Hostname: page-component-745bb68f8f-kw2vx Total loading time: 0 Render date: 2025-02-05T02:21:40.295Z Has data issue: false hasContentIssue false

Generic stabilizers for simple algebraic groups acting on orthogonal and symplectic Grassmannians

Published online by Cambridge University Press:  24 January 2025

Aluna Rizzoli*
Affiliation:
King’s College London, Strand, London WC2R 2LS, UK;

Abstract

We consider faithful actions of simple algebraic groups on self-dual irreducible modules and on the associated varieties of totally singular subspaces, under the assumption that the dimension of the group is at least as large as the dimension of the variety. We prove that in all but a finite list of cases, there is a dense open subset where the stabilizer of any point is conjugate to a fixed subgroup, called the generic stabilizer. We use these results to determine whether there exists a dense orbit. This in turn lets us complete the answer to the problem of determining all pairs of maximal connected subgroups of a classical group with a dense double coset.

Type
Algebra
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press

1 Introduction and statement of results

Let G be a simple algebraic group over an algebraically closed field K of characteristic p, where we take $p=\infty $ if K has characteristic zero. Let V be a nontrivial irreducible $KG$ -module of dimension d. For $1\leq k\leq d$ , the Grassmannian variety $\mathcal {G}_k(V)$ consists of all k-dimensional subspaces of V and is isomorphic to the variety $SL(V)/P$ , where P is a maximal parabolic subgroup of $SL(V)$ stabilising a k-dimensional subspace. Assume that the module V is self-dual. Then the group G preserves a nondegenerate bilinear form, which is either symmetric or alternating (unless $p = 2$ when it is both). If the form is symmetric and G preserves an associated quadratic form, we say that the module V is orthogonal, and we say it is symplectic otherwise. For $1\leq k\leq \frac {d}{2}$ , we denote by $\mathcal {S}_k(V)$ the variety of totally singular k-dimensional subspaces of V. Any such variety is irreducible unless $k=\frac {d}{2}$ and V is orthogonal, in which case the two $SO(V)$ -orbits on $\mathcal {S}_k(V)$ are its irreducible components, which we shall denote by $\mathcal {S}_{k'}(V)$ and $\mathcal {S}_{k"}(V)$ , or $\mathcal {S}_{k}'(V)$ and $\mathcal {S}_{k}"(V)$ . Note that there is no intrinsic way of choosing which of the two orbits is labeled $\mathcal {S}_{k'}(V)$ , and therefore, such choice is arbitrary and usually simply dependent on the order of consideration. As V is orthogonal or symplectic, each such irreducible variety is isomorphic to $SO(V)/P$ or $Sp(V)/P$ , where P is a parabolic subgroup of $SO(V)$ or $Sp(V)$ (maximal unless V is orthogonal and $k=\frac {d}{2}-1$ ), and the elements of the variety are orthogonal Grassmannians or symplectic Grassmannians.

If G acts faithfully on a variety X, we say that the action has generic stabilizer S if there exists a nonempty open subset $U\subseteq X$ such that the stabilizer $G_u$ is conjugate to S for all $u\in U$ .

We say that the action has semi-generic stabilizer S if there exists a nonempty open subset $U\subseteq X$ such that $G_u$ is isomorphic to S for all $u\in U$ .

In general, we say that G acting on X (not necessarily faithfully) has a (semi-)generic stabilizer if $G/G_X$ , where $G_X$ denotes the kernel of the action, has a (semi-)generic stabilizer for its faithful action on X. Note that when $X=\mathcal {G}_k(V)$ or $X=\mathcal {S}_k(V)$ , the kernel $G_X$ is precisely the center of G.

In characteristic zero, generic stabilizers exist under mild hypotheses. In this case, it follows from [Reference Andreev, Vinberg and Élashvili1] that the generic stabilizer is positive dimensional if and only if $\dim G\geq \dim X$ . However, even in characteristic $0$ , there are examples of actions with no generic stabilizers (see [Reference Richardson25]).

If $p<\infty $ , there are even more instances where generic stabilizers do not exist, as there is no analogue of Richardson’s result. For example, in [Reference Martin21, Example 8.3], we find a construction for an $SL_2(K)$ -action on an affine variety in positive characteristic, with no generic stabilizer.

Again, let G be a simple algebraic group over an algebraically closed field K of characteristic p and V be a nontrivial irreducible $KG$ -module. Recently, Guralnick and Lawther have solved the generic stabilizer problem for the action on $X=\mathcal {G}_k(V)$ . In [Reference Guralnick and Lawther10], they proved that if $X=V$ or $X=\mathcal {G}_k(V)$ , then the action of G on X has a generic stabilizer, unless $G=B_3$ , $p=2$ , V is the spin module for G, and $k=4$ , in which case the action has a semi-generic stabilizer but not a generic stabilizer. They showed that for such actions, the generic stabilizer is in general trivial, and they otherwise determined all nontrivial (semi-)generic stabilizers explicitly.

In this paper, we treat the action of G on the orthogonal and symplectic Grassmannians of self-dual irreducible G-modules (i.e., $X = \mathcal {S}_k(V)$ ). We only deal with the case $\dim G\geq \dim X$ . The reason for this is two-fold. We will be interested in applications to questions about the existence of dense orbits and dense double cosets, for which we only need to be concerned with the cases where $\dim G\geq \dim X$ . Secondly, as shown in [Reference Guralnick and Lawther10], the strategy for dealing with the situation $\dim G<\dim X$ presents entirely different challenges. The case $\dim G < \dim X$ shall be the subject of future work. Note that (semi-)generic stabilizers for the action on $\mathcal {S}_k(V)$ are generally going to be radically different from the ones for the action on $\mathcal {G}_k(V)$ . Indeed, it is often going to be the case that the generic stabilizer for the action on $\mathcal {G}_k(V)$ is finite, while the generic stabilizer for the action on $\mathcal {S}_k(V)$ is positive-dimensional. We shall now state our first result. The modules V are denoted by their highest weight, the groups G by their Dynkin diagram.

Theorem 1. Let G be a simple algebraic group over an algebraically closed field of characteristic p, and V a self-dual nontrivial irreducible G-module of dimension d and highest weight $\lambda $ . For $1\leq k\leq \frac {d}{2}$ such that $\dim G\geq \dim \mathcal {S}_k(V)$ , if the action of G on $\mathcal {S}_k(V)$ has no generic stabilizer, then $p=2$ and $(G,\lambda ,k)$ is one of the following:

  1. (i) $(E_7,\lambda _7,2)$ ;

  2. (ii) $(D_6,\lambda _6,2)$ ;

  3. (iii) $(A_5,\lambda _3,2)$ ;

  4. (iv) $(B_4,\lambda _4,8)$ .

In the first three cases, the action of G on $\mathcal {S}_k(V)$ has no generic stabilizer but does have a semi-generic one. In the last case, the action of G on $\mathcal {S}_k'(V)$ has a generic stabilizer, but the action of G on $\mathcal {S}_k"(V)$ only has a semi-generic one.

While Theorem 1 guarantees the existence, in Theorem 2, we shall explicitly determine the structure of every (semi-)generic stabilizer. The proof of Theorem 1 involves a quick reduction to a finite list of families of cases to be considered, followed by a case-by-case analysis using many of the methods adopted in [Reference Guralnick and Lawther10]. This case-by-case analysis is the subject of the majority of this paper.

Before stating our remaining results, we shall set up some more notation. If G is semisimple, let T be a fixed maximal torus, and $\Phi $ the root system for G with respect to T, described by its Dynkin diagram. The root system has positive roots $\Phi ^+$ and a base $\Delta =\{\alpha _1,\alpha _2,\dots ,\alpha _n\}$ of simple roots. For the simple algebraic groups, the ordering of the simple roots is taken according to Bourbaki [Reference Bourbaki4]. For a $KG$ -module V and a weight $\mu $ of G, we write $V_\mu $ for the $\mu $ -weight space of V. If G acts on a set X, for $x\in X$ , we denote by $G_x$ the stabilizer in G of x. We use P to denote a parabolic subgroup containing a Borel subgroup $B\geq T$ and $P_k$ to denote the maximal parabolic subgroup obtained by deleting the k-th node of the Dynkin diagram for G. Similarly, we use $P_{i,j}$ to denote the parabolic subgroup obtained by deleting the i-th and j-th nodes of the Dynkin diagram for G. We use $T_i$ to denote an i-dimensional torus, $Sym(n)$ and $Alt(n)$ to denote the symmetric group and the alternating group on a set of size n, and $Dih(2n)$ for a dihedral group of order $2n$ . Throughout the paper, we work modulo field twists and exceptional isogenies. So, for example, we only treat one of $C_n$ or $B_n$ in characteristic $2$ , and we always assume that the highest weight $\lambda $ is not a multiple of p.

We now set up further notation to better encapsulate the exact setup we will adopt. As in [Reference Guralnick and Lawther10], we define a quadruple to be a $4$ -tuple of the form $(G,\lambda ,p,k)$ with the following properties:

  1. (i) G is a simple algebraic group over an algebraically closed field of characteristic p;

  2. (ii) $V=V_G(\lambda )$ is an irreducible G-module;

  3. (iii) $1\leq k\leq \frac {\dim V}{2}$ .

We say that a quadruple $(G,\lambda ,p,k)$ is small if

$$\begin{align*}\dim G\geq \dim\mathcal{G}_k(V).\end{align*}$$

We say that a quadruple $(G,\lambda ,p,k)$ is self-dual if $V_G(\lambda )$ is self-dual, in which case the quadruple is ts-small if

$$\begin{align*}\dim G\geq \dim\mathcal{S}_k(V).\end{align*}$$

We say that the quadruple $(G,\lambda ,p,k)$ has a (semi-)generic stabilizer if the action of G on $X=\mathcal {G}_k(V)$ has a (semi-)generic stabilizer. For a self-dual quadruple $(G,\lambda ,p,k)$ , we say it has a (semi-)generic $ts$ -stabilizer if the action of G on $X=\mathcal {S}_k(V)$ has a (semi-)generic stabilizer. In this paper, we classify (semi-)generic $ts$ -stabilizers of $ts$ -small quadruples.

Our main result will be given in a single table (Table 1). In the first column, we have the type of our simple algebraic group G; in the second column, the highest weight of the irreducible G-module V; in the third column, we list the rank $\ell $ of G; in the fourth column, we have conditions on p; and in the fifth column, we have the particular k which specifies which variety $\mathcal {S}_k(V)$ we are acting on. We then have a column listing the (semi-)generic stabilizer for the action, denoted by $C_{\mathcal {S}_k(V)}$ in which we use an asterisk to indicate whether the stabilizer is not generic but semi-generic. In the columns ‘Orth?’ and ‘Dense?’, we indicate whether the module V is orthogonal and whether there exists a dense orbit for the action on $\mathcal {S}_k(V)$ . In the last column, we give the number of the proposition within the paper where the relevant information is obtained.

Table 1 Generic $ts$ -stabilizers for $ts$ -small quadruples.

Theorem 2. The (semi-)generic $ts$ -stabilizer for a $ts$ -small quadruple is given in Table 1. In addition, the existence or nonexistence of a dense orbit is indicated.

We are quickly able to determine whether a dense orbit on X exists, as the set of points in X that have stabilizers of minimal dimension is open in X (see, for example, [Reference Martin21, Lemma 2.1]). This means that the dimension of a (semi-)generic stabilizer is actually the minimal dimension of any stabilizer, and if this is larger than $\dim G-\dim X$ , there is no dense orbit. There is extensive interest in the literature around the existence of a dense orbit. In particular, if G is a reductive algebraic group and V is an irreducible rational G-module such that G has a dense orbit on V, the pair $(G,V)$ is called a prehomogeneous vector space, often shortened to $PV$ -space. A classification of $PV$ -spaces was determined in characteristic zero in [Reference Sato and Kimura29] and extended to positive characteristic in [Reference Zhijie32][Reference Zhijie33]. The fact that this classification covers all semisimple algebraic groups implies that the density question is already understood for the action on $\mathcal {G}_k(V)$ . Indeed, G has a dense orbit on $\mathcal {G}_k(V)$ if and only if $GL_k\otimes G$ has a dense orbit on $V_1\otimes V$ , where $V_1$ is the natural module for $GL_k$ . The next result summarises the answer to the dense-orbit question for the action on orthogonal and symplectic Grassmannians.

Theorem 3. Let G be a simple algebraic group over an algebraically closed field of characteristic p, and V a self-dual nontrivial irreducible G-module of dimension d and highest weight $\lambda $ . For $1\leq k\leq \frac {d}{2}$ , the action of G on $\mathcal {S}_k(V)$ has a dense orbit if and only if $\dim G\geq \dim \mathcal {S}_k(V)$ , the zero-weight space of V has dimension at most $2$ , and $(G,\lambda ,p,k)$ is not one of the following:

  1. (i) $(A_5,\lambda _3,any,2)$ ;

  2. (ii) $(B_4,\lambda _4,any,7)$ ;

  3. (iii) $(B_4,\lambda _4,any,8")$ ;

  4. (iv) $(D_6,\lambda _6,any,2)$ ;

  5. (v) $(E_7,\lambda _7,any,2)$ .

Remark 1. The only $ts$ -small quadruples $(G,\lambda ,p,k)$ with a zero-weight space of dimension at least $3$ have $k=1$ , with either $V_G(\lambda )$ a composition factor of the adjoint module for G, or $G=C_\ell $ and $\lambda = \lambda _2$ .

Denote by $Cl(V)$ a classical group with natural module V. Given that $X=\mathcal {G}_k(V)$ and $X=\mathcal {S}_k(V)$ are varieties of cosets of the form $Cl(V)/P$ , where P is a parabolic subgroup of $Cl(V)$ , the action of G on X has a dense orbit if and only if there is a dense $(G,P)$ -double coset in $Cl(V)$ . We now seek to complete the answer to the question of existence of a dense double coset in the following sense. In Theorem 5, we classify pairs $(G,H)$ of closed maximal connected subgroups of $Cl(V)$ such that there exists a dense $(G,H)$ -double coset in $Cl(V)$ . In order to do this, we first need to determine which semisimple groups acting tensor-decomposably on an irreducible module V have a dense orbit on $\mathcal {S}_k(V)$ . The possibilities for such subgroups are given by Proposition 2.2. They are $SO_n\otimes SO_m\leq SO_{mn}$ , $Sp_{2n}\otimes SO_m\leq Sp_{2mn}$ , $Sp_{2n}\otimes Sp_{2m}\leq SO_{4mn}$ , with $p\neq 2$ if one of the factors is an orthogonal group acting on an odd-dimensional vector space. The strategy is similar to the one employed for the proof of Theorem 2, although we only care about the connected component of (semi-)generic stabilizers. This is achieved in the following theorem.

Theorem 4. Let $G=Cl(V_1)\otimes Cl(V_2)$ be a maximal subgroup of either $SO(V)$ or $Sp(V)$ with $V=V_1\otimes V_2$ and suppose $k\leq \frac {1}{2}\dim V_1\dim V_2$ . Then the action of G on $\mathcal {S}_k(V)$ has a dense orbit if and only if $(G,k)$ is one of the following:

  1. (i) $(Sp(V_1)\otimes SO(V_2),1)$ and $(K^*G,V)$ is a prehomogeneous vector space, as classified in [Reference Zhijie32][Reference Zhijie33];

  2. (ii) $(Sp_2\otimes Sp_{2n},1)$ with $n\geq 1$ ;

  3. (iii) $(Sp_2\otimes Sp_{2n},2)$ with $n\geq 2$ ;

  4. (iv) $(Sp_2\otimes Sp_{2n},3)$ with $n\geq 3$ ;

  5. (v) $(Sp_2\otimes Sp_{2n},k)$ with $k=(2n)'$ or $k=(2n)"$ and $1\leq n\leq 3$ ;

  6. (vi) $(Sp_2\otimes Sp_{2n},2n-1)$ with $1\leq n\leq 3$ ;

  7. (vii) $(Sp_4\otimes Sp_{2n},1)$ with $n\geq 2$ .

We are now ready to classify pairs $(G,H)$ of closed maximal connected subgroups of $Cl(V)$ such that there exists a dense $(G,H)$ -double coset in $Cl(V)$ . One particular class of such pairs (see case $(iii)(b)$ in the following theorem) arises from spherical subgroups. These are reductive subgroups G such that there is a dense $(G,B)$ -double coset, where B is a Borel subgroup. Spherical subgroups have been classified by Krämer in characteristic $0$ and by Knop and Röhrle in [Reference Knop and Röhrle14] for arbitrary characteristic.

Theorem 5. Let $\Gamma $ be a classical group $Cl(V)$ . Let $G,H$ be a pair of maximal connected subgroups of $\Gamma $ . Then there exists a dense $(G,H)$ -double coset in $\Gamma $ if and only if one of the following holds:

  1. (i) G and H are both reductive and $\Gamma = GH$ . Such factorizations were classified in [Reference Liebeck, Saxl and Seitz17].

  2. (ii) G and H are both parabolic subgroups.

  3. (iii) After possibly interchanging G and H, we have that H is parabolic, G is reductive and one of the following holds:

    1. (a) $\Gamma =SL(V)$ , $H=P_k$ (or $\Gamma =Sp(V)$ , $H=P_k$ with $k=1$ ) and $(GL_k\otimes G,K^k\otimes V)$ is a prehomogeneous vector space, as classified in [Reference Sato and Kimura29][Reference Zhijie32][Reference Zhijie33];

    2. (b) G is the stabilizer of a subspace X of V and either X is nondegenerate or $p=2$ and X is nonsingular of dimension $1$ with V orthogonal. In this case G is spherical, so H is arbitrary.

    3. (c) $G=Cl(V_1)\otimes Cl(V_2)$ , $V = V_1\otimes V_2$ , and G has a dense orbit on the variety of totally singular subspaces corresponding to $\Gamma /H$ , as detailed in Theorem 4.

    4. (d) G is simple and irreducible on V and has a dense orbit on the variety of totally singular subspaces corresponding to $\Gamma /H$ , as detailed in Theorem 3.

Remark 2. Dropping the maximality assumption presents considerable challenges. Even the case where both G and H are reductive does not have a general solution, as the results in [Reference Brundan6] require a technical condition on G and H.

Remark 3. The double coset density question remains open when $\Gamma $ is an exceptional group. In case $(iii)$ of Theorem 5, the only information available for $\Gamma $ exceptional is when G is of maximal rank (see [Reference Duckworth8]). The complete classification will be the subject of forthcoming work.

Let us conclude with a related question. Given two maximal connected subgroups of $Cl(V)$ , are there finitely many double cosets? Of course, the existence of finitely many double cosets implies the existence of a dense double coset, while the opposite need not be true. For example, $G=A_{\ell }$ for $\ell \geq 8$ has a dense orbit on $X=\mathcal {G}_2(V)$ where $V=V_{G}(\lambda _2)$ , but also has infinitely many orbits ([Reference Guralnick, Liebeck, Macpherson and Seitz11]). Therefore, there is a dense $(G,P_2)$ -double coset in $SL(V)$ , as well as infinitely many double cosets.

However, it was shown in [Reference Guralnick, Liebeck, Macpherson and Seitz11] that a simple group G having finitely many orbits on $X=\mathcal {G}_1(V)$ , where V is an irreducible G-module, is equivalent to G having a dense orbit on X. This result was replicated for self-dual modules when $X=\mathcal {S}_1(V)$ and $X=\mathcal {S}_2(V)$ in [Reference Rizzoli26][Reference Rizzoli27].

Strikingly, we shall conjecture that there is only one exception to this equivalence when X is an orthogonal or symplectic Grassmannian.

Conjecture 1. Let G be a simple connected irreducible subgroup of either $SO(V)$ or $Sp(V)$ . Then unless $(G,\lambda ,p,k) = (C_3,\lambda _2,p,7)$ with $p\neq 3$ , the action of G on $\mathcal {S}_k(V)$ has a dense orbit if and only if G acts on $\mathcal {S}_k(V)$ with finitely many orbits.

The paper will be structured as follows. In Section 2, we shall consolidate the notation and background material, develop the tools for our analysis and determine the complete list of $ts$ -small quadruples, which we divide into three tables (see Proposition 2.17). We will devote a section to each table.

Given a $ts$ -small quadruple, we then proceed to determine its (semi-)generic $ts$ -stabilizer. In Section 3, we handle the cases where we already had finitely many orbits on all k-spaces. In Section 4, we deal with the cases having a large zero-weight space, and in Section 5, we handle the remaining cases. This completes the proof of Theorem 1 and Theorem 2.

We then shift our attention to the double coset density question. In Section 6, we prove Theorem 4. Finally, in Section 7, we use all of the previous results to prove Theorem 5.

2 Preliminary results

In this section, we gather some useful results, develop notation and conclude with a complete list of $ts$ -small quadruples.

2.1 Bilinear forms

We start by fixing the notation for the action of an orthogonal group on its natural module $V_{nat}$ . To do this, we use the standard notation for its root system: we take an orthonormal basis $\epsilon _1,\dots ,\epsilon _\ell $ of the $\ell $ -dimensional Euclidean space and take simple roots $\alpha _i = \epsilon _i-\epsilon _{i+1}$ for $i<\ell $ and $\alpha _\ell = \epsilon _\ell $ or $\epsilon _{\ell -1}+\epsilon _\ell $ according as $G =B_\ell $ or $D_\ell $ .

If $G=D_\ell $ , then $V_{nat}$ has (hyperbolic) basis $e_1,f_1,\dots ,e_\ell ,f_\ell $ on which root elements act by

$$ \begin{align*} x_{\epsilon_i-\epsilon_j}(t) &: e_j\mapsto e_j+te_i,\quad f_i\mapsto f_i-tf_j, \\ x_{-\epsilon_i+\epsilon_j}(t) &: e_i\mapsto e_i+te_j,\quad f_j\mapsto f_j-tf_i, \\ x_{\epsilon_i+\epsilon_j}(t) &: f_j\mapsto f_j+te_i,\quad f_i\mapsto f_i-te_j, \\ x_{-\epsilon_i-\epsilon_j}(t) &: e_j\mapsto e_j-tf_i,\quad e_i\mapsto e_i+tf_j, \end{align*} $$

while fixing the basis vectors that are not listed.

If $G=B_\ell $ , then $V_{nat}$ has (hyperbolic) basis $v_0,e_1,f_1,\dots ,e_\ell ,f_\ell $ on which root elements act by

$$ \begin{align*} x_{\epsilon_i-\epsilon_j}(t) &: e_j\mapsto e_j+te_i,\quad f_i\mapsto f_i-tf_j, \\ x_{-\epsilon_i+\epsilon_j}(t) &: e_i\mapsto e_i+te_j,\quad f_j\mapsto f_j-tf_i, \\ x_{\epsilon_i+\epsilon_j}(t) &: f_j\mapsto f_j+te_i,\quad f_i\mapsto f_i-te_j, \\ x_{-\epsilon_i-\epsilon_j}(t) &: e_j\mapsto e_j-tf_i,\quad e_i\mapsto e_i+tf_j, \\ x_{\epsilon_i}(t) &: v_0\mapsto v_0+2te_i,\quad f_i\mapsto f_i-tv_0-t^2e_i, \\ x_{-\epsilon_i}(t) &: v_0\mapsto v_0-2tf_i,\quad e_i\mapsto e_i+tv_0-t^2f_i, \end{align*} $$

while fixing the basis vectors that are not listed.

The following two results describe the structure of maximal connected subgroups of classical groups. Recall that by $Cl(V)$ , we denote a classical group with natural module V.

Theorem 2.1 [Reference Liebeck and Seitz18].

Let H be a closed connected subgroup of $G=Cl(V)$ . Then one of the following holds:

  1. (i) $H\leq Stab_G(X)$ with $X\leq V$ a proper nonzero subspace which is either totally singular or nondegenerate, or $p=2$ , $G=SO(V)$ and X is nonsingular of dimension $1$ ;

  2. (ii) $V=V_1\otimes V_2$ and H lies in a subgroup of the form $Cl(V_1)\otimes Cl(V_2)$ acting naturally on $V_1\otimes V_2$ with $\dim V_i\geq 2$ for $i=1,2$ ;

  3. (iii) H is a simple algebraic group acting irreducibly on V and $V|_H$ is tensor indecomposable.

The possibilities for the second case of Theorem 2.1 are given by the following proposition.

Proposition 2.2 [Reference Liebeck and Seitz18, Prop 2.2].

Suppose $V=V_1\otimes V_2$ and $f_i$ is a nondegenerate bilinear form on $V_i$ .

  1. (i) There is a unique nondegenerate bilinear form $f=f_1\otimes f_2$ on V such that $f(u_1\otimes u_2,v_1\otimes v_2)=f_1(u_1,v_1)f_2(u_2,v_2)$ for all $u_i,v_i\in V_i$ .

  2. (ii) f is symmetric if $f_1,f_2$ are both alternating or both symmetric, and f is alternating otherwise.

  3. (iii) f is preserved by $I(V_1)\circ I(V_2)$ acting naturally on the tensor product, where $I(V_i)$ is the stabilizer in $GL(V_i)$ of $f_i$ .

  4. (iv) If $p=2$ , then there is a unique quadratic form Q on V, with associated bilinear form f, such that $Q(v_1\otimes v_2)=0$ for all $v_i\in V_i$ and Q is preserved by $Sp(V_1)\otimes Sp(V_2)$ .

The following lemma gives the dimension of the symplectic and orthogonal Grassmannians we are acting on.

Lemma 2.3. Let V be either a symplectic or orthogonal geometry of dimension d over an algebraically closed field. Then

$$\begin{align*}\dim \mathcal{S}_k(V)= kd-\frac{3k^2+\epsilon_V k }{2},\end{align*}$$

where $\epsilon _V$ is $1$ or $-1$ according as V is orthogonal or symplectic.

Proof. If V is orthogonal with $d=2\ell $ and $k=\ell -1$ , we have $\dim \mathcal {S}_k(V)=\dim D_\ell -\dim P_{\ell -1,\ell }$ . In all other cases, the dimension is simply given by $\dim Cl(V) / P_k=\dim Cl(V)-\dim P_k$ for $Cl(V)=B_{\frac {d-1}{2}},C_{\frac {d}{2}},$ or $D_{\frac {d}{2}}$ as appropriate.

Recall that a $ts$ -small quadruple is a $4$ -tuple of the form $(G,\lambda ,p,k)$ , with $V=V_G(\lambda )$ a self-dual irreducible G-module, and $\dim G\geq \dim \mathcal {S}_k(V)$ . Lemma 2.3 gives the following dimension bound.

Lemma 2.4. Let $(G,\lambda ,p,k)$ be a $ts$ -small quadruple. Suppose that $V=V_G(\lambda )$ has dimension d. Then

$$\begin{align*}\dim G\geq kd-\frac{3k^2+\epsilon_V k }{2},\end{align*}$$

where $\epsilon _V$ is $1$ or $-1$ according as V is orthogonal or symplectic.

Proof. This follows directly from Lemma 2.3.

A self-dual module is either orthogonal or symplectic. The following lemma provides a useful criterion in odd characteristic.

Lemma 2.5 [Reference Steinberg30, Lemma 78-79].

Let G be a simple simply connected algebraic group and $V=V_G(\lambda )$ a self-dual G-module in characteristic $p\neq 2$ . Then if $Z(G)$ has no element of order $2$ , the module V is orthogonal.

Otherwise, let z be the only element of order $2$ in $Z(G)$ , except for the case $G=D_\ell $ with even $\ell $ , where z is the element of $Z(G)$ such that $G/\langle z \rangle \simeq SO_{2\ell }(k)$ . Then the module V is orthogonal if $\lambda (z) = 1$ , and symplectic otherwise.

Note that the value $\lambda (z)$ can be computed by [Reference Lübeck20, Appendix A.2]. If $p=2$ , for certain modules, we can determine whether V is symplectic or orthogonal thanks to [Reference Korhonen15]. More explicit descriptions of some of these forms can be found in [Reference Garibaldi and Nakano9] and [Reference Babic and Chernousov3].

When dealing with maximal totally singular subspaces of an orthogonal module V of dimension $2\ell $ , we need to be able to distinguish between the two $D_\ell $ -orbits on $\mathcal {S}_\ell (V)$ . The following lemma provides an easy way to do so.

Lemma 2.6. Let V be an orthogonal module of even dimension and U, W two maximal totally singular subspaces of V. Then U and W are in the same $SO(V)$ -orbit if and only if $\dim U - \dim U\cap W$ is even.

Proof. This follows quickly from [Reference Aschbacher2, 22.14(2)].

We conclude this section with a lemma concerning orthogonality of weight spaces in a self-dual irreducible module.

Lemma 2.7. Let $V = V_G(\lambda )$ be a self-dual irreducible $KG$ -module. Let $(\cdot ,\cdot )$ be an irreducible symmetric or alternating bilinear form on V preserved by G. Then the following hold:

  1. (a) For any two weights $\mu ,\nu $ such that $\mu \neq -\nu $ , the weight spaces $V_\mu $ and $V_{\nu }$ are orthogonal to each other;

  2. (b) $V_{\mu }$ is totally singular for all nonzero weights $\mu $ ;

  3. (c) $V_0$ is nondegenerate.

