega(B,{\mathfrak{p}}^{\prime}) is equivalent to the claim, for any BB-module MM, that MBCM\!\downarrow^{B}_{C} is projective if and only if MBtM\!\downarrow^{B}_{\langle t\rangle} is projective. A similar equivalence will work to show φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}). Since φ\varphi extends ψ\psi, we have C=φ(u(t))C=\varphi(u(\langle t\rangle)) also as a subalgebra of A.A.

Let MM be any AA-module, and let N=MABN=M\downarrow^{A}_{B} be its restriction. Assuming ψΩ(B,𝔭)\psi\in\Omega(B,{\mathfrak{p}}^{\prime}), we have immediately that MAC=NBCM\!\downarrow^{A}_{C}=N\!\downarrow^{B}_{C} is projective if and only if MAu(t)=NBu(t)M\!\downarrow^{A}_{u(\langle t\rangle)}=N\!\downarrow^{B}_{u(\langle t\rangle)} is projective. Thus φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}).

Now we have V=k𝔤t=W𝔤𝔤,V=k\!\uparrow^{{\mathfrak{g}}}_{\langle t\rangle}=W\!\uparrow^{\mathfrak{g}}_{{\mathfrak{g}}^{\prime}}, and we have assumed that Wψ1𝔤tW^{\psi^{-1}}\!\!\!\downarrow^{{\mathfrak{g}}^{\prime}}_{\langle t\rangle} is not isomorphic to any nJpsnJ_{p^{s}}. By definition Vφ1V^{\varphi^{-1}} is the base change of VV along Aφ1AA\xrightarrow{\varphi^{-1}}A. Since V=W𝔤𝔤=ABWV=W\!\uparrow^{\mathfrak{g}}_{{\mathfrak{g}}^{\prime}}=A\otimes_{B}W, we then have Vφ1V^{\varphi^{-1}} is the base change of WW along the composition BAφ1A.B\hookrightarrow A\xrightarrow{\varphi^{-1}}A. Since φ\varphi extends ψ\psi, we also have φ1\varphi^{-1} extends ψ1\psi^{-1}. Hence we have isomorphism Vφ1Wψ1𝔤𝔤.V^{\varphi^{-1}}\cong W^{\psi^{-1}}\!\uparrow_{{\mathfrak{g}}^{\prime}}^{\mathfrak{g}}.

By Lemma 4.1.4 we have Vφ1𝔤tV^{\varphi^{-1}}\downarrow^{{\mathfrak{g}}}_{\langle t\rangle} is not isomorphic to any mJpsmJ_{p^{s}}, as we know this restriction result is equivalent to the CGM polynomial identity being witnessed by tt, by Lemma 4.1.5.

What remains is to show that 𝒳(A,V)={𝔭}.{\mathscr{X}}(A,V)=\{{\mathfrak{p}}\}. For this, assuming again that 𝔤{\mathfrak{g}}^{\prime} is central in 𝔤{\mathfrak{g}}, the Nichols-Zoeller basis shows that, supposing tt is of rank rr, Vt[p]r1V\downarrow_{\langle t^{[p]^{r-1}}\rangle} is not projective. Therefore 𝔭=𝔭(t){\mathfrak{p}}={\mathfrak{p}}(t) belongs to 𝒳(A,V){\mathscr{X}}(A,V). A simple argument using the PBW basis (reviewed in 4.4.1) shows, for any restricted Lie algebra 𝔤{\mathfrak{g}}, that if V=k𝔤tV=k\uparrow^{{\mathfrak{g}}}_{\langle t\rangle} for some nonzero t𝒩(𝔤),t\in{\mathscr{N}}({\mathfrak{g}}), then 𝒳(u(𝔤),V){𝔭(t)},{\mathscr{X}}(u({\mathfrak{g}}),V)\subseteq\{{\mathfrak{p}}(t)\}, so we are done. ∎

Corollary 4.1.7.

