194
Views
0
CrossRef citations to date
0
Altmetric
Research Articles

Minimum codimension of eigenspaces in irreducible representations of simple classical linear algebraic groups

Pages 2558-2597 | Received 16 Dec 2022, Accepted 02 Jan 2024, Published online: 23 Jan 2024

Abstract

Let k be an algebraically closed field of characteristic p0, let G be a simple simply connected classical linear algebraic group of rank l and let T be a maximal torus in G with rational character group X(T). For a nonzero p-restricted dominant weight λX(T), let V be the associated irreducible kG-module. We define νG(V) as the minimum codimension of any eigenspace on V for any non-central element of G. In this paper, we determine lower-bounds for νG(V) for G of type Al and dim(V)l32, and for G of type Bl,Cl, or Dl and dim(V)4l3. Moreover, we give the exact value of νG(V) for G of type Al with l15; for G of type Bl or Cl with l14; and for G of type Dl with l16.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1 Introduction

Let k be an algebraically closed field of characteristic p0, let V be a finite-dimensional k-vector space and let H be a group acting linearly on V. For hH denote by Vh(μ) the eigenspace corresponding to the eigenvalue μk* of h on V. Set νH(V)=min{dim(V)-dim(Vh(μ))|hHZ(H) and μk*}.

In [Citation2], one can find the classification of groups H acting linearly, irreducibly and primitively on a vector space V (over a field of characteristic zero) that contain an element h for which νH(V) is small when compared to dim(V). The following year, Hall, Liebeck and Seitz, [Citation6], expanded on Gordeev’s result by working over algebraically closed fields of arbitrary characteristic, and they proved that, in the case of linear algebraic groups, if H is classical, we have νH(V)n8(2l+1), where l is the rank of H and V is a faithful rational irreducible kH-module of dimension n; while, if H is not of classical type, then νH(V)>n12. Now, with the lower-bounds for νH(V) known, the following natural step was to start the classification of pairs (H, V) with bounded νH(V) from above, in particular the pairs (H, V) with νH(V)=1 or νH(V)=2 have been of great interest, see for example [Citation9, Citation10, Citation23]. In [Citation4], the irreducible subgroups H of GL(V), where V is a finite-dimensional k-vector space of dimension n > 1, which act primitively and tensor-indecomposably on V and νH(V)max{2,n2} have been classified.

Let G be a simple simply connected classical linear algebraic group of rank l with l1 over k and let V be a nontrivial rational irreducible tensor-indecomposable kG-module. In this paper, we determine νG(V) in the following cases:

  1. G is of type Al with l15 and dim(V)l32;

  2. G is of type Bl with l14 and dim(V)4l3;

  3. G is of type Cl with l14 and dim(V)4l3;

  4. G is of type Dl with l16 and dim(V)4l3.

Moreover, for the groups of smaller rank and their corresponding irreducible tensor-indecomposable modules with dimensions satisfying the above bounds, we improve the known lower-bounds for νG(V) (see [Citation4, Theorem 8.4]). The origin of this paper is the PhD thesis of the author, in which the classification of pairs (G, V) with νG(V)dim(V) was established. We now state the main results of this paper. The notation used will be introduced in Section 2.

Theorem 1.1.

Let k be an algebraically closed field of characteristic p0 and let G be a simple simply connected classical linear algebraic group of rank l over k. When G is of type Bl, we assume that p2. Let T be a maximal torus in G and let V=LG(λ), where λ is a nonzero p-restricted dominant weight. When G is of type Al with l1 assume that dim(V)l32. In all other cases (G of type Cl with l2; Bl with l3; Dl with l4) assume that dim(V)4l3. The value of νG(V) is as given .

Table 1 The value of νG(V).

In the following section, we fix the notation and terminology used throughout the text. In Section 3 we go over preliminary results, we establish an algorithm for calculating νG(V), and, for each type of classical group G, we determine the complete list of kG-modules that are candidates for Theorem 1.1.

The proof of Theorem 1.1 is given in Sections 4–7, where each section is dedicated to one of the types of classical groups.

2 Notation

Throughout the text k is an algebraically closed field of characteristic p0. Note that when we write pp0, for some prime p0, we allow p = 0. Let G be a simple simply connected classical linear algebraic group of rank l and let T be a maximal torus in G with rational character group X(T). Let Y(T) be the group of rational cocharacters of T and let , be the perfect natural pairing on X(T)×Y(T). We denote by Φ the root system of G corresponding to T and by Δ={α1,,αl}a set of simple roots in Φ, where we use the standard Bourbaki labeling, as given in [Citation7, 11.4, p.58]. Let Φ+ be the set of positive roots of G. Following [Citation1, Section 2.1], we fix a total order on Φ: for α,βΦ we have α≼β if and only if α=β, or βα=i=1raiαi with 1rl,aiZ,1ir, and ar > 0.

For each αΦ, let hαY(T) be its corresponding coroot, let Uα be its corresponding root subgroup and let xα:kUα be an isomorphism of algebraic groups with the property that txα(c)t1=xα(α(t)c) for all tT and all ck. Let Gs be the set of semisimple elements of G and let Gu be the set of unipotent elements of G. Any sT can be written s=αiΔhαi(cαi), where cαik*, respectively any uGu can be written as u=αΦ+xα(cα), where cαk and the product respects . Lastly, let B be the positive Borel subgroup of G, W be the Weyl group of G corresponding to T, and w0W be the longest word.

The set of dominant weights of G with respect to Δ is denoted by X(T)+, and the set of p-restricted dominant weights by X(T)p+. We adopt the usual convention that when p = 0, all weights are p-restricted. For λX(T)+, we denote by LG(λ) the irreducible kG-module of highest weight λ. Further, we denote by ωi, 1il, the fundamental dominant weight of G with respect to αi.

All representations and all modules of a linear algebraic group are assumed to be rational and nonzero. For a kG-module V we will use the notation V=W1|W2||Wm to express that V has a composition series V=V1V2VmVm+1=0 with composition factors WiVi/Vi+1,1im. Further, by Vm we denote the direct sum VV, in which V occurs m times. When p > 0, we denote by V(pi) the kG-module obtained from V by twisting the action of G with the Frobenius endomorphism i times, see [Citation16, Section 16.2]. Lastly, we will denote the natural module of G by W.

For mZ0, we define εm:Z0{0,1} by εm(n)=0 if m∤n and εm(n)=1 if m|n.

3 Preliminary results

To begin, we prove the following result which gives us the strategy we will use for calculating νG(V).

Proposition 3.1.

Let G be a simple linear algebraic group and let V be an irreducible kG-module. Then: νG(V)=dim(V)-max{maxsTZ(G)dim(Vs(μ)),maxuGu{1}dim(Vu(1))}.

Proof.

Let ρ:GGL(V) be the associated representation and let gGZ(G). We write the Jordan decomposition of g=gsgu=gugs, where gsGs and guGu. By [Citation16, Theorem 2.5], ρ(g)=ρ(gs)ρ(gu)=ρ(g)sρ(g)u is the Jordan decomposition of ρ(g) in GL(V). We choose a basis of V with the property that ρ(g) is written in its Jordan normal form. Then, with respect to this basis, ρ(g)s is the diagonal matrix whose entries are just the diagonal entries of ρ(g), while ρ(g)u is the unipotent matrix obtained from the Jordan normal form of ρ(g) by dividing all entries of each Jordan block by the diagonal element. We distinguish the following two cases:

Case 1: Assume gsZ(G). First, we remark that gu1, as gZ(G). Secondly, as gsZ(G), it follows that ρ(g)s=diag(c,c,,c) for some ck*. Thereby, c is the sole eigenvalue of ρ(g) on V and we have dim(Vg(c))=dim(Vgu(1))maxuGu{1}dim(Vu(1)).

Case 2: Assume gsZ(G). Then, since ρ(g)s is a diagonal matrix with entries the diagonal entries of ρ(g), we determine that ρ(g) and ρ(g)s have the same eigenvalues on V and, for any such eigenvalue ck* we have dim(Vg(c))dim(Vgs(c))maxsGsZ(G)dim(Vs(μ))=maxsTZ(G)dim(Vs(μ)), where the last equality follows by [Citation16, Corollary 4.5 and Theorem 4.4]. □

3.1 Group isogenies and irreducible modules

In this section, we will assume that the simple algebraic group G is not simply connected, and we let G˜ be its simply connected cover. Fix a central isogeny ϕ:G˜G with ker(ϕ)Z(G˜) and 0. Let T˜ be a maximal torus in G˜ with the property that ϕ(T˜)=T and, similarly, let B˜ be the Borel subgroup of G˜ given by ϕ(B˜)=B. Let λX(T)+. Since X(T)X(T˜), we will denote by λ˜ the weight λ when viewing it as an element of X(T˜). By [Citation8, II.2.10], as λX(T)+, it follows that λ˜X(T˜)+. Moreover, by the same result, we have that LG(λ) is a simple kG˜-module and LG(λ)LG˜(λ˜) as kG˜-modules.

Lemma 3.2.

We have maxs˜T˜Z(G˜)dim((LG˜(λ˜))s˜(μ˜))=maxsTZ(G)dim((LG(λ))s(μ)) and maxu˜G˜u{1}dim((LG˜(λ˜))u˜(1)) =maxuGu{1}dim((LG(λ))u(1)). In particular, we have νG˜(LG˜(λ˜))=νG(LG(λ)).

Proof.

Let g˜G˜Z(G˜) and let μ˜k* be an eigenvalue of g˜ on LG˜(λ˜). Let g=ϕ(g˜) and note that gGZ(G). Denote by μk* the eigenvalue of g˜ on LG(λ) corresponding to μ˜ under LG(λ)LG˜(λ˜). We have that: dim((LG˜(λ˜))g˜(μ˜))=dim((LG(λ))g˜(μ))=dim((LG(λ))ϕ(g˜)(μ))=dim((LG(λ))g(μ))max(g,μ)(GZ(G))×k*dim((LG(λ))g(μ)).

Lastly, let (g*,μ*)GZ(G)×k* be such that dim((LG(λ))g*(μ*))=max(g,μ)(GZ(G))×k*dim((LG(λ))g(μ)). As the map ϕ:G˜G is surjective, let g*˜ be an arbitrary preimage of g* in G˜ and μ*˜k* be the eigenvalue of g*˜ on LG˜(λ˜) corresponding to μ* under LG(λ)LG˜(λ˜). Then: dim((LG˜(λ˜))g*˜(μ*˜))=max(g,μ)(GZ(G))×k*dim((LG(λ))g(μ)).

The following result justifies why we only treat groups of type Bl and their respective modules over fields of characteristic different to 2.

Lemma 3.3.

Let p = 2. Let B, respectively C, be a simple simply connected linear algebraic group of type Bl, respectively of type Cl. Let ωiB,1il, respectively ωiC,1il, be the fundamental dominant weights of B, respectively of C. Then, for any 2-restricted dominant weight i=1ldiωiB of B we have that: νB(LB(i=1ldiωiB))=νC(LC(i=1l1diωiC)(2)LC(ωl)).

Proof.

As p = 2, there exists an exceptional isogeny ϕ:CB between the two groups, see [Citation21, Theorem 28]. Consequently, we can induce irreducible kC-modules from irreducible kB-modules by twisting with the isogeny ϕ. Therefore, we have: LB(i=1ldiωiB)LC(2i=1l1diωiC+dlωlC)LC(i=1l1diωiC)(2)LC(dlωlC).

Remark 3.4.

In view of Lemma 3.3, for any 2-restricted dominant weight μ=i=1l1diωiB of B, we have νB(LB(μ))=νC(LC(2λ))=νC(LC(λ)(2))=νC(LC(λ)), where λ=i=1l1diωiC. Similarly, for the weight ωlB, we have νB(LB(ωlB))=νC(LC(ωlC)). Lastly, in the case of weights of the form μ=i=1ldiωiB, where dl=1 and there exists 1il1 such that di = 1, in order to determine νB(LB(μ)) it suffices to calculate νC(LC(i=1l1diωiC)(2)LC(ωl)).

3.2 Restriction to Levi subgroups

We return to the situation where G is simply connected. For each 1il, let Pi be the maximal parabolic subgroup of G corresponding to Δi:=Δ{αi}, and let Li=T,U±α1,, U±αi1,U±αi+1, ,U±αl be a Levi subgroup of Pi. The root system of Li is Φi=Φ(Zα1++Zαi1+Zαi+1++Zαl), in which Δi is a set of simple roots. Now, we have that Li=Z(Li)°[Li,Li], where Z(Li)°=(jiker(αj))° is a one-dimensional subtorus of G and [Li,Li] is a semisimple simply connected linear algebraic group of rank l1, see [Citation16, Proposition 12.14]. Lastly, let Ti=T[Li,Li] be a maximal torus in [Li,Li], contained in the Borel subgroup Bi=B[Li,Li]. We will abuse notation and denote the fundamental dominant weights of Li corresponding to Δi by ω1,,ωi1,ωi+1,,ωl.

Let λX(T)+,λ=i=1ldiωi, let V=LG(λ) be the associated irreducible kG-module, and let Λ(V) be the set of weights of V. Fix some 1il. We say that a weight μΛ(V) has αi-level j if μ=λjαiricrαr, where crZ0. The maximum αi-level of weights in V will be denoted by ei(λ). By [Citation8, II, Proposition 2.4(b)], we have that ei(λ) is equal to the αi-level of w0(λ). Now, consider the Levi subgroup Li of Pi. For each 0jei(λ), define the subspace Vj:=γNΔiVλjαiγ of V and note that Vj is invariant under Li. Then, as a k[Li,Li]-module, V admits the following decomposition: V|[Li,Li]=j=0ei(λ)Vj,where, by [Citation20, Proposition], V0=γNΔiVλγ is the irreducible k[Li,Li]-module of highest weight λ|Ti.

Lemma 3.5.

Assume V is a self-dual kG-module. Then, for all 0jei(λ)2, we have Vei(λ)j(Vj)*, as k[Li,Li]-modules.

Proof.

We note that, as V is self-dual, we have w0(λ)=λ and V is equipped with a nondegenerate bilinear form (,). Let μ,μΛ(V) be such that μμ. Let vVμ and vVμ. Then (v,v)=(t·v,t·v)=(μ(t)v,μ(t)v)=(μ+μ)(t)(v,v), for all tT. Therefore (v,v)=0, as μμ, and so VμVμ. Moreover, as (,) is nondegenerate, it follows that μΛ(V) for all μΛ(V).

Secondly, let μΛ(V) be a weight of αi-level j, where 0jei(λ). We will show that μ has αi-level ei(λ)j. On one hand, we know that ei(λ) is equal to the αi-level of w0(λ), thus w0(λ)=λei(λ)αiriarαr, where arZ0. On the other hand, as μ=λjαiricrαr, for crZ0, we have μ=λ+jαi+ricrαr =λ(ei(λ)j)αiribrαr, where brZ0 for all ri. Thus, μ has αi-level equal to ei(λ)j. In particular, as Vμ(Vμ) for all μμ, it follows that (Vj)rei(λ)jVr.

Lastly, as V|[Li,Li]=j=0ei(λ)Vj is self-dual, it follows that V|[Li,Li]j=0ei(λ)(Vj)*. Furthermore, as V is equipped with a nondegenerate bilinear form, we have that (Vj)*V/(Vj), for all 0jei(λ). As (Vj)rei(λ)jVr, it follows that dim((Vj)*)dim(Vei(λ)j). By the same argument, this time applied to Vei(λ)j, we determine that dim((Vei(λ)j)*)dim(Vj). Therefore, dim((Vj)*)=dim(Vei(λ)j), thus (Vj)=rei(λ)jVr, and we conclude that (Vj)*Vei(λ)j. □

Remark 3.6.