Proof. Let $v,u$ be two weight vectors in non-opposite weight spaces. Then there exists $t\in T$ such that $t.v = \kappa _1 v$ and $t.u = \kappa _2 u$ with $\kappa _1\neq \kappa _2^{-1}$ . Therefore, $(u,v) = \kappa _1\kappa _2 (u,v)$ , which implies $(u,v) = 0$ , proving $(a)$ .

By part $(a)$ with $\mu = \nu $ , we have that $V_\mu $ is totally-isotropic. Similarly to the previous paragraph, if V is an orthogonal module, there exists $\kappa \neq \pm 1$ such that $Q(v) = \kappa ^2 Q(v)$ , implying $Q(v) = 0$ for all $v\in V_{\mu }$ . This proves $(b)$ .

By $(a)$ , the zero weight space $V_0$ is orthogonal to all nonzero weight spaces. Therefore, any singular vector v in the radical of $V_0$ is a singular vector in the radical of V, which implies $v=0$ since the form is nondegenerate. This proves that $V_0$ is itself nondegenerate.

2.2 Clifford theory

Let $G\leq GL(V)$ be a subgroup acting completely reducibly and homogeneously on V. The following lemma shows that G must preserve a tensor product structure on V.

Lemma 2.8 [Reference Kleidman and Liebeck13, Lemma 4.4.3].

Let V be an irreducible $KG$ -module and suppose that $S\leq G$ acts completely reducibly and homogeneously on V, with $s\geq 2$ irreducible summands of dimension r. Then the following hold:

  1. (i) there is a tensor decomposition $V=V_1\otimes V_2$ (where $\dim V_1 = r,\, \dim V_2 = s$ ), such that $S\leq GL(V_1)\otimes 1$ and $C_{GL(V)}(S) = 1\otimes GL(V_2)$ ;

  2. (ii) $C_{GL(V)}(C_{GL(V)}(S))=GL(V_1)\otimes 1$ ;

  3. (iii) $N_{GL(V)}(S) = N_{GL(V_1)}(S)\otimes GL(V_2) $ ;

  4. (iv) the irreducible $KS$ -submodules of V are precisely the subspaces $V_1\otimes v$ , where $0\neq v\in V_2$ .

The following lemma shows that no cyclic extension of G can act irreducibly on V.

Lemma 2.9. Let V be an irreducible $KG$ -module and suppose that $S\leq G$ acts completely reducibly and homogeneously on V, with $s\geq 2$ irreducible summands of dimension r. Let $S\langle \tau \rangle \leq G$ be a cyclic extension of S. Then $S\langle \tau \rangle $ does not act irreducibly on V.

Proof. We apply Lemma 2.8 to get $V=V_1\otimes V_2$ with $\dim V_1 = r$ , $S\leq GL(V_1)\otimes 1$ and $\tau =\tau _1\otimes \tau _2\in GL(V_1)\otimes GL(V_2)$ . Since $\tau _2$ stabilises at least one $1$ -space of $V_2$ , by part $(iv)$ of Lemma 2.8, the element $\tau $ stabilises one irreducible $KS$ -submodule of V.

Lemma 2.10. Let V be an irreducible $KG$ -module and suppose that $S\leq G$ acts homogeneously on V as the sum of $2$ irreducible summands. Let $S\langle \tau \rangle \leq G$ be a cyclic extension of S. Then $S\langle \tau \rangle $ fixes $1$ , $2$ , or all $KS$ -submodules of V.

Proof. From the proof of Lemma 2.9, we have $V=V_1\otimes V_2$ , with $\dim V_2 =2$ , $S\leq GL(V_1)\otimes 1$ and $\tau =\tau _1\otimes \tau _2\in GL(V_1)\otimes GL(V_2)$ . Then $\tau _2$ stabilises either $1$ , $2$ or all $1$ -spaces of $V_2$ , concluding.

2.3 Generic stabilizers

In this section, we gather some of the essential lemmas that will allow us to determine (semi-)generic stabilizers. As mentioned in the introduction, a (semi-)generic stabilizer realises the minimum dimension of any stabilizer. This follows from the next two results.

Lemma 2.11 [Reference Newstead22, Lemma 3.7].

Let an algebraic group G act on a quasi-projective variety X. For any $t\in \mathbb {N}\cup \{0\}$ , the set $\{x\in X\lvert \dim G_x\geq t\}$ is closed.

Corollary 2.12. Let an algebraic group G act on an irreducible quasi-projective variety X with (semi-)generic stabilizer S. Then for all $x\in X$ , we have $\dim G_x\geq \dim S$ .

Proof. By assumption, there exists an open set Y such that all elements of Y have stabilizer isomorphic to S. By Lemma 2.11, we then find that $\dim G_x\geq \dim S$ for all $x\in X$ .

Let us consider some of the methods used in [Reference Guralnick and Lawther10]. In particular, we are interested in the localization to a subvariety approach [Reference Guralnick and Lawther10, §4.1]. Let X be a variety on which a simple algebraic group G acts. Let Y be a subvariety of X and $x\in X$ . The transporter in G of x into Y is

$$\begin{align*}\mathrm{Tran}_G(x,Y)=\{g\in G:g.x\in Y\}.\end{align*}$$

Let $\phi :G\times X\rightarrow X$ be the orbit map.

Lemma 2.13 [Reference Guralnick and Lawther10, Lemma 4.1].

For $y\in Y$ , the following hold:

  1. (i) $\dim \mathrm {Tran_G}(y,Y) = \dim \phi ^{-1} (y)$ ;

  2. (ii) $\operatorname {\mathrm {codim}} \mathrm {Tran_G}(y,Y)=\dim (\overline {G.y})-\dim (\overline {G.y\cap Y}).$

If Y is a subvariety of X, a point $y\in Y$ is called Y-exact if

$$\begin{align*}\operatorname{\mathrm{codim}} \mathrm{Tran_G}(y,Y)=\operatorname{\mathrm{codim}} Y.\end{align*}$$

Lemma 2.14 [Reference Guralnick and Lawther10, Lemma 4.3].

Let $\hat {Y}$ be a dense open subset of Y. Suppose that all points in $\hat {Y}$ are Y-exact. Then $\phi (G\times \hat {Y})$ contains a dense open subset of X.

Such a set $\hat {Y}$ is sufficiently representative of the G-action on X, which leads to the following lemma.

Lemma 2.15 [Reference Guralnick and Lawther10, Lemma 4.4].

Let $\hat {Y}$ be a dense open subset of Y. Let C be a subgroup of G containing $G_X$ . Suppose that for each $y\in \hat {Y}$ , the following is true:

  1. (i) y is Y-exact;

  2. (ii) $G_y$ is conjugate to C.

Then $C/G_X$ is the generic stabilizer in the action of G on X.

We similarly derive a criterion for proving that there does not exist a generic stabilizer.

Lemma 2.16. Assume that Y is not finite, and let $\hat {Y}$ be a dense open subset of Y. Suppose that for each $y\in \hat {Y}$ , the following is true:

  1. (i) y is Y-exact;

  2. (ii) for all $y'\in \hat {Y}\setminus \{y\}$ , we have that $G_y$ is not conjugate to $G_{y'}$ .

Then there is no generic stabilizer in the action of G on X.

Proof. By Lemma 2.14, there is a dense open subset $U_1$ of X, contained in $\phi (G\times \hat {Y})$ which by assumption is the union of G-orbits with pairwise non-conjugate stabilizers. Given any such orbit $G.y$ for some $y\in \hat {Y}$ , we have

$$\begin{align*}\dim(\overline{G.y}) =\dim(\overline{G.y})-\dim (\overline{G.y\cap Y}) = \operatorname{\mathrm{codim}} Y <\dim X, \end{align*}$$

since y is Y-exact combined with $\dim Y\geq 1$ and Lemma 2.13.

Assume that there is a generic stabilizer in the action of G on X. Then there is a dense open subset $U_2$ of X such that $G_{x_1}$ and $G_{x_2}$ are conjugate for all $x_1,x_2\in U_2$ . Taking the intersection of $U_1$ and $U_2$ , we get an open dense subset U of X with the same property. Therefore, given any two $x_1,x_2\in U$ , we must have $G.x_1 = G.x_2$ (i.e., U consists of a single G-orbit). This implies that there is a dense orbit, contradicting $\dim (\overline {G.y}) < \dim X$ .

2.4 Spin modules

We set up spin modules following [Reference Popov23]. Let $\{e_1,\dots ,e_n,e_{n+1},\dots , e_{2n}\}=\{e_1,\dots ,e_n,f_{1},\dots , f_{n}\}$ be a standard basis for the $2n$ -dimensional K-vector space $V=V_{2n}$ with quadratic form Q and bilinear form $(\cdot ,\cdot )$ , such that $\{e_i,e_{n+i}\}=\{ e_i,f_i\}$ are hyperbolic pairs for $i\leq n$ . Let $L,M$ be the totally singular subspaces $\langle e_1,\dots ,e_n\rangle $ and $\langle f_1,\dots ,f_n\rangle $ , respectively.

We denote by C the Clifford algebra of $(V,Q)$ . This is an associative algebra over K generated by V, in which $v^2=Q(v)$ for every $v\in V$ . It has the structure of a graded module over K. Let $\phi ':C\rightarrow C $ , sending $x\mapsto x'$ , be the involution of C fixing every element of V (i.e., the anti-automorphism sending a product $\prod _{i=1}^k v_i\in C$ to $ \prod _{i=1}^k v_{n-i+1}$ ). We denote by $C^+$ and $C^-$ the sums of homogeneous submodules of C of even and odd degrees, respectively. Then $C=C^+\oplus C^-$ . In particular, $C^+$ is a subalgebra of C invariant under $\phi '$ .

The Clifford group is $G^*=\{s\in C|s$ is invertible in C and $sVs^{-1}=V\}$ . The even Clifford group is $(G^*)^+=G^*\cap C^+$ . The spin group $Spin_{2n}$ is $\{s\in (G^*)^+|ss'=1\}$ .

The vector representation of the Clifford group $G^*$ is given by $\Theta : G^*\rightarrow Aut(V,Q)$ , such that $\Theta (s)\cdot v=svs^{-1}$ . The restriction of $\Theta $ to $Spin_{2n}$ is the natural representation of $Spin_{2n}$ . The root subgroups of $Spin_{2n}$ are parametrised by pairs $(i,j)$ with $i+j\neq 2n+1$ ; the root subgroup parametrised by the pair $(i,j)$ consists of elements of the form $1+\lambda e_ie_j$ , where $1+\lambda e_ie_j$ acts on a vector $v\in V$ by $v\mapsto v+\lambda (e_j,v)e_i-\lambda (e_i,v)e_j$ .

Put $e_L=e_1e_2\dots e_n$ and $e_M=e_{n+1}e_{n+2}\dots e_{2n}$ . We denote by $C_W$ the subalgebra of C generated by the elements of a subspace $W\subset V_{2n}$ . Then $ce:M$ is a minimal left ideal in C, and the correspondence $x\mapsto xe_M$ generates an isomorphism $C_L\rightarrow ce:M$ of vector spaces. So for any $s\in C,x\in C_L$ , there exists a unique element $y\in C_L$ for which $sxe_M=ye_M$ . Setting $\rho (s)\cdot x = s\cdot x=y$ gives us the spinor representation $\rho $ of the algebra C in $C_L$ . Let $X=C_L\cap C^+$ . Then restricting $\rho $ to $Spin_{2n}$ , we get the half-spin representation of $Spin_{2n}$ in X.

An element of X is called a spinor. The restriction to $B_{n-1}$ is the spin representation for $B_{n-1}$ .

Table 2 $ts$ -small quadruples with finitely many orbits on $\boldsymbol{\mathcal {G}_k(V)}$ .

2.5 List of ts-small quadruples

The following result lists all $ts$ -small quadruples. We will then be able to prove Theorem 1 and Theorem 2 by proceeding case-by-case.

Proposition 2.17. Let $(G,\lambda ,p,k)$ be a $ts$ -small quadruple and $V = V_G(\lambda )$ . Then, precisely one of the following is true:

  1. (i) G has finitely many orbits on $\mathcal {G}_k(V)$ , and $(G,\lambda ,p,k)$ is in Table 2;

  2. (ii) $k=1$ , and either V is a composition factor of $\mathrm {Lie}(G)$ or $G=C_\ell $ and $\lambda = \lambda _2$ , as in Table 3;

  3. (iii) $(G,\lambda ,p,k)$ is in Table 4.

Table 3 Infinite families of $ts$ -small quadruples.

Table 4 Remaining $ts$ -small quadruples.

Proof. In [Reference Guralnick, Liebeck, Macpherson and Seitz11], we find a complete list of modules with finitely many orbits on k-spaces, of which every self-dual one leads to a $ts$ -small quadruple. Now assume that $(G,\lambda ,p,k)$ is a $ts$ -small quadruple where G does not have finitely many orbits on $\mathcal {G}_k(V)$ . In [Reference Rizzoli26, Thm 3.1], we have a complete list of such quadruples with $k=1$ and V orthogonal, while in [Reference Rizzoli27, Prop. 4.1], we have a complete list of such quadruples for $k\geq 2$ .

The proof follows from combining the three lists of $ts$ -small quadruples, with the addition of the self-dual small quadruples where V is symplectic and $k=1$ .

3 Quadruples with finitely many orbits on $\mathcal {G}_k(V)$

In this section, we handle the $ts$ -small quadruples where we already have finitely many orbits on all k-spaces (i.e., the ones listed in Table 2). It follows directly that there is a dense orbit for the action on totally singular k-spaces, and producing the generic stabilizer reduces to finding a stabilizer of appropriate dimension.

Proposition 3.1. Let G be one of $A_1$ , $B_\ell $ with $p\neq 2$ , $C_\ell $ or $D_\ell $ . Let $\lambda =\lambda _1$ . Then the $ts$ -small quadruple $(G,\lambda ,p,k)$ has generic $ts$ -stabilizer $P_k$ , unless $k=\ell -1$ and $G=D_\ell $ , in which case it has generic $ts$ -stabilizer $P_{\ell -1,\ell }$ .

Proof. Note that if $G=D_\ell $ , by our convention, the values of k are $1,\dots ,\ell -1,\ell ',\ell "$ . In all cases, in the statement of the proposition, the group G is transitive on $\mathcal {S}_k(V)$ . It is well known that the maximal parabolic subgroups of a classical group are stabilizers of totally singular subspaces; therefore, these must be the generic stabilizers. Unless we are in type D and $k=\ell -1$ , the stabilizer of $y\in \mathcal {S}_k(V)$ is a conjugate of $P_k$ . If $G=D_\ell $ and $k=\ell -1$ , the stabilizer of $y\in \mathcal {S}_k(V)$ is a conjugate of $P_{\ell -1,\ell }$ .

Proposition 3.2. The generic $ts$ -stabilizers for the $ts$ -small quadruples in Table 5 are as given.

Table 5 $ts$ -small quadruples with $k=1$ and V symplectic.

Proof. In each of these cases, $k=1$ and the module V is symplectic. Therefore, $\mathcal {S}_k(V)=\mathcal {G}_k(V)$ , and the result follows directly from [Reference Guralnick and Lawther10]. In the last column of Table 5, we give a reference for each individual case.

Proposition 3.3. Let $G=A_1$ and $\lambda = \lambda _1+p^i\lambda _1$ with $p>0$ and $i>0$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $T_1$ .

Proof. Let $q=p^i$ and let $\sigma =\sigma _q$ be the standard Frobenius morphism acting on K as $t\mapsto t^\sigma =t^q$ and on G as $x_{\pm \alpha _1}(t)\mapsto x_{\pm \alpha _1}(t^q)$ . Let $G=SL_2(K)$ . We can view V as the space $M_{2\times 2}(K)$ of $2\times 2$ matrices on which G acts by $g.v=gv(g^\sigma )^T$ for $v\in M_{2\times 2}(K)$ and $g\in G$ . Since G preserves the determinant of v for all $v\in M_{2\times 2}(K)$ , we can take the quadratic form $Q:V\rightarrow K$ as $Q(v)=\det v$ . The singular $1$ -spaces of V are therefore the $1$ -spaces spanned by matrices with determinant $0$ . Let y be the singular $1$ -space spanned by $\left (\begin {matrix} 0 & 1\\ 0 & 0 \\ \end {matrix}\right )$ . Then $\left (\begin {matrix} a & b\\ c & d \\ \end {matrix}\right )\in G_y$ if and only if $b=c=0$ and $d=a^{-1}$ . Since $\dim G - \dim \mathcal {S}_1(V) = 1 =\dim G_y$ , the element y is in a dense orbit for the G-action on $\mathcal {S}_1(V)$ . Therefore, $C_{\mathcal {S}_1(V)} = T_1$ .

Proposition 3.4. Let $G=A_2$ and $\lambda = \lambda _1+\lambda _2$ with $p=3$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_2T_1$ .

Proof. By [Reference Rizzoli27, Lemma 5.5], there is $y\in \mathcal {S}_1(V)$ with $G_y=U_2T_1$ . Since $\dim G - \dim \mathcal {S}_1(V)=8-5 =\dim G_y$ , the element y is in a dense orbit for the G-action on $\mathcal {S}_1(V)$ . Therefore, $C_{\mathcal {S}_1(V)} = U_2T_1$ .

Proposition 3.5. Let $G=A_5$ and $\lambda = \lambda _3$ with $p=2$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_8A_2T_1$ .

Proof. By [Reference Revoy24, 2.3.1(II)], there is $y\in \mathcal {S}_1(V)$ with $G_y=U_8A_2T_1$ . Since $\dim G - \dim \mathcal {S}_1(V)=35-18 =\dim G_y$ , the element y is in a dense orbit for the G-action on $\mathcal {S}_1(V)$ . Therefore, $C_{\mathcal {S}_1(V)} = U_8A_2T_1$ .

Proposition 3.6. Let $G=B_3$ , $\lambda = \lambda _3$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_6A_2T_1$ .

Proof. By [Reference Liebeck, Saxl and Seitz17, Thm B], the group G is transitive on $\mathcal {S}_1(V)$ . The generic stabilizer is the $P_3$ -parabolic (i.e. $C_{\mathcal {S}_1(V)} = U_6A_2T_1$ ).

Proposition 3.7. Let $G=B_3$ , $\lambda = \lambda _3$ with $k=2$ or $k=3$ . Then the quadruple $(G,\lambda ,p,k)$ has generic $ts$ -stabilizer $U_5A_1A_1T_1$ if $k=2$ and $U_3A_2T_1$ if $k=3$ .

Proof. By [Reference Liebeck, Saxl and Seitz17, Thm B], the group G is transitive on $\mathcal {S}_4'(V)$ . Let $W\in \mathcal {S}_4'(V)$ . The group G is the group of fixed points of a triality automorphism of $D_4 = Cl(V)$ . Therefore, $G_W$ is isomorphic to the generic stabilizer for the action on $\mathcal {S}_1(V)$ , i.e. $U_6A_2T_1$ . Then it is easy to see ([Reference Guralnick, Liebeck, Macpherson and Seitz11, Lemma 3.5]) that $G_W$ acts on $\mathcal {G}_2(W)$ with two orbits, one with stabilizer $U_5A_1A_1T_1$ and one with stabilizer $U_7A_1A_1T_1$ . Since every totally singular $2$ -space is contained in an element of $\mathcal {S}_4'(V)$ , we conclude that there are at most two G-orbits on $\mathcal {S}_2(V)$ . Since $\dim G-\dim \mathcal {S}_2(V)=12$ , there must be a $12$ -dimensional stabilizer for the G-action on $\mathcal {S}_2(V)$ . The only possibility is therefore $C_{\mathcal {S}_2(V)} = U_5A_1A_1T_1$ .

Similarly, $G_W$ acts on $\mathcal {G}_3(W)$ with two orbits, one with stabilizer $U_3A_2T_1$ and one with stabilizer $U_8A_1T_2$ . Since $\dim G-\dim U_3A_2T_1 = \dim \mathcal {S}_3(V)$ , we conclude that $C_{\mathcal {S}_3(V)} = U_3A_2T_1$ .

Proposition 3.8. Let $G=B_\ell $ , $\lambda = \lambda _\ell $ with $\ell =4$ or $\ell =5$ , with $p=2$ if $\ell =5$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_7G_2T_1$ if $\ell =4$ and $U_{14}B_2T_1$ if $\ell =5$ .

Proof. By [Reference Igusa12, Prop. 5, Prop. 6] (when $p\neq 2$ ) and [Reference Guralnick, Liebeck, Macpherson and Seitz11, Lemma 2.11] (when $p=2$ ), there is only one orbit on $\mathcal {G}_1(V)$ with a $22$ -dimensional stabilizer when $\ell =4$ , with structure $U_7G_2T_1$ , and only one orbit on $\mathcal {G}_1(V)$ with $25$ -dimensional stabilizer when $\ell =5$ , with structure $U_{14}B_2T_1$ . Since $\dim G-\dim \mathcal {S}_1(V) = 22$ when $\ell =4$ and $25$ when $\ell =4$ , we conclude that $C_{\mathcal {S}_1(V)} = U_7G_2T_1$ when $\ell =4$ and $U_{14}B_2T_1$ when $\ell =5$ .

Proposition 3.9. Let $G=C_3$ , $\lambda = \lambda _2$ with $p=3$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_6A_1T_1$ .

Proof. By [Reference Rizzoli27, Lemma 5.15], there is $y\in \mathcal {S}_1(V)$ with $G_y=U_6A_1T_1$ . Since $\dim G - \dim \mathcal {S}_1(V)=21-11 =\dim G_y$ , the element y is in a dense orbit for the G-action on $\mathcal {S}_1(V)$ . Therefore, $C_{\mathcal {S}_1(V)} = U_6A_1T_1$ .

Proposition 3.10. Let $G=D_6$ , $\lambda = \lambda _6$ with $p=2$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_{14}B_3T_1$ .

Proof. By the proof of [Reference Guralnick, Liebeck, Macpherson and Seitz11, Lemma 2.11], there is only one orbit on $\mathcal {G}_1(V)$ with a $36$ -dimensional stabilizer, with structure $U_{14}B_3T_1$ . Since $\dim G - \dim \mathcal {S}_1(V)=66-30 =\dim G_y$ , the element y is in a dense orbit for the G-action on $\mathcal {S}_1(V)$ . Therefore, $C_{\mathcal {S}_1(V)} = U_{14}B_3T_1$ .

Proposition 3.11. Let $G=G_2$ , $\lambda = \lambda _1$ with $p\neq 2$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_{5}A_1T_1$ .

Proof. By [Reference Liebeck, Saxl and Seitz17, Thm. A], G is transitive on singular $1$ -spaces of V. Therefore, a representative can be taken to be the $1$ -space spanned by the highest weight vector, with stabilizer $P_1=U_{5}A_1T_1$ .

Proposition 3.12. Let $G=G_2$ , $\lambda = \lambda _1$ . Then the quadruple $(G,\lambda ,p,2)$ has generic $ts$ -stabilizer $U_{3}A_1T_1$ .

Proof. Let $\mu _1,\mu _2,\mu _3$ be the positive weights $2\alpha _1+\alpha _2,\alpha _1+\alpha _2,\alpha _1$ , respectively. Given weight vectors $v_{\mu _1},v_{-\mu _2} $ , let $y:=\langle v_{\mu _1},v_{-\mu _2}\rangle $ , an element of $\mathcal {S}_2(V)$ . Since $n_2n_1n_2n_1n_2.\langle v_{\mu _1} \rangle = \langle v_{-\mu _2}\rangle $ , we can easily determine that $G_{\langle v_{\mu _1}\rangle } \cap G_{\langle v_{-\mu _2}\rangle } = P_1\cap P_1^{n_2n_1n_2n_1n_2} = U_3T_2$ , by checking which root subgroups are in common between $P_1$ and $P_1^{n_2n_1n_2n_1n_2}$ . Therefore, $G_y\leq U_3A_1T_1$ . Since $\dim \mathcal {S}_2(V) = 7$ (in both cases $p\neq 2$ and $p=2$ ), the minimum dimension of the stabilizer of any totally singular $2$ -space is $7$ . Therefore, $G_y = U_3A_1T_1 = C_{\mathcal {S}_2(V)}$ .

Proposition 3.13. Let $G=G_2$ , $\lambda = \lambda _1$ with $p=2$ . Then the quadruple $(G,\lambda ,p,3)$ has generic $ts$ -stabilizer $A_2$ .

Proof. By the proof of [Reference Guralnick, Liebeck, Macpherson and Seitz11, Lemma 3.4], there is an $A_2$ -subgroup which is the stabilizer of $y\in \mathcal {S}_3(V)$ . Since $\dim G - \dim \mathcal {S}_3(V)=14-6 =\dim G_y$ , the element y is in a dense orbit for the G-action on $\mathcal {S}_3(V)$ . Therefore, $C_{\mathcal {S}_3(V)} = A_2$ .

Proposition 3.14. Let $G=F_4$ , $\lambda = \lambda _4$ with $p=3$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_{14}G_2T_1$ .

Proof. By [Reference Brundan6, Lemma 4.13], there is $y\in \mathcal {S}_1(V)$ such that $G_y=U_{14}G_2T_1$ . Since $\dim G - \dim \mathcal {S}_1(V)=29 =\dim G_y$ , the element y is in a dense orbit for the G-action on $\mathcal {S}_1(V)$ . Therefore, $C_{\mathcal {S}_1(V)} = U_{14}G_2T_1$ .

Proposition 3.15. Let $G=E_7$ , $\lambda = \lambda _7$ with $p=2$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $U_{26}F_4T_1$ .

Proof. By [Reference Liebeck and Saxl16, Lemma 4.3], there is only one orbit on $\mathcal {G}_1(V)$ with a $79$ -dimensional stabilizer, with structure $U_{26}F_4T_1$ . Since $\dim G - \dim \mathcal {S}_1(V)=133-54 =\dim U_{26}F_4T_1$ , we must have $C_{\mathcal {S}_1(V)} = U_{26}F_4T_1$ .

4 Infinite families of quadruples

In this section, we handle the cases that appear in Table 3. The following two propositions provide a reduction to a finite list of cases.

Proposition 4.1. Let V be a composition factor of $\mathrm {Lie}(G)$ , with $p\neq 2$ if $G=B_\ell $ or $G=C_\ell $ , and assume that the $0$ -weight space $V_0$ is at least $3$ -dimensional. Then $C_{\mathcal {S}_1(V)}=C_{\mathcal {G}_1(V)}$ .

Proof. The composition factors of $\mathrm {Lie}(G)$ are listed in [Reference Liebeck and Seitz19, Prop. 1.10]. By assumption on V, we have $V=\mathrm {Lie}(G)/Z$ , where Z is the centre of $\mathrm {Lie}(G)$ . If $p=2$ , and $G=A_\ell $ with $\ell \equiv 1\ \mod 4$ , or $G=D_\ell $ with $\ell \equiv 2\ \mod 4$ , or $G=E_7$ , then the module V is symplectic and $\mathcal {S}_1(V) = \mathcal {G}_1(V)$ . In all other cases, the module V is orthogonal. The proof closely mimics [Reference Guralnick and Lawther10, Lemma 4.5(ii)], but in the interest of clarity, it is fully reproduced with the appropriate changes. Note that our setup corresponds to the specific case $\theta = 1$ in [Reference Guralnick and Lawther10, Lemma 4.5(ii)], which in particular means $G=H$ in the proof of [Reference Guralnick and Lawther10, Lemma 4.5(ii)]. Let

$$\begin{align*}W^\ddagger = \{ w\in W:\exists \xi\in K^*,\forall v\in V_0, w.v=\xi v\},\end{align*}$$

and let $N^\ddagger $ be the pre-image of $W^\ddagger $ under the quotient map $N\rightarrow W$ . Let Y be $\mathcal {S}_1(V_0)$ and let $\hat {Y}_1$ be the subset of Y consisting of $1$ -spaces of $V_0$ spanned by images of regular semisimple elements in $\mathrm {Lie}(T)$ . Since Z does not contain regular semisimple elements, the set $\hat {Y}_1$ is nonempty, and thus dense in Y. Since $\dim V_0\geq 3$ , the span of Y is the full $V_0$ . Therefore, any element in N which fixes all $y\in Y$ must be in $N^\ddagger $ . Thus, given $w\in W\setminus W^\ddagger $ , take $n\in N$ with $nT=w$ ; the set of elements of Y fixed by n is a proper closed subvariety of Y. Let $\hat {Y}_2$ be the complement of the union of these subvarieties as w runs over $W\setminus W^\ddagger $ . Set $\hat {Y} = \hat {Y}_1\cap \hat {Y}_2$ , a dense open subset of Y.