Let 𝔤{\mathfrak{g}}^{\prime} be a restricted Lie algebra meeting the same hypotheses as Lemma 4.1.5, and 𝔤{\mathfrak{g}}^{\prime\prime} any restricted Lie algebra. Then 𝔤=𝔤𝔤{\mathfrak{g}}={\mathfrak{g}}^{\prime}\oplus{\mathfrak{g}}^{\prime\prime} also meets the hypotheses of Lemma 4.1.5 and therefore 𝔤{\mathfrak{g}} does not satisfy Property PA.

Proof.

We have isomorphism of algebras u(𝔤)u(𝔤)u(𝔤).u({\mathfrak{g}})\cong u({\mathfrak{g}}^{\prime})\otimes u({\mathfrak{g}}^{\prime\prime}). Therefore an isotropy ψΩ(𝔤,𝔭)\psi\in\Omega({\mathfrak{g}}^{\prime},{\mathfrak{p}}^{\prime}) extends to an automorphism φ:u(𝔤)u(𝔤)\varphi:u({\mathfrak{g}})\to u({\mathfrak{g}}), defined by φ=ψu(𝔤).\varphi=\psi\otimes u({\mathfrak{g}}^{\prime\prime}). Since 𝔤{\mathfrak{g}}^{\prime} is a central subalgebra of 𝔤{\mathfrak{g}}, we apply Corollary 4.1.6, and get that fgfg meets the hypothesis of Lemma 4.1.5. ∎

4.2 Representation type of abelian restricted Lie algebras

We begin by recalling the structure theorem for abelian restricted Lie algebras over an algebraically closed field kk. We denote 𝔫n{\mathfrak{n}}_{n} the pp-nilpotent cyclic Lie algebra of dimension nn, i.e.

𝔫n=x1,,xnxi[p]=xi+1, where xn+1=0.{\mathfrak{n}}_{n}=\langle x_{1},\dots,x_{n}\mid x_{i}^{[p]}=x_{i+1},\text{ where }x_{n+1}=0\rangle.

Denote 𝔱=xx[p]=x{\mathfrak{t}}=\langle x\mid x^{[p]}=x\rangle the 11-dimensional torus.

Theorem 4.2.1.

(Seligman, 1967 [Sel67]) Let 𝔤{\mathfrak{g}} be an ablian restricted Lie algebra of finite dimension over kk. Then 𝔤{\mathfrak{g}} has a direct sum decomposition as

𝔤𝔱ri1𝔫isi{\mathfrak{g}}\cong{\mathfrak{t}}^{r}\oplus\sum_{i\geq 1}{\mathfrak{n}}_{i}^{s_{i}}

for some r0r\geq 0 and finitely many nonzero si0.s_{i}\geq 0.

For any restricted Lie algebra 𝔤{\mathfrak{g}}, the algebra 𝔱𝔤{\mathfrak{t}}\oplus{\mathfrak{g}} has the same representation type as 𝔤{\mathfrak{g}}. This is because u(𝔱)u({\mathfrak{t}}) is isomorphic to a direct product of pp many copies of kk, so u(𝔱𝔤)u(𝔱)u(𝔤)u({\mathfrak{t}}\oplus{\mathfrak{g}})\cong u({\mathfrak{t}})\otimes u({\mathfrak{g}}) is a direct product of pp many copies of u(𝔤).u({\mathfrak{g}}). Therefore, in classifying abelian Lie algebras 𝔤{\mathfrak{g}} according to representation type, we may reduce to the nullcone 𝒩(𝔤){\mathscr{N}}({\mathfrak{g}}), which for 𝔤{\mathfrak{g}} abelian is a Lie subalgebra. By Seligman’s structure theorem, 𝒩(𝔤){\mathscr{N}}({\mathfrak{g}}) is a direct sum of copies of 𝔫n,n1{\mathfrak{n}}_{n},\ n\geq 1.

Theorem 4.2.2.