Applying Lemma 3.5, let V=LG(λ), where λ=i=1ldiωiX(T)+. As V*LG(w0(λ)), it follows that V is self-dual if w0(λ)=λ. Thus, for groups of type Al, V is self-dual if di=dl+1i for all 1il. For groups of type Bl and Cl, as w0=1, all irreducible kG-modules are self-dual. Lastly, for groups of type Dl with l even, all irreducible kG-modules are self-dual, while for groups of type Dl with l odd, V is self-dual if dl1=dl.

In what follows, we give a formula for e1(λ), the maximum α1-level of weights in LG(λ), for the classical linear algebraic groups. Further, for groups of type Cl, we also give a formula for el(λ).

Lemma 3.7.

Let G be of type Al and let λ=i=1ldiωiX(T)+. Then e1(λ)=j=1ldj.

Proof.

In order to determine e1(λ) we have to calculate the α1-level of w0(λ). We have that (3.1) w0(λ)=λ(λw0(λ))=λr=1ldr(ωrw0(ωr))=λr=1ldr(ωr+ωlr+1).(3.1)

Using [Citation7, , p. 69], we write the ωi’s, 1il, in terms of the simple roots αj, 1jl, and we see that for 1rl2, we have ωr+ωlr+1=j=1r1jαj+rj=rlr+1αj+j=lr+2l(l+1j)αj; and if l is odd, we have ωl+12=12[α1+2α2++l12·αl12+l+12·αl+12+l12·αl+12+1++αl]. Substituting in (3.1), we determine that e1(λ)=j=1ldj. □

Lemma 3.8.

Let G be of type Cl and let λ=i=1ldiωiX(T)+. Then e1(λ)=2j=1ldj and el(λ)=j=1ljdj.

Proof.

Note that we have w0=1, hence w0(λ)=λ. We write the ωi’s, 1il, in terms of the simple roots αj, 1jl, see [Citation7, , p. 69], and we get: w0(λ)=λ=λ2λ=λ2(d1++dl)α12(d1+2j=2ldj)α2(j=1ljdj)αl.

We remark that the coefficient of each αi is a nonnegative integer and the result follows. □

Lemma 3.9.

Let G be of type Bl and let λ=i=1ldiωiX(T)+. Then e1(λ)=2[j=il1dj]+dl.

Proof.

We have that w0=1, hence w0(λ)=λ. Writing the fundamental dominant weights ωi in terms of the simple roots αj, we see that: w0(λ)=λ=λ2λ=λ2i=1l[j=1i1jdj+i(j=il1dj+12dl)]αi,therefore e1(λ)=2[j=il1dj]+dl. □

Lemma 3.10.

Let G be of type Dl and let λ=i=1ldiωiX(T)+. Then e1(λ)=2[j=1l2dj+12i(dl1+dl)].

Proof.

We first assume that l is even. Then w0=1, hence w0(λ)=λ, and so w0(λ)=λ2j=1l2djωj2dl1ωl12dlωl=λr=1l22[j=1r1jdj+rj=rl2dj+12r(dl1+dl)]αr[j=1l2jdj+12(ldl1+(l2)dl)]αl1[j=1l2jdj+12((l2)dl1+ldl)]αl.

Thus e1(λ)=2[j=1l2dj+12(dl1+dl)]. We now assume that l is odd. We note that w0(ωj)=ωj, for all 1jl2, w0(ωl1)=ωl and w0(ωl)=ωl1. It follows that: w0(λ)=j=1l2djωjdl1ωldlωl1=λ2j=1l2djωj(dl1+dl)(ωl1+ωl)=λr=1l22[j=1r1jdj+rj=rl2dj+12r(dl1+dl)]αr[j=1l2jdj+12(l1)(dl1+dl)](αl1+αl)and so e1(λ)=2[j=1l2dj+12(dl1+dl)]. □

3.3 The algorithm for calculating νG(V)

Let V=LG(λ) for some λ=i=1ldiωiX(T)+. Consider the restriction V|[Li,Li]=j=0ei(λ)Vj, where 1il and Vj=γNΔiVλjαiγ, for all 1jei(λ). In view of Proposition 3.1, in order to determine νG(V), one has to calculate maxsTZ(G)dim(Vs(μ)) and maxuGu{1}dim(Vu(1)). In this section we will outline an algorithm for calculating maxsTZ(G)dim(Vs(μ)) and maxuGu{1}dim(Vu(1)).

First, let sTZ(G). Then, in particular, sLi and so s=z·h, where zZ(Li)° and h[Li,Li]. As zZ(Li)° and Z(Li)° is a one-dimensional torus, there exists ck* and krZ,1rl, such that z=r=1lhαr(ckr). Moreover, we have αj(z)=1 for all 1jl, ji. On the other hand, as h[Li,Li], we have h=rihαr(ar), where ark* for all ri. Now, as zZ(Li)°, z acts on each Vj, 0jei(λ), as scalar multiplication by szj, where: (3.2) szj:=(λjαiγ)(z)=(λjαi)(r=1lhαr(ckr))=r=1l(ckrdr)·r=1lcjkrαi,αr.(3.2)

Lastly, let μ1j,,μtjj,tj1, be the distinct eigenvalues of h on Vj, 0jei(λ), and let n1j,,ntjj be their respective multiplicities. Then, as s=z·h, it follows that the eigenvalues of s on Vj are szjμ1j,,szjμtjj and they are distinct, as the μrj’s are, with respective multiplicities n1j,,ntjj. This proves the following:

Lemma 3.11.

Let sTZ(G),s=z·h with zZ(Li)° and h[Li,Li]. Let μ1j,,μtjj,tj1, be the distinct eigenvalues of h on Vj, 0jei(λ), with respective multiplicities n1j,,ntjj. Then:

  1. z acts on Vj as scalar multiplication by szj, where szj is given in (3.2);

  2. the distinct eigenvalues of s on Vj are szjμ1j,,szjμtjj, with multiplicities n1j,,ntjj;

  3. the eigenvalues of s on V are szjμ1j,,szjμtjj,0jei(λ), with respective multiplicities at least n1j,,ntjj.

An algorithm for calculating maxsTZ(G)dim(Vs(μ))

First, assume that s admits an eigenvalue μ on V with the property that dim(Vsj(μ))=dim(Vj) for some 0jei(λ). In this case sZ(Li)°Z(G),s=r=1lhαr(ckr), and so it acts on each Vj as scalar multiplication by szj. Therefore, the eigenvalues of s on V, not necessarily distinct, are sz0,,szei(λ). We also remark that the szj’s are not all equal, as sZ(G). We have that dim(Vs(μ))=jIμdim(Vj), where Iμ{0,,ei(λ)} is such that for any jIμ we have μ=szj. Therefore, maxsZ(Li)°Z(G)dim(Vs(μ))= maxsZ(Li)°Z(G)jIμdim(Vj), where the calculation of the latter maximum is straightforward. Secondly, assume that dim(Vsj(μ))<dim(Vj) for all eigenvalues μ of s on V and all 0jei(λ). This case is solved inductively. We write s=z·h, where zZ(Li)° and h[Li,Li]. Recall that [Li,Li] is a semisimple simply connected group of rank l1. We have that dim(Vs(μ))j=0ei(λ)dim(Vhj(μhj)), where μ=szjμhj and dim(Vhj(μhj))<dim(Vj) for all 0jei(λ). Therefore, maxsTZ(Li)°dim(Vs(μ))j=0ei(λ)maxh,μhdim(Vhj(μh)), and we use induction to determine this upper-bound.

We now let uGu{1}. We will denote by k[u] the group algebra of u over k.

Lemma 3.12.

Let uG be a unipotent element and let V be a finite-dimensional kG-module. Let V=MtMt1M1M0=0, where t1, be a filtration of k[u]-submodules of V. Then: dim(Vu(1))i=1tdim((Mi/Mi1)u(1)).

Moreover, suppose that for each i, we have a u-invariant decomposition Mi=Mi1Mi1 with Mi1Mi/Mi1 as k[u]-modules. Then dim(Vu(1))=i=1tdim((Mi/Mi1)u(1)).

Proof.

For each 1it, we fix a basis in Mi with the property that the matrix (u)Mi/Mi1 associated to the action of u on Mi/Mi1 is upper-triangular. Then, the matrix (u)V of the action of u on V is the block upper-triangular matrix: (u)V=((u)M1 0 (u)M2/M1 00 (u)M3/M2 000 (u)Mt/Mt1).

Using (u)V, we calculate the matrix of the action of uidV on V: (uidV)V=((uidM1)M10(uidM2/M1)M2/M100000(uidMt/Mt1)Mt/Mt1),where (uidMi/Mi1)Mi/Mi1 is the matrix of the action of uidMi/Mi1 on Mi/Mi1, 1it, with respect to the basis of Mi we have previously fixed. It follows that: rank(uidV)i=1trank((uidMi/Mi1)Mi/Mi1) and, consequently, we have dim(ker(uidV)) i=1tdim(ker((uidV)|Mi/Mi1)). Now, as Vu(1)=ker(uidV) we determine that dim(Vu(1))i=1tdim((Mi/Mi1)u(1)).

Lastly, for all 1it, assume that there exists a k[u]-submodule Mi1 of Mi such that Mi=Mi1Mi1. Then V|k[u]=M0Mt1M1M2/M1Mt/Mt1, and so there exists a basis of V with the property that: (uidV)V=((uidM1)M10000(uidM2/M1)M2/M100000000(uidMt/Mt1)Mt/Mt1),

thereby rank(uidV)=i=1trank((uidMi/Mi1)Mi/Mi1). Arguing as above, we establish that dim(Vu(1)) =i=1tdim((Mi/Mi1)u(1)). □

An algorithm for calculating maxuGu{1}dim(Vu(1))

Let Pi=Li·Qi=T,Uβ|βΦi·Uα|αΦ+Φi be the Levi decomposition of the maximal parabolic subgroup Pi of G. Let uGu, u=αΦ+xα(cα), where the product respects the total order on Φ and cαk. Now, as uB and BPi, it follows that u admits a decomposition u=αΦixα(cα)·αΦ+Φixα(cα), where each of the products respects and cαk, for all αΦ+. We set uLi=αΦixα(cα) and uQi=αΦ+Φixα(cα), and we note that uLiLi and uQiQi. Recall that V=LG(λ) for some λX(T)+ and that V|[Li,Li]=j=0ei(λ)Vj. Let μΛ(V), with corresponding weight space Vμ, and let αΦ. As UαVμrZ0Vμ+rα, see [Citation16, Lemma 15.4], we have uLi·VjVj,uQi·Vjr=0jVr and (uQi1)·Vjr=0j1Vr, for all 0jei(λ). Therefore, V admits a filtration V=Mei(λ)Mei(λ)1M1M00 of k[u]-submodules, where Mj=r=0jVr for all 0jei(λ). We see that u acts on each Mj/Mj1, 1jei(λ), as uLi and so, by Lemma 3.12, we determine that dim(Vu(1))j=0ei(λ)dim(VuLij(1))=dim(VuLi(1)). Therefore, if we identify the kLi-composition factors of each Vj, 0jei(λ), then using already proven results and Lemma 3.12, we can establish an upper-bound for each dim(VuLij(1)). Now, assuming that uLi1, the upper-bound we obtain for dim(VuLi(1)), hence for dim(Vu(1)), will be strictly smaller than dim(V). Lastly, we remark that if u=uLi, i.e. uQi=1, then u·VjVj, for all 0jei(λ), and thus, by Lemma 3.12, it follows that dim(Vu(1))=dim(VuLi(1)).

We end this section with two lemmas concerning the behavior of unipotent elements. The first one is due to Guralnick and Lawther, [Citation5], and tells us which unipotent conjugacy classes in G afford the largest dimensional eigenspaces.

Lemma 3.13.

[Citation5, p.19 and Lemmas 1.4.1 and 1.4.4] We have dim(Vu2(1))dim(Vu1(1)), if u1Gu belongs to a unipotent conjugacy class of root elements and u2Gu belongs to any nontrivial unipotent class, unless e(Φ)>1 and one of the following holds:

  1. G=Cl, p = 2, u1 belongs to the unipotent conjugacy class of xαl(1) and u2 belongs to the unipotent conjugacy class of xα1(1).

  2. G=Cl, u1 belongs to the unipotent conjugacy class of xα1(1) and u2 belongs to the unipotent conjugacy class of xαl(1).

  3. G=Bl, u1 belongs to the unipotent conjugacy class of xαl(1) and u2 belongs to the unipotent conjugacy class of xα1(1).

The second lemma gives us dim(2(V)u(1)), when char(k)=2, V is a finite-dimensional k-vector space and the unipotent element u acts as a single Jordan block on GL(V). For each i0, let Vi be the indecomposable k[u]-module with dim(Vi)=i and on which u acts as the full Jordan block Ji of size i. Note that {Vi|i0} is a set of representatives of the isomorphisms classes of indecomposable k[u]-modules.

Lemma 3.14.

Let k be a field of characteristic p = 2 and let V be a vector space of dimension i1 over k. Let u be a unipotent element acting as a single Jordan block in GL(V). Then dim((2(V))u(1))=i2.

Proof.

We will prove the result by induction on i1. First, we note that both cases i = 1 and i = 2 follow directly from the structure of 2(V). Hence, we assume that i3 and that the result holds for all 1r<i. Let m be the unique nonnegative integer for which 2m1<i2m and set q=2m. Now, up to isomorphism, there exist exactly q indecomposable k[u]-modules: V1,V2,,Vq, where dim(Vj)=j and u acts on Vj as Jj. Therefore, as k[u]-modules, we have VVi. Now, by [Citation3, Theorem 2], we have 2(Vi)=2(Vqi)(iq21)VqV3q2i, and so (3.3) dim((2(Vi))u(1))=dim((2(Vqi))u(1))+(iq21)dim((Vq)u(1))+dim((V3q2i)u(1)).(3.3)

As 3q2i<q and as u acts as a single Jordan block on Vq and V3q2i, respectively, it follows that dim((Vq)u(1))=1 and dim((V3q2i)u(1))=1. Furthermore, we note that, as q2<i, we have qi<i and, by applying induction, it follows that dim((2(Vqi))u(1))=qi2. Substituting in (3.3) we obtain dim((2(Vi))u(1))=qi2+iq21+1=i2. □

3.4 The list of modules

Lemma 3.15.

Let k be an algebraically closed field of characteristic p0 and let G be a simple simply connected classical linear algebraic group. Let λX(T)p+,λ0.

  1. Let G be of type Al with l1 and assume that dim(LG(λ))l32. If l15, then, up to duality of the corresponding kG-module, we have λΛAl:={ω1,ω2,2ω1,ω1+ωl,ω3,3ω1,ω1+ω2}. If l14, the additional λ’s are given in .

  2. Let G be of type Cl with l2 and assume that dim(LG(λ))4l3. If l14, we have that λΛCl:={ω1,ω2,2ω1,ω3, 3ω1,ω1+ω2}. If l13, the additional λ’s are given in .

  3. Let p2 and let G be of type Bl with l3. Assume that dim(LG(λ))4l3. If l14, we have λΛBl:={ω1,ω2,2ω1,ω3, 3ω1,ω1+ω2}. If l13, the additional λ’s are given in .

