Classification of finite subgroups of PGL

6 The Maximal Abelian Subgroup Class Equation

6.1 A Finite Subgroup of \(\pmb {L}\)

We now return to the realm of finite groups and consider \(G\) to be an arbitrary finite subgroup of \(L\). We will still continue to use \(Z\) to denote the centre of \(L\), and will use \(Z(G)\) whenever we refer to the centre of \(G\).

Observe that if \(Z\) is not contained in \(G\), then \(Z\) must contain a non-identity element, thus \(|Z| = 2\) and \(p \neq 2\) by Lemma 5.16. Recall that \(L\) has a unique element of order 2 by Lemma 5.17, \(- I_L\), which is not in \(G\), therefore \(G\) has no element of order 2.

By Cauchy’s Theorem, which says that if a prime \(p\) divides the order of a finite group, then the group contains an element of order \(p\), we deduce that 2 does not divide the order of \(G\).

This means that \(|G|\) and \(|Z|\) are relatively prime, so \(G \cap Z = \{ I_L \} \) and we can use Corollary 3.21 to show that \(GZ \cong G \times Z\). This shows that regardless of whether \(G\) contains \(Z\) or not, its structure is uniquely determined by \(GZ\), so it suffices to only consider the case when \(Z \subset G\).

6.2 Maximal Abelian Subgroups

Definition 6.1 Maximal Abelian Subgroup
#

Let \(H\) and \(J\) be subgroups of a group \(G\) where \(H\) is abelian. \(H\) is called maximal abelian if \(J\) is not abelian whenever \(H \subsetneq J\).

Definition 6.2 Elementary Abelian
#

A group \(G\) is said to be elementary abelian if it is abelian and every non-trivial element has order \(p\), where \(p\) is prime.

Definition 6.3
#

Let \(\mathfrak {M}\) denote the set of all maximal abelian subgroups of \(G\).

Maximal abelian subgroups play an important role in determining the structure of \(G\). In particular, every element in \(G\) must be contained in some maximal abelian subgroup, since every element commutes at least with itself and \(Z\). This will allow us to decompose \(G\) into the conjugacy classes of these maximal abelian subgroups. Note also that unless \(G=Z\), \(Z\) is not a maximal abelian subgroup, because for each \(x \in G \! \setminus \! Z\), \(\langle Z,x \rangle \) is clearly a larger abelian subgroup than \(Z\).

We will shortly prove an important theorem regarding the maximal abelian subgroups of \(G\), but in order to do so we require the following two lemmas.

Lemma 6.4

If \(G\) is a finite group of order \(p^m\) where \(p\) is prime and \(m{\gt}0\), then \(p\) divides \(|Z(G)|\).

Proof

Let \(C(x)\) be the set of elements of \(G\) which are conjugate in \(G\) to \(x\), we call this the conjugacy class of \(x\). Bhattacharya shows that the set of all conjugacy classes form a partition of \(G\) [ 2 , p.112 ] . Now consider the following rearranged class equation of \(G\), where \(S\) is a subset of \(G\) containing exactly one element from each conjugacy class not contained in \(Z(G)\).

\begin{equation} \label{cen2} |G| - \sum _{x \in S} [G:N_G(x)] = |Z(G)|. \end{equation}
1

Since \(|G| = p^m\), each subgroup of \(G\) is of order \(p^k\) for some \(k \leq m\). In particular each \(N_G(x)\) has order \(p^k\) and is strictly contained in \(G\) since \(x \not\in Z(G)\) by assumption. Thus each \([G:N_G(x)] {\gt} 1\), and are therefore divisible by \(p\). Since \(p\) divides the left hand side of (1), it must also divide the right, thus \(p\) divides \(|Z(G)|\).

Lemma 6.5

Every finite subgroup of a multiplicative group of a field is cyclic.

Proof

See [ .

Theorem 6.6

Let \(G\) be an arbitrary finite subgroup of \(L\) containing \(Z\).
(i) If \(x \in G \! \setminus \! Z\) then we have \(C_G(x) \in \mathfrak {M}\).
(ii) For any two distinct subgroups \(A\) and \(B\) of \(\mathfrak {M}\), we have

\begin{align*} A \cap B = Z. \end{align*}

(iii) An element \(A\) of \(\mathfrak {M}\) is either a cyclic group whose order is relatively prime to \(p\), or of the form \(Q \times Z\) where \(Q\) is an elementary abelian Sylow \(p\)-subgroup of \(G\).
(iv) If \(A \in \mathfrak {M}\) and \(|A|\) is relatively prime to \(p\), then we have \([N_G(A): A] \leq 2\). Furthermore, if \([N_G(A): A] = 2\), then there is an element \(y\) of \(N_G(A) \! \setminus \! A\) such that,

\begin{align*} yxy^{-1} = x^{-1} \qquad \forall x \in A.\end{align*}

(v) Let \(Q\) be a Sylow \(p\)-subgroup of \(G\). If \(Q \neq \{ I_G\} \), then there is a cyclic subgroup \(K\) of \(G\) such that \(N_G(Q) = QK\). If \(|K| {\gt} |Z|\), then \(K \in \mathfrak {M}\).

Proof

(i) Let \(x\) be chosen arbitrarily from \(G \! \setminus \! Z\). Then by Corollary 5.24, \(C_L(x)\) is abelian. By definition, \(C_G(x) = C_L(x) \cap G\), and using the elementary fact that the intersection of 2 groups is itself a group, we have \(C_G(x) {\lt} C_L(x)\). Now since every subgroup of an abelian group is abelian, \(C_G(x)\) is also abelian.

Now let \(J\) be a maximal abelian subgroup of \(G\) containing \(C_G(x)\). Since \(J\) is abelian and \(x \in C_G(x) \subset J\), we have \(jx=xj\), \(\forall j \in J\), thus \(J \subset C_G(x)\). Therefore \(J=C_G(x)\) and \(C_G(x) \in \mathfrak {M}\).

(ii) Consider \(x \in A \cap B\). Since both \(A\) and \(B\) are abelian, \(x\) commutes with each \(a \in A\) and \(b \in B\) and thus \(C_G(x)\) contains both \(A\) and \(B\). If \(x \in G \setminus Z\), then \(C_G(x) \in \mathfrak {M}\) by (i) and because \(A\) and \(B\) are distinct we have \(A \subsetneq A \cup B \subset C_G(x)\). This contradicts the fact that \(A\) is maximum abelian and thus \(x \in Z\). Finally, note that Z is contained in every maximal abelian subgroup, since otherwise we would have the contradiction that \(\langle A, Z \rangle \) would generate a larger abelian subgroup than \(A\). Hence \(A \cap B = Z\).

(iii) First consider the trivial case of \(G=Z\). Here \(G\) is the only element of \(\mathfrak {M}\). If \(p \neq 2\) then \(|G|=2\) and \(G\) is a cyclic group whose order is relatively prime to \(p\). If \(p=2\) then \(G = I_G\) which is trivially a \(S_p\)-subgroup.

Now assume \(G \neq Z\). Since \(Z \not\in \mathfrak {M}\), each \(A \in \mathfrak {M}\) contains at least one \(x \not\in Z\). By Proposition 5.18 this \(x\) is conjugate to either \(d_\omega \) or \(\pm t_\lambda \) in