Let 𝔤{\mathfrak{g}} be an abelian restricted Lie algebra of finite dimension over kk, and let nn be the dimension of 𝒩(𝔤){\mathscr{N}}({\mathfrak{g}}).

  1. I.

    If 𝒩(𝔤){\mathscr{N}}({\mathfrak{g}}) is cyclic (i.e. isomorphic to 𝔫n{\mathfrak{n}}_{n}), then 𝔤{\mathfrak{g}} is of finite representation type.

  2. II.

    If 𝒩(𝔤){\mathscr{N}}({\mathfrak{g}}) is not cyclic and pn=4p^{n}=4 (i.e. p=n=2p=n=2), then 𝔤{\mathfrak{g}} is of tame representation type.

  3. III.

    In any other case pn>4p^{n}>4 and we have 𝔤{\mathfrak{g}} is of wild representation type.

Proof.

Since 𝒩(𝔤){\mathscr{N}}({\mathfrak{g}}) is a Lie subalgebra with enveloping algebra isomorphic as an associative algebra to the group algebra of some finite abelian pp-group over kk, we may appeal directly to modular representation theory of finite groups. It has long been known (see Bondarenko and Drozd [BonDrozd82]) that a group is of wild representation type over kk if and only if its Sylow pp-subgroup is not cyclic, with abelianization of order >4,>4, with the only tame pp-groups appearing in characteristic p=2p=2. In particular the only abelian pp-group of tame representation type is the Klein 44-group, and any noncyclic abelian pp-group of order >4>4 is of wild representation type. ∎

Corollary 4.2.3.

Let 𝔤{\mathfrak{g}} be an abelian restricted Lie algebra of wild representation type with no nontrivial wild direct summands (for any decomposition 𝔤𝔤𝔤,{\mathfrak{g}}\cong{\mathfrak{g}}^{\prime}\oplus{\mathfrak{g}}^{\prime\prime}, if 𝔤{\mathfrak{g}}^{\prime} is of wild representation type then 𝔤=0{\mathfrak{g}}^{\prime\prime}=0).

  1. I.

    If p=2,p=2, then 𝔤=𝔫1𝔫1𝔫1,{\mathfrak{g}}={\mathfrak{n}}_{1}\oplus{\mathfrak{n}}_{1}\oplus{\mathfrak{n}}_{1}, or 𝔤=𝔫n𝔫m{\mathfrak{g}}={\mathfrak{n}}_{n}\oplus{\mathfrak{n}}_{m} for n+m3,n+m\geq 3, and n,m1n,m\geq 1.

  2. II.

    If p>2,p>2, then 𝔤=𝔫n𝔫m{\mathfrak{g}}={\mathfrak{n}}_{n}\oplus{\mathfrak{n}}_{m} for n,m1n,m\geq 1.

4.3 No wild abelian Lie algebra satisfies Property PC

We have proven in Theorem 3.2.4 that no abelian algebra 𝔤{\mathfrak{g}} of dimension 22 with trivial restriction 𝔤[p]=0{\mathfrak{g}}^{[p]}=0 may satisfy Property PA for p>2p>2. We will show how to extend this result to all wild abelian Lie algebras.

Proposition 4.3.1.

Let 𝔤{\mathfrak{g}} be an abelian restricted Lie algebra of wild representation type, with no nontrivial wild direct summands as in Corollary 4.2.3. If 𝔤𝔫n𝔫n{\mathfrak{g}}\not\cong{\mathfrak{n}}_{n}\oplus{\mathfrak{n}}_{n} or if p2p\neq 2, then 𝔤{\mathfrak{g}} meets the hypotheses of Lemma 4.1.5 for a nilpotent t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}) of order 1: there exists t𝒩1(𝔤)t\in{\mathscr{N}}_{1}({\mathfrak{g}}) with u(t)u(𝔤)u(\langle t\rangle)\hookrightarrow u({\mathfrak{g}}) representing 𝔭𝒳(A){\mathfrak{p}}\in{\mathscr{X}}(A), and an isotropy φΩ(u(𝔤),𝔭)\varphi\in\Omega(u({\mathfrak{g}}),{\mathfrak{p}}) such that Vφ1𝔤tV^{\varphi^{-1}}\downarrow^{{\mathfrak{g}}}_{\langle t\rangle} is not trivial (isomorphic to some nJ1nJ_{1}) and not projective (isomorphic to some nJpnJ_{p}). Therefore 𝔤{\mathfrak{g}} does not satisfy Property PA.