Let $y\in \hat {Y}$ . By [Reference Guralnick, Liebeck, Macpherson and Seitz11, Lemma 2.1], two elements of $\mathcal {G}_1(V_0)$ are in the same G-orbit if and only if they are in the same W-orbit. Therefore, $G.y\cap Y$ is finite and $\dim (\overline {G.y\cap Y})=0$ . Also, since y is spanned by a regular semisimple element, we have $G_y^0 = T$ , and therefore, $\dim (\overline {G.y})=\dim G - \dim T $ . Finally, note that since $V = \mathrm {Lie(G)}/Z$ , we have $(C_G(V_0))^0 = T$ , and therefore, $\dim V-\dim V_0=\dim G-\dim T$ (see the proof of [Reference Guralnick, Liebeck, Macpherson and Seitz11, Lemma 2.4]). Thus,

$$ \begin{align*} \dim \mathcal{S}_1(V)-\dim (\overline{G.y}) &= \dim V- 2 -\dim (\overline{G.y}) =\\ &= \dim G-\dim T+\dim V_0 -2 -\dim (\overline{G.y}) =\\ &= \dim Y - \dim(\overline{G.y\cap Y}) = \dim Y + \operatorname{\mathrm{codim}} \mathrm{Tran}_G(y,Y)-\dim (\overline{G.y}),\end{align*} $$

where the last step uses Lemma 2.13. By definition, this proves that y is Y-exact. The conditions of Lemma 2.15 hold and $C_{\mathcal {S}_1(V)} = T.W^\ddagger = C_{\mathcal {G}_1(V)}$ .

Proposition 4.2. Let $G=C_\ell $ , $\lambda =\lambda _2$ . Furthermore, assume that the $0$ -weight space $V_0$ is at least $3$ -dimensional. Then $C_{\mathcal {S}_1}(V)=C_{\mathcal {G}_1}(V) = A_1^\ell $ .

Proof. If $p=2$ and $\ell \equiv 2\ \mod 4$ , then the module V is symplectic and $\mathcal {S}_1(V)=\mathcal {G}_1(V)$ . Therefore, assume that when $p=2$ , we have $\ell \not \equiv 2\ \mod 4$ , which implies that the module V is orthogonal. We use the setup of [Reference Guralnick and Lawther10, Prop. 5.13] combined with the approach of [Reference Guralnick and Lawther10, Lemma 4.5(ii)], which we saw in action in Proposition 4.1. Inside $\bigwedge ^2 V_{nat}$ , we have submodules $X_1 = \{ \sum _{i<j}\rho _{ij}e_i\wedge e_j+\sum _{i<j}\sigma _{ij}f_i\wedge f_j+\sum _{i,j}\tau _{ij}e_i\wedge f_j:\sum _i\tau _{ii}=0\}$ and $X_2=\langle \sum _i e_i\wedge f_i \rangle $ . If $p\nmid \ell $ then $V=X_1$ , otherwise $X_2<X_1$ and $V=X_1/X_2$ . In all cases, $V=X_1/(X_1\cap X_2)$ . Let $x_i = e_i\wedge f_i$ . The $0$ -weight space is $V_0 = \{ \sum a_ix_i+(X_1\cap X_2):\sum a_i = 0\}$ . Then G fixes a nondegenerate quadratic form on V such that $Q( a_ix_i+(X_1\cap X_2)) = \sum a_i^2 +\sum _{i<j}a_ia_j$ (see [Reference Garibaldi and Nakano9, p. 8.1.2] when $p=2$ ). Let $Y = \mathcal {S}_1(V_0)$ and let

$$\begin{align*}\hat{Y}_1 = \left\{ \langle v \rangle\in Y , v=\sum a_ix_i+(X_1\cap X_2): a_i\neq a_j\text{ if }i\neq j\right\},\end{align*}$$

a dense subset of Y. Then $y\in \hat {Y}_1$ is fixed by $C=A_1^\ell = \bigcap _{i} G_{\langle e_i,f_i \rangle }$ . Any minimal connected overgroup of C in G is isomorphic to $C_2 A_1^{\ell -2}$ , which does not fix $y\in \hat {Y}_1$ because of the condition on the coefficients. Therefore, for any $y\in \hat {Y}_1$ , we have $(G_y)^0 = C$ . Let $N = N_G(C)/C$ , a group isomorphic to $Sym(\ell )$ . Let

$$\begin{align*}N^\ddagger = \{ n\in N:\exists \xi\in K^*,\forall v\in V_0, n.v=\xi v\}.\end{align*}$$

Now assume that $n\in N$ fixes all $y\in Y$ . Since $\dim V_0\geq 3$ , the span of Y is the full $V_0$ . Therefore, any element in N which fixes all $y\in Y$ must be in $N^\ddagger $ . Thus, the set of elements fixed by $n\in N\setminus N^\ddagger $ is a proper closed subvariety of Y. Let $\hat {Y}_2$ be the complement of the union of these subvarieties as n runs over $N\setminus N^\ddagger $ . Set $\hat {Y} = \hat {Y}_1\cap \hat {Y}_2$ , a dense open subset of Y. Finally, note that the proof of [Reference Guralnick and Lawther10, Prop. 5.13] shows that $N^\ddagger = 1$ . Let $y\in \hat {Y}$ . Two elements of $\mathcal {G}_1(V_0)$ are in the same G-orbit if and only if they are in the same N-orbit, and therefore, $\dim (\overline {G.y\cap Y})=0$ . Also, $\dim V-\dim V_0=\dim G-\dim C$ , by the proof of [Reference Guralnick, Liebeck, Macpherson and Seitz11, Lemma 2.4]. Therefore, as in Proposition 4.1, we get

$$ \begin{align*} \dim \mathcal{S}_1(V)-\dim (\overline{G.y})&= \dim V- 2 -\dim (\overline{G.y}) =\\ &= \dim G-\dim C+\dim V_0 -2 -\dim (\overline{G.y}) = \\ &=\dim Y - \dim(\overline{G.y\cap Y}),\end{align*} $$

proving that y is Y-exact. The conditions of Lemma 2.15 hold and $C_{\mathcal {S}_1(V)} = C.N^\ddagger = C = A_1^\ell $ .

Proposition 4.3. Let $G=B_2$ and $\lambda = 2\lambda _2$ with $p\neq 2$ , or $G=A_2$ and $\lambda = \lambda _1+\lambda _2$ with $p\neq 3$ , or $G=A_3$ and $\lambda _1+\lambda _3$ with $p=2$ , or $G=G_2$ and $\lambda = \lambda _2$ with $p\neq 3$ . Then $C_{\mathcal {S}_1(V)}$ is respectively $T_2.\mathbb {Z}_4$ , $T_2.\mathbb {Z}_3$ , $T_3.Alt(4)$ , $T_2.\mathbb {Z}_6$ .

Proof. In all of these cases, the module V is a composition factor of $\mathrm {Lie}(G)$ and the $0$ -weight space is $2$ -dimensional, and we will show that the generic stabilizer is the stabilizer of one of the two singular $1$ -spaces of $V_0$ .

Let $G=B_2$ and $\lambda = 2\lambda _2$ with $p\neq 2$ . Then $V=\mathrm {Lie}(G)$ , and we can take $V_0 = \{\mathrm {diag}(a,b,0,-b,-a) :a,b\in K\}$ . Since $p\neq 2$ , the group G fixes the nondegenerate quadratic form Q induced by the Killing form. Let $v=\mathrm {diag}(a,b,0,-b,-a)$ be a singular element of $V_0$ . Then since $Q(v)=0$ , we know that $a^2+b^2=0$ . Since v is regular semisimple, we must have $(G_{\langle v\rangle })^0 = T_2$ . We then find that $W_{\langle v\rangle }=\langle w\rangle $ , where w is an element of order $4$ sending $\mathrm {diag}(a,b,0,-b,-a)\mapsto \mathrm {diag}(b,-a,0,a,-b) $ . Since $\dim G-\dim \mathcal {S}_1(V) = 2 =\dim G_{\langle v\rangle }$ , we conclude that $C_{\mathcal {S}_1(V)} = T_2.\mathbb {Z}_4$ .

Let $G=A_2$ and $\lambda =\lambda _1+\lambda _2$ with $p\neq 3$ . Then $V=\mathrm {Lie}(G)$ , and we can take $V_0 = \{\mathrm {diag}(a,b,-a-b):a,b\in K\}$ . If $p\neq 2$ , a nondegenerate symmetric bilinear form preserved by G is given by the Killing form. If $p=2$ , we find an explicit description of a nondegenerate quadratic form preserved by G in [Reference Babic and Chernousov3, §5.1]. Let $v=\mathrm {diag}(a,b,-a-b)$ be a singular element of $V_0$ , which implies that $a^2+b^2+ab = 0$ . Since v is regular semisimple, we must have $(G_{\langle v\rangle })^0 = T_2$ . As $a^2+b^2+ab = 0$ , we then find that $W_{\langle v\rangle }=\langle w\rangle $ , where w is a $3$ -cycle in W. Since $\dim G-\dim \mathcal {S}_1(V) = 2 =\dim G_{\langle v\rangle }$ , we conclude that $C_{\mathcal {S}_1(V)} = T_2.\mathbb {Z}_3$ .

Let $G=A_3$ and $\lambda =\lambda _1+\lambda _3$ with $p=2$ . Then $V=\mathfrak {sl}_4/\langle I\rangle $ , where I is the identity $4\times 4$ matrix, and we can take $V_0 = \{\mathrm {diag}(a,b,a+b,0)+\langle I\rangle :a,b\in K\}$ . Let $v=\mathrm {diag}(a,b,a+b,0)+\langle I \rangle $ be a singular element of $V_0$ , which implies that $a^2+b^2+ab = 0$ (again see [Reference Babic and Chernousov3, §5.1]). Since $\mathrm {diag}(a,b,a+b,0)$ is regular semisimple, we must have $(G_{\langle v\rangle })^0 = T_3$ . As $a^2+b^2+ab = 0$ , we then find that $W_{\langle v\rangle }=\langle \tau _1,\tau _2,w\rangle $ , where w is a $3$ -cycle in W sending $\mathrm {diag}(a,b,a+b,0)\mapsto \mathrm {diag}(a+b,a,b,0)$ , $\tau _1$ is an element of order $2$ sending $\mathrm {diag}(a,b,a+b,0)\mapsto \mathrm {diag}(0,a+b,b,a)$ and $\tau _2$ is an element of order $2$ sending $\mathrm {diag}(a,b,a+b,0)\mapsto \mathrm {diag}(a+b,0,a,b)$ . Since $\dim G-\dim \mathcal {S}_1(V) = 3 =\dim G_{\langle v\rangle }$ , we conclude that $C_{\mathcal {S}_1(V)} = T_3.Alt(4)$ .

Let $G=G_2$ and $\lambda = \lambda _2$ with $p\neq 3$ . Then $V=\mathrm {Lie}(G)$ . We view G as a subgroup of $B_3$ , so that we can take $V_0 =\{ \mathrm {diag}(a,b,-a-b,0,-a,-b,a+b):a,b\in K\}$ . Let $v=\mathrm {diag}(a,b,-a-b,0,-a,-b,a+b)$ be a singular element of $V_0$ , which implies that $a^2+b^2+ab = 0$ (see [Reference Babic and Chernousov3, §5.1]). Since v is regular semisimple, we must have $(G_{\langle v\rangle })^0 = T_2$ . As $a^2+b^2+ab = 0$ , we then find that $W_{\langle v\rangle }=\langle \tau , w\rangle $ , where $\tau $ is an element of order $2$ sending $v\mapsto -v$ and w is an element of order $3$ sending $\mathrm {diag}(a,b,-a-b,0,a+b,-b,-a)\mapsto \mathrm {diag}(-a-b,a,b,0,-b,-a,a+b)$ . Since $\dim G-\dim \mathcal {S}_1(V) = 2 =\dim G_{\langle v\rangle }$ , we conclude that $C_{\mathcal {S}_1(V)} = T_2.\mathbb {Z}_6$ .

Proposition 4.4. Let $G=C_3$ and $\lambda = \lambda _2$ with $p\neq 3$ , or $G=C_4$ and $\lambda = \lambda _2$ with $p=2$ , or $G=D_4$ and $\lambda =\lambda _2$ with $p=2$ . Then $C_{\mathcal {S}_1(V)}$ is respectively $A_1^3.\mathbb {Z}_3$ , $A_1^4.Alt(4)$ , $T_4.(2^3.Alt(4))$ .

Proof. In all of these cases, the module V is a composition factor of $\bigwedge ^2 V_{nat}$ , the $0$ -weight space is $2$ -dimensional, and the generic stabilizer is the stabilizer of one of the two singular $1$ -spaces of $V_0$ . Let $X_1,X_2$ be as in the proof of Proposition 4.2, so that $V = X_1/X_2$ .

Let $G=C_3$ and $\lambda = \lambda _2$ with $p\neq 3$ . Then $X_2 =0$ , and we can take

$$\begin{align*}V_0 = \{a e_1\wedge f_1+b e_2\wedge f_2 -(a+b)e_3\wedge f_3 :a,b,\in K\}.\end{align*}$$

Let $v=a e_1\wedge f_1+b e_2\wedge f_2 -(a+b)e_3\wedge f_3$ be a singular element of $V_0$ . Then since $Q(v)=0$ , we know that $a^2+b^2+ab=0$ . Since $a\neq b$ , we must have $(G_{\langle v\rangle })^0 = A_1^3$ as in the proof of Proposition 4.2. We then find that $W_{\langle v\rangle }=\langle w\rangle $ , where w is a $3$ -cycle in W. Since $\dim G-\dim \mathcal {S}_1(V) = 9 =\dim G_{\langle v\rangle }$ , we conclude that $C_{\mathcal {S}_1(V)} = A_1^3.\mathbb {Z}_3$ .

The remaining two cases are entirely similar, with the result for $D_4$ being derived from $C_4$ , since $V_{D_4}(\lambda _2) = V_{C_4}(\lambda _2)\downarrow D_4$ when $p=2$ .

5 Remaining quadruples

This section is where we deal with the remaining cases (i.e., all the possibilities listed in Table 4).

Proposition 5.1. The generic $ts$ -stabilizers for the $ts$ -small quadruples in Table 6 are as given.

Table 6 Remaining $ts$ -small quadruples with known generic $ts$ -stabilizer.

Proof. In each of these cases, except for $(B_6,\lambda _6,p ,1)$ with $p\neq 2$ , it was proven in [Reference Rizzoli26] or [Reference Rizzoli27] that G has a dense orbit on $\mathcal {S}_k(V)$ . The structure of the generic stabilizer is as given in [Reference Rizzoli26][Reference Rizzoli27], with the appropriate reference listed in Table 6. Note that for the case $(B_6,\lambda _6,2,1)$ , Proposition 5.2 clarifies the lackluster explanation found in [Reference Rizzoli26].

Finally, in the case $(B_6,\lambda _6,p,1)$ with $p\neq 2$ , the module V is symplectic, and we have $\mathcal {S}_k(V) = \mathcal {G}_k(V)$ , with the result following from [Reference Guralnick and Lawther10].

Proposition 5.2. Let $G= B_6$ , $\lambda = \lambda _6$ with $p= 2$ . Then the quadruple $(G,\lambda ,p,1)$ has generic $ts$ -stabilizer $(U_5A_1)^2.\mathbb {Z}_2$ .

Proof. This result is already listed in [Reference Rizzoli26, Thm. 1], albeit without a full explanation. The quadruple $(D_5,\lambda _5,2,1)$ has generic stabilizer $(G_2G_2).\mathbb {Z}_2$ by [Reference Guralnick and Lawther10]. By [Reference Rizzoli26, Lemma 5.17], we deduce that $C_{\mathcal {S}_1(V)}$ is isomorphic to the generic stabilizer of the action of $(G_2G_2).\mathbb {Z}_2$ on $V_{G_2}(\lambda _1)\oplus V_{G_2}(\lambda _1)$ . Since $G_2$ is transitive on nonzero vectors of $V_{G_2}(\lambda _1)$ , the generic stabilizer for this action is easily seen to be $(U_5A_1)^2.\mathbb {Z}_2$ .

Proposition 5.3. Let $G= G_2$ , $\lambda = \lambda _1$ with $p\neq 2$ . Then the quadruple $(G,\lambda ,p,3)$ has generic $ts$ -stabilizer $A_2$ .

Proof. As in Proposition 3.13, we know that there is $y\in \mathcal {S}_3(V)$ with $G_y = A_2$ . Since $\dim G - \dim A_2 = 6 = \dim \mathcal {S}_3(V)$ , we conclude that $C_{\mathcal {S}_3(V)} = A_2$ .

We now relax slightly the condition that the group acting should be simple: we allow a product of isomorphic simple groups, possibly extended by a graph automorphism. If the connected group is of the form $G_1G_2G_3$ , we write $\lambda = \mu _1\otimes \mu _2\otimes \mu _3$ to mean $V_G(\lambda ) = V_{G_1}(\mu _1)\otimes V_{G_2}(\mu _2)\otimes V_{G_3}(\mu _3)$ , where each $\mu _i$ is a dominant weight for $G_i$ .

Lemma 5.4. Let $G=A_1^3$ , $\lambda =\lambda _1\otimes \lambda _1\otimes \lambda _1$ with $p=2$ . Then there is an open dense subset $\hat {Y}$ of $\mathcal {S}_2(V)$ such that for all $y_1\neq y_2\in \hat {Y}$ , we have $G_{y_1}\simeq G_{y_2} = U_1.\mathbb {Z}_2$ , and $G_{y_1}$ is not conjugate to $G_{y_2}$ .

Proof. This is proved in [Reference Rizzoli27, Lemma 5.69].

Lemma 5.5. Let $G=A_1^3$ , $\lambda =\lambda _1\otimes \lambda _1\otimes \lambda _1$ with $p\neq 2$ . The quadruple $(G,\lambda ,p,2)$ has generic $ts$ -stabilizer $\mathbb {Z}_2\times \mathbb {Z}_2$ .

Proof. We can recover this result by a slight change of the proof of [Reference Guralnick and Lawther10, Prop. 6.7]. Let $G=SL_2(K)^3$ with basis $e,f$ for $V_{nat}$ , so that $V=V_{nat}\otimes V_{nat}\otimes V_{nat}$ . Then, like in the proof of [Reference Guralnick and Lawther10, Prop. 6.7], given $\mathbf {a}=(a_1,a_2,a_3,a_4)$ , let

$$\begin{align*}v^{(1)}&=a_1e\otimes e\otimes e+ a_2e\otimes f\otimes f+a_3f\otimes e\otimes f+a_4f\otimes f\otimes e,\\v^{(2)}&=a_1f\otimes f\otimes f+ a_2f\otimes e\otimes e+a_3e\otimes f\otimes e+a_4e\otimes e\otimes f. \end{align*}$$

Then $y_{\mathbf {a}}:=\langle v^{(1)},v^{(2)} \rangle $ is totally singular if and only if $a_1^2+a_2^2+a_3^2+a_4^2 = 0$ . In the proof of [Reference Guralnick and Lawther10, Prop. 6.7], they then proceed to define $\hat {Y}$ as a dense subset of $Y:=\{ y_{\mathbf {a}}:\mathbf {a}\neq (0,0,0,0)\}$ by requiring certain polynomials on the coefficients to be nonzero. One of these conditions is that $a_1^2+a_2^2+a_3^2+a_4^2 \neq 0$ . However, this is later only used for the case $p=2$ . Therefore, we can modify the definition of $\hat {Y}$ in the proof of [Reference Guralnick and Lawther10, Prop. 6.7] and still end up with an open dense subset of $\mathcal {G}_2(V)$ where all stabilizers are conjugate to $\mathbb {Z}_2.\mathbb {Z}_2$ . The difference is that now this set will also contain totally singular $2$ -spaces, and therefore, $C_{\mathcal {S}_2(V)}=C_{\mathcal {G}_2(V)}=\mathbb {Z}_2.\mathbb {Z}_2$ .

Proposition 5.6. Let $G=E_7$ and $\lambda =\lambda _7$ , or $G=D_6$ and $\lambda =\lambda _6$ , or $G=A_5$ and $\lambda =\lambda _3$ . Then the quadruple $(G,\lambda ,p,2)$ has no generic $ts$ -stabilizer if $p=2$ but has a semi-generic $ts$ -stabilizer. If $p\neq 2$ , we have $C_{\mathcal {S}_2(V)}=C_{\mathcal {G}_2(V)}$ .

Proof. We shall describe how to use the proof of [Reference Guralnick and Lawther10, Proposition 6.28] to reach the conclusion. In [Reference Guralnick and Lawther10, Proposition 6.28], the authors determine the generic stabilizer for the G-action on all $2$ -spaces. They do so in the following manner. They define a certain $8$ -space $V_{[0]}$ spanned by pairs of opposite weight vectors. This $8$ -space is the fixed point space of a subgroup A of G, where $A=D_4,A_1^3,T_2$ according to whether $G=E_7,D_6,A_5$ , respectively. They define a dense subset $\hat {Y}_1$ of $Y:=\mathcal {G}_2(V_{[0]})$ with the property that for any $y\in Y$ , we have $\mathrm {Tran}_G(y,Y) = A A_1^3.Sym(3)$ , where $A_1^3$ acts on $V_{[0]}$ as $\lambda _1\otimes \lambda _1\otimes \lambda _1$ . The set $\hat {Y}_1$ is defined by requiring certain expressions in terms of the coefficients of the given $V_{[0]}$ basis to be nonzero. Here, the key observation is that these conditions do not exclude all totally singular $2$ -spaces of $V_{[0]}$ , and therefore, $\hat {Y}^{\mathcal {S}}_1:=\hat {Y}_1\cap \mathcal {S}_2(V_{[0]})$ is a dense subset of $Y^{\mathcal {S}}:=\mathcal {S}_2(V_{[0]})$ . Given $y\in \hat {Y}^{\mathcal {S}}_1 $ , since $\mathrm {Tran}_G(y,Y) = A A_1^3.Sym(3)$ , we also have $\mathrm {Tran}_G(y,Y^{\mathcal {S}}) = A A_1^3.Sym(3)$ . Furthermore, they show that $G_y = A(A_1^3)_y$ for all $y\in \hat {Y}_1$ .

Assume $p\neq 2$ . Then by Lemma 5.5, there exists a dense open subset $\hat {Y}^{\mathcal {S}}_2$ of $Y^{\mathcal {S}}$ such that every stabilizer is $A_1^3$ -conjugate to $\mathbb {Z}_2.\mathbb {Z}_2$ . Taking the intersection with $\hat {Y}^{\mathcal {S}}_1$ , we get an open dense subset $\hat {Y}^{\mathcal {S}}$ of $Y^{\mathcal {S}}$ . For all $y\in \hat {Y}^{\mathcal {S}}$ , we know that $G_y$ is conjugate to $A.\mathbb {Z}_2.\mathbb {Z}_2$ . In each case, the codimension of the transporter of $y\in \hat {Y}^{\mathcal {S}}$ into $Y^{\mathcal {S}}$ is equal to the codimension of $Y^{\mathcal {S}}$ in $\mathcal {S}_2(V)$ . Therefore, every $y\in \hat {Y}^{\mathcal {S}}$ is $Y^{\mathcal {S}}$ -exact. By Lemma 2.15, we conclude that $C_{\mathcal {S}_2(V)} $ is $D_4.\mathbb {Z}_2.\mathbb {Z}_2,A_1^3.\mathbb {Z}_2.\mathbb {Z}_2,T_2.\mathbb {Z}_2.\mathbb {Z}_2$ according to whether $G=E_7,D_6,A_5$ , respectively. These generic stabilizers are the same as for the action on all $2$ -spaces.

Now assume that $p=2$ . By Lemma 5.4, there exists a dense open subset $\hat {Y}^{\mathcal {S}}_2$ of $Y^{\mathcal {S}}$ such that every stabilizer has a $1$ -dimensional connected component and stabilizers are pairwise non-conjugate. Taking the intersection with $\hat {Y}^{\mathcal {S}}_1$ , we get a dense open subset $\hat {Y}^{\mathcal {S}}_3$ of $Y^{\mathcal {S}}$ . Let $y_1,y_2\in \hat {Y}^{\mathcal {S}}_3$ and assume that $x.y_1 = y_2$ . Then $x\in N_G(A)$ , and since $V_{{0}}$ is the fixed space of A, we must have $x\in \mathrm {Tran}_G(y_1,Y^{\mathcal {S}}) = A A_1^3.Sym(3)$ . Therefore, $y_1$ and $y_2$ must be in the same $A_1^3.Sym(3)$ -orbit. However by construction, $y_1$ and $y_2$ are not in the same $A_1^3$ -orbit, and therefore, there exists a dense open subset $\hat {Y}^{\mathcal {S}}$ of $Y^{\mathcal {S}}$ , contained in $\hat {Y}^{\mathcal {S}}_3$ , such that any two distinct elements have non-conjugate stabilizers. In each case, the codimension of the transporter of $y\in \hat {Y}^{\mathcal {S}}$ into $Y^{\mathcal {S}}$ is equal to the codimension of $Y^{\mathcal {S}}$ in $\mathcal {S}_2(V)$ . Therefore, every $y\in \hat {Y}^{\mathcal {S}}$ is $Y^{\mathcal {S}}$ -exact. By Lemma 2.16, there is no generic stabilizer in the action of G on $\mathcal {S}_2(V)$ . In fact, Lemma 2.14 shows that we have semi-generic stabilizers $D_4.U_1.\mathbb {Z}_2,A_1^3.U_1.\mathbb {Z}_2,T_2.U_1.\mathbb {Z}_2$ according to whether $G=E_7,D_6,A_5$ , respectively.

Lemma 5.7. Let $G=A_2$ , $\lambda =\lambda _1+\lambda _2$ with $p\neq 3$ . Let $\tau $ be a graph automorphism of G. Then the quadruple $(G\langle \tau \rangle , \lambda ,p,2)$ has generic $ts$ -stabilizer $Sym(3)$ .

Proof. Take $G = SL_3(K)$ acting on $V = \mathfrak {sl}_3(K)$ by conjugation. Here, $Z(G) = \langle \mathrm {diag} (\omega ,\omega ,\omega ) \rangle $ , where $\omega $ is a nontrivial third-root of unity. Note that $Z(G)$ acts trivially on V. Let $\tau $ be the graph automorphism acting on G as $g\mapsto g^{-T}$ and on V as $v\mapsto -v^T$ . We have that G fixes a nondegenerate quadratic form on V given by

$$\begin{align*}Q\left( (m_{ij})_{ij} \right) = m_{11}^2+m_{22}^2+m_{11}m_{22}+\sum_{i<j}m_{ij}m_{ji}.\end{align*}$$

For $1\leq i,j\leq 3$ , let $e_{ij}$ denote a $3\times 3$ matrix with a $1$ in position $(i,j)$ and zeroes everywhere else. There are three G-orbits on $\mathcal {S}_1(V)$ , which we label as $\Delta _1$ , $\Delta _2$ and $\Delta _3$ , respectively with representatives $\langle e_{13}\rangle $ , $\langle e_{12}+e_{23}\rangle $ , $\langle e_{11}+\omega e_{22}+\omega ^2 e_{33}\rangle $ . This follows directly from considering the Jordan Canonical Form of elements in V. The stabilizers are respectively $B = U_3T_2$ , $Z(G).U_2T_1$ and $T_2.\mathbb {Z}_3$ .

Let

$$\begin{align*}u_{bc} = \left(\begin{matrix} & 1 \\ & & b \\c & \\ \end{matrix}\right),\quad v_{ad} = \left(\begin{matrix} & & 1 \\ a & \\ & d \\ \end{matrix}\right); \end{align*}$$
$$\begin{align*}Y = \left\lbrace \langle u_{bc}, v_{ad}\rangle : a+c+bd = 0 \right\rbrace.\end{align*}$$

The set Y is a $3$ -dimensional subvariety of $\mathcal {S}_2(V)$ . Let

$$\begin{align*}\hat{Y} = \left\lbrace \langle u_{bc}, v_{ad}\rangle\in Y : abcd \neq 0,\, \frac{(bd-c)^2}{bcd}\neq 0,-\frac{3}{2},-3 \right\rbrace, \end{align*}$$

where we disregard the expression $\frac {(bd-c)^2}{bcd}\neq -\frac {3}{2}$ if $p = 2$ . Then $\hat {Y}$ is a dense subset of Y. Let $y = \langle u_{bc}, v_{ad}\rangle \in \hat {Y}$ . Then $\langle u_{bc} \rangle $ and $\langle v_{ad}\rangle $ are in $\Delta _3$ since they are spanned by rank- $3$ matrices. Now consider $v = u_{bc}+\lambda v_{ad}$ . We have $\det v = bc+ad \lambda ^3$ , which implies that there are precisely three $1$ -spaces of y not belonging to $\Delta _3$ . It is clear that none of these have rank $1$ , and therefore, all three of these $1$ -spaces belong to $\Delta _2$ . Let $\lambda _1,\lambda _2,\lambda _3$ be the three distinct roots of $q(x) = bc+ad x^ 3 $ , so that $\langle u_{bc}+\lambda _i v_{ad}\rangle \in \Delta _2$ . Then

$$\begin{align*}G_y\leq (G_{\langle u_{bc}+\lambda_1 v_{ad}\rangle }\cap G_{\langle u_{bc}+\lambda_2 v_{ad}\rangle }).Sym(3). \end{align*}$$

Let $g^* = \operatorname {\mathrm {diag}}(1,\omega ,\omega ^2)\in G$ . Then $g^*. u_{bc} = \omega ^2 u_{bc}$ and $g^*. v_{ad} = \omega v_{ad}$ , implying $g^*\in G_y$ and $Z(G)\langle g^* \rangle \simeq Z(G).\mathbb {Z}_3\leq G_y$ . Take $\mu ,\,\nu \in K$ with $\mu ^3 = \frac {1}{ac}$ and $\nu ^3 = \frac {ab}{d}$ , and let

$$\begin{align*}g^\ddagger =\operatorname{\mathrm{diag}} (\mu,\nu,{(\mu\nu)^{-1}})\tau ,\end{align*}$$

an element of $N_{G\langle \tau \rangle }(\langle g^* \rangle )$ fixing y. All elements of the form $h\tau $ with $h\in T$ are conjugate under T. Therefore, for any $y_1,y_2\in \hat {Y}$ , we know that $(G\langle \tau \rangle )_{y_1}$ and $(G\langle \tau \rangle )_{y_2}$ contain a subgroup conjugate to $Z(G).\langle g^*\rangle \langle g^\ddagger \rangle \simeq Z(G).Sym(3)$ .

