1 Introduction

In this paper, motivated by recent and less recent results about the Möbius function \(\mu \) for the subgroup lattice \(\mathcal {L}(G)\) of a finite group G, we give a result which relates the Möbius function for a subgroup G of \(\textrm{GL}(n,q)\) to two simplicial complexes: one defined from the lattice of the subspaces fixed by a reducible subgroup \(H\le G\) and the second from the lattice \(\mathcal {L}(G)\) of the subgroups of G.

We will introduce in the Preliminaries all the definitions and details useful for reading the paper.

In his PhD thesis [8], Shareshian considers the problem of computing \(\mu (1,G)\) for several finite classical groups G; the idea is to approximate \(\mu (1,G)\) through a good function \(f_{G,n,p}(u,1)\,\), such that:

$$\begin{aligned} \mu (1,G(n,\,p^u))=f_{G,n,p}(u,1)+\sum _{K\in \mathcal {C}_9}\mu (1,K)\,. \end{aligned}$$
(1)

Here, \(G=G(n,p^u)\) denotes a family of finite classical groups with the same defining classical form, which act in a natural way on the vector space V of finite dimension n over the finite field \(\mathbb {F}_q\) of order \(q=p^u\,\). If \(\,\mathcal {C}_1,\dots ,\mathcal {C}_8,\mathcal {C}_9\,\) are the Aschbacher classes of maximal subgroups of a finite classical group (see [4]), \(\mathcal {C}_9\) is the class of almost-simple groups not belonging to the first 8 classes of “geometric type” and the function \(f_{G,n,p}(u,1)\,\) provides an estimate of \(\mu (1,G)\) with respect to the contributions given by the subgroups of G which belong to the classes \(\mathcal {C}_i\,\), for \(i\in \{1,\dots ,8\}\).

Actually, Shareshian’s approach focuses on the first class \(\mathcal {C}_1(G)\), that is, the class of reducible subgroups of G.

In particular, the reducible subgroups of G contribute to \(f_{G,n,p}(u,1)\) through the computation of the Möbius function of

$$\begin{aligned} \widehat{\mathcal {I}}_1(G)&:=\{ K\le G\mid K\le M \text { for some } M\in \mathcal {C}_1(G) \} \cup \{G\}\,, \end{aligned}$$

which is obtained by adjoining the maximum G to the order ideal

$${\mathcal {I}}_1(G):=\{ K\le G\mid K\le M \text { for some } M\in \mathcal {C}_1(G) \},$$

that is,

$$\mu _{\widehat{\mathcal {I}}_1(G)}(1,G)= -\sum _{\begin{array}{c} K\in \mathcal {I}_1(G) \end{array}}\mu (1,K)\,$$

and

$$\begin{aligned} \mu (1,G)=\mu _{\widehat{\mathcal {I}}_1(G)}(1,G) -\sum _{\begin{array}{c} K<G \\ K\notin \mathcal {I}_1(G) \end{array}}\mu (1,K)\,. \end{aligned}$$
(2)

In this paper, we will consider irreducible subgroups G of the general linear group \(\textrm{GL}(n,q)\), that is, groups of linear automorphisms of a vector space V of dimension n over the finite field \(\mathbb {F}_q\) with q elements, which fix no non-trivial subspace of V. In this hypothesis, we will take a reducible subgroup H of G (that is, H fixes some proper subspace of V) and we will work on the analogue of \(\mu _{\widehat{\mathcal {I}}_1(G)}(1,G)\), namely \(\mu _{\widehat{\mathcal {I}}(G,H)}(H,G)\), so that

$$\begin{aligned} \mu (H,G)=\mu _{\widehat{\mathcal {I}}(G,H)}(H,G) -\sum _{\begin{array}{c} K\notin \mathcal {I}(G,H)\\ H\le K<G \end{array}}\mu (H,K)\, \end{aligned}$$
(3)

(see Sects. 2 and 3 for all precise definitions).