Proof.

By Corollary 4.2.3 we have three cases to consider. But by Lemma 3.2.4, we have already covered the case where p>2p>2 and n=m=1n=m=1 for 𝔫n𝔫m{\mathfrak{n}}_{n}\oplus{\mathfrak{n}}_{m}. For the remaining cases we may assume that n+m3n+m\geq 3 in any characteristic.

Let p=2p=2 and assume 𝔤=x,y,zx[2]=y[2]=z[2]=0,{\mathfrak{g}}=\langle x,y,z\mid x^{[2]}=y^{[2]}=z^{[2]}=0\rangle, so A=u(𝔤)=k[x,y,z]/(x2,y2,z2)A=u({\mathfrak{g}})=k[x,y,z]/(x^{2},y^{2},z^{2}). Define 𝔥{\mathfrak{h}} to be the subalgebra x\langle x\rangle and define φAut(A)\varphi\in\operatorname{Aut}(A) by

φ(x)=x+yz,φ(y)=y,φ(z)=z.\varphi(x)=x+yz,\quad\varphi(y)=y,\quad\varphi(z)=z.

Then φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}) where 𝔭{\mathfrak{p}} is represented by u(𝔥)A,u({\mathfrak{h}})\hookrightarrow A, and Vφ1V^{\varphi^{-1}} is not annihilated by xx. Therefore the restriction Vφ1𝔤𝔥V^{\varphi^{-1}}\downarrow^{{\mathfrak{g}}}_{{\mathfrak{h}}} is neither trivial nor projective.

Now assume 𝔤=𝔫n𝔫m{\mathfrak{g}}={\mathfrak{n}}_{n}\oplus{\mathfrak{n}}_{m}, for n1n\geq 1 and mmax{n,2}.m\geq\operatorname{max}\{n,2\}. Denote xn+1=ym+1=0x_{n+1}=y_{m+1}=0 and take bases for cyclic summands

𝔫n=x1,,xnxi[p]=xi+1,𝔫m=y1,,ymyi[p]=yi+1,{\mathfrak{n}}_{n}=\langle x_{1},\dots,x_{n}\mid x_{i}^{[p]}=x_{i+1}\rangle,\quad{\mathfrak{n}}_{m}=\langle y_{1},\dots,y_{m}\mid y_{i}^{[p]}=y_{i+1}\rangle,

so A=u(𝔤)=k[x,y]/(xpn,ypm)A=u({\mathfrak{g}})=k[x,y]/(x^{p^{n}},y^{p^{m}}) for x=x1,y=y1.x=x_{1},y=y_{1}. We take 𝔥{\mathfrak{h}} to be the subalgebra xn=xpn1.\langle x_{n}=x^{p^{n-1}}\rangle. Now define φAut(A)\varphi\in\operatorname{Aut}(A) by

φ(x)=x+y2,φ(y)=y.\varphi(x)=x+y^{2},\quad\varphi(y)=y.

Then φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}), where 𝔭{\mathfrak{p}} is represented by u(𝔥)A.u({\mathfrak{h}})\hookrightarrow A. We have φ(xn)=(x+y2)pn1=xn+y2pn1\varphi(x_{n})=(x+y^{2})^{p^{n-1}}=x_{n}+y^{2p^{n-1}}. But 2pn1<pm2p^{n-1}<p^{m} provided either m>nm>n or p>2p>2 so Vφ1V^{\varphi^{-1}} is not annihilated by xnx_{n}. Hence Vφ1𝔤𝔥V^{\varphi^{-1}}\!\!\!\downarrow^{{\mathfrak{g}}}_{{\mathfrak{h}}} is neither trivial nor projective.