  4. Let G be of type Dl with l4 and assume that dim(LG(λ))4l3. If l16, we have λΛDl:={ω1,ω2,2ω1,ω3, 3ω1,ω1+ω2}. If l15, the additional λ’s are given in (up to duality or outer automorphisms of the corresponding kG-module).

Table 2 The nonzero weights λX(T)p+ΛAl with dim(LG(λ))l32.

Table 3 The nonzero weights λX(T)p+ΛCl with dim(LG(λ))4l3.

Table 4 The nonzero weights λX(T)p+ΛBl with dim(LG(λ))4l3.

Table 5 The nonzero weights λX(T)p+ΛDl with dim(LG(λ))4l3.

Proof.

The result follows by [Citation17, Theorem 1.2] and [Citation14]. □

4 Proof of Theorem 1.1 for groups of type Al

Let G be a simple simply connected linear algebraic group of type Al with l1. In view of Proposition 3.1, it is sufficient to know maxsTZ(G)dim(Vs(μ)) and maxuGu{1}dim(Vu(1)) in order to determine νG(V), where V is any irreducible kG-module. In this section we prove Theorems 4.1 and 4.2 which provide the values of maxuGu{1}dim(Vu(1)) and maxsTZ(G)dim(Vs(μ)) for all kG-modules V=LG(λ) with λX(T)p+,λ0, and dim(LG(λ))l32. As a corollary, the part of Theorem 1.1 concerning simple simply connected linear algebraic groups of type Al will follow.

Theorem 4.1.

Let k be an algebraically closed field of characteristic p0 and let G be a simple simply connected linear algebraic group of type Al with l1. Let V=LG(λ), where λX(T)p+ and λ0, be such that dim(V)l32. The value of maxuGu{1}dim(Vu(1)) is as given in .

Table 6 The value of maxuGu{1}dim(Vu(1)) for groups of type Al.

Proof.

To begin, recall that we have denoted by W the natural module of G, see end of Section 2. Let (V,p,l) be a triplet featured in . In view of Lemma 3.13, for any 1il, we have maxuGu{1}dim(Vu(1))=dim(Vxαi(1)(1)). Thus, in what follows, we will focus on calculating dim(Vxαi(1)(1)),1il. To ease notation, we reference each triplet (V,p,l) featured in by its corresponding number in column ‘Ref’.

4.1.1: Follows because VW as kG-modules, and xα1(1) acts on W as J2J1l1.

4.1.2: Note that V2(W) by [Citation18, Proposition 4.2.2]. We write W=W1W2, where dim(W1)=2 and xα1(1) acts on W1 as J2; and dim(W2)=l1 and xα1(1) acts trivially on W2. Since 2(W1W2)2(W1)[W1W2]2(W2), we get dim((2(W))xα1(1)(1))= dim((2(W1))xα1(1)(1))+ dim((W1W2)xα1(1)(1))+dim((2(W2))xα1(1)(1)). As xα1(1) acts as a single Jordan block on 2(W1), using [Citation15, Lemma 3.4], respectively Lemma 3.14 when p = 2, we get dim((2(W1))xα1(1)(1))=1. Further, as xα1(1) acts trivially on W2, we have dim((2(W2))xα1(1)(1))= l23l+22. Lastly, as xα1(1) acts on W1W2 as J2J1l1, we have dim((W1W2)xα1(1)(1))=l1, by [Citation15, Lemma 3.4]. It follows that dim((2(W))xα1(1)(1))=l2l+22.

4.1.3: Note that VS2(W) by [Citation18, Proposition 4.2.2]. We write W=W1W2, where dim(W1)=2 and xα1(1) acts on W1 as J2; and dim(W2)=l1 and xα1(1) acts trivially on W2. Then, since S2(W1W2)S2(W1)[W1W2]S2(W2), we argue as above to show dim((S2(W))xα1(1)(1))=(l+12).

4.1.4: Follows from [Citation18, Proposition 4.6.10] and [Citation11, Theorem 6.1], arguing as in 4.1.2 and 4.1.3.

4.1.5, 4.1.6, 4.1.8, and 4.1.9: Follow from [Citation18, Proposition 4.2.2], arguing as in 4.1.2 and 4.1.3.

4.1.7: First, assume p3. By [Citation18, Proposition 4.6.10], we have that LG(ω1)LG(ω2)VLG(ω3), and so dim(Vxαl(1)(1))=dim((LG(ω1)LG(ω2))xαl(1)(1))-dim((LG(ω3))xαl(1)(1)). As xαl(1) acts on LG(ω1) as J2J1l1 and on LG(ω2) as J2l1J1l23l+42, see 4.1.2, one shows that dim((LG(ω1)LG(ω2))xαl(1)(1))=l3l2+4l22. The result now follows by 4.1.1, 4.1.2, and 4.1.5.

Now, let p = 3. Set λ=ω1+ω2 and L=L1. By Lemma 3.7, we have e1(λ)=2, therefore V|[L,L]=V0V1V2. First, by [Citation20, Proposition], we have V0LL(ω2). Secondly, the weight (λα1)|T1=2ω2 admits a maximal vector in V1, thus V1 has a composition factor isomorphic to LL(2ω2). Lastly, the weight (λ2α1α2)|T1=ω2+ω3 admits a maximal vector in V2, thus V2 has a composition factor isomorphic to LL(ω2+ω3). By dimensional considerations, we determine that V|[L,L]LL(ω2)LL(2ω2)LL(ω2+ω3).

By 4.1.1 and 4.1.3, we have dim(Vxαl(1)(1))=l1+(l2)+dim((LL(ω2+ω3))xαl(1)(1)). Recursively and using 4.1.5, for the base case of l=3, we get dim(Vxαl(1)(1)) =3+j=2l1 (j+12)+ j=2l1j= (l+23)l+1.

Theorem 4.2.

Let k be an algebraically closed field of characteristic p0 and let G be a simple simply connected linear algebraic group of type Al with l1. Let V=LG(λ), where λX(T)p+ and λ0, be such that dim(V)l32. The value of maxsTZ(G)dim(Vs(μ)) is as given in .

Table 7 The value of maxsTZ(G)dim(Vs(μ)) for groups of type Al.

In order to prove Theorem 4.2, in addition to the algorithm for calculating maxsTZ(G)dim(Vs(μ)) outlined in Section 3.3, we will sometimes use the following algorithm from [Citation5, Section 2.2], which gives lower bounds for νG(V). We give a brief description of it in what follows. Let Ψ be a standard subsystem of Φ and let W(Ψ) be its Weyl group. We define rΨ=|W:W(Ψ)|·|ΦΨ|2|Φs|. For λ=i=1laiωiX(T)+, set Ψ(λ)=αi|ai=0, and for λX(T)p+, define sλ=0λλrλ. By [Citation5, Prop. 2.2.1], we have νG(LG(λ))sλ. Now, as sλ=0≼μ≼λrμ, it will prove extremely useful to give a formula for rωi, 1il. To this end, we first note that Ψ(ωi),1il, is of type Ai1Ali, thus |Ψ(ωi)|=l22li+2i2+l2i and |ΦΨ(ωi)|=2i(li+1). Moreover, as |W(Ψ(ωi))|=i!(li+1)!, we have |W:W(Ψ(ωi))|=(l+1i), therefore (4.1) rωi=(l+1i)·2i(li+1)2l(l+1)=(l1i1).(4.1)

Let sTZ(G). To improve readability, we will use the expression “s is as in (Hs)” to mean that s satisfies the following: s=diag(μ1·In1,μ2·In2,,μm·Inm), where m2,μiμj for all i < j, ln1nm1 and i=1mμini=1. Note that any sTZ(G) is as in (Hs).

Proposition 4.3.

Let V=LG(ω1). Then maxsTZ(G)dim(Vs(μ))=l, where the maximum is attained if and only if s is conjugate to diag(d,d, ,d,dl) with dl+11 and μ=d.

Proof.

As VW as kG-modules, for all (s,μ)(TZ(G))×k* we have dim(Vs(μ))l. Now, equality holds if and only if s and μ are as in the statement of the result. □

Proposition 4.4.

Let l3 and let V=LG(ω2). Then maxsTZ(G)dim(Vs(μ))=l(l1)2+εl(3), where the maximum is attained if and only if

  1. l=3 and s is conjugate to diag(d,d,±d1,±d1) with d2±1, and μ=±1.

  2. l=4 and s is conjugate to diag(d,d,d,e,e) with de and d3=e2, and μ=de.

  3. l4 and s is conjugate to diag(d,,d,dl) with dl+11, and μ=d2.

Proof.

Let sTZ(G) be as in (Hs). As V2(W), see [Citation18, Proposition 4.2.2], we deduce that the eigenvalues of s on V, not necessarily distinct, have one of the following forms: (4.2) {μi2,1im, with multiplicity at least ni(ni1)2;μiμj,1i<jm, with multiplicity at least ninj.(4.2)

In order for s to have an eigenvalue of the form μj2, there has to exist some 1im with ni2. Suppose there exists such i, and consider the eigenvalue μi2 of s on V. Now, since the μr’s are distinct, it follows that μi2μiμj for all ij, hence: dim(Vs(μi2))l(l+1)2(l+1ni)ni.

Let l=3 and assume dim(Vs(μi2))4. Then 2(4ni)ni=(ni2)220, which does not hold as ni3. Thus, we let l4 and assume that dim(Vs(μi2))l(l1)2. Then (lni)(1ni)0 and, since lni2, the inequality holds if and only if ni=l. Hence, m = 2, n1=l,n2=1,μ2=μ1l and μ1l+11, as μ1μ2 and μ1lμ2=1. This gives dim(Vs(μi2))l(l1)2 for all sTZ(G), where equality holds if and only if i = 1 and s is conjugate to diag(μ1,,μ1,μ1l) with μ1l+11, as in (3). This completes the case of eigenvalues of the form μj2 of s on V.

Fix i<j and consider the eigenvalue μiμj of s on V. Since the μr’s are distinct, we remark that: {μiμjμi2 and μiμjμj2;μiμjμiμr, for i<rm and rj, and μiμjμrμi, for 1r<i;μiμjμrμj, for 1r<j and ri, and μiμjμjμr, for j<rm.

By (4.2), these account for at least ni(ni1)2+nj(nj1)2+(ni+nj)(l+1ninj) eigenvalues of s on V different to μiμj. It follows that: dim(Vs(μiμj))l(l+1)2ni(ni1)2nj(nj1)2(ni+nj)(l+1ninj).

As in the previous case, we begin with l=3. Then, since ni+nj4, we have (ni,nj){(3,1),(2,2),(2,1),(1,1)}. Assume that dim(Vs(μiμj))4. Then 2ni(ni1)2nj(nj1)2(ni+nj)(4ninj)0, which holds if and only if ni=nj=2, i.e. m = 2, n1=n2=2,μ2=±μ11, as μ12μ22=1, and μ12±1, as μ1μ2. Thus dim(Vs(μiμj))4 for all sTZ(G) and all i < j, where equality holds if and only if s is conjugate to diag(μ1,μ1,±μ11,±μ11) with μ12±1, as in (1). We now let l4 and assume dim(Vs(μiμj))l(l1)2. Then: (4.3) (lninj)(1ninj)ni(ni1)2nj(nj1)20.(4.3)

Since ninj1, we have ni(ni1)2+nj(nj1)20 and 1ninj<0, therefore, by inequality (4.3), it follows that lninj0. If ni+nj=l, then, for inequality (4.3) to hold, we must have ni(ni1)2+nj(nj1)2=0, hence ni=nj=1, contradicting l4. On the other hand, if ni+nj=l+1, then m = 2 and, by (4.3), we determine that n1(3n1)(n21)(n22)0. Now, the inequality holds if and only if n1=3 and n2=2, as l4. In this case, l=4 and s=diag(μ1,μ1,μ1,μ2,μ2) with μ1μ2 and μ13=μ22. Therefore, dim(Vs(μiμj))l(l1)2 for all sTZ(G) and all i < j, where equality holds if and only if l=4 and s is conjugate to diag(μ1,μ1,μ1,μ2,μ2) with μ1μ2 and μ13=μ22, as in (2). □

Proposition 4.5.

Assume p2 and let V=LG(2ω1). Then maxsTZ(G)dim(Vs(μ))=l2+l+22, where the maximum is attained if and only if s is conjugate to diag(d,,d,dl) with dl+1=1, and μ=d2.

Proof.

Let sTZ(G) be as in (Hs). Since, VS2(W), see [Citation18, Proposition 4.2.2], the eigenvalues of s on V, not necessarily distinct, have one of the following forms: (4.4) {μi2,1im, with multiplicity at least ni(ni+1)2;μiμj,1i<jm, with multiplicity at least ninj.(4.4)

Fix some i and consider the eigenvalue μi2 of s on V. Since the μr’s are distinct, we deduce that: dim(Vs(μi2))(l+1)(l+2)2ni(l+1ni).

Assume dim(Vs(μi2))l2+l+22. Then (lni)(1ni)0, which holds if and only if ni{1,l}. In both cases, we get dim(Vs(μi2))l2+l+22, where equality holds if and only if μi2=μj2 for all ji, i.e. m = 2 and s=diag(μ1,,μ1,μ1l) with μ1l+1=1.

Fix i < j and consider the eigenvalue μiμj of s on V. Since the μr’s are distinct, we have {μiμjμi2 and μiμjμj2;μiμjμiμr, for i<rm and rj, and μiμjμrμi, for 1r<i;μiμjμrμj, for 1r<j and ri, and μiμjμjμr, for j<rm.

By (4.4), these account for at least ni(ni+1)2+nj(nj+1)2+(ni+nj)(l+1ninj) eigenvalues of s on V which are different to μiμj. Hence, we have dim(Vs(μiμj))(l+1)(l+2)2ni(ni+1)2nj(nj+1)2(ni+nj)(l+1ninj). Assume dim(Vs(μiμj))l2+l+22. Then: (4.5) (lninj)(1ninj)ni(ni+1)+nj(nj+1)20.(4.5)

As ninj1, we have ni(ni+1)+nj(nj+1)22, therefore (lninj)(1ninj)20. If ni+njl, then (lninj)(1ninj)0 and inequality (4.5) does not hold. If ni+nj=l+1, then m = 2, n2=l+1n1 and inequality (4.5) does not hold, as (lninj)(1ninj)ni(ni+1)+nj(nj+1)2=[(ln1)2+(n11)2+l+1]2<0. Therefore, dim(Vs(μiμj))<l2+l+22 for all sTZ(G) and all i < j. We conclude that maxsTZ(G)dim(Vs(μ)) =l2+l+22. □

Proposition 4.6.

Let l2 and V=LG(ω1+ωl). Then maxsTZ(G)dim(Vs(μ)) =l2εp(l+1)+εp(3)εl(2), where the maximum is attained if and only if

  1. p2,l=2,μ=1 and s is conjugate to diag(d,d,d2) with d3=1.

  2. p3,l=2, μ = 1 and s is conjugate to diag(d,d,d2) with d31.

  3. l3, μ = 1 and s is conjugate to diag(d,,d,dl) with dl+11.

Proof.