The subject of this paper is somehow motivated by the following conjecture:

Conjecture 1.1

(Mann, [6]) Let G be a PFG group and \(\mu \) the Möbius function on the lattice of open subgroups of G. Then, \(|\mu (H,G)|\) is bounded by a polynomial function in the index |G : H| and the number of subgroups of G of index m with \(\mu (H,G)\ne 0\) grows at most polynomially in m.

Indeed, although the problem is still open in its general setting, it was reduced by Lucchini in [5] to the study of similar growth conditions for finite almost-simple groups.

The following theorem is the main result of the present paper:

Theorem 4.5

Consider a vector space V of finite dimension over \(\mathbb {F}_q\,\). Let G be an irreducible subgroup of \(\,\textrm{GL}(V)\) and \(H\le G\). Then,

$$\begin{aligned} -\mu _{\widehat{\mathcal {I}}(G,H)}(H,G)=\sum _{E\in \Psi '(G,H)}(-1)^{|E|}\,=\sum _{X\in \Psi (G,H)}(-1)^{|X|}\,=-\tilde{\chi }(\Delta _1 )=-\tilde{\chi }(\Delta _2). \end{aligned}$$
(4)

\({\widehat{\mathcal {I}}(G,H)}\), \({\Psi '(G,H)}\), \({\Psi (G,H)}\), \(\Delta _i\), and \(\tilde{\chi }(\Delta _i)\) (for \(i=1,\,2\)) are defined in Sect. 3, also for irreducible subgroups H. This will allow us to avoid the restriction to only reducible subgroups H of G in the statement of the Theorem.

In the final section, we will use Theorem 4.5 to deal with \(\mu (H,G)\) in some particular case. In [1], Theorem 4.5 is used to attack Conjecture 1.1 for some class of subgroups H of linear and projective groups.

We thank Andrea Lucchini and Johannes Siemons for many useful discussions.

2 Preliminaries

In this paper, all the groups and sets are finite.

For main results about posets and lattices, we refer to [9]. Here, we just recall some basic fact, useful for reading the paper.

Definition 2.1

Let \(\mathcal {P}\) be a finite poset. The Möbius function associated with \(\mathcal {P}\) is the map \(\,\mu _{\mathcal {P}}:\mathcal {P}\times \mathcal {P}\rightarrow \mathbb {Z}\) satisfying

$$\mu _{\mathcal {P}}(x,y)=0\quad \text { unless }\;x\le y,$$

and defined recursively for \(x\le y\,\) by

$$\begin{aligned} \mu _{\mathcal {P}}(x,x)=1\; \quad \text { and }\quad \; \sum \limits _{x\le t\le y}{\mu _{\mathcal {P}}(x,t)}=0 \;\text { if }\, x<y\,. \end{aligned}$$
(5)

Notation

If \(\mathcal {P}\) is the subgroup lattice \(\mathcal {L}(G)\) of G, we will write \(\mu (H,K)\) instead of \(\mu _{\mathcal {L}(G)}(H,K)\).

Definition 2.2

An (abstract) simplicial complex \(\Delta \) on a vertex set T is a collection \(\Delta \) of subsets of T satisfying the two following conditions:

  • if \(t\in T\), then \(\{t\} \in \Delta \);

  • if \(F\in \Delta \) and \(G\subseteq F\), then \( G\in \Delta \).

An element \(F\in \Delta \) is called a face of \(\Delta \), and the dimension of F is defined to be \(|F|-1\). In particular, the empty set \(\emptyset \) is a face of \(\Delta \) (provided \(\Delta \ne \emptyset \)) of dimension \(-1\).

Definition 2.3

The Euler characteristic \(\chi (\Delta )\) of the simplicial complex \(\Delta \) is so defined:

$$\chi (\Delta ):=\, \sum _{i\ge 0}(-1)^i\mathcal {F}_i\,=\,\mathcal {F}_0 - \mathcal {F}_1 + \mathcal {F}_2 - \mathcal {F}_3 + \dots $$

where \(\mathcal {F}_i\) denotes the number of the i-faces that is of the faces of dimension i, in \(\Delta \).