We now proceed in the following way. We show that the stabilizer in $G_{\langle u_{bc}+\lambda _1 v_{ad}\rangle }$ of y is $Z(G)$ . This in turn implies that $G_y = Z(G).\mathbb {Z}_3$ . Since $\langle u_{bc}+\lambda _1 v_{ad} \rangle \in \Delta _2$ , we are able to find an element of G sending $u_{bc}+\lambda _1 v_{ad} \mapsto e_{12}+e_{23}$ . This is achieved by a scalar multiple of

$$\begin{align*}R = \left(\begin{matrix} 0 & 0 & 1 \\ c &\lambda_1 d & 0 \\ -\frac{bc}{\lambda_1} & c & -\lambda_1 a\\ \end{matrix}\right).\end{align*}$$

We now have

$$\begin{align*}R.y = \left\langle \left(\begin{matrix} a_1 & 0 & 1 \\ a_2 & a_1 & 0 \\ a_3 & -a_2 & -2 a_1\\ \end{matrix}\right) , e_{12}+e_{23}\right\rangle,\end{align*}$$

where $a_1=\lambda _1a,\,a_2 = \frac {a}{d}(bd-c) = \frac {c^2}{d}-b^2d$ , and $a_3 = \frac {3abc}{\lambda _1 d} = -3a_1^2$ . By assumption, $a_1a_2a_3\neq 0$ . Now let $g\in G_{\langle e_{12}+e_{23} \rangle }$ . Multiplying by an element of $Z(G)$ , we can assume that $g=ns$ , where $s = \mathrm {diag}(\frac {1}{t},1,t)$ and $n = \left (\begin {matrix} 1 & n_1 & n_2 \\ 0 & 1 & n_1 \\0 & 0 & 1\\ \end {matrix}\right ).$ Then

$$ \begin{align*} (m_{ij})_{ij}&:=g. \left(\begin{matrix} a_1 & 0 & 1 \\ a_2 & a_1 & 0 \\-3a_1^2 & -a_2 & -2 a_1\\ \end{matrix}\right) =\\ &= \left(\begin{matrix} a_1 + a_2 t n_1 - 3 a_1^2 t^2n_2 & m_{12} & m_{13} \\ t (a_2 - 3 a_1^2 t n_1) & a_1 - 2 a_2 t n_1 + 3 a_1^2 t^2 n_1^2 & m_{23} \\ -3 a_1^2 t^2 & -t (a_2 - 3 a_1^2 t n_1) & -2 a_1 + a_2 t n_1 - 3 a_1^2t^2 (n_1^2- n_2)\\ \end{matrix}\right),\end{align*} $$

where

$$ \begin{align*} m_{12}&=t (a_2 (-n_1^2 - n_2) + 3 a_1^2 t n_1 n_2),\\ m_{13}&= \frac{1}{t^2} + a_2 t n_1^3 - 3 a_1n_2 - 3 a_1^2 t^2 n_2 (n_1^2 - n_2),\\ m_{23}&= -3 a_1 n_1 - 3 a_1^2 t^2 (n_1^2-n_2) + a_2 t (n_1^2 - n_2).\end{align*} $$

Assume that g fixes $R.y$ . Since $m_{31}=-3a_1^2t^2$ , we must have $(m_{ij})_{ij} = t^2 R.y + \alpha (e_{12}+e_{23})$ for some $\alpha \in K$ . Therefore, $m_{21} = t^2 a_2$ which implies $n_1 = \frac {a_2(1-t)}{3 a_1^2 t}.$ Similarly, we must have $m_{11}=m_{22}$ which implies $n_2 = \frac {a_2 n_1}{a_1^2 t }-n_1^2$ . Then $m_{11} = t^2 a_1$ implies that either $t=\pm 1$ or $a_2^2= 3a_1^3$ . Assume that $a_2^2= 3a_1^3$ . Since $a_2=\frac {a}{d}(bd-c)$ and $a_1=\lambda _1 a$ , the equation $a_2^2= 3a_1^3$ implies $(bd-c)^2 = -3bcd$ , contrary to the definition of $\hat {Y}$ . Therefore, $t=\pm 1$ . If $t=1$ , we immediately get $n_1=n_2=0$ , concluding. Assume therefore that $t=-1$ and $p\neq 2$ . Then $m_{12}=m_{23}$ forces $2a_2^2 = 3 a_1^3$ , which implies $ 2(bd-c)^2 = -3bcd$ , which is impossible by assumption on $\hat {Y}$ .

This concludes the proof that $G\langle \tau \rangle _y = Z(G)\langle g^*\rangle \langle g^\ddagger \rangle $ . Now for any $y\in \hat {Y}$ , since $G_y= Z(G)\langle g^*\rangle $ , any element in $\operatorname {\mathrm {Tran}} _G(y,Y)$ must be in $N( \langle g^*\rangle )$ . We know that $N( \langle g^*\rangle ) = T_2.\mathbb {Z}_3$ , and it is easy to check that $T_2\in \operatorname {\mathrm {Tran}}_G(y,Y)$ . Therefore, $\dim \operatorname {\mathrm {Tran}}_G(y,Y) =2 $ , and then since $\dim \mathcal {S}_2(V)-\dim G = 1$ , the set $\hat {Y}$ is Y-exact. By Lemma 2.15, we conclude that the quadruple $(G\langle \tau \rangle , \lambda ,p,2)$ has generic $ts$ -stabilizer $Sym(3)$ .

Proposition 5.8. Let $G=F_4$ and $\lambda =\lambda _4$ , or $G=C_3$ and $\lambda =\lambda _2$ , with $p\neq 3$ . Then the quadruple $(G,\lambda ,p,2)$ has generic $ts$ -stabilizer $A_2.Sym(3)$ or $T_1.Sym(3)$ , respectively.

Proof. This is entirely similar to the proof of Proposition 5.6 and relies on the construction used in [Reference Guralnick and Lawther10, Prop. 6.26]. All we have to observe is that the set $\hat {Y}_1$ defined in the proof of [Reference Guralnick and Lawther10, Prop. 6.26] does indeed contain totally singular $2$ -spaces. This follows from the observation in the proof of [Reference Guralnick and Lawther10, Prop. 6.25], where the authors need to show that the set $\hat {Y}_1$ is nonempty. They do so by saying that $v^{(1)}=a_{33}e_{\gamma _{33}}+a_{12}e_{\gamma _{12}}+a_{21}e_{\gamma _{21}}$ , $v^{(2)}=b_{11}e_{\gamma _{11}}+b_{23}e_{\gamma _{23}}+b_{32}e_{\gamma _{32}}$ , and $v^{(3)}=e_{\gamma _{22}}+e_{\gamma _{31}}+e_{\gamma _{13}}$ span a $3$ -space in $\hat {Y}_1$ if $(a_{12}b_{23}-a_{33}b_{11})(a_{21}b_{32}-a_{12}b_{23})(a_{33}b_{11}-a_{21}b_{32})\neq 0$ . Clearly, there are totally singular $2$ -spaces $\langle v^{(1)},v^{(2)}\rangle $ with coefficients satisfying this condition, and therefore, the set $\hat {Y}_1$ defined in the proof of [Reference Guralnick and Lawther10, Prop. 6.26] does contain totally singular $2$ -spaces.

Once we understand this, the generic stabilizer is, respectively, $A_2.X$ , $T_1.X$ , where X is the generic stabilizer for the action of $A_2.\mathbb {Z}_2$ on $\mathcal {S}_2(\lambda _1+\lambda _2)$ . By Lemma 5.7, we conclude.

Proposition 5.9. Let $G=A_2$ and $\lambda =\lambda _1+\lambda _2$ with $p=3$ . Then $C_{\mathcal {S}_3}(V) = T_2.\mathbb {Z}_3$ .

Proof. Let $\alpha _1,\alpha _2$ be the fundamental roots for $A_2$ and let $\alpha _3=\alpha _1+\alpha _2$ . The adjoint module $\mathrm {Lie}(G)$ has the Chevalley basis $e_{\alpha _3},e_{\alpha _2},e_{\alpha _1},h_{\alpha _1},h_{\alpha _2},e_{-\alpha _1},e_{-\alpha _2},e_{-\alpha _3}$ . We write $v_1v_2$ for the Lie product of vectors $v_1,v_2\in \mathrm {Lie}(G)$ . We assume that the structure constants are as described by the matrix

$$\begin{align*}\left(\begin{matrix} 0 & 0 & 0 &-1 & 1 & 0 \\ 0 & 0 & -1 &0 & 0 & 1 \\0 & 1 & 0 &0 & 0 & -1\\ 1 & 0 & 0 &0 & -1 & 0 \\ -1 & 0 & 0 &1 & 0 & 0\\ 0 & -1 & 1 &0 & 0 & 0\\ \end{matrix}\right),\end{align*}$$

where the rows and columns are in the order $\alpha _3,\alpha _2,\alpha _1,-\alpha _1-\alpha _2,-\alpha _3$ . By [Reference Rizzoli27, Lemma 5.4], we can explicitly construct our highest weight irreducible module as

$$ \begin{align*}V_{G}(\lambda_1+\lambda_2)= \mathrm{Lie}(G)/\langle h_{\alpha_1}-h_{\alpha_2} \rangle.\end{align*} $$

In a slight abuse of notation, we omit writing the quotient, so that v actually stands for $v+\langle h_{\alpha _1}-h_{\alpha _2} \rangle $ . We order the basis for $V_{G}(\lambda _1+\lambda _2 )$ as $e_{\alpha _3},e_{\alpha _2},e_{\alpha _1},h_{\alpha _1},e_{-\alpha _1},e_{-\alpha _2},e_{-\alpha _3}$ . With respect to this ordering, using standard formulas found in [Reference Carter7, §4.4], we find the matrices denoting the transformations $x_{\pm \alpha _1}(t),x_{\pm \alpha _2}(t),x_{\pm \alpha _3}(t)$ , as well as $h_{\alpha _1}(\kappa )$ and $h_{\alpha _2}(\kappa )$ . These are straightforward calculations, and we therefore only state the results.

$h_{\alpha _1}(\kappa )= \mathrm {diag}(\kappa ,\kappa ^{-1},\kappa ^2,1,\kappa ^{-2},\kappa ,\kappa ^{-1})$ and $h_{\alpha _2}(\kappa )= \mathrm {diag}(\kappa ,\kappa ^2,\kappa ^{-1},1,\kappa ,\kappa ^{-2},\kappa ^{-1})$ .

Let $(\cdot ,\cdot ):V\times V \rightarrow K$ be the nondegenerate symmetric bilinear form given by $(e_{\alpha _i},e_{-\alpha _j})=\delta _{ij}$ , $(e_{\alpha _i},e_{\alpha _j})=(e_{-\alpha _i},e_{-\alpha _j})=0$ , $(h_{\alpha _1},e_{\pm \alpha _i})=0$ and $(h_{\alpha _1},h_{\alpha _1})=-1$ , where $1\leq i,j\leq 3$ . Then G fixes this form, as can be seen by just checking the action of the generators.

We need some information about the action of G on singular $1$ -spaces. Let $T=\langle h_{\alpha _1}(\kappa ),h_{\alpha _2}(\kappa ):\kappa \in K^*\rangle $ be the standard maximal torus and $B=\langle T,x_{\alpha _1}(t),x_{\alpha _2}(t):t\in K\rangle $ a Borel subgroup. Then by [Reference Rizzoli27, Lemma 5.5], the group G has $2$ orbits on singular vectors in V, with representatives $x=e_{\alpha _3}$ and $y=e_{\alpha _1}+e_{\alpha _2}$ . Furthermore, $G_x=U_3T_1$ , $G_y=U_2$ , $G_{\langle x \rangle }=U_3T_2=B$ and $G_{\langle y \rangle }=U_2T_1\leq B$ .

We now define a totally singular $3$ -space that we will show has a $2$ -dimensional stabilizer, therefore belonging to a dense orbit. Let

$$\begin{align*}W_3 = \langle e_{\alpha_1}, e_{\alpha_2},e_{-\alpha_3}\rangle.\end{align*}$$

Let $\Delta $ be the G-orbit with representative $\langle x\rangle $ , where $x = e_{\alpha _3}$ . We start by observing that

$$\begin{align*}\mathcal{G}_1(W_3)\cap \Delta = \{\langle e_{\alpha_1} \rangle, \langle e_{\alpha_2} \rangle,\langle e_{-\alpha_3}\rangle\}.\end{align*}$$

It suffices to show that every other $1$ -space of $W_3$ is in the same orbit as $\langle y\rangle $ , where $y = e_{\alpha _1}+ e_{\alpha _2}$ . This is clear for $\langle e_{\alpha _1}+ \lambda e_{\alpha _2}\rangle $ when ${\lambda \neq 0} $ . The element $x_{-\alpha _1}(t)x_{-\alpha _2}(-t)x_{-\alpha _3}(t^2)$ sends $y\mapsto y+t^3e_{-\alpha _3}$ ; therefore, $\langle e_{\alpha _1}+\lambda e_{\alpha _2}+\gamma e_{-\alpha _3}\rangle $ is in the orbit of $\langle y \rangle $ for $\lambda ,\gamma \neq 0$ . Finally, $n_{\alpha _2}(1)n_{\alpha _1}(1). \langle y\rangle =\langle e_{\alpha _1}+e_{-\alpha _3} \rangle $ and $n_{\alpha _1}(-1)n_{\alpha _2}(-1). \langle y\rangle =\langle e_{\alpha _2}+e_{-\alpha _3} \rangle $ , concluding.

Since $\dim \mathcal {S}_3(V) = 6$ , it suffices to prove that $G_{W_3} = T.\mathbb {Z}_3$ . Since y only contains three $1$ -spaces that are in the orbit $\Delta $ , the connected component of $G_{W_3}$ is simply

$$\begin{align*}G_{\langle e_{\alpha_1}\rangle}\cap G_{\langle e_{\alpha_2}\rangle}\cap G_{\langle e_{-\alpha_3}\rangle}, \end{align*}$$

which is easily seen to be the maximal torus T. Finally, one checks that the elements of order $2$ of $W=N_G(T)/T$ do not fix $W_3$ , while the elements of order $3$ do. Therefore, $G_{W_3} \simeq T_2.\mathbb {Z}_3$ , and indeed, $C_{\mathcal {S}_3}(V) = T_2.\mathbb {Z}_3$ .

Proposition 5.10. Let $G=A_2$ and $\lambda = \lambda _1+\lambda _2$ with $p\neq 3$ . Then $C_{\mathcal {S}^{\prime }_4(V)}=C_{\mathcal {S}^{\prime \prime }_4(V)}=T_2.\mathbb {Z}_3$ .

Proof. The group G is stable under a triality automorphism of $D_4$ ; therefore, the $(G,P_3)$ -double cosets in $D_4$ are in bijection with the $(G,P_1)$ -double cosets in $D_4$ , as are the $(G,P_4)$ -double cosets. By Proposition 4.3, the group G acts on $\mathcal {S}_1(V)$ with generic stabilizer $T_2.\mathbb {Z}_3$ , concluding.

Proposition 5.11. Let $G=B_3$ and $\lambda = \lambda _3$ . Then $C_{\mathcal {S}^{\prime }_4(V)}=C_{\mathcal {S}^{\prime \prime }_4(V)}=C_{\mathcal {S}_1(V)}=U_6A_2T_1$ .

Proof. Similarly to Proposition 5.10, the $(G,P_3)$ -double cosets in $D_4$ are in bijection with the $(G,P_1)$ -double cosets in $D_4$ , as are the $(G,P_4)$ -double cosets. Therefore, $C_{\mathcal {S}^{\prime }_4(V)}=C_{\mathcal {S}^{\prime \prime }_4(V)}=C_{\mathcal {S}_1(V)}=P_3(G)$ , as claimed.

Proposition 5.12. Let $G=B_4$ and $\lambda =\lambda _4$ . Then $C_{\mathcal {S}_3(V)} = A_1$ .

Proof. By [Reference Guralnick and Lawther10, Prop. 6.21], we already know that $D_5$ has a dense orbit on all $3$ -spaces of $V_{D_5}(\lambda _5)$ . We construct the module $V_{D_5}(\lambda _5)$ in the same way as in [Reference Guralnick and Lawther10, Prop. 6.21] and then consider the restriction to G. Let $\beta _1,\dots ,\beta _6$ be the simple roots of a group of type $E_6$ and let $D_5<E_6$ have simple roots $\alpha _1=\beta _1, \alpha _2=\beta _3, \alpha _3=\beta _4, \alpha _4=\beta _5, \alpha _5=\beta _2$ . Then we may take

$$\begin{align*}V_{D_5}(\lambda_5) = \left\langle e_{\alpha}:\alpha = \sum m_i\beta_i, m_6 = 1\right\rangle < \mathrm{Lie}(E_6).\end{align*}$$

With this notation, [Reference Guralnick and Lawther10, Prop. 6.21] shows that if we write

$$\begin{align*}\gamma_2 = 101111,\, \gamma_3 = 011111,\,\gamma_4 = 111111,\,\gamma_5=011211,\,\gamma_6= 111211,\,\gamma_7 =011221 ,\end{align*}$$

then

$$\begin{align*}W_3:= \langle e_{\gamma_2}+e_{\gamma_3}, e_{\gamma_4}+e_{\gamma_5}, e_{\gamma_6}+e_{\gamma_7}\rangle \end{align*}$$

has stabilizer $A_1A_1$ in $D_5/Z(D_5)$ . The generators for the first $A_1$ are simply $X_{\pm \rho }$ , where $\rho $ is the longest root of $D_5$ , while the second $A_1$ is generated by

$$ \begin{align*} x(t) & = x_{\beta_1}(-t)x_{\beta_4}(2t)x_{\beta_5}(t)x_{\beta_2}(3t)x_{\beta_4+\beta_5}(-t^2)x_{\beta_2+\beta_4}(3t^2)x_{\beta_2+\beta_4+\beta_5}(4t^3) \text{ for } t\in K; \\ T_1& =\{h_{\beta_1}(\kappa)h_{\beta_2}(\kappa^3)h_{\beta_4}(\kappa^4)h_{\beta_5}(\kappa^3):\kappa\in K^*\};\\ n & = n_{\beta_1}n_{\beta_4}n_{\beta_2\beta_4\beta_5}^{-1}. \end{align*} $$

Let $\{e_1,\dots ,e_5,f_5,\dots , f_1\}$ be the standard basis for the natural $D_5$ -module, and let $G=(D_5)_{\langle e_2+f_2 \rangle }$ . We will show that $W_3$ is totally singular and has stabilizer $A_1$ in $G/Z(G)$ . Let $T=\langle h_{\beta _i}(\kappa _i): 1\leq i\leq 5, \kappa _i\in K^* \rangle $ be the standard maximal torus of $D_5$ . Then $T_G:=T\cap G= \langle h_{\beta _i}(\kappa _i): 1\leq i\leq 5, \kappa _i\in K^* , \kappa _1=\kappa _3 \rangle $ , since $\alpha _1=\beta _1$ and $\alpha _2=\beta _3$ . Then the $T_G$ -weights on $\langle e_{\gamma _i}\rangle _{2\leq i\leq 7}$ are respectively given by $\frac {\kappa _1}{\kappa _2},\frac {\kappa _2}{\kappa _4},\frac {\kappa _1\kappa _2}{\kappa _4},\frac {\kappa _4}{\kappa _1\kappa _5},\frac {\kappa _4}{\kappa _5},\frac {\kappa _5}{\kappa _1}$ . No two such weights form a pair of opposite weights, and therefore by Lemma 2.7, the subspace $W_3$ is totally singular.

Finally, we find that the diagonal subgroup of $A_1A_1$ with positive root subgroup $x(t)x_{\rho }(t)$ fixes $\langle e_2+f_2\rangle $ . By maximality of this diagonal $A_1$ in $A_1A_1$ , this means that $(A_1A_1)\cap G =A_1$ . Therefore, $\dim G - \dim \mathcal {S}_3(V) = 3 = \dim G_{W_3}$ , which implies that $C_{\mathcal {S}_3(V)} = A_1$ .

In what follows, we refer the reader back to Section 2.4 for the relevant notation on spin modules. For the following propositions, let $G=B_4 = (D_5)_{e_5-f_5}$ , $\lambda =\lambda _4$ and order the basis of V as

$$ \begin{align*} v_1&= 1, & v_5&=e_1e_5, & v_9&=e_1e_2e_3e_5, & v_{13}&=e_1e_4,\\ v_2&=e_1e_2, & v_6&=e_2 e_5, &v_{10}&=e_1e_2e_4e_5, & v_{14}&=e_2e_4,\\ v_3&=e_1e_3, &v_7&=e_3e_5, & v_{11}&=e_1e_3e_4e_5,& v_{15}&=e_3e_4,\\ v_4&=e_2e_3, & v_8&=e_4e_5, & v_{12}&=e_2e_3e_4e_5,& v_{16}&=e_1e_2e_3e_4.\end{align*} $$

Lemma 5.13. The quadratic form given by the matrix $\operatorname {\mathrm {antidiag}} (1,-1,1,-1,1,-1,1,-1,0,0,0,0,0,0,0,0)$ defines a nondegenerate quadratic form on $V_{B_4}(\lambda _4)$ fixed by the $B_4$ -action.

Proof. The pairs $v_i,v_{17-i}$ are pairs of opposite weight vectors. We can assume that $Q(v_1+v_{16})=1$ and use the $B_4$ -action to determine whether $Q(v_i+v_{17-i})$ is $1$ or $-1$ . Let $g=1+e_1e_2\in B_4$ . Then $g.v_1 = 1+e_1e_2 = v_1+v_2$ and $g.v_{15}=e_3e_4+e_1e_2e_3e_4=v_{15}+v_{16}$ . Therefore, $0=Q(v_1+v_{15}) = Q(v_1+v_2+v_{15}+v_{16})=1+Q(v_2+v_{15})$ , as claimed. The same approach shows that $Q(v_3+v_{14})=1$ and $Q(v_4+v_{13})=-1$ . To conclude, let $g= (1+e_1f_5)(1+e_1e_5)$ which is an element of $B_4$ since it fixes $e_5-f_5$ . Then $0=Q(v_1+v_{12}) = Q(v_1+v_5+v_{12}-v_{16})=-1+Q(v_5+v_{12})$ . The remaining cases follow similarly.

Lemma 5.14. For $\alpha \in \Phi (B_4)$ , the root elements $x_{\alpha }(t)$ are written in terms of spinors as follows:

$$ \begin{align*} x_{\epsilon_i-\epsilon_j}(t) &= 1+t e_i f_j, \\ x_{-\epsilon_i+\epsilon_j}(t) &= 1+t e_j f_i, \\ x_{\epsilon_i+\epsilon_j}(t) &= 1+t e_i e_j, \\ x_{-\epsilon_i-\epsilon_j}(t) &= 1-t f_i f_j, \\ x_{\epsilon_i}(t) &= (1+t e_i e_5)(1+t e_i f_5), \\ x_{-\epsilon_i}(t) &= (1-t f_i e_5)(1-t f_i f_5). \end{align*} $$

Proof. The action on $V_{nat}$ of the elements on the left-hand sides of each equation is as described at the beginning of Section 2.1. The action on $V_{nat}$ of the elements on the right-hand sides of each equation is described in Section 2.4. The result follows by comparing the two actions.

Proposition 5.15. Let $G=B_4$ , $\lambda =\lambda _4$ . Then $C_{\mathcal {S}^{\prime }_8(V)}=A_2.\mathbb {Z}_2$ .

Proof. Suppose that $p\neq 3$ . Let $\beta _1 = \alpha _1$ , $\beta _2 = \alpha _2$ , $\beta _3 = \alpha _3$ , $\beta _4 = -\alpha _0$ , where $\alpha _0$ is the longest root in $\Phi ^+(B_4)$ . Then $\{\beta _i\}_i$ is the base of a root system of type $D_4$ . Let D be the corresponding $D_4$ -subgroup of $B_4$ . Then $V\downarrow D = \lambda _3+\lambda _4=V_8+V_8'$ . Let $\omega $ be a nontrivial third-root of unity. Let $\tau $ be the composition of $h_{\beta _2}(\omega )$ with the triality automorphism of D sending $x_{\beta _i}(t)\mapsto x_{\beta _{\sigma .i}}(t)$ for $i=1,3,4$ and $\sigma = (134)$ . Then the fixed points in D under the triality automorphism $\tau $ form an irreducible $A_2$ -subgroup of D. Using the structure constants inherited from $B_4$ , let A be the irreducible $A_2$ -subgroup of D given by

$$ \begin{align*} A =\langle &x_{\beta_1}(t)x_{\beta_3}(\omega^2 t)x_{\beta_4}(\omega t), \\ &x_{-\beta_1}(t)x_{-\beta_3}(\omega t)x_{-\beta_4}(\omega^2 t),\\ &x_{\beta_1+\beta_2}(t)x_{\beta_2+\beta_3}(-\omega t)x_{\beta_2+\beta_4}(\omega^2 t),\\ &x_{-\beta_1-\beta_2}(t)x_{-\beta_2-\beta_3}(-\omega^2 t)x_{-\beta_2-\beta_4}(\omega t) : t \in K\rangle.\end{align*} $$

By Lemma 5.14, this is the same as

$$ \begin{align*} A =\langle &(1+te_1f_2)(1+\omega^2 t e_3f_4)(1-\omega t f_1f_2), \\ &(1+te_2f_1)(1+\omega t e_4f_3)(1+\omega^2 t e_1e_2),\\ &(1+te_1f_3)(1-\omega t e_2f_4)(1-\omega^2 t f_1f_3),\\ &(1+te_3f_1)(1-\omega^2 t e_4f_2)(1+\omega t e_1e_3) : t \in K\rangle.\end{align*} $$

With this setup, we have $V_8\downarrow A \simeq V_8'\downarrow A\simeq V_{A_2}(\lambda _1+\lambda _2)$ . Then A fixes all $8$ -spaces of the form $\{v+\lambda \phi (v):v\in V_8\}$ where $\phi $ is an A-module isomorphism between $V_8$ and $V_8'$ . Given our explicit generators for A, it is easy to verify that we can take $\phi $ acting as

$$ \begin{align*} e_1e_2e_3e_4 &\mapsto e_2e_3e_4e_5, & e_2e_3 &\mapsto e_4e_5, & e_3e_4 &\mapsto \omega^2 e_1e_3e_4e_5,\\ e_1e_2 &\mapsto \omega e_2e_5, & e_1e_4 &\mapsto e_1e_2e_3e_5, & 1 &\mapsto e_1e_5,\\ e_1e_3 &\mapsto \omega^2 e_3e_5, & e_2e_4 &\mapsto \omega e_1e_2e_4e_5. & \end{align*} $$

When $\lambda \neq 0$ , the group A must be the connected component of the stabilizer of $\{v+\lambda \phi (v):v\in V_8\}$ , since the only minimal connected overgroup of A in G is D, which only fixes the $8$ -spaces $V_8$ and $V_8'$ . Also, $N_G(A)=N_{D.\mathbb {Z}_2}(A) = Z(G).A.\mathbb {Z}_2 = Z(G).A\langle \tau _2\rangle $ , where $\tau _2$ acts as a graph automorphism on $D_4$ and $A_2$ , swapping $V_8$ and $V_8 '$ . Explicit calculations show that we can take $\tau _2 = h_{\alpha _1}(-1) n $ , where $n=n_1n_2n_1n_3n_4n_3n_2n_1$ , for $n_i = n_{\alpha _i}$ . One then checks that $\tau _2$ fixes $\{v+\lambda \phi (v):v\in V_8\}$ when $\lambda ^2 = -1$ (i.e., when $\{v+\lambda \phi (v):v\in V_8\}$ is totally singular). Since $ \dim G -\dim \mathcal {S}_8'(V) = 8 = \dim A$ , we conclude that $C_{\mathcal {S}_8'(V)} = A_2.\mathbb {Z}_2$ .