First, we note that if εp(l+1)=0, then WW*VLG(0), while, if εp(l+1)=1, then WW*LG(0)|V|LG(0), see [Citation18, Proposition 4.6.10]. Let sTZ(G) be as in (Hs). We determine that the eigenvalues of s on V, not necessarily distinct, have one of the following forms: {1 with multiplicity i=1mni21εp(l+1);μiμj1 and μi1μj, where 1i<jm, each with multiplicity at least ninj.

We first consider the eigenvalue 1 of s on V. Since the μi’s are distinct, it follows that: dim(Vs(1))=i=1mni21εp(l+1)=l2+2l2i<jninjεp(l+1).

Assume dim(Vs(1))l2εp(l+1)+εp(3)εl(2). Since l=i=1mni1, we have that: (4.6) 2(1n2)(n11)+2i=3mni(1j=1i1nj)εp(3)εl(2)0,(4.6) which holds if and only if either l=2,p3, m = 2, n2=1 and n1=2; or l3, m = 2, n2=1 and n1=l. In both cases, as μ1μ2 and μ1lμ2=1, we get μ2=μ1l and μ1l+11. Thus, dim(Vs(1))l2εp(l+1)+εp(3)εl(2) for all sTZ(G) and equality holds if and only if p, l, s and μ are as in (2), or (3).

We now fix i < j and consider the eigenvalue μiμj1 of s on V. In the case when μiμj1μi1μj, one shows that dim(Vs(μiμj1))l(l+1)2<l2εp(l+1)+εp(3)εl(2). We thus assume that p2 and μiμj1=1. Since the μr’s are distinct, we remark that: {μiμr11 and μi1μr1, where i<rm,rj; and μr1μi1 and μrμi11, where 1r<i;μrμj11 and μr1μj1, where 1r<j,ri; and μj1μr1 and μjμr11, where j<rm.

By (4.6), it follows that dim(Vs(1))dim(V)-dim(Vs(1))2(ni+nj)(l+1ninj). Assume dim(Vs(1))l2εp(l+1)+εp(3)εl(2). Then: (4.7) 2(lninj)(1ninj)r=1mnr2+1+εp(l+1)εp(3)εl(2)0(4.7) and, for it to hold, we must have (lninj)(1ninj)>0, i.e. m = 2 and n1+n2=l+1. Substituting in (4.7) gives (n21)2+n1(2n1)+εp(l+1)εp(3)εl(2)0, and we get n1=2 and n2{1,2}. One checks that the inequality holds only for n2=1. Thus, we have dim(Vs(1))l2εp(l+1)+εp(3)εl(2), where equality holds if and only if p, l, s and μ are as in (1). □

Proposition 4.7.

Let l5 and V=LG(ω3). Then maxsTZ(G)dim(Vs(μ)) (l3)+2.

Proof.

Set λ=ω3 and L=L1. By Lemma 3.7, we have e1(λ)=1, therefore V|[L,L]=V0V1. By [Citation20, Proposition], we have V0LL(ω3) and, since the weight (λα1α2α3)|T1=ω4 admits a maximal vector in V1, by dimensional considerations, it follows that: (4.8) V|[L,L]LL(ω3)LL(ω4).(4.8)

Let sTZ(G). If dim(Vsi(μ))=dim(Vi) for some eigenvalue μ of s on V, where i = 0, 1, then sZ(L)°Z(G) and acts on Vi as scalar multiplication by cl2i(l+1). As cl+11, we have dim(Vs(μ))(l3) for all eigenvalues μ of s on V. We thus assume that dim(Vsi(μ))<dim(Vi) for all eigenvalues μ of s on V and for both i = 0, 1. We write s=z·h, where zZ(L)° and h[L,L]. First, let l=5. Using (4.8) and Proposition 4.4, we determine that dim(Vs(μ))12 for all μ. Therefore, maxsTZ(G)dim(Vs(μ))12. We now assume that l6. By (4.8) and Proposition 4.4, we have dim(Vs(μ))(l1)(l2)2+dim((LL(ω4))h(μh)). Recursively and using the result for l=5, we get dim(Vs(μ))j=5l1j(j1)2+12=(l3)+2 for all eigenvalues μ of s on V. □

Proposition 4.8.

Assume p2,3 and let V=LG(3ω1). Then maxsTZ(G)dim(Vs(μ)) =(l+23)+l.

Proof.

We will apply the algorithm from [Citation5, Section 2.2] (described earlier in this section) to determine a lower bound for νG(V). Afterwards, we will show that this bound is attained. By [Citation14], the sub-dominant weights in V are 3ω1,ω1+ω2 and ω3. Therefore, s3ω1=r3ω1+rω1+ω2+rω3. As Ψ(ω1+ω2) is of type Al2, we have |ΦΨ(ω1+ω2)s|=2(2l1) and |W:W(Ψ(ω1+ω2))|=l(l+1), hence rω1+ω2=2l1. Lastly, using (4.1), we get s3ω1=l2+l+22, therefore νG(V)l2+l+22.

By Lemma 3.7, we have e1(λ)=3, therefore V|[L1,L1]=V0V1V2V3. Now, we argue as we did in the proof of Proposition 4.7 to determine the composition factors of each Vi. It will follow that dim(V0)=1,dim(V1)=l,dim(V2)=(l+12) and dim(V3)=(l+23). Let sZ(L1)°Z(G). We note that s acts on each Vi as scalar multiplication by c3li(l+1). For cl+1=1, we have dim(Vs(c3))=(l+23)+l, therefore codim(Vs(c3))=l2+l+22. Since νG(V)l2+l+22, it follows that maxsTZ(G)dim(Vs(μ)) =(l+23)+l. □

Proof of Theorem 4.2.

To ease notation, we will reference each triplet (V,p,l) featured in by its corresponding number in column “Ref”. Note that 4.2.1–4.2.6 have been solved in Propositions 4.3–4.8. The remaining cases are more straightforward: one has to first determine the structure of V|[L1,L1] and then apply the algorithm of Section 3.3. In what follows, we will only indicate the results that are used in the inductive step of the algorithm and mention if there are small special cases to consider.

4.2.7: If p = 3, the proof is analogous to that of Proposition 4.8. Thus, we assume p3. The result follows recursively. The base case of l=3 follows by Propositions 4.3, 4.5, and 4.6. Note that one has to treat the case when s=z·h, where zZ(L1)° and h[L1,L1] is conjugate to hα2(d)hα3(d2) with d31, separately. For l4, by Propositions 4.3, 4.4, and 4.5, we have dim(Vs(μ))l1+[(l1)2+l1+22(l1)2(l1)+22εp(2)]+(1+εp(2))(l12)+dim((LL1(ω2+ω3))h(μh)). Recursively, it follows that dim(Vs(μ))113εp(2)+j=3l1[j2+j+1] j=3l1εp(2)=l3+2l3lεp(2).

4.2.8: Follows by Proposition 4.7.

4.2.9: Follows by 4.2.8. □

4.1 Supplementary results

At this point, we have completed the proofs of Theorems 4.1 and 4.2. However, we will require additional result for the groups of type Al in the proofs of Theorems 5.1, 5.2, and 7.1, 7.2, respectively. We collect these in Theorems 4.9 and 4.10.

Theorem 4.9.

Let k be an algebraically closed field of characteristic p0 and let G be a simple simply connected linear algebraic group of type Al with l1. Let V=LG(λ), where λX(T)p+ and λ0. We have that

Proof.

Let (V,p,l) be a triplet listed in . Let sTZ(G) and let μ be an eigenvalue of s on V. To calculate upper-bounds for dim(Vs(μ)) in cases 4.3.24.3.8, one will first determine the structure of V|[L1,L1], and then apply the algorithm of Section 3.3. Because the proofs are this straightforward, we only indicate the results that are used in the inductive step of the algorithm and mention if there are small special cases to consider. To ease notation, we will reference each triplet (V,p,l) by its corresponding number in column “Ref”.

Table 8 The supplementary maxsTZ(G)dim(Vs(μ)) for groups of type Al.

4.3.1: The eigenvalues of s on V, not necessarily distinct, are μ1m,μ1m2,,μ1m+2, μ1m, where μ1±1 and, consequently, μ1iμ1i2 for all m+2im.

Assume m is even. If μ±1, then dim(Vs(μ))dim(V)-dim(Vs(1))2m2, as V is self-dual. Let μ = 1. If μ1i=1 for some 2im, then, as μ121 and μ1iμ1i2, at most m24+ε of the eigenvalues μ1m,μ1m2,,μ14 can equal 1, where ε=1ε4(m2). We deduce that dim(Vs(1))ε+m2. If μ=1, then at most m4+ξ of the eigenvalues μ1m,μ1m2,,μ12 can equal –1, where ξ=1ε4(m). We deduce that dim(Vs(1))ξ+m2. The case of m odd is analogous.

To show that equality holds, consider s=diag(μ1,μ11)TZ(G), where μ12=1. Then, since μ12=1, we have dim(Vs(μ1m))=1+m2.

4.3.2: When p = 2 and s=z·h, where zZ(L1)° and h[L1,L1], one shows that the eigenvalues of s on V have the form: c5d±1, c2, c1d±3,c1d±1 and c4d±2, where d1 and ck*. Therefore maxsTZ(G)dim(Vs(μ))=4. When p2, the result follows by Propositions 4.3, 4.5, and 4.8.

4.3.3: When s=z·h with zZ(L1)° and h[L1,L1], one has to show that dim(V2(μh2))52εp(5) for all μh2. The result then follows by Propositions 4.3, 4.5, and 4.8 and 4.3.1.

4.3.4: When p = 2 and s=z·h with zZ(L1)° and h[L1,L1], one has to show that dim((LL1(2ω2+ω3))h(μh))4 for all μh of h. The result then follows by Propositions 4.3, 4.5, 4.6, and 4.3.2.

4.3.5: Follows by Propositions 4.5 and 4.6.

4.3.6: Follows by Proposition 4.8 and 4.3.2.

4.3.7: When s=z·h with zZ(L1)° and h[L1,L1], one has to show that the result holds in two special cases. First case happens when p2 and h conjugate to hα2(d)hα3(d2) with d3=1. This is solved by showing that the eigenvalues of h on V1, respectively on V2, are ±1 with dim(Vh1(1))=8 and dim(V1(1))=10, respectively ±d with dim(Vh2(d))=10 and dim(Vh2(d))=11. The second case occurs when p = 2 and, by Proposition 4.3, 4.3.2 and the structure of V1, we have dim(Vs(μ))13. One has to show that equality does not hold. If it did, then, by Proposition 4.3, h would be conjugate to hα2(d)hα3(d2) with d31, and we would have c7d2=c3d3=c1d2, where ck*. However, since p = 2, we get d3=1, a contradiction. Outside the two special cases, the result follows by Propositions 4.3, 4.5, 4.8, 4.6, and 4.3.2.

4.3.8: Follows by Propositions 4.3, 4.5, 4.6, 4.3.2, and 4.3.3. □

Theorem 4.10.

Let k be an algebraically closed field of characteristic p0 and let G be a simple simply connected linear algebraic group of type Al with l1. Let V=LG(λ), where λX(T)p+ and λ0. We have that

Proof.

Let (V,p,l) be a triple listed in . To begin, note that, by Lemma 3.13, we have maxuGu{1}dim(Vu(1)) =dim(Vxαl(1)(1)). Now, to calculate dim(Vxαl(1)(1)) we use the structure of V|[L1,L1], Lemma 3.12 and the algorithm of Section 3.3. Since the proofs are very similar to the ones of Theorem 4.1, we will only mention the results used in the inductive step of the algorithm.

Table 9 The supplementary maxuGu{1}dim(Vu(1)) for groups of type Al.

4.4.1: Follows from [Citation22, Theorem 1.9].

4.4.2: If p = 2, we have VLG(ω1)(2)LG(ω2), and xα2(1) acts as J2J1 on both LG(ω1)(2) and LG(ω2). If p2, Theorem 4.1[4.1.1, 4.1.3, 4.1.6] gives the result.

4.4.3: Follows by Theorem 4.1[4.1.1, 4.1.3, 4.1.6] and the fact that dim((LL1(4ω2))xα2(1))=1+εp(3).

4.4.4: Follows by Theorem 4.1[4.1.1, 4.1.3, 4.1.7] and 4.4.2.

4.4.5: Follows by Theorem 4.1[4.1.3, 4.1.4].

4.4.6: Follows by Theorem 4.1[4.1.6] and 4.4.2.

4.4.7: When p2, it follows by Theorem 4.1[4.1.1, 4.1.3, 4.1.4, 4.1.6] and 4.4.2. When p = 2, one shows that xα3(1) acts on LL(3ω2) as J24J1, and the result follows by Theorem 4.1[4.1.1] and 4.4.2.

4.4.8: Follows by Theorem 4.1[4.1.1, 4.1.3, 4.1.4], 4.4.2 and 4.4.3. □

5 Proof of Theorem 1.1 for groups of type Cl

Let G be a simple simply connected linear algebraic group of type Cl with l2. This section is dedicated to Theorems 5.1 and 5.2, which give the values of maxuGu{1}dim(Vu(1)) and maxsTZ(G)dim(Vs(μ)) for all kG-modules V=LG(λ) with λX(T)p+,λ0, and dim(LG(λ))4l3. As a corollary, the part of Theorem 1.1 concerning simple simply connected linear algebraic groups of type Cl will follow.

Theorem 5.1.

Let k be an algebraically closed field of characteristic p0 and let G be a simple simply connected linear algebraic group of type Cl with l2. Let V=LG(λ), where λX(T)p+ and λ0, be such that dim(V)4l3. The value of maxuGu{1}dim(Vu(1)) is as given in .

Table 10 The value of maxuGu{1}dim(Vu(1)) for groups of type Cl.

Proof.

Recall that we have denoted by W the natural module of G. To begin, let (V,p,l) be a triplet listed in . Remark that, by Lemma 3.13, we have maxuGu{1}dim(Vu(1)) =maxi=j,ldim(Vxαi(1)(1)), where jl. For 5.1.7 - 5.1.31, to calculate maxi=j,ldim(Vxαi(1)(1)),jl, we use: the structure of either one, or both of V|[L1,L1] and V|[Ll,Ll]; Lemma 3.12 and the algorithm of Section 3.3. When the proofs are very similar to ones of Theorem 4.1, we will only mention the results used in the inductive step.

5.1.1: Note that xα1(1), respectively xαl(1), acts on W as J22J12l4, respectively as J2J12l2.

5.1.2: Note that 2(W)LG(0)V, if εp(l)=0, while if εp(l)=1, we have 2(W)LG(0)|V|LG(0), see [Citation18, Lemma 4.8.2]. By the proof of Theorem 4.1[4.1.2] we get dim((2(W))xαl(1)(1))=2l23l+2. To calculate dim((2(W))xα1(1)(1)), we write W=W1W2, where dim(W1)=4 and xα1(1) acts as J22 on W1; and dim(W2)=2l4 and xα1(1) acts trivially on W2. Using [Citation15, Lemma 3.4] and Lemma 3.14, we determine that dim((2(W))xα1(1)(1))=2l25l+6. Lastly, by [Citation18, Lemma 4.8.2] and [Citation11, Corollary 6.2], or [Citation12, Theorem B] if p = 2, we determine that dim(Vxαl(1)(1)) =2l23l+1εp(l) and dim(Vxα1(1)(1))=2l25l+5εp(l)+εp(2)εl(2).

5.1.3: Follows by [Citation18, Proposition 4.2.2] and Theorem 4.1[4.1.3].