We recall here the definition of an order ideal:

Definition 2.4

Let \((\mathcal {P},\le )\) be a poset. An order ideal of \(\mathcal {P}\) is a subset \(I\subseteq \mathcal {P}\) such that

$$\begin{aligned} \forall \,x\in I,\; \forall \,t\in \mathcal {P}\qquad t\le x \,\Rightarrow \, t\in I\,. \end{aligned}$$
(6)

In particular, if A is a subset of \(\mathcal {P}\), then the set

$$\mathcal {P}_{\le A}:= \{s\in \mathcal {P} \mid s\le a \text { for some } a\in A\}\subseteq \mathcal {P}$$

is the order ideal of \(\mathcal {P}\) generated by A.

Observe that a simplicial complex on a set of vertices T is just an order ideal of the Boolean algebra \(B_T\) that contains all one element subsets of T.

Definition 2.5

Let \(\Delta \) be a simplicial complex, and \(\chi (\Delta )\) be the Euler characteristic of \(\Delta \). The reduced Euler characteristic \(\tilde{\chi }(\Delta )\) of \(\Delta \) is defined by \(\tilde{\chi }(\emptyset )=0\), and \(\tilde{\chi }(\Delta )=\chi (\Delta )-1\) if \(\Delta \ne \emptyset \). See also Equation (3.22) in [9].

It is possible to build a simplicial complex from a poset \(\mathcal {P}\) in the following way:

The order complex \(\Delta (\mathcal {P})\) of \(\mathcal {P}\) is defined as the simplicial complex whose vertices are the elements of \(\mathcal {P}\) and whose k-dimensional faces are the chains \(a_0\prec a_1\prec \cdots \prec a_k\) of length k of distinct elements \(a_0,\ldots ,a_k\in \mathcal {P}\).

Now, denote by \(\widehat{\mathcal {P}}\) the finite poset obtained from \(\mathcal {P}\) by adjoining a least element \(\hat{0}\) and a greatest element \(\hat{1}\). The Möbius function \(\mu _{\widehat{\mathcal {P}}}(\hat{0},\hat{1})\) is related to \(\tilde{\chi }(\Delta (\mathcal {P}))\) by a well-known result by Hall in [3] about the computation of \(\mu _{\widehat{\mathcal {P}}}(\hat{0},\hat{1})\) by means of the chains of even and odd length between \(\hat{0}\) and \(\hat{1}\).

Proposition 2.6

(see [9], Proposition 3.8.6) Let \(\mathcal {P}\) be a finite poset. Then,

$$ \mu _{\widehat{\mathcal {P}}}(\hat{0},\hat{1})=\tilde{\chi }(\Delta (\mathcal {P})). $$

3 The ideal \(\mathcal {I}(G,H)\) and the complexes \(\Delta _i\) in Theorem 4.5

In this section, and in the rest of the paper, G is an irreducible subgroup of the general linear group \(\textrm{GL}(n,q)\) over the vector space \(V=\mathbb {F}_q^n\) of finite dimension n over the finite field \(\mathbb {F}_q\) with q elements. We consider the natural action of G on the set of subspaces of V.

We define the order ideal \(\mathcal {I}(G,H)\) and the simplicial complexes \(\Delta _i\) (for \(i=1,2\)) that we consider in Theorem 4.5. In Remark 2, we will explicitly observe that the two complexes \(\Delta _i\) are not the order simplicial complex rising from \(\mathcal {I}(G,H)\).

