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

Adjoint Jordan blocks for simple algebraic groups of type CC_{\ell} in characteristic two

Mikko Korhonen SUSTech International Center for Mathematics, Southern University of Science and Technology, Shenzhen 518055, Guangdong, P. R. China [email protected] In memory of Irina Suprunenko
Abstract.

Let GG be a simple algebraic group over an algebraically closed field KK with Lie algebra 𝔤\mathfrak{g}. For unipotent elements uGu\in G and nilpotent elements e𝔤e\in\mathfrak{g}, the Jordan block sizes of Ad(u)\operatorname{Ad}(u) and ad(e)\operatorname{ad}(e) are known in most cases. In the cases that remain, the group GG is of classical type in bad characteristic, so charK=2\operatorname{char}K=2 and GG is of type BB_{\ell}, CC_{\ell}, or DD_{\ell}.

In this paper, we consider the case where GG is of type CC_{\ell} and charK=2\operatorname{char}K=2. As our main result, we determine the Jordan block sizes of Ad(u)\operatorname{Ad}(u) and ad(e)\operatorname{ad}(e) for all unipotent uGu\in G and nilpotent e𝔤e\in\mathfrak{g}. In the case where GG is of adjoint type, we will also describe the Jordan block sizes on [𝔤,𝔤][\mathfrak{g},\mathfrak{g}].

Support by Shenzhen Science and Technology Program (Grant No. RCBS20210609104420034).

1. Introduction

Let GG be a simple algebraic group over an algebraically closed field KK, and denote the Lie algebra of GG by 𝔤\mathfrak{g}. Recall that an element uGu\in G is unipotent, if f(u)f(u) is a unipotent linear map for every rational representation f:GGL(W)f:G\rightarrow\operatorname{GL}(W). Similarly e𝔤e\in\mathfrak{g} is said to be nilpotent, if df(e)\mathrm{d}f(e) is a nilpotent linear map for every rational representation f:GGL(W)f:G\rightarrow\operatorname{GL}(W).

We denote the adjoint representations of GG and 𝔤\mathfrak{g} by Ad:GGL(𝔤)\operatorname{Ad}:G\rightarrow\operatorname{GL}(\mathfrak{g}) and ad:𝔤𝔤𝔩(V)\operatorname{ad}:\mathfrak{g}\rightarrow\mathfrak{gl}(V), respectively. In this paper, we will consider the following two problems.

Problem 1.1.

Let uGu\in G be a unipotent element. What is the Jordan normal form of Ad(u)\operatorname{Ad}(u)?

Problem 1.2.

Let e𝔤e\in\mathfrak{g} be a nilpotent element. What is the Jordan normal form of ad(e)\operatorname{ad}(e)?

In most cases, the answer to both questions is known. When GG is simply connected of exceptional type, the Jordan block sizes were computed in the unipotent case by Lawther [21, 22] and in the nilpotent case by Stewart [33]. In the case where GG is exceptional of adjoint type, both questions are easily settled using the results of Lawther and Stewart, see [20, Lemma 3.1].

When GG is of type AA_{\ell}, solutions to both questions follow from the main results of [16] and [19], see [19, Remark 1.3]. For GG of type BB_{\ell}, CC_{\ell}, or DD_{\ell} in good characteristic, the Jordan block sizes are described by well-known results on decompositions of tensor products, symmetric squares, and exterior squares — see for example [19, Remark 3.5].

Thus it remains to solve Problem 1.1 and Problem 1.2 in the case where GG is of type BB_{\ell}, CC_{\ell}, or DD_{\ell} in characteristic charK=2\operatorname{char}K=2. In this paper, we consider the case where GG is of type CC_{\ell}. We make the following assumption for the rest of this paper.

Assume that charK=2\operatorname{char}K=2.

Let GG be a simple algebraic group of type CC_{\ell} over KK. As our main result, we determine the Jordan block sizes of Ad(u)\operatorname{Ad}(u) and ad(e)\operatorname{ad}(e) for all unipotent uGu\in G and nilpotent e𝔤e\in\mathfrak{g}. In type CC_{\ell} either GG is simply connected or GG is adjoint; we deal with these two possibilities as follows.

Let GscG_{sc} be simply connected and simple of type CC_{\ell} with Lie algebra 𝔤sc\mathfrak{g}_{sc}, and let GadG_{ad} be simple of adjoint type CC_{\ell} with Lie algebra 𝔤ad\mathfrak{g}_{ad}. There exists an isogeny φ:GscGad\varphi:G_{sc}\rightarrow G_{ad}, which induces a bijection between the unipotent variety of GscG_{sc} and GadG_{ad}. Similarly the differential dφ:𝔤sc𝔤ad\mathrm{d}\varphi:\mathfrak{g}_{sc}\rightarrow\mathfrak{g}_{ad} induces a bijection between the nilpotent cone of 𝔤sc\mathfrak{g}_{sc} and 𝔤ad\mathfrak{g}_{ad}.

Thus for the solution of Problem 1.1 and Problem 1.2 in type CC_{\ell}, it will suffice to consider the Jordan block sizes of unipotent uGscu\in G_{sc} on 𝔤sc\mathfrak{g}_{sc} and 𝔤ad\mathfrak{g}_{ad}, and the Jordan block sizes of nilpotent e𝔤sce\in\mathfrak{g}_{sc} on 𝔤sc\mathfrak{g}_{sc} and 𝔤ad\mathfrak{g}_{ad}.

We can assume that Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) with Lie algebra 𝔤sc=𝔰𝔭(V)\mathfrak{g}_{sc}=\mathfrak{sp}(V), where dimV=2\dim V=2\ell. It is well known (Lemma 8.1) that 𝔤scS2(V)\mathfrak{g}_{sc}\cong S^{2}(V)^{*} as GscG_{sc}-modules. The Jordan block sizes of the action of unipotent uGL(V)u\in\operatorname{GL}(V) on S2(V)S^{2}(V) and nilpotent e𝔤𝔩(V)e\in\mathfrak{gl}(V) on S2(V)S^{2}(V) are described in [18]. Taking the dual does not change the Jordan block sizes, so the Jordan block sizes on 𝔤sc\mathfrak{g}_{sc} are known by previous results.

In our main results, we will describe the Jordan block sizes on 𝔤ad\mathfrak{g}_{ad} in terms of the Jordan block sizes on 𝔤sc\mathfrak{g}_{sc}. To state the result in the unipotent case, we first need describe the classification of unipotent conjugacy classes in Sp(V)\operatorname{Sp}(V), which in characteristic two is due to Hesselink [8]. We discuss this in some more detail in Section 5.

Let CqC_{q} be a cyclic group of order qq, where q=2αq=2^{\alpha} for some α0\alpha\geq 0. Then there are a total of qq indecomposable K[Cq]K[C_{q}]-modules V1V_{1}, \ldots, VqV_{q} up to isomorphism, where dimVi=i\dim V_{i}=i and a generator of CqC_{q} acts on ViV_{i} with a single i×ii\times i unipotent Jordan block. For convenience we will denote V0=0V_{0}=0. If WW is a vector space over KK, we denote W0=0W^{0}=0, and Wd=WWW^{d}=W\oplus\cdots\oplus W (dd summands) for an integer d>0d>0.

Suppose then that uGL(V)u\in\operatorname{GL}(V) is unipotent, so uu has order qq for some q=2αq=2^{\alpha}. We denote the group algebra of u\langle u\rangle by K[u]K[u]. Then VK[u]Vd1n1VdtntV\downarrow K[u]\cong V_{d_{1}}^{n_{1}}\oplus\cdots\oplus V_{d_{t}}^{n_{t}} for some integers 0<d1<<dt0<d_{1}<\cdots<d_{t}, where ni>0n_{i}>0 for all 1it1\leq i\leq t. Equivalently, the Jordan normal form of uu has Jordan blocks of sizes d1d_{1}, \ldots, dtd_{t}, and a block of size did_{i} has multiplicity nin_{i}.

For unipotent uSp(V)u\in\operatorname{Sp}(V), we have a decomposition VK[u]=U1UtV\downarrow K[u]=U_{1}\perp\cdots\perp U_{t}, where UiU_{i} are orthogonally indecomposable K[u]K[u]-modules. Here orthogonally indecomposable means that if Ui=UU′′U_{i}=U^{\prime}\perp U^{\prime\prime} as K[u]K[u]-modules, then U=0U^{\prime}=0 or U′′=0U^{\prime\prime}=0. The orthogonally indecomposable K[u]K[u]-modules fall into two types: one denoted by V(m)V(m) (for mm even) and another denoted by W(m)W(m) (for m1m\geq 1). We define these in Section 5, for now we only mention that V(m)VmV(m)\cong V_{m} and W(m)VmVmW(m)\cong V_{m}\oplus V_{m} as K[u]K[u]-modules.

Our first main result is the following. (Below we denote by ν2\nu_{2} the 22-adic valuation on the integers, so ν2(a)\nu_{2}(a) is the largest integer k0k\geq 0 such that 2k2^{k} divides aa.)

Theorem 1.3.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent, with orthogonal decomposition

VK[u]=1itW(mi)1jsV(2kj).V\downarrow K[u]=\sum_{1\leq i\leq t}W(m_{i})\perp\sum_{1\leq j\leq s}V(2k_{j}).

Denote by α0\alpha\geq 0 the largest integer such that 2αmi,kj2^{\alpha}\mid m_{i},k_{j} for all ii and jj. Then the following hold:

  1. (i)

    Suppose that s=0s=0. Then:

    1. (a)

      If α=0\alpha=0, then 𝔤sc𝔤ad\mathfrak{g}_{sc}\cong\mathfrak{g}_{ad} as K[u]K[u]-modules.

    2. (b)

      If α>0\alpha>0, then

      𝔤sc\displaystyle\mathfrak{g}_{sc} V2αV\displaystyle\cong V_{2^{\alpha}}\oplus V^{\prime}
      𝔤ad\displaystyle\mathfrak{g}_{ad} V1V2α1V\displaystyle\cong V_{1}\oplus V_{2^{\alpha}-1}\oplus V^{\prime}

      for some K[u]K[u]-module VV^{\prime}.

  2. (ii)

    Suppose that s>0s>0, and let β=max1jsν2(kj)\beta=\max_{1\leq j\leq s}\nu_{2}(k_{j}). Then:

    1. (a)

      If β=ν2(kj)\beta=\nu_{2}(k_{j}) for all 1js1\leq j\leq s and ν2(mi)>β\nu_{2}(m_{i})>\beta for all 1it1\leq i\leq t, then 𝔤sc𝔤ad\mathfrak{g}_{sc}\cong\mathfrak{g}_{ad} as K[u]K[u]-modules.

    2. (b)

      If β>ν2(kj)\beta>\nu_{2}(k_{j}) for some 1js1\leq j\leq s, or ν2(mi)β\nu_{2}(m_{i})\leq\beta for some 1it1\leq i\leq t, then

      𝔤sc\displaystyle\mathfrak{g}_{sc} V2αV2βV\displaystyle\cong V_{2^{\alpha}}\oplus V_{2^{\beta}}\oplus V^{\prime}
      𝔤ad\displaystyle\mathfrak{g}_{ad} V2α1V2β+1V\displaystyle\cong V_{2^{\alpha}-1}\oplus V_{2^{\beta}+1}\oplus V^{\prime}

      for some K[u]K[u]-module VV^{\prime}.

To describe our results in the nilpotent case, we recall the classification of nilpotent orbits in 𝔰𝔭(V)\mathfrak{sp}(V), due to Hesselink [8]. For more details we refer to Section 6.

Suppose that q=2αq=2^{\alpha} for some α0\alpha\geq 0. Let 𝔴q\mathfrak{w}_{q} be the abelian 22-Lie algebra over KK generated by a single nilpotent element e𝔴qe\in\mathfrak{w}_{q} such that e[2α]=0e^{[2^{\alpha}]}=0 and e[2α1]0e^{[2^{\alpha-1}]}\neq 0. There are a total of qq indecomposable 𝔴q\mathfrak{w}_{q}-modules W1W_{1}, \ldots, WqW_{q} up to isomorphism, where dimWi=i\dim W_{i}=i and ee acts on WiW_{i} with a single i×ii\times i nilpotent Jordan block. Throughout we will denote W0=0W_{0}=0.

Consider then a nilpotent linear map e𝔤𝔩(V)e\in\mathfrak{gl}(V), and denote the 22-Lie subalgebra generated by ee with K[e]K[e]. Then K[e]𝔴qK[e]\cong\mathfrak{w}_{q}, where qq is the smallest power of two such that eq=0e^{q}=0. We have VK[e]Wd1n1WdtntV\downarrow K[e]\cong W_{d_{1}}^{n_{1}}\oplus\cdots\oplus W_{d_{t}}^{n_{t}} for some integers 0<d1<<dt0<d_{1}<\cdots<d_{t}, where ni>0n_{i}>0 for all 1it1\leq i\leq t. As in the unipotent case, this amounts to the statement that the Jordan normal form of ee has Jordan blocks of sizes d1d_{1}, \ldots, dtd_{t}, and a block of size did_{i} has multiplicity nin_{i}.

For nilpotent e𝔰𝔭(V)e\in\mathfrak{sp}(V), we have an orthogonal decomposition VK[e]=U1UtV\downarrow K[e]=U_{1}\perp\cdots\perp U_{t}, where the UiU_{i} are orthogonally indecomposable K[e]K[e]-modules. Here orthogonally indecomposable is defined as in the unipotent case. For the orthogonally indecomposable K[e]K[e]-modules, there are several different types: V(m)V(m) (for mm even), Wk(m)W_{k}(m) (for 0<k<m/20<k<m/2 and m>2m>2), and W(m)W(m) (for m1m\geq 1). We give the definitions and more details in Section 6, for now we just note that V(m)WmV(m)\cong W_{m}, Wk(m)WmWmW_{k}(m)\cong W_{m}\oplus W_{m}, and W(m)WmWmW(m)\cong W_{m}\oplus W_{m} as K[e]K[e]-modules.

Our main result in the nilpotent case is the following.

Theorem 1.4.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent, with orthogonal decomposition

VK[e]=1itW(mi)1jtWkj(j)1rt′′V(2dr).V\downarrow K[e]=\sum_{1\leq i\leq t}W(m_{i})\perp\sum_{1\leq j\leq t^{\prime}}W_{k_{j}}(\ell_{j})\perp\sum_{1\leq r\leq t^{\prime\prime}}V(2d_{r}).

Denote by α0\alpha\geq 0 the largest integer such that 2αmi,j,2dr2^{\alpha}\mid m_{i},\ell_{j},2d_{r} for all ii, jj, and rr. Then the following statements hold:

  1. (i)

    Suppose that VK[e]=1itW(mi)V\downarrow K[e]=\sum_{1\leq i\leq t}W(m_{i}). Then:

    1. (a)

      If α=0\alpha=0, then 𝔤sc𝔤ad\mathfrak{g}_{sc}\cong\mathfrak{g}_{ad} as K[e]K[e]-modules.

    2. (b)

      If α>0\alpha>0, then

      𝔤sc\displaystyle\mathfrak{g}_{sc} W2αV\displaystyle\cong W_{2^{\alpha}}\oplus V^{\prime}
      𝔤ad\displaystyle\mathfrak{g}_{ad} W1W2α1V\displaystyle\cong W_{1}\oplus W_{2^{\alpha}-1}\oplus V^{\prime}

      for some K[e]K[e]-module VV^{\prime}.

  2. (ii)

    Suppose that VK[e]V\downarrow K[e] is not of the form 1itW(mi)\sum_{1\leq i\leq t}W(m_{i}). Then:

    1. (a)

      If α=1\alpha=1, then 𝔤sc𝔤ad\mathfrak{g}_{sc}\cong\mathfrak{g}_{ad} as K[e]K[e]-modules.

    2. (b)

      If α1\alpha\neq 1, then

      𝔤sc\displaystyle\mathfrak{g}_{sc} W1W2αV\displaystyle\cong W_{1}\oplus W_{2^{\alpha}}\oplus V^{\prime}
      𝔤ad\displaystyle\mathfrak{g}_{ad} W2W2α1V\displaystyle\cong W_{2}\oplus W_{2^{\alpha}-1}\oplus V^{\prime}

      for some K[e]K[e]-module VV^{\prime}.

Example 1.5.

In Table 1 and Table 2, we illustrate Theorem 1.3 and Theorem 1.4 in the case where GG is of type CC_{\ell} for 242\leq\ell\leq 4. In the tables we have also included the Jordan block sizes on [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}], which we prove as an intermediate result in Proposition 9.5 and Proposition 10.5. Furthermore, the values of α\alpha and β\beta that appear in Theorem 1.3 and Theorem 1.4 are included.

In the tables, we use the notation

d1n1,,dtntd_{1}^{n_{1}},\ldots,d_{t}^{n_{t}}

for 0<d1<<dt0<d_{1}<\cdots<d_{t} and ni>0n_{i}>0 to denote that the Jordan block sizes are d1d_{1}, \ldots, dtd_{t}, and a block of size did_{i} has multiplicity nin_{i}. For the orthogonal decompositions in the first column, notation such as V(m)kV(m)^{k} denotes an orthogonal direct sum V(m)V(m)V(m)\perp\cdots\perp V(m) (kk summands).

In particular, from Theorem 1.3 and Theorem 1.4 we get the number of Jordan blocks of Ad(u)\operatorname{Ad}(u) and ad(e)\operatorname{ad}(e), which is equal to the dimension of the Lie algebra centralizer of uu and ee, respectively.

Corollary 1.6.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent and denote α\alpha, ss as in Theorem 1.3. Then the following hold:

  1. (i)

    Suppose that s=0s=0. Then

    dim𝔤adu={dim𝔤scu, if α=0.dim𝔤scu+1, if α>0.\dim\mathfrak{g}_{ad}^{u}=\begin{cases}\dim\mathfrak{g}_{sc}^{u},&\text{ if }\alpha=0.\\ \dim\mathfrak{g}_{sc}^{u}+1,&\text{ if }\alpha>0.\end{cases}
  2. (ii)

    Suppose that s>0s>0. Then

    dim𝔤adu={dim𝔤scu1, if α=0 and ν2(kj)>0 for some j.dim𝔤scu1, if α=0 and ν2(mi)=0 for some i.dim𝔤scu, if α=0 and ν2(kj)=0,ν2(mi)>0 for all i and j.dim𝔤scu, if α>0.\dim\mathfrak{g}_{ad}^{u}=\begin{cases}\dim\mathfrak{g}_{sc}^{u}-1,&\text{ if }\alpha=0\text{ and }\nu_{2}(k_{j})>0\text{ for some }j.\\ \dim\mathfrak{g}_{sc}^{u}-1,&\text{ if }\alpha=0\text{ and }\nu_{2}(m_{i})=0\text{ for some }i.\\ \dim\mathfrak{g}_{sc}^{u},&\text{ if }\alpha=0\text{ and }\nu_{2}(k_{j})=0,\nu_{2}(m_{i})>0\text{ for all }i\text{ and }j.\\ \dim\mathfrak{g}_{sc}^{u},&\text{ if }\alpha>0.\end{cases}
Corollary 1.7.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent and denote α\alpha as in Theorem 1.3. Then the following hold:

  1. (i)

    Suppose that VK[e]=1itW(mi)V\downarrow K[e]=\sum_{1\leq i\leq t}W(m_{i}). Then

    dim𝔤ade={dim𝔤sce, if α=0.dim𝔤sce+1, if α>0.\dim\mathfrak{g}_{ad}^{e}=\begin{cases}\dim\mathfrak{g}_{sc}^{e},&\text{ if }\alpha=0.\\ \dim\mathfrak{g}_{sc}^{e}+1,&\text{ if }\alpha>0.\end{cases}
  2. (ii)

    Suppose that VK[e]V\downarrow K[e] is not of the form 1itW(mi)\sum_{1\leq i\leq t}W(m_{i}). Then

    dim𝔤ade={dim𝔤sce1, if α=0.dim𝔤sce, if α>0.\dim\mathfrak{g}_{ad}^{e}=\begin{cases}\dim\mathfrak{g}_{sc}^{e}-1,&\text{ if }\alpha=0.\\ \dim\mathfrak{g}_{sc}^{e},&\text{ if }\alpha>0.\end{cases}

In the case of regular elements, this amounts to the following. (An element xGx\in G or x𝔤x\in\mathfrak{g} is regular, if dimCG(x)=rankG\dim C_{G}(x)=\operatorname{rank}G. In the case of G=Sp(V)G=\operatorname{Sp}(V), a regular unipotent element is characterized by VK[u]=V(2)V\downarrow K[u]=V(2\ell), and a regular nilpotent element is characterized by VK[e]=V(2)V\downarrow K[e]=V(2\ell).)

Corollary 1.8.

Suppose that GG is of type CC_{\ell} in characteristic two (adjoint or simply connected). Let uGu\in G be a regular unipotent element and e𝔤e\in\mathfrak{g} a regular nilpotent element. Then dim𝔤u=+1\dim\mathfrak{g}^{u}=\ell+1 and dim𝔤e=2\dim\mathfrak{g}^{e}=2\ell.

For the proofs of our main results, our basic approach is as follows. We will first describe the Jordan block sizes of unipotent uGscu\in G_{sc} and nilpotent e𝔤sce\in\mathfrak{g}_{sc} on 𝔤sc/Z(𝔤sc)\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}), in terms of the Jordan block sizes on 𝔤sc\mathfrak{g}_{sc}. This is attained in Section 9 and Section 10.

It is known that 𝔤sc/Z(𝔤sc)[𝔤ad,𝔤ad]\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc})\cong[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] as GscG_{sc}-modules (Lemma 8.2), so we then have the Jordan block sizes on [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}] as well. In Section 11 and Section 12 we will describe the Jordan block sizes of unipotent uGscu\in G_{sc} and nilpotent e𝔤sce\in\mathfrak{g}_{sc} on 𝔤ad\mathfrak{g}_{ad}, in terms of the Jordan block sizes on [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}]. Combining these results, we get the Jordan block sizes of uu and ee on 𝔤ad\mathfrak{g}_{ad}, in terms of the Jordan block sizes on 𝔤sc\mathfrak{g}_{sc} (Theorem 1.3, Theorem 1.4).

The other sections of this paper are organized as follows. The notation, terminology, and some preliminary results are stated in Section 2 and Section 3. In Section 4, we state results on Jordan block sizes of unipotent and nilpotent elements on tensor products, symmetric squares, and exterior squares. The classification of unipotent classes in Sp(V)\operatorname{Sp}(V) and nilpotent orbits in 𝔰𝔭(V)\mathfrak{sp}(V) is described in Section 5 and Section 6, respectively.

In Section 7, we discuss the Chevalley construction of simple algebraic groups, and in particular the action of Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) on 𝔤ad\mathfrak{g}_{ad}. We then make some observations about the structure of 𝔤sc\mathfrak{g}_{sc} and 𝔤ad\mathfrak{g}_{ad} as a GscG_{sc}-module in Section 8.

As mentioned earlier, in Section 9 and Section 10 we describe the Jordan block sizes of unipotent and nilpotent elements on 𝔤sc/Z(𝔤sc)\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}), in terms of the Jordan block sizes on 𝔤sc\mathfrak{g}_{sc}. In Section 11 and Section 12 we similarly describe the Jordan block sizes on [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}], in terms of Jordan block sizes on 𝔤ad\mathfrak{g}_{ad}. These results allow us to prove our main results, and the proofs of the results stated in this introduction are given in Section 13.