5.1.4: The result for l=3 follows by 5.1.1 and 5.1.2. Assume l4. We note that if εp(l1)=0, we have 3(W)VLG(ω1), while, if εp(l1)=1, then 3(W)LG(ω1)|V|LG(ω1), as kG-modules, see [Citation18, Lemma 4.8.2]. For xα1(1), we argue as in 5.1.2 to calculate dim((3(W))xα1(1)(1)), and we deduce dim(Vxα1(1)(1))(2l13)2l2+9l11(2l2)εp(l1). For xαl(1), using Theorem 4.1[4.1.5], one shows that dim(Vxαl(1)(1))(2l13)1(2l1)εp(l1), where equality holds for εp(l1)=0. To show that equality also holds for εp(l1)=1, we determine the structure of V|[L1,L1] and use 5.1.2.

5.1.5: Follows by [Citation18, Proposition 4.2.2], arguing as in 5.1.2.

5.1.6: Let l=2. When p2, the result follows by Theorem 4.1[4.1.1, 4.1.3]. When p = 2, we have V=LG(ω1)LG(ω2), as kG-modules, see [Citation19, (1.6)]. Using the Jordan form of xαi(1), i = 1, 2, on LG(ω1) and [Citation12, Theorem B], one shows that maxuGu{1}dim(Vu(1))=8.

Let l3. If p = 3, the result follows by 5.1.1, 5.1.3 and the result for l=2. Assume p3. First, we argue recursively, using 5.1.1, 5.1.2, 5.1.3 and the result for l=2, to show that dim(Vxαi(1)(1))2(2l3)(2l1)εp(2l+1)+2lεp(2),i=l1,l. Lastly, assume that p2. By [Citation18, Lemmas 4.9.1 and 4.9.2], we know that V is a composition factor of the kG-module W2(W), where ch(W2(W))=χ(ω1+ω2)+χ(ω3)+2χ(ω1). Note that by χ(λ) we understand the character of the Weyl kG-module of highest weight λX(T). Therefore, in view of [Citation18, Lemmas 4.8.2 and 4.9.2], we have dim(Vu(1))dim((W2(W))u(1))-dim((LG(ω3))u(1))(2+εp(l1)+εp(2l+1))dim((LG(ω1))u(1)) for all uGu. For xαl(1), it follows that dim(Vxαl(1)(1))2(2l3)(2l1)εp(2l+1).

5.1.7: Follows by Theorem 4.1[4.1.1, 4.1.3, 4.1.6].

5.1.8: Follows by Theorems 4.1[4.1.1, 4.1.3, 4.1.6] and 4.10[4.4.1].

5.1.9: Because we will need to know maxuGu{1}dim((LG(3ω2))u(1)) also when p7, we will not limit ourselves to the case p = 7 and, instead, we will only assume that p2,3. In this case, by Theorems 4.1[4.1.3, 4.1.6] and 4.10[4.4.1], we get maxuGu{1}dim((LG(3ω2))u(1)) 10+2εp(5)3εp(7), where equality holds for p5.

5.1.10: As above, we will only assume that p2. Then, by Theorems 4.1[4.1.1, 4.1.3, 4.1.6] and 4.10[4.4.1], we deduce that maxuGu{1}dim(Vu(1))154εp(3).

5.1.11: Follows by 5.1.1, 5.1.2, 5.1.3, and 5.1.6.

5.1.12: Follows by 5.2.6 and 5.2.7.

5.1.13: As in 5.1.9, we only assume p2. Then, by 5.1.3, 5.1.6, and 5.1.7, we determine that maxuGu{1}dim((LG(2ω3))u(1)) 4015εp(5)+4εp(3), where equality holds for p3.

5.1.14: Again, we only assume p2. Then, by 5.1.1, 5.1.2, 5.1.3, 5.1.6, and 5.1.7, we determine that maxuGu{1}dim(Vu(1)) 50εp(7).

5.1.15: Follows by 5.1.2 and 5.1.4.

5.1.16: Follows recursively, using 5.1.11 and 5.1.30.

5.1.17: Follows by 5.1.4, 5.1.6 and 5.1.11.

5.1.18: Follows by 5.1.1, 5.1.2, 5.1.6, and 5.1.11.

5.1.19: Follows by 5.1.2, 5.1.4, 5.1.15, and 5.1.30.

5.1.20: Follows by 5.1.4 and 5.1.15.

5.1.21: Follows by 5.1.2, 5.1.4 and 5.1.19.

5.1.22: Follows by 5.1.4, 5.1.19, 5.1.20, and 5.1.30.

5.1.23: Follows by 5.1.19 and 5.1.20.

5.1.24: Follows by 5.1.2, 5.1.4 and 5.1.21.

5.1.25 and 5.1.26: Follow by 5.1.22 and 5.1.23.

5.1.27: Follows by 5.1.21 and 5.1.22.

5.1.28: Follows by 5.1.2, 5.1.4, and 5.1.24.

5.1.29: Follows by 5.1.2, 5.1.4, and 5.1.28.

5.1.30: Follows recursively, using 5.1.4.

5.1.31: Note that VLG(ω1)(2)LG(ωl) and LG(ω1+ωl)LG(ω1)LG(ωl), by [Citation19, (1.6)]. The result follows by 5.1.11 and 5.1.16. □

Theorem 5.2.

Let k be an algebraically closed field of characteristic p0 and let G be a simple simply connected linear algebraic group of type Cl with l2. Let V=LG(λ), where λX(T)p+ and λ0, be such that dim(V)4l3. The value of maxsTZ(G)dim(Vs(μ)) is as given in .

Table 11 The value of maxsTZ(G)dim(Vs(μ)) for groups of type Cl.

Let sTZ(G). As in Section 4, to improve readability, we will say “s is as in (Hs)” to express the fact that s satisfies the following: s=diag(μ1·In1,,μm·Inm,μm1·Inm,,μ11·In1), where μiμj±1 for all i < j, i=1mni=l and n1nm1; and if m = 1, then μ1±1. Note that any sTZ(G) is as in (Hs).

Proposition 5.3.

Let V=LG(ω1). Then maxsTZ(G)dim(Vs(μ)) =2l2, where the maximum is attained if and only if μ=±1 and s is conjugate to diag(±1,, ±1,d,d1,±1,,±1) with d±1.

Proof.

The proof is analogous to that of Proposition 4.3. □

Proposition 5.4.

Let V=LG(ω2). Then

  1. l=2 and maxsTZ(G)dim(Vs(μ))=42εp(2), where the maximum is attained if and only if

  • (1.1) p2,μ=1 and s is conjugate to diag(1,1,1,1).

  • (1.2) p = 2, μ = 1 and s is conjugate to diag(d,d,d1,d1) with d1.

  • (1.3) p = 2, μ=d±1 and s is conjugate to diag(d,1,1,d1) with d1.

  1. l=3 and maxsTZ(G)dim(Vs(μ))=8, where the maximum is attained if and only if

  • (2.1) p3, μ = 1 and s is conjugate to diag(d,d,d,d1,d1,d1) with d21.

  • (2.2) p2,μ=1 and s is conjugate to ±diag(1,1,1,1,1,1).

  1. l=4 and maxsTZ(G)dim(Vs(μ))=162εp(2), where the maximum is attained if and only if

  • (3.1) p2,μ=1 and s is conjugate to diag(1,1,1,1,1,1,1,1).

  • (3.2) p = 2, μ = 1 and s is conjugate to diag(d,d,d,d,d1,d1,d1,d1) with d1.

  • (3.3) p = 2, μ = 1 and s is conjugate to diag(1,1,1,d,d1,1,1,1) with d1.

  1. l5 and maxsTZ(G)dim(Vs(μ))=2l25l+3εp(l), where the maximum is attained if and only if μ = 1 and s is conjugate to ±diag(1,,1,d,d1,1,,1) with d1.

Proof.

To ease notation, define Bl,l2, in the following way: B2=42εp(2); B3=8; B4=162εp(2); and Bl=2l25l+3εp(l) for all l5. Let sTZ(G) be as in (Hs). Using the structure of 2(W) as a kG-module, see [Citation18, Lemma 4.8.2], we deduce that the eigenvalues of s on V, not necessarily distinct, have one of the following forms: (5.1) {μi2 and μi2,1im, each with multiplicity at least ni(ni1)2;μiμj and μi1μj1,1i<jm, each with multiplicity at least ninj;μiμj1 and μi1μj,1i<jm, each with multiplicity at least ninj;1 with multiplicity at least i=1mni21εp(l).(5.1)

Let μk* be an eigenvalue of s on V. If μμ1, then, as V is self-dual, we have dim(Vs(μ))dim(V)-dim(Vs(1))2dim(V)(l1εp(l))2Bl, where equality holds if and only if l, p, s and μ are as in (1.3). Thus, for the remainder of the proof, we assume that μ=±1.

Let m = 1, i.e. n1=l and μ121. Using (5.1), we show that dim(Vs(±1))Bl, where equality holds if and only if l, p, s and μ are as in (1.2), (2.1), and (3.2). Thus, for the remainder of the proof, we assume that m2.

If l=2, then m = 2, i.e. n1=n2=1. Using (5.1), we determine that dim(Vs(±1))B2, where equality holds if and only if p, s and μ are as in (1.1). If l=3, then m3. If m = 2, i.e. n1=2 and n2=1, then dim(Vs(±1))B3 where equality holds if and only if p, s and μ are as in (2.2). If m = 3, by (5.1), we get dim(Vs(1))=2εp(l); and dim(Vs(1))12. Further, as –1 can equal at most one eigenvalue of the form μiμj and at most one of the form μiμj1, we deduce that dim(Vs(1))4. For the remainder of the proof, we assume that l4.

For the eigenvalue μ = 1, as μiμj±1 for all i<j, we have μi±1μj±11 for all i < j, thus: (5.2) dim(Vs(1))2l2l1εp(l)4i<jninj.(5.2)

Let l=4 and assume that dim(Vs(1))B4. It follows that, 11+εp(l)4i<jninj0 and, as r=1mnr=4,m2 and nrnq for all r < q, we get i<jninj3. Thus, 11+εp(l)4i<jninj0 holds if and only if p = 2, m = 2, n1=3 and n2=1. Substituting in (5.2), gives dim(Vs(1))14, where equality holds if and only if s is as in (3.3). We now let l5 and assume that dim(Vs(1))Bl. By (5.2), we get: (5.3) l1i<jninj0.(5.3)

As l=i=1mni, it follows that i=1m2ni(1i<jnj)+(nm11)(1nm)0. But i=1m2ni(1i<jnj)0 and (nm11)(1nm)0, as ni1 for all 1im, therefore inequality (5.3) holds if and only if m = 2, n2=1 and n1=l1. Substituting in (5.2), gives dim(Vs(1))Bl, where equality holds if and only if all eigenvalues of s on V different to μ1±1μ2±1 are equal to 1, i.e. if and only if s is as in (4).

Lastly, let μ=1. We remark that dim(Vs(1))dim(V)-dim(Vs(1))2l2lr=1mnr2, see (5.1). If μiμj1 for all i < j, we have dim(Vs(1))2l2lr=1mnr22i<jninj =l2l<Bl. We thus assume that there exist i < j such that μiμj=1. Then μi1μj1=1 and, since the μi’s are distinct, we have that: (5.4) {μi21 and μj21, hence μi21 and μj21;μiμr1 and μi1μr11, where i<rm,rj; and μrμi1 and μr1μi11, where 1r<i;μrμj1 and μr1μj11, where 1r<j,ri; and μjμr1 and μj1μr11, where j<rm.(5.4)

By (5.1), all of the above account for at least ni(ni1)+nj(nj1)+2(ni+nj)(lninj) additional eigenvalues of s on V different to –1. This gives: (5.5) dim(Vs(1))2l2lr=1mnr2ni(ni1)nj(nj1)2(ni+nj)(lninj).(5.5)

Let l=4 and assume dim(Vs(1))B4. Then (ni,nj){(3,1),(2,2),(2,1),(1,1)} and we have 12ri,jnr27(ni+nj)+4ninj0. For (ni,nj){(3,1),(2,1)} the inequality does not hold. For (ni,nj)=(2,2) we get dim(Vs(1))16, see (5.5), where equality holds if and only if s is as in (3.1). For (ni,nj)=(1,1) we have dim(Vs(1))18ri,jnr2, see (5.5). Since we are assuming dim(Vs(1))B4, we must have nr = 1 for all ri,j, therefore m = 4 and ni = 1 for all 1i4. Substituting in (5.5) gives dim(Vs(1))16, where equality holds if and only if all eigenvalues of s on V different to 1 and the ones listed in (5.4) are equal to –1. However, as at most one eigenvalue of the form μiμr1,ri,j can equal –1, we see that the condition for equality cannot be satisfied. This completes the case of l=4. Thus, let l5 and assume dim(Vs(1))Bl. It follows that: (5.6) l(4ninj)3+εp(l)ri,jnr2(ninj)2(ni+nj)(lninj1)0.(5.6)

If lninj1<0, then, as r=1mnr=l, we have m = 2 and so l=n1+n2. Substituting in (5.6) gives l(5l)3+εp(l)ri,jnr2(2n1l)20, which does not hold as l5. If lninj10, then, for (5.6) to hold, we must have l(4ninj)>0, hence (ni,nj){(2,1),(1,1)}. If (ni,nj)=(2,1), inequality (5.6) does not hold. If (ni,nj)=(1,1), substituting in (5.6) gives 3+εp(l)ri,jnr20. One checks that this inequality holds if and only if l{5,6}, nr = 1 for all 1rl, and εp(6)=1 when l=6. In both cases, we can assume without loss of generality that μ1μ2=1. As the μr’s are distinct, at most one eigenvalue of each of the forms μ1μr1,μ11μr, μ2μr1 and μ21μr,3rl, can equal –1. This gives an additional 4(l3) eigenvalues of s on V that are different to –1. Consequently, we have dim(Vs(1))2l210l+20<Bl. This completes the proof. □

Proposition 5.5.

Assume p2 and let V=LG(2ω1). Then maxsTZ(G)dim(Vs(μ))=2l23l+4, where the maximum is attained if and only if

  1. l=2,μ=1 and s is conjugate to diag(d,d,d1,d1) with d2=1.

  2. l2, μ = 1 and s is conjugate to ±diag(1,,1,1,1,1,,1).

Proof.

Let sTZ(G) be as in (Hs). We note that VS2(W), see [Citation18, Proposition 4.2.2], therefore the eigenvalues of s on V, not necessarily distinct, have one of the following forms: (5.7) {μi2 and μi2,1im, each with multiplicity at least ni(ni+1)2;μiμj and μi1μj1,1i<jm, each with multiplicity at least ninj;μiμj1 and μi1μj,1i<jm, each with multiplicity at least ninj;1 with multiplicity at least i=1mni2.(5.7)

The cases of μμ1; sTZ(G) with m = 1; and sTZ(G) with m2 and μ = 1 are handled as in the proof of Proposition 5.4. The only case we have left to consider is sTZ(G) with m2 and μ=1. We note that dim(Vs(1))2l2+li=1mni2. If μiμj1 for all i < j, then, by (5.7), there are at least 2i<jninj additional eigenvalues of s on V different to –1. This gives: (5.8) dim(Vs(1))2l2+li=1mni22i<jninj=2l2+l(i=1mni)2=l2+l.(5.8)

Therefore dim(Vs(1))2l23l+4, where equality holds if and only if l=2 and all eigenvalues of s on V different to 1, μ1μ2 and μ11μ21 are equal to –1. But then, by (5.7), we must have μ12=μ1μ21, a contradiction. We thus assume that there exist i < j such that μiμj=1. Then μi1μj1=1, and: {μi21,μi21 and μj21,μj21;μiμr1 and μi1μr11, where i<rm,rj; and μrμi1 and μr1μi11, where 1r<i;μjμr1 and μj1μr11, where j<rm; and μrμj1 and μr1μj11, where 1r<j,ri.