Now assume that $p=3$ . This time let A be the $A_2$ -subgroup of G obtained via $V_{A_2}(\lambda _1)\otimes V_{A_2}(\lambda _2)$ . The subgroup A acts indecomposably on the natural module for G, as $1/7/1$ . More concretely, we can realise A as the subgroup generated by

$$ \begin{align*} x_{\gamma_1}(t)&:=x_{\alpha_2}(t)x_{\alpha_4}(-t)= (1+t e_2f_3) (1-t e_4e_5)(1-t e_4f_5),\\ x_{-\gamma_1}(t)&:=x_{-\alpha_2}(t)x_{-\alpha_4}(-t)= (1+t e_3f_2) (1+t f_4e_5)(1+t f_4f_5),\\ x_{\gamma_2}(t)&:=x_{\alpha_2+\alpha_3}(t)x_{\alpha_3+\alpha_4}(t)x_{\alpha_0-\alpha_2}(t)= (1+t e_2f_4) (1+t e_3e_5)(1+t e_3f_5)(1+t e_1e_3),\\ x_{-\gamma_2}(t)&:=x_{-\alpha_2-\alpha_3}(t)x_{-\alpha_3-\alpha_4}(t)x_{\alpha_1+\alpha_2}(-t)= (1+t e_4f_2) (1-t f_3e_5)(1-t f_3f_5)(1-t e_1f_3),\end{align*} $$

as t varies over K. Now let W be the $8$ -space spanned by vectors

$$\begin{align*}\begin{array}{lll} e_3e_5, & e_2e_5-e_3e_4, & e_2e_4,\\ 1-e_1e_5, & e_1e_2e_3e_4+e_2e_3e_4e_5, & e_2e_3+e_1e_2e_3e_5,\\ e_1e_2+e_1e_3e_4e_5, & e_1e_4-e_4e_5. & \end{array}\end{align*}$$

By Lemma 5.13, the subspace W is totally singular. Simple calculations show that $A\leq G_{W}$ . Furthermore, let i be a square root of $-1$ and

$$\begin{align*}\tau = h_{\alpha_1}(-1)h_{\alpha_2}(-1)h_{\alpha_3}(-1)h_{\alpha_4}(i)n_{0122}.\end{align*}$$

One checks that $\tau \in N_{G}(A_2)$ and $\tau \in G_W$ . Now let M be a minimal connected overgroup of A that fixes W. Since there are no irreducible subgroups of G containing A, we must have $A\leq M\leq P_1 = G_{\langle e_1 \rangle } = U_7B_3T_1$ . Since the projection of A on $B_3$ is an irreducible $A_2<B_3$ , the projection of M on $B_3$ is either $A_2$ or $G_2$ . As there is only one conjugacy class of $G_2$ ’s in $P_1$ , corresponding to the $G_2\leq B_3$ , the last case is not possible. Therefore, $M\leq U_7A_2T_1$ , and as the $A_2$ is acting irreducibly on the $U_7$ , we must have $M = U_7A_2$ . It is, however, straightforward to check that $ U_7\not \leq G_W$ . This proves that $A=(G_W)^0$ . The final step is to show that $C_G(A) = Z(G)$ . One way to do this is to consider the centralizer $C_1$ of $T\cap A = \langle h_{\alpha _2}(\kappa )h_{\alpha _4}(\kappa ),h_{\alpha _3}(\kappa )\rangle _{\kappa \in K^*}$ , a maximal torus of A. We find that $C_1=\langle T,X_{\pm 1111},n_0\rangle \simeq A_1T_3.\mathbb {Z}_2$ , where $n_0$ is an element of $N_G(T)$ sending each root to its negative. The centralizer of A must be contained in $P_1$ , as $\langle e_1 \rangle $ is the only $1$ -space stabilised by A. We have $C_1\cap P_1 = \langle T,X_{1111}\rangle $ , and at this point, it is easy to see that $C_G(A)=C_{C_1\cap P_1}(A) = Z(G)$ . Therefore, $G_W=Z(G).A.\mathbb {Z}_2$ , and by dimensional considerations, $A_2.\mathbb {Z}_2\leq G/Z(G)$ is the generic stabilizer for the G-action on $\mathcal {S}_8'(V)$ .

Proposition 5.16. Let $G=B_4$ , $\lambda =\lambda _4$ . Then $C_{\mathcal {S}_7(V)}=T_2.\mathbb {Z}_2$ .

Proof. By Proposition 5.15, there is a dense G-orbit on $\mathcal {S}^{\prime }_8(V)$ , with stabilizer $A_2.\mathbb {Z}_2$ . Fix an $8$ -space y in this orbit. Then $G_y = A_2.\mathbb {Z}_2 = A_2\langle \tau \rangle $ acts on y as on $\mathrm {Lie}(A_2)$ , with $\tau $ acting on $\mathrm {Lie}(A_2)$ by transposition. When $p\neq 3$ , the quadruple $(A_2.\mathbb {Z}_2,\lambda _1+\lambda _2,p,1)$ has generic stabilizer $T_2.\mathbb {Z}_2$ , as the open set for the $A_2$ -action is constructed like in Lemma 4.1 starting from elements in $\mathrm {Lie}(T)$ which are fixed by $\tau $ . The same is actually true also when $p=3$ . In this case, the $A_2\langle \tau \rangle $ module y is not irreducible, but we can still build an open dense subset of $ \mathcal {G}_1(y)$ consisting of orbits of regular semisimple elements, such that all stabilizers are conjugate to $T_2.\mathbb {Z}_2$ . Since the action on $\mathcal {G}_1(y)$ is isomorphic to the action on $\mathcal {G}_7(y)$ , there is an open dense subset $\hat {X}$ of $X:=\mathcal {G}_7(y)$ such that for all $x\in \hat {X}$ , the stabilizer $(A_2.\mathbb {Z}_2)_x$ is $A_2.\mathbb {Z}_2$ -conjugate to $T_2.\mathbb {Z}_2$ . Note that $X\subset \mathcal {S}_7(V)$ , and since every element of $\mathcal {S}_7(V)$ is a subspace of precisely one element of $\mathcal {S}^{\prime }_8(V)$ , we must have $\mathrm {Tran}_G(x,X) = G_y$ for all $x\in X$ . As $\dim G - \dim \mathrm {Tran}_G(x,X) = 28 = \dim \mathcal {S}_7(V) - \dim X$ , the set $\hat {X}$ is X-exact. By Lemma 2.15, we conclude that $C_{\mathcal {S}_7(V)}=T_2.\mathbb {Z}_2$ .

Proposition 5.17. Let $G=B_4$ , $\lambda = \lambda _4$ with $p\neq 2$ . Then $C_{\mathcal {S}^{\prime \prime }_8(V)}=A_1^3$ .

Proof. Let $V_{nat}$ be the natural module for G and let $V_1\oplus V_2\oplus V_3$ be an orthogonal decomposition of $V_{nat}$ such that $\dim V_i= 3$ . Let $S=A_1^3$ be the the connected component of the stabilizer of this orthogonal decomposition. Then $N_G(S)$ is a maximal subgroup of G isomorphic to $(\mathbb {Z}_2^2\times S).Sym(3)$ . The group S acts homogeneously on V as the sum of two copies of the $8$ -dimensional irreducible S-module $\lambda _1\otimes \lambda _1\otimes \lambda _1$ . Let Y be the $1$ -dimensional variety of nontrivial proper S-submodules of V (i.e. the set of all $8$ -dimensional S-submodules of V). Since $p\neq 2$ , any such $8$ -space must be totally singular. We will now show that no element of $N_G(S)/S$ acts trivially on Y. Let $\tau $ be a pre-image under the the canonical projection $N_G(S) \rightarrow S$ of one of the $3$ nontrivial reflections in $(\mathbb {Z}_2^2\times S)/S$ . Then $\tau $ lies in $N_G(A_1D_3)$ , acting as a graph automorphism on $D_3$ . Now, $A_1D_3$ acts on V as $(\lambda _1\otimes \lambda _2) \oplus (\lambda _1\otimes \lambda _3)$ , and therefore, $\tau $ swaps these two $8$ -spaces. Similarly, a $2$ -cycle $\tau \in (S.Sym(3))/S$ corresponds to an element in the $D_3$ component of $A_1D_3$ acting as a graph automorphism of $A_1^2\leq D_3$ . Here, $\tau $ does fix $(\lambda _1\otimes \lambda _2)$ and $ (\lambda _1\otimes \lambda _3)$ , although they are not isomorphic $S\langle \tau \rangle $ -modules. A $3$ -cycle $\tau \in (S.Sym(3))/S$ also acts nontrivially on Y since it is a product of two $2$ -cycles that do not have the same fixed points. Finally, no product of a reflection with a transposition can act trivially, again because they do not fix the same points. Now by Lemma 2.10, there is a dense subset $\hat {Y}$ of Y on which no element of $N_G(S)/S$ has fixed points. Furthermore, there are only three proper connected subgroups of G that properly contain S, all isomorphic to $A_1D_3$ acting on V as $(\lambda _1\otimes \lambda _2) \oplus (\lambda _1\otimes \lambda _3)$ . Any such $A_1D_3$ only fixes two $8$ -spaces, which are not contained in Y since they are the fixed points of the $2$ -cycles in $Sym(3)$ . Let $y\in \hat {Y}$ . We have shown that $G_y = S$ , and therefore, $\mathrm {Tran}_G(y,Y) = N_G(S)$ . We then get $\dim G - \dim \mathrm {Tran}_G(y,Y) = 27 = \dim \mathcal {S}^{\prime \prime }_8(V) -\dim Y$ . Thus, the set $\hat {Y}$ is Y-exact, and by Lemma 2.15, we conclude that $C_{\mathcal {S}^{\prime \prime }_8(V)}=A_1^3$ .

Proposition 5.18. Let $G = B_4$ , $\lambda = \lambda _4$ with $p=2$ . Then the quadruple $(G,\lambda ,p,8")$ has no generic $ts$ -stabilizer, but has a semi-generic $ts$ -stabilizer $A_1^3$ .

Proof. Given the standard parabolic $P_1=U_7B_3T_1 = U_7L$ , let $X\leq L'$ be a subgroup isomorphic to an $A_1^3$ acting as $2\perp 2\perp 2$ on $V_{B_3}(\lambda _1)$ . Here, L acts on the abelian unipotent radical $U_7$ by fixing the longest short-root subgroup $X_{1111}$ and as $V_{B_3}(\lambda _1)$ on $U_7/X_{1111}$ . Then X has a $3$ -dimensional $1$ -cohomology on $U_7$ , corresponding to the conjugacy classes of $A_1^3$ -subgroups of $U_7X$ . We can parametrise this by pairing the root subgroups generating X with the highest and lowest weight vectors for the action on $U_7/X_{1111}$ . More precisely, take

$$\begin{align*}X = \langle X_{\pm 0111},X_{\pm 0011},X_{\pm 0001}\rangle\end{align*}$$

and define

$$ \begin{align*} A_1^{(1)}(\lambda)&:=\langle x_{0111}(t)x_{1222}(\lambda t),x_{-0111}(t)x_{1000}(\lambda t) \rangle_{t\in K},\\ A_1^{(2)}(\lambda)&:=\langle x_{0011}(t)x_{1122}(\lambda t),x_{-0011}(t)x_{1100}(\lambda t) \rangle_{t\in K},\\ A_1^{(3)}(\lambda)&:=\langle x_{0001}(t)x_{1112}(\lambda t),x_{-0001}(t)x_{1110}(\lambda t) \rangle_{t\in K},\\ X_{abc} &:= \langle A_1^{(1)}(a),A_1^{(2)}(b),A_1^{(3)}(c) \rangle.\end{align*} $$

Each $A_1^{(i)}(\lambda )$ is a connected subgroup of $P_1$ of type $A_1$ . Furthermore, $A_1^{(i)}(\lambda )$ and $A_1^{(j)}(\mu )$ commute if $i\neq j$ , which means that $X_{abc}$ is isomorphic to $A_1^3$ . We can write the given generators for $A_1^{(i)}(\lambda )$ in a nice compact form in the Clifford algebra – namely,

$$\begin{align*}A_1^{(i-1)}(\lambda) = \langle (1+t e_ie_5) (1+t e_if_5)(1+\lambda t e_ie_1),(1+t f_ie_5) (1+t f_if_5)(1+\lambda t f_ie_1)\rangle_{t\in K},\end{align*}$$

where $i\in \{2,3,4\}$ . Then $\mathcal {C}:=\{X_{abc}\}_{a,b,c\in K}$ is a set of representatives for the conjugacy classes of $A_1^3$ -subgroups of $U_7X$ .

Now consider an arbitrary $X_{abc}$ . We proceed to show that $X_{abc}$ acts homogeneously on V as a sum of two irreducible $8$ -spaces. Since $P_1$ fixes the (totally singular) $8$ -space

$$\begin{align*}V_1:= \left\langle \begin{array}{lll} e_1e_2, &e_1e_3, & e_1e_4,\\ e_1e_5,& e_1e_2e_3e_5,&e_1e_2e_4e_5,\\e_1e_3e_4e_5,&e_1e_2e_3e_4 \end{array}\right\rangle,\end{align*}$$

so does $X_{abc}$ . Secondly, let

$$\begin{align*}W_{abc}:= \left\langle \begin{array}{lll} e_1e_3e_4e_5+e_3e_4,& (a+1)e_1e_2e_3e_4+e_2e_3e_4e_5,&e_2e_3+(a+b+1)e_1e_2e_3e_5,\\e_3e_5+(1+b)e_1e_3,& (1+a+c)e_1e_2e_4e_5+e_2e_4,&(c+1)e_1e_4+e_4e_5,\\1+(b+c+1)e_1e_5, &e_2e_5+(1+a+b+c)e_1e_2 \end{array}\right\rangle.\end{align*}$$

A simple check using the generators of $X_{abc}$ shows that $W_{abc}$ is fixed by $X_{abc}$ and it is isomorphic to $\lambda _1\otimes \lambda _1\otimes \lambda _1$ as an $A_1^3$ -module. Since $p=2$ , it is not guaranteed that $W_{abc}$ is totally singular. Indeed, $W_{abc}$ is totally singular if and only if $a+b+c=0$ , by a direct check using Lemma 5.13. We now consider the subset $\mathcal {C}^*$ of $\mathcal {C}$ given by triples $(a,b,c)$ with $a+b+c=0$ such that $a,b,c$ are all distinct. Under these conditions on $(a,b,c)$ , we know that an element of $\mathcal {C}^*$ acts homogeneously on V as a sum of two totally singular $8$ -spaces and acts indecomposably on $V_{B_4}(\lambda _1)$ as $1/(2\perp 2\perp 2)/1$ .

We will now show that $X_{abc}\in \mathcal {C}^*$ is the connected component of the stabilizer in G of $W_{abc}$ . Since X acts indecomposably as $(2\perp 2 \perp 2) /1$ on $U_7$ , so does $X_{abc}$ . In particular, recall that $X_{1111}$ is fixed by $X_{abc}$ . Therefore, if u is a nontrivial element in $U_7$ , we must have $X_{1111} \cap \langle u,X_{abc} \rangle \neq 1$ . A direct check shows that no nontrivial element of $X_{1111}$ stabilises $W_{abc}$ , implying $(U_7)_{W_{abc}} = 1$ . Let M be a minimal connected overgroup of $X_{abc}$ , such that $M\leq G_{W_{abc}}$ . If M has a larger projection $\overline {M}$ onto $L'$ than $X_{abc}$ , it means that either $\overline {M}=A_1B_2$ or $\overline {M}=L'$ . In the latter case, $M=U_7L'$ , which is absurd; therefore, assume that $\overline {M}=A_1B_2$ . Without loss of generality, take $\overline {M} = \langle X_{\pm 0111},X_{\pm 0010},X_{\pm 0001}\rangle .$ Let $u\in U_7$ and $\kappa \in K^*$ such that $ux_{0010}(1)h_{\alpha _1}(\kappa )\in M$ . Since $M\cap U_7 = 1$ , we must have $[u,X_{0010}] =[u,h_{\alpha _1}(\kappa )] = 1$ . We cannot have $u\in X_{1222}$ since a direct check shows that $ux_{0010}(1)h_{\alpha _1}(\kappa )$ does not fix $W_{abc}$ ; so we must have $\kappa = 1$ and $u\in \langle X_{1000},X_{1110},X_{1111},X_{1122},X_{1222}\rangle $ . Similarly, since

$$\begin{align*}[X_{0010},A_1^{(3)}(c)] =[X_{0010},A_1^{(1)}(a)]=1,\end{align*}$$

we also get $[u,X_{\pm 0011}] =[u,X_{\pm 0111}] =1$ . Thus, $u\in \langle X_{1110},X_{1111}\rangle $ . Now assume that $x:=x_{1110}(t_1)x_{1111}(t_2)x_{0010}(1)$ fixes $W_{abc}$ . Since $x.(e_1e_3e_4e_5+e_3e_4)\in W_{abc}$ , we find that $t_1=t_2=0$ . Therefore, $x = x_{0010}(1)$ . Since $x.(e_4e_5+(c+1)e_1e_4)\in W_{abc}$ , we find that $b=c$ , which is absurd by our choice of $(a,b,c)$ . This completes the proof that $(G_{W_{abc}})^0=X_{abc}$ when $a,b,c$ are all distinct.

The radical of the Levi $B_3T_1$ acts by scalar multiplication on $(a,b,c)$ . Therefore, the subset $\mathcal {C}^{**}$ of $\mathcal {C}^*$ defined by the further condition $a=1$ contains $A_1^3$ -subgroups which are pairwise non-conjugate in $U_7XT_1$ . Let $Y = \{W_{abc}:X_{abc}\in \mathcal {C}^{**} \}$ , a $1$ -dimensional variety of totally singular $8$ -spaces. By [Reference Stewart31, Prop. 3.5.2 (D)], G-fusion of elements of $\mathcal {C}^{**}$ is controlled by $N_{L'}(X)/X\simeq Sym(3)$ . We can be even more precise, and like in [Reference Stewart31, Lemma 4.1.3] deduce that if $X_{abc}^g = X_{a'b'c'}$ , then $g\in U_7N_{B_3T_1}(X)$ . Therefore, $X_{abc}$ is G-conjugate to $X_{a'b'c'}$ if and only if $(a',b',c') = t(\pi (a),\pi (b),\pi (c))$ for some $t\in K$ and $\pi \in Sym(\{a,b,c\})$ . Also, if $a,b,c$ are pairwise distinct and are not of the form $a,\mu a,\mu ^2 a$ where $\mu $ is a root of $x^2+x+1$ , we must have $N_G(X_{abc}) = U_1X_{abc}$ .

Therefore, there is a dense subset $\hat {Y}$ of Y such that any two distinct elements in $\hat {Y}$ have non-conjugate stabilizers in G, isomorphic to $A_1^3$ . Let $y\in \hat {Y}$ . Then by construction, $A_1^3\leq \mathrm {Tran_G}(y,Y) \leq A_1^3.Sym(3)$ , and by dimensional considerations, $\hat {Y}$ is Y-exact. By Lemma 2.16, we conclude that there is no generic stabilizer, and by Lemma 2.14, we conclude that there is a semi-generic stabilizer isomorphic to $A_1^3$ .

There are now two cases left in order to complete the proof of Theorem 2. These are given by the $ts$ -small quadruples $(C_2,2\lambda _1,p,5)$ ( $p\neq 2$ ) and $(C_3,\lambda _2,p,7)$ ( $p\neq 3$ ). These two cases present considerable challenges and similarities to each other. They are the subject of the next two sections.

5.1 The case $(C_2, 2\lambda _1,p,5)$

In this section, we handle the case of $C_2$ acting on maximal totally singular subspaces of its adjoint module. We shall prove that this action has a dense orbit with finite generic stabilizer. We resort to making extensive use of computational methods in Magma, with the relevant code being listed in Appendix A as well as being made available on the author’s GitHub [Reference Rizzoli28].

Suppose that $p\neq 2$ . Let $G=Sp_4(K)$ , with fundamental roots $\alpha _1,\alpha _2$ , where $\alpha _1$ is short. Let $\alpha _3 = \alpha _1+\alpha _2$ and $\alpha _4 = 2\alpha _1+\alpha _2$ . Order the standard basis of the natural module $V_{nat}$ as $(e_1,e_2,f_2,f_1)$ and let $V = \mathrm {Lie}(G)\leq \mathfrak {sl}_4(K)$ , on which G is acting by conjugation. Let $e_{\pm \alpha _1},e_{\pm \alpha _2},e_{\pm \alpha _3},e_{\pm \alpha _4},h_{\alpha _1},h_{\alpha _2}$ be the corresponding Chevalley basis, where $e_{\alpha _1},e_{ \alpha _2},e_{\alpha _3},e_{\alpha _4}$ are respectively the matrices

$$\begin{align*}\left(\begin{matrix} &1 & \\ & &0 \\ & & & -1\\ & & \\ \end{matrix}\right),\, \left(\begin{matrix} &0 & \\ & &1 \\ & & & 0\\ & & \\ \end{matrix}\right),\,\left(\begin{matrix} & & 1 \\ & & &1 \\ & & \\ & & \\ \end{matrix}\right),\,\left(\begin{matrix} & & & 1\\ & & \\ & & \\ & & \\ \end{matrix}\right),\end{align*}$$

the elements $e_{-\alpha _1},e_{-\alpha _2},e_{-\alpha _3},e_{-\alpha _4}$ are respectively their transposes, and

$$\begin{align*}h_{\alpha_1} = \operatorname{\mathrm{diag}}(1,-1,1,-1)\, ,h_{\alpha_2} = \operatorname{\mathrm{diag}}(0,1,-1,0),h_{\alpha_3} = \operatorname{\mathrm{diag}}(1,1,-1,-1), h_{\alpha_4} = \operatorname{\mathrm{diag}}(1,0,0,-1).\end{align*}$$

Let T be the standard maximal torus of G. The module V is orthogonal, with quadratic form given by

$$\begin{align*}Q(v)=\mathrm{Trace}(v^2).\end{align*}$$

For i a square root of $-1$ , and $\zeta $ a square root of $-2$ , let $W_{(i,\zeta )}$ be the totally singular $5$ -space of V spanned by

$$ \begin{align*} v^{(0)} &= h_{\alpha_4}+ih_{\alpha_2}, \\ v^{(1)} &=e_{\alpha_1}+\zeta e_{\alpha_2}, \\ v^{(2)} &= e_{\alpha_3}+\zeta e_{-\alpha_4}, \\ v^{(3)} &= e_{-\alpha_3}+\zeta e_{\alpha_4}, \\ v^{(4)} &= e_{-\alpha_1}+\zeta e_{-\alpha_2}. \end{align*} $$

Let $W_{5}$ be the $5$ -space of V spanned by

$$ \begin{align*} u^{(0)} &= h_{\alpha_4}+2h_{\alpha_2}, \\ u^{(1)} &=e_{\alpha_1}+3 e_{\alpha_2}, \\ u^{(2)} &= e_{\alpha_2}+3 e_{-\alpha_4}, \\ u^{(3)} &= e_{-\alpha_3}, \\ u^{(4)} &= e_{-\alpha_1}+e_{-\alpha_2}+3e_{\alpha_4}, \end{align*} $$

a totally singular subspace if $p=5$ .

Let

$$\begin{align*}\tau = \left(\begin{matrix} 0& 1& 0 & 0\\ 0& 0& 0& -1\\ 1 & 0& 0& 0\\ 0& 0 & 1 & 0\\ \end{matrix}\right), \text{ and } x= \begin{cases} \operatorname{\mathrm{diag}} (\omega^{-1},\omega^{-2},\omega^2,\omega) \text{ with } \omega^5 = 1,\omega\neq 1,& \text{if } p\neq 5;\\ \begin{pmatrix} 1&2 & 1 & 1\\ & 1&1 &4 \\ & & 1&3 \\ & & & 1\\ \end{pmatrix},& \text{if } p = 5. \end{cases} \end{align*}$$

Furthermore, let

$$\begin{align*}\tau^*= \begin{cases} \tau,& \text{if } p\neq 5;\\ \operatorname{\mathrm{diag}} (\alpha,2\alpha,4\alpha,3\alpha),\text{ where } \alpha^2 = 2, & \text{if } p = 5. \end{cases} \end{align*}$$

Finally, let

$$\begin{align*}S^* = \langle x,\tau^* \rangle, \text{ and } W^*= \begin{cases} W_{(i,\zeta)},& \text{if } p\neq 5;\\ W_5,& \text{if } p = 5. \end{cases} \end{align*}$$

With this setup, it is easy to check that $S^* \leq G_{W^*}$ .

Proposition 5.19. Let $G=C_2$ , $\lambda = 2\lambda _1$ with $p = 5$ . Then $C_{\mathcal {S}_5'(V)}=\mathbb {Z}_4$ .

Proof. We use the setup of [Reference Guralnick and Lawther10, Lemma 4.8(i)] and its proof. Let $h_0 = \operatorname {\mathrm {diag}} (-1,-2,2,1)$ , a regular semisimple element of $\mathrm {Lie}(T)$ , and set $\mathfrak {G} = \langle h_0 \rangle $ . For a subspace U of $\mathrm {Lie}(G)$ , write $\mathrm {Ann}_{\mathrm {Lie}(G)}(U)$ for the subspace $\{v\in \mathrm {Lie}(G):[v,U]\leq U\}.$ A straightforward calculation shows that $\mathrm {Ann}_{\mathrm {Lie}(G)}(W_{(3,\zeta )}) = \mathfrak {G}$ . Let $S = G_{W_{(3,\zeta )}}$ and take $g\in S$ . We have $\mathfrak {G} = \mathrm {Ann}_{\mathrm {Lie}(G)}(g.W_{(3,\zeta )})=g.\mathfrak {G}$ . Therefore, $g.\mathfrak {G} = \mathfrak {G}$ , which is easily seen to imply $g\in T.\langle \tau \rangle .$ A direct calculation shows that $T\cap S = \pm 1$ , which implies that $S = \langle \tau \rangle = Z(G).\mathbb {Z}_4$ . Since $\dim G - \dim S = \dim G = \dim \mathcal {S}_5'(V)$ , we conclude that $C_{\mathcal {S}_5'(V)}=\mathbb {Z}_4$ .

Remark 5.20. Note that the subspace $W_{2,\zeta }$ does not belong to the same $D_5$ -orbit as $W_{3,\zeta }$ by Lemma 2.6; however, it also does not have a finite stabilizer. Indeed, it is not difficult to see that it has a stabilizer isomorphic to $U_3T_2$ .

The following lemma describes the subgroup structure of $Sp_4(q)$ .

Lemma 5.21. [Reference Bray, Holt and Roney-Dougal5, §8.2].

Assume that $p<\infty $ and let $q = p^e$ for some $e\in \mathbb {Z}_{\geq 1}$ . Then the maximal subgroups of $Sp_4(q)$ (q odd) are as in Table 7, and the maximal subgroups of $SL_2(q)$ are as in Table 8. In both cases, see [Reference Bray, Holt and Roney-Dougal5] for more details, including the precise notation.

Table 7 Maximal subgroups $M_q$ of $Sp_4(q)$ (q odd).

Table 8 Maximal subgroups $M_q$ of $SL_2(q)$ .

Lemma 5.22. Assume $p=5$ . Let $H = G_{\langle u^{(1)} \rangle }$ . Then $H_{W^*} = S^* \simeq Z(G).\mathbb {Z}_5.\mathbb {Z}_4$ .

Proof. The element $u^{(1)}$ is a regular nilpotent element, and a simple calculation shows that $H = U_2T_1$ , where

$$\begin{align*}U_2 = \left\{\left(\begin{matrix} 1&-b &-b^2 & a \\ & 1& 2b &b^2 \\ & & 1&b \\ & & & 1\\ \end{matrix}\right): a,b\in K\right\},\,T_1 = \{ \operatorname{\mathrm{diag}}(\kappa^3,\kappa,\kappa^{-1},\kappa^{-3}):\kappa \in K^*\}.\end{align*}$$

Let $g =\left (\begin {smallmatrix} 1&-b &-b^2 & a \\ & 1& 2b &b^2 \\ & & 1&b \\ & & & 1\\ \end {smallmatrix}\right )\left (\begin {smallmatrix} \kappa ^3&& \\ & \kappa & \\ & &\kappa ^{-1} \\ & & & \kappa ^{-3}\\ \end {smallmatrix}\right )\in H_{W^*}$ . We have $g u^{(0)} g^{-1} = u^{(0)}-b u^{(1)} +(b^3+3a)e_{\alpha _4}$ . This forces $a = 3b^3$ . Also, $gu^{(3)}g^{-1} = \frac {1}{\kappa ^4}(u^{(3)}-b^2 u^{(0)}+2b^3 u^{(1)}+2b u^{(4)}+(b^5-b)e_{\alpha _4 })$ , forcing $b=b^5$ . Finally, $gu^{(2)}g^{-1}= \frac {1}{\kappa ^6} (-b^3 u^{(0)}+b(1-\kappa ^8)e_{\alpha _3} +u^{(2)}+3b u^{(3)}+3b^2 u^{(4)}-b^4 u^{(1)}+(\kappa ^8-1)e_{\alpha _2}+b^2(\kappa ^8-1)e_{\alpha _4})$ , implying $\kappa ^8 = 1$ . This allows us to conclude that $H_{W^*} = \langle x , \tau ^* \rangle =S^*$ , as claimed.

Lemma 5.23. Let $H = N_G(\langle x \rangle )$ . Then $H_{W^*}= S^* \simeq Z(G).\mathbb {Z}_5.\mathbb {Z}_4$ .