Table 1. For Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) simply connected of type CC_{\ell} with 242\leq\ell\leq 4 and charK=2\operatorname{char}K=2, Jordan block sizes of unipotent uGscu\in G_{sc} on 𝔤sc\mathfrak{g}_{sc}, [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}], and 𝔤ad\mathfrak{g}_{ad}. See Example 1.5.
\ell VK[u]V\downarrow K[u] 𝔤scK[u]\mathfrak{g}_{sc}\downarrow K[u] [𝔤ad,𝔤ad]K[u][\mathfrak{g}_{ad},\mathfrak{g}_{ad}]\downarrow K[u] 𝔤adK[u]\mathfrak{g}_{ad}\downarrow K[u] α\alpha β\beta
22 V(4)V(4) 2,422,4^{2} 1,421,4^{2} 2,422,4^{2} 11 11
V(2)2V(2)^{2} 12,241^{2},2^{4} 1,241,2^{4} 12,241^{2},2^{4} 0 0
W(1)V(2)W(1)\perp V(2) 14,231^{4},2^{3} 13,231^{3},2^{3} 12,241^{2},2^{4} 0 0
W(2)W(2) 12,241^{2},2^{4} 13,231^{3},2^{3} 14,231^{4},2^{3} 11 -
W(1)2W(1)^{2} 1101^{10} 191^{9} 1101^{10} 0 -
33 V(6)V(6) 1,4,821,4,8^{2} 4,824,8^{2} 1,4,821,4,8^{2} 0 0
V(2)V(4)V(2)\perp V(4) 1,22,441,2^{2},4^{4} 22,442^{2},4^{4} 2,3,442,3,4^{4} 0 11
V(2)3V(2)^{3} 13,291^{3},2^{9} 12,291^{2},2^{9} 13,291^{3},2^{9} 0 0
W(1)2V(2)W(1)^{2}\perp V(2) 111,251^{11},2^{5} 110,251^{10},2^{5} 19,261^{9},2^{6} 0 0
W(1)V(4)W(1)\perp V(4) 13,2,441^{3},2,4^{4} 12,2,441^{2},2,4^{4} 12,3,441^{2},3,4^{4} 0 11
W(1)V(2)2W(1)\perp V(2)^{2} 15,281^{5},2^{8} 14,281^{4},2^{8} 13,291^{3},2^{9} 0 0
W(3)W(3) 1,22,441,2^{2},4^{4} 22,442^{2},4^{4} 1,22,441,2^{2},4^{4} 0 -
W(1)W(2)W(1)\perp W(2) 15,281^{5},2^{8} 14,281^{4},2^{8} 15,281^{5},2^{8} 0 -
W(1)3W(1)^{3} 1211^{21} 1201^{20} 1211^{21} 0 -
44 V(8)V(8) 4,844,8^{4} 3,843,8^{4} 4,844,8^{4} 22 22
V(2)V(6)V(2)\perp V(6) 12,2,4,62,821^{2},2,4,6^{2},8^{2} 1,2,4,62,821,2,4,6^{2},8^{2} 12,2,4,62,821^{2},2,4,6^{2},8^{2} 0 0
V(4)2V(4)^{2} 22,482^{2},4^{8} 1,2,481,2,4^{8} 22,482^{2},4^{8} 11 11
V(2)2V(4)V(2)^{2}\perp V(4) 12,25,461^{2},2^{5},4^{6} 1,25,461,2^{5},4^{6} 1,24,3,461,2^{4},3,4^{6} 0 11
V(2)W(3)V(2)\perp W(3) 12,25,461^{2},2^{5},4^{6} 1,25,461,2^{5},4^{6} 26,462^{6},4^{6} 0 0
W(1)V(2)3W(1)\perp V(2)^{3} 16,2151^{6},2^{15} 15,2151^{5},2^{15} 14,2161^{4},2^{16} 0 0
W(1)3V(2)W(1)^{3}\perp V(2) 122,271^{22},2^{7} 121,271^{21},2^{7} 120,281^{20},2^{8} 0 0
W(2)V(4)W(2)\perp V(4) 12,25,461^{2},2^{5},4^{6} 13,24,461^{3},2^{4},4^{6} 13,23,3,461^{3},2^{3},3,4^{6} 11 11
V(2)4V(2)^{4} 14,2161^{4},2^{16} 13,2161^{3},2^{16} 14,2161^{4},2^{16} 0 0
W(1)2V(4)W(1)^{2}\perp V(4) 110,2,461^{10},2,4^{6} 19,2,461^{9},2,4^{6} 19,3,461^{9},3,4^{6} 0 11
W(1)2V(2)2W(1)^{2}\perp V(2)^{2} 112,2121^{12},2^{12} 111,2121^{11},2^{12} 110,2131^{10},2^{13} 0 0
W(1)V(6)W(1)\perp V(6) 14,4,62,821^{4},4,6^{2},8^{2} 13,4,62,821^{3},4,6^{2},8^{2} 12,2,4,62,821^{2},2,4,6^{2},8^{2} 0 0
W(1)V(2)V(4)W(1)\perp V(2)\perp V(4) 14,24,461^{4},2^{4},4^{6} 13,24,461^{3},2^{4},4^{6} 13,23,3,461^{3},2^{3},3,4^{6} 0 11
W(4)W(4) 22,482^{2},4^{8} 22,3,472^{2},3,4^{7} 1,22,3,471,2^{2},3,4^{7} 22 -
W(1)W(3)W(1)\perp W(3) 14,22,34,441^{4},2^{2},3^{4},4^{4} 13,22,34,441^{3},2^{2},3^{4},4^{4} 14,22,34,441^{4},2^{2},3^{4},4^{4} 0 -
W(2)2W(2)^{2} 14,2161^{4},2^{16} 15,2151^{5},2^{15} 16,2151^{6},2^{15} 11 -
W(1)2W(2)W(1)^{2}\perp W(2) 112,2121^{12},2^{12} 111,2121^{11},2^{12} 112,2121^{12},2^{12} 0 -
W(1)4W(1)^{4} 1361^{36} 1351^{35} 1361^{36} 0 -
Table 2. For Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) simply connected of type CC_{\ell} with 242\leq\ell\leq 4 and charK=2\operatorname{char}K=2, Jordan block sizes of nilpotent e𝔤sce\in\mathfrak{g}_{sc} on 𝔤sc\mathfrak{g}_{sc}, [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}], and 𝔤ad\mathfrak{g}_{ad}. See Example 1.5.
\ell VK[e]V\downarrow K[e] 𝔤scK[e]\mathfrak{g}_{sc}\downarrow K[e] [𝔤ad,𝔤ad]K[e][\mathfrak{g}_{ad},\mathfrak{g}_{ad}]\downarrow K[e] 𝔤adK[e]\mathfrak{g}_{ad}\downarrow K[e] α\alpha
22 V(4)V(4) 12,421^{2},4^{2} 12,3,41^{2},3,4 1,2,3,41,2,3,4 22
V(2)2V(2)^{2} 12,241^{2},2^{4} 13,231^{3},2^{3} 12,241^{2},2^{4} 11
W(1)V(2)W(1)\perp V(2) 14,231^{4},2^{3} 13,231^{3},2^{3} 12,241^{2},2^{4} 0
W(2)W(2) 12,241^{2},2^{4} 13,231^{3},2^{3} 14,231^{4},2^{3} 11
W(1)2W(1)^{2} 1101^{10} 191^{9} 1101^{10} 0
33 V(6)V(6) 13,2,821^{3},2,8^{2} 14,821^{4},8^{2} 13,2,821^{3},2,8^{2} 11
W1(3)W_{1}(3) 15,441^{5},4^{4} 14,441^{4},4^{4} 13,2,441^{3},2,4^{4} 0
V(2)V(4)V(2)\perp V(4) 13,2,441^{3},2,4^{4} 14,441^{4},4^{4} 13,2,441^{3},2,4^{4} 11
W(1)2V(2)W(1)^{2}\perp V(2) 111,251^{11},2^{5} 110,251^{10},2^{5} 19,261^{9},2^{6} 0
W(1)V(2)2W(1)\perp V(2)^{2} 15,281^{5},2^{8} 14,281^{4},2^{8} 13,291^{3},2^{9} 0
W(1)V(4)W(1)\perp V(4) 15,441^{5},4^{4} 14,441^{4},4^{4} 13,2,441^{3},2,4^{4} 0
V(2)3V(2)^{3} 13,291^{3},2^{9} 14,281^{4},2^{8} 13,291^{3},2^{9} 11
W(1)W(2)W(1)\perp W(2) 15,281^{5},2^{8} 14,281^{4},2^{8} 15,281^{5},2^{8} 0
W(3)W(3) 15,441^{5},4^{4} 14,441^{4},4^{4} 15,441^{5},4^{4} 0
W(1)3W(1)^{3} 1211^{21} 1201^{20} 1211^{21} 0
44 V(8)V(8) 14,841^{4},8^{4} 14,7,831^{4},7,8^{3} 13,2,7,831^{3},2,7,8^{3} 33
W1(4)W_{1}(4) 14,481^{4},4^{8} 14,3,471^{4},3,4^{7} 13,2,3,471^{3},2,3,4^{7} 22
W(1)W1(3)W(1)\perp W_{1}(3) 16,23,461^{6},2^{3},4^{6} 15,23,461^{5},2^{3},4^{6} 14,24,461^{4},2^{4},4^{6} 0
V(2)V(6)V(2)\perp V(6) 14,22,62,821^{4},2^{2},6^{2},8^{2} 15,2,62,821^{5},2,6^{2},8^{2} 14,22,62,821^{4},2^{2},6^{2},8^{2} 11
V(4)2V(4)^{2} 14,481^{4},4^{8} 14,3,471^{4},3,4^{7} 13,2,3,471^{3},2,3,4^{7} 22
V(2)2V(4)V(2)^{2}\perp V(4) 14,24,461^{4},2^{4},4^{6} 15,23,461^{5},2^{3},4^{6} 14,24,461^{4},2^{4},4^{6} 11
W(1)V(6)W(1)\perp V(6) 16,2,62,821^{6},2,6^{2},8^{2} 15,2,62,821^{5},2,6^{2},8^{2} 14,22,62,821^{4},2^{2},6^{2},8^{2} 0
W(1)V(2)V(4)W(1)\perp V(2)\perp V(4) 16,23,461^{6},2^{3},4^{6} 15,23,461^{5},2^{3},4^{6} 14,24,461^{4},2^{4},4^{6} 0
W(1)2V(4)W(1)^{2}\perp V(4) 112,461^{12},4^{6} 111,461^{11},4^{6} 110,2,461^{10},2,4^{6} 0
W(1)2V(2)2W(1)^{2}\perp V(2)^{2} 112,2121^{12},2^{12} 111,2121^{11},2^{12} 110,2131^{10},2^{13} 0
W(1)3V(2)W(1)^{3}\perp V(2) 122,271^{22},2^{7} 121,271^{21},2^{7} 120,281^{20},2^{8} 0
W(2)V(4)W(2)\perp V(4) 14,24,461^{4},2^{4},4^{6} 15,23,461^{5},2^{3},4^{6} 14,24,461^{4},2^{4},4^{6} 11
V(2)4V(2)^{4} 14,2161^{4},2^{16} 15,2151^{5},2^{15} 14,2161^{4},2^{16} 11
W(1)V(2)3W(1)\perp V(2)^{3} 16,2151^{6},2^{15} 15,2151^{5},2^{15} 14,2161^{4},2^{16} 0
V(2)W(3)V(2)\perp W(3) 16,23,461^{6},2^{3},4^{6} 15,23,461^{5},2^{3},4^{6} 14,24,461^{4},2^{4},4^{6} 0
W(1)2W(2)W(1)^{2}\perp W(2) 112,2121^{12},2^{12} 111,2121^{11},2^{12} 112,2121^{12},2^{12} 0
W(2)2W(2)^{2} 14,2161^{4},2^{16} 15,2151^{5},2^{15} 16,2151^{6},2^{15} 11
W(1)W(3)W(1)\perp W(3) 18,34,441^{8},3^{4},4^{4} 17,34,441^{7},3^{4},4^{4} 18,34,441^{8},3^{4},4^{4} 0
W(4)W(4) 14,481^{4},4^{8} 14,3,471^{4},3,4^{7} 15,3,471^{5},3,4^{7} 22
W(1)4W(1)^{4} 1361^{36} 1351^{35} 1361^{36} 0

2. Notation and terminology

We will use the following notation and terminology throughout the paper.

2.1. Generalities

We will always assume that KK is an algebraically closed field of characteristic two.

For a KK-vector space VV and integer n>0n>0, we denote Vn=VVV^{n}=V\oplus\cdots\oplus V, where VV occurs nn times. Furthermore, we denote V0=0V^{0}=0.

We will always use GG to denote a group, and all the K[G]K[G]-modules that we consider will be finite-dimensional. Suppose that V=U1U2UtUt+1=0V=U_{1}\supset U_{2}\supset\cdots\supset U_{t}\supset U_{t+1}=0 is the socle filtration of a K[G]K[G]-module VV, so Zi:=Ui/Ui+1=soc(V/Ui+1)Z_{i}:=U_{i}/U_{i+1}=\operatorname{soc}(V/U_{i+1}) for all 1it1\leq i\leq t. Then we will denote this by V=Z1|Z2||ZtV=Z_{1}|Z_{2}|\cdots|Z_{t}, and VV is said to be uniserial if ZiZ_{i} is irreducible for all 1it1\leq i\leq t.

We denote by ν2\nu_{2} the 22-adic valuation on the integers, so ν2(a)\nu_{2}(a) is the largest integer k0k\geq 0 such that 2k2^{k} divides aa.

2.2. Algebraic groups

Suppose that GG is a simple linear algebraic group over KK. In the context of algebraic groups, the notation that we use will be as in [14]. We will denote the Lie algebra of GG by 𝔤\mathfrak{g}.

When GG is an algebraic group, by a GG-module we will always mean a finite-dimensional rational K[G]K[G]-module. Fix a maximal torus TT of GG with character group X(T)X(T), and a base Δ={α1,,α}\Delta=\{\alpha_{1},\ldots,\alpha_{\ell}\} for the root system of GG, where =rankG\ell=\operatorname{rank}G. We will always use the standard Bourbaki labeling of the simple roots αi\alpha_{i}, as given in [11, 11.4, p. 58]. We denote the dominant weights with respect to Δ\Delta by X(T)+X(T)^{+}, and the fundamental dominant weight corresponding to αi\alpha_{i} is denoted by ϖi\varpi_{i}. For a dominant weight λX(T)+\lambda\in X(T)^{+}, we denote the rational irreducible K[G]K[G]-module with highest weight λ\lambda by LG(λ)L_{G}(\lambda).

An element uGu\in G is unipotent, if f(u)f(u) is a unipotent linear map for every rational representation f:GGL(W)f:G\rightarrow\operatorname{GL}(W). Since charK=2\operatorname{char}K=2, equivalently uu is unipotent if and only if uu has order 2α2^{\alpha} for some α0\alpha\geq 0.

Similarly e𝔤e\in\mathfrak{g} is said to be nilpotent, if df(e)\mathrm{d}f(e) is a nilpotent linear map for every rational representation f:GGL(W)f:G\rightarrow\operatorname{GL}(W). Equivalently ee is nilpotent if and only if e[2α]=0e^{[2^{\alpha}]}=0 for some α>0\alpha>0, where xx[2]x\mapsto x^{[2]} is the canonical pp-mapping on 𝔤\mathfrak{g}.

We will mostly be concerned with the case where G=Sp(V)G=\operatorname{Sp}(V), in which case GG is of type CC_{\ell}, where dimV=2\dim V=2\ell. In this case uSp(V)u\in\operatorname{Sp}(V) is unipotent if and only if uu is a unipotent linear map on VV. Furthermore 𝔤=𝔰𝔭(V)\mathfrak{g}=\mathfrak{sp}(V), and e𝔤e\in\mathfrak{g} is nilpotent if and only if ee is a nilpotent linear map on VV.

2.3. Actions of unipotent elements

Let q=2αq=2^{\alpha} for some α0\alpha\geq 0. Then a cyclic group Cq=uC_{q}=\langle u\rangle has qq indecomposable K[Cq]K[C_{q}]-modules, up to isomorphism. We denote them by V1V_{1}, \ldots, VqV_{q}, where dimVi=i\dim V_{i}=i and uu acts on ViV_{i} as a single i×ii\times i unipotent Jordan block. For convenience we denote V0=0V_{0}=0.

Let uu be an element of a group GG and let VV be a finite-dimensional K[G]K[G]-module on which uu acts as a unipotent linear map. We denote Vu:={vV:uv=v}V^{u}:=\{v\in V:uv=v\}, which is the fixed point space of uu on VV. Note that dimVu\dim V^{u} is the number of Jordan blocks of uu on VV.

If uGL(V)u\in\operatorname{GL}(V) is unipotent, we denote by K[u]K[u] the group algebra of u\langle u\rangle. Then VK[u]=Vd1n1VdtntV\downarrow K[u]=V_{d_{1}}^{n_{1}}\oplus\cdots\oplus V_{d_{t}}^{n_{t}}, where 0<d1<<dt0<d_{1}<\cdots<d_{t} are the Jordan block sizes of uu on VV, and nin_{i} is the multiplicity of a Jordan block of size did_{i}.

2.4. Actions of nilpotent elements

Let q=2αq=2^{\alpha} for some α0\alpha\geq 0. We denote by 𝔴q\mathfrak{w}_{q} the abelian 22-Lie algebra over KK generated by a nilpotent element ee with e[2α]=0e^{[2^{\alpha}]}=0 and e[2α1]0e^{[2^{\alpha-1}]}\neq 0. Then as a KK-vector space

𝔴q=0i<αe[2i].\mathfrak{w}_{q}=\bigoplus_{0\leq i<\alpha}\langle e^{[2^{i}]}\rangle.

There are a total of qq indecomposable 𝔴q\mathfrak{w}_{q}-modules, up to isomorphism. We denote them by W1W_{1}, \ldots, WqW_{q}, where dimWi=i\dim W_{i}=i and ee acts on WiW_{i} as a single i×ii\times i nilpotent Jordan block. For convenience we denote W0=0W_{0}=0.

If VV is a finite-dimensional module for a Lie algebra 𝔤\mathfrak{g} and e𝔤e\in\mathfrak{g} acts on VV as a nilpotent element, we denote Ve:={vV:ev=0}V^{e}:=\{v\in V:ev=0\}. Then dimVe\dim V^{e} is the number of Jordan blocks of ee on VV.

If e𝔤𝔩(V)e\in\mathfrak{gl}(V) is nilpotent, we denote by K[e]K[e] the 22-Lie algebra generated by ee. Then K[e]𝔴qK[e]\cong\mathfrak{w}_{q}, where qq is the smallest power of two such that eq=0e^{q}=0. Furthermore VK[e]=Wd1n1WdtntV\downarrow K[e]=W_{d_{1}}^{n_{1}}\oplus\cdots\oplus W_{d_{t}}^{n_{t}}, where 0<d1<<dt0<d_{1}<\cdots<d_{t} are the Jordan block sizes of ee on VV, and nin_{i} is the multiplicity of a Jordan block of size did_{i}.

3. Jordan normal forms on subspaces and quotients

In the proofs of our main results, the basic tool that we use to describe Jordan block sizes are the following two lemmas.

Lemma 3.1 ([16, Lemma 3.3]).

Let e𝔤𝔩(V)e\in\mathfrak{gl}(V) be a nilpotent linear map. Suppose that WVW\subseteq V is a subspace invariant under ee such that dimV/W=1\dim V/W=1. Let m1m\geq 1 be such that Kerem1W\operatorname{Ker}e^{m-1}\subseteq W and KeremW\operatorname{Ker}e^{m}\not\subseteq W. Then

V\displaystyle V WmV\displaystyle\cong W_{m}\oplus V^{\prime}
W\displaystyle W Wm1V\displaystyle\cong W_{m-1}\oplus V^{\prime}

for some K[e]K[e]-module VV^{\prime}.

Lemma 3.2 ([17, Lemma 3.4]).

Let e𝔤𝔩(V)e\in\mathfrak{gl}(V) be a nilpotent linear map. Suppose that WVW\subseteq V is a subspace invariant under ee such that dimW=1\dim W=1. Let m1m\geq 1 be such that Imem1W\operatorname{Im}e^{m-1}\supseteq W and ImemW\operatorname{Im}e^{m}\not\supseteq W. Then

V\displaystyle V WmV\displaystyle\cong W_{m}\oplus V^{\prime}
V/W\displaystyle V/W Wm1V\displaystyle\cong W_{m-1}\oplus V^{\prime}

for some K[e]K[e]-module VV^{\prime}.

Remark 3.3.

Recall that we define W0=0W_{0}=0. Thus if m=1m=1 in Lemma 3.1, we have

VW1W,V\cong W_{1}\oplus W,

and so the Jordan normal form of ee on WW is given by removing a Jordan block of size 11 from the Jordan normal form of ee on VV. (Similar remarks apply to Lemma 3.2.)

4. Decomposition of tensor products and symmetric squares

Let q=2αq=2^{\alpha} for some α0\alpha\geq 0. In this section, we state some basic results about the decomposition of tensor products, exterior squares, and symmetric squares of K[Cq]K[C_{q}]-modules and 𝔴q\mathfrak{w}_{q}-modules.

4.1. Unipotent case

Let uu be a generator for the cyclic group CqC_{q}. To describe the indecomposable summands of tensor products VWV\otimes W of K[Cq]K[C_{q}]-modules, it is clear that it suffices to do so in the case where VV and WW are indecomposable.

The decomposition of VmVnV_{m}\otimes V_{n} into indecomposable summands has been studied extensively in all characteristics, see for example (in chronological order) [31], [28], [24], [30], [26], [10], [27], [12], and [1]. Although there is no closed formula in general, there are various recursive formulae which can be used to determine the indecomposable summands efficiently. We are assuming charK=2\operatorname{char}K=2, in which case we have the following result.

Theorem 4.1 ([7, (2.5a)], [6, Lemma 1, Corollary 3]).

Let 0<mnq0<m\leq n\leq q and suppose that q/2<nqq/2<n\leq q. Then the following statements hold:

  1. (i)

    If n=qn=q, then VmVnVqmV_{m}\otimes V_{n}\cong V_{q}^{m} as K[Cq]K[C_{q}]-modules.

  2. (ii)

    If m+n>qm+n>q, then VmVnVqn+mq(VqnVqm)V_{m}\otimes V_{n}\cong V_{q}^{n+m-q}\oplus(V_{q-n}\otimes V_{q-m}) as K[Cq]K[C_{q}]-modules.

  3. (iii)

    If m+nqm+n\leq q, then VmVnVqdtVqd1V_{m}\otimes V_{n}\cong V_{q-d_{t}}\oplus\cdots\oplus V_{q-d_{1}} as K[Cq]K[C_{q}]-modules, where VmVqnVd1VdtV_{m}\otimes V_{q-n}\cong V_{d_{1}}\oplus\cdots\oplus V_{d_{t}}.

With Theorem 4.1, every tensor product VmVnV_{m}\otimes V_{n} is either described explicitly (case (i)), or in terms of another tensor product VmVnV_{m^{\prime}}\otimes V_{n^{\prime}} with 0<mn<n0<m^{\prime}\leq n^{\prime}<n. Thus by repeated applications of Theorem 4.1, we can rapidly find the indecomposable summands of VmVnV_{m}\otimes V_{n} for any given mm and nn.

For exterior squares and symmetric squares, similarly it suffices to consider the indecomposable case. This follows from the fact that for all K[Cq]K[C_{q}]-modules VV and WW, we have isomorphisms

2(VW)\displaystyle\wedge^{2}(V\oplus W) 2(V)2(W)VW\displaystyle\cong\wedge^{2}(V)\oplus\wedge^{2}(W)\oplus V\otimes W
S2(VW)\displaystyle S^{2}(V\oplus W) S2(V)S2(W)VW\displaystyle\cong S^{2}(V)\oplus S^{2}(W)\oplus V\otimes W

as K[Cq]K[C_{q}]-modules. For the decomposition of 2(Vn)\wedge^{2}(V_{n}) and S2(Vn)S^{2}(V_{n}), we have the following results.

Theorem 4.2 ([6, Theorem 2]).

Suppose that q/2<nqq/2<n\leq q. Then we have

2(Vn)2(Vqn)Vqnq/21V3q/2n\wedge^{2}(V_{n})\cong\wedge^{2}(V_{q-n})\oplus V_{q}^{n-q/2-1}\oplus V_{3q/2-n}

as K[Cq]K[C_{q}]-modules.

Theorem 4.3 ([18, Theorem 1.3]).

Suppose that q/2<nqq/2<n\leq q. Then we have

S2(Vn)2(Vqn)Vqnq/2Vq/2S^{2}(V_{n})\cong\wedge^{2}(V_{q-n})\oplus V_{q}^{n-q/2}\oplus V_{q/2}

as K[Cq]K[C_{q}]-modules.

Similarly to Theorem 4.1, with Theorem 4.2 and Theorem 4.3 we can quickly decompose 2(Vn)\wedge^{2}(V_{n}) and S2(Vn)S^{2}(V_{n}) for any given nn.

Lemma 4.4.

Let n>0n>0 be an integer. Then the following hold:

  1. (i)

    dim2(Vn)u=n2\dim\wedge^{2}(V_{n})^{u}=\lfloor\frac{n}{2}\rfloor.

  2. (ii)

    dimS2(Vn)u=n2+1\dim S^{2}(V_{n})^{u}=\lfloor\frac{n}{2}\rfloor+1.

Proof.

We first consider (i). If n=1n=1, then 2(V1)=0\wedge^{2}(V_{1})=0 has dimension 0 and thus (i) holds. Suppose then that n>1n>1 and proceed by induction on nn. Let qq be a power of two such that q/2<nqq/2<n\leq q. If n=qn=q, it follows from Theorem 4.2 that 2(Vn)Vqnq/21Vq/2\wedge^{2}(V_{n})\cong V_{q}^{n-q/2-1}\oplus V_{q/2}, so dim2(Vn)u=nq/2=n/2\dim\wedge^{2}(V_{n})^{u}=n-q/2=n/2. If q/2<n<qq/2<n<q, we have

2(Vn)2(Vqn)Vqnq/21V3q/2n\wedge^{2}(V_{n})\cong\wedge^{2}(V_{q-n})\oplus V_{q}^{n-q/2-1}\oplus V_{3q/2-n}

by Theorem 4.2. Then by induction

dim2(Vn)u=qn2+nq/2=n2,\dim\wedge^{2}(V_{n})^{u}=\left\lfloor\frac{q-n}{2}\right\rfloor+n-q/2=\left\lfloor\frac{n}{2}\right\rfloor,

as claimed by (i).

Next we will prove (ii). If n=1n=1, then S2(V1)=V1S^{2}(V_{1})=V_{1} so (ii) holds. Suppose then that n>1n>1, and let qq be a power of two such that q/2<nqq/2<n\leq q. If n=qn=q, then S2(Vn)Vqnq/2Vq/2S^{2}(V_{n})\cong V_{q}^{n-q/2}\oplus V_{q/2}, and thus dimS2(Vn)u=nq/2+1=n/2+1\dim S^{2}(V_{n})^{u}=n-q/2+1=n/2+1. Suppose then that q/2<n<qq/2<n<q. It follows from Theorem 4.3 and (i) that

dimS2(Vn)u=qn2+nq/2+1=n2+1,\dim S^{2}(V_{n})^{u}=\left\lfloor\frac{q-n}{2}\right\rfloor+n-q/2+1=\left\lfloor\frac{n}{2}\right\rfloor+1,

as claimed by (ii).∎

Lemma 4.5.

Let >0\ell>0 and define α=ν2()\alpha=\nu_{2}(\ell). Then the smallest Jordan block size in S2(V2)S^{2}(V_{2\ell}) is 2α2^{\alpha}, occurring with multiplicity one.

Proof.

(cf. [17, Lemma 4.12]) If =2α\ell=2^{\alpha}, it follows from Theorem 4.3 that S2(V2)=V2α+12αV2αS^{2}(V_{2\ell})=V_{2^{\alpha+1}}^{2^{\alpha}}\oplus V_{2^{\alpha}}, so the claim holds. If 2α\ell\neq 2^{\alpha}, we have q/2<2<qq/2<2\ell<q for some q=2βq=2^{\beta}. Then

(4.1) S2(V2)=2(Vq2)Vq2q/2Vq/2S^{2}(V_{2\ell})=\wedge^{2}(V_{q-2\ell})\oplus V_{q}^{2\ell-q/2}\oplus V_{q/2}

by Theorem 4.3.

Now ν2(q2)=2α+1\nu_{2}(q-2\ell)=2^{\alpha+1} since q>2α+1q>2^{\alpha+1}, so by [17, Lemma 4.12] the smallest Jordan block size in 2(Vq2)\wedge^{2}(V_{q-2\ell}) is 2α2^{\alpha}, occurring with multiplicity one. Furthermore 2α<q/22^{\alpha}<q/2 because q/2<2<qq/2<2\ell<q, so the lemma follows from (4.1).∎

4.2. Nilpotent case

As in the previous section, to determine the indecomposable summands of tensor products, exterior squares, and symmetric squares of 𝔴q\mathfrak{w}_{q}-modules, it suffices to do so in the indecomposable case.

For the indecomposable summands of WmWnW_{m}\otimes W_{n}, it turns out that we get the same decomposition as for VmVnV_{m}\otimes V_{n} in the unipotent case.

Proposition 4.6 ([4, Section III], [25, Corollary 5 (a)]).

Let 0<n,mq0<n,m\leq q and suppose that we have VmVnVr1VrtV_{m}\otimes V_{n}\cong V_{r_{1}}\oplus\cdots\oplus V_{r_{t}} as K[Cq]K[C_{q}]-modules for some r1,,rt>0r_{1},\ldots,r_{t}>0. Then WmWnWr1WrtW_{m}\otimes W_{n}\cong W_{r_{1}}\oplus\cdots\oplus W_{r_{t}} as 𝔴q\mathfrak{w}_{q}-modules.

Thus we can apply Theorem 4.1 to find the decomposition of WmWnW_{m}\otimes W_{n} into indecomposable summands.

Following [5, p. 231], we call the consecutive-ones binary expansion of an integer n>0n>0 the alternating sum n=1ir(1)i+12βin=\sum_{1\leq i\leq r}(-1)^{i+1}2^{\beta_{i}} such that β1>>βr0\beta_{1}>\cdots>\beta_{r}\geq 0 and rr is minimal. (Here βr1>βr+1\beta_{r-1}>\beta_{r}+1 if r>1r>1.)