By (5.7), these amount to at least ni(ni+1)+nj(nj+1)+2(ni+nj)(lninj) additional eigenvalues of s on V different to –1. This gives: (5.9) dim(Vs(1))2l2+lr=1mnr2ni(ni+1)nj(nj+1)2(ni+nj)(lninj).(5.9)

If dim(Vs(1))2l23l+4, it follows that l(4ninj)ri,jnr2(ninj)2(ni+nj)(l+1ninj)40,

which does not hold. This completes the proof of the proposition. □

Proposition 5.6.

Let l=4 and V=LG(ω3). Then maxsTZ(G)dim(Vs(μ))=282εp(3)8εp(2).

Proof.

First, one determines the structure of V|[L1,L1] and then applies the algorithm from Section 3.3, using Theorem 5.2[5.2.4 for l=3] and Propositions 5.4 and 5.5. For p = 3, it follows that maxsTZ(G)dim(Vs(μ))=26, while, for p3, we get maxsTZ(G)dim(Vs(μ)) 302εp(2) and that there exist (s,μ)(Z(L)°Z(G))×k* with dim(Vs(μ))=288εp(2). In what follows we will show that maxsTZ(G)dim(Vs(μ))=288εp(2).

Assume there exist (s,μ)(TZ(G))×k* with dim(Vs(μ))>288εp(2). Then sZ(L1)°, and we write s=z·h with zZ(L1)° and h[L1,L1]. If p2, by Proposition 5.4, h must be conjugate to one of {diag(1,1,1,1,1,1,1,1), diag(1,1,1,1,1,1,1,1), diag(1,d,d,d,d1,d1,d1)  with d21}. In all cases, one shows that dim(Vs(μ))28 for all μ. Let p = 2. In view of Proposition 5.4, assume h is conjugate to one of {diag(1,1,1,d,d1,1,1,1) with d1,diag(1,d,d,d,d1,d1, d1,1) with d1,diag(1,d,d,e,e1,d1,d1,1) with d1,e±1}. Using the weight structure of V|[L1,L1], one shows that dim(Vs(μ))20 for all μ. On the other hand, if h belongs to a different conjugacy class, then by Propositions 5.4, 5.5 and the weight structure of V0, it follows that dim(Vs(μ))18 for all μ. □

Proposition 5.7.

Let l5 and V=LG(ω3). Then maxsTZ(G)dim(Vs(μ))(2l23)+102(l1)εp(l1) +4εp(3)10εp(2).

Proof.

To prove the result, we determine the structure of V|[L1,L1] and apply the algorithm of Section 3.3. Note that the case when sZ(L1)° is handled recursively. We write s=z·h, where zZ(L1)° and h[L1,L1]. For l=5, we use Propositions 5.3, 5.4, and 5.6, to show that maxsTZ(G)dim(Vs(μ))66+4εp(3)18εp(2). For l6, using the result for l=5, one shows that: dim(Vs(μ))4(l1)28(l1)+42(l1)εp(l1)+(2(l1)2)εp(l2)+dim((LL(ω4))h(μh))4j=5l1j28j=5l1j+j=5l14j=5l12jεp(j)+j=5l12(j1)εp(j1)+66+4εp(3)18εp(2)=(2l23)+102(l1)εp(l1)+4εp(3)10εp(2).

Proof of Theorem 5.2.

The proofs of 5.2.1, 5.2.2, and 5.2.3 are covered in Propositions 5.3, 5.4, and 5.5. We have made this choice as they require more in-depth analysis. The proofs of the remaining results are much more straightforward: for each triplet (V,p,l) we first determine the structure of V|[Li,Li], i = 1 or i = 2, and then apply the algorithm of Section 3.3.

5.2.4: Propositions 5.3 and 5.4 give the result for l=3, while Proposition 5.6, respectively 5.7, gives the result for l=4, respectively l5.

5.2.5: The proof is analogous to that of Proposition 5.7, using Propositions 5.3 and 5.5.

5.2.6: Let l=2. If p = 5, the result follows from Propositions 4.3, 4.5, and 4.8. If p = 2, one identifies the eigenvalues of s on V and determines that maxsTZ(G)dim(Vs(μ))=6. If p2,5, one uses the weight structure of V1 and Proposition 4.3 to prove the result. Assume l3. If p = 3, the result follows recursively, using Propositions 5.3, 5.5, and 5.2.6. If p3, the proof is analogous to that of Proposition 5.7 and uses 5.2.4 and 5.2.30.

5.2.7: Follows by Propositions 4.3, 4.5, and 4.8.

5.2.8: Follows by Propositions 4.5 and 4.8. Note that when sZ(L1)°, one has to show that there do not exist μ with dim(Vs(μ))>12. If this were the case, then, as V is self-dual, we would have μ=±1 and dim(Vsi(±1))=2 for all 0i6,i3. We write s=z·h with zZ(L1)° and h[L1,L1]. Then dim(Vhi(μhi))=2 for all i3, where μhi=μ·ci3. In particular, by Propositions 4.5 and 4.8, we have μh2=1, thereby c2=1, and μh1=d±1 with d2=1, thereby μ=d±1, contradicting μ=±1.

5.2.9: Because we will need to know maxsTZ(G)dim((LG(3ω2))s(μ)) also when p7, we will not limit ourselves to the case p = 7 and, instead, only assume that p2,3. Using Propositions 4.5, 4.8, and Theorem 4.9[4.3.1], we deduce that maxsTZ(G)dim(Vs(μ))=20εp(7).

5.2.10: As above, we only assume p2. Using Proposition 4.5 we deduce that maxsTZ(G)dim(Vs(μ))=204εp(3). Note that when sZ(L2)° (s=z·h, where zZ(L2)° and h[L2,L2]), one has to show that dim(Vh1(μh1))5εp(3) and dim(Vh2(μh2))62εp(3) for all μhi, i = 1, 2. Lastly, equality is shown to hold for s=diag(1,1,1,1)TZ(G) and μ=1.

5.2.11: Follows by Propositions 5.3, 5.4, 5.5, and 5.2.6. Note that when sZ(L1)°, one has to show that dim(Vh2(μh2))124εp(3)8εp(2) for all μh2, where s=z·h with zZ(L1)° and h[L1,L1].

5.2.12: Follows by 5.2.6 and 5.2.7.

5.2.13: As in 5.2.9, we only assume p2. Using Proposition 5.5, 5.2.6, and 5.2.7, we deduce that maxsTZ(G)dim((LG(2ω3))s(μ))=5213εp(5).

5.2.14: We only assume p2. Using Propositions 5.3, 5.4, 5.5, 5.2.6, and 5.2.7, we deduce that maxsTZ(G)dim(Vs(μ))=50εp(7). Note that one has to show that there do not exist (s,μ)(TZ(G))×k* with dim(Vs(μ))>50εp(7). If this were the case, then sZ(L1)° (s=z·h with zZ(L1)° and h[L1,L1]) and h would be conjugate to one of {diag(1,1,1,1,1,1),diag(1,d,d,d1,d1,1) with  d2=1}. In both cases, one shows that dim(Vs(μ))50εp(7).

5.2.15: Follows by Proposition 5.4 and 5.2.4.

5.2.16: Follows recursively, using 5.2.11 and 5.2.30.

5.2.17: Follows by 5.2.4, 5.2.6, and 5.2.11.

5.2.18: Follows by Propositions 5.3, 5.4, 5.2.6, and 5.2.11.

5.2.19: Follows by Propositions 5.4, 5.6, 5.2.30, and 5.2.15.

5.2.20: Follows by Proposition 5.6 and 5.2.15.

5.2.21: Follows by Propositions 5.4, 5.7 and 5.2.19.

5.2.22: Follows by Proposition 5.7, 5.2.19, 5.2.20, and 5.2.30.

5.2.23: Follows by 5.2.19 and 5.2.20.

5.2.24: Follows by Propositions 5.4, 5.7, and 5.2.21.

5.2.25 and 5.2.26: Follow by 5.2.22 and 5.2.23.

5.2.27: Follows by 5.2.21 and 5.2.22.

5.2.28: Follows by Propositions 5.4, 5.7, and 5.2.24.

5.2.29: Follows by Propositions 5.4, 5.7, and 5.2.28.

5.2.30: Follows by 5.2.4.

5.2.31: Proof is analogous to that of Proposition 5.7, and uses 5.2.4 and 5.2.30. □

6 Proof of Theorem 1.1 for groups of type Bl

Let k be an algebraically closed field of characteristic p2 and let G, respectively G˜, be a simple adjoint, respectively simply connected, linear algebraic group of type Bl with l3. We fix a central isogeny ϕ:G˜G with ker(ϕ)Z(G˜) and 0, and let T˜, respectively B˜, be a maximal torus, respectively a Borel subgroup, in G˜ with the property that ϕ(T˜)=T, respectively ϕ(B˜)=B. We denote by X(T˜),Z(G˜),G˜u, Δ˜={α˜1,,α˜l} and ω˜1,,ω˜l the character group of T˜, the center of G˜, the set of unipotent elements in G˜, the set of simple roots in G˜ corresponding to B˜, and the fundamental dominant weights of G˜ corresponding to Δ˜. We denote by L˜ia Levi subgroup of the maximal parabolic subgroup P˜i of G˜ corresponding to Δ˜i=Δ˜{α˜i}, 1il, and we let T˜i=T˜[L˜i,L˜i]. Now, for λ˜X(T˜)a p-restricted dominant weight, we let V˜=LG˜(λ˜) and we have V˜|[L˜i,L˜i]=j=0ei(λ˜)V˜j, where V˜j=γ˜N0Δ˜iV˜λ˜jα˜iγ˜, 0jei(λ˜).

This section is dedicated to Theorems 6.1 and 6.2 which give the values of maxu˜G˜u{1}dim(V˜u˜(1)) and maxs˜T˜Z(G˜)dim(V˜s˜(μ˜)) for all kG˜-modules V˜=LG˜(λ˜), with λ˜X(T˜)a nonzero p-restricted dominant weight, and dim(LG˜(λ˜))4l3. As a corollary, the part of Theorem 1.1 concerning simple simply connected linear algebraic groups of type Bl will follow.

Theorem 6.1.

Let k be an algebraically closed field of characteristic p2 and let G˜ be a simple simply connected linear algebraic group of type Bl with l3. Let V˜=LG˜(λ˜), where λ˜X(T˜) is a nonzero p-restricted dominant weight, be such that dim(V˜)4l3. The value of maxu˜G˜u{1}dim(V˜u˜(1)) is as given in .

Table 12 The value of maxu˜G˜u{1}dim(V˜u˜(1)) for groups of type Bl.

Proof.

For 6.1.16.1.5 we will deduce the result for V˜=LG˜(λ˜) by calculating maxuGu{1}dim((LG(λ))u(1)) =maxi=1,ldim((LG(λ))xαi(1)(1)), see Section 3.1 and Lemmas 3.2 and 3.13. The proofs for 6.1.6 - 6.1.17 are much more straightforward: first one determines the structure of V˜|[L˜1,L˜1], and then applies the algorithm of Section 3.3. Once more, by Lemma 3.13, we have maxu˜G˜u{1}dim(V˜u˜(1))=maxi=l1,ldim(V˜xα˜j(1)(1)).

6.1.1: Note that V˜LG(ω1) as kG˜-modules and LG(ω1)W as kG-modules. We have that xα1(1), respectively xαl(1), acts on W as J3J12l2, respectively as J22J12l3.

6.1.2: Note that V˜LG(ω2) as kG˜-modules and LG(ω2)2(W) as kG-modules, see [Citation18, Proposition 4.2.2]. For xα1(1), we write W=W1W2, where dim(W1)=3 and xα1(1) acts as J3 on W1; and dim(W2)=2l2 and xα1(1) acts trivially on W2. Using [Citation15, Lemma 3.4], we show that dim((LG(ω2))xα1(1)(1))=2l23l+2. Similarly, for xαl(1), we write W=W1W2, where dim(W1)=4 and xαl(1) acts as J22 on W1; and dim(W2)=2l3 and xαl(1) acts trivially on W2. One shows that dim((LG(ω2))xαl(1)(1))=2l23l+4.

6.1.3: Note that V˜LG(2ω1) as kG˜-modules. Moreover, by [Citation18, Proposition 4.7.3], we have S2(W)LG(2ω1)LG(0) if εp(2l+1)=0, and S2(W)LG(0)|LG(2ω1)|LG(0) if εp(2l+1)=1. We argue as in 6.1.2 to show that dim((S2(W))xαi(1)(1))=2l2l+1,i=1,l. Then, dim((LG(2ω1))xαi(1)(1))=2l2lεp(2l+1),i=1,l, by [Citation11, Corollary 6.3].

6.1.4: For l=3, we use Theorem 5.1[5.1.1]. Assume l4. Note that V˜LG(ω3) as kG˜-modules and that 3(W)LG(ω3) as kG-modules, see [Citation18, Proposition 4.2.2]. We now argue as in 6.1.2 to obtain the result.

6.1.5: Note that V˜LG(3ω1) as kG˜-modules and that S3(W)LG(3ω1)W if εp(2l+3)=0, respectively S3(W)W|LG(3ω1)|W if εp(2l+3)=1, see [Citation18, Propositions 4.7.4]. We argue in 6.1.2 to show that maxuGu{1}dim((LG(3ω1))u(1)) 4l3l3(2l1)εp(2l+3). Now, one has to establish the structure of LG(3ω1)|[L1,L1] and apply the algorithm of Section 3.3. Equality will follow recursively, using 6.1.1 and 6.1.3.

6.1.6: For l=3, it follows by Theorem 5.1[5.1.2, 5.1.3, 5.1.7, 5.1.10]. For l4, it follows recursively by 6.1.1, 6.1.2, 6.1.3 and the result for l=3.

6.1.7: Follows by Theorem 5.1[5.1.2, 5.1.3].

6.1.8, 6.1.9, and 6.1.10: Follow by Theorem 5.1[5.1.5, 5.1.10].

6.1.11: Follows by 6.1.2, 6.1.4, and 6.1.7.

6.1.12 and 6.1.13: Follow by 6.1.2, 6.1.4, and 6.1.11.

6.1.14: Follows by 6.1.2, 6.1.4, and 6.1.12.

6.1.15: Follows by 6.1.2, 6.1.4, and 6.1.14.

6.1.16: Follows recursively, by 6.1.17 and Theorem 5.1[5.1.1, 5.1.6] for the base case.

6.1.17: Follows recursively, using 6.1.4 to prove the base case. □

Theorem 6.2.

Let k be an algebraically closed field of characteristic p2 and let G˜ be a simple simply connected linear algebraic group of type Bl with l3. Let V˜=LG˜(λ˜), where λ˜X(T˜) is a nonzero p-restricted dominant weight, be such that dim(V˜)4l3. The value of maxs˜T˜Z(G˜)dim(V˜s˜(μ˜)) is as given in .

Table 13 The value of maxs˜T˜Z(G˜)dim(V˜s˜(μ˜)) for groups of type Bl.