Given a subgroup G of \(\textrm{GL}(n,q)\) and a subgroup H of G, put

$$\mathcal {L}(G)_{\ge H}:=\{K\le G\,\mid \,H\le K\}$$

and

$$\begin{aligned} \mathcal {C}(G,H)&:=\{\,\textrm{stab}_G(W)\mid 0< W< V\,,\,H\subseteq \textrm{stab}_G(W) \,\}. \end{aligned}$$

Definition 3.1

The reducible subgroup ideal in \(\mathcal {L}(G)_{\ge H}\) is the order ideal of \(\mathcal {L}(G)_{\ge H}\) generated by \(\mathcal {C}(G,H)\,\). Namely,

$$\begin{aligned} \mathcal {I}(G,H)&=\{ K\le G\mid H\le K\le M \text { for some } M\in \mathcal {C}(G,H) \}\,. \end{aligned}$$

Remark 1

If H is reducible, that is, H fixes some non-trivial subspace W of V, then \(H\in \mathcal {I}(G,H)\). Otherwise, if H is irreducible, \(H\notin \mathcal {I}(G,H)\) and \(\mathcal {I}(G,H)\) is empty. The subspaces fixed by H are said to be H-invariant.

Definition 3.2

If H is reducible, we set

$$\begin{aligned} \widehat{\mathcal {I}}(G,H)&:= \mathcal {I}(G,H)\cup \{G\}\, \end{aligned}$$

by adjoining the maximum G to \(\mathcal {I}(G,H)\), which has minimum H. Otherwise, if H is irreducible, we set \(\widehat{\mathcal {I}}(G,H):=\{H,G\}\) by adjoining the minimum H and the maximum G to the empty poset \(\emptyset \).

Remark 2

Here, we just note that the poset \(\mathcal {I}(G,H)\) has already a minimum if H is reducible, so that \(\mu _{\widehat{\mathcal {I}}(G,H)}(H,G)\) is not, in general, the reduced Euler characteristic of the order complex \(\Delta ({\mathcal {I}}(G,H))\) of \(\mathcal {I}(G,H)\).

To define the simplicial complexes \(\Delta _i\) of Theorem 4.5, we begin with fixing some more notation. We denote by \(\mathcal {S}(V,H)\) the lattice of H-invariant subspaces of V and define

$$\mathcal {S}(V,H)^*:=\mathcal {S}(V,H)\setminus \{0,V\}.$$

Moreover, given an irreducible group \(G\le \textrm{GL}(V)\), and \(H\le G\), we will consider the following three sets:

  1. (a)

    \(\Psi (G,H):=\{X\subseteq \mathcal {C}(G,H)\,\mid \,\bigcap _{M\in X}M\ne H\,\}\,\);

  2. (b)

    \(\Psi (G,H)^\complement :=\{Y\subseteq \mathcal {C}(G,H)\,\mid \, \bigcap _{M\in Y}M= H\,\}\,\);

  3. (c)

    \(\Psi '(G,H):=\{E\subseteq \mathcal {S}(V,H)^*\,\mid \,\bigcap _{W\in E}\textrm{stab}_G(W)\ne H\}.\)

Observe that \(\emptyset \in \Psi (G,H)\) and \(\emptyset \in \Psi '(G,H)\), but \(\emptyset \notin \Psi (G,H)^\complement \,\).

Remark 3

If H is an irreducible subgroup of G, then \(\mathcal {S}(V,H)^*=\emptyset \) and \(\Psi '(G,H)=\{\emptyset \}=\Psi (G,H)\). Since for irreducible H we have that \(\widehat{\mathcal {I}}(G,H)=\{H,G\}\), then Theorem 4.5 is trivially verified in this case.

Definition 3.3

The simplicial complexes \(\Delta _i\) of Theorem 4.5 are so defined:

\(\Delta _1\):

  • The set of vertices \(T_1\) is given by the subspaces \(W\in \mathcal {S}(V,H)^*\) for which \(H\ne \textrm{stab}_G(W)\);

  • the set of faces of \(\Delta _1 \) is given by \(\Psi '(G,H)\).

\(\Delta _2\):

  • The set of vertices \(T_2\) is given by the subgroups \(M\in \mathcal {C}(G,H)\) such that \(H\ne M\);

  • the set of faces of \(\Delta _2 \) is given by \(\Psi (G,H)\).