The decomposition of VnVnV_{n}\otimes V_{n} into indecomposable summands can be given explicitly in terms of the consecutive-ones binary expansion of nn [5, Theorem 15]. Such descriptions can also be given for 2(Vn)\wedge^{2}(V_{n}) and S2(Vn)S^{2}(V_{n}), by using Theorem 4.2 and Theorem 4.3.

For the decomposition of WnWnW_{n}\otimes W_{n}, 2(Wn)\wedge^{2}(W_{n}), and S2(Wn)S^{2}(W_{n}) we have the following result.

Theorem 4.7 ([18, Theorem 1.6, Theorem 1.7, Theorem 3.7]).

Let n>0n>0 be an integer, with consecutive-ones binary expansion n=1ir(1)i+12βin=\sum_{1\leq i\leq r}(-1)^{i+1}2^{\beta_{i}}, where β1>>βr0\beta_{1}>\cdots>\beta_{r}\geq 0. For 1kr1\leq k\leq r , define dk:=2βk+k<ir(1)k+i2βi+1d_{k}:=2^{\beta_{k}}+\sum_{k<i\leq r}(-1)^{k+i}2^{\beta_{i}+1}. Then

WnWn\displaystyle W_{n}\otimes W_{n} 1krW2βkdk\displaystyle\cong\bigoplus_{1\leq k\leq r}W_{2^{\beta_{k}}}^{d_{k}}
2(Wn)\displaystyle\wedge^{2}(W_{n}) 1krβk>0W2βk1dk/2\displaystyle\cong\bigoplus_{\begin{subarray}{c}1\leq k\leq r\\ \beta_{k}>0\end{subarray}}W_{2^{\beta_{k}}-1}^{d_{k}/2}
S2(Wn)\displaystyle S^{2}(W_{n}) W1n/21krβk>0W2βkdk/2\displaystyle\cong W_{1}^{\lceil n/2\rceil}\oplus\bigoplus_{\begin{subarray}{c}1\leq k\leq r\\ \beta_{k}>0\end{subarray}}W_{2^{\beta_{k}}}^{d_{k}/2}

as 𝔴q\mathfrak{w}_{q}-modules.

5. Unipotent elements in Sp(V)\operatorname{Sp}(V)

Consider Sp(V)\operatorname{Sp}(V) with dimV=2\dim V=2\ell. In this section, we will recall the description of unipotent conjugacy classes in Sp(V)\operatorname{Sp}(V), due to Hesselink [8]. For more details, see for example [8], [23, Chapter 4, Chapter 6], or [17, Section 6].

For a group GG, a bilinear K[G]K[G]-module (W,β)(W,\beta) is a finite-dimensional K[G]K[G]-module WW equipped with a GG-invariant bilinear form β\beta. Two bilinear K[G]K[G]-modules (W,β)(W,\beta) and (W,β)(W^{\prime},\beta^{\prime}) are said to be isomorphic, if there exists an isomorphism WWW\rightarrow W^{\prime} of K[G]K[G]-modules which is also an isometry.

Let u,uSp(V)u,u^{\prime}\in\operatorname{Sp}(V) be unipotent. Let qq be a power of two such that uq=(u)q=1u^{q}=(u^{\prime})^{q}=1, so that VK[u]V\downarrow K[u] and VK[u]V\downarrow K[u^{\prime}] are K[Cq]K[C_{q}]-modules. Then uu and uu^{\prime} are conjugate in Sp(V)\operatorname{Sp}(V) if and only if VK[u]VK[u]V\downarrow K[u]\cong V\downarrow K[u^{\prime}] as bilinear K[Cq]K[C_{q}]-modules.

For uSp(V)u\in\operatorname{Sp}(V) unipotent, it is clear that we can write VK[u]=U1UtV\downarrow K[u]=U_{1}\perp\cdots\perp U_{t}, where UiU_{i} are orthogonally indecomposable K[u]K[u]-modules. Here orthogonally indecomposable means that if Ui=UU′′U_{i}=U^{\prime}\perp U^{\prime\prime} as K[u]K[u]-modules, then U=0U^{\prime}=0 or U′′=0U^{\prime\prime}=0. There are two basic types of orthogonally indecomposable K[u]K[u]-modules, which we can define as follows. (Similar definitions are given in [23, Section 6.1].)

Definition 5.1.

For 1\ell\geq 1, we define the module V(2)V(2\ell) as follows. Let n=2n=2\ell, and suppose that VV has basis v1v_{1}, \ldots, vnv_{n} with b(vi,vj)=1b(v_{i},v_{j})=1 if i+j=n+1i+j=n+1 and 0 otherwise. Define u:VVu:V\rightarrow V by

uv1\displaystyle uv_{1} =v1\displaystyle=v_{1}
uvi\displaystyle uv_{i} =vi+vi1++v1 for all 1<i+1\displaystyle=v_{i}+v_{i-1}+\cdots+v_{1}\text{ for all }1<i\leq\ell+1
uvi\displaystyle uv_{i} =vi+vi1 for all +1<in.\displaystyle=v_{i}+v_{i-1}\text{ for all }\ell+1<i\leq n.

Then uSp(V)u\in\operatorname{Sp}(V), and we define V(2)V(2\ell) as the bilinear K[u]K[u]-module VK[u]V\downarrow K[u].

Definition 5.2.

For 1\ell\geq 1, we define the module W()W(\ell) as follows. Let n=2n=2\ell, and suppose that VV has basis v1v_{1}, \ldots, vnv_{n} with b(vi,vj)=1b(v_{i},v_{j})=1 if i+j=n+1i+j=n+1 and 0 otherwise. Define u:VVu:V\rightarrow V by

uv1\displaystyle uv_{1} =v1\displaystyle=v_{1}
uvi\displaystyle uv_{i} =vi+vi1++v1 for all 1<i\displaystyle=v_{i}+v_{i-1}+\cdots+v_{1}\text{ for all }1<i\leq\ell
uv+1\displaystyle uv_{\ell+1} =v+1\displaystyle=v_{\ell+1}
uvi\displaystyle uv_{i} =vi+vi1 for all +1<in.\displaystyle=v_{i}+v_{i-1}\text{ for all }\ell+1<i\leq n.

Then uSp(V)u\in\operatorname{Sp}(V), and we define W()W(\ell) as the bilinear K[u]K[u]-module VK[u]V\downarrow K[u].

The fact that the modules are isomorphic to those described by Hesselink [8, Proposition 3.5] is seen as follows.

  • For the module V(2)V(2\ell) in Definition 5.1 we have V(2)K[u]=V2V(2\ell)\downarrow K[u]=V_{2\ell}, so this agrees with [8, Proposition 3.5].

  • In Definition 5.2, we have totally singular decomposition V=WZV=W\oplus Z, where W=v1,,vW=\langle v_{1},\ldots,v_{\ell}\rangle and Z=v+1,,vnZ=\langle v_{\ell+1},\ldots,v_{n}\rangle are K[u]K[u]-modules with WZVW\cong Z\cong V_{\ell}. From this it follows that VV is isomorphic to the module W()W(\ell) defined by Hesselink, see for example [17, Lemma 6.12].

The classification of unipotent conjugacy classes in Sp(V)\operatorname{Sp}(V) is based on the following result.

Theorem 5.3 ([8, Proposition 3.5]).

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent such that VK[u]V\downarrow K[u] is orthogonally indecomposable. Then VK[u]V\downarrow K[u] is isomorphic to V(2)V(2\ell) or W()W(\ell), where dimV=2\dim V=2\ell.

By Theorem 5.3, for every unipotent uSp(V)u\in\operatorname{Sp}(V) we have an orthogonal decomposition

VK[u]=U1Ut,V\downarrow K[u]=U_{1}\perp\cdots\perp U_{t},

where for all 1it1\leq i\leq t we have UiV(2i)U_{i}\cong V(2\ell_{i}) or UiW(i)U_{i}\cong W(\ell_{i}) for some integer i1\ell_{i}\geq 1.

In general there can be several different ways to decompose VK[u]V\downarrow K[u] into orthogonally indecomposable summands, and even the number of summands is not uniquely determined. This is due to the fact that for even mm, we have an isomorphism

W(m)V(m)V(m)V(m)V(m)W(m)\perp V(m)\cong V(m)\perp V(m)\perp V(m)

of bilinear K[u]K[u]-modules. However, there are normal forms which are uniquely determined, such as the Hesselink normal form [8, 3.7] [17, Theorem 6.4] or the distinguished normal form defined by Liebeck and Seitz in [23, p. 61].

6. Nilpotent elements in 𝔰𝔭(V)\mathfrak{sp}(V)

We consider G=Sp(V)G=\operatorname{Sp}(V) with Lie algebra 𝔰𝔭(V)\mathfrak{sp}(V), where dimV=2\dim V=2\ell. We recall the classification of nilpotent orbits in 𝔰𝔭(V)\mathfrak{sp}(V) due to Hesselink [8]. For more details, we refer to [8] and [23, Chapter 4, Chapter 5].

For a nilpotent element e𝔰𝔭(V)e\in\mathfrak{sp}(V), define the index function χV:00\chi_{V}:\mathbb{Z}_{\geq 0}\rightarrow\mathbb{Z}_{\geq 0} corresponding to ee by

χV(m):=min{n0:b(en+1v,env)=0 for all vKerem}.\chi_{V}(m):=\operatorname{min}\{n\geq 0:b(e^{n+1}v,e^{n}v)=0\text{ for all }v\in\operatorname{Ker}e^{m}\}.

Let 0<d1<<dt0<d_{1}<\cdots<d_{t} be the Jordan block sizes of ee, and let nin_{i} be the multiplicity of Jordan block size did_{i} for ee. By a result of Hesselink [8, Theorem 3.8], the nilpotent orbit of ee is determined by the integers did_{i}, nin_{i}, and the function χV\chi_{V}.

Hesselink also proved that it suffices to only consider the values of χV\chi_{V} on the Jordan block sizes d1d_{1}, \ldots, dtd_{t}.

Theorem 6.1 ([8, 3.9]).

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent with index function χ=χV\chi=\chi_{V} on VV. Let 0<d1<<dt0<d_{1}<\cdots<d_{t} be the Jordan block sizes of ee, and let nin_{i} be the multiplicity of a Jordan block of size did_{i} for ee. Then the following statements hold:

  1. (i)

    The nilpotent orbit of ee is determined by the symbol (d1χ(d1)n1,,dtχ(dt)nt)({d_{1}}_{\chi(d_{1})}^{n_{1}},\ldots,{d_{t}}_{\chi(d_{t})}^{n_{t}}).

  2. (ii)

    χ(d1)χ(dt)\chi(d_{1})\leq\cdots\leq\chi(d_{t}) and d1χ(d1)dtχ(dt)d_{1}-\chi(d_{1})\leq\cdots\leq d_{t}-\chi(d_{t}).

  3. (iii)

    0χ(di)di/20\leq\chi(d_{i})\leq d_{i}/2 for all 1it1\leq i\leq t.

  4. (iv)

    χ(di)=di/2\chi(d_{i})=d_{i}/2 if nin_{i} is odd.

Remark 6.2.

Conversely, consider integers did_{i}, nin_{i}, χ(di)\chi(d_{i}) with 0<d1<<dt0<d_{1}<\cdots<d_{t} and i=1tnidi=dimV\sum_{i=1}^{t}n_{i}d_{i}=\dim V, such that conditions (ii) – (iv) of Theorem 6.1 hold. Then there exists a nilpotent element e𝔰𝔭(V)e\in\mathfrak{sp}(V) with corresponding symbol (d1χ(d1)n1,,dtχ(dt)nt)({d_{1}}_{\chi(d_{1})}^{n_{1}},\ldots,{d_{t}}_{\chi(d_{t})}^{n_{t}}) [8, 3.9].

Similarly to the unipotent case, we can phrase the classification in terms of bilinear modules. For a Lie algebra 𝔴\mathfrak{w}, a bilinear 𝔴\mathfrak{w}-module (W,β)(W,\beta) is a finite-dimensional 𝔴\mathfrak{w}-module WW equipped with a 𝔴\mathfrak{w}-invariant bilinear form β\beta, so β(Xv,w)+β(v,Xw)=0\beta(Xv,w)+\beta(v,Xw)=0 for all X𝔴X\in\mathfrak{w} and v,wWv,w\in W. Two bilinear 𝔴\mathfrak{w}-modules (W,β)(W,\beta) and (W,β)(W^{\prime},\beta^{\prime}) are said to be isomorphic, if there exists an isomorphism WWW\rightarrow W^{\prime} of 𝔴\mathfrak{w}-modules which is also an isometry.

Let e,e𝔰𝔭(V)e,e^{\prime}\in\mathfrak{sp}(V) be nilpotent. Choose a power of two qq such that eq=(e)q=0e^{q}=(e^{\prime})^{q}=0, so that VK[e]V\downarrow K[e] and VK[e]V\downarrow K[e^{\prime}] are 𝔴q\mathfrak{w}_{q}-modules. Then ee and ee^{\prime} are conjugate under the action of Sp(V)\operatorname{Sp}(V) if and only if VK[e]VK[e]V\downarrow K[e]\cong V\downarrow K[e^{\prime}] as bilinear 𝔴q\mathfrak{w}_{q}-modules.

For nilpotent e𝔰𝔭(V)e\in\mathfrak{sp}(V), it is clear that we can write VK[e]=V1VtV\downarrow K[e]=V_{1}\perp\cdots\perp V_{t}, where ViV_{i} are orthogonally indecomposable K[e]K[e]-modules. (Here orthogonally indecomposable is defined similarly to the group case.) By the next lemma, the index function χV\chi_{V} is determined by its restriction to the orthogonally indecomposable summands of VK[e]V\downarrow K[e].

Lemma 6.3 ([23, Lemma 5.2]).

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent and assume VK[e]=W1W2V\downarrow K[e]=W_{1}\perp W_{2} as K[e]K[e]-modules. Then χV(m)=max{χW1(m),χW2(m)}\chi_{V}(m)=\max\{\chi_{W_{1}}(m),\chi_{W_{2}}(m)\} for all m0m\geq 0.

The orthogonally indecomposable modules were classified by Hesselink. In the case of 𝔰𝔭(V)\mathfrak{sp}(V), there are three types of orthogonally indecomposable modules, defined as follows. (Similar definitions are given in [23, Section 5.1].)

Definition 6.4.

For 1\ell\geq 1, we define the module V(2)V(2\ell) as follows. Let n=2n=2\ell, and suppose that VV has basis v1v_{1}, \ldots, vnv_{n} with b(vi,vj)=1b(v_{i},v_{j})=1 if i+j=n+1i+j=n+1 and 0 otherwise. Define e:VVe:V\rightarrow V by

ev1\displaystyle ev_{1} =0\displaystyle=0
evi\displaystyle ev_{i} =vi1 for all 1<in.\displaystyle=v_{i-1}\text{ for all }1<i\leq n.

Then e𝔰𝔭(V)e\in\mathfrak{sp}(V), and we define V(2)V(2\ell) as the bilinear K[e]K[e]-module VK[e]V\downarrow K[e].

Definition 6.5.

For 1\ell\geq 1, we define the module W()W(\ell) as follows. Let n=2n=2\ell, and suppose that VV has basis v1v_{1}, \ldots, vnv_{n} with b(vi,vj)=1b(v_{i},v_{j})=1 if i+j=n+1i+j=n+1 and 0 otherwise. Define e:VVe:V\rightarrow V by

ev1\displaystyle ev_{1} =0,\displaystyle=0, evi\displaystyle ev_{i} =vi1 for all 1<i.\displaystyle=v_{i-1}\text{ for all }1<i\leq\ell.
ev+1\displaystyle ev_{\ell+1} =0,\displaystyle=0, evi\displaystyle ev_{i} =vi1 for all +1<in.\displaystyle=v_{i-1}\text{ for all }\ell+1<i\leq n.

Then e𝔰𝔭(V)e\in\mathfrak{sp}(V), and we define W()W(\ell) as the bilinear K[e]K[e]-module VK[e]V\downarrow K[e].

Definition 6.6.

For 1\ell\geq 1 and 0<k</20<k<\ell/2 we define the module Wk()W_{k}(\ell) as follows. Let n=2n=2\ell, and suppose that VV has basis v1v_{1}, \ldots, vnv_{n} with b(vi,vj)=1b(v_{i},v_{j})=1 if i+j=n+1i+j=n+1 and 0 otherwise. Define e:VVe:V\rightarrow V by

ev1\displaystyle ev_{1} =0,\displaystyle=0, ev+1\displaystyle ev_{\ell+1} =0\displaystyle=0
evnk+1\displaystyle ev_{n-k+1} =vnk+vk\displaystyle=v_{n-k}+v_{k} evi\displaystyle ev_{i} =vi1 for all i{1,+1,nk+1}\displaystyle=v_{i-1}\text{ for all }i\not\in\{1,\ell+1,n-k+1\}

Then e𝔰𝔭(V)e\in\mathfrak{sp}(V), and we define Wk()W_{k}(\ell) as the bilinear K[e]K[e]-module VK[e]V\downarrow K[e].

The fact that the modules in Definition 6.46.6 agree with those described by Hesselink in [8, Proposition 3.5] is seen as follows.

  • In Definition 6.4, this is clear from the fact that VK[e]=W2V\downarrow K[e]=W_{2\ell}, so VK[e]=V(2)V\downarrow K[e]=V(2\ell) as defined by Hesselink.

  • In Definition 6.5, we have VK[e]=WWV\downarrow K[e]=W_{\ell}\oplus W_{\ell}. It is easy to see that b(ev,v)=0b(ev,v)=0 for all vVv\in V, so χV(m)=0\chi_{V}(m)=0 for all m0m\geq 0. It follows from [8, Proposition 3.5, Theorem 3.8] that VK[e]V\downarrow K[e] is isomorphic to the module W()W(\ell) defined by Hesselink.

  • In Definition 6.6 we have used the representative given in [20], which gives Wk()W_{k}(\ell) by [20, Lemma 3.4].

Theorem 6.7 ([8, Proposition 3.5]).

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent such that VK[e]V\downarrow K[e] is orthogonally indecomposable. Then VK[e]V\downarrow K[e] is isomorphic to V(2)V(2\ell), W()W(\ell), or Wk()W_{k}(\ell) for some 0<k</20<k<\ell/2. These modules are characterized by the following properties:

VK[e]V\downarrow K[e] Jordan normal form on VV χV\chi_{V}
V(2)V(2\ell) W2W_{2\ell} χV(2)=\chi_{V}(2\ell)=\ell
W()W(\ell) WWW_{\ell}\oplus W_{\ell} χV()=0\chi_{V}(\ell)=0
Wk()W_{k}(\ell) (0<k</20<k<\ell/2) WWW_{\ell}\oplus W_{\ell} χV()=k\chi_{V}(\ell)=k

By Theorem 6.7, for every nilpotent e𝔰𝔭(V)e\in\mathfrak{sp}(V) we have an orthogonal decomposition

VK[e]=U1Ut,V\downarrow K[e]=U_{1}\perp\cdots\perp U_{t},

where for all 1it1\leq i\leq t we have UiV(2i)U_{i}\cong V(2\ell_{i}), UiW(i)U_{i}\cong W(\ell_{i}), or UiWki(i)U_{i}\cong W_{k_{i}}(\ell_{i}) (0<ki<i/20<k_{i}<\ell_{i}/2) for some integer i1\ell_{i}\geq 1.

As in the unipotent case, the orthogonally indecomposable summands and their number is not uniquely determined. For example, by Lemma 6.3 and Theorem 6.1 (i), we have isomorphisms

W(d)V(d)\displaystyle W(d)\perp V(d^{\prime}) Wd/2(d)V(d)\displaystyle\cong W_{d^{\prime}/2}(d)\perp V(d^{\prime}) for d>d>0 even;\displaystyle\text{ for }d>d^{\prime}>0\text{ even};
W(d)V(d)\displaystyle W(d)\perp V(d) V(d)V(d)V(d)\displaystyle\cong V(d)\perp V(d)\perp V(d) for d>0 even;\displaystyle\text{ for }d>0\text{ even};

of bilinear K[e]K[e]-modules.

We end this section with two observations about orthogonally indecomposable modules of the form 1itW(mi)\sum_{1\leq i\leq t}W(m_{i}).

Lemma 6.8.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent. Then the following statements are equivalent:

  1. (i)

    There is an orthogonal decomposition VK[e]=1itW(mi)V\downarrow K[e]=\sum_{1\leq i\leq t}W(m_{i}) for some integers m1m_{1}, \ldots, mtm_{t}.

  2. (ii)

    There is a totally singular decomposition V=WZV=W\oplus Z, where WW and ZZ are K[e]K[e]-submodules of VV.

  3. (iii)

    b(ev,v)=0b(ev,v)=0 for all vVv\in V.

Proof.

(cf. [17, Lemma 6.12]) We prove that (i) \Rightarrow (ii) \Rightarrow (iii) \Rightarrow (i).

(i) \Rightarrow (ii): It is clear from Definition 6.5 that each W(mi)W(m_{i}) has a decomposition W(mi)=WiZiW(m_{i})=W_{i}\oplus Z_{i}, where WiW_{i} and ZiZ_{i} are totally singular K[e]K[e]-submodules with WiWmiZiW_{i}\cong W_{m_{i}}\cong Z_{i}. Therefore (i) implies (ii).

(ii) \Rightarrow (iii): We have b(ew,w)=b(ez,z)=0b(ew,w)=b(ez,z)=0 for all wWw\in W and zZz\in Z since WW and ZZ are ee-invariant and totally singular. Thus b(e(w+z),w+z)=b(ew,z)+b(ez,w)=0b(e(w+z),w+z)=b(ew,z)+b(ez,w)=0 for all wWw\in W and zZz\in Z, since e𝔰𝔭(V)e\in\mathfrak{sp}(V).

(iii) \Rightarrow (i): Suppose that b(ev,v)=0b(ev,v)=0 for all vVv\in V. Then for the index function of ee we have χV(m)=0\chi_{V}(m)=0 for all m1m\geq 1. Therefore every orthogonally indecomposable summand of VK[e]V\downarrow K[e] must be of the form W(m)W(m) for some m1m\geq 1 (Theorem 6.7 and Lemma 6.3), so (i) holds.∎

Lemma 6.9.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent. Then VK[e2]=1itW(mi)V\downarrow K[e^{2}]=\sum_{1\leq i\leq t}W(m_{i}) for some integers m1m_{1}, \ldots, mtm_{t}.

Proof.

We have b(e2v,v)+b(ev,ev)=0b(e^{2}v,v)+b(ev,ev)=0 for all vVv\in V since e𝔰𝔭(V)e\in\mathfrak{sp}(V), so b(e2v,v)=0b(e^{2}v,v)=0 for all vVv\in V. Now the claim follows from Lemma 6.8.∎

7. Chevalley construction

In this section, we will recall basics of the Chevalley construction of Lie algebras and simple algebraic groups, and in particular how it applies for groups of type CC_{\ell} in characteristic two. For more details, see for example [32] or [11, Chapter VII]. We will also make some preliminary observations about the actions of unipotent and nilpotent elements on the adjoint Lie algebra 𝔤ad\mathfrak{g}_{ad} of type CC_{\ell}.

7.1. Chevalley construction

Let 𝔤\mathfrak{g} be a finite-dimensional simple Lie algebra over \mathbb{C}. Fix a Cartan subalgebra 𝔥\mathfrak{h} of 𝔤\mathfrak{g}, and let Φ\Phi be the corresponding root system, so

𝔤=𝔥αΦ𝔤α,\mathfrak{g}=\mathfrak{h}\oplus\bigoplus_{\alpha\in\Phi}\mathfrak{g}_{\alpha},

where 𝔤α={X𝔤:[H,X]=α(H)X for all H𝔥.}.\mathfrak{g}_{\alpha}=\{X\in\mathfrak{g}:[H,X]=\alpha(H)X\text{ for all }H\in\mathfrak{h}.\}. The Killing form κ\kappa is non-degenerate on 𝔥\mathfrak{h}, so for all αΦ\alpha\in\Phi there exists Hα𝔥H_{\alpha}^{\prime}\in\mathfrak{h} such that κ(H,Hα)=α(H)\kappa(H,H_{\alpha}^{\prime})=\alpha(H) for all H𝔥H\in\mathfrak{h}. We define Hα:=2κ(Hα,Hα)HαH_{\alpha}:=\frac{2}{\kappa(H_{\alpha}^{\prime},H_{\alpha}^{\prime})}H_{\alpha}^{\prime} for all αΦ\alpha\in\Phi.

It was shown by Chevalley [2, Théorème 1] that one can choose Xα𝔤αX_{\alpha}\in\mathfrak{g}_{\alpha} such that the following properties hold:

  1. (a)

    [Xα,Xα]=Hα[X_{\alpha},X_{-\alpha}]=H_{\alpha} for all αΦ\alpha\in\Phi,

  2. (b)

    If α,βΦ\alpha,\beta\in\Phi and α+βΦ\alpha+\beta\in\Phi, then [Xα,Xβ]=±(r+1)Xα+β[X_{\alpha},X_{\beta}]=\pm(r+1)X_{\alpha+\beta}, where r0r\geq 0 is the largest integer such that βrα\beta-r\alpha is a root.

  3. (c)

    If α,βΦ\alpha,\beta\in\Phi and α+βΦ\alpha+\beta\not\in\Phi, then [Xα,Xβ]=0[X_{\alpha},X_{\beta}]=0.

For a choice of root vectors XαX_{\alpha} satisfying (a) – (c) above, we define 𝔤\mathfrak{g}_{\mathbb{Z}} to be the \mathbb{Z}-span of all XαX_{\alpha} and HαH_{\alpha} for αΦ\alpha\in\Phi. Let Δ\Delta be a base for Φ\Phi and let Φ+\Phi^{+} be the corresponding system of positive roots. Then {Xα:αΦ}{Hα:αΔ}\{X_{\alpha}:\alpha\in\Phi\}\cup\{H_{\alpha}:\alpha\in\Delta\} is a \mathbb{Z}-basis of 𝔤\mathfrak{g}_{\mathbb{Z}}, called a Chevalley basis for 𝔤\mathfrak{g}.

Fix a Chevalley basis for 𝔤\mathfrak{g}. Let 𝒰\mathscr{U}_{\mathbb{Z}} be the corresponding Kostant \mathbb{Z}-form, which is the subring of the universal enveloping algebra generated by 11 and Xαkk!\frac{X_{\alpha}^{k}}{k!} for all αΦ\alpha\in\Phi and k1k\geq 1.

Let VV be a faithful finite-dimensional 𝔤\mathfrak{g}-module over \mathbb{C}. We denote the set of weights of 𝔥\mathfrak{h} in VV by Λ(V)\Lambda(V). Then ΦΛ(V)Λ\mathbb{Z}\Phi\subseteq\mathbb{Z}\Lambda(V)\subseteq\Lambda, where Λ\Lambda is the weight lattice.

A lattice in VV is the \mathbb{Z}-span of a basis of VV. We say that a lattice LVL\subseteq V is admissible, if LL is 𝒰\mathscr{U}_{\mathbb{Z}}-invariant.