Proof. Assume $p\neq 5$ . Since x is a regular semisimple element, it is easy to see that $H= T.\langle \tau \rangle $ , and one quickly finds that $T_{W^*} = \pm \langle x\rangle $ . Since $\tau \in G_{W^*}$ , we conclude that $H_{W^*} = \langle x,\tau \rangle = Z(G).\langle x \rangle .\mathbb {Z}_4$ . If $p=5$ , we have $H\leq \langle C_G(x),T \rangle $ . It is easy to see that $C_G(x)$ is the unipotent radical of $G_{\langle u^{(1)}\rangle }$ and that

$$\begin{align*}N_T(\langle x \rangle) = \langle \tau^* \rangle \leq \{\operatorname{\mathrm{diag}}(\kappa^3,\kappa,\kappa^{-1},\kappa^{-3}):\kappa \in K^*\}.\end{align*}$$

Therefore, $H\leq G_{\langle u^{(1)} \rangle }$ , and we can conclude by Lemma 5.22.

Lemma 5.24. Suppose that $H\leq G$ is a reducible subgroup of G containing $\langle x \rangle $ . Then $H_{W^*}\leq S^*$ .

Proof. If $p\neq 5$ , the only $1$ -spaces of $V_{nat}$ stabilised by the semisimple element x are spanned by a standard basis vector; therefore, $H\leq G_U$ , where $U\leq V_{nat}$ is a $1$ -space, a totally singular $2$ -space or a nondegenerate $2$ -space, in each case spanned by standard basis vectors. However, if $p=5$ , then $H\leq G_U$ , where $U = \langle e_1\rangle $ or $U=\langle e_1,e_2\rangle $ , as x is a regular unipotent element contained in the standard Borel subgroup. In all the cases where U is totally singular (i.e., $G_U\simeq P_1$ or $G_U\simeq P_2$ ), it is easily seen that $G_U$ stabilises a unique $6$ -space of V. We intersect this $6$ -space with $W^*$ , identifying a $1$ -space spanned by a regular nilpotent element v that must be stabilised by $H_{W^*}$ . This then reduces the problem to computing the stabilizer of $W^*$ within a $U_2T_1$ . If $p=5$ , we find that $\langle v \rangle = \langle u^{(1)} \rangle $ , concluding by Lemma 5.22.

Therefore, from now on, assume that $p\neq 5$ . If $H = G_{\langle e_1 \rangle }$ , then H stabilises $\langle e_{\pm \alpha _2},e_{\alpha _1},e_{\alpha _3},e_{\alpha _4 },h_{\alpha _2}\rangle $ , which intersects $W^*$ in $\langle v^{(1)} \rangle $ . We find that $G_{\langle v^{(1)} \rangle } = U_2T_1$ , where

$$\begin{align*}U_2 = \left\{\left(\begin{matrix} 1&b &\frac{b^2\zeta}{2} & a \\ & 1& \zeta b &-\frac{b^2\zeta}{2} \\ & & 1&-b \\ & & & 1\\ \end{matrix}\right): a,b\in K\right\},\,T_1 = \{ \operatorname{\mathrm{diag}}(\kappa^3,\kappa,\kappa^{-1},\kappa^{-3}):\kappa \in K^*\}.\end{align*}$$

A direct calculation like in the proof of Lemma 5.22 then shows that $(U_2T_1)_{W^*}= \pm \langle x \rangle $ . The other cases with $G_U\simeq P_1$ or $G_U\simeq P_2$ are similarly dealt with. Here, we report just the intersection $\langle v \rangle $ . If $U = \langle e_2 \rangle $ , then $\langle v \rangle = \langle v^{(2)}\rangle $ ; if $U = \langle f_1 \rangle $ , then $\langle v \rangle = \langle v^{(4)}\rangle $ ; if $U = \langle f_2 \rangle $ , then $\langle v \rangle = \langle v^{(3)}\rangle $ ; if $U = \langle e_1,e_2 \rangle $ , then $\langle v \rangle = \langle v^{(1)}\rangle $ ; if $U = \langle e_1,f_2 \rangle $ , then $\langle v \rangle = \langle v^{(3)}\rangle $ ; if $U = \langle e_2,f_1 \rangle $ , then $\langle v \rangle = \langle v^{(4)}\rangle $ .

It remains to consider the case $H\leq G_{\langle e_1,f_1 \rangle }$ . Here, H fixes the subspace $\langle h_{\alpha _1},h_{\alpha _2},e_{\pm \alpha _2},e_{\pm \alpha _4 } \rangle $ , which intersects $W^*$ in $\langle v^{(0)}\rangle $ . Therefore, $H_{W^*}\leq N_G(T)$ , and we conclude by Lemma 5.23.

Lemma 5.25. Suppose that $\langle x \rangle \leq H.\mathbb {Z}_2 < G$ , where H is an arbitrary subgroup of G. Then $(H.\mathbb {Z}_2)_{W^*}\leq S^*$ if and only if $H_{W^*}\leq S^*$ .

Proof. The forward direction is trivial. Suppose that $H_{W^*}\leq \langle x , \tau ^* \rangle = S^*$ . Since x has order $5$ , we must have $\langle x\rangle \leq H$ . Since $H_{W^*}\leq S^*$ , the subgroup $\langle x \rangle $ is the unique subgroup of order $5$ in $H_{W^*}$ . Therefore, since $H_{W^*} \lhd (H.\mathbb {Z}_2)_{W^*}$ , the subgroup $\langle x \rangle $ is normal in $(H.\mathbb {Z}_2)_{W^*}$ . By Lemma 5.23, we know that the stabilizer of $W^*$ in $N_G(\langle x\rangle )$ is $S^*$ , concluding.

Proposition 5.26. Let $G=C_2$ , $\lambda = 2\lambda _1$ with $p\neq 2$ . Then $C_{\mathcal {S}_5'(V)}=\mathbb {Z}_5.\mathbb {Z}_4$ and $C_{\mathcal {S}_5"(V)} = \mathbb {Z}_{5/(p,5)}.\mathbb {Z}_4$ .

Proof. Let $S = G_{W^*}$ . We shall prove that $S\leq N_G(\langle x \rangle )$ . This will conclude the proof of the proposition as follows. By Lemma 5.23, we have $(N_G(\langle x \rangle )_{W^*} = S^*$ , which then implies $S=S^*$ and that $W^*$ is in a dense G-orbit on one of the two $D_5$ -orbits on $\mathcal {S}_5(V)$ . If $p\neq 5$ , then $W_{(i,\zeta )}$ and $W_{(-i,\zeta )}$ intersect in a $4$ -dimensional subspace, and therefore, by Lemma 2.6, they belong to distinct $D_5$ -orbits on $\mathcal {S}_5(V)$ . They each have stabilizer $S^*$ , concluding the $p\neq 5$ case. If $p=5$ , the subspace $W^*$ intersects $W_{(3,\zeta )}$ trivially, which by Lemma 2.6 implies that $W^*$ and $W_{(3,\zeta )}$ belong to distinct $D_5$ -orbits on $\mathcal {S}_5(V)$ . Again, $G_{W^*} = S^*$ , concluding.

In order to prove that $S\leq N_G(\langle x \rangle )$ , we show that for all $p<\infty $ and $e\in \mathbb {Z}_{\geq 1}$ , if

$$\begin{align*}\langle x \rangle \leq R \leq Sp_4(p^e)=Sp_4(q)<G, \text{ with }R\not\leq N_G(\langle x \rangle),\end{align*}$$

then R does not stabilise $W^*$ . Note that this is indeed sufficient, since if $g\in S\setminus N_G(\langle x \rangle )$ , then there must exist $e\in \mathbb {Z}_{\geq 1}$ such that $g \in Sp_4(p^e)$ , with $R = \langle x,g\rangle $ satisfying the condition above. We shall make extensive use of maximal subgroups of $Sp_4(q)$ , often combined with exhaustive computations in Magma. The commented code is made available both in the Appendix as well as on the author’s Github [Reference Rizzoli28]. The $p=\infty $ case then follows from the $p<\infty $ case since they are equivalent once p is sufficiently large.

Suppose that

$$\begin{align*}\langle x \rangle \leq R \leq M_q < Sp_4(q)<G, \text{ with }R\not\leq N_G(\langle x \rangle),\end{align*}$$

where $M_q$ is a maximal subgroup of $Sp_4(q)$ , as listed in Table 7. The goal is to prove that R does not stabilize W.

If $M_q$ is as in one of the first $6$ rows of Table 7, then by Lemma 5.25, we can assume that R is reducible, and Lemma 5.24 implies that R does not stabilize W. If $M_q = Sp_4(q_0).(2,r)$ where r is prime and $q = q_0^r$ , then Lemma 5.25 allows us to reduce to one of the other cases.

Suppose that $M_q$ is the double cover of $Alt(6)$ , $Sym(6)$ or $Alt(7)$ , in which case $q = p$ . An exhaustive search using Magma shows that $Z(G).\langle x \rangle \leq R^* \leq R$ where $R^* = Z(G).Alt(5)$ , the double cover of $Alt(5)$ , isomorphic to $SL_2(5)$ . The general strategy adopted for this exhaustive search is the following. We set up R as an abstract group. Then for all conjugacy classes of elements of order $5$ of R, we take a representative $g_5$ and go through all subgroups of R that contain $g_5$ , determining which ones do not normalise $\langle g_5\rangle $ . By Lemma 5.24, we can assume that $R^*$ is irreducible in G. If $p\neq 5$ , since $p\neq 2$ by assumption and $p\neq 3$ by choice of $M_q$ , we can use ordinary character theory to show that $R^*$ does not fix any $5$ -space of V. The subgroup $R^*$ must be embedded in G via its unique irreducible symplectic character $\chi $ of degree $4$ . We then verify that $S^2(\chi ) = \chi _4+\chi _5+\chi _7$ , where $\chi _4,\chi _5,\chi _7$ are irreducible characters of degrees $3,3,4$ . Therefore, $R^*$ fixes no $5$ -space of V. See Listing 3 for the corresponding Magma code. If $p = 5$ , we can use a direct construction of $2.Alt(6)\leq Sp_4(5)$ to check that $R^*$ acts on V with composition factors of dimensions $3,3,3,1$ . Again, this means that R fixes no $5$ -space of V. See Listing 3 for the Magma code proving this.

Suppose that $M_q = Z(G).\mathbb {Z}_2^4.Sym(5)$ , the normalizer of an extraspecial subgroup of G of minus type. Similarly to the previous case, an exhaustive search shows that R must contain $R^* = Z(G).\mathbb {Z}_2^4.\langle x \rangle $ or $R^* = Z(G).Alt(5)$ . By Lemma 5.24, we can assume that $R^*$ is irreducible in G. In the second case, we have already seen that $R^*$ does not fix a $5$ -space of V when $p\neq 3$ . If $p=3$ , the same holds, which can be checked directly in $Sp_4(3)$ by taking an explicit construction of $Z(G).Alt(5)$ . Therefore, assume that $R^* = Z(G).\mathbb {Z}_2^4.\langle x \rangle $ , a group with GAP Id $(160,199)$ . If $p\neq 5$ , we can use the ordinary characters of $R^*$ to check that $R^*$ does not fix any totally singular $5$ -spaces of V. The subgroup $R^*$ must be embedded in G via its unique irreducible symplectic character $\chi $ of degree $4$ . One then finds that $S^2(\chi ) = \psi _1+\psi _2$ , where $\psi _1$ and $\psi _2$ are distinct self-dual irreducible characters of degree $5$ . Therefore, $R^*$ stabilises exactly two nondegenerate $5$ -spaces of V. See Listing 2 for the Magma code. If $p=5$ , we can use a direct construction of $\mathbb {Z}_2.\mathbb {Z}_2^4.Alt(5)\leq Sp_4(5)$ to show that the $KR^*$ -module $V\downarrow R^*$ has two self-dual non-isomorphic composition factors, implying that $R^*$ does not stabilise a totally singular $5$ -space. See details of the computations in Listing 4.

It remains to consider the case $M_q = SL_2(q)$ with $p\geq 5$ and $q\geq 7$ , as in the last row of Table 7. In order to handle this case, we consider the subgroup structure of the maximal subgroups of $SL_2(q)$ , as classified in Table 8. First note that $SL_2(q)$ does not fix a $5$ -space of V, as if $p = 5$ it acts on V with composition factors of dimensions $4,3,3$ , while if $p>5$ , it acts on V with composition factors of dimension $7$ and $3$ . Therefore, assume that $R\leq M_q^* < SL_2(q) $ , where $M_q^*$ is a maximal subgroup of $SL_2(q)$ , as described by Table 8. If $M_q^*$ is as in one of the first five rows of Table 8, then by Lemma 5.25, we can assume that R is reducible, and Lemma 5.24 implies that R does not fix $W^*$ . If $M_q^* = SL_2(q_0).(2,r)$ where r is prime and $q = q_0^r$ , then Lemma 5.25 allows us to reduce to one of the other cases. Since $p\neq 2$ , the case $M_q^* = PSL_2(q_0) $ is excluded, while since x has order $5$ , the cases $M_q^* = \mathbb {Z}_2.\mathbb {Z}_2^2.Sym(3)$ and $M_q^* = \mathbb {Z}_2.\mathbb {Z}_2^2.\mathbb {Z}_3$ are not possible. The only other possibility is $M_q^* = SL_2(5)$ with $p\neq 5$ , as per the last two rows of Table 8. In this case, we must have $R = M_q^* = SL_2(5)$ , which we have already dealt with. This completes the case-by-case analysis.

5.2 The case $(C_3, \lambda _2,p,7)$

In this section, we handle the last remaining case needed to complete the proof of Theorem 2. In particular, we shall prove that the $ts$ -small quadruples $(C_3, \lambda _2,p,7')$ and $(C_3, \lambda _2,p,7")$ have a finite generic stabilizer. The strategy is entirely similar to the one used for the $(C_2, 2\lambda _1,p,5)$ case. We shall, however, make even greater use of computational methods, sometimes resorting to solving large systems of equations using Magma. Again, the code can be found in Appendix A, as well as on the author’s GitHub [Reference Rizzoli28].

Suppose that $p\neq 3$ . Let $G=Sp_6(K)$ and order the standard basis of the natural module $V_{nat}$ as $(e_1,e_2,e_3,f_3,f_2,f_1)$ . Let V be the submodule of $\bigwedge ^2 V_{nat}$ defined by

$$\begin{align*}V=\langle e_i\wedge e_j,f_i\wedge f_j,e_i\wedge f_j, \sum \alpha_{i}e_i\wedge f_i : i\neq j, \sum \alpha_i=0 \rangle. \end{align*}$$

Then $V=V_{G}(\lambda _2)$ . Let $\omega $ be a primitive cube root of unity and let $(v_1,\dots ,v_{14})$ be the ordered basis of V given by

$$ \begin{align*} v_1&= e_1\wedge e_2, & v_5&=e_2\wedge f_3, & v_{10}&=e_3\wedge f_2, \\ v_2&=e_1\wedge e_3, & v_6&=e_1 \wedge f_2, &v_{11}&=e_3\wedge f_1, \\ v_3&=e_2\wedge e_3, &v_7&=e_1\wedge f_1+\omega e_2\wedge f_2 +\omega^2 e_3\wedge f_3, & v_{12}&=f_2\wedge f_3,\\ v_4&=e_1\wedge f_3, & v_8&=e_1\wedge f_1+\omega^2 e_2\wedge f_2 +\omega e_3\wedge f_3, & v_{13}&=f_1\wedge f_3,\\ & & v_9&=e_2\wedge f_1, & v_{14}&=f_1\wedge f_2.\end{align*} $$

Then it is easy to check that G fixes a nondegenerate quadratic form on V, given by

$$\begin{align*}Q\left(\sum_1^{14}\alpha_i v_i\right)= \sum_1^7 \alpha_i \alpha_{15-i}. \end{align*}$$

Let T be the standard maximal torus of G. For i a primitive fourth root of unity, let $W_{(\omega ,i)}$ be the totally singular $7$ -space of V spanned by

$$ \begin{align*} v^{(0)} &= e_1\wedge f_1+\omega e_2\wedge f_2 +\omega^2 e_3\wedge f_3, \\ v^{(1)} &=e_2\wedge f_3+i e_1\wedge f_2, \\ v^{(2)} &= e_1\wedge f_3-i e_2\wedge e_3, \\ v^{(3)} &= f_1\wedge f_2+i e_1\wedge e_3, \\ v^{(4)} &= e_1\wedge e_2+i f_1\wedge f_3, \\ v^{(5)} &= e_3\wedge f_1-i f_2\wedge f_3, \\ v^{(6)} &= e_3\wedge f_2+i e_2\wedge f_1. \end{align*} $$

Let $W_{7}$ be the $7$ -space of V spanned by

$$ \begin{align*} u^{(0)} &= e_1\wedge f_1+4 e_2\wedge f_2 +2 e_3\wedge f_3, \\ u^{(1)} &=e_1\wedge e_2 +3 e_3\wedge f_1+3 f_2\wedge f_3, \\ u^{(2)} &= e_2\wedge e_3+4 f_1\wedge f_2, \\ u^{(3)} &= e_1\wedge f_3 +4 f_1\wedge f_2, \\ u^{(4)} &= e_2\wedge f_3+2 e_1\wedge f_2,\\ u^{(5)} &= e_2\wedge f_1 +5 e_3 \wedge f_2, \\ u^{(6)} &= f_1\wedge f_3, \end{align*} $$

a totally singular subspace if $p=7$ . Let

$$\begin{align*}\tau = \left(\begin{matrix} 0& 0& 0 & 0 & -1 & 0\\ 0& 0& 1 & 0 & 0 & 0\\ 1& 0& 0 & 0 & 0 & 0\\ 0& 0& 0 & 0 & 0 & 1\\ 0& 0& 0 & 1 & 0 & 0\\ 0& 1& 0 & 0 & 0 & 0\\ \end{matrix}\right), \text{ and } x= \begin{cases} \operatorname{\mathrm{diag}} (\zeta^{-1},\zeta^{-2},\zeta^{-3},\zeta^{3},\zeta^2,\zeta) \text{ with } \zeta^7 = 1,\zeta\neq 1,& \text{if } p\neq 7;\\ \begin{pmatrix} 1& 1& 4 & 6 & 2 & 1\\ & 1& 1 & 4 & 1 & 5\\ & & 1 & 1 & 3 & 6\\ & & & 1 & 6 & 4\\ & & & & 1 & 6\\ & & & & & 1\\ \end{pmatrix},& \text{if } p = 7. \end{cases} \end{align*}$$

Furthermore, let

$$\begin{align*}\tau^*= \begin{cases} \tau,& \text{if } p\neq 7;\\ \operatorname{\mathrm{diag}} (\alpha,5\alpha,4\alpha,6\alpha,2\alpha,3\alpha),\text{ where } \alpha^2 = 5, & \text{if } p = 7. \end{cases} \end{align*}$$

Finally, let

$$\begin{align*}S^\dagger = \langle x,(\tau^*)^4 \rangle,\, S^* = \langle x,\tau^* \rangle, \text{ and } W^*= \begin{cases} W_{(\omega,i)},& \text{if } p\neq 7;\\ W_7,& \text{if } p = 7. \end{cases} \end{align*}$$

With this setup, it is easy to check that $S^\dagger \leq S^* \leq G_{W^*}$ .

Proposition 5.27. Let $G=C_3$ , $\lambda = \lambda _2$ with $p = 7$ . Then $C_{\mathcal {S}_7'(V)}=\mathbb {Z}_6$ .

Proof. We use the setup of [Reference Guralnick and Lawther10, Lemma 4.8(i)] and its proof. Let $h_0 = \operatorname {\mathrm {diag}} (-1,-2,-3,3,2,1)$ , a regular semisimple element of $\mathrm {Lie}(T)$ , and set $\mathfrak {G} = \langle h_0 \rangle $ . For a subspace U of $\mathrm {Lie}(G)$ , write $\mathrm {Ann}_{\mathrm {Lie}(G)}(U)$ for the subspace $\{v\in \mathrm {Lie}(G):[v,U]\leq U\}.$ A straightforward calculation shows that $\mathrm {Ann}_{\mathrm {Lie}(G)}(W_{(2,i)}) = \mathfrak {G}$ . Let $S = G_{W_{(2,i)}}$ and take $g\in S$ . We have $\mathfrak {G} = \mathrm {Ann}_{\mathrm {Lie}(G)}(g.W_{(2,i)})=g.\mathfrak {G}$ . Therefore, $g.\mathfrak {G} = \mathfrak {G}$ , which is easily seen to imply $g\in T.\langle \tau \rangle .$ A direct calculation shows that $T\cap S = \pm 1$ , which implies that $S = \langle \tau \rangle = Z(G).\mathbb {Z}_6$ . Since $\dim G - \dim S = \dim G = \dim \mathcal {S}_7'(V)$ , we conclude that $C_{\mathcal {S}_7'(V)}=\mathbb {Z}_6$ .

Lemma 5.28. Suppose that $H\leq G$ is a reducible subgroup of G containing $S^\dagger $ . Then $H_{W^*}\leq S^*$ .

Proof. Suppose that $p=7$ . Since $x\in S^\dagger $ is a regular unipotent element contained in the standard Borel subgroup B, we have that H is contained in $G_{\langle e_1 \rangle }$ , $G_{\langle e_1,e_2 \rangle }$ or $G_{\langle e_1,e_2,e_3 \rangle }$ . We consider each of these cases and deduce that $H_{W^*}\leq B$ . We then use Magma to directly show that $B_{W^*} = S^*$ , concluding as required. Suppose that $H\leq G_{\langle e_1,e_2 \rangle }$ . The group $G_{\langle e_1,e_2 \rangle }$ stabilises $U_6 = \langle e_1\wedge e_2,e_1\wedge e_3,e_2\wedge e_3,e_1\wedge f_3,e_2\wedge f_3, e_1\wedge f_1+e_2\wedge f_2 -2 e_3\wedge f_3\rangle $ , and therefore, $H_{W^*}$ must stabilise $W^* \cap U_6 = \langle e_1\wedge f_3-e_2\wedge e_3\rangle $ as well as $W^*\cap (U_6)^\perp = \langle e_1\wedge f_3-e_2\wedge e_3,e_2\wedge f_3 +2 e_1 \wedge f_2\rangle $ . The latter implies that $H_{W^*}$ stabilises $\langle e_1,e_2,e_3,f_3,f_2 \rangle $ and therefore also its radical $\langle e_1 \rangle $ . Let $g\in H_{W^*}$ . Since g stabilises $\langle e_1\rangle $ , $\langle e_1,e_2\rangle $ , $\langle e_1,e_2 \rangle ^\perp $ and $\langle e_1\wedge f_3-e_2\wedge e_3 \rangle $ , it is easy to see that $g.e_3 \in \langle e_1,e_2,e_3 \rangle $ . Therefore, $g\in B$ and $H_{W^*}\leq B$ . Now consider the case $H\leq G_{\langle e_1 \rangle }$ . Any element of H must also stabilise $\langle e_1 \rangle ^\perp = \langle e_1,e_2,e_3,f_3,f_2\rangle $ , and similarly to the previous case, we find that $H_{W^*}$ stabilises $\langle e_1\wedge f_3-e_2\wedge e_3,e_2\wedge f_3 +2 e_1 \wedge f_2\rangle $ . Let $g = (a_{ij})_{ij} \in H_{W^*}$ . We have $g.(e_1\wedge f_3-e_2\wedge e_3) = a_{11}e_1\wedge (a_{14}e_1+a_{24}e_2+a_{34}e_3+a_{44}f_3+a_{54}f_2)-(a_{12}e_1+a_{22}e_2+a_{32}e_3+a_{42}f_3+a_{52}f_2)\wedge (a_{13}e_1+a_{23}e_2+a_{33}e_3+a_{43}f_3+a_{53}f_2)$ . Since $g.(e_1\wedge f_3-e_2\wedge e_3) = \gamma _1(e_1\wedge f_3-e_2\wedge e_3) + \gamma _2(e_2\wedge f_3 +2 e_1 \wedge f_2)$ for some $\gamma _1,\gamma _2\in K$ , we must have

$$\begin{align*}\det \begin{pmatrix} a_{22} & a_{23} \\ a_{52} & a_{53} \end{pmatrix} = \det \begin{pmatrix} a_{32} & a_{33} \\ a_{42} & a_{43} \end{pmatrix} = \det \begin{pmatrix} a_{32} & a_{33} \\ a_{52} & a_{53} \end{pmatrix} =\det \begin{pmatrix} a_{42} & a_{43} \\ a_{52} & a_{53} \end{pmatrix} = 0.\end{align*}$$

If $(a_{52},a_{53}) \neq (0,0)$ , we get

$$\begin{align*}\gamma_1= \det \begin{pmatrix} a_{22} & a_{23} \\ a_{32} & a_{33} \end{pmatrix} =0= \det \begin{pmatrix} a_{22} & a_{23} \\ a_{42} & a_{43} \end{pmatrix} = \gamma_2 , \end{align*}$$

a contradiction. Therefore, $a_{52} = a_{53}=0$ . Considering the image of the second basis vector $e_2\wedge f_3 +2 e_1 \wedge f_2$ , we similarly find that $a_{52} = a_{54} = 0$ . Therefore, g stabilises $\langle e_1,e_2,e_3,f_3 \rangle $ , and therefore also its radical $ \langle e_1,e_2 \rangle $ , reducing to the case $H\leq G_{\langle e_1,e_2 \rangle }$ . The case $H\leq G_{\langle e_1,e_2,e_3 \rangle }$ follows similarly. This proves that $H_{W^*}\leq B$ . It remains to show that $B_{W^*}=S^*$ . Given $g\in B$ , we can write it as

$$\begin{align*}g &=h_{\alpha_1}(t_1)h_{\alpha_2}(t_2)h_{\alpha_3}(t_3)x_{100}(a_1)x_{110}(a_2)x_{010}(a_3)x_{221}(a_4)x_{121}(a_5)\\&\quad x_{111}(a_6)x_{021}(a_7)x_{011}(a_8)x_{001}(a_9),\end{align*}$$

where $t_1,t_2,t_3\in K^*$ and $a_i\in K$ for $1\leq i \leq 9$ . Let U be the subspace of V with basis given by

$$\begin{align*}u_1,\dots, u_7 = e_1\wedge e_3,\, e_1\wedge f_2,\,e_1\wedge f_1-e_2\wedge f_2,\,e_3\wedge f_2,\,e_3\wedge f_1,\,f_2\wedge f_3,\,f_1\wedge f_2.\end{align*}$$

Then $V = W^* \oplus U$ . For each basis vector $u^{(i)}$ of $W^*$ , write $g.u^{(i)}$ as $w^*+ \sum _{j=1}^7 f_{ij}u_i$ , where $w^* \in W^*$ and we view $f_{ij}$ as an element of $\mathbb {F}_7[t_1^{\pm 1},t_2^{\pm 1}, t_3^{\pm 1},a_1,\dots ,a_9]$ . Then to determine $B_{W^*}$ it suffices to determine the zero locus of the ideal $I\otimes K \leq K[t_1^{\pm 1},t_2^{\pm 1}, t_3^{\pm 1},a_1,\dots ,a_9]$ , where $I \leq \mathbb {F}_7[t_1^{\pm 1},t_2^{\pm 1}, t_3^{\pm 1},a_1,\dots ,a_9]$ is the ideal generated by $\{f_{ij}:1\leq i,j\leq 7\}$ . We do this by first determining a Groebner basis for I using Magma. See Listing 5 for the code that does this. We find that I is generated by

$$ \begin{gather*} a_1 + 6a_9,\, a_2 + 3a_9^2,\, a_3 + 6a_9,\, a_4 + a_9^5,\, a_5 + 6a_9^4,\, a_6 + a_9^3,\, a_7 + 2a_9^3,\, a_8 + 3a_9^2,\, a_9^7 + 6a_9,\, \\ t_1 + 6t_2^{-2}t_3^{-3},\, t_2 + 6t_2^{-2},\, t_3 + 6t_3^{-3},\, t_1^{-1} + 6(t_2t_3)^{-1},\, t_2^{-3} + 6,\, t_3^{-4} + 6. \end{gather*} $$

It is now easy to find the zero locus of $I\otimes K$ , to determine that

$$ \begin{gather*} h_{\alpha_1}(t_1)h_{\alpha_2}(t_2)h_{\alpha_3}(t_3)x_{100}(a_1)x_{110}(a_2)x_{010}(a_3)x_{221}(a_4)\\x_{121}(a_5)x_{111}(a_6)x_{021}(a_7)x_{011}(a_8)x_{001}(a_9) \in B_{W^*} \\ \iff a_1\in \mathbb{F}_7,\, a_9 = a_1,\, a_2 = 4a_1^2,\, a_3 = a_1,\, a_4 = 6a_1^5,\, a_5 = a_1^4, \, a_6 = -a_1^3,\, a_7 = 5a_1^3,\, a_8 = 4a_1^2,\, \\ t_2^3 = t_3^4 = 1, t_1 = t_2t_3. \end{gather*} $$

Taking $a_1=t_1=t_2=t_3 = 1$ , we get the element x, while taking $a_1 = 0$ we get the subgroup generated by $\tau ^*$ . This completes the proof that $B_{W^*} = S^*$ .