The remaining case of p=2p=2 and m=n2m=n\geq 2 has been excluded from the above Proposition, as one finds it is necessary to use a t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}) of order nn, not simply order 1. Indeed, have A=u(𝔤)=k[x,y]/(x2n,y2n)A=u({\mathfrak{g}})=k[x,y]/(x^{2^{n}},y^{2^{n}}), and any 1-dimensional restricted Lie subalgebra 𝔥𝔤{\mathfrak{h}}\subset{\mathfrak{g}} is generated by t=ax2n1+by2n1t=ax^{2^{n-1}}+by^{2^{n-1}} for a,ba,b not both 0. We may assume a=1,b=0a=1,b=0. Any φA\varphi\in A fixing the corresponding point 𝔭=[1:0]1,{\mathfrak{p}}=[1:0]\in{\mathbb{P}}^{1}, must have φ(x)=x+ξ\varphi(x)=x+\xi for a higher order term ξ\xi, in which case φ\varphi fixes x2n1A.x^{2^{n-1}}\in A. This case is dealt with in the next proposition.

Proposition 4.3.2.

Let kk be an algebraically closed field of characteristic p=2p=2, and let 𝔤=𝔫n𝔫n{\mathfrak{g}}={\mathfrak{n}}_{n}\oplus{\mathfrak{n}}_{n} for n2n\geq 2, with A=u(𝔤)=k[x,y]/(x2n,y2n).A=u({\mathfrak{g}})=k[x,y]/(x^{2^{n}},y^{2^{n}}). Then x𝒩n(𝔤)x\in{\mathscr{N}}_{n}({\mathfrak{g}}), so take 𝔥=x{\mathfrak{h}}=\langle x\rangle be the cyclic Lie subalgebra of 𝔤{\mathfrak{g}} of dimension nn. The associated prime 𝔭=𝔭(x)𝒳(A),{\mathfrak{p}}={\mathfrak{p}}(x)\in{\mathscr{X}}(A), written in coordinates dual to the basis x2n1,y2n1x^{2^{n-1}},y^{2^{n-1}} for 𝒩1(𝔤),{\mathscr{N}}_{1}({\mathfrak{g}}), is the point 𝔭=[1:0]𝒳(A)=1.{\mathfrak{p}}=[1:0]\in{\mathscr{X}}(A)={\mathbb{P}}^{1}. Let V=k𝔥𝔤V=k\uparrow_{{\mathfrak{h}}}^{{\mathfrak{g}}} be the induced module of the trivial 𝔥{\mathfrak{h}} module. Then there exists an isotropy φΩ(A,𝔭)\varphi\in\Omega(A,{\mathfrak{p}}) such that Vφ1V^{\varphi^{-1}} is not isomorphic to any nJ2snJ_{2^{s}}. Therefore 𝔤{\mathfrak{g}} does not satisfy Property PA.

Proof.

Define an automorphism φAut(A)\varphi\in\operatorname{Aut}(A) by

φ(x)=x+y2n11,φ(y)=y.\varphi(x)=x+y^{2^{n-1}-1},\quad\varphi(y)=y.

Then φΩ(A,𝔭).\varphi\in\Omega(A,{\mathfrak{p}}). Now we examine the representation Vφ1𝔥𝔤,V^{\varphi^{-1}}\downarrow_{{\mathfrak{h}}}^{\mathfrak{g}}, which has its decomposition into Jordan blocks determined by the action of xAx\in A on Vφ1V^{\varphi^{-1}}. The action of xx on Vφ1V^{\varphi^{-1}} is a matrix agreeing with the action of y2n11y^{2^{n-1}-1} on V,V, the induced module. The module V=u(𝔤)u(𝔥)kV=u({\mathfrak{g}})\otimes_{u({\mathfrak{h}})}\otimes k has a kk-linear basis of elements yi1,y^{i}\otimes 1, for 0i2n1.0\leq i\leq 2^{n-1}. The Jordan decomposition of y2n11y^{2^{n-1}-1} consists of two blocks isomorphic to J2,J_{2}, with kk-linear bases {y2n11,y1}\{y^{2^{n-1}}\otimes 1,y\otimes 1\} and {y2n111,11}\{y^{2^{n-1}-1}\otimes 1,1\otimes 1\}. The other blocks are all isomorphic to J1.J_{1}. In particular not all the blocks are of the same size so we are done. ∎