Let FF be an algebraically closed field of characteristic p>0p>0, and LL an admissible lattice in VV. Set VF:=FLV_{F}:=F\otimes_{\mathbb{Z}}L. Then for all X𝒰X\in\mathscr{U}_{\mathbb{Z}}, the reduction modulo pp of XX is the FF-linear map defined by 1X:VFVF1\otimes X^{\prime}:V_{F}\rightarrow V_{F}, where X:LLX^{\prime}:L\rightarrow L is the action of XX on LL.

In particular, for all αΦ\alpha\in\Phi and k0k\geq 0 we have a linear map Xα,k:VFVFX_{\alpha,k}:V_{F}\rightarrow V_{F} which is the reduction modulo pp of Xαk/k!X_{\alpha}^{k}/k! on VFV_{F}. Since XαX_{\alpha} acts nilpotently on VV, we can define for all tFt\in F the root element xα(t)x_{\alpha}(t) as the exponential xα(t):=k0tkXα,kx_{\alpha}(t):=\sum_{k\geq 0}t^{k}X_{\alpha,k}. We have xα(t)GL(VF)x_{\alpha}(t)\in\operatorname{GL}(V_{F}), and the Chevalley group (over FF) corresponding to VV and LL is defined as

G(V,L):=xα(t):αΦ,tF.G(V,L):=\langle x_{\alpha}(t):\alpha\in\Phi,t\in F\rangle.

Then G(V,L)G(V,L) is a simple algebraic group over FF with root system Φ\Phi [32, Theorem 6]. Furthermore, we have a maximal torus

T=hα(t):αΔ,tK×,T=\langle h_{\alpha}(t):\alpha\in\Delta,t\in K^{\times}\rangle,

where hα(t)h_{\alpha}(t) is defined as in [32, Lemma 19, p. 22]. Then the weights of TT on VV can be identified with Λ(V)\Lambda(V), and the character group X(T)X(T) can be identified with Λ(V)\mathbb{Z}\Lambda(V) [32, p. 39].

We say that G(V,L)G(V,L) is simply connected if Λ(V)=Λ\mathbb{Z}\Lambda(V)=\Lambda, and adjoint if Λ(V)=Φ\mathbb{Z}\Lambda(V)=\mathbb{Z}\Phi.

Lemma 7.1 ([32, Corollary 1, p. 41]).

Suppose that VV^{\prime} is another faithful finite-dimensional 𝔤\mathfrak{g}-module with admissible lattice LL^{\prime}, and denote the corresponding root elements in G(V,L)G(V^{\prime},L^{\prime}) by xα(t)x_{\alpha}^{\prime}(t). Then:

  1. (i)

    If G(V,L)G(V^{\prime},L^{\prime}) is simply connected, there exists a morphism φ:G(V,L)G(V,L)\varphi:G(V^{\prime},L^{\prime})\rightarrow G(V,L) of algebraic groups with xα(t)xα(t)x_{\alpha}^{\prime}(t)\mapsto x_{\alpha}(t) for all αΦ\alpha\in\Phi and tFt\in F.

  2. (ii)

    If G(V,L)G(V^{\prime},L^{\prime}) is adjoint, there exists a morphism φ:G(V,L)G(V,L)\varphi:G(V,L)\rightarrow G(V^{\prime},L^{\prime}) of algebraic groups with xα(t)xα(t)x_{\alpha}(t)\mapsto x_{\alpha}^{\prime}(t) for all αΦ\alpha\in\Phi and tFt\in F.

The Lie algebra of G(V,L)G(V,L) is identified as follows. The stabilizer of LL in 𝔤\mathfrak{g} and 𝔥\mathfrak{h} is given by [32, Corollary 2, p. 16]

𝔥L\displaystyle\mathfrak{h}^{L} ={H𝔥:μ(H) for all μΛ(V)},\displaystyle=\{H\in\mathfrak{h}:\mu(H)\in\mathbb{Z}\text{ for all }\mu\in\Lambda(V)\},
𝔤L\displaystyle\mathfrak{g}^{L} =𝔥LαΦXα.\displaystyle=\mathfrak{h}^{L}\oplus\bigoplus_{\alpha\in\Phi}\mathbb{Z}X_{\alpha}.

In particular 𝔤L\mathfrak{g}^{L} and 𝔥L\mathfrak{h}^{L} only depend on the weights in VV, not the choice of the admissible lattice LL. Furthermore, under the adjoint action 𝔤L\mathfrak{g}^{L} is an admissible lattice in 𝔤\mathfrak{g} [11, Proposition 27.2].

Then 𝔤L\mathfrak{g}^{L} is a \mathbb{Z}-Lie algebra which acts on LL. The Chevalley algebra

𝔤(V,L):=F𝔤L\mathfrak{g}(V,L):=F\otimes_{\mathbb{Z}}\mathfrak{g}^{L}

is a Lie algebra over FF which acts faithfully on VFV_{F}, and 𝔤(V,L)𝔤𝔩(VF)\mathfrak{g}(V,L)\subseteq\mathfrak{gl}(V_{F}) is precisely the Lie algebra of G(V,L)G(V,L). The adjoint action of G(V,L)G(V,L) can be realized by applying the Chevalley construction with V=𝔤V^{\prime}=\mathfrak{g} and L=𝔤LL^{\prime}=\mathfrak{g}^{L}, and then taking the morphism G(V,L)G(V,L)G(V,L)\rightarrow G(V^{\prime},L^{\prime}) of Lemma 7.1 (ii).

Note that then the Lie algebra 𝔤(V,L)\mathfrak{g}(V,L) is a GscG_{sc}-module for a simply connected Chevalley group GscG_{sc} with root system Φ\Phi (Lemma 7.1 (i)). In all cases, the structure of 𝔤(V,L)\mathfrak{g}(V,L) as a Lie algebra and as a GscG_{sc}-module is described by Hogeweij in [9]. Here the composition factors of 𝔤(V,L)\mathfrak{g}(V,L) are completely determined by VV, but the submodule structure depends on the choice of the admissible lattice LL.

In the simply connected case we have 𝔤L=𝔤\mathfrak{g}^{L}=\mathfrak{g}_{\mathbb{Z}}. In the adjoint case we denote 𝔤ad:=𝔤L\mathfrak{g}_{\mathbb{Z}}^{ad}:=\mathfrak{g}^{L}, so 𝔤ad\mathfrak{g}_{\mathbb{Z}}^{ad} is the \mathbb{Z}-span of 𝔤\mathfrak{g}_{\mathbb{Z}} and 𝔥ad:={H𝔥:α(H) for all αΦ}\mathfrak{h}^{ad}:=\{H\in\mathfrak{h}:\alpha(H)\in\mathbb{Z}\text{ for all }\alpha\in\Phi\}.

7.2. Chevalley construction for type CC_{\ell}

We will now setup the Chevalley construction for groups of type CC_{\ell} over KK. (Recall that by KK we always denote an algebraically closed field of characteristic two.)

Let VV_{\mathbb{C}} be a \mathbb{C}-vector space of dimension n=2n=2\ell, with basis v1v_{1}, \ldots, vnv_{n}. We define a non-degenerate alternating bilinear form (,)(-,-) on VV_{\mathbb{C}} by (vi,vni+1)=1=(vni+1,vi)(v_{i},v_{n-i+1})=1=-(v_{n-i+1},v_{i}) for all 1i1\leq i\leq\ell, and (vi,vj)=0(v_{i},v_{j})=0 if i+jn+1i+j\neq n+1.

Let 𝔰𝔭(V)={X𝔤𝔩(V):(Xv,w)+(v,Xw)=0 for all v,wV}\mathfrak{sp}(V_{\mathbb{C}})=\{X\in\mathfrak{gl}(V_{\mathbb{C}}):(Xv,w)+(v,Xw)=0\text{ for all }v,w\in V_{\mathbb{C}}\}, so 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}) is a simple Lie algebra of type CC_{\ell}. Let 𝔥\mathfrak{h}_{\mathbb{C}} be the Cartan subalgebra formed by the diagonal matrices in 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}). Then 𝔥={diag(h1,,h,h,,h1):hi}\mathfrak{h}_{\mathbb{C}}=\{\operatorname{diag}(h_{1},\ldots,h_{\ell},-h_{\ell},\ldots,-h_{1}):h_{i}\in\mathbb{C}\}.

For 1i1\leq i\leq\ell, define linear maps εi:𝔥\varepsilon_{i}:\mathfrak{h}_{\mathbb{C}}\rightarrow\mathbb{C} by εi(h)=hi\varepsilon_{i}(h)=h_{i} where hh is a diagonal matrix with diagonal entries (h1,,h,h,,h1)(h_{1},\ldots,h_{\ell},-h_{\ell},\ldots,-h_{1}). Then

Φ={±(εi±εj):1i<j}{±2εi:1i}\Phi=\{\pm(\varepsilon_{i}\pm\varepsilon_{j}):1\leq i<j\leq\ell\}\cup\{\pm 2\varepsilon_{i}:1\leq i\leq\ell\}

is the root system for 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}), and

Φ+={εi±εj:1i<j}{2εi:1i}\Phi^{+}=\{\varepsilon_{i}\pm\varepsilon_{j}:1\leq i<j\leq\ell\}\cup\{2\varepsilon_{i}:1\leq i\leq\ell\}

is a system of positive roots. The set of simple roots corresponding to Φ+\Phi^{+} is Δ={α1,,α}\Delta=\{\alpha_{1},\ldots,\alpha_{\ell}\}, where αi=εiεi+1\alpha_{i}=\varepsilon_{i}-\varepsilon_{i+1} for 1i<1\leq i<{\ell} and α=2ε\alpha_{\ell}=2\varepsilon_{\ell}.

For all i,ji,j let Ei,jE_{i,j} be the linear endomorphism on VV_{\mathbb{C}} such that Ei,j(vj)=viE_{i,j}(v_{j})=v_{i} and Ei,j(vk)=0E_{i,j}(v_{k})=0 for kjk\neq j. Throughout we will use the following Chevalley basis of 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}), which is taken from [13, Section 11, p. 38].

Xεiεj\displaystyle X_{\varepsilon_{i}-\varepsilon_{j}} =Ei,jEnj+1,ni+1\displaystyle=E_{i,j}-E_{n-j+1,n-i+1} for all ij,\displaystyle\text{ for all }i\neq j,
Xεi+εj\displaystyle X_{\varepsilon_{i}+\varepsilon_{j}} =Ej,ni+1+Ei,nj+1\displaystyle=E_{j,n-i+1}+E_{i,n-j+1} for all ij,\displaystyle\text{ for all }i\neq j,
X(εi+εj)\displaystyle X_{-(\varepsilon_{i}+\varepsilon_{j})} =Enj+1,i+Eni+1,j\displaystyle=E_{n-j+1,i}+E_{n-i+1,j} for all ij,\displaystyle\text{ for all }i\neq j,
X2εi\displaystyle X_{2\varepsilon_{i}} =Ei,ni+1\displaystyle=E_{i,n-i+1} for all i,\displaystyle\text{ for all }i,
X2εi\displaystyle X_{-2\varepsilon_{i}} =Eni+1,i\displaystyle=E_{n-i+1,i} for all i.\displaystyle\text{ for all }i.
Hαi\displaystyle H_{\alpha_{i}} =[Xαi,Xαi]\displaystyle=[X_{\alpha_{i}},X_{-\alpha_{i}}] for all 1i.\displaystyle\text{ for all }1\leq i\leq\ell.

Let Λ𝔥\Lambda\subset\mathfrak{h}_{\mathbb{C}}^{*} be the weight lattice. The maps ε1\varepsilon_{1}, \ldots, ε\varepsilon_{\ell} form \mathbb{Z}-basis for Λ\Lambda, and for 1i1\leq i\leq\ell, the iith fundamental highest weight is equal to ϖi:=ε1++εi\varpi_{i}:=\varepsilon_{1}+\cdots+\varepsilon_{i}.

As in the previous subsection, we denote by 𝔤\mathfrak{g}_{\mathbb{Z}} be the \mathbb{Z}-span of the Chevalley basis above. Then 𝔤ad\mathfrak{g}_{\mathbb{Z}}^{ad} is the \mathbb{Z}-span of 𝔤\mathfrak{g}_{\mathbb{Z}} and HH, where H𝔥H\in\mathfrak{h}_{\mathbb{C}} is the diagonal matrix H=diag(1/2,,1/2,1/2,,1/2)H=\operatorname{diag}(1/2,\ldots,1/2,-1/2,\ldots,-1/2). In terms of the Chevalley basis, we have

H=12(Hα1+2Hα2++Hα).H=\frac{1}{2}\left(H_{\alpha_{1}}+2H_{\alpha_{2}}+\cdots+\ell H_{\alpha_{\ell}}\right).

7.3. Simply connected groups of type CC_{\ell}

Let VV_{\mathbb{Z}} be the \mathbb{Z}-span of the basis v1v_{1}, \ldots, vnv_{n} of VV_{\mathbb{C}}. It is clear that VV_{\mathbb{Z}} is an admissible lattice. Denote V:=KVV:=K\otimes_{\mathbb{Z}}V_{\mathbb{Z}}. By abuse of notation we denote vi:=1viv_{i}:=1\otimes v_{i} for all 1in1\leq i\leq n, so v1v_{1}, \ldots, vnv_{n} is a basis of VV. The alternating bilinear form (,)(-,-) induces an alternating bilinear form b:V×VKb:V\times V\rightarrow K on VV, with b(vi,vj)=1b(v_{i},v_{j})=1 if i+j=n+1i+j=n+1 and b(vi,vj)=0b(v_{i},v_{j})=0 otherwise.

Then the Chevalley group corresponding to VV_{\mathbb{C}} and VV_{\mathbb{Z}} is simply connected of type CC_{\ell}, and it is precisely the symplectic group Sp(V)\operatorname{Sp}(V) corresponding to bb [29, 5]. We denote Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V). Then the Lie algebra of GscG_{sc} is

𝔤sc=𝔰𝔭(V)={X𝔤𝔩(V):b(Xv,w)+b(v,Xw)=0 for all v,wV}.\mathfrak{g}_{sc}=\mathfrak{sp}(V)=\{X\in\mathfrak{gl}(V):b(Xv,w)+b(v,Xw)=0\text{ for all }v,w\in V\}.

We denote by 𝔤ad:=K𝔤ad\mathfrak{g}_{ad}:=K\otimes_{\mathbb{Z}}\mathfrak{g}_{\mathbb{Z}}^{\operatorname{ad}} the adjoint Lie algebra of type CC_{\ell}. For any 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}})-module WW with admissible lattice LL, we will consider KLK\otimes_{\mathbb{Z}}L as a GscG_{sc}-module via the action provided by Lemma 7.1 (i).

7.4. Lie algebra of adjoint type CC_{\ell}

To compute with the action of Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) on 𝔤ad\mathfrak{g}_{ad}, it will be convenient to use the following well-known identification of 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}) with the symmetric square S2(V)S^{2}(V_{\mathbb{C}}).

Lemma 7.2.

For x,yVx,y\in V_{\mathbb{C}}, define a linear map ψx,y:VV\psi_{x,y}:V_{\mathbb{C}}\rightarrow V_{\mathbb{C}} by v(y,v)x+(x,v)yv\mapsto(y,v)x+(x,v)y. Then we have an isomorphism S2(V)𝔰𝔭(V)S^{2}(V_{\mathbb{C}})\rightarrow\mathfrak{sp}(V_{\mathbb{C}}) of 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}})-modules, defined by xyψx,yxy\mapsto\psi_{x,y} for all x,yVx,y\in V_{\mathbb{C}}.

Proof.

We have an isomorphism τ:VV𝔤𝔩(V)\tau:V_{\mathbb{C}}\otimes V_{\mathbb{C}}^{*}\rightarrow\mathfrak{gl}(V_{\mathbb{C}}) of 𝔤𝔩(V)\mathfrak{gl}(V_{\mathbb{C}})-modules, where τ(vf)\tau(v\otimes f) is the linear map wf(w)vw\mapsto f(w)v for all vVv\in V_{\mathbb{C}} and fVf\in V_{\mathbb{C}}^{*}. Moreover, we have an isomorphism τ:VVVV\tau^{\prime}:V_{\mathbb{C}}\otimes V_{\mathbb{C}}\rightarrow V_{\mathbb{C}}\otimes V_{\mathbb{C}}^{*} of 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}})-modules defined by xyxfyx\otimes y\mapsto x\otimes f_{y}, where fy(v)=(y,v)f_{y}(v)=(y,v) for all vVv\in V.

Now identifying S2(V)VVS^{2}(V_{\mathbb{C}})\subseteq V_{\mathbb{C}}\otimes V_{\mathbb{C}} via xyxy+yxxy\mapsto x\otimes y+y\otimes x, the restriction of ττ\tau\tau^{\prime} to S2(V)S^{2}(V_{\mathbb{C}}) is a map ψ:S2(V)𝔤𝔩(V)\psi:S^{2}(V_{\mathbb{C}})\rightarrow\mathfrak{gl}(V_{\mathbb{C}}) defined by xyψx,yxy\mapsto\psi_{x,y}. Then ψ\psi is an injective morphism of 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}})-modules, and it is clear that ψx,y𝔰𝔭(V)\psi_{x,y}\in\mathfrak{sp}(V_{\mathbb{C}}) for all x,yVx,y\in V_{\mathbb{C}}. Since S2(V)S^{2}(V_{\mathbb{C}}) and 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}) have the same dimension, we conclude that ψ\psi defines an isomorphism S2(V)𝔰𝔭(V)S^{2}(V_{\mathbb{C}})\rightarrow\mathfrak{sp}(V_{\mathbb{C}}) of 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}})-modules.∎

We denote by ψ:S2(V)𝔰𝔭(V)\psi:S^{2}(V_{\mathbb{C}})\rightarrow\mathfrak{sp}(V_{\mathbb{C}}) the isomorphism ψ(xy)=ψx,y\psi(xy)=\psi_{x,y} as in Lemma 7.2. Let 1i,j1\leq i,j\leq\ell with iji\neq j. Then under the map ψ\psi, the root vectors X±εi±εjX_{\pm\varepsilon_{i}\pm\varepsilon_{j}} in the Chevalley basis correspond to elements of S2(V)S^{2}(V_{\mathbb{C}}) as follows.

ψ(vivnj+1)\displaystyle\psi\left(v_{i}v_{n-j+1}\right) =Xεiεj\displaystyle=-X_{\varepsilon_{i}-\varepsilon_{j}}
ψ(vivj)\displaystyle\psi\left(v_{i}v_{j}\right) =Xεi+εj\displaystyle=X_{\varepsilon_{i}+\varepsilon_{j}}
ψ(vni+1vnj+1)\displaystyle\psi\left(v_{n-i+1}v_{n-j+1}\right) =X(εi+εj)\displaystyle=-X_{-(\varepsilon_{i}+\varepsilon_{j})}
ψ(12vi2)\displaystyle\psi\left(\frac{1}{2}v_{i}^{2}\right) =X2εi\displaystyle=X_{2\varepsilon_{i}}
ψ(12vni+12)\displaystyle\psi\left(\frac{1}{2}v_{n-i+1}^{2}\right) =X2εi\displaystyle=-X_{-2\varepsilon_{i}}

Furthermore, define

δ:=12i=1vivni+1.\delta:=\frac{1}{2}\sum_{i=1}^{\ell}v_{i}v_{n-i+1}.

Then ψ(δ)=H\psi(\delta)=-H.

We define

Lsc\displaystyle L_{sc} :=-span of vivj and 12vi2 for all 1i,jn;\displaystyle:=\mathbb{Z}\text{-span of }v_{i}v_{j}\text{ and }\frac{1}{2}v_{i}^{2}\text{ for all }1\leq i,j\leq n;
Lad\displaystyle L_{ad} :=-span of Lsc and δ;\displaystyle:=\mathbb{Z}\text{-span of }L_{sc}\text{ and }\delta;

so that ψ(Lsc)=𝔤\psi(L_{sc})=\mathfrak{g}_{\mathbb{Z}} and ψ(Lad)=𝔤ad\psi(L_{ad})=\mathfrak{g}_{\mathbb{Z}}^{ad}.

Then LscLadL_{sc}\subset L_{ad} are admissible lattices in S2(V)S^{2}(V_{\mathbb{C}}), since 𝔤\mathfrak{g}_{\mathbb{Z}} and 𝔤ad\mathfrak{g}_{\mathbb{Z}}^{ad} are admissible lattices in 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}).

From the Chevalley construction K𝔤=𝔤scK\otimes_{\mathbb{Z}}\mathfrak{g}_{\mathbb{Z}}=\mathfrak{g}_{sc} and K𝔤ad=𝔤adK\otimes_{\mathbb{Z}}\mathfrak{g}_{\mathbb{Z}}^{ad}=\mathfrak{g}_{ad} as GscG_{sc}-modules, so ψ\psi induces isomorphisms

(7.1) ψ:𝔤scKLsc\displaystyle\psi^{\prime}:\mathfrak{g}_{sc}\rightarrow K\otimes_{\mathbb{Z}}L_{sc}
ψ′′:𝔤adKLad\displaystyle\psi^{\prime\prime}:\mathfrak{g}_{ad}\rightarrow K\otimes_{\mathbb{Z}}L_{ad}

of GscG_{sc}-modules.

By [9, Lemma 2.2] we have dimZ(𝔤sc)=1\dim Z(\mathfrak{g}_{sc})=1 and Z(𝔤sc)Z(\mathfrak{g}_{sc}) is spanned by 12H1\otimes 2H. Similarly by [9, Table 1], we have dim𝔤ad/[𝔤ad,𝔤ad]=1\dim\mathfrak{g}_{ad}/[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]=1 and [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}] is spanned by 1v1\otimes v with v𝔤v\in\mathfrak{g}_{\mathbb{Z}}. Thus under the isomorphisms in (7.1), we get

(7.2) ψ(Z(𝔤sc))\displaystyle\psi^{\prime}(Z(\mathfrak{g}_{sc})) =12δ,\displaystyle=\langle 1\otimes 2\delta\rangle,
ψ′′([𝔤ad,𝔤ad])\displaystyle\psi^{\prime\prime}([\mathfrak{g}_{ad},\mathfrak{g}_{ad}]) =1v:vLsc.\displaystyle=\langle 1\otimes v:v\in L_{sc}\rangle.

Note that

(7.3) Lad/Lsc={Lsc,δ+Lsc}/2,L_{ad}/L_{sc}=\{L_{sc},\delta+L_{sc}\}\cong\mathbb{Z}/2\mathbb{Z},

so for the linear map π:KLscKLad\pi:K\otimes_{\mathbb{Z}}L_{sc}\rightarrow K\otimes_{\mathbb{Z}}L_{ad} induced by the inclusion LscLadL_{sc}\subset L_{ad}, we have Kerπ=1(2δ)\operatorname{Ker}\pi=\langle 1\otimes(2\delta)\rangle and Imπ=1v:vLsc\operatorname{Im}\pi=\langle 1\otimes v:v\in L_{sc}\rangle.

7.5. Action of unipotent elements on 𝔤ad\mathfrak{g}_{ad}, orthogonally indecomposable case

Let Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) as in the previous sections, and let uSp(V)u\in\operatorname{Sp}(V) be a unipotent element.

We will describe how to construct a representative for the conjugacy class of uu in terms of root elements, and how to compute the action of uu on 𝔤adKLad\mathfrak{g}_{ad}\cong K\otimes_{\mathbb{Z}}L_{ad}. We first do this in the case where VK[u]V\downarrow K[u] is orthogonally indecomposable, so VK[u]=V(2)V\downarrow K[u]=V(2\ell) or VK[u]=W()V\downarrow K[u]=W(\ell) (Theorem 5.3). In this case it is well known that by replacing uu with a conjugate, we can take

u={xα1(1)xα(1), if VK[u]=V(2).xα1(1)xα1(1), if VK[u]=W().u=\begin{cases}x_{\alpha_{1}}(1)\cdots x_{\alpha_{\ell}}(1),&\text{ if }V\downarrow K[u]=V(2\ell).\\ x_{\alpha_{1}}(1)\cdots x_{\alpha_{\ell-1}}(1),&\text{ if }V\downarrow K[u]=W(\ell).\end{cases}

For αΦ\alpha\in\Phi, let

x(α):=1+Xα+Xα22!𝒰.x_{\mathbb{Z}}^{(\alpha)}:=1+X_{\alpha}+\frac{X_{\alpha}^{2}}{2!}\in\mathscr{U}_{\mathbb{Z}}.

Then xα(1)x_{\alpha}(1) is the reduction modulo pp of the action of x(α)x_{\mathbb{Z}}^{(\alpha)} on VV_{\mathbb{Z}}. Furthermore, since Xα3S2(V)=0X_{\alpha}^{3}\cdot S^{2}(V_{\mathbb{C}})=0, the action of xα(1)x_{\alpha}(1) on 𝔤adKLad\mathfrak{g}_{ad}\cong K\otimes_{\mathbb{Z}}L_{ad} is the reduction modulo pp of the action of x(α)x_{\mathbb{Z}}^{(\alpha)} on LadL_{ad}.

Therefore we define

u:={x(α1)x(α), if VK[u]=V(2).x(α1)x(α1), if VK[u]=W().u_{\mathbb{Z}}:=\begin{cases}x_{\mathbb{Z}}^{(\alpha_{1})}\cdots x_{\mathbb{Z}}^{(\alpha_{\ell})},&\text{ if }V\downarrow K[u]=V(2\ell).\\ x_{\mathbb{Z}}^{(\alpha_{1})}\cdots x_{\mathbb{Z}}^{(\alpha_{\ell-1})},&\text{ if }V\downarrow K[u]=W(\ell).\end{cases}

If VK[u]=V(2)V\downarrow K[u]=V(2\ell), we get

uv1\displaystyle u_{\mathbb{Z}}\cdot v_{1} =v1,\displaystyle=v_{1},
uvi\displaystyle u_{\mathbb{Z}}\cdot v_{i} =vi+vi1++v1 if 1<i+1,\displaystyle=v_{i}+v_{i-1}+\cdots+v_{1}\text{ if }1<i\leq\ell+1,
uvi\displaystyle u_{\mathbb{Z}}\cdot v_{i} =vivi1 if +1<i2.\displaystyle=v_{i}-v_{i-1}\text{ if }\ell+1<i\leq 2\ell.

Similarly for VK[u]=W()V\downarrow K[u]=W(\ell), we get

uv1\displaystyle u_{\mathbb{Z}}\cdot v_{1} =v1,\displaystyle=v_{1},
uvi\displaystyle u_{\mathbb{Z}}\cdot v_{i} =vi+vi1++v1 if 1<i,\displaystyle=v_{i}+v_{i-1}+\cdots+v_{1}\text{ if }1<i\leq\ell,
uv+1\displaystyle u_{\mathbb{Z}}\cdot v_{\ell+1} =v+1\displaystyle=v_{\ell+1}
uvi\displaystyle u_{\mathbb{Z}}\cdot v_{i} =vivi1 if +1<i2.\displaystyle=v_{i}-v_{i-1}\text{ if }\ell+1<i\leq 2\ell.

Thus the action of uu on VV is exactly as described in Definition 5.1 or Definition 5.2.