We now turn to the case $p\neq 7$ . The only proper nonzero subspaces of V stabilised by $S^\dagger $ are $\langle e_1,e_2,f_3 \rangle $ and $\langle e_3,f_1,f_2 \rangle $ . Suppose that $H\leq G_{\langle e_1,e_2,f_3 \rangle }$ . Then it is easy to see that $H_{W^*}$ stabilises the subspace $W^\ddagger = \langle v^{(0)},v^{(1)},v^{(2)},v^{(4)} \rangle $ . Let

$$\begin{align*}g = u^- n t u,\end{align*}$$

where

$$ \begin{align*} u &= x_{100}(a_1)x_{110}(a_2)x_{010}(a_3)x_{221}(a_4)x_{121}(a_5)x_{111}(a_6)x_{021}(a_7)x_{011}(a_8)x_{001}(a_9), \\ t &=h_{\alpha_1}(t_1)h_{\alpha_2}(t_2)h_{\alpha_3}(t_3), \\ n &\in \{1,n_1,n_2,n_1n_2,n_2n_1,n_1n_2n_1\}, \\ u^- &=x_{100}(b_1)x_{110}(b_2)x_{010}(b_3)x_{221}(b_4)x_{121}(b_5)x_{111}(b_6)x_{021}(b_7)x_{011}(b_8)x_{001}(b_9),\\ u^- &\in \langle X_\alpha : n^{-1}.\alpha \in \Phi^-,\alpha\in \Phi^+ \rangle, \end{align*} $$

for $a_1,\dots ,a_9,b_1,\dots ,b_9\in K$ and $t_1,t_2,t_3\in K^*$ . Then g is an arbitrary element of $G_{\langle e_1,e_2,e_3 \rangle }$ , written in terms of its Bruhat decomposition, and therefore, $g^{n_3}$ is an arbitrary element of $G_{\langle e_1,e_2,f_3 \rangle }$ .

For each possible n, we use Magma to solve the system of equations corresponding to $g^{n_3} \in G_{W^\ddagger }$ , similarly to how we did for $p=7$ . More care is now required for the setup of the computations in Magma since the characteristic is arbitrary. What we do is find a Groebner basis over $\mathbb {Q}$ and also output the list of primes that the algorithm divided by in its various steps. If p is not in such list, we can use the Groebner basis to easily solve the system; otherwise, we simply run the Groebner basis algorithm again over $\mathbb {F}_p$ . This is done as per Listing 6 and Listing 7. We find that $g^{n_3}$ does not stabilise $W^\ddagger $ when $n \in \{n_1,n_2,n_1n_2n_2\}$ . When $n\in \{1,n_1n_2,n_2n_1\}$ , we find that $g^{n_3}$ stabilises $W^\ddagger $ if and only if $u = u^- = 1$ and $t_1,t_2,t_3$ satisfy the following:

$$ \begin{align*} t_1 &= t_2^{-2}t_3^{-1}, & t_2^7 &= 1, & t_3^2 &= 1, &\text{ when } n &= 1; \\ t_1 &= -t_2^{-2}t_3^{-1}, & t_2^7 &= 1, & t_3^2 &= 1, &\text{ when } n &= n_1n_2; \\ t_1 &= t_2^{-2}t_3^{-1}, & t_2^7 &= -1, & t_3^2 &= 1, &\text{ when } n &= n_2n_1. \end{align*} $$

This is easily seen to be equivalent to $g^{n_3}\in S^*$ . The case $H\leq G_{\langle e_3,f_1,f_2 \rangle }$ follows similarly, with an arbitrary element of $G_{\langle e_3,f_1,f_2 \rangle }$ being given by $g^{n_1n_2n_3n_2n_1n_2n_3n_2}$ .

Lemma 5.29. Let $H = N_G(S^\dagger )$ . Then $H_{W^*}= S^*$ .

Proof. Assume $p\neq 7$ . Since x is a regular semisimple element, we have $H\leq N_G(T)$ , from which it is easy to check that $H \leq T.\langle \tau ^* \rangle $ . A quick calculation shows that $T_{W^*} = \pm \langle x \rangle $ , concluding as required.

If $p = 7$ , we have $H\leq B$ , where B is the standard Borel subgroup of G. Then we conclude by Lemma 5.28.

Lemma 5.30. Suppose that $S^\dagger \leq H.\mathbb {Z}_2 < G$ , where H is an arbitrary subgroup of G. Then $(H.\mathbb {Z}_2)_{W^*}\leq S^*$ if and only if $H_{W^*}\leq S^*$ .

Proof. The forward direction is trivial. Suppose that $H_{W^*}\leq S^*$ . Since $S^\dagger $ has order $21$ , we must have $S^\dagger \leq H$ . Since $H_{W^*}\leq S^*$ , the subgroup $S^\dagger $ is the unique subgroup of order $21$ in $H_{W^*}$ . Therefore, $H_{W^*} \lhd (H.\mathbb {Z}_2)_{W^*}$ implies that $S^\dagger $ is normal in $(H.\mathbb {Z}_2)_{W^*}$ . By Lemma 5.29, we know that the stabilizer of $W^*$ in $N_G(S^\dagger )$ is $S^*$ , concluding.

Proposition 5.31. Let $G=C_3$ , $\lambda = \lambda _2$ with $p\neq 3$ . Then $C_{\mathcal {S}_7'(V)}=\mathbb {Z}_7.\mathbb {Z}_6$ and $C_{\mathcal {S}_7"(V)} = \mathbb {Z}_{7/(p,7)}.\mathbb {Z}_6$ .

Proof. Let $S = G_{W^*}$ . We shall prove that $S\leq N_G(S^\dagger )$ . This will conclude the proof of the proposition as follows. By Lemma 5.29, we have $(N_G(S^\dagger ))_{W^*} = S^*$ , which then implies $S=S^*$ and that $W^*$ is in a dense G-orbit on one of the two $D_7$ -orbits on $\mathcal {S}_7(V)$ . If $p\neq 7$ , then $W_{(\omega ,i)}$ and $W_{(\omega ^2,i)}$ intersect in a $6$ -dimensional subspace, and therefore, by Lemma 2.6, they belong to distinct $D_7$ -orbits on $\mathcal {S}_7(V)$ . They each have stabilizer $S^*$ , concluding the $p\neq 7$ case. If $p=7$ , the subspace $W^*$ intersects $W_{(2,i)}$ trivially, which by Lemma 2.6 implies that $W^*$ and $W_{(4,\zeta )}$ belong to distinct $D_7$ -orbits on $\mathcal {S}_7(V)$ . Again $G_{W^*} = S^*$ , concluding.

In order to prove that $S\leq N_G(\langle x \rangle )$ , we show that for all $p<\infty $ and $e\in \mathbb {Z}_{\geq 1}$ , if

$$\begin{align*}S^\dagger \leq R \leq Sp_6(p^e)=Sp_6(q)<G, \text{ with }R\not\leq N_G(S^\dagger),\end{align*}$$

then R does not stabilise $W^*$ . Note that this is indeed sufficient since if $g\in S\setminus N_G (S^\dagger )$ , then there must exist $e\in \mathbb {Z}_{\geq 1}$ such that $g \in Sp_6(p^e)$ , with $R = \langle S,g\rangle $ satisfying the condition above. We shall make extensive use of maximal subgroups of $Sp_6(q)$ , combined with computations in Magma. The $p=\infty $ case then follows from the $p<\infty $ case since they are equivalent once p is sufficiently large.

Suppose that

$$\begin{align*}S^\dagger \leq R \leq M_q < Sp_6(q)<G, \text{ with }R\not\leq N_G(S^\dagger),\end{align*}$$

where $M_q$ is a maximal subgroup of $Sp_6(q)$ , as listed in [Reference Bray, Holt and Roney-Dougal5, Table 8.28, Table 8.29]. By Lemma 5.28, we can assume that $M_q$ is an irreducible subgroup, with order divisible by $|S^\dagger | = 21$ , and by Lemma 5.30, we can assume that $M_q$ does not contain a reducible subgroup of index $2$ . The goal is to prove that R does not stabilize $W^*$ . Going through (Reference Bray, Holt and Roney-Dougal5, Table 8.28,Table 8.29) for the maximal subgroups of $Sp_6(q)$ , and (Reference Bray, Holt and Roney-Dougal5, Table 8.30,Table 8.31,Table 8.32,Table 8.33,Table 8.34) for the maximal subgroups of $SO_6^+(q)$ , $SO_6^-(q)$ and $G_2(q)$ in even characteristic, we reduce to having to consider the cases in the following table:

If $M_q$ is $Sp_6(q_0).\mathbb {Z}_2$ , then by Lemma 5.30, we reduce to one of the other cases.

Suppose that $M_q = (Sp_2(q)^3).Sym(3)$ . By Lemma 5.30, we can assume that $R\leq (Sp_2(q)^3).\mathbb {Z}_3$ . Since x has order $7$ , it must be contained in $Sp_2(q)^3$ . Therefore, $p\neq 7$ since $Sp_2(q)^3$ does not contain a regular unipotent element. Then $Sp_2(q)^3$ must be the stabilizer of the orthogonal sum $\langle e_1 ,f_1 \rangle \perp \langle e_2 ,f_2 \rangle \perp \langle e_3 ,f_3 \rangle $ in $Sp_6(q)$ . This implies that $(Sp_2(q)^3).\mathbb {Z}_3 = (Sp_2(q)^3).\langle \tau ^4 \rangle $ . Since $\tau ^4 \in S^\dagger $ , it remains to determine the stabilizer of $W^*$ in $(Sp_2(q)^3)$ , which is easily seen to be $\langle x,\tau ^3 \rangle \leq S^*$ , concluding.

Suppose that $M_q = Sp_2(q)\otimes GO_3(q)$ with $p\neq 2$ . Then $M_q$ does not contain a regular unipotent element, and therefore, $p \neq 7$ . Comparing the actions of $M_q$ and $S^\dagger $ on V, we also see that $M_q$ does not contain $S^\dagger $ .

Suppose that $M_q = Z(G).PSL_2(7).\mathbb {Z}_2$ . By Lemma 5.30, we can assume that $R\leq Z(G).PSL_2(7)$ . Here, we use Magma to determine that the only possibility for R is $Z(G).PSL_2(7)$ itself, which acts on V as a sum of a $6$ -dimensional irreducible and an $8$ -dimensional irreducible when $p\neq 7$ , and with composition factors of dimension $5,1,3,5$ when $p=7$ , therefore not stabilising $W^*$ . The Magma code used here and for the next cases can be found in Listing 8.

Suppose that $M_q = Z(G).PSL_2(13)$ . Then $M_q$ does not contain a subgroup of order $21$ , a contradiction. Suppose that $M_q = (Z(G)\times U_3(3)).\mathbb {Z}_2$ . Clearly, we can reduce to the case $R\leq U_3(3)$ . Here, we find that the only possibility for R is $PSL_2(7)$ , concluding like for the $M_q = Z(G).PSL_2(7).\mathbb {Z}_2$ case.

Suppose that $M_q = Z(G).J_2$ . Then we find that all the possibilities for R contain one of $PSL_2(7)$ , $SL_2(7)$ , $U_3(3)$ , concluding via the previous analysis. Finally, if $M_q = SL_2(q)$ , a similar treatment using the maximal subgroups of $SL_2(q)$ allows us to conclude.

6 Proof of Theorem 4

In this section, we shall prove Theorem 4. Unlike with previous work, we shall not be interested in determining the exact structure of the (semi-)generic stabilizers, if they exist. Instead, we will often resort to finding a dense open subset of the variety where the stabilizers have a certain minimal dimension in order to exclude the possibility of a dense orbit, in a very similar fashion to the work we have done to determine (semi-)generic stabilizers. The following lemma is a crucial tool for narrowing down the cases we will have to consider.

Lemma 6.1. Let $H= Cl(V_1)\otimes Cl(V_2)\leq Cl(V_1\otimes V_2)=Cl(V)$ . Assume that $\dim V_2 = k\dim V_1$ for some $k\geq 1$ . Let $G = Cl(V_1)\otimes Cl(V_2')\leq Cl(V_1\otimes V_2')=Cl(V')$ with $\dim V_2'\geq \dim V_2$ and $Cl(V),Cl(V')= SO(V),SO(V')$ or $Cl(V),Cl(V')= Sp(V),Sp(V')$ . Then H has a dense orbit on $\mathcal {S}_k(V)$ if and only if G has a dense orbit on $\mathcal {S}_k(V')$ .

Proof. Write $\dim V_1 = d_1$ , $\dim V_2 = d_2 = kd_1$ , $\dim V_2' = d_2'$ ; then $\dim V = d_1d_2$ , $\dim V' = d_1d_2'$ . We may assume $V_2\leq V_2'$ ; let $V_2"$ be the orthogonal complement to $V_2$ in $V_2'$ , so that $\dim V_2" = d_2'-d_2$ . Let $\epsilon _V$ be $1$ or $-1$ according as V and $V'$ are both orthogonal or both symplectic, and similarly $\epsilon _{V_2}$ be $1$ or $-1$ according as $V_2$ , $V_2'$ and $V_2"$ are all orthogonal or all symplectic. Then

$$ \begin{align*} \dim G - \dim \mathcal{S}_k(V') &= \dim Cl(V_1)+\dim Cl(V_2')-kd_1d_2'+\frac{3k^2+\epsilon_V k}{2}, \\ \dim H - \dim \mathcal{S}_k(V) &= \dim Cl(V_1)+\dim Cl(V_2)-kd_1d_2+\frac{3k^2+\epsilon_V k}{2}; \end{align*} $$

thus,

$$ \begin{align*} &(\dim G - \dim \mathcal{S}_k(V'))-(\dim H-\dim \mathcal{S}_k(V)) \\ = &\quad \dim Cl(V_2')-\dim Cl(V_2)-kd_1(d_2'-d_2)\\ = &\quad \frac{1}{2}d_2'(d_2'-\epsilon_{V_2})-\frac{1}{2}d_2(d_2-\epsilon_{V_2})-d_2(d_2'-d_2)\\ = & \quad\frac{1}{2}(d_2'-d_2)(d_2'-d_2-\epsilon_{V_2})\\ = &\quad \dim Cl(V_2"). \end{align*} $$

Now let $v_1,\dots ,v_{d_1}$ be a fixed basis of $V_1$ . Given $y\in \mathcal {S}_k(V')$ , choose a basis $x_1,\dots ,x_k$ of y and write each $x_j$ uniquely as $\sum _{i=1}^{d_1} v_i\otimes u_{ij}$ with each $u_{ij}\in V_2'$ ; set $\mathrm {supp}_2(y) = \langle u_{ij}:1\leq i\leq d_1,\, 1\leq j\leq k\rangle \leq V_2'$ . Define

$$\begin{align*}Y = \{y\in \mathcal{S}_k(V'):\mathrm{supp}_2(y)\text{ is nondegenerate of dimension }d_2\}.\end{align*}$$

The set Y is dense in $\mathcal {S}_k(V')$ because the set $\{ y\in \mathcal {S}_k(V'): \dim \mathrm {supp}_2(y) <d_2\}$ is a proper closed subvariety of $\mathcal {S}_k(V')$ , and nondegenerate $d_2$ -spaces are dense in the variety of all $d_2$ -spaces in $V_2'$ ; likewise, $Y\cap \mathcal {S}_k(V)$ is dense in $\mathcal {S}_k(V)$ . Moreover, all nondegenerate $d_2$ -spaces in $V_2'$ lie in a single $Cl(V_2')$ -orbit; thus, given $y\in Y$ , by applying an element of G, we may assume that $\mathrm {supp}_2(y) = V_2$ , and then the stabilizer of y in G must fix $V_2$ and hence $V_2"$ , whence $G_y = H_y \times Cl(V_2")$ , so that $\dim G_y-\dim H_y = \dim Cl(V_2")$ , and hence,

$$\begin{align*}\dim \mathcal{S}_k(V')-(\dim G-\dim G_y) = \dim \mathcal{S}_k(V)-(\dim H-\dim H_y).\end{align*}$$

Now if H has a dense orbit on $\mathcal {S}_k(V)$ , it must meet $Y\cap \mathcal {S}_k(V)$ ; thus, it contains some $y\in Y\cap \mathcal {S}_k(V)$ for which the right side of the above equation is zero, whence the left side is also zero and so G has a dense orbit on $\mathcal {S}_k(V')$ . Conversely, if G has a dense orbit on $\mathcal {S}_k(V')$ , it must meet Y, and therefore, $Y\cap \mathcal {S}_k(V)$ ; thus, it contains some $y\in Y\cap \mathcal {S}_k(V)$ for which the left side of the above equation is zero, whence the right side is also zero and H has a dense orbit on $\mathcal {S}_k(V)$ .

Lemma 6.2. Let $G=SO(V_1)\otimes SO(V_2)$ with $3\leq \dim V_1\leq \dim V_2$ . Then G does not have a dense orbit on $\mathcal {S}_k(V_1\otimes V_2)$ for all $1\leq k\leq \frac {1}{2}\dim V_1\dim V_2$ .

Proof. Let $m=\dim V_1$ , $n=\dim V_2$ and assume that G does have a dense orbit on $\mathcal {S}_k(V_1\otimes V_2)$ . Then

$$\begin{align*}\dim G = \frac{1}{2}(m^2+n^2-m-n)\geq mnk-\frac{3k^2+k}{2} = \dim\mathcal{S}_k(V_1\otimes V_2). \end{align*}$$

We first show that $mk\leq n$ . If $k=1$ this is immediate, so assume $k>1$ . Write $n= am$ for some $a\geq 1$ . Then if we define $g:\mathbb {R}\rightarrow \mathbb {R}$ by

$$\begin{align*}g(x) =m^2 + m^2 x^2 - m - m x - 2 m^2 x k + 3 k^2 + k,\end{align*}$$

we have $g(a) = 2(\dim G-\dim \mathcal {S}_k(V_1\otimes V_2))\geq 0$ . The discriminant of $g(x)$ is $(-m-2m^2k)^2-4m^2(m^2-m+3k^2+k) = m^2h(m)$ , where we define $h:\mathbb {R}\rightarrow \mathbb {R}$ by

$$\begin{align*}h(x) = 1-4k-12k^2+4(k+1)x+4(k^2-1)x^2.\end{align*}$$

In turn, the discriminant of $h(x)$ is $16(k+1)^2-16(k^2-1)(1-4k-12k^2)=16(k+1)(12k^3-8k^2-4k+2)\geq 0$ , so that the equation $y = h(x)$ has real roots; the positive root is

$$\begin{align*}m = \frac{-(k+1)+\sqrt{(k+1)(12k^3-8k^2-4k+2)}}{2(k^2-1)},\end{align*}$$

which is easily seen to be always less than $2$ . Thus, as $m\geq 3$ , we have $h(m)>0$ , so that $g(x)$ has positive discriminant, and therefore, the equation $y=g(x)$ has real roots

$$\begin{align*}r_1,r_2 = \frac{1+2km\pm\sqrt{h(m)}}{2m},\end{align*}$$

where $r_1<r_2$ . We claim that $r_1<a$ . Since $a\geq 1$ , the claim is certainly true if $r_1<1$ ; we have $r_1<1 \Leftrightarrow h(m)> (2km-2m+1)^2 \Leftrightarrow 2km^2-2m^2+2m-3k^2-k>0$ . If $k\leq 4$ , the last inequality holds as $m\geq 3$ , so we may assume $k\geq 5$ . If we had $k\leq \frac {1}{2}m^2$ , this would force $m\geq 4$ , and then

$$ \begin{align*} 2km^2-2m^2+2m-3k^2-k &= \frac{2}{5}(k-5)m^2+2m+k\left(\frac{8}{5}m^2-3k-1\right) \\ & \geq \frac{2}{5}(k-5)m^2+2m+k\left(\frac{8}{5}m^2-\frac{3}{2}m^2-1\right) \\ & = \frac{2}{5}(k-5)m^2+2m+k\left(\frac{1}{10}m^2-1\right) \\ &> 0, \end{align*} $$

so we may assume $k\geq \frac {1}{2}m^2$ . Write $k = \theta m^2$ ; as $k\leq \frac {1}{2}mn=\frac {1}{2}am^2$ , we have $\frac {1}{2}<\theta \leq \frac {1}{2}a$ . Then $r_1<2\theta $ if and only if

$$\begin{align*}1+2\theta m^3-\sqrt{4\theta^2m^6-12\theta^2 m^4 +4\theta m^3-4(1+\theta)m^2+4m+1}<4\theta m,\end{align*}$$

which reduces to

$$\begin{align*}m(\theta^2(m^2-4)-\theta-1)+(1+2\theta)>0,\end{align*}$$

which is true for all $m\geq 3$ and $\theta> \frac {1}{2}$ . Thus, we do have $r_1<a$ , as claimed; so as $g(a)\geq 0$ , we must have $r_2<a$ , whence $a>\frac {r_1+r_2}{2}= \frac {1+2km}{2m}>k$ as required.

Let $U\leq V_2$ be a nondegenerate subspace of dimension $k\dim V_1$ . By Lemma 6.1, we know that $SO(V_1)\otimes SO(U)$ has a dense orbit on $\mathcal {S}_k(V\otimes U)$ . The dimension requirement is that $ m^2-m+m^2k^2-mk \geq 2 m^2k^2-3k^2-k$ , which is absurd when $m\geq 3$ and $k\geq 1$ . This contradicts G having a dense orbit on $\mathcal {S}_k(V_1\otimes V_2)$ .

Lemma 6.3. Let $G=Sp(V_1)\otimes SO(V_2)$ with $4\leq \dim V_1\leq \dim V_2$ or $3\leq \dim V_2\leq \dim V_1$ . Then G does not have a dense orbit on $\mathcal {S}_k(V_1\otimes V_2)$ for all $2\leq k\leq \frac {1}{2}\dim V_1\dim V_2$ .

Proof. This is similar to Lemma 6.2. Note that here $k\geq 2$ .

Lemma 6.4. Let $G=Sp(V_1)\otimes Sp(V_2)$ with $4\leq \dim V_1\leq \dim V_2$ . Then if G has a dense orbit on $\mathcal {S}_k(V_1\otimes V_2)$ , we must have $k=1$ .

Proof. Let $m=\frac {\dim V_1}{2}$ , $n = \frac {\dim V_2}{2}$ and assume that G does have a dense orbit on $\mathcal {S}_k(V_1\otimes V_2)$ . Then

$$\begin{align*}\dim G = 2m^2+2n^2+m+n\geq 4mnk-\frac{3k^2+k}{2} = \dim\mathcal{S}_k(V_1\otimes V_2), .\end{align*}$$

Similar calculations as in Lemma 6.2 show that $k\dim V_1\leq \dim V_2$ . By Lemma 6.1, we then deduce that $Sp(V_1)\otimes Sp(U_2)$ has a dense orbit on $\mathcal {S}_k(V_1\otimes U_2)$ , where $\dim U_2 = k\dim V_1$ . Then dimensional considerations rule out $k\geq 2$ .

Proposition 6.5. Let $G=Sp(V_1)\otimes Sp(V_1)$ with $\dim V_1\geq 6$ . Then G does not have a dense orbit on $\mathcal {S}_1(V_1\otimes V_1)$ .

Proof. Let $2n = \dim V_1$ and $e_1,\dots ,e_n,f_n,\dots ,f_1$ be the standard basis of $V_1$ . Let

$$\begin{align*}Y=\left\{\left\langle \sum_{i=1}^{n} a_i e_i\otimes f_i\right\rangle: \sum a_i^2 = 0\right\},\end{align*}$$

a subvariety of $\mathcal {S}_1(V_1\otimes V_1)$ . Let

$$\begin{align*}\hat{Y}=\left\{\left\langle \sum_{i=1}^{n} a_i e_i\otimes f_i\right\rangle: \sum a_i^2 = 0, a_i\neq a_j\text{ if }i\neq j \right\},\end{align*}$$

a dense subset of Y. Let $y\in \hat {Y}$ . Then $\mathrm {Tran}_G(y,Y)$ contains an $A_1^n$ stabilising all elements of y, which projects onto each $Sp(V_1)$ as $\bigcap Sp(V_1)_{\langle e_i,f_i\rangle }$ . By assumption on the $a_i$ ’s, it is easy to see that this is the connected component of $\mathrm {Tran}_G(y,Y)$ . Since $\dim G-\dim \mathrm {Tran}_G(y,Y) = 4n^2+2n -3n = 4n^2-2-(n-2)=\dim \mathcal {S}_1(V_1\otimes V_1) - \dim Y$ , we find that y is Y-exact. Therefore, by Lemma 2.14, we conclude that there is an open dense subset of the variety of singular $1$ -spaces of $V_1\otimes V_1$ such that all stabilizers are $3n$ -dimensional. Therefore, $3n$ is the lower bound for the dimension of the stabilizer of any singular $1$ -space. Since $\dim G-3n = 4n^2-2-(n-2)> \dim \mathcal {S}_1(V_1\otimes V_1)$ when $n\geq 3$ , we conclude that G does not have a dense orbit on $\mathcal {S}_1(V_1\otimes V_1)$ .

Lemma 6.6. Let $G=Sp_2\otimes Sp(V_2)\leq SO(V)$ with $\dim V = 2\dim V_2 \geq 20$ . Suppose that G has a dense orbit on $\mathcal {S}_k(V)$ for some $2\leq k\leq \dim V_2$ . Then k is either $2,3,4,{\dim V_2}-1$ , $\dim V_2'$ or $\dim V_2"$ .

Proof. Calculations similar to the ones in Lemma 6.2 show that if $k<\dim V_2-1$ , we have $2k\leq \dim V_2$ . We can then apply Lemma 6.1 to determine that $Sp_2\otimes Sp_{2k}=Sp(V_1)\otimes Sp(U)$ must have a dense orbit on $\mathcal {S}_k(V_1\otimes U)$ . Dimensional considerations then give $ 3+2k^2+k\geq 4k^2-\frac {3}{2}k^2-\frac {k}{2}$ , which implies $k\leq 4$ .

Lemma 6.7. Let $G=Sp_2\otimes Sp(V_2)\leq SO(V)$ with $\dim V = 2\dim V_2\leq 16$ . Suppose that G has a dense orbit on $\mathcal {S}_k(V)$ for some $2\leq k\leq \dim V_2-2$ . Then either $k=2$ or $(k,\dim V_2)$ is one of $(3,6),(3,8),(4,6),(4,8),(6,8)$ .

Proof. This is simply a matter of checking the dimension of $\mathcal {S}_k(V)$ in all finitely many cases.

Proposition 6.8. Let $G=Sp_2\otimes Sp_{2n} \leq SO_{4n}=SO(V)$ . Then G has a dense orbit on $\mathcal {S}_{2n}'(V)$ (and $\mathcal {S}_{2n}"(V)$ ) if and only if $n = 1,2,3$ .

Proof. When $n=1,2$ the group G is spherical in $SO(V)$ . Therefore, assume that $n\geq 3$ . Let $ V_1=\langle e,f\rangle $ and $V_2=\langle e_1,\dots ,e_n, f_n,\dots ,f_1\rangle $ so that $G=Sp(V_1)\otimes Sp(V_2)$ and the given bases are the standard bases for $V_1$ and $V_2$ . Given $\mathbf {a} = (a_1,\dots ,a_n)\in K^n$ , define $W_{\mathbf {a}} = \langle (e+a_i f)\otimes e_i, (e+a_i f)\otimes f_i:1\leq i\leq n\rangle $ , a $2n$ -dimensional totally singular subspace of V. Define

$$\begin{align*}Y = \{W_{\mathbf{a}}: \mathbf{a}\in K^n\},\end{align*}$$

an n-dimensional subvariety of $\mathcal {S}_{2n}'(V)$ . Let

$$\begin{align*}\hat{Y} = \{W_{\mathbf{a}}: \mathbf{a}\in K^n,a_1\dots a_n\neq 0,\,a_i\neq a_j \text{ for all }i\neq j\},\end{align*}$$

a dense subset of Y. Let $y\in \hat {Y}$ . It is easily seen that $\mathrm {Tran}_G(y,Y)$ has connected component $A_1^{n+1} = Sp(V_1)\otimes \bigcap Sp(V_2)_{\langle e_i+f_i \rangle }$ , while $G_y^0 = 1\otimes \bigcap Sp(V_2)_{\langle e_i+f_i \rangle }\simeq A_1^n$ . Since

$$\begin{align*}\dim G-\dim \mathrm{Tran}_G(y,Y) = \dim\mathcal{S}_{2n}'(V)-\dim Y,\end{align*}$$

we have that $\hat {Y}$ is Y-exact. Therefore, by Lemma 2.14 and Corollary 2.12, we know that $\dim A_1^n$ is the minimum dimension for the stabilizer of any $y\in \mathcal {S}_{2n}'(V)$ . Dimensional considerations rule out $n\geq 4$ , while for $n=3$ , we have $\dim G-\dim A_1^3 = 15=\dim \mathcal {S}_{2n}'(V).$ By Lemma 2.6, changing the definition of $W_{\mathbf {a}}$ by swapping the first two generators $(e+a_1f)\otimes e_1,(e+a_1f)\otimes f_1$ with $e\otimes (e_1+a_1f_1),f\otimes (e_1+a_1f_1)$ leads to the same result for the action on $\mathcal {S}_{2n}"(V)$ .