We explicitly observe what happens in the special case when, for some proper non-trivial subspace W of V, the subgroup \(H=\textrm{stab}_G(W)\) is maximal with respect to the property of being a stabilizer of a proper non-trivial subspace of V in G. In this case, we note that, by definition, \(T_1=T_2=\emptyset \) and the set of faces of \(\Delta _1\) and \(\Delta _2\) is \(\{\emptyset \}\). Then, Theorem 4.5 is trivially verified.

4 Computing \(\mu _{\widehat{\mathcal {I}}(G,H)}(H,G)\)

In order to prove Theorem 4.5, we need Proposition 4.2 that gives a link between \(\Psi (G,H)\) and \(\Psi '(G,H)\) and also shows that the reduced Euler characteristics of the complexes \(\Delta _i\) coincide. The proof of Proposition 4.2 follows at once from Lemma 4.1.

Lemma 4.1

Let T be a subgroup of a finite group L acting on a finite set X, and let \(X'\subseteq X\) be a subset such that \(T\le L_x\,\) for all \(x\in X'\). (As usual, \(L_x\) denotes the stabilizer of x in L.) Set

  • \(\mathcal {L}:=\{L_x\,\mid \,x\in X'\}\,\);

  • \(\mathcal {R}:=\{E\subseteq {\mathcal {L}}\,\mid \, \bigcap _{K\in E}K\ne T\}\,\);

  • \(\mathcal {S}:=\{Q\subseteq X'\,\mid \, \bigcap _{x\in Q}L_x\ne T\}\,\).

Then,

$$\sum _{E\in \mathcal {R}}(-1)^{|E|}=\sum _{Q\in \mathcal {S}}(-1)^{|Q|}.$$

Proof

If \(Q\in \mathcal {S}\) and \(E\in \mathcal {R}\), set

$$\,\mathcal {L}_Q=\{L_x\mid x\in Q\}\;\text { and }\; \mathcal {S}_E=\{Q\in \mathcal {S}\, \mid \, E=\,\mathcal {L}_Q\}.$$

It is immediate to realize that

$$\mathcal {S}=\bigsqcup _{E\in \mathcal {R}}\mathcal {S}_E$$

is the disjoint union of all the \(\mathcal {S}_E\). So, it suffices to show that for each \(E\in \mathcal {R}\) the following identity is verified:

$$\begin{aligned} (-1)^{|E|}=\sum _{Q\in \mathcal {S}_E}(-1)^{|Q|}\,. \end{aligned}$$
(7)

For \(E=\emptyset \), the identity (7) is trivially true. Now, fix a non-empty \(E\in \mathcal {R}\), and for each \(K\in E\) define

$$X'_K=\{x\in X'\,\mid \, L_x=K\}\subseteq X'.$$

Let \(Q\in \mathcal {S}_E\) and observe that Q can be represented as the following disjoint union:

$$Q=\bigsqcup _{K\in E}Q_K,$$

where \(\,Q_K=\{x\in Q\,\mid \, L_x=K\}\subseteq X'_K\,\) and \(Q_K\ne \emptyset \). (This property characterizes the elements Q of \(\mathcal {S}_E\).) Since \(\sum _{\emptyset \ne Q_K\subseteq X'_K}(-1)^{|Q_K|}{=-1}\) (see Remark 4), we get

$$\begin{aligned}\sum _{Q\in \mathcal {S}_E}(-1)^{|Q|}&=\sum _{Q\in \mathcal {S}_E}(-1)^{\sum _{K\in E} |Q_K|}=\prod _{K\in E}\left( \sum _{\emptyset \ne Q_K\subseteq X'_K}(-1)^{|Q_K|}\right) \\&=\prod _{K\in E}(-1)=(-1)^{|E|}\,\end{aligned}$$

and we obtain the identity (7). \(\square \)

Proposition 4.2

Let V be a vector space of finite dimension over \(\mathbb {F}_q\,\) and consider a subgroup \(H\le G\le \textrm{GL}(V)\). We have:

$$\begin{aligned} \sum _{E\in \Psi '(G,H)}(-1)^{|E|}=\sum _{{X}\in \Psi (G,H)}(-1)^{|{X}|}\,. \end{aligned}$$
(8)

Equivalently, \(\tilde{\chi }(\Delta _1 )=\tilde{\chi }(\Delta _2)\).