Before we begin, we recall that G is a simple adjoint linear algebraic group of type Bl, and that we have denoted by W the natural module of G. Let sTZ(G). As in Section 4, we will say“s is as in (Hs)” to mean that s satisfies the following: s=diag(μ1·In1,,μm·Inm,1·In,μm1·Inm,,μ11·In1) with μiμj±1,i<j,μi1,1im and n+2i=1mni=2l+1, where 1n2l1 and ln1nm1. Note that every sTZ(G) is as in (Hs).

Proposition 6.3.

Let V˜=LG˜(ω˜1). Then maxs˜T˜Z(G˜)dim(V˜s˜(μ˜))=2l.

Proof.

Set V=LG(ω1). We note that V˜V as kG˜-modules and that VW as kG-modules. The proof now follows that of Proposition 4.3. □

Proposition 6.4.

Let V˜=LG˜(ω˜2). Then maxs˜T˜Z(G˜)dim(V˜s˜(μ˜))=2l2l.

Proof.

Set V=LG(ω2) and note that V˜V as kG˜-modules. Further, by [Citation18, Proposition 4.2.2], we have V2(W) as kG-modules. Let sTZ(G) be as in (Hs). The eigenvalues of s on V, not necessarily distinct, have one of the following forms: (6.1) {μi2 and μi2, where 1im, each with multiplicity at least ni(ni1)2;μiμj and μi1μj1, where 1i<jm, each with multiplicity at least ninj;μiμj1 and μi1μj, where 1i<jm, each with multiplicity at least ninj;μi and μi1, where 1im, each with multiplicity at least nni;1 with multiplicity at least n(n1)2+r=1mnr2.(6.1)

Let μk* be an eigenvalue of s on V. If μμ1, one shows that dim(Vs(μ))2l2+l2<2l2l. We thus assume for the remainder of the proof that μ=±1.

Assume m = 1. Then μ11, as sZ(G). By (6.1), the eigenvalues of s on V, not necessarily distinct, are μ1±2 each with multiplicity at least n1(n11)2; μ1±1 each with multiplicity at least nn1; and 1 with multiplicity at least n(n1)2+n12. Let μ = 1. Since n=2l+12n1, we have dim(Vs(1))2l2l2(ln1)(2n11)2l2l, where equality holds if and only if n1=l and μ1=1. Now, let μ=1. If μ1=1, then dim(Vs(1))=2nn1, while, if μ12=1, then dim(Vs(1))=n1(n11), therefore dim(Vs(1))<2l2l for all s with m = 1. We thus assume that m2.

Let μ = 1. Since μiμj±1 for all i<j, by (6.1), we determine that dim(Vs(1))2l2+l4i<jninj 2ni=1mni. Assume dim(Vs(1))2l2l. Then, as 2i=1mni=2l+1n, it follows that (6.2) (2ln)(1n)4i<jninj0,(6.2) which does not hold. Therefore dim(Vs(1))<2l2l for all sTZ(G) with m2.

Lastly, let μ=1. If μi1 for all i, then dim(Vs(1))2l2+ln(n1)2r=1mnr22nr=1mnr. Suppose that dim(Vs(1))2l2l. Then 2ln(n1)2r=1mnr22nr=1mnr0. Since 2r=1mnr= 2l+1n, we must have (2ln)(1n)n(n1)2r=1mnr20, which does not hold. We thus assume there exist i such that μi=1. Then, since the μi’s are distinct and different to 1, by (6.1), we determine that dim(Vs(1))2l2+ln(n1)2 r=1mnr22nrinr4nirinrni(ni1). Suppose dim(Vs(1))2l2l. Then: (6.3) 2ln(n1)2r=1mnr22nrinr4nirinrni(ni1)0.(6.3)

We have that r=1mnr2r=1mnr, as nr1, and that 2l=2r=1mnr+n1. Substituting in (6.3) gives rinr(12n)+ ni(2ni4rinr)(n1)(n2)20, which does not hold. Therefore dim(Vs(1))<2l2l for all sTZ(G) with m2. □

Proposition 6.5.

Let V˜=LG˜(2ω˜1). Then maxs˜T˜Z(G˜)dim(V˜s˜(μ˜))=2l2+lεp(2l+1).

Proof.

Set V=LG(2ω1). Note that V˜V as kG˜-modules and that S2(W)VLG(0) if εp(2l+1)=0; while S2(W)LG(0)|V|LG(0) if εp(2l+1)=1, see [Citation18, Proposition 4.7.3]. Let sTZ(G) be as in (Hs). The eigenvalues of s on V, not necessarily distinct, have one of the following forms: (6.4) {μi2 and μi2,1im, each with multiplicity at least ni(ni+1)2;μiμj and μi1μj1,1i<jm, each with multiplicity at least ninj;μiμj1 and μi1μj,1i<jm, each with multiplicity at least ninj;μi and μi1,1im, each with multiplicity at least nni;1 with multiplicity at least r=1mnr2+n(n+1)21εp(2l+1).(6.4)

The cases μ=μ1 and sTZ(G) with m = 1 are handled as in the proof of Proposition 6.4. The only case we have left is that of sTZ(G) with m2. For μ = 1, as μi±1μj±11 for all i<j, by (6.4), we determine that dim(Vs(1))2l2+3lεp(2l+1)2ni=1mni4i<jninj<2l2+lεp(2l+1), see (6.2). Thus, let μ=1. Suppose that μi1 for all i. Then, by (6.4), we have dim(Vs(1))2l2+3l+1 r=1mnr2 n(n+1)2 2nr=1mnr. One shows, as in the corresponding case in the proof of Proposition 6.4, that dim(Vs(1))<2l2+lεp(2l+1). Lastly assume that there exists i such that μi=1. Then μr±11 for all ri and, by (6.4), we get dim(Vs(1))2l2+3l+1r=1mnr2 n(n+1)2 2nrinrni(ni+1). Once more, as in the proof of Proposition 6.4, we show that dim(Vs(1)) <2l2+lεp(2l+1). □

Proof of Theorem 6.2.

The results 6.2.1, 6.2.2, and 6.2.3 are covered in Propositions 6.3, 6.4, and 6.5, respectively. The proofs for 6.2.46.1.17 are much more straightforward: first one determines the structure of V˜|[L˜1,L˜1] and then applies the algorithm of Section 3.3.

6.2.4: The case l=3 follows by Proposition 5.3, while the case l=4 follows by Propositions 6.3, 6.4 and 6.2.7. Note that in the latter one has to show that for s˜=hα˜1(1)hα˜3(1)hα˜4(d) with d2=1, we have dim(V˜s˜(1))=56. The case l5 follows recursively, by Propositions 6.3, 6.4 and the result for l=4. Moreover, one shows that for l even and s˜=hα˜1(1)hα˜3(1)hα˜l1(1)hα˜l(c) with c2=1, respectively l odd and s˜=hα˜1(1)hα˜3(1) hα˜l2(1)hα˜l(c) with c2=1, we have dim(V˜s˜(1))=4l36l2+2l3.

6.2.5: For l=3, it follows by Proposition 5.4 and Theorem 5.2[5.2.7, 5.2.9]. Also, one shows that equality holds for s˜=hα˜1(1)hα˜3(c) with c2=1 and μ˜=1. For l4, the result is obtained recursively, using Propositions 6.3, 6.5 and the result for l=3. Moreover, one shows that equality holds for μ˜=1,l even and s˜=hα˜1(1)hα˜3(1)hα˜l1(1)hα˜l(c) with c2=1, respectively l odd and s˜=hα˜1(1)hα˜3(1) hα˜l2(1)hα˜l(c) with c2=1.

6.2.6: For l=3, it follows by Propositions 5.4, 5.5 and Theorem 5.2[5.2.7, 5.2.10]. For l4, it follows recursively from 6.3, 6.4, 6.5 and the result for l=3.

6.2.7: Follows by Propositions 5.4 and 5.5. Note that when s˜=z˜·h˜, where z˜Z(L˜1)° and h˜[L˜1,L˜1], one has to first show that dim(V˜h˜1(μ˜h˜1))8 for all μ˜h˜1.

6.2.8: Follows by Proposition 5.3 and Theorem 5.2 [5.2.10].

6.2.9 and 6.2.10: Follow by Theorem 5.2[5.2.5, 5.2.10].

6.2.11: Follows by Proposition 6.4 and 6.2.7. Moreover, one shows that there do not exist (s˜,μ˜)(T˜Z(G˜))×k* with dim(V˜s˜(μ˜))=75.

6.2.12 and 6.2.13: Follow by Proposition 6.4, 6.2.4, and 6.2.11.

6.2.14: Follows by Proposition 6.4, 6.2.4, and 6.2.12.

6.2.15: Follows by Proposition 6.4, 6.2.4, and 6.2.14.

6.2.16: Follows recursively, by 6.2.17, Proposition 5.3 and Theorem 5.2[5.2.6] for the base case.

6.2.17: Follows recursively, using 6.2.4 to prove the base case. □

7 Proof of Theorem 1.1 for groups of type Dl

Let k be an algebraically closed field of characteristic p0, let W be a 2l-dimensional k-vector space equipped with a nondegenerate quadratic form Q and let G=SO(W,Q). Note that G is a simple algebraic group of type Dl. We let G˜ be a simply connected cover of G, and we fix a central isogeny ϕ:G˜G with ker(ϕ)Z(G˜) and 0. As in Section 6, we will denote by X˜ the object in G˜ corresponding to the object X in G under ϕ. For example, T˜ is a maximal torus in G˜ with ϕ(T˜)=T.

Let a be a nondegenerate alternating bilinear form on W and let H=Sp(W,a). Note that H is a simple simply connected linear algebraic group of type Cl. Let TH be a maximal torus in H. Note that we can choose a symplectic basis and an orthogonal basis in W such that T=TH. Lastly, we denote by ω1H,,ωlH the fundamental dominant weights of H.

This section is dedicated to Theorems 7.1 and 7.2 which give the values of maxu˜G˜u{1}dim(V˜u˜(1)) and maxs˜T˜Z(G˜)dim(V˜s˜(μ˜)) for all kG˜-modules V˜=LG˜(λ˜), with λ˜X(T˜)a nonzero p-restricted dominant weight, and dim(LG˜(λ˜))4l3. As a corollary, the part of Theorem 1.1 concerning simple simply connected linear algebraic groups of type Dl will follow.

Theorem 7.1.

Let k be an algebraically closed field of characteristic p0 and let G˜ be a simple simply connected linear algebraic group of type Dl with l4. Let V˜=LG˜(λ˜), where λ˜X(T˜) is a nonzero p-restricted dominant weight, be such that dim(V˜)4l3. The value of maxu˜G˜u{1}dim(V˜u˜(1)) is as given in

Table 14 The value of maxu˜G˜u{1}dim(V˜u˜(1)) for groups of type Dl.

Proof.

For 7.1.17.1.5, 7.1.7, 7.1.10, 7.1.19, and 7.1.24 we will deduce the result for V˜=LG˜(λ˜) by calculating maxuGu{1}dim((LG(λ))u(1))=dim((LG(λ))xαl(1)(1)), see Section 3.1 and Lemmas 3.2 and 3.13. The proofs for the remaining results are more straightforward: first one determines the structure of V˜|[L˜1,L˜1], and then applies the algorithm of Section 3.3. Once more, we will use Lemma 3.13, by which we have maxu˜G˜u{1}dim(V˜u˜(1))=dim(V˜xα˜l(1)(1)).

7.1.1: Note that V˜LG(ω1) as kG˜-modules and LG(ω1)W as kG-modules. Moreover, xαl(1) acts on W as J22J12l4.

7.1.2: Note that V˜LG(ω2) as kG˜-modules. The result follows by [Citation18, Proposition 4.2.2], respectively by Lemma 7.3 if p = 2, the proof of Theorem 5.1[5.1.2] and, when ε2(l)=1, by [Citation12, Theorem B].

7.1.3: Note that V˜LG(2ω1) as kG˜-modules. The result follows by [Citation18, Propositions 4.7.3], the proof of Theorem 5.1[5.1.3] and [Citation11, Corollary 6.2].

7.1.4: Note that V˜LG(ω3) as kG˜-modules. If p2, the result follows by [Citation18, Proposition 4.2.2] and Theorem 5.1[5.1.4]. Assume p = 2. Then LH(ω3H)LG(ω3) as kG-modules, see [Citation19, ]. Moreover, 3(W)LH(ω3H)LH(ω1H) if ε2(l1)=0 and 3(W)LH(ω1H)|LH(ω3H)|LH(ω1H) if ε2(l1)=1, see [Citation18, Lemma 4.8.2]. Then maxu˜G˜u{1}dim(V˜u˜(1)) 4l318l2+44l423(1+ε2(l1))(2l2). Equality is shown recursively, using the structure of LG(ω3)|[L1,L1], 7.1.1 and 7.1.2.

7.1.5: Note that V˜LG(3ω1) as kG˜-modules. If εp(l+1)=0, then S3(W)LG(3ω1)LG(ω1), while, if εp(l+1)=1, then S3(W)LG(ω1)|LG(3ω1)|LG(ω1) as kG-modules, see [Citation18, Proposition 4.7.4]. Then maxu˜G˜u{1}dim(V˜u˜(1)) (2l3)(2l2)εp(l+1). Equality is shown recursively, using the structure of LG(3ω1)|[L1,L1], 7.1.1, 7.1.3 and Theorem 5.1[5.1.2, 5.1.9, 5.1.14].

7.1.6: The result follows recursively, using 7.1.1 and 7.1.3 (and 7.1.2 when p3). To prove the base case of l=4, we need Theorems 4.1[4.1.2] (and [4.1.4] when p3) and 4.10[4.3.4, 4.3.5].

7.1.7: Note that V˜LG(ω3+ω4) as kG˜-modules. If p = 2, then Λ3(W)LG(ω3+ω4)W as kG-modules, see [Citation19, ] and [Citation18, Lemma 4.8.2]. The result now follows by Theorem 5.1[5.1.4]. If p2, it follows by the structure of V˜|[L˜1,L˜1] and Theorem 4.1[4.1.2, 4.1.3, 4.1.4].

7.1.8: Follows by Theorem 4.10[4.4.4, 4.4.5, 4.4.8].

7.1.9: Follows by Theorems 4.1[4.1.1, 4.1.7] and 4.10[4.4.7].

7.1.10: Follows by [Citation19, ] and Theorem 5.1[5.1.18].

7.1.11: Follows by 7.1.3 and 7.1.7.

7.1.12: Follows by 7.1.2, 7.1.3, and 7.1.7.

7.1.13: Follows by 7.1.1 and 7.1.7.

7.1.14: Follows by 7.1.7.

7.1.15: Follows by 7.1.11 and 7.1.12.

7.1.16: Follows by 7.1.4, 7.1.11, 7.1.12, and 7.1.25.

7.1.17: Follows by 7.1.13 and 7.1.25.

7.1.18: Follows by 7.1.2, 7.1.4, and 7.1.12.

7.1.19: Follows by [Citation19, ] and Theorem 5.1[5.1.19].

7.1.20: Follows by 7.1.17 and 7.1.25.

7.1.21: Follows by 7.1.2, 7.1.4, and 7.1.18.

7.1.22: Follows by 7.1.2, 7.1.4, and 7.1.21.

7.1.23: Follows by 7.1.20 and 7.1.25.

7.1.24: Follows by [Citation19, ] and Theorem 5.1[5.1.29].