Since Xαi3S2(V)=0X_{\alpha_{i}}^{3}\cdot S^{2}(V_{\mathbb{C}})=0, for the action on S2(V)S^{2}(V_{\mathbb{C}}) we have

x(α)(vw)=(x(α)v)(x(α)w)x_{\mathbb{Z}}^{(\alpha)}\cdot(vw)=(x_{\mathbb{Z}}^{(\alpha)}v)(x_{\mathbb{Z}}^{(\alpha)}w)

for all v,wVv,w\in V_{\mathbb{C}}. Thus

u(vw)=(uv)(uw)u_{\mathbb{Z}}\cdot(vw)=(u_{\mathbb{Z}}v)(u_{\mathbb{Z}}w)

for all v,wVv,w\in V_{\mathbb{C}}.

7.6. Action of unipotent elements on 𝔤ad\mathfrak{g}_{ad}, general case

For all unipotent elements uGscu\in G_{sc}, in general we can proceed as follows. We have an orthogonal decomposition VK[u]=U1UtV\downarrow K[u]=U_{1}\perp\cdots\perp U_{t}, where dimUi>0\dim U_{i}>0 and UiU_{i} is an orthogonally indecomposable K[u]K[u]-module for all 1it1\leq i\leq t. Write dimUi=ni=2i\dim U_{i}=n_{i}=2\ell_{i} for all 1it1\leq i\leq t.

Then u=u1utu=u_{1}\cdots u_{t}, where uiSp(Ui)u_{i}\in\operatorname{Sp}(U_{i}) is the action of uu on UiU_{i}. For all 1it1\leq i\leq t, we have UiK[ui]W(i)U_{i}\downarrow K[u_{i}]\cong W(\ell_{i}) or UiK[ui]V(2i)U_{i}\downarrow K[u_{i}]\cong V(2\ell_{i}) (Theorem 5.3).

Relabel the basis v1v_{1}, \ldots, vnv_{n} of VV_{\mathbb{C}} as

v1(1),,vn1(1),,v1(t),,vnt(t),v_{1}^{(1)},\ldots,v_{n_{1}}^{(1)},\ldots,v_{1}^{(t)},\ldots,v_{n_{t}}^{(t)},

where (vi(j),vi(j))=1(v_{i}^{(j)},v_{i^{\prime}}^{(j^{\prime})})=1 if j=jj=j^{\prime} and i+i=nji+i^{\prime}=n_{j}, and 0 otherwise. Then the Cartan subalgebra 𝔥\mathfrak{h} consists of diagonal matrices of the form

h=diag(a1(1),,a1(1),a1(1),,a1(1),,a1(t),,at(t),at(t),,a1(t)).h=\operatorname{diag}(a_{1}^{(1)},\ldots,a_{\ell_{1}}^{(1)},-a_{\ell_{1}}^{(1)},\ldots,-a_{1}^{(1)},\ \ldots,\ a_{1}^{(t)},\ldots,a_{\ell_{t}}^{(t)},-a_{\ell_{t}}^{(t)},\ldots,-a_{1}^{(t)}).

We define then for 1it1\leq i\leq t and 1ki1\leq k\leq\ell_{i} the linear map εk(i):𝔥\varepsilon_{k}^{(i)}:\mathfrak{h}\rightarrow\mathbb{C} by hak(i)h\mapsto a_{k}^{(i)}.

Denote by V(i)V_{\mathbb{Z}}^{(i)} (resp. V(i)V_{\mathbb{C}}^{(i)}) the \mathbb{Z}-span (resp. \mathbb{C}-span) of v1(i),,vni(i)v_{1}^{(i)},\ldots,v_{n_{i}}^{(i)}, so we have

V\displaystyle V_{\mathbb{Z}} =V(1)V(t),\displaystyle=V_{\mathbb{Z}}^{(1)}\oplus\cdots\oplus V_{\mathbb{Z}}^{(t)},
V\displaystyle V_{\mathbb{C}} =V(1)V(t).\displaystyle=V_{\mathbb{C}}^{(1)}\perp\cdots\perp V_{\mathbb{C}}^{(t)}.

Since V=KVV=K\otimes_{\mathbb{Z}}V_{\mathbb{Z}}, we can assume that Ui=KV(i)U_{i}=K\otimes_{\mathbb{Z}}V_{\mathbb{Z}}^{(i)} for all 1it1\leq i\leq t.

We have 𝔰𝔭(V(i))𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}^{(i)})\subseteq\mathfrak{sp}(V_{\mathbb{C}}), and 𝔰𝔭(V(i))\mathfrak{sp}(V_{\mathbb{C}}^{(i)}) has a root system Φ(i)\Phi^{(i)} of type CiC_{\ell_{i}} with simple roots Δ(i)={β1,,βi}\Delta^{(i)}=\{\beta_{1},\ldots,\beta_{\ell_{i}}\}, where

βj={εj(i)εj+1(i), if 1j<i,2εi(i), if j=i.\beta_{j}=\begin{cases}\varepsilon_{j}^{(i)}-\varepsilon_{j+1}^{(i)},&\text{ if }1\leq j<\ell_{i},\\ 2\varepsilon_{\ell_{i}}^{(i)},&\text{ if }j=\ell_{i}.\end{cases}

Note that {Xα:αΦ(i)}{Hα:αΔ(i)}\{X_{\alpha}:\alpha\in\Phi^{(i)}\}\cup\{H_{\alpha}:\alpha\in\Delta^{(i)}\} is a Chevalley basis of 𝔰𝔭(V(i))\mathfrak{sp}(V_{\mathbb{C}}^{(i)}), and the corresponding Kostant \mathbb{Z}-form 𝒰(i)\mathscr{U}_{\mathbb{Z}}^{(i)} is a subring of 𝒰\mathscr{U}_{\mathbb{Z}}. Then V(i)V_{\mathbb{Z}}^{(i)} is an admissible lattice for 𝔰𝔭(V(i))\mathfrak{sp}(V_{\mathbb{C}}^{(i)}), and applying the Chevalley construction we get Sp(Ui)\operatorname{Sp}(U_{i}).

As in the orthogonally indecomposable case, we define

u,i:={x(β1)x(βi), if UiK[ui]=V(2i).x(β1)x(βi1), if UiK[ui]=W(i).u_{\mathbb{Z},i}:=\begin{cases}x_{\mathbb{Z}}^{(\beta_{1})}\cdots x_{\mathbb{Z}}^{(\beta_{\ell_{i}})},&\text{ if }U_{i}\downarrow K[u_{i}]=V(2\ell_{i}).\\ x_{\mathbb{Z}}^{(\beta_{1})}\cdots x_{\mathbb{Z}}^{(\beta_{\ell_{i}-1})},&\text{ if }U_{i}\downarrow K[u_{i}]=W(\ell_{i}).\end{cases}

Then u,i𝒰(i)u_{\mathbb{Z},i}\in\mathscr{U}_{\mathbb{Z}}^{(i)}, and the reduction modulo pp of the action of u,iu_{\mathbb{Z},i} on V(i)V_{\mathbb{Z}}^{(i)} is precisely uiSp(Ui)u_{i}\in\operatorname{Sp}(U_{i}).

Thus we can define

u:=u,1u,t𝒰,u_{\mathbb{Z}}:=u_{\mathbb{Z},1}\cdots u_{\mathbb{Z},t}\in\mathscr{U}_{\mathbb{Z}},

so that uu is the reduction modulo pp of the action of uu_{\mathbb{Z}} on VV_{\mathbb{Z}}. Furthermore, the action of uu on 𝔤ad\mathfrak{g}_{ad} is the reduction modulo pp of the action of uu_{\mathbb{Z}} on LadL_{ad}.

7.7. Action of nilpotent elements on 𝔤ad\mathfrak{g}_{ad}

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent. Similarly to the unipotent case, we can find e𝔤e_{\mathbb{Z}}\in\mathfrak{g}_{\mathbb{Z}} such that ee is the reduction modulo pp of the action of ee_{\mathbb{Z}} on VV_{\mathbb{Z}}.

We first consider the orthogonally indecomposable case, so VK[e]=V(2)V\downarrow K[e]=V(2\ell), VK[e]=W()V\downarrow K[e]=W(\ell), or VK[e]=Wk()V\downarrow K[e]=W_{k}(\ell) for some 0<k</20<k<\ell/2 (Theorem 6.7). In this case we define

e:={Xα1++Xα, if VK[e]=V(2).Xα1++Xα1, if VK[e]=W().Xα1++Xα1+X2εk, if VK[e]=Wk().e_{\mathbb{Z}}:=\begin{cases}X_{\alpha_{1}}+\cdots+X_{\alpha_{\ell}},&\text{ if }V\downarrow K[e]=V(2\ell).\\ X_{\alpha_{1}}+\cdots+X_{\alpha_{\ell-1}},&\text{ if }V\downarrow K[e]=W(\ell).\\ X_{\alpha_{1}}+\cdots+X_{\alpha_{\ell-1}}+X_{2\varepsilon_{k}},&\text{ if }V\downarrow K[e]=W_{k}(\ell).\end{cases}

Then for the reduction modulo pp of ee_{\mathbb{Z}}, the action of V=KVV=K\otimes_{\mathbb{Z}}V_{\mathbb{Z}} is exactly as in Definition 6.46.6. (Here the expression for Wk()W_{k}(\ell) is taken from [20, Lemma 3.4].)

In the general case, we have an orthogonal decomposition VK[e]=U1UtV\downarrow K[e]=U_{1}\perp\cdots\perp U_{t}, where UiU_{i} is an orthogonally indecomposable K[e]K[e]-module, with dimUi=2i\dim U_{i}=2\ell_{i}. We have e=e1++ete=e_{1}+\cdots+e_{t}, where ei𝔰𝔭(Ui)e_{i}\in\mathfrak{sp}(U_{i}) is the action of ee on UiU_{i}. In the notation of the previous section, we take Ui=KV(i)U_{i}=K\otimes_{\mathbb{Z}}V_{\mathbb{Z}}^{(i)}. We can then define

e:=e,1++e,t,e_{\mathbb{Z}}:=e_{\mathbb{Z},1}+\cdots+e_{\mathbb{Z},t},

where e,i𝔤𝔰𝔭(V(i))e_{\mathbb{Z},i}\in\mathfrak{g}_{\mathbb{Z}}\cap\mathfrak{sp}(V_{\mathbb{C}}^{(i)}) is defined as in the orthogonally indecomposable case, such that eie_{i} is the reduction modulo pp of e,ie_{\mathbb{Z},i}.

Then ee is the reduction modulo pp of ee_{\mathbb{Z}}. Moreover, given any irreducible 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}})-module WW_{\mathbb{C}} and admissible lattice LWL\subset W_{\mathbb{C}}, the action of ee on KLK\otimes_{\mathbb{Z}}L is precisely the reduction modulo pp of the action of ee_{\mathbb{Z}} on LL.

7.8. Elements of 𝒰\mathscr{U}_{\mathbb{Z}} acting nilpotently

In this section, we consider X𝒰X\in\mathscr{U}_{\mathbb{Z}} which act nilpotently on S2(V)S^{2}(V_{\mathbb{C}}). Then by reduction modulo pp, we get an action X^\widehat{X} on 𝔤sc\mathfrak{g}_{sc}, and X~\widetilde{X} on 𝔤ad\mathfrak{g}_{ad}. We will describe when Im(X^)Z(𝔤sc)\operatorname{Im}(\widehat{X})\supseteq Z(\mathfrak{g}_{sc}) and when Ker(X~)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{X})\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]. For our purposes this is mostly relevant for X=(u1)mX=(u_{\mathbb{Z}}-1)^{m} in the unipotent case, and X=emX=e_{\mathbb{Z}}^{m} in the nilpotent case, for integer m1m\geq 1.

Lemma 7.3.

Let X𝒰X\in\mathscr{U}_{\mathbb{Z}} be such that XX acts nilpotently on S2(V)S^{2}(V_{\mathbb{C}}), and denote the action of XX on 𝔤sc\mathfrak{g}_{sc} by X^\widehat{X}. Then Im(X^)Z(𝔤sc)\operatorname{Im}(\widehat{X})\supseteq Z(\mathfrak{g}_{sc}) if and only if there exists vLscv\in L_{sc} such that Xv2δmod2LscX\cdot v\equiv 2\delta\mod 2L_{sc}.

Proof.

Identifying 𝔤sc=KLsc\mathfrak{g}_{sc}=K\otimes_{\mathbb{Z}}L_{sc} via (7.1), we have Z(𝔤sc)=12δZ(\mathfrak{g}_{sc})=\langle 1\otimes 2\delta\rangle by (7.2). Thus if vLscv\in L_{sc} is such that Xv2δmod2LscX\cdot v\equiv 2\delta\mod 2L_{sc}, we have X^(1v)=12δ\widehat{X}\cdot(1\otimes v)=1\otimes 2\delta.

Conversely, suppose that Im(X^)Z(𝔤sc)\operatorname{Im}(\widehat{X})\supseteq Z(\mathfrak{g}_{sc}). Then there exists w𝔤scw\in\mathfrak{g}_{sc} with X^w=12δ\widehat{X}\cdot w=1\otimes 2\delta. Since X^=1X\widehat{X}=1\otimes X^{\prime} where XX^{\prime} is the action of XX on LscL_{sc}, we can assume that w=1vw=1\otimes v with vLscv\in L_{sc}. Then

12δ=X^w=1(Xv)1\otimes 2\delta=\widehat{X}\cdot w=1\otimes(X\cdot v)

implies that Xv2δmod2LscX\cdot v\equiv 2\delta\mod 2L_{sc}.∎

For actions of unipotent elements, Lemma 7.3 gives the following.

Lemma 7.4.

Let uGscu\in G_{sc} be unipotent, and suppose that uu is the reduction modulo pp of u𝒰u_{\mathbb{Z}}\in\mathscr{U}_{\mathbb{Z}}. Denote the action of uu on 𝔤sc\mathfrak{g}_{sc} by u^\widehat{u}. Let m1m\geq 1 be an integer. Then Im(u^1)mZ(𝔤sc)\operatorname{Im}(\widehat{u}-1)^{m}\supseteq Z(\mathfrak{g}_{sc}) if and only if there exists vLscv\in L_{sc} such that (u1)m(v)2δmod2Lsc(u_{\mathbb{Z}}-1)^{m}\cdot(v)\equiv 2\delta\mod 2L_{sc}.

Proof.

Follows from Lemma 7.3, with X=(u1)mX=(u_{\mathbb{Z}}-1)^{m}. ∎

Lemma 7.5.

Let X𝒰X\in\mathscr{U}_{\mathbb{Z}} be such that XX acts nilpotently on S2(V)S^{2}(V_{\mathbb{C}}), and denote the action of XX on 𝔤ad\mathfrak{g}_{ad} by X~\widetilde{X}. Then the following statements are equivalent:

  1. (i)

    Ker(X~)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{X})\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] .

  2. (ii)

    There exists vLscv\in L_{sc} such that X(δ+v)2LadX\cdot(\delta+v)\in 2L_{ad}.

  3. (iii)

    There exists vLscv\in L_{sc} such that one of the following holds:

    1. (a)

      X(δ+v)0mod2LscX\cdot(\delta+v)\equiv 0\mod{2L_{sc}}.

    2. (b)

      X(δ+v)2δmod2LscX\cdot(\delta+v)\equiv 2\delta\mod{2L_{sc}}.

Proof.

We first prove that (i) and (ii) are equivalent. We identify 𝔤ad=KLad\mathfrak{g}_{ad}=K\otimes_{\mathbb{Z}}L_{ad} via (7.1), in which case 𝔤ad=1δ,[𝔤ad,𝔤ad]\mathfrak{g}_{ad}=\langle 1\otimes\delta,[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]\rangle as KK-vector spaces. Here [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}] is the subspace spanned by 1v1\otimes v with vLscv\in L_{sc}, as seen in (7.2).

If there exists vLscv\in L_{sc} such that X(δ+v)2LadX\cdot(\delta+v)\in 2L_{ad}, we have

X~(1δ+1v)=1(X(δ+v))=0,\widetilde{X}\cdot(1\otimes\delta+1\otimes v)=1\otimes(X\cdot(\delta+v))=0,

so Ker(X~)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{X})\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

Conversely, suppose that Ker(X~)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{X})\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]. Then there exists w[𝔤ad,𝔤ad]w\in[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] such that X~(1δ+w)=0\widetilde{X}\cdot(1\otimes\delta+w)=0. We have X~=1X\widetilde{X}=1\otimes X^{\prime} where XX^{\prime} is the action of XX on LadL_{ad}, so we can assume that w=1vw=1\otimes v for some vLscv\in L_{sc}. Then

0=X~(1δ+1v)=1(X(δ+v))0=\widetilde{X}\cdot(1\otimes\delta+1\otimes v)=1\otimes(X\cdot(\delta+v))

implies that X(δ+v)2LadX\cdot(\delta+v)\in 2L_{ad}.

Next we show that (ii) and (iii) are equivalent. It follows from (7.3) that 2Lad/2LscLad/Lsc/22L_{ad}/2L_{sc}\cong L_{ad}/L_{sc}\cong\mathbb{Z}/2\mathbb{Z} contains only two elements, 2Lsc2L_{sc} and 2δ+2Lsc2\delta+2L_{sc}. Therefore X(δ+v)2LadX\cdot(\delta+v)\in 2L_{ad} if and only if (iii)(a) or (iii)(b) holds. ∎

Next we make some observations that will allow us to reduce the proofs of our main results to the orthogonally indecomposable case. Continue with the notation as in Section 7.6, so V=U1UtV=U_{1}\perp\cdots\perp U_{t} with Ui=KV(i)U_{i}=K\otimes_{\mathbb{Z}}V_{\mathbb{Z}}^{(i)}, and 𝒰(i)\mathscr{U}_{\mathbb{Z}}^{(i)} is the Kostant \mathbb{Z}-form for 𝔰𝔭(V(i))\mathfrak{sp}(V_{\mathbb{C}}^{(i)}). Define

δi=121jivj(i)v2ij+1(i)\delta_{i}=\frac{1}{2}\sum_{1\leq j\leq\ell_{i}}v_{j}^{(i)}v_{2\ell_{i}-j+1}^{(i)}

for all 1it1\leq i\leq t. (Note that δ=δ1++δt\delta=\delta_{1}+\cdots+\delta_{t}.) Furthermore, for 1it1\leq i\leq t we define

Lsc(i)\displaystyle L_{sc}^{(i)} :=-span of vj(i)vk(i) and 12(vj(i))2 for all 1j,kni.\displaystyle:=\mathbb{Z}\text{-span of }v_{j}^{(i)}v_{k}^{(i)}\text{ and }\frac{1}{2}\left(v_{j}^{(i)}\right)^{2}\text{ for all }1\leq j,k\leq n_{i}.
Lad(i)\displaystyle L_{ad}^{(i)} :=-span of Lsc(i) and δi.\displaystyle:=\mathbb{Z}\text{-span of }L_{sc}^{(i)}\text{ and }\delta_{i}.

Then

(7.4) Lsc=Lsc(1)Lsc(t)iiV(i)V(i)L_{sc}=L_{sc}^{(1)}\oplus\cdots\oplus L_{sc}^{(t)}\oplus\bigoplus_{i\neq i^{\prime}}V_{\mathbb{Z}}^{(i)}V_{\mathbb{Z}}^{(i^{\prime})}

as \mathbb{Z}-modules.

Moreover Lad(i)S2(V(i))L_{ad}^{(i)}\subset S^{2}(V_{\mathbb{C}}^{(i)}) is an admissible lattice for the action of 𝔰𝔭(V(i))\mathfrak{sp}(V_{\mathbb{C}}^{(i)}), and

𝔤ad(i):=KLad(i)\mathfrak{g}_{ad}^{(i)}:=K\otimes_{\mathbb{Z}}L_{ad}^{(i)}

is an adjoint Lie algebra of type CiC_{\ell_{i}}. Here Sp(Ui)\operatorname{Sp}(U_{i}) acts on 𝔤ad(i)\mathfrak{g}_{ad}^{(i)} via the Chevalley construction.

Lemma 7.6.

Let X=X1++XtX=X_{1}+\cdots+X_{t}, where Xi𝒰(i)X_{i}\in\mathscr{U}_{\mathbb{Z}}^{(i)} acts nilpotently on S2(V)S^{2}(V_{\mathbb{C}}) for all 1it1\leq i\leq t. Denote by X~\widetilde{X} the linear map acting on 𝔤ad\mathfrak{g}_{ad}, given by reducing the action of XX on LadL_{ad} modulo pp. For 1it1\leq i\leq t, let Xi~\widetilde{X_{i}} be the linear map acting on 𝔤ad(i)\mathfrak{g}_{ad}^{(i)}, given by reducing the action of XiX_{i} on Lad(i)L_{ad}^{(i)} modulo pp.

Then the following statements hold.

  1. (i)

    Ker(X~)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{X})\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if there exist w1w_{1}, \ldots, wtw_{t} with wiLsc(i)w_{i}\in L_{sc}^{(i)} such that one of the following holds:

    1. (a)

      Xi(δi+wi)0mod2Lsc(i)X_{i}\cdot(\delta_{i}+w_{i})\equiv 0\mod{2L_{sc}^{(i)}} for all 1it1\leq i\leq t.

    2. (b)

      Xi(δi+wi)2δimod2Lsc(i)X_{i}\cdot(\delta_{i}+w_{i})\equiv 2\delta_{i}\mod{2L_{sc}^{(i)}} for all 1it1\leq i\leq t.

  2. (ii)

    If Ker(X~)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{X})\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}], then Ker(Xi~)[𝔤ad(i),𝔤ad(i)]\operatorname{Ker}(\widetilde{X_{i}})\not\subseteq[\mathfrak{g}_{ad}^{(i)},\mathfrak{g}_{ad}^{(i)}] for all 1it1\leq i\leq t.

Proof.

For (i), suppose first that there exist w1w_{1}, \ldots, wtw_{t} with wiLsc(i)w_{i}\in L_{sc}^{(i)} such that (i)(a) or (i)(b) holds. For w=w1++wtw=w_{1}+\cdots+w_{t}, we have

X(δ+w)=X1(δ1+w1)++Xt(δt+wt).X\cdot(\delta+w)=X_{1}\cdot(\delta_{1}+w_{1})+\cdots+X_{t}\cdot(\delta_{t}+w_{t}).

Thus X(δ+w)0mod2LscX\cdot(\delta+w)\equiv 0\mod{2L_{sc}} if (a) holds, and X(δ+w)2δmod2LscX\cdot(\delta+w)\equiv 2\delta\mod{2L_{sc}} if (b) holds. Therefore Ker(X~)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{X})\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] by Lemma 7.5.

Conversely, suppose that Ker(X~)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{X})\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]. Then by Lemma 7.5, there exists wLscw\in L_{sc} such that X(δ+w)0mod2LscX\cdot(\delta+w)\equiv 0\mod{2L_{sc}} or X(δ+w)2δmod2LscX\cdot(\delta+w)\equiv 2\delta\mod{2L_{sc}}. By (7.4) we can write w=w1++wt+ww=w_{1}+\cdots+w_{t}+w^{\prime}, where wiLsc(i)w_{i}\in L_{sc}^{(i)} for all 1it1\leq i\leq t and wiiV(i)V(i)w^{\prime}\in\bigoplus_{i\neq i^{\prime}}V_{\mathbb{Z}}^{(i)}V_{\mathbb{Z}}^{(i^{\prime})}.

Note that Lsc(1)Lsc(t)L_{sc}^{(1)}\oplus\cdots\oplus L_{sc}^{(t)} and iiV(i)V(i)\bigoplus_{i\neq i^{\prime}}V_{\mathbb{Z}}^{(i)}V_{\mathbb{Z}}^{(i^{\prime})} are 𝒰(i)\mathscr{U}_{\mathbb{Z}}^{(i)}-invariant. Thus

X(δ+w)=X1(δ1+w1)++Xt(δt+wt)+XwX\cdot(\delta+w)=X_{1}\cdot(\delta_{1}+w_{1})+\cdots+X_{t}\cdot(\delta_{t}+w_{t})+X\cdot w^{\prime}

with Xi(δi+wi)Lad(i)X_{i}\cdot(\delta_{i}+w_{i})\in L_{ad}^{(i)} for all 1it1\leq i\leq t, and XwiiV(i)V(i)X\cdot w^{\prime}\in\bigoplus_{i\neq i^{\prime}}V_{\mathbb{Z}}^{(i)}V_{\mathbb{Z}}^{(i^{\prime})}. Since 2δLsc(1)Lsc(t)2\delta\in L_{sc}^{(1)}\oplus\cdots\oplus L_{sc}^{(t)}, it follows that Xw=0X\cdot w^{\prime}=0. Moreover we have assumed that X(δ+w)0mod2LscX\cdot(\delta+w)\equiv 0\mod{2L_{sc}} or X(δ+w)2δmod2LscX\cdot(\delta+w)\equiv 2\delta\mod{2L_{sc}}, so either (i)(a) or (i)(b) holds.

Claim (ii) follows from (i) and Lemma 7.5. ∎

For the action of unipotent elements, we have the following result.

Lemma 7.7.

Let u=u1utSp(V)u=u_{1}\cdots u_{t}\in\operatorname{Sp}(V) be unipotent with u=u,1u,t𝒰u_{\mathbb{Z}}=u_{\mathbb{Z},1}\cdots u_{\mathbb{Z},t}\in\mathscr{U}_{\mathbb{Z}} as in Section 7.6. Denote the action of uu on 𝔤ad\mathfrak{g}_{ad} by u~\widetilde{u}, and denote the action of uiu_{i} on 𝔤ad(i)\mathfrak{g}_{ad}^{(i)} by ui~\widetilde{u_{i}}.

Let m1m\geq 1 be an integer. Then the following statements hold.

  1. (i)

    Ker(u~1)m[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{m}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if there exists w1w_{1}, \ldots, wtw_{t} with wiLsc(i)w_{i}\in L_{sc}^{(i)} such that one of the following holds:

    1. (a)

      (u,i1)m(δi+wi)0mod2Lsc(i)(u_{\mathbb{Z},i}-1)^{m}\cdot(\delta_{i}+w_{i})\equiv 0\mod{2L_{sc}^{(i)}} for all 1it1\leq i\leq t.

    2. (b)

      (u,i1)m(δi+wi)2δimod2Lsc(i)(u_{\mathbb{Z},i}-1)^{m}\cdot(\delta_{i}+w_{i})\equiv 2\delta_{i}\mod{2L_{sc}^{(i)}} for all 1it1\leq i\leq t.

  2. (ii)

    If Ker(u~1)m[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{m}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}], then Ker(ui~1)m[𝔤ad(i),𝔤ad(i)]\operatorname{Ker}(\widetilde{u_{i}}-1)^{m}\not\subseteq[\mathfrak{g}_{ad}^{(i)},\mathfrak{g}_{ad}^{(i)}] for all 1it1\leq i\leq t.

Proof.

It follows from Lemma 7.5 that Ker(u~1)m[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{m}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if there exists wLscw\in L_{sc} such that (u1)(δ+w)0mod2Lad(u_{\mathbb{Z}}-1)\cdot(\delta+w)\equiv 0\mod{2L_{ad}}.