4.4 A family of nonabelian wild Lie algebras

Assume for this section that p>2.p>2. Let 𝔤n{\mathfrak{g}}_{n} be the Heisenberg Lie algebra of dimension 2n+12n+1, having presentation

𝔤n=xi,yi,z;1in[xi,yj]=δijz,[z,xi]=[z,yi]=0,xi[p]=yi[p]=z[p]=0.{\mathfrak{g}}_{n}=\langle x_{i},y_{i},z;1\leq i\leq n\mid[x_{i},y_{j}]=\delta_{ij}z,\ [z,x_{i}]=[z,y_{i}]=0,\ x_{i}^{[p]}=y_{i}^{[p]}=z^{[p]}=0\rangle.

We have canonical embeddings 𝔤n𝔤n+1{\mathfrak{g}}_{n}\subset{\mathfrak{g}}_{n+1} of Lie algebras by keeping the indexing of xi,yix_{i},y_{i} the same. Assuming p>2p>2, we have that each 𝔤n{\mathfrak{g}}_{n} is of wild representation type. In this section we will first show that 𝔤1{\mathfrak{g}}_{1} does not satisfy Property PA by an argument of polynomial identities à la Lemma 4.1.5, and then that this can be extended to any 𝔤n{\mathfrak{g}}_{n} via Lemma 4.1.4 (note each subalgebra 𝔤n𝔤n+1{\mathfrak{g}}_{n}\subset{\mathfrak{g}}_{n+1} is central).

Since [[𝔤n,𝔤n],𝔤n]=0[[{\mathfrak{g}}_{n},{\mathfrak{g}}_{n}],{\mathfrak{g}}_{n}]=0 and a basis for 𝔤n{\mathfrak{g}}_{n} is annihilated by the [p][p] restriction mapping, it follows that 𝔤n[p]=0.{\mathfrak{g}}_{n}^{[p]}=0. Thus 𝒩1(𝔤n)=𝔤n,{\mathscr{N}}_{1}({\mathfrak{g}}_{n})={\mathfrak{g}}_{n}, and we may identify 𝒳(𝔤n){\mathscr{X}}({\mathfrak{g}}_{n}) with (𝔤n),{\mathbb{P}}({\mathfrak{g}}_{n}), i.e. the variety 2n{\mathbb{P}}^{2n} of 11-dimensional subspaces of 𝔤n{\mathfrak{g}}_{n}.

4.4.1.

For abelian restricted Lie algebras 𝔤{\mathfrak{g}} and their corresponding infinitesimal group schemes G~\widetilde{G}, the equivalence relation on general π\pi-points is straightforward. On one hand the structure of cohomology is easier to deal with, using well-known constructions for minimal resolutions. On the other hand even our note 2.3.2 is easier to apply in the case of abelian Lie algebras: the induced module V=kt𝔤V=k\uparrow_{\langle t\rangle}^{{\mathfrak{g}}} from a subgroup t(𝒩1(𝔤))\langle t\rangle\in{\mathbb{P}}({\mathscr{N}}_{1}({\mathfrak{g}})) is easily shown to be supported only at the corresponding point 𝔭(t)𝒳(G~){\mathfrak{p}}(t)\in{\mathscr{X}}(\widetilde{G}), by restricting along each subalgebra in 1(𝒩1(𝔤)){\mathbb{P}}^{1}({\mathscr{N}}_{1}({\mathfrak{g}})). The equivalence class of π\pi-points corresponding to 𝔭(t){\mathfrak{p}}(t) is therefore {αα(V) is not projective}.\{\alpha\mid\alpha^{*}(V)\text{ is not projective}\}.