7.1.25: Follows recursively, using 7.1.1. □

Theorem 7.2.

Let k be an algebraically closed field of characteristic p0 and let G˜ be a simple simply connected linear algebraic group of type Dl with l4. Let V˜=LG˜(λ˜), where λ˜X(T˜) is a nonzero p-restricted dominant weight, be such that dim(V˜)4l3. The value of maxs˜T˜Z(G˜)dim(V˜s˜(μ˜)) is as given in .

Table 15 The value of maxs˜T˜Z(G˜)dim(V˜s˜(μ˜)) for groups of type Dl.

Before we begin, recall that H=Sp(W,a) is a simple simply connected linear algebraic group of type Cl with maximal torus TH with the property that T=TH. Thus, any sT is conjugate in H to an element sHTH of the form sH=diag(μ1·In1,μ2·In2,,μm·Inm,μm1·Inm,,μ21·In2,μ11·In1) with μiμj±1 for all i < j, i=1mni=l and ln1nm1.

Lemma 7.3.

Let p = 2. If ε2(l)=0, then 2(W)LG(ω2)LG(0), as kG-modules. If ε2(l)=1, then the kG-module 2(W) has three composition factors one isomorphic to LG(ω2) and two to LG(0).

Proof.

By [Citation19, 1.15], the kG-module 2(W) admits a unique nontrivial composition factor of highest weight ω2. Since dim(LG(ω2))=2l2l-gcd(2,l), see [Citation13, ], we determine that, if ε2(l)=0, then 2(W) has two composition factors: one isomorphic to LG(ω2) and one to LG(0), hence 2(W)LG(ω2)LG(0), by [Citation8, II.2.14]. If ε2(l)=1, then 2(W) has three composition factors: one isomorphic to LG(ω2) and two to LG(0). □

Proposition 7.4.

Let V˜=LG˜(ω˜2). Then maxs˜T˜Z(G˜)dim(V˜s˜(μ˜))=2l25l+4(1+ε2(l))εp(2).

Proof.

Let sTZ(G) and note that, in particular, we have sTHZ(H). Let μ be an eigenvalue of s on V. Now, when p2, as V2(W), we have dim(Vs(μ))=dim((2(W))s(μ)). Further, by [Citation18, Lemma 4.8.2] which gives the structure of 2(W) as a kH-module, we have dim(Vs(μ))=dim((LH(ω2H))s(μ)) for μ1, and dim(Vs(1))=dim((LH(ω2H))s(1))+1+εp(l). We now use Proposition 5.4 to get the result. Similarly, when p = 2, by the structure of 2(W) as a kG-module, we determine that dim(Vs(μ))=dim((2(W))s(μ)) for μ1, and dim(Vs(1))=dim((2(W))s(1))1ε2(l). Arguing as in the previous case, we determine that dim(Vs(μ))=dim((LH(ω2H))s(μ)) for all μ, and the result follows by Proposition 5.4. □

Proposition 7.5.

Let p2 and V˜=LG˜(2ω˜1). Then maxs˜T˜Z(G˜)dim(V˜s˜(μ˜))=2l23l+3εp(l).

Proof.

Set V=LG(2ω1) and note that V˜V as kG˜-modules. Further, if εp(l)=0, then S2(W)=VLG(0), while, if εp(l)=1, then S2(W)=LG(0)|V|LG(0), see [Citation18, Propositions 4.7.3].

Let sTZ(G) and note that, in particular, sTHZ(H). Let μ be an eigenvalue of s on V. Arguing as in the proof of Proposition 7.4, we show that dim(Vs(μ))=dim((LH(2ω1H))s(μ)) for μ1; and dim(Vs(1))=dim((LH(2ω1H))s(1))1εp(l). The result for μ = 1 is given by Proposition 5.5. For μ such that μμ1 we have dim(Vs(μ))dim(V)2<2l23l+3εp(l), as V is self-dual. Thus, for the remainder of the proof, we let μ=1.

As sTHZ(H), we write s=diag(μ1·In1,,μm·Inm,μm1·Inm,,μ11·In1), where μiμj±1 for all i < j, i=1mni=l and n1nm1. If μiμj1 for all i<j, then dim(Vs(1))l2+l<2l23l+3εp(l), see inequality (5.8). If there exists i < j such that μiμj=1, then dim(Vs(1))2l2+lr=1mnr2ni(ni+1)nj(nj+1) 2(ni+nj)(lninj), see inequality (5.9). Assume that dim(Vs(1))2l23l+3εp(l). Then: l(4ninj)3+εp(l)ri,jnr2(ninj)2(ni+nj)(l+1ninj)0.

As l+1>ni+nj, we must have (ni,nj){(1,1),(2,1)}. For (ni,nj)=(1,1), we get 1+εp(l) ri,jnr20, which does not hold. Similarly, if (ni,nj)=(2,1), then 2l+2+εp(l)ri,jnr20, which does not hold. Thus, dim(Vs(1))<2l23l+3εp(l) for all sTZ(G). □

Proposition 7.6.

Let l4 and let V˜=LG˜(ω˜1+ω˜2). If p = 3, then maxs˜T˜Z(G˜)dim(V˜s˜(μ˜))=(2l3)+(2l2)(1ε3(2l1)); while if p3, then maxs˜T˜Z(G˜)dim(V˜s˜(μ˜))=24(l3)+(2l2)(4εp(2l1)2εp(2)).

Proof.

The result follows recursively, using the algorithm of Section 3.3, the structure of V˜|[L˜1,L˜1], Proposition 7.5, Theorem 7.2[7.2.1] and, additionally, Proposition 7.4 when p3. In what follows we prove that the base case holds.

Let l=4. If p = 3, we use Propositions 4.4 and Theorem 4.9[4.3.4, 4.3.5] to prove the result. Note that when s˜=z˜·h˜ with z˜Z(L˜1)° and h˜[L˜1,L˜1], one has to treat the case when h˜ is conjugate to hα˜2(±1)hα˜3(d1)hα˜4(d) with d2±1 separately. Thus, assume p3. The composition factors of V˜|[L˜1,L˜1] are as follows: one isomorphic to LL˜1(ω˜2+ω˜3+ω˜4), 4εp(7)+εp(5) to LL˜1(ω˜2), two to LL˜1(2ω˜2),2+2εp(2) to LL˜1(ω˜3+ω˜4) and 22εp(7)+2εp(2) to LL˜1(0).

When sZ(L˜1)°, i.e. s˜=z˜·h˜ with z˜Z(L˜1)° and h˜[L˜1,L˜1], one has to eliminate the cases when there exists μ˜h˜ with dim((LL˜1(ω˜2))h˜(μ˜h˜))=4 and dim((LL˜1(ω˜2))h˜(μ˜h˜))=3. If p2, the result then follows from Proposition 4.6 and Theorem 4.9[4.3.4, 4.3.5]. Thus, assume p = 2 and let L˜11 be a Levi subgroup of the maximal parabolic subgroup P˜11 of L˜1 corresponding to Δ˜11=Δ˜1{α˜3}. We once again abuse notation and denote by ω˜2 and ω˜4 the fundamental dominant weights of L˜11 corresponding to the simple roots α˜2 and α˜4. One shows that if h˜Z(L˜11)°Z(L˜1), then dim((LL˜1(ω˜2+ω˜3+ω˜4))h˜(μ˜h˜))24 for all μ˜h˜. On the other hand, if h˜Z(L˜11)°, then dim((LL˜1(ω˜2+ω˜3+ω˜4))h˜(μ˜h˜))2dim((LL˜11(ω˜2+ω˜4))h˜11(μ˜h˜11))+2dim((LL˜11(2ω˜2+ω˜4))h˜11(μ˜h˜11))+10dim((LL˜11(ω˜2))h˜11(μ˜h˜11)), where h˜=z˜11·h˜11 with z˜11Z(L˜11)° and h˜11[L˜11,L˜11]. One treats the case when there exists μ˜h˜11 such that dim((LL˜11(ω˜2))h˜11(μ˜h˜11))=2 separately, and afterwards concludes that dim(V˜s˜(μ˜))70. □

Proof of Theorem 7.2.

The proofs of 7.2.2, 7.2.3, and 7.2.6 are given in Propositions 7.4–7.6, respectively.

7.2.1: We argue as in the proofs of Theorem 7.1[7.1.1] and Proposition 4.3.

7.2.4: If p = 2, we argue as we did in the proof of Theorem 7.1[7.1.4] and use Theorem 5.2[5.2.4]. If p2, the result follows recursively, using [Citation18, Proposition 4.2.2], Theorem 5.2[5.2.4], Proposition 7.4, 7.2.1, and 7.2.7 for the base case of l=5.

7.2.5: It follows recursively, using Proposition 7.5, 7.2.1, and, for the base case of l=4, Proposition 4.4 and Theorem 4.9[4.3.5, 4.3.6].

7.2.7: If p = 2, we argue as in the proof of Theorem 7.1[7.1.7] and use Theorem 5.2[5.2.4]. When p2, we use Propositions 4.4–4.6, to prove the result. Note that when s˜=z˜·h˜ with z˜Z(L˜1)° and h˜[L˜1,L˜1], one has to treat the case when h˜ is conjugate to hα˜2(d)hα˜3(d2)hα˜4(d3) with d41 separately.

7.2.8: Follows by Theorem 4.9[4.3.4, 4.3.5, 4.3.8].

7.2.9: Follows by Proposition 4.3 and Theorems 4.2[4.2.7] and 4.9[4.3.7].

7.2.10: Follows by [Citation19, ] and Theorem 5.2[5.2.18].

7.2.11: Follows by Proposition 7.5 and 7.2.7.

7.2.12: Follows by Propositions 7.4, 7.5, and 7.2.7.

7.2.13: Follows by 7.2.1 and 7.2.7.

7.2.14: Follows by 7.2.7.

7.2.15: Follows by 7.2.11 and 7.2.12.

7.2.16: Follows by 7.2.4, 7.2.11, 7.2.12, and 7.2.25.

7.2.17: Follows by 7.2.13 and 7.2.25.

7.2.18: Follows by Proposition 7.4, 7.2.4, and 7.2.12.

7.2.19: Follows by [Citation19, ] and Theorem 5.2[5.2.19].

7.2.20: Follows by 7.2.17 and 7.2.25.

7.2.21: Follows by Proposition 7.4, 7.2.4, and 7.2.18.

7.2.22: Follows by Proposition 7.4, 7.2.4, and 7.2.21.

7.2.23: Follows by 7.2.20 and 7.2.25.

7.2.24: Follows by [Citation19, ] and Theorem 5.2[5.2.29].

7.2.25: Follows recursively, using 7.2.1. Note that when l=5 and s˜Z(L˜1)°, one has to prove that: there do not exist μ˜ such that dim(V˜s˜(μ˜))>10; and that for s˜=hα˜4(d)hα˜5(d) with d1 we have dim(V˜s˜(1))=10. □

Acknowledgments

The author is immensely grateful to Donna Testerman for her guidance. The author would also like to thank Simon Goodwin, Martin Liebeck and Adam Thomas for many helpful discussions and comments.

Disclosure statement

The author reports there are no competing interests to declare.

Additional information

Funding

This work was supported by the Swiss National Science Foundation, grant number FNS 200020 175571, and by the Engineering and Physical Sciences Research Council, grant number EP/R018952/1.

References

  • Carter, R. W. (1989). Simple Groups of Lie Type. Wiley Classics Library. London: Wiley.
  • Gordeev, N. L. (1991). Coranks of elements of linear groups and the complexity of algebras of invariants. Leningrad Math. J. 2:245–267.
  • Gow, R., Laffey, T. J. (2006). On the decomposition of the exterior square of an indecomposable module of a cyclic p-group. J. Group Theory 9(5):659–672.
  • Guralnick, R. M., Saxl, J. (2003). Generation of finite almost simple groups by conjugates. J. Algebra 268(2):519–571. DOI: 10.1016/S0021-8693(03)00182-0.
  • Guralnick, R. M., Lawther, R. (2019). Generic stabilizers in actions of simple algebraic groups i: modules and the first grassmanian varieties. arXiv: Group Theory.
  • Hall, J. I., Liebeck, M. W., Seitz, G. M. (1992). Generators for finite simple groups, with applications to linear groups. Q. J. Math. 43(4):441–458. DOI: 10.1093/qmathj/43.4.441.
  • Humphreys, J. E. (1972). Introduction to Lie Algebras and Representation Theory. Graduate Texts in Mathematics. New York: Springer.
  • Jantzen, J. C. (2007). Representations of Algebraic Groups. Mathematical Surveys and Monographs. Providence, RI: American Mathematical Society.
  • Kac, V. G., Watanabe, K. (1982). Finite linear groups whose ring of invariants is a complete intersection. Bull. Amer. Math. Soc. 6:221–223. DOI: 10.1090/S0273-0979-1982-14989-8.
  • Kemper, G., Malle, G. (1997). The finite irreducible linear groups with polynomial ring of invariants. Transformation Groups 2:57–89. DOI: 10.1007/BF01234631.
  • Korhonen, M. (2019). Jordan blocks of unipotent elements in some irreducible representations of classical groups in good characteristic. Proc. Amer. Math. Soc. 147(10):4205–4219. DOI: 10.1090/proc/14570.
  • Korhonen, M. (2020). Hesselink normal forms of unipotent elements in some representations of classical groups in characteristic two. J. Algebra 559:268–319. DOI: 10.1016/j.jalgebra.2020.03.035.
  • Lübeck, F. (2001). Small degree representations of finite Chevalley groups in defining characteristic. LMS J. Comput. Math. 4:135–169. http://www.math.rwth-aachen.de/Frank.Luebeck/chev/WMSmall/index.html. DOI: 10.1112/S1461157000000838.
  • Lübeck, F. (2001). Tables of weight multiplicities.
  • Liebeck, M. W., Seitz, G. M. (2012). Unipotent and Nilpotent Classes in Simple Algebraic Groups and Lie Algebras. Mathematical Surveys and Monographs. Providence, RI: American Mathematical Society.
  • Malle, G., Testerman, D. (2011). Linear Algebraic Groups and Finite Groups of Lie Type. Cambridge Studies in Advanced Mathematics. Cambridge, UK: Cambridge University Press.
  • Martínez, A. L. (2018). Low-dimensional irreducible rational representations of classical algebraic groups. DOI: 10.48550/ARXIV.1811.07019.
  • McNinch, G. J. (1998). Dimensional criteria for semisimplicity of representations. Proc. London Math. Soc. 76(1):95–149. DOI: 10.1112/S0024611598000045.
  • Seitz, G. M. (1987). The Maximal Subgroups of Classical Algebraic Groups, 365. Providence, RI: American Mathematical Society.
  • Smith, S. D. (1982). Irreducible modules and parabolic subgroups. J. Algebra 75(1):286–289. DOI: 10.1016/0021-8693(82)90076-X.
  • Steinberg, R. (2016). Lectures on Chevalley Groups. University Lecture Series, Vol. 66. Providence, RI: American Mathematical Society.
  • Suprunenko, I. D. (1983). Preservation of systems of weights of irreducible representations of an algebraic group and a Lie algebra of type A with bounded higher weights in reduction modulo p. Vestsi Akad. Na uk BSSR, Ser. Fiz.-Mat. Na uk (2):18–22.
  • Verbitsky, M. (1999). Holomorphic symplectic geometry and orbifold singularities. Asian J. Math. 4:553–564. DOI: 10.4310/AJM.2000.v4.n3.a4.