By (7.4) we can write every wLscw\in L_{sc} in the form w=w1++wt+ww=w_{1}+\cdots+w_{t}+w^{\prime}, where wiLsc(i)w_{i}\in L_{sc}^{(i)} for all 1it1\leq i\leq t and wiiV(i)V(i)w^{\prime}\in\bigoplus_{i\neq i^{\prime}}V_{\mathbb{Z}}^{(i)}V_{\mathbb{Z}}^{(i^{\prime})}. Since u,iu_{\mathbb{Z},i} acts trivially on V(i)V_{\mathbb{Z}}^{(i^{\prime})} for iii\neq i^{\prime}, we have

(u1)m(δ+w)=(u,11)(δ1+w1)++(u,t1)(δt+wt)+(u1)w.(u_{\mathbb{Z}}-1)^{m}\cdot(\delta+w)=(u_{\mathbb{Z},1}-1)\cdot(\delta_{1}+w_{1})+\cdots+(u_{\mathbb{Z},t}-1)\cdot(\delta_{t}+w_{t})+(u_{\mathbb{Z}}-1)\cdot w^{\prime}.

Here (u,i1)(δi+wi)Lad(i)(u_{\mathbb{Z},i}-1)\cdot(\delta_{i}+w_{i})\in L_{ad}^{(i)} for all 1it1\leq i\leq t, and (u1)wiiV(i)V(i)(u_{\mathbb{Z}}-1)\cdot w^{\prime}\in\bigoplus_{i\neq i^{\prime}}V_{\mathbb{Z}}^{(i)}V_{\mathbb{Z}}^{(i^{\prime})}. Thus the result follows similarly to the proof of Lemma 7.6.∎

8. Lie algebras of type CC_{\ell} as GscG_{sc}-modules

In the notation of Section 7.3, let Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) be a simply connected simple algebraic group of type CC_{\ell}, with Lie algebra 𝔤sc=𝔰𝔭(V)\mathfrak{g}_{sc}=\mathfrak{sp}(V). In this section, we will make some initial observations about the structure of 𝔤sc\mathfrak{g}_{sc} and 𝔤ad\mathfrak{g}_{ad} as GscG_{sc}-modules.

Lemma 8.1.

Let Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) with Lie algebra 𝔤sc=𝔰𝔭(V)\mathfrak{g}_{sc}=\mathfrak{sp}(V). Then 𝔤scS2(V)\mathfrak{g}_{sc}\cong S^{2}(V)^{*} as GscG_{sc}-modules.

Proof.

A proof is given in [3, 5.4], alternatively this follows from the isomorphism KLsc𝔤scK\otimes_{\mathbb{Z}}L_{sc}\cong\mathfrak{g}_{sc} given in Section 7.4.∎

Lemma 8.2.

We have 𝔤sc/Z(𝔤sc)[𝔤ad,𝔤ad]\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc})\cong[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] as GscG_{sc}-modules.

Proof.

Let GadG_{ad} be a group of adjoint type CC_{\ell}, and let ψ:GscGad\psi:G_{sc}\rightarrow G_{ad} be an isogeny as in Lemma 7.1. Then the map dψ:𝔤sc𝔤ad\mathrm{d}\psi:\mathfrak{g}_{sc}\rightarrow\mathfrak{g}_{ad} is a morphism of GscG_{sc}-modules, with Kerdψ=Z(𝔤sc)\operatorname{Ker}\mathrm{d}\psi=Z(\mathfrak{g}_{sc}) [9, Lemma 2.2]. The image of dψ\mathrm{d}\psi contains all root elements of 𝔤ad\mathfrak{g}_{ad}, so by [9, Table 1] it is equal to [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}]. Therefore 𝔤sc/Kerdψ=𝔤sc/Z(𝔤sc)[𝔤ad,𝔤ad]\mathfrak{g}_{sc}/\operatorname{Ker}\mathrm{d}\psi=\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc})\cong[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] as GscG_{sc}-modules.∎

Let φ\varphi the linear map φ:S2(V)K\varphi:S^{2}(V)\rightarrow K defined by φ(xy)=b(x,y)\varphi(xy)=b(x,y) for all x,yVx,y\in V. Since bb is GscG_{sc}-invariant, it follows that φ\varphi is a surjective morphism of GscG_{sc}-modules.

Lemma 8.3.

Let Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) with Lie algebra 𝔤sc=𝔰𝔭(V)\mathfrak{g}_{sc}=\mathfrak{sp}(V). Then Kerφ(𝔤sc/Z(𝔤sc))\operatorname{Ker}\varphi\cong\left(\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc})\right)^{*} as GscG_{sc}-modules.

Proof.

It is clear that φ\varphi is surjective, so we have a short exact sequence

0KerφS2(V)K00\rightarrow\operatorname{Ker}\varphi\rightarrow S^{2}(V)\rightarrow K\rightarrow 0

of GscG_{sc}-modules. This induces a short exact sequence

0KS2(V)(Kerφ)00\rightarrow K\rightarrow S^{2}(V)^{*}\rightarrow\left(\operatorname{Ker}\varphi\right)^{*}\rightarrow 0

of GscG_{sc}-modules. Now S2(V)𝔤scS^{2}(V)^{*}\cong\mathfrak{g}_{sc} as GscG_{sc}-modules (Lemma 8.1), and 𝔤sc\mathfrak{g}_{sc} has a unique trivial submodule since Z(𝔤sc)KZ(\mathfrak{g}_{sc})\cong K [9, Table 1]. Thus we conclude that (Kerφ)𝔤sc/Z(𝔤sc)\left(\operatorname{Ker}\varphi\right)^{*}\cong\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}) as GscG_{sc}-modules, from which the lemma follows. ∎

Lemma 8.4.

Let G=Sp(V)G=\operatorname{Sp}(V) be simply connected and simple of type CC_{\ell}. Then:

  1. (i)

    There exists a uniserial GG-module WW with W=LG(0)|LG(2ϖ1)|LG(0)W=L_{G}(0)|L_{G}(2\varpi_{1})|L_{G}(0).

  2. (ii)

    A GG-module WW as in (i) is unique up to isomorphism.

  3. (iii)

    Let uGu\in G be a unipotent element. Then

    dimWu={dimVu+1, if dimVu is odd.dimVu+2, if dimVu is even.\dim W^{u}=\begin{cases}\dim V^{u}+1,&\text{ if }\dim V^{u}\text{ is odd.}\\ \dim V^{u}+2,&\text{ if }\dim V^{u}\text{ is even.}\end{cases}
Proof.

Let G=SO(V)G^{\prime}=\operatorname{SO}(V^{\prime}) with dimV=2+1\dim V^{\prime}=2\ell+1, so GG^{\prime} is a simple algebraic group of adjoint type BB_{\ell}. Let τ:GG\tau:G\rightarrow G^{\prime} be an exceptional isogeny as in [32, Theorem 28]. We can embed GG^{\prime} into a simple algebraic group SO(W)\operatorname{SO}(W) of type D+1D_{\ell+1} as the stabilizer of a nonsingular vector, see for example [23, Section 6.8]. Here dimW=2+2\dim W=2\ell+2, and as in [23, Section 6.8], we identify V=vWV^{\prime}=\langle v\rangle^{\perp}\subset W, where vWv\in W is a nonsingular vector.

It is straightforward to see that WW is a uniserial GG^{\prime}-module with

W=LG(0)|LG(ϖ1)|LG(0),W=L_{G^{\prime}}(0)|L_{G^{\prime}}(\varpi_{1}^{\prime})|L_{G^{\prime}}(0),

where ϖ1\varpi_{1}^{\prime} is the first fundamental highest weight for GG^{\prime}. Then the twist of WW by τ\tau is a uniserial GG-module WτW^{\tau} as in (i). For a unipotent element uGu\in G, the fixed point space of uu on the Frobenius twist LG(2ϖ1)LG(ϖ1)[1]L_{G}(2\varpi_{1})\cong L_{G}(\varpi_{1})^{[1]} is the same as on LG(ϖ1)VL_{G}(\varpi_{1})\cong V, because the Frobenius endomorphism preserves unipotent conjugacy classes. Therefore (iii) holds for WW by [15, Lemma 3.8].

It remains to check that WW is unique. To this end, note first that

ExtG1(K,LG(2ϖ1))HomG(2(V),K)\operatorname{Ext}_{G}^{1}(K,L_{G}(2\varpi_{1}))\cong\operatorname{Hom}_{G}(\wedge^{2}(V)^{*},K)

[14, II.2.14] and [3, 5.4]. Here HomG(2(V),K)2(V)GK\operatorname{Hom}_{G}(\wedge^{2}(V)^{*},K)\cong\wedge^{2}(V)^{G}\cong K, since VV has a unique GG-invariant alternating bilinear form up to a scalar. Thus there exists a unique nonsplit extension

0KZLG(2ϖ1)0,0\rightarrow K\rightarrow Z\rightarrow L_{G}(2\varpi_{1})\rightarrow 0,

up to isomorphism of GG-modules.

Since ExtG1(K,K)=ExtG2(K,K)=0\operatorname{Ext}_{G}^{1}(K,K)=\operatorname{Ext}_{G}^{2}(K,K)=0 [14, II.4.11] and ExtG1(K,LG(2ϖ1))K\operatorname{Ext}_{G}^{1}(K,L_{G}(2\varpi_{1}))\cong K, we have ExtG1(K,Z)K\operatorname{Ext}_{G}^{1}(K,Z)\cong K. Hence there exists a unique nonsplit extension

0ZWK0,0\rightarrow Z\rightarrow W\rightarrow K\rightarrow 0,

up to isomorphism of GG-modules. Every WW as in (i) is such an extension, so we conclude that WW is unique up to isomorphism.∎

Lemma 8.5.

Let Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V), so GscG_{sc} is simply connected and simple of type CC_{\ell}. Assume that \ell is even. Then there is a short exact sequence

0LG(ϖ2)𝔤adW00\rightarrow L_{G}(\varpi_{2})\rightarrow\mathfrak{g}_{ad}\rightarrow W\rightarrow 0

of GscG_{sc}-modules, where WW is as in Lemma 8.4 (i).

Proof.

As observed in [3, 5.6], in this case as a GscG_{sc}-module 𝔤ad\mathfrak{g}_{ad} is uniserial with 𝔤ad=LG(ϖ2)|LG(0)|LG(2ϖ1)|LG(0)\mathfrak{g}_{ad}=L_{G}(\varpi_{2})|L_{G}(0)|L_{G}(2\varpi_{1})|L_{G}(0). Thus the result follows from Lemma 8.4 (ii). ∎

Lemma 8.6.

Let G=Sp(V)G=\operatorname{Sp}(V) be simply connected and simple of type CC_{\ell}. Assume that 2mod4\ell\equiv 2\mod{4} and let uGu\in G be unipotent with VK[u]=V(2)V\downarrow K[u]=V(2\ell). Then dim𝔤adudim𝔤scu\dim\mathfrak{g}_{ad}^{u}\leq\dim\mathfrak{g}_{sc}^{u}.

Proof.

It follows from Lemma 4.4 that dim2(V)u=\dim\wedge^{2}(V)^{u}=\ell. Because 2mod4\ell\equiv 2\mod{4}, the smallest Jordan block size of uu on 2(V)\wedge^{2}(V) is equal to 22 [17, Lemma 4.12], and by [17, Theorem B] we have dimLG(ϖ2)u=1\dim L_{G}(\varpi_{2})^{u}=\ell-1.

For WW as in Lemma 8.5 (i), we have dimWu=2\dim W^{u}=2 by Lemma 8.5 (iii). Thus it follows from Lemma 8.5 that dim𝔤adudimLG(ϖ2)u+dimWu=+1\dim\mathfrak{g}_{ad}^{u}\leq\dim L_{G}(\varpi_{2})^{u}+\dim W^{u}=\ell+1. By Lemma 8.1 and Lemma 4.4 (ii) we have dim𝔤scu=+1\dim\mathfrak{g}_{sc}^{u}=\ell+1, so the result follows.∎

Remark 8.7.

As a corollary of our results, we will see later that equality dim𝔤adu=dim𝔤scu\dim\mathfrak{g}_{ad}^{u}=\dim\mathfrak{g}_{sc}^{u} holds in Lemma 8.6, without assumptions on \ell (Corollary 1.8).

9. Jordan block sizes of unipotent elements on 𝔤sc/Z(𝔤sc)\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc})

Let Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) be simply connected of type CC_{\ell}, and let bb be the GscG_{sc}-invariant alternating bilinear form defining GscG_{sc}. We have 𝔤scS2(V)\mathfrak{g}_{sc}\cong S^{2}(V)^{*} by Lemma 8.1, so for every unipotent element uGu\in G the Jordan block sizes of uu on 𝔤sc\mathfrak{g}_{sc} are known by the results described in Section 4. In this section, we will describe the Jordan block sizes of unipotent elements uGu\in G on 𝔤sc/Z(𝔤sc)\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}), in terms of Jordan block sizes of uu on 𝔤sc\mathfrak{g}_{sc}.

Throughout this section, we will denote by φ\varphi the linear map φ:S2(V)K\varphi:S^{2}(V)\rightarrow K defined by φ(xy)=b(x,y)\varphi(xy)=b(x,y) for all x,yVx,y\in V. Since bb is GscG_{sc}-invariant, it follows that φ\varphi is a surjective morphism of GscG_{sc}-modules. By Lemma 8.3, the Jordan normal form of a unipotent element uGu\in G on 𝔤sc/Z(𝔤sc)\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}) is the same as on Kerφ\operatorname{Ker}\varphi.

We will then describe the Jordan block sizes of unipotent elements uGu\in G on Kerφ\operatorname{Ker}\varphi. By Lemma 3.1, this amounts to finding the largest integer m0m\geq 0 such that Ker(u~1)mKerφ\operatorname{Ker}(\widetilde{u}-1)^{m}\subseteq\operatorname{Ker}\varphi, where u~\widetilde{u} is the action of uu on S2(V)S^{2}(V).

Lemma 9.1.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent. Suppose that we have a orthogonal decomposition V=U1UtV=U_{1}\perp\cdots\perp U_{t} as K[u]K[u]-modules. Denote the action of uu on S2(V)S^{2}(V) by u~\widetilde{u}, and the action of uu on S2(Ui)S^{2}(U_{i}) by ui~\widetilde{u_{i}}. Let m0m\geq 0 be an integer.

Then Ker(u~1)mKerφ\operatorname{Ker}(\widetilde{u}-1)^{m}\subseteq\operatorname{Ker}\varphi if and only if Ker(ui~1)mKerφ\operatorname{Ker}(\widetilde{u_{i}}-1)^{m}\subseteq\operatorname{Ker}\varphi for all 1it1\leq i\leq t.

Proof.

The orthogonal decomposition V=U1UtV=U_{1}\perp\cdots\perp U_{t} into K[u]K[u]-modules induces a decomposition

S2(V)=1itS2(Ui)1i<jtUiUjS^{2}(V)=\bigoplus_{1\leq i\leq t}S^{2}(U_{i})\perp\bigoplus_{1\leq i<j\leq t}U_{i}U_{j}

into K[u]K[u]-modules, where UiUj={xy:xUi,yUj}U_{i}U_{j}=\{xy:x\in U_{i},y\in U_{j}\}. We have UiUjKerφU_{i}U_{j}\subseteq\operatorname{Ker}\varphi for all 1i<jt1\leq i<j\leq t, from which the lemma follows.∎

With Lemma 9.1, we reduce to the case where VK[u]V\downarrow K[u] is orthogonally indecomposable. We first consider the case where VK[u]=V(2)V\downarrow K[u]=V(2\ell), so n=2n=2\ell. Let v1v_{1}, \ldots, vnv_{n} be a basis as in the definition of V(2)V(2\ell) (Definition 5.1). Then b(vi,vj)=1b(v_{i},v_{j})=1 if i+j=n+1i+j=n+1, and 0 otherwise. Furthermore, the action of uu is defined by

uv1\displaystyle uv_{1} =v1,\displaystyle=v_{1},
uvi\displaystyle uv_{i} =vi+vi1++v1 for all 2i+1,\displaystyle=v_{i}+v_{i-1}+\cdots+v_{1}\text{ for all }2\leq i\leq\ell+1,
uvi\displaystyle uv_{i} =vi+vi1 for all +1<in.\displaystyle=v_{i}+v_{i-1}\text{ for all }\ell+1<i\leq n.

We will denote vj=0v_{j}=0 for all j0j\leq 0 and j>nj>n.

Define γ=1ivivn+1i\gamma=\sum_{1\leq i\leq\ell}v_{i}v_{n+1-i}. We have a short exact sequence

0V[2]S2(V)2(V)0,0\rightarrow V^{[2]}\rightarrow S^{2}(V)\rightarrow\wedge^{2}(V)\rightarrow 0,

where V[2]V^{[2]} is the subspace generated by v2v^{2} for all vVv\in V. As noted in [17, Section 9], the image of γ\gamma in 2(V)\wedge^{2}(V) is fixed by the action of Sp(V)\operatorname{Sp}(V). Thus uγ=γ+xu\cdot\gamma=\gamma+x for some xV[2]x\in V^{[2]}, and more precisely we have the following.

Lemma 9.2.

Let uu and γ\gamma be as above. Then uγ=γ+1ivi2u\cdot\gamma=\gamma+\sum_{1\leq i\leq\ell}v_{i}^{2}, and γ+v+12\gamma+v_{\ell+1}^{2} is fixed by the action of uu.

Proof.

Denote si:=j0vijvnis_{i}:=\sum_{j\geq 0}v_{i-j}v_{n-i}. First we note that

u1i<vivni+1\displaystyle u\cdot\sum_{1\leq i<\ell}v_{i}v_{n-i+1} =1i<(j0vij)(vni+1+vni)\displaystyle=\sum_{1\leq i<\ell}\left(\sum_{j\geq 0}v_{i-j}\right)\left(v_{n-i+1}+v_{n-i}\right)
=1i<(si+si1+vivni+1)\displaystyle=\sum_{1\leq i<\ell}\left(s_{i}+s_{i-1}+v_{i}v_{n-i+1}\right)
(9.1) =s1+1i<vivni+1.\displaystyle=s_{\ell-1}+\sum_{1\leq i<\ell}v_{i}v_{n-i+1}.

By another calculation, we get

(9.2) uvv+1=(1jvj)(1j+1vj)=s1+vv+1+1jvj2.u\cdot v_{\ell}v_{\ell+1}=\left(\sum_{1\leq j\leq\ell}v_{j}\right)\left(\sum_{1\leq j\leq\ell+1}v_{j}\right)=s_{\ell-1}+v_{\ell}v_{\ell+1}+\sum_{1\leq j\leq\ell}v_{j}^{2}.

Adding (9.1) and (9.2) together, we conclude that uγ=γ+1ivi2u\cdot\gamma=\gamma+\sum_{1\leq i\leq\ell}v_{i}^{2}.

Since uv+12=(1i+1vi)2=1i+1vi2u\cdot v_{\ell+1}^{2}=\left(\sum_{1\leq i\leq\ell+1}v_{i}\right)^{2}=\sum_{1\leq i\leq\ell+1}v_{i}^{2}, it follows that γ+v+12\gamma+v_{\ell+1}^{2} is fixed by the action of uu. ∎

Lemma 9.3.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent such that VK[u]=V(2)V\downarrow K[u]=V(2\ell). Let u~\widetilde{u} be the action of uu on S2(V)S^{2}(V), and denote α=ν2()\alpha=\nu_{2}(\ell). Then the following hold:

  1. (i)

    Ker(u~1)2α1Kerφ\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi.

  2. (ii)

    Ker(u~1)2αKerφ\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.

Proof.

By Lemma 4.5 the smallest Jordan block size of u~\widetilde{u} is 2α2^{\alpha}, so (i) follows from Lemma 3.1.

For (ii), we first consider the case where α=0\alpha=0, so \ell is odd. In the notation used before the lemma, by Lemma 9.2 the vector γ+v+12\gamma+v_{\ell+1}^{2} is fixed by the action of uu. Furthermore φ(γ+v+12)=0\varphi(\gamma+v_{\ell+1}^{2})=\ell\neq 0, so Ker(u~1)Kerφ\operatorname{Ker}(\widetilde{u}-1)\not\subseteq\operatorname{Ker}\varphi, as claimed.

Suppose then that α>0\alpha>0. We have VK[u2α]=V(/2α1)2αV\downarrow K[u^{2^{\alpha}}]=V(\ell/2^{\alpha-1})^{2^{\alpha}} by [17, Lemma 6.13]. Combining this with Lemma 9.1 and the fact that (u~1)2α=u~2α1(\widetilde{u}-1)^{2^{\alpha}}=\widetilde{u}^{2^{\alpha}}-1, the claim follows from the α=0\alpha=0 case.∎

Lemma 9.4.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent such that VK[u]=W()V\downarrow K[u]=W(\ell). Let u~\widetilde{u} be the action of uu on S2(V)S^{2}(V), and denote α=ν2()\alpha=\nu_{2}(\ell). Then the following hold:

  1. (i)

    Ker(u~1)2α1Kerφ\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi.

  2. (ii)

    Ker(u~1)2αKerφ\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.

Proof.

We have V=WWV=W\oplus W^{*} as K[u]K[u]-modules, where WW and WW^{*} are totally isotropic subspaces. This induces a decomposition

S2(V)=S2(W)S2(W)WWS^{2}(V)=S^{2}(W)\oplus S^{2}(W^{*})\oplus WW^{*}

of K[u]K[u]-modules, where WWWWWW^{*}\cong W\otimes W^{*} is the subspace {xy:xW,yW}\{xy:x\in W,y\in W^{*}\}. We have S2(W),S2(W)KerφS^{2}(W),S^{2}(W^{*})\subseteq\operatorname{Ker}\varphi and the smallest Jordan block size in WWW\otimes W^{*} is 2α2^{\alpha} [16, Lemma 4.2], so (i) follows from Lemma 3.1.

For (ii), we first consider the case where α=0\alpha=0, so \ell is odd. Choose a basis v1v_{1}, \ldots, vv_{\ell} of WW, and let w1w_{1}, \ldots, ww_{\ell} be the corresponding dual basis in WW^{*}, so b(vi,wj)=δi,jb(v_{i},w_{j})=\delta_{i,j} for all 1i,j1\leq i,j\leq\ell. Then γ=1iviwi\gamma^{\prime}=\sum_{1\leq i\leq\ell}v_{i}w_{i} is fixed by the action of uu, see for example [16, Lemma 3.7]. We have φ(γ)=0\varphi(\gamma^{\prime})=\ell\neq 0, so Ker(u~1)Kerφ\operatorname{Ker}(\widetilde{u}-1)\not\subseteq\operatorname{Ker}\varphi, as claimed.

Suppose then that α>0\alpha>0. We have (u~1)2α=u~2α1(\widetilde{u}-1)^{2^{\alpha}}=\widetilde{u}^{2^{\alpha}}-1, and furthermore VK[u2α]=W(/2α)2αV\downarrow K[u^{2^{\alpha}}]=W(\ell/2^{\alpha})^{2^{\alpha}} by [17, Lemma 6.12]. Thus as in the proof of Lemma 9.3, the claim follows from Lemma 9.1 and the α=0\alpha=0 case.∎

Proposition 9.5.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent, with orthogonal decomposition

VK[u]=1itW(mi)1jsV(2kj).V\downarrow K[u]=\sum_{1\leq i\leq t}W(m_{i})\perp\sum_{1\leq j\leq s}V(2k_{j}).

Denote by α0\alpha\geq 0 the largest integer such that 2αmi,kj2^{\alpha}\mid m_{i},k_{j} for all ii and jj. Then

𝔤sc\displaystyle\mathfrak{g}_{sc} V2αV\displaystyle\cong V_{2^{\alpha}}\oplus V^{\prime}
𝔤sc/Z(𝔤sc)\displaystyle\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}) V2α1V\displaystyle\cong V_{2^{\alpha}-1}\oplus V^{\prime}

for some K[u]K[u]-module VV^{\prime}.

Proof.

Let u~\widetilde{u} be the action of uu on S2(V)S^{2}(V). By Lemma 8.1, Lemma 8.3, and Lemma 3.1, the proposition is equivalent to the statement that Ker(u~1)2α1Kerφ\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi and Ker(u~1)2αKerφ\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.

By Lemma 9.1, Lemma 9.3, and Lemma 9.4, we have Ker(u~1)2α1Kerφ\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi. Next note that either α=ν2(mi)\alpha=\nu_{2}(m_{i}) for some ii, or α=ν2(kj)\alpha=\nu_{2}(k_{j}) for some jj. It follows then from Lemma 9.1, together with Lemma 9.4 (if α=ν2(mi)\alpha=\nu_{2}(m_{i})) or Lemma 9.3 (if α=ν2(kj)\alpha=\nu_{2}(k_{j})) that Ker(u~1)2αKerφ\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.∎

10. Jordan block sizes of nilpotent elements on 𝔤sc/Z(𝔤sc)\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc})

We continue with the setup of the previous section. Let Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) be simply connected of type CC_{\ell} with Lie algebra 𝔤sc=𝔰𝔭(V)\mathfrak{g}_{sc}=\mathfrak{sp}(V). In this section, we will describe the Jordan block sizes of nilpotent elements e𝔰𝔭(V)e\in\mathfrak{sp}(V) on 𝔤sc/Z(𝔤sc)(Kerφ)\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc})\cong(\operatorname{Ker}\varphi)^{*} (Lemma 8.3). As in the previous section, we describe the Jordan block sizes of ee on Kerφ\operatorname{Ker}\varphi in terms of Jordan block sizes of ee on S2(V)S^{2}(V). Note that the Jordan normal form of ee on S2(V)S^{2}(V) is known by the results described in Section 4.

We begin by reducing the calculation to the orthogonally indecomposable case. After this we consider the different orthogonally indecomposable K[e]K[e]-modules in turn, and by combining the results we obtain Proposition 10.5, which is analogous to Proposition 9.5.

Lemma 10.1.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be unipotent. Suppose that we have a orthogonal decomposition V=U1UtV=U_{1}\perp\cdots\perp U_{t} as K[e]K[e]-modules. Denote the action of ee on S2(V)S^{2}(V) by e~\widetilde{e}, and the action of ee on S2(Ui)S^{2}(U_{i}) by ei~\widetilde{e_{i}}. Let m0m\geq 0 be an integer.

Then Ker(e~)mKerφ\operatorname{Ker}(\widetilde{e})^{m}\subseteq\operatorname{Ker}\varphi if and only if Ker(ei~)mKerφ\operatorname{Ker}(\widetilde{e_{i}})^{m}\subseteq\operatorname{Ker}\varphi for all 1it1\leq i\leq t.

Proof.

Follows with the same proof as Lemma 9.1.∎

Lemma 10.2.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent such that VK[e]=W()V\downarrow K[e]=W(\ell). Let e~\widetilde{e} be the action of ee on S2(V)S^{2}(V), and denote α=ν2()\alpha=\nu_{2}(\ell). Then the following hold:

  1. (i)

    Ker(e~)2α1Kerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi.