The latter approach is adaptable to the following: let 𝔤{\mathfrak{g}} be any finite dimensional restricted Lie algebra and A=u(𝔤)A=u({\mathfrak{g}}). Let 𝔭=𝔭(t){\mathfrak{p}}={\mathfrak{p}}(t) for some nonzero t𝒩(𝔤)t\in{\mathscr{N}}({\mathfrak{g}}), and let V=k𝔤tV=k\uparrow^{{\mathfrak{g}}}_{\langle t\rangle} be the induced module. Say tt is of order rr and so t\langle t\rangle is rr-dimensional. Identifying 𝒳(A)=(𝒩1(𝔤)){\mathscr{X}}(A)={\mathbb{P}}({\mathscr{N}}_{1}({\mathfrak{g}})) by Friedlander and Parshall [FrPar86], we want to show that 𝒳(A,V){𝔭}.{\mathscr{X}}(A,V)\subset\{{\mathfrak{p}}\}. Without loss of generality assume {𝔭}𝒳(A)\{{\mathfrak{p}}\}\subsetneq{\mathscr{X}}(A), and let 𝔮=𝔭(s)𝒳(A){\mathfrak{q}}={\mathfrak{p}}(s)\in{\mathscr{X}}(A) be distinct from 𝔭{\mathfrak{p}}, for some nonzero s𝒩1(𝔤)s\in{\mathscr{N}}_{1}({\mathfrak{g}}). In particular s,t,t[p],t[p]r1s,t,t^{[p]},\dots t^{[p]^{r-1}} is a linearly independent set. Extend this to an ordered basis

s1<<sns_{1}<\dots<s_{n}

for 𝔤{\mathfrak{g}}, by assuming s=s1s=s_{1} and sni=t[p]is_{n-i}=t^{[p]^{i}} for 0ir1.0\leq i\leq r-1. By the PBW theorem, u(𝔤)u({\mathfrak{g}}) has a kk-basis of ordered monomials in the coordinates sjs_{j}, 1jn1\leq j\leq n. Therefore, for the trivial t\langle t\rangle-module kk, the induced module

V=k𝔤t=u(𝔤)u(t)kV=k\uparrow^{{\mathfrak{g}}}_{\langle t\rangle}=u({\mathfrak{g}})\otimes_{u(\langle t\rangle)}k

has a kk-basis of simple tensors sα1s^{\alpha}\otimes 1, for ordered monomials sαs^{\alpha} in the coordinates sjs_{j} for 1jnr.1\leq j\leq n-r. Hence VsV\downarrow_{\langle s\rangle} is free over u(s)=k[s]/sp,u(\langle s\rangle)=k[s]/s^{p}, since

ss1sβ1=s1+1sβ1s\cdot s_{1}^{\ell}s^{\beta}\otimes 1=s_{1}^{\ell+1}s^{\beta}\otimes 1

for any ordered monomial sβs^{\beta} in the coordinates sjs_{j} for 2jnr.2\leq j\leq n-r. We conclude 𝔮𝒳(A,V).{\mathfrak{q}}\not\in{\mathscr{X}}(A,V).