Proposition 6.9. Let $G=Sp_2\otimes Sp_{2n} \leq SO(4n)=SO(V)$ . Then G has a dense orbit on $\mathcal {S}_{2n-1}(V)$ if and only if $n = 1,2,3$ .

Proof. Suppose that G has a dense orbit on $\mathcal {S}_{2n-1}(V)$ . Every $y\in \mathcal {S}_{2n-1}(V)$ is contained in precisely one element of $\mathcal {S}_{2n}'(V)$ and one of $\mathcal {S}_{2n}"(V)$ . Let $\mathcal {O}$ be the dense orbit of G on $\mathcal {S}_{2n-1}(V)$ ; then its complement $\mathcal {S}_{2n-1}(V) \setminus \mathcal {O}$ is contained in a proper closed subvariety X of $\mathcal {S}_{2n-1}(V)$ . Let Z be the set of elements of $\mathcal {S}^{\prime }_{2n}(V)$ all of whose hyperplanes lie in X; then Z is a proper closed subvariety of $\mathcal {S}^{\prime }_{2n}(V)$ , so its complement $\mathcal {S}^{\prime }_{2n}(V) \setminus Z$ is a dense subset of $\mathcal {S}^{\prime }_{2n}(V)$ with the property that any of its elements has a hyperplane lying in $\mathcal {O}$ . Thus, given two elements of $\mathcal {S}^{\prime }_{2n}(V) \setminus Z$ , we can choose hyperplanes within them and an element of G which sends one hyperplane to the other and therefore one element of $\mathcal {S}^{\prime }_{2n}(V) \setminus Z$ to the other; so $\mathcal {S}^{\prime }_{2n}(V) \setminus Z$ lies in a single G-orbit, and therefore, G has a dense orbit on $\mathcal {S}^{\prime }_{2n}(V)$ . Replacing $\mathcal {S}^{\prime }_{2n}(V)$ by $\mathcal {S}^{\prime \prime }_{2n}(V)$ shows that G also has a dense orbit on $\mathcal {S}^{\prime \prime }_{2n}(V)$ . By Proposition 6.8, we therefore have $n=1,2,3$ . When $n=1,2$ the group G is spherical in $SO(V)$ . Therefore, assume that $n= 3$ . Given y in the dense G-orbit on $\mathcal {S}_{6}'(V)$ , the group induced by $G_y$ on y is $A_1^3$ acting as a sum of three natural modules for $A_1$ . Therefore, $G_y$ has a dense orbit on $5$ -spaces of y, concluding that G has a dense orbit on $\mathcal {S}_{5}(V)$ .

Proposition 6.10. Let $G=Sp_2\otimes Sp_6 \leq SO(12)=SO(V)$ . Then G has a dense orbit on $\mathcal {S}_3(V)$ .

Proof. Let $ V_1=\langle e,f\rangle $ and $V_2=\langle e_1,e_2,e_3,f_3,f_2,f_1\rangle $ so that $G=Sp(V_1)\otimes Sp(V_2)$ and the given bases are the standard bases for $V_1$ and $V_2$ . Let

$$\begin{align*}W = \langle e\otimes e_1+f\otimes e_2,\quad e\otimes f_2+f\otimes f_1,\quad e\otimes (e_2+e_3)+f\otimes (f_3-f_2)\rangle, \end{align*}$$

a totally singular $3$ -space of V. Let $S=G_W.$ Let $g\in S$ such that $g=1\otimes g_1$ . Then $g_1$ fixes $\langle e_1,f_2,e_2+e_3\rangle $ and $\langle e_2,f_1,f_3-f_2\rangle $ , and consequently their radicals $\langle e_1\rangle $ and $\langle f_1 \rangle $ . One then quickly reaches the conclusion that $g_1$ must also fix $\langle e_2 \rangle $ , $\langle f_2 \rangle $ , $\langle e_2+e_3 \rangle $ and $\langle f_3-f_2\rangle $ . However, this would mean that g acts on W by sending $e\otimes (e_2+e_3)+f\otimes (f_3-f_2)\mapsto \lambda e\otimes (e_2+e_3)+\lambda ^{-1}f\otimes (f_3-f_2)$ , implying that $g_1 =\pm 1$ . To conclude, we observe that given $g = g_1\otimes 1$ , by a simple application of Witt’s Lemma, we can find $g' = 1\otimes g_2$ such that $gg'\in S$ . Therefore, $S^0=Sp_2$ . Since $\dim G-\dim S = 21 = \dim \mathcal {S}_3(V)$ , we conclude that G has a dense orbit on $\mathcal {S}_3(V)$ .

Proposition 6.11. Let $G=Sp_2\otimes Sp_6 \leq SO_{12}=SO(V)$ . Then G has no dense orbit on $\mathcal {S}_4(V)$ .

Proof. Let $ V_1=\langle e,f\rangle $ and $V_2=\langle e_1,e_2,e_3,f_3,f_2,f_1\rangle $ so that $G=Sp(V_1)\otimes Sp(V_2)$ and the given bases are the standard bases for $V_{nat}$ . The stabilizer of an element in a dense orbit would have dimension $1$ . We will show that already the group $G_2:=1\otimes Sp_6 < G$ acts on $\mathcal {S}_4(V)$ with stabilizers that are at least $3$ -dimensional. Let $W_{abcd}$ be the totally singular $4$ -space spanned by

$$\begin{align*}\begin{array}{ll} e\otimes e_1+f\otimes (ae_1+be_2+ce:3+de_4), & e\otimes e_2+f\otimes (be_1+de_2),\\ e\otimes f_1+f\otimes (af_1+bf_2+cf_3+df_4), & e\otimes f_2+f\otimes (bf_1+df_2).\\ \end{array}\end{align*}$$

Let $Y = \{ W_{abcd}:a,b,c,d\in K\}$ , a $4$ -dimensional subvariety of $\mathcal {S}_4(V)$ . Let $\hat {Y} = \{ W_{abcd}:a,b,c,d\in K^*\}$ , a dense subset of Y. Take $y=W_{abcd}\in \hat {Y}$ . Since $y\subseteq e\otimes \langle e_1,f_2,e_2,f_2\rangle +f\otimes V_6$ , the stabilizer $(G_2)_y$ must preserve $\langle e_1,f_1,e_2,f_2\rangle $ , and therefore its orthogonal complement $\langle e_3,f_3\rangle $ . Since for all $g\in (G_2)_y$ we must have $g.(e\otimes e_2+f\otimes (be_1+de_2))\in y$ and $g.(e\otimes f_2+f\otimes (bf_1+df_2))\in y$ , we get that $\langle e_2,f_2\rangle $ must also be preserved by $(G_2)_y$ . Therefore, $(G_2)_y\leq A_1^3$ . It is now immediate to see that the image of any $e_i$ or $f_i$ in $\langle e_i,f_i\rangle $ completely determines the element $g\in (G_2)_y$ . Since the standard diagonal subgroup $A:=A_1\leq A_1^3$ fixes y, we must then have $(G_2)_y = A$ . Now assume that $g\in \mathrm {Tran}_{G_2}(y,Y)$ . Again, we have $g\in A_1^3$ , and since A clearly fixes any element of Y, it is the stabilizer of $g.y$ . Therefore, $g\in N_{A_1^3}(A)$ , which is a finite extension of A. This shows that $\dim \mathrm {Tran}_{G_2}(y,Y) = 3$ , and therefore, $\operatorname {\mathrm {codim}} \mathrm {Tran}_{G_2}(y,Y) = 18 = 22-4 = \operatorname {\mathrm {codim}} Y$ . This shows that all points in $\hat {Y}$ are Y-exact, and Lemma 2.14 allows us to conclude that $3=\dim A_1$ is the lower bound for the dimension of any stabilizer for the $G_2$ -action on $\mathcal {S}_4(V)$ . In particular, this proves that G has no dense orbit on $\mathcal {S}_4(V)$ .

Proposition 6.12. Let $G=Sp_2\otimes Sp_8 \leq SO_{16}=SO(V)$ . Then G has no dense orbit on $\mathcal {S}_4(V)$ and on $\mathcal {S}_6(V)$ .

Proof. Let $ V_1=\langle e,f\rangle $ and $V_2=\langle e_1,e_2,e_3,e_4,f_4,f_3,f_2,f_1\rangle $ so that $G=Sp(V_1)\otimes Sp(V_2)$ and the given bases are the standard bases for $V_1$ and $V_2$ . The stabilizer of an element in a dense orbit would have dimension $2$ . We will show that already the group $G_2:=1\otimes Sp_8 < G$ acts on $\mathcal {S}_4(V)$ with stabilizers that are at least $3$ -dimensional. Given $ \mathbf {a}\in K^7$ , let $W_{\mathbf {a}}$ be the totally singular $4$ -space spanned by vectors

$$\begin{align*}\begin{array}{ll} e\otimes e_1+f\otimes (a_1e_1+a_2e_2+a_3e_3+a_4e_4), & e\otimes e_2+f\otimes (a_2e_1+a_5e_2+a_6e_3+a_7e_4),\\ e\otimes f_1+f\otimes (a_1f_1+a_2f_2+a_3f_3+a_4f_4), & e\otimes f_2+f\otimes (a_2f_1+a_5f_2+a_6f_3+a_7f_4).\\ \end{array}\end{align*}$$

Let $Y = \{ W_{\mathbf {a}}:\mathbf {a}\in K^7\}$ , a $7$ -dimensional subvariety of $\mathcal {S}_4(V)$ . Let $\hat {Y}_1 = \{ W_{\mathbf {a}}:\mathbf {a}\in (K^*)^7\}$ , a dense subset of Y. The standard diagonal $A_1\leq A_1^4 = \bigcap (G_2)_{\langle e_i,f_i\rangle }$ fixes any $y\in Y$ . Call this $A_1$ -subgroup A. Let $y\in \hat {Y}$ . We will now prove that the connected component of $(G_2)_y$ is A. We begin by observing that $(G_2)_y$ fixes $\langle e_1,f_1,e_2,f_2 \rangle $ , and therefore, $(G_2)_y\leq C_2C_2 $ . Let $\pi _i (A)$ ( $i=1,2$ ) denote the projection of A onto each $C_2$ . Assume $p\neq 2$ . Then $\pi _i (A)$ is a diagonal $A_1$ in $C_2$ , which lies in two opposite parabolic subgroups of $C_2$ , acting irreducibly on their unipotent radicals, and in precisely one Levi subgroup $L_i$ . Also, $\pi _i (A)$ is maximal in infinitely many $A_1^2$ ’s stabilising a sum of two nondegenerate $2$ -spaces. Let M be a minimal connected overgroup of A in $C_2C_2$ . Then $\pi _i(M)$ is one of $A_1$ , $A_1^2$ , $L_i$ , $U_3A_1$ . If $\pi _1(M) \simeq \pi _2(M)$ , then M is diagonal in $C_2C_2$ by minimality. Suppose that $\pi _1(M) \not \simeq \pi _2(M)$ . If $\pi _i(M)\not \leq L_i$ for $i=1$ or $i=2$ , then M contains $\pi _1 (A)\times \pi _2(A)$ . Summarising, the minimal connected ovegroups of A in $C_2C_2$ are as follows:

  1. (i) $A_1^2$ diagonal in $C_2C_2$ ;

  2. (ii) $A_1^2 = \pi _1 (A)\times \pi _2(A)$ ;

  3. (iii) $A T_1$ , where $T_1< Z(L_1)\times Z(L_2)$ is a $1$ -dimensional torus;

  4. (iv) $U_3A$ diagonal in $C_2C_2$ .

Assume we are in the first case and $M=A_1^2$ is diagonal in $C_2C_2$ . Then $\pi _1 (A)$ is maximal in an $A_1^2$ fixing $\langle e_1+\lambda _1 e_2,f_1+\lambda _1 f_2\rangle \perp \langle e_1-\lambda _1^{-1} e_2,f_1-\lambda _1^{-1} f_2\rangle $ , while $\pi _2 (A)$ is maximal in an $A_1^2$ fixing $\langle e_3+\lambda _2 e_4,f_3+\lambda _2 f_4\rangle \perp \langle e_3-\lambda _2^{-1} e_4,f_3-\lambda _2^{-1} f_4\rangle $ , for some $\lambda _1,\lambda _2\in K^*$ . Let $y\in \hat {Y}_1$ and assume that M fixes y. Considering a $1$ -dimensional torus in M but not in A, one finds four independent equations in terms of the entries of $\mathbf {a}$ and $\lambda _1,\lambda _2$ that all need to be satisfied since M fixes y. As the variety of diagonal $A_1^2$ ’s from case $(i)$ is $2$ -dimensional, the elements y of Y such that there is some such diagonal $A_1^2$ fixing y lie in a subvariety of Y whose codimension is at least $2$ . Therefore, there is a dense subset $\hat {Y}_2$ of Y with the property that no minimal connected overgroup of A of type $(i)$ fixes some $y\in \hat {Y}_2$ . The same can be quickly deduced for the other cases. This shows that there exists a dense subset $\hat {Y}$ of Y such that A is the connected component of the stabilizer of any $y\in \hat {Y}$ . Now take $y\in \hat {Y}$ and $g\in \mathrm {Tran}_{G_2}(y,Y)$ . Again, we find $g\in C_2C_2$ , and since A clearly fixes all elements of Y, it must be the connected component of the stabilizer of $g.y$ . Therefore, $g\in N_{C_2C_2}(A)$ . The connected component of $N_{C_2C_2}(A)$ is $AT_2$ , where $T_2$ is a $2$ -dimensional torus, and therefore, $\operatorname {\mathrm {codim}} \mathrm {Tran}_{G_2}(y,Y) \geq 31 $ . However, $\operatorname {\mathrm {codim}} Y = 31$ , which means that $\operatorname {\mathrm {codim}} \mathrm {Tran}_{G_2}(y,Y) = 31 $ and that all points of $\hat {Y}$ are Y-exact. By 2.14, this proves that the minimal dimension for the stabilizer of any totally singular $4$ -space of V is $3$ .

If instead $p=2$ , the reasoning is similar, with the difference that $\pi _i(A)$ is contained in a single parabolic subgroup of $C_2$ and in no Levi subgroup, and has connected centralizer $U_1$ . Therefore, the minimal connected ovegroups of A in $C_2C_2$ are as follows:

  1. (i) $A_1^2$ diagonal in $C_2C_2$ ;

  2. (ii) $A_1^2 = \pi _1 (A)\times \pi _2(A)$ ;

  3. (iii) $U_1 A $ , where $U_1\leq C_{C_2}(\pi _1(A))\times C_{C_2}(\pi _2(A))$ is a $1$ -dimensional unipotent subgroup.

The same analysis then concludes. The case $\mathcal {S}_6(V)$ is entirely similar.

Proof of Theorem 4.

Recall that $V=V_1\otimes V_2$ . If $ G = SO(V_1) \otimes SO(V_2) \leq SO(V) $ , Lemma 6.2 shows that $ G $ has no dense orbit on $ \mathcal {S}_k(V) $ . If $ G = Sp(V_1) \otimes SO(V_2) $ , Lemma 6.3 likewise shows that $ G $ has no dense orbit on $ \mathcal {S}_k(V) $ if $ k \geq 2 $ ; if instead $ k = 1 $ , then $ \mathcal {S}_k(V) = \mathcal {G}_k(V) $ , and so $ G $ has a dense orbit on $ \mathcal {S}_k(V) $ if and only if $ K^*G $ has a dense orbit on $ V $ , that is, if and only if $ (K^*G, V) $ is a prehomogeneous vector space, giving case (i) in the statement of Theorem 4.

For the remainder of the argument, assume $ G = Sp(V_1) \otimes Sp(V_2) \leq SO(V) $ with $ \dim V_1 \leq \dim V_2 $ , and $ G $ has a dense orbit on $ \mathcal {S}_k(V) $ . If $ k = 1 $ , Proposition 6.5 and Lemma 6.1 between them show that we must have $ \dim V_1 = 2 $ or $4$ ; in both possibilities, [Reference Rizzoli26, Thm. 3] shows that $ G $ does indeed have a dense orbit on $ \mathcal {S}_k(V) $ , giving cases (ii) and (vii) respectively in the statement of Theorem 4. Now suppose $ k \geq 2 $ . Lemma 6.4 shows that we must have $ \dim V_1 = 2 $ ; write $ \dim V_2 = 2n $ . If $ n \geq 5 $ , Lemma 6.6 shows that $ k = 2, 3, 4, 2n - 1, (2n)' $ or $ (2n)" $ ; if instead $ n \leq 4 $ , Lemma 6.7 shows that either $ k = 2 $ or $ (k, 2n) = (3, 6), (3, 8), (4, 6), (4, 8) $ or $ (6, 8) $ . If $ k = (2n)' $ or $ (2n)" $ , Proposition 6.8 shows that we must have $ n = 1, 2, 3 $ (and that then $ G $ does indeed have a dense orbit), giving case (v); if $ k = 2n - 1 $ , Proposition 6.9 shows the same thing, giving case (vi). Propositions 6.11 and 6.12 show that $ (k, 2n) \neq (4, 6), (4, 8) $ or $ (6, 8) $ , and then Lemma 6.1 shows that $ k \neq 4 $ . If $ k = 3 $ , Proposition 6.10 shows that $ G $ does have a dense orbit if $ n = 3 $ , and then Lemma 6.1 shows that the same is true if $ n> 3 $ , giving case (iv). Finally, if $ k = 2 $ , the sphericality of $ Sp_2 \otimes Sp_4 $ shows that $ G $ does have a dense orbit if $ n = 2 $ , and then Lemma 6.1 shows that the same is true if $ n> 2 $ , giving case (iii).

7 Proof of Theorem 5

We conclude with the proof of Theorem 5.

Proof of Theorem 5.

By Theorem 2.1, either both G and H are parabolic subgroups, and by the Bruhat decomposition we have finitely many $(G,H)$ -double cosets in $\Gamma $ ; or they are both reductive, and by [Reference Brundan6, Thm. A], there is a dense double coset if and only if there is a single double coset (hence a factorization $\Gamma = GH$ ); or one of the two subgroups, say H, is a maximal parabolic, and G is reductive. It remains to deal with this last case, so assume that $H =P_k$ is a maximal parabolic and G is reductive. If $\Gamma = SL(V)$ or if $\Gamma = Sp(V)$ with $k=1$ , there is a dense $(G,H)$ -double coset in $\Gamma $ if and only if G acts on the Grassmannian $\mathcal {G}_k(V)$ with a dense orbit. This is equivalent to $GL_k\otimes G$ acting on $K^k\otimes V$ with a dense orbit – which is equivalent to the pair $(GL_k\otimes G, K^k\otimes V)$ being a prehomogeneous vector space, as classified by [Reference Sato and Kimura29][Reference Zhijie32][Reference Zhijie33].

Now assume that $\Gamma \neq SL(V)$ and $k\neq 1$ if $\Gamma = Sp(V)$ . There are $4$ options for G. The first option is for G to be simple and irreducible on V, in which case there is a dense $(G,H)$ -double coset if and only if G has a dense orbit on the variety of totally singular subspaces corresponding to $\Gamma /H$ , as classified by Theorem 3.

The second option is for G to be the connected component of the stabilizer of an orthogonal sum (i.e., $G=Sp(V_1)\times Sp(V_2)$ and $\Gamma = Sp(V_1\perp V_2)$ or $G=SO(V_1)\times SO(V_2)$ and $\Gamma = SO(V_1\perp V_2)$ ), and in both cases, G is a spherical subgroup of $\Gamma $ . Thirdly, we can have G being the stabilizer of a degenerate but nonsingular $1$ -space of V, where $p=2$ and V is orthogonal. In this case, G is again spherical in $\Gamma $ .

Lastly, G can be semisimple but not simple, acting irreducibly and tensor decomposably on V. The possibilities for $(G,k)$ are then given by Theorem 4.

A Magma code

Acknowledgements

The author is extremely grateful to Prof. Martin Liebeck to Prof. Donna Testerman for their lasting support and helpful conversations. The author would also like to thank the anonymous reviewer for their very careful reading of the manuscript and their many suggestions for fixing gaps and improving the overall quality.

Competing interest

The author have no competing interest to declare.

Financial Support

The initial part of this work was undertaken during the author’s tenure as INI-Simons Post Doctoral Research Fellow. The author would like to thank INI and DPMMS for support and hospitality during this fellowship, which was supported by the Simons Foundation (Award ID 316017) and by EPSRC (grant number EP/R014604/1).

The work was further supported by the Swiss National Science Foundation (grant number 207730).

References

Andreev, E. M., Vinberg, É. B. and Élashvili, A. G., ‘Orbits of greatest dimension in semi-simple linear Lie groups’, Funct. Anal. Appl. 1 (1967), 257261.CrossRefGoogle Scholar
Aschbacher, M., Finite Group Theory (Cambridge Studies in Advanced Mathematics) second edn. (Cambridge University Press, Cambridge, 2000).CrossRefGoogle Scholar
Babic, A. and Chernousov, V., ‘Lower bounds for essential dimensions in characteristic 2 via orthogonal representations’, Pacific J. Math. 279 (2015), 3763. issn: 0030-8730. doi: 10.2140/pjm.2015.279.37.CrossRefGoogle Scholar
Bourbaki, N., Lie Groups and Lie Algebras: Chapters 4-6 (Elements de mathematique [series]) pts. 46 (Springer Berlin Heidelberg, 2008). isbn: 9783540691716.Google Scholar
Bray, J. N., Holt, D. F. and Roney-Dougal, C. M., The Maximal Subgroups of the Low-Dimensional Finite Classical Groups (London Mathematical Society Lecture Note Series) Vol. 407 (Cambridge University Press, Cambridge, 2013). isbn: 978-0-521-13860-4. doi: 10.1017/CBO9781139192576.CrossRefGoogle Scholar
Brundan, J., ‘Double coset density in classical algebraic groups’, Trans. Amer. Math. Soc. 352 (2000), 14051436. issn: 0002-9947. doi: 10.1090/S0002-9947-99-02258-8.CrossRefGoogle Scholar
Carter, R. W., Simple Groups of Lie Type (Wiley, Chichester, 1989). isbn: 0471137359.Google Scholar
Duckworth, W. E., ‘Infiniteness of double coset collections in algebraic groups’, J. Algebra 273(2) (2004), 718733. issn: 0021-8693. doi: https://doi.org/10.1016/j.jalgebra.2003.08.011.CrossRefGoogle Scholar
Garibaldi, S. and Nakano, D. K., ‘Bilinear and quadratic forms on rational modules of split reductive groups’, Canad. J. Math. 68(2) (2016), 395421. doi: 10.4153/CJM-2015-042-5.CrossRefGoogle Scholar
Guralnick, R. and Lawther, R., Generic Stabilizers in Actions of Simple Algebraic Groups (Memoirs of the American Mathematical Society) (American Mathematical Society, 2024).CrossRefGoogle Scholar
Guralnick, R. M., Liebeck, M. W., Macpherson, D. and Seitz, G. M., ‘Modules for algebraic groups with finitely many orbits on subspaces’, J. Algebra 196 (1997), 211250. issn: 0021-8693. doi: 10.1006/jabr.1997.7068.CrossRefGoogle Scholar
Igusa, J., ‘A classification of spinors up to dimension twelve’, Amer. J. Math. 92 (1970), 9971028. issn: 0002-9327. doi: 10.2307/2373406.CrossRefGoogle Scholar
Kleidman, P. and Liebeck, M., The Subgroup Structure of the Finite Classical Groups (London Mathematical Society Lecture Note Series) vol. 129 (Cambridge University Press, Cambridge, 1990). isbn: 0-521-35949-X. doi: 10.1017/CBO9780511629235.CrossRefGoogle Scholar
Knop, F. and Röhrle, G., ‘Spherical subgroups in simple algebraic groups’, Compos. Math. 151 (2015), 12881308. doi: 10.1112/S0010437X1400791X.CrossRefGoogle Scholar
Korhonen, M., ‘Invariant forms on irreducible modules of simple algebraic groups’, J. Algebra 480 (2017), 385422. issn: 0021-8693. doi: 10.1016/j.jalgebra.2017.02.031.CrossRefGoogle Scholar
Liebeck, M. W. and Saxl, J., ‘On the orders of maximal subgroups of the finite exceptional groups of Lie type’, Proc. Lond. Math. Soc. s3-55.2 (1987), 299330. issn: 0024-6115. doi: 10.1093/plms/s3-55˙2.299. eprint: https://academic.oup.com/plms/article-pdf/s3-55/2/299/4522254/s3-552-299.pdf.CrossRefGoogle Scholar
Liebeck, M. W., Saxl, J. and Seitz, G. M., ‘Factorizations of simple algebraic groups’, Trans. Amer. Math. Soc. 348 (1996), 799822. issn: 0002-9947. doi: 10.1090/S0002-9947-96-01447-X.CrossRefGoogle Scholar
Liebeck, M. W. and Seitz, G. M., ‘On the subgroup structure of classical groups’, Invent. Math. 134 (1998), 427453. issn: 0020-9910. doi: 10.1007/s002220050270.CrossRefGoogle Scholar
Liebeck, M. W. and Seitz, G. M., ‘On the subgroup structure of exceptional groups of Lie type’, Trans. Amer. Math. Soc. 350 (1998), 34093482.CrossRefGoogle Scholar
Lübeck, F., ‘Small degree representations of finite Chevalley groups in defining characteristic’, LMS J. Comput. Math. 4 (2001), 135169. issn: 1461-1570. doi: 10.1112/S1461157000000838.CrossRefGoogle Scholar
Martin, B., ‘Generic stabilisers for actions of reductive groups’, Pacific J. Math. 279(1) (2015), 397422.CrossRefGoogle Scholar
Newstead, P. E., Introduction to Moduli Problems and Orbit Spaces (Tata Institute of Fundamental Research Lectures on Mathematics and Physics) vol. 51 (Tata Institute of Fundamental Research, Bombay; Narosa Publishing House, New Delhi, 1978). isbn: 0-387-08851-2.Google Scholar
Popov, V. L., ‘Classification of spinors of dimension fourteen’, Trans. Moscow Math. Soc. 1 (1980), 181232.Google Scholar
Revoy, P., ‘Trivecteurs de rang 6’, in Colloque sur les formes quadratiques (Montpellier, 1977) (Mémoires de la Société Mathématique de France 59) (Société mathématique de France, 1979), 141155. doi: 10.24033/msmf.256.Google Scholar
Richardson, R. W., ‘Deformations of Lie subgroups and the variation of isotropy subgroups’, Acta Math. 129(1) (1972), 3573.CrossRefGoogle Scholar
Rizzoli, A., ‘Finite singular orbit modules for algebraic groups’, J. Algebra 570 (2021), 75118. issn: 0021-8693. doi: https://doi.org/10.1016/j.jalgebra.2020.11.006.CrossRefGoogle Scholar
Rizzoli, A., ‘Modules for algebraic groups with finitely many orbits on totally singular 2-spaces’, J. Algebra 606 (2022), 104169. issn: 0021-8693. doi: https://doi.org/10.1016/j.jalgebra.2022.05.006.CrossRefGoogle Scholar
Rizzoli, A., Magma code used for computations. Version 0.1. 2024. url: https://github.com/alunik/generic-stabs.Google Scholar
Sato, M. and Kimura, T., ‘A classification of irreducible prehomogeneous vector spaces and their relative invariants’, Nagoya Math. J. 65 (1977), 1155. doi: 10.1017/S0027763000017633.CrossRefGoogle Scholar
Steinberg, R., Lectures on Chevalley Groups (University Lecture Series) vol. 66 (American Mathematical Society, Providence, RI, 2016). isbn: 978-1-4704-3105-1. doi: 10.1090/ulect/066.CrossRefGoogle Scholar
Stewart, D. I., The Reductive Subgroups of F (Memoirs of the American Mathematical Society) (American Mathematical Society, 2013). isbn: 9780821898734.Google Scholar
Zhijie, C., ‘A classification of irreducible prehomogeneous vector spaces over an algebraically closed field of characteristic 2 (I)’, Acta Math. Sin. 2(2) (1986), 168177. doi: 10.1007/BF02564878.CrossRefGoogle Scholar
Zhijie, C., ‘A classification of irreducible prehomogeneous vector spaces over an algebraically closed field of characteristic $p$ II’, Chinese Ann. Math. Ser. A 9(1) (1988), 1022. issn: 1000-8314.Google Scholar
Figure 0

Table 1 Generic $ts$-stabilizers for $ts$-small quadruples.

Figure 1

Table 2 $ts$-small quadruples with finitely many orbits on $\boldsymbol{\mathcal {G}_k(V)}$.

Figure 2

Table 3 Infinite families of $ts$-small quadruples.

Figure 3

Table 4 Remaining $ts$-small quadruples.

Figure 4

Table 5 $ts$-small quadruples with $k=1$ and V symplectic.

Figure 5

Table 6 Remaining $ts$-small quadruples with known generic $ts$-stabilizer.

Figure 6

Table 7 Maximal subgroups $M_q$ of $Sp_4(q)$ (q odd).

Figure 7

Table 8 Maximal subgroups $M_q$ of $SL_2(q)$.