  2. (ii)

    Ker(e~)2αKerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.

Proof.

We proceed similarly to the proof of Lemma 9.4. By definition of W()W(\ell), we have a totally singular decomposition V=WWV=W\oplus W^{*}, where WW and WW^{*} are K[e]K[e]-modules on which ee acts with a single Jordan block of size \ell. This induces a decomposition

S2(V)=S2(W)S2(W)WWS^{2}(V)=S^{2}(W)\oplus S^{2}(W^{*})\oplus WW^{*}

of K[e]K[e]-modules, where WWWWWW^{*}\cong W\otimes W^{*}. Thus

Ker(e~)=Ker(e~S2(W))Ker(e~S2(W))Ker(e~WW).\operatorname{Ker}(\widetilde{e})=\operatorname{Ker}(\widetilde{e}_{S^{2}(W)})\oplus\operatorname{Ker}(\widetilde{e}_{S^{2}(W)})\oplus\operatorname{Ker}(\widetilde{e}_{WW^{*}}).

We have S2(W),S2(W)KerφS^{2}(W),S^{2}(W^{*})\subseteq\operatorname{Ker}\varphi since WW and WW^{*} are totally singular. Furthermore, the smallest Jordan block size in WWW\otimes W^{*} is 2α2^{\alpha} by Proposition 4.6 and [16, Lemma 4.2], so (i) follows from Lemma 3.1.

Next we consider (ii). We have VK[e2α]=W(/2α)2αV\downarrow K[e^{2^{\alpha}}]=W(\ell/2^{\alpha})^{2^{\alpha}} by Lemma 6.9. Therefore by Lemma 10.1, it suffices to consider the case where α=0\alpha=0, in which case \ell is odd. In this case, choose a basis v1v_{1}, \ldots, vv_{\ell} of WW, and let w1w_{1}, \ldots, ww_{\ell} be the corresponding dual basis in WW^{*}, so b(vi,wj)=δi,jb(v_{i},w_{j})=\delta_{i,j} for all 1i,j1\leq i,j\leq\ell. Then γ=1iviwi\gamma=\sum_{1\leq i\leq\ell}v_{i}w_{i} is annihilated by ee and φ(γ)=0\varphi(\gamma)=\ell\neq 0, so Ker(e~)Kerφ\operatorname{Ker}(\widetilde{e})\not\subseteq\operatorname{Ker}\varphi, as required.∎

Lemma 10.3.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent such that VK[e]=Wk()V\downarrow K[e]=W_{k}(\ell), where 0<k</20<k<\ell/2. Let e~\widetilde{e} be the action of ee on S2(V)S^{2}(V), and denote α=ν2()\alpha=\nu_{2}(\ell). Then the following hold:

  1. (i)

    Ker(e~)2α1Kerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi.

  2. (ii)

    Ker(e~)2αKerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.

Proof.

Let v1v_{1}, \ldots, vnv_{n} be the basis used in the definition of Wk()W_{k}(\ell). Then b(vi,vj)=1b(v_{i},v_{j})=1 if i+j=n+1i+j=n+1 and 0 otherwise. Furthermore, the action of ee is defined by

ev1\displaystyle ev_{1} =0,\displaystyle=0,
ev+1\displaystyle ev_{\ell+1} =0\displaystyle=0
evi\displaystyle ev_{i} =vi1 for all i{1,+1,nk+1}\displaystyle=v_{i-1}\text{ for all }i\not\in\{1,\ell+1,n-k+1\}
evnk+1\displaystyle ev_{n-k+1} =vnk+vk.\displaystyle=v_{n-k}+v_{k}.

We will denote vj=0v_{j}=0 for all j0j\leq 0 and j>nj>n.

Let W=vi:1iW=\langle v_{i}:1\leq i\leq\ell\rangle and Z=eivn:0i<Z=\langle e^{i}v_{n}:0\leq i<\ell\rangle. Then V=WZV=W\oplus Z, where WW and ZZ are K[e]K[e]-modules on which ee acts as a single Jordan block of size \ell. We take w1w_{1}, \ldots, ww_{\ell} as a basis of ZZ, where wi:=ei1vnw_{i}:=e^{i-1}v_{n} for all 1i1\leq i\leq\ell.

For the claims, we will first consider the case where α=0\alpha=0, so \ell is odd. In this case (i) is trivial. For (ii), note that the vector γ=1jvjwj\gamma=\sum_{1\leq j\leq\ell}v_{j}w_{j} is annihilated by the action of ee. For all 1j1\leq j\leq\ell, we have wj=vnj+1w_{j}=v_{n-j+1} or wj=vnj+1+vrw_{j}=v_{n-j+1}+v_{r} for some 1rk1\leq r\leq k. Therefore φ(γ)=0\varphi(\gamma)=\ell\neq 0, so Ker(e~)Kerφ\operatorname{Ker}(\widetilde{e})\not\subseteq\operatorname{Ker}\varphi, as claimed.

Suppose then that α>0\alpha>0. For (i), we will first prove that Ker(e~)Kerφ\operatorname{Ker}(\widetilde{e})\subseteq\operatorname{Ker}\varphi. To this end, note that the decomposition V=WZV=W\oplus Z induces a decomposition

S2(V)=S2(W)S2(Z)WZS^{2}(V)=S^{2}(W)\oplus S^{2}(Z)\oplus WZ

of K[e]K[e]-modules, where WZ={wz:wW,zZ}WZ=\{wz:w\in W,z\in Z\} is isomorphic to WZW\otimes Z. It follows from [18, Proof of Theorem 1.6] that S2(W)e=W[2]S^{2}(W)^{e}=W^{[2]} and S2(Z)e=Z[2]S^{2}(Z)^{e}=Z^{[2]}, so

Ker(e~)=W[2]Z[2]Ker(e~WZ).\operatorname{Ker}(\widetilde{e})=W^{[2]}\oplus Z^{[2]}\oplus\operatorname{Ker}(\widetilde{e}_{WZ}).

It is clear that W[2],Z[2]KerφW^{[2]},Z^{[2]}\subseteq\operatorname{Ker}\varphi, and Ker(e~WZ)Kerφ\operatorname{Ker}(\widetilde{e}_{WZ})\subseteq\operatorname{Ker}\varphi since the smallest Jordan block size in WZW\otimes Z is 2α2^{\alpha} (Proposition 4.6 and [16, Lemma 4.2]).

Thus Ker(e~)Kerφ\operatorname{Ker}(\widetilde{e})\subseteq\operatorname{Ker}\varphi. This also proves (i) in the case where α=1\alpha=1, so suppose next that α>1\alpha>1. We have VK[e2α1]=W(/2α1)2α1V\downarrow K[e^{2^{\alpha-1}}]=W(\ell/2^{\alpha-1})^{2^{\alpha-1}} by Lemma 6.9. Since /2α1\ell/2^{\alpha-1} is even, it follows from Lemma 10.2 and Lemma 10.1 that

(10.1) Ker(e~)2α1Kerφ.\operatorname{Ker}(\widetilde{e})^{2^{\alpha-1}}\subseteq\operatorname{Ker}\varphi.

In S2(W)S^{2}(W), S2(Z)S^{2}(Z), and WZW\otimes Z all Jordan block sizes of ee are powers of two (Theorem 4.7). In particular e~\widetilde{e} has no Jordan blocks of size 2α1<d<2α2^{\alpha-1}<d<2^{\alpha}, so by (10.1) and Lemma 3.1 we conclude that Ker(e~)2α1Kerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi, as required.

It remains to prove (ii) in the case where α>0\alpha>0. To this end, note first that we have VK[e2α]=W(/2α)2αV\downarrow K[e^{2^{\alpha}}]=W(\ell/2^{\alpha})^{2^{\alpha}} by Lemma 6.9. It follows then from Lemma 10.2 and Lemma 10.1 that Ker(e~)2αKerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.∎

Lemma 10.4.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent such that VK[e]=V(2)V\downarrow K[e]=V(2\ell). Let e~\widetilde{e} be the action of ee on S2(V)S^{2}(V), and denote α=ν2(2)\alpha=\nu_{2}(2\ell). Then the following hold:

  1. (i)

    Ker(e~)2α1Kerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi.

  2. (ii)

    Ker(e~)2αKerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.

Proof.

We have Ker(e~)=V[2]\operatorname{Ker}(\widetilde{e})=V^{[2]} by [18, Proof of Theorem 1.6], so Ker(e~)Kerφ\operatorname{Ker}(\widetilde{e})\subseteq\operatorname{Ker}\varphi. Since ee has no Jordan blocks of size 1<d<2α1<d<2^{\alpha} on S2(V)S^{2}(V) (Theorem 4.7), it follows from Lemma 3.1 that Ker(e~)2α1Kerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi.

Next we consider claim (ii). Since α=ν2(2)>0\alpha=\nu_{2}(2\ell)>0, we have VK[e2α]=W(/2α1)2α1V\downarrow K[e^{2^{\alpha}}]=W(\ell/2^{\alpha-1})^{2^{\alpha-1}} by Lemma 6.9. Because /2α1\ell/2^{\alpha-1} is odd, we conclude from Lemma 10.2 and Lemma 10.1 that Ker(e~)2αKerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.∎

Proposition 10.5.

Let e𝔰𝔭(V)e\in\mathfrak{sp}(V) be nilpotent, with orthogonal decomposition

VK[e]=1itW(mi)1jtWkj(j)1rt′′V(2dr).V\downarrow K[e]=\sum_{1\leq i\leq t}W(m_{i})\perp\sum_{1\leq j\leq t^{\prime}}W_{k_{j}}(\ell_{j})\perp\sum_{1\leq r\leq t^{\prime\prime}}V(2d_{r}).

Denote by α0\alpha\geq 0 the largest integer such that 2αmi,j,2dr2^{\alpha}\mid m_{i},\ell_{j},2d_{r} for all ii, jj, and rr.

Then

𝔤sc\displaystyle\mathfrak{g}_{sc} W2αV\displaystyle\cong W_{2^{\alpha}}\oplus V^{\prime}
𝔤sc/Z(𝔤sc)\displaystyle\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}) W2α1V\displaystyle\cong W_{2^{\alpha}-1}\oplus V^{\prime}

for some K[e]K[e]-module VV^{\prime}.

Proof.

Let e~\widetilde{e} be the action of ee on S2(V)S^{2}(V). By Lemma 8.1, Lemma 8.3, and Lemma 3.1, the proposition is equivalent to the statement that Ker(e~)2α1Kerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}-1}\subseteq\operatorname{Ker}\varphi and Ker(e~)2αKerφ\operatorname{Ker}(\widetilde{e})^{2^{\alpha}}\not\subseteq\operatorname{Ker}\varphi.

Thus similarly to the proof of Proposition 9.5, the result follows from Lemma 10.1, together with Lemma 10.2, Lemma 10.3, and Lemma 10.4.∎

11. Jordan block sizes of unipotent elements on 𝔤ad\mathfrak{g}_{ad}

Let Gsc=Sp(V)G_{sc}=\operatorname{Sp}(V) be simply connected and simple of type CC_{\ell}. In this section, we will describe the Jordan block sizes of unipotent elements uGscu\in G_{sc} acting on 𝔤ad\mathfrak{g}_{ad}. Our approach is to describe the Jordan block sizes of uu on 𝔤ad\mathfrak{g}_{ad} in terms of the Jordan block sizes on [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}], using Lemma 3.1. Since the Jordan block sizes of uu on [𝔤ad,𝔤ad]𝔤sc/Z(𝔤sc)[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]\cong\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}) (Lemma 8.2) are known by the results of Section 9, we then get an explicit description of the Jordan block sizes of uu on 𝔤ad\mathfrak{g}_{ad}.

Throughout this section we will consider 𝔤ad\mathfrak{g}_{ad} as constructed in Section 7.27.4, and we use the notation established there.

Lemma 11.1.

Let uSp(V)u\in\operatorname{Sp}(V) be the reduction modulo pp of u=x(α1)x(α1)u_{\mathbb{Z}}=x_{\mathbb{Z}}^{(\alpha_{1})}\cdots x_{\mathbb{Z}}^{(\alpha_{\ell-1})} as in Section 7.5, so that VK[u]=W()V\downarrow K[u]=W(\ell). Then the following hold:

  1. (i)

    (u1)δ=0(u_{\mathbb{Z}}-1)\cdot\delta=0.

  2. (ii)

    There exists vLscv\in L_{sc} such that (u1)(δ+v)2δmod2Lsc(u_{\mathbb{Z}}-1)\cdot(\delta+v)\equiv 2\delta\mod{2L_{sc}} if and only if \ell is even.

Proof.

For claim (i), consider Xαi=Ei,i+1Eni,ni+1X_{\alpha_{i}}=E_{i,i+1}-E_{n-i,n-i+1} for 1i11\leq i\leq\ell-1 from the Chevalley basis of 𝔰𝔭(V)\mathfrak{sp}(V_{\mathbb{C}}). It is clear that Xαiδ=0X_{\alpha_{i}}\cdot\delta=0, so x(αi)δ=δx_{\mathbb{Z}}^{(\alpha_{i})}\cdot\delta=\delta for all 1i11\leq i\leq\ell-1, from which (i) follows.

For (ii), we have (u1)(δ+v)=(u1)v(u_{\mathbb{Z}}-1)\cdot(\delta+v)=(u_{\mathbb{Z}}-1)\cdot v for all vLscv\in L_{sc} by (i). By Lemma 7.4, there exists vLscv\in L_{sc} such that (u1)(v)2δmod2Lsc(u_{\mathbb{Z}}-1)\cdot(v)\equiv 2\delta\mod{2L_{sc}} if and only if Im(u^1)Z(𝔤sc)\operatorname{Im}(\widehat{u}-1)\supseteq Z(\mathfrak{g}_{sc}), where u^\widehat{u} is the action of uu on 𝔤sc\mathfrak{g}_{sc}. It follows from Proposition 9.5 and Lemma 3.2 that Im(u^1)Z(𝔤sc)\operatorname{Im}(\widehat{u}-1)\supseteq Z(\mathfrak{g}_{sc}) if and only if \ell is even, so we conclude that (ii) holds.∎

Lemma 11.2.

Assume that \ell is odd. Let uSp(V)u\in\operatorname{Sp}(V) be the reduction modulo pp of u=x(α1)x(α)u_{\mathbb{Z}}=x_{\mathbb{Z}}^{(\alpha_{1})}\cdots x_{\mathbb{Z}}^{(\alpha_{\ell})} as in Section 7.5, so that VK[u]=V(2)V\downarrow K[u]=V(2\ell). Then the following hold:

  1. (i)

    There exists vLscv\in L_{sc} such that (u1)(δ+v)2δmod2Lsc(u_{\mathbb{Z}}-1)\cdot(\delta+v)\equiv 2\delta\mod{2L_{sc}}.

  2. (ii)

    There does not exist vLscv\in L_{sc} such that (u1)(δ+v)0mod2Lsc(u_{\mathbb{Z}}-1)\cdot(\delta+v)\equiv 0\mod{2L_{sc}}.

Proof.

For (i), first we calculate

uδ\displaystyle u_{\mathbb{Z}}\cdot\delta =x(α1)x(α1)(δ+12v2)\displaystyle=x_{\mathbb{Z}}^{(\alpha_{1})}\cdots x_{\mathbb{Z}}^{(\alpha_{\ell-1})}\cdot(\delta+\frac{1}{2}v_{\ell}^{2}) (by x(α)δ=δ+12v2)\displaystyle(\text{by }x_{\mathbb{Z}}^{(\alpha_{\ell})}\cdot\delta=\delta+\frac{1}{2}v_{\ell}^{2})
=δ+12(v1++v)2\displaystyle=\delta+\frac{1}{2}\left(v_{1}+\cdots+v_{\ell}\right)^{2} (by Lemma 11.1 (i))\displaystyle(\text{by Lemma \ref{lemma:deltaWLunip} (i)})
=δ+1i12vi2+1i<jvivj.\displaystyle=\delta+\sum_{1\leq i\leq\ell}\frac{1}{2}v_{i}^{2}+\sum_{1\leq i<j\leq\ell}v_{i}v_{j}.

Furthermore

u12v+12=12(v1++v+1)2=1i+112vi2+1i<j+1vivj,u_{\mathbb{Z}}\cdot\frac{1}{2}v_{\ell+1}^{2}=\frac{1}{2}\left(v_{1}+\cdots+v_{\ell+1}\right)^{2}=\sum_{1\leq i\leq\ell+1}\frac{1}{2}v_{i}^{2}+\sum_{1\leq i<j\leq\ell+1}v_{i}v_{j},

so

(11.1) u(δ+12v+12)δ+12v+12+1iviv+1mod2Lsc.u_{\mathbb{Z}}\cdot\left(\delta+\frac{1}{2}v_{\ell+1}^{2}\right)\equiv\delta+\frac{1}{2}v_{\ell+1}^{2}+\sum_{1\leq i\leq\ell}v_{i}v_{\ell+1}\mod{2L_{sc}}.

Next define

γ:=1j(1)/2v2j(vn2j+1+vn2j+2).\gamma:=\sum_{1\leq j\leq(\ell-1)/2}v_{2j}\left(v_{n-2j+1}+v_{n-2j+2}\right).

For 1j(1)/21\leq j\leq(\ell-1)/2, we denote sj:=1t2jvtvn2js_{j}:=\sum_{1\leq t\leq 2j}v_{t}v_{n-2j}. Then modulo 2Lsc2L_{sc}, we have

uγ\displaystyle u_{\mathbb{Z}}\cdot\gamma 1j(1)/2(1t2jvt)(vn2j+2+vn2j)\displaystyle\equiv\sum_{1\leq j\leq(\ell-1)/2}\left(\sum_{1\leq t\leq 2j}v_{t}\right)\left(v_{n-2j+2}+v_{n-2j}\right)
1j(1)/2(sj+sj1+v2j1vn2j+2+v2jvn2j+2)\displaystyle\equiv\sum_{1\leq j\leq(\ell-1)/2}\left(s_{j}+s_{j-1}+v_{2j-1}v_{n-2j+2}+v_{2j}v_{n-2j+2}\right)
s(1)/2+1j(1)/2v2j1vn2j+2+1j(1)/2v2jvn2j+2\displaystyle\equiv s_{(\ell-1)/2}+\sum_{1\leq j\leq(\ell-1)/2}v_{2j-1}v_{n-2j+2}+\sum_{1\leq j\leq(\ell-1)/2}v_{2j}v_{n-2j+2}
s(1)/2+1j(1)/2v2j1vn2j+2+(γ+1j(1)/2v2jvn2j+1)\displaystyle\equiv s_{(\ell-1)/2}+\sum_{1\leq j\leq(\ell-1)/2}v_{2j-1}v_{n-2j+2}+\left(\gamma+\sum_{1\leq j\leq(\ell-1)/2}v_{2j}v_{n-2j+1}\right)
γ+s(1)/2+1j1vjvnj+1\displaystyle\equiv\gamma+s_{(\ell-1)/2}+\sum_{1\leq j\leq\ell-1}v_{j}v_{n-j+1}
γ+s(1)/2+(2δ+vv+1)\displaystyle\equiv\gamma+s_{(\ell-1)/2}+\left(2\delta+v_{\ell}v_{\ell+1}\right)
γ+2δ+1jvjv+1.\displaystyle\equiv\gamma+2\delta+\sum_{1\leq j\leq\ell}v_{j}v_{\ell+1}.

Combining this with (11.1), we get

u(δ+12v+12+γ)δ+12v+12+γ+2δmod2Lsc.u_{\mathbb{Z}}\cdot\left(\delta+\frac{1}{2}v_{\ell+1}^{2}+\gamma\right)\equiv\delta+\frac{1}{2}v_{\ell+1}^{2}+\gamma+2\delta\mod{2L_{sc}}.

We conclude then that (i) holds, with v=12v+12+γv=\frac{1}{2}v_{\ell+1}^{2}+\gamma.

For claim (ii), suppose for the sake of contradiction that there exists vLscv\in L_{sc} such that (u1)(δ+v)0mod2Lsc(u_{\mathbb{Z}}-1)\cdot(\delta+v)\equiv 0\mod{2L_{sc}}. Then it follows from (i) that there exists wLscw\in L_{sc} such that (u1)w2δmod2Lsc(u_{\mathbb{Z}}-1)\cdot w\equiv 2\delta\mod{2L_{sc}}. By Lemma 7.4, we have Im(u^1)Z(𝔤sc)\operatorname{Im}(\widehat{u}-1)\supseteq Z(\mathfrak{g}_{sc}), where u^\widehat{u} is the action of uu on 𝔤sc\mathfrak{g}_{sc}. On the other hand Im(u^1)Z(𝔤sc)\operatorname{Im}(\widehat{u}-1)\not\supseteq Z(\mathfrak{g}_{sc}) by Proposition 9.5 and Lemma 3.2, so we have a contradiction.∎

Lemma 11.3.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent such that VK[u]=1itW(i)V\downarrow K[u]=\sum_{1\leq i\leq t}W(\ell_{i}), where i1\ell_{i}\geq 1 for all 1it1\leq i\leq t. Then Ker(u~1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

Proof.

Follows from Lemma 7.7 (i) and Lemma 11.1 (i). ∎

Lemma 11.4.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent such that VK[u]=1itW(mi)1jsV(2kj)V\downarrow K[u]=\sum_{1\leq i\leq t}W(m_{i})\perp\sum_{1\leq j\leq s}V(2k_{j}), where kj1k_{j}\geq 1 is odd for all 1js1\leq j\leq s. Assume that s>0s>0.

Denote the action of uu on 𝔤ad\mathfrak{g}_{ad} by u~\widetilde{u}. Then Ker(u~1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if mim_{i} is even for all 1it1\leq i\leq t.

Proof.

In view of Lemma 11.2 (ii) and Lemma 7.7 (i), we have Ker(u~1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if Lemma 7.7 (i)(b) holds. Then by Lemma 11.1 (ii), we conclude that Ker(u~1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if mim_{i} is even for all 1it1\leq i\leq t.∎

Lemma 11.5.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent such that VK[u]=V(2)V\downarrow K[u]=V(2\ell). Let u~\widetilde{u} be the action of uu on 𝔤ad\mathfrak{g}_{ad}, and denote α=ν2()\alpha=\nu_{2}(\ell). Then the following hold:

  1. (i)

    Ker(u~1)2α1[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}-1}\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

  2. (ii)

    Ker(u~1)2α[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

Proof.

We begin with the proof of (i), which is trivially true when α=0\alpha=0, so suppose that α>0\alpha>0.

We first consider α=1\alpha=1, in which case 2mod4\ell\equiv 2\mod{4}. Then

dim𝔤adu\displaystyle\dim\mathfrak{g}_{ad}^{u} dim𝔤scu\displaystyle\leq\dim\mathfrak{g}_{sc}^{u} by Lemma 8.6
=dim[𝔤ad,𝔤ad]u\displaystyle=\dim[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]^{u} by Lemma 8.2 and Proposition 9.5
dim𝔤adu.\displaystyle\leq\dim\mathfrak{g}_{ad}^{u}.

Thus we conclude that dim𝔤adu=dim𝔤scu=dim[𝔤ad,𝔤ad]u.\dim\mathfrak{g}_{ad}^{u}=\dim\mathfrak{g}_{sc}^{u}=\dim[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]^{u}. It follows then from Lemma 3.1 that Ker(u~1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}], as claimed by (i).

Suppose then that α>1\alpha>1. We have VK[u2α1]=V(/2α2)2α1V\downarrow K[u^{2^{\alpha-1}}]=V(\ell/2^{\alpha-2})^{2^{\alpha-1}} by [17, Lemma 6.13]. It follows then from Lemma 7.7 (ii) and the case α=1\alpha=1 that

Ker(u~1)2α1=Ker(u~2α11)[𝔤ad,𝔤ad].\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha-1}}=\operatorname{Ker}(\widetilde{u}^{2^{\alpha-1}}-1)\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

In particular, we have

(11.2) Ker(u~1)[𝔤ad,𝔤ad].\operatorname{Ker}(\widetilde{u}-1)\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

Next we note that it follows from Lemma 8.2, Proposition 9.5, and Lemma 4.5 that the smallest Jordan block size of uu on [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}] is equal to 2α12^{\alpha}-1. In particular, there are no Jordan blocks of size 1d<2α11\leq d<2^{\alpha}-1 for uu on [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}], which combined with Lemma 3.1 and (11.2) implies Ker(u~1)2α1[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}-1}\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

For (ii), note that VK[u2α]=V(/2α1)2αV\downarrow K[u^{2^{\alpha}}]=V(\ell/2^{\alpha-1})^{2^{\alpha}} by [17, Lemma 6.13]. Thus it follows from Lemma 11.4 that Ker(u~1)2α=Ker(u~2α1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}=\operatorname{Ker}(\widetilde{u}^{2^{\alpha}}-1)\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].∎

Proposition 11.6.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent, with orthogonal decomposition VK[u]=1itW(mi)1jsV(2kj)V\downarrow K[u]=\sum_{1\leq i\leq t}W(m_{i})\perp\sum_{1\leq j\leq s}V(2k_{j}). Assume that s>0s>0, and let α=max1jsν2(kj)\alpha=\max_{1\leq j\leq s}\nu_{2}(k_{j}). Let u~\widetilde{u} be the action of uu on 𝔤ad\mathfrak{g}_{ad}. Then the following statements hold:

  1. (i)

    Ker(u~1)2α1[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}-1}\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

  2. (ii)

    Ker(u~1)2α[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if α=ν2(kj)\alpha=\nu_{2}(k_{j}) for all 1js1\leq j\leq s, and ν2(mi)>α\nu_{2}(m_{i})>\alpha for all 1it1\leq i\leq t.

  3. (iii)

    Ker(u~1)2α+1[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}+1}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

Proof.

Claim (i) follows from Lemma 7.7 and Lemma 11.5.

Next we will prove (ii). In the case where α=0\alpha=0 we have kjk_{j} odd for all 1js1\leq j\leq s. Then the claim of (ii) is that Ker(u~1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if mim_{i} is even for all 1it1\leq i\leq t, which is precisely Lemma 11.4.

Suppose then that α>0\alpha>0. Write mi=mi2α+mi′′m_{i}=m_{i}^{\prime}2^{\alpha}+m_{i}^{\prime\prime} with 0mi′′<2α0\leq m_{i}^{\prime\prime}<2^{\alpha} and kj=kj2α1+kj′′k_{j}=k_{j}^{\prime}2^{\alpha-1}+k_{j}^{\prime\prime} with 0kj′′<2α10\leq k_{j}^{\prime\prime}<2^{\alpha-1}. Then for the restrictions of the summands in VK[u2α]V\downarrow K[u^{2^{\alpha}}], we have

W(mi)K[u2α]\displaystyle W(m_{i})\downarrow K[u^{2^{\alpha}}] =W(mi)2αmi′′W(mi+1)mi′′\displaystyle=W(m_{i}^{\prime})^{2^{\alpha}-m_{i}^{\prime\prime}}\perp W(m_{i}^{\prime}+1)^{m_{i}^{\prime\prime}}
V(2kj)K[u2α]\displaystyle V(2k_{j})\downarrow K[u^{2^{\alpha}}] ={V(kj/2α1)2α, if ν2(kj)=α.W(kj)2α1kj′′W(kj+1)kj′′, if ν2(kj)<α.\displaystyle=\begin{cases}V(k_{j}/2^{\alpha-1})^{2^{\alpha}},&\text{ if }\nu_{2}(k_{j})=\alpha.\\ W(k_{j}^{\prime})^{2^{\alpha-1}-k_{j}^{\prime\prime}}\perp W(k_{j}^{\prime}+1)^{k_{j}^{\prime\prime}},&\text{ if }\nu_{2}(k_{j})<\alpha.\end{cases}

for all 1it1\leq i\leq t and 1js1\leq j\leq s, by [17, Lemma 3.1, Lemma 6.12, Lemma 6.13]. In the above we denote W(0)=0W(0)=0.