If VV is known to be not projective over u(𝔤)u({\mathfrak{g}}), then we know in fact that 𝒳(A,V)={𝔭},{\mathscr{X}}(A,V)=\{{\mathfrak{p}}\}, as the support must be nonempty. Recall the algebra 𝔤{\mathfrak{g}} is called unipotent if u(𝔤)u({\mathfrak{g}}) is a local ring (having a unique maximal left ideal). Whenever 𝔤{\mathfrak{g}} is unipotent, the induced module VV can not be projective because VV has dimension pnr,p^{n-r}, which is strictly smaller than the rank 1 free module, the smallest projective. If 𝔤=𝔤𝔲{\mathfrak{g}}={\mathfrak{g}}^{\prime}\oplus\mathfrak{u} for some unipotent 𝔲\mathfrak{u}, and t𝒩(𝔲)t\in{\mathscr{N}}(\mathfrak{u}), then the induced module VV also can not be projective. Thus if 𝔤{\mathfrak{g}} is any abelian restricted Lie algebra, we have another proof that 𝒳(A,V)={𝔭}{\mathscr{X}}(A,V)=\{{\mathfrak{p}}\} using Seligman’s structure theorem, since any sum of pp-nilpotent cyclic Lie algebras is unipotent. But we have in fact done enough work to compute when two π\pi-points are equivalent in some important nonabelian cases as well, without resorting to resolutions!

4.4.2.

(Induced modules for 𝔤1{\mathfrak{g}}_{1}) Here we will give matrices describing the induced modules Vr=V1,𝔤(x)V_{r}=V_{1,{\mathfrak{g}}}(x) for 𝔤=𝔤1{\mathfrak{g}}={\mathfrak{g}}_{1}, where x=x1𝔤x=x_{1}\in{\mathfrak{g}} and JrJ_{r} denotes the unique indecomposable x\langle x\rangle-representation of dimension rr. We also denote y=y1,A=u(𝔤)y=y_{1},A=u({\mathfrak{g}}), and D=u(x)k[x]/xpD=u(\langle x\rangle)\cong k[x]/x^{p}. Now we have JrD/DxrJ_{r}\cong D/Dx^{r}, where DxrDx^{r} is the left-ideal, and hence VrA/Axr.V_{r}\cong A/Ax^{r}. We choose y<z<xy<z<x as an ordered basis for 𝔤{\mathfrak{g}}, so ordered monomials yizjx, 0i,j,p1,y^{i}z^{j}x^{\ell},\ \ 0\leq i,j,\ell\leq p-1, are a basis for AA. The action of xx on AA is given by

xyizjx=yizjx+1+iyi1zj+1x.x\cdot y^{i}z^{j}x^{\ell}=y^{i}z^{j}x^{\ell+1}+iy^{i-1}z^{j+1}x^{\ell}.

Now we’ll give matrices for the action of xx on Vr,V_{r}, to find the DD-module structure of each VrAD.V_{r}\downarrow^{A}_{D}. We put a lexicographical order on the basis of representing monomials yizjx,0i,jp1, 0r1,y^{i}z^{j}x^{\ell},\quad 0\leq i,j\leq p-1,\ \ 0\leq\ell\leq r-1, for VrA/AxrV_{r}\cong A/Ax^{r} by the following relations:

yizjx<yizjx\displaystyle y^{i}z^{j}x^{\ell}<y^{i}z^{j^{\prime}}x^{\ell}\quad j>ji,,\displaystyle\iff\quad j>j^{\prime}\qquad\forall i,\ell,
yizjx<yizjx\displaystyle y^{i}z^{j}x^{\ell}<y^{i^{\prime}}z^{j^{\prime}}x^{\ell}\quad i<ij,j,\displaystyle\iff\quad i<i^{\prime}\qquad\forall j,j^{\prime},\ell
yizjx<yizjx\displaystyle y^{i}z^{j}x^{\ell}<y^{i^{\prime}}z^{j^{\prime}}x^{\ell^{\prime}}\quad >i,i,j,j.\displaystyle\iff\quad\ell>\ell^{\prime}\qquad\forall i,i^{\prime},j,j^{\prime}.

Recall our notation 𝔑p\mathfrak{N}_{p} for the p×pp\times p upper triangular nilpotent Jordan block of rank p1p-1. Now we define the block matrix 𝔐\mathfrak{M}, with p×pp\times p blocks, each of size p×pp\times p

𝔐={\mathfrak{M}}=0{0}𝔑p{\mathfrak{N}_{p}}0{0}{\cdots}0{0}0{0}0{0}2𝔑p{2\mathfrak{N}_{p}}{\cdots}0{0}