Proof

Consider the natural action of G on the set of subspaces of V. By Lemma 4.1, the equality follows at once from the definitions of \(\mathcal {S}(V,H)^*\), \(\mathcal {C}(G,H)\), \(\Psi (G,H)\) and \(\Psi '(G,H)\,\).

With previous notation, we take \(T=H\), \(L=G\), \(\mathcal {R}=\Psi (G,H)\), \(\mathcal {S}=\Psi '(G,H)\). \(\square \)

To prove Theorem 4.5, we need Proposition 4.4 which is achieved through Theorem 4.3 (Crosscut Theorem, see [9, Corollary 3.9.4]) and Remark 4.

Remark 4

For every finite set A of cardinality \(n>0\), we have

$$\sum _{S\subseteq A}(-1)^{|S|}=\sum _{k=0}^n\left( {\begin{array}{c}n\\ k\end{array}}\right) (-1)^k=(1-1)^n=0.$$

Theorem 4.3

(Crosscut Theorem) Let L be a finite lattice with minimum \(\hat{0}\) and maximum \(\hat{1}\), so that \(\hat{0}\ne \hat{1}\). Let M be the set of all coatoms in L. Let \(X\subseteq L\) be a subset such that \(M \subseteq X\) and \(\hat{1} \notin X\).

Given \(\mathcal {Y}:= \{Y\subseteq X\,\mid \, Y\ne {\emptyset }\, \text { and }\, \bigcap _{Y\in \mathcal {Y}}Y=\hat{0}\}\), the following equality holds:

$$\mu _L(\hat{0}, \hat{1}) =\sum _{Y\in \mathcal {Y}}(-1)^{\mid Y\mid }.$$

Proposition 4.4

Let V be a vector space of finite dimension over \(\mathbb {F}_q\,\). Let \(H\le G\le \textrm{GL}(V)\). Then, we have that

$$\begin{aligned} \mu _{\widehat{\mathcal {I}}(G,H)}(H,G)=\sum _{Y\in \Psi (G,H)^\complement }(-1)^{|Y|}\,. \end{aligned}$$
(9)

Proof

We observe that \(\widehat{\mathcal {I}}(G,H)\subseteq \mathcal {L}(G)_{\ge H}\) is a lattice because the join of two subgroups \(K_1,K_2\in \mathcal {I}(G,H)\) is either in \(\mathcal {I}(G,H)\) or equal to G. The meet of \(K_1,K_2\in \mathcal {I}(G,H)\) is \(K_1\cap K_2\,\). Hence, \(\widehat{\mathcal {I}}(G,H)\) is a finite lattice, whose set of coatoms is contained in \(\mathcal {C}(G,H)\), and its maximum \(G\notin \mathcal {C}(G,H)\), because G is assumed to be irreducible. Since

$$\Psi (G,H)^\complement =\{Y\subseteq \mathcal {C}(G,H)\,\mid \,Y\ne \emptyset \;\text { and }\; \bigcap _{M\in Y}M= H\,\},$$

by Theorem 4.3 we immediately obtain (9). \(\square \)

Now, observe that the disjoint union \(\Psi (G,H)\cup \Psi (G,H)^\complement \) is the power set of \(\mathcal {C}(G,H)\), so that by Remark 4 we have

$$\begin{aligned} \sum _{X\in \Psi (G,H)}(-1)^{|X|} + \sum _{Y\in \Psi (G,H)^\complement }(-1)^{|Y|} = 0. \end{aligned}$$
(10)

If we put together Eqs. (8), (9), and (10), we get:

Theorem 4.5

Let V be a vector space of finite dimension over \(\mathbb {F}_q\,\). Let \(H\le G\le \textrm{GL}(V)\). Then,

$$\begin{aligned} -\mu _{\widehat{\mathcal {I}}(G,H)}(H,G)=\sum _{E\in \Psi '(G,H)}(-1)^{|E|}\,. \end{aligned}$$
(11)

Proof

By (10), we have

$$\sum _{{X}\in \Psi (G,H)}(-1)^{|{X}|}=-\sum _{{Y}\in \Psi (G,H)^\complement }(-1)^{|{Y}|}.$$

Then, by Proposition 4.2 and Proposition 4.4,

$$\begin{aligned} \sum _{E\in \Psi '(G,H)}(-1)^{|E|} =\sum _{{X}\in \Psi (G,H)}(-1)^{|X|}&=-\sum _{{Y}\in \Psi (G,H)^\complement }(-1)^{|{Y}|} \\ \\&=-\mu _{\widehat{\mathcal {I}}(G,H)}(H,G)\,. \end{aligned}$$

\(\square \)

5 Final remark

Going back to Conjecture 1.1, we observe that the knowledge of

$$\sum _{E\in \Psi '(G,H)}(-1)^{|E|}\,$$

coming from (11) can be exploited to estimate the value \(\mu (H,G)\) of the Möbius function \(\mu \) of G for \(H\le G\), at least in some particular case. Here, we just give an example for a particular reducible H, taking \(G=\textrm{GL}(n,q)\).

Following the idea suggested by Shareshian in [8], one could write

$$\begin{aligned} \mu (H,G)=f_{G,n,q}(H)+\sum _{K\in \mathcal {C}_9}\mu (H,K)\,, \end{aligned}$$
(12)

where \(f_{G,n,q}(H)\) depends on the classes \(\mathcal {C}_i(G,H)\), for \(i=1,\dots ,8\), in Aschbacher’s classification. In some lucky case, H is not contained in maximal subgroups belonging to classes \(\mathcal (C_i),\, i\ne 1,\, 9\). This happens, for example, in the following case:

Let \(V\simeq \mathbb {F}_q^n\,\) be a vector space of finite dimension n over \(\mathbb {F}_q\,\) and fix the following basis of \(V\,\):

$$\mathcal {E}:=\{w_1,\dots ,w_{m},\, v_{m+1},\dots , v_{n}\},$$

so that

$$\begin{aligned} V = \langle w_1,\dots ,w_{m}\rangle \oplus \langle v_{m+1},\dots , v_n\rangle \, \end{aligned}$$

If \(W=\langle w_1,\dots ,w_{m}\rangle \), H is the subgroup of \(\textrm{GL}(n,q)\) acting as \(\textrm{GL}(m,q)\) on W and fixing all the elements \(v_{m+1},\dots , v_n\).

We do not give in this context the details of the proof of the following theorem, which needs many technical arguments. In [1], all the details are given.

Theorem 5.1

([2]) Let \(G=\textrm{GL}(n,q)\), and let \(H\le G\) be such that

$$H=\textrm{GL}(m,q)\oplus I_{n-m}.$$

Let \(q=p\,\) be an odd prime, and let the dimension n be prime. If \(\,n-m+1\,\) is prime, then

$$\mu _{\widehat{\mathcal {I}}(G)}(H,G)=0=f_{G,n,p}(H)\,$$

so that

$$\begin{aligned} \mu (H,G)=\sum _{\begin{array}{c} K\in \mathcal {C}_9(G,H)\,,\, H\subseteq K \end{array}}\mu (H,K)\,. \end{aligned}$$

In general, we do not have much information about the ninth class. Just to give an example, we considered the groups of low dimension studied by Schröder in her PhD thesis ([7]), and we saw that in dimension \(n=13\) also class \(\mathcal {C}_9(G,H)\) is empty for \(p>5\). In this case, \(\mu (H,G)=0\).

All the details and data needed to prove Theorem 5.1 are contained in [2] and are available if requested.