We have α=ν2(kj)\alpha=\nu_{2}(k_{j}) for some jj, so V(kj/2α1)V(k_{j}/2^{\alpha-1}) appears as a summand of VK[u2α]V\downarrow K[u^{2^{\alpha}}]. Thus by Lemma 11.4, we have Ker(u~1)2α=Ker(u~2α1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}=\operatorname{Ker}(\widetilde{u}^{2^{\alpha}}-1)\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] if and only if all the summands of the form W(m)W(m) in VK[u2α]V\downarrow K[u^{2^{\alpha}}] have mm even.

Claim 1: Assume that ν2(mi)α\nu_{2}(m_{i})\leq\alpha for some ii. Then Ker(u~1)2α[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].
It suffices to show that W(mi)K[u2α]W(m_{i})\downarrow K[u^{2^{\alpha}}] has a summand W(m)W(m) with mm odd. If 0<mi′′<2α0<m_{i}^{\prime\prime}<2^{\alpha}, then W(mi)K[u2α]W(m_{i})\downarrow K[u^{2^{\alpha}}] has W(mi)W(m_{i}^{\prime}) and W(mi+1)W(m_{i}^{\prime}+1) as a summand. If mi′′=0m_{i}^{\prime\prime}=0, then W(mi)K[u2α]=W(mi)2αW(m_{i})\downarrow K[u^{2^{\alpha}}]=W(m_{i}^{\prime})^{2^{\alpha}} with mim_{i}^{\prime} odd, since ν2(mi)α\nu_{2}(m_{i})\leq\alpha.

Claim 2: Assume that ν2(kj)<α\nu_{2}(k_{j})<\alpha for some jj. Then Ker(u~1)2α[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].
It suffices to show that V(2kj)K[u2α]V(2k_{j})\downarrow K[u^{2^{\alpha}}] has a summand W(m)W(m) with mm odd. If 0<kj′′<2α10<k_{j}^{\prime\prime}<2^{\alpha-1}, then V(2kj)K[u2α]V(2k_{j})\downarrow K[u^{2^{\alpha}}] has W(kj)W(k_{j}^{\prime}) and W(kj+1)W(k_{j}^{\prime}+1) as a summand. Suppose then that kj′′=0k_{j}^{\prime\prime}=0. Then V(2kj)K[u2α]=W(kj)2α1V(2k_{j})\downarrow K[u^{2^{\alpha}}]=W(k_{j}^{\prime})^{2^{\alpha-1}}, with kjk_{j}^{\prime} is odd since ν2(kj)<α\nu_{2}(k_{j})<\alpha.

The “only if” part of (ii) follows from Claim 1 and Claim 2. Conversely, if α=ν2(kj)\alpha=\nu_{2}(k_{j}) for all 1js1\leq j\leq s and ν2(mi)>α\nu_{2}(m_{i})>\alpha for all 1it1\leq i\leq t, we have

VK[u2α]=1itW(mi/2α)2α1jsV(kj/2α1).V\downarrow K[u^{2^{\alpha}}]=\sum_{1\leq i\leq t}W(m_{i}/2^{\alpha})^{2^{\alpha}}\perp\sum_{1\leq j\leq s}V(k_{j}/2^{\alpha-1}).

Thus Ker(u~1)2α=Ker(u~2α1)[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}}=\operatorname{Ker}(\widetilde{u}^{2^{\alpha}}-1)\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] by Lemma 11.4.

For (iii), let u=u,1u,t+s𝒰u_{\mathbb{Z}}=u_{\mathbb{Z},1}\cdots u_{\mathbb{Z},t+s}\in\mathscr{U}_{\mathbb{Z}} be as in Section 7.5, so that uu is the reduction modulo pp of the action of uu_{\mathbb{Z}} on VV_{\mathbb{Z}}. Denote δ1\delta_{1}, \ldots, δt+s\delta_{t+s} as in Section 7.5. It follows from Lemma 7.7 (i), Lemma 11.1, and Lemma 11.5 that there exist wiLsc(i)w_{i}\in L_{sc}^{(i)} such that

(u,i1)2α(δi+wi)\displaystyle(u_{\mathbb{Z},i}-1)^{2^{\alpha}}\cdot(\delta_{i}+w_{i}) 0mod2Lsc(i)\displaystyle\equiv 0\mod{2L_{sc}^{(i)}} if 1it,\displaystyle\text{ if }1\leq i\leq t,
(u,i1)2α(δi+wi)\displaystyle(u_{\mathbb{Z},i}-1)^{2^{\alpha}}\cdot(\delta_{i}+w_{i}) 2δimod2Lsc(i)\displaystyle\equiv 2\delta_{i}\mod{2L_{sc}^{(i)}} if t+1it+s.\displaystyle\text{ if }t+1\leq i\leq t+s.

Then

(u1)2α(δ+w1++wt+s)2δt+1++2δt+smod2Lsc.(u_{\mathbb{Z}}-1)^{2^{\alpha}}\cdot(\delta+w_{1}+\cdots+w_{t+s})\equiv 2\delta_{t+1}+\cdots+2\delta_{t+s}\mod{2L_{sc}}.

On the other hand, we have u,iδi=δi+ziu_{\mathbb{Z},i}\cdot\delta_{i}=\delta_{i}+z_{i} for some ziLsc(i)z_{i}\in L_{sc}^{(i)}, as seen in the beginning of the proof of Lemma 11.2. Thus (u1)2δi0mod2Lsc(i)(u_{\mathbb{Z}}-1)\cdot 2\delta_{i}\equiv 0\mod{2L_{sc}^{(i)}} for all t+1it+st+1\leq i\leq t+s, and consequently

(u1)2α+1(δ+w1++wt+s)0mod2Lsc.(u_{\mathbb{Z}}-1)^{2^{\alpha}+1}\cdot(\delta+w_{1}+\cdots+w_{t+s})\equiv 0\mod{2L_{sc}}.

It follows then from Lemma 7.7 that Ker(u~1)2α+1[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{u}-1)^{2^{\alpha}+1}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}], as claimed by (iii).∎

Proposition 11.7.

Let uSp(V)u\in\operatorname{Sp}(V) be unipotent, with orthogonal decomposition VK[u]=1itW(mi)1jsV(2kj)V\downarrow K[u]=\sum_{1\leq i\leq t}W(m_{i})\perp\sum_{1\leq j\leq s}V(2k_{j}). Then the following statements hold:

  1. (i)

    Suppose that s=0s=0. Then

    𝔤adV1[𝔤ad,𝔤ad]\mathfrak{g}_{ad}\cong V_{1}\oplus[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]

    as K[u]K[u]-modules.

  2. (ii)

    Suppose that s>0s>0, and let α=max1jsν2(kj)\alpha=\max_{1\leq j\leq s}\nu_{2}(k_{j}). Then:

    1. (a)

      If α=ν2(kj)\alpha=\nu_{2}(k_{j}) for all 1js1\leq j\leq s and ν2(mi)>α\nu_{2}(m_{i})>\alpha for all 1it1\leq i\leq t, then

      𝔤ad\displaystyle\mathfrak{g}_{ad} V2αV\displaystyle\cong V_{2^{\alpha}}\oplus V^{\prime}
      [𝔤ad,𝔤ad]\displaystyle[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] V2α1V\displaystyle\cong V_{2^{\alpha}-1}\oplus V^{\prime}

      for some K[u]K[u]-module VV^{\prime}.

    2. (b)

      If α>ν2(kj)\alpha>\nu_{2}(k_{j}) for some 1js1\leq j\leq s, or ν2(mi)α\nu_{2}(m_{i})\leq\alpha for some 1it1\leq i\leq t, then

      𝔤ad\displaystyle\mathfrak{g}_{ad} V2α+1V\displaystyle\cong V_{2^{\alpha}+1}\oplus V^{\prime}
      [𝔤ad,𝔤ad]\displaystyle[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] V2αV\displaystyle\cong V_{2^{\alpha}}\oplus V^{\prime}

      for some K[u]K[u]-module VV^{\prime}.

Proof.

Claim (i) follows from Lemma 11.3 and Lemma 3.1. Similarly (ii) follows from Proposition 11.6 and Lemma 3.1. ∎

12. Jordan block sizes of nilpotent elements on 𝔤ad\mathfrak{g}_{ad}

Continuing with the setup of the previous section, in this section we describe the Jordan block sizes of nilpotent e𝔤sce\in\mathfrak{g}_{sc} on 𝔤ad\mathfrak{g}_{ad}, in terms of the Jordan block sizes on [𝔤ad,𝔤ad][\mathfrak{g}_{ad},\mathfrak{g}_{ad}]. The basic approach is similar to the previous section, but the proofs will be more simple due to the fact that VK[e2]V\downarrow K[e^{2}] is always has an orthogonal decomposition of the form 1itW(mi)\sum_{1\leq i\leq t}W(m_{i}) (Lemma 6.9).

Lemma 12.1.

Let e𝔤sce\in\mathfrak{g}_{sc} be nilpotent such that VK[e]=1itW(mi)V\downarrow K[e]=\sum_{1\leq i\leq t}W(m_{i}), and let e~\widetilde{e} be the action of ee on 𝔤ad\mathfrak{g}_{ad}. Then Kere~[𝔤ad,𝔤ad]\operatorname{Ker}\widetilde{e}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

Proof.

Let e=e,1++e,t𝔤e_{\mathbb{Z}}=e_{\mathbb{Z},1}+\cdots+e_{\mathbb{Z},t}\in\mathfrak{g}_{\mathbb{Z}} be as in Section 7.7, so that ee is the reduction modulo pp of ee_{\mathbb{Z}}. Then

e,i=Xαti+1++Xαti+mi1,e_{\mathbb{Z},i}=X_{\alpha_{t_{i}+1}}+\cdots+X_{\alpha_{t_{i}+m_{i}-1}},

where t1=0t_{1}=0 and ti=m1++mi1t_{i}=m_{1}+\cdots+m_{i-1} for all 1<it1<i\leq t.

As noted earlier (proof of Lemma 11.1), we have Xαiδ=0X_{\alpha_{i}}\cdot\delta=0 for all 1i<1\leq i<\ell. Therefore eδ=0e_{\mathbb{Z}}\cdot\delta=0, from which it follows that Kere~[𝔤ad,𝔤ad]\operatorname{Ker}\widetilde{e}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] (Lemma 7.5). ∎

Lemma 12.2.

Let e𝔤sce\in\mathfrak{g}_{sc} be nilpotent with VK[e]=V(2)V\downarrow K[e]=V(2\ell) or VK[e]=Wk()V\downarrow K[e]=W_{k}(\ell) for some 0<k</20<k<\ell/2. Let e~\widetilde{e} be the action of ee on 𝔤ad\mathfrak{g}_{ad}. Then Kere~[𝔤ad,𝔤ad]\operatorname{Ker}\widetilde{e}\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

Proof.

Let e𝔤e_{\mathbb{Z}}\in\mathfrak{g}_{\mathbb{Z}} be as in Section 7.7, so that ee is the reduction modulo pp of ee_{\mathbb{Z}}. Then

e=Xα1++Xα1+X2εr,e_{\mathbb{Z}}=X_{\alpha_{1}}+\cdots+X_{\alpha_{\ell-1}}+X_{2\varepsilon_{r}},

where r=r=\ell if VK[e]=V(2)V\downarrow K[e]=V(2\ell), and r=kr=k if VK[e]=Wk()V\downarrow K[e]=W_{k}(\ell).

Suppose that Kere~[𝔤ad,𝔤ad]\operatorname{Ker}\widetilde{e}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]. Then it follows from Lemma 7.5 that there exists vLscv\in L_{sc} such that e(δ+v)2Lade_{\mathbb{Z}}\cdot(\delta+v)\in 2L_{ad}. We have eδ=12vr2,e_{\mathbb{Z}}\cdot\delta=\frac{1}{2}v_{r}^{2}, so

(12.1) 12vr2=ev+2w\frac{1}{2}v_{r}^{2}=-e_{\mathbb{Z}}\cdot v+2w

for some wLadw\in L_{ad}.

Note that

e12vi2=±vivi1e_{\mathbb{Z}}\cdot\frac{1}{2}v_{i}^{2}=\pm v_{i}v_{i-1}

for all 1in1\leq i\leq n. Therefore eLscS2(V)e_{\mathbb{Z}}L_{sc}\subseteq S^{2}(V_{\mathbb{Z}}). Since also 2LadS2(V)2L_{ad}\subseteq S^{2}(V_{\mathbb{Z}}), it follows from (12.1) that 12vr2S2(V)\frac{1}{2}v_{r}^{2}\in S^{2}(V_{\mathbb{Z}}), which is impossible. We have a contradiction, so we conclude that Kere~[𝔤ad,𝔤ad]\operatorname{Ker}\widetilde{e}\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].∎

Lemma 12.3.

Let e𝔤sce\in\mathfrak{g}_{sc} be nilpotent, and let e~\widetilde{e} be the action of ee on 𝔤ad\mathfrak{g}_{ad}. Then Kere~[𝔤ad,𝔤ad]\operatorname{Ker}\widetilde{e}\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}], except possibly when VK[e]=1itW(i)V\downarrow K[e]=\sum_{1\leq i\leq t}W(\ell_{i}) for some 1\ell_{1}, \ldots, t\ell_{t}.

Proof.

Write VK[e]=U1UtV\downarrow K[e]=U_{1}\perp\cdots\perp U_{t}, where UiU_{i} is orthogonally indecomposable for all 1it1\leq i\leq t, and dimUi=2i\dim U_{i}=2\ell_{i} with i>0\ell_{i}>0. Let e=e,1++e,t𝔤e_{\mathbb{Z}}=e_{\mathbb{Z},1}+\cdots+e_{\mathbb{Z},t}\in\mathfrak{g}_{\mathbb{Z}} be as in Section 7.7, so that ee is the reduction modulo pp of ee_{\mathbb{Z}}.

The result follows from Lemma 7.6 (ii) (with Xi=e,iX_{i}=e_{\mathbb{Z},i}) and Lemma 12.2.∎

Lemma 12.4.

Let e𝔤sce\in\mathfrak{g}_{sc} be nilpotent, and let e~\widetilde{e} be the action of ee on 𝔤ad\mathfrak{g}_{ad}. Then Ker(e~)2[𝔤ad,𝔤ad]\operatorname{Ker}(\widetilde{e})^{2}\not\subseteq[\mathfrak{g}_{ad},\mathfrak{g}_{ad}].

Proof.

We have VK[e2]=1itW(mi)V\downarrow K[e^{2}]=\sum_{1\leq i\leq t}W(m_{i}) by Lemma 6.9, so the result follows from Lemma 12.1. ∎

Proposition 12.5.

Let e𝔤sce\in\mathfrak{g}_{sc} be nilpotent. Then the following hold:

  1. (i)

    If VK[e]=1itW(mi)V\downarrow K[e]=\sum_{1\leq i\leq t}W(m_{i}), then

    𝔤adW1[𝔤ad,𝔤ad]\mathfrak{g}_{ad}\cong W_{1}\oplus[\mathfrak{g}_{ad},\mathfrak{g}_{ad}]

    as K[e]K[e]-modules.

  2. (ii)

    If VK[e]V\downarrow K[e] is not of the form 1itW(mi)\sum_{1\leq i\leq t}W(m_{i}), then

    𝔤ad\displaystyle\mathfrak{g}_{ad} W2V\displaystyle\cong W_{2}\oplus V^{\prime}
    [𝔤ad,𝔤ad]\displaystyle[\mathfrak{g}_{ad},\mathfrak{g}_{ad}] W1V\displaystyle\cong W_{1}\oplus V^{\prime}

    for some K[e]K[e]-module VV^{\prime}.

Proof.

In case (i), the claim follows from Lemma 12.1 and Lemma 3.1. Similarly (ii) follows from Lemma 12.3, Lemma 12.4, and Lemma 3.1.∎

13. Proofs of main results

We can now prove the main results stated in the introduction; all of them are straightforward consequences of the results from previous sections.

Proof of Theorem 1.3.

By Proposition 9.5, we have

(13.1) 𝔤sc\displaystyle\mathfrak{g}_{sc} V2αV\displaystyle\cong V_{2^{\alpha}}\oplus V^{\prime}
𝔤sc/Z(𝔤sc)\displaystyle\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}) V2α1V\displaystyle\cong V_{2^{\alpha}-1}\oplus V^{\prime}

for some K[u]K[u]-module VV^{\prime}.

If s=0s=0, then by Proposition 11.7 and Lemma 8.2 we have 𝔤adV1𝔤sc/Z(𝔤sc)V1V2α1V\mathfrak{g}_{ad}\cong V_{1}\oplus\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc})\cong V_{1}\oplus V_{2^{\alpha}-1}\oplus V^{\prime} as K[u]K[u]-modules. Thus (i) holds.

For (ii), suppose that s>0s>0 and let β=max1jsν2(kj)\beta=\max_{1\leq j\leq s}\nu_{2}(k_{j}). Consider first the case where (ii)(a) holds, so ν2(kj)=β\nu_{2}(k_{j})=\beta for all 1js1\leq j\leq s and ν2(mi)>β\nu_{2}(m_{i})>\beta for all 1it1\leq i\leq t. Then α=β\alpha=\beta, so by Proposition 11.7 and Lemma 8.2 we have

𝔤ad\displaystyle\mathfrak{g}_{ad} V2αV′′\displaystyle\cong V_{2^{\alpha}}\oplus V^{\prime\prime}
𝔤sc/Z(𝔤sc)\displaystyle\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}) V2α1V′′\displaystyle\cong V_{2^{\alpha}-1}\oplus V^{\prime\prime}

for some K[u]K[u]-module V′′V^{\prime\prime}. We have VV′′V^{\prime}\cong V^{\prime\prime} by (13.1), so 𝔤ad𝔤sc\mathfrak{g}_{ad}\cong\mathfrak{g}_{sc} as K[u]K[u]-modules.

We consider then the case where (ii)(b) holds, so either ν2(kj)<β\nu_{2}(k_{j})<\beta for some 1js1\leq j\leq s, or ν2(mi)β\nu_{2}(m_{i})\leq\beta for some 1it1\leq i\leq t. Then by Proposition 11.7 and Lemma 8.2, we have

𝔤ad\displaystyle\mathfrak{g}_{ad} V2β+1V′′\displaystyle\cong V_{2^{\beta}+1}\oplus V^{\prime\prime}
𝔤sc/Z(𝔤sc)\displaystyle\mathfrak{g}_{sc}/Z(\mathfrak{g}_{sc}) V2βV′′\displaystyle\cong V_{2^{\beta}}\oplus V^{\prime\prime}

for some K[u]K[u]-module V′′V^{\prime\prime}. By (13.1) we have V′′V2α1V′′′V^{\prime\prime}\cong V_{2^{\alpha}-1}\oplus V^{\prime\prime\prime} and VV2βV′′′V^{\prime}\cong V_{2^{\beta}}\oplus V^{\prime\prime\prime} for some K[u]K[u]-module V′′′V^{\prime\prime\prime}. Thus

𝔤sc\displaystyle\mathfrak{g}_{sc} V2αV2βV′′′,\displaystyle\cong V_{2^{\alpha}}\oplus V_{2^{\beta}}\oplus V^{\prime\prime\prime},
𝔤ad\displaystyle\mathfrak{g}_{ad} V2α1V2β+1V′′′\displaystyle\cong V_{2^{\alpha}-1}\oplus V_{2^{\beta}+1}\oplus V^{\prime\prime\prime}

so claim (ii)(b) holds.∎

Proof of Theorem 1.4.

Similarly to Theorem 1.3, the result follows from Proposition 10.5, Lemma 8.2, and Proposition 12.5. ∎

Proof of Corollary 1.6 and Corollary 1.7.

Immediate from Theorem 1.3 and Theorem 1.4. ∎

Proof of Corollary 1.8.

For a regular unipotent uSp(V)u\in\operatorname{Sp}(V) we have VK[u]=V(2)V\downarrow K[u]=V(2\ell) [23, p. 61]. Then dim𝔤scu=+1\dim\mathfrak{g}_{sc}^{u}=\ell+1 by Lemma 8.1 and Lemma 4.4. It follows then from Corollary 1.6 that dim𝔤adu=dim𝔤scu=+1\dim\mathfrak{g}_{ad}^{u}=\dim\mathfrak{g}_{sc}^{u}=\ell+1.

For regular nilpotent e𝔰𝔭(V)e\in\mathfrak{sp}(V) we have similarly VK[e]=V(2)V\downarrow K[e]=V(2\ell) [23, p. 60]. It follows from Lemma 8.1 and [18, Corollary 4.2] that dim𝔤sce=2\dim\mathfrak{g}_{sc}^{e}=2\ell. Then dim𝔤ade=dim𝔤sce=2\dim\mathfrak{g}_{ad}^{e}=\dim\mathfrak{g}_{sc}^{e}=2\ell by Corollary 1.7, as claimed. ∎

References

  • [1] M. J. J. Barry. Decomposing tensor products and exterior and symmetric squares. J. Group Theory, 14(1):59–82, 2011.
  • [2] C. Chevalley. Sur certains groupes simples. Tohoku Math. J. (2), 7:14–66, 1955.
  • [3] M. F. Dowd and P. Sin. On representations of algebraic groups in characteristic two. Comm. Algebra, 24(8):2597–2686, 1996.
  • [4] R. M. Fossum. One-dimensional formal group actions. In Invariant theory (Denton, TX, 1986), volume 88 of Contemp. Math., pages 227–397. Amer. Math. Soc., Providence, RI, 1989.
  • [5] S. P. Glasby, C. E. Praeger, and B. Xia. Decomposing modular tensor products: ‘Jordan partitions’, their parts and pp-parts. Israel J. Math., 209(1):215–233, 2015.
  • [6] R. Gow and T. J. Laffey. On the decomposition of the exterior square of an indecomposable module of a cyclic pp-group. J. Group Theory, 9(5):659–672, 2006.
  • [7] J. A. Green. The modular representation algebra of a finite group. Illinois J. Math., 6:607–619, 1962.
  • [8] W. H. Hesselink. Nilpotency in classical groups over a field of characteristic 22. Math. Z., 166(2):165–181, 1979.
  • [9] G. M. D. Hogeweij. Almost-classical Lie algebras. I, II. Nederl. Akad. Wetensch. Indag. Math., 44(4):441–452, 453–460, 1982.
  • [10] X. D. Hou. Elementary divisors of tensor products and pp-ranks of binomial matrices. Linear Algebra Appl., 374:255–274, 2003.
  • [11] J. E. Humphreys. Introduction to Lie algebras and representation theory. Springer-Verlag, New York-Berlin, 1972. Graduate Texts in Mathematics, Vol. 9.
  • [12] K.-i. Iima and R. Iwamatsu. On the Jordan decomposition of tensored matrices of Jordan canonical forms. Math. J. Okayama Univ., 51:133–148, 2009.
  • [13] J. C. Jantzen. Darstellungen halbeinfacher algebraischer Gruppen und zugeordnete kontravariante Formen. Bonn. Math. Schr., (67):v+124, 1973.
  • [14] J. C. Jantzen. Representations of algebraic groups, volume 107 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, second edition, 2003.
  • [15] M. Korhonen. Unipotent elements forcing irreducibility in linear algebraic groups. J. Group Theory, 21(3):365–396, 2018.
  • [16] M. Korhonen. Jordan blocks of unipotent elements in some irreducible representations of classical groups in good characteristic. Proc. Amer. Math. Soc., 147(10):4205–4219, 2019.
  • [17] M. Korhonen. Hesselink normal forms of unipotent elements in some representations of classical groups in characteristic two. J. Algebra, 559:268–319, 2020.
  • [18] M. Korhonen. Decomposition of exterior and symmetric squares in characteristic two. Linear Algebra Appl., 624:349–363, 2021.
  • [19] M. Korhonen. Jordan blocks of nilpotent elements in some irreducible representations of classical groups in good characteristic. J. Pure Appl. Algebra, 225(8):Paper No. 106694, 10, 2021.
  • [20] M. Korhonen, D. I. Stewart, and A. R. Thomas. Representatives for unipotent classes and nilpotent orbits. Comm. Algebra, 50(4):1641–1661, 2022.
  • [21] R. Lawther. Jordan block sizes of unipotent elements in exceptional algebraic groups. Comm. Algebra, 23(11):4125–4156, 1995.
  • [22] R. Lawther. Correction to: “Jordan block sizes of unipotent elements in exceptional algebraic groups”. Comm. Algebra, 26(8):2709, 1998.
  • [23] M. W. Liebeck and G. M. Seitz. Unipotent and nilpotent classes in simple algebraic groups and Lie algebras, volume 180 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2012.
  • [24] J. D. McFall. How to compute the elementary divisors of the tensor product of two matrices. Linear and Multilinear Algebra, 7(3):193–201, 1979.
  • [25] C. W. Norman. On Jordan bases for two related linear mappings. J. Math. Anal. Appl., 175(1):96–104, 1993.
  • [26] C. W. Norman. On the Jordan form of the tensor product over fields of prime characteristic. Linear and Multilinear Algebra, 38(4):351–371, 1995.
  • [27] C. W. Norman. On Jordan bases for the tensor product and Kronecker sum and their elementary divisors over fields of prime characteristic. Linear Multilinear Algebra, 56(4):415–451, 2008.
  • [28] T. Ralley. Decomposition of products of modular representations. Bull. Amer. Math. Soc., 72:1012–1013, 1966.
  • [29] R. Ree. On some simple groups defined by C. Chevalley. Trans. Amer. Math. Soc., 84:392–400, 1957.
  • [30] J.-C. Renaud. The decomposition of products in the modular representation ring of a cyclic group of prime power order. J. Algebra, 58(1):1–11, 1979.
  • [31] B. Srinivasan. The modular representation ring of a cyclic pp-group. Proc. London Math. Soc. (3), 14:677–688, 1964.
  • [32] R. Steinberg. Lectures on Chevalley groups, volume 66 of University Lecture Series. American Mathematical Society, Providence, RI, 2016.
  • [33] D. I. Stewart. On the minimal modules for exceptional Lie algebras: Jordan blocks and stabilizers. LMS J. Comput. Math., 19(1):235–258, 2016. Corrected version in arXiv:1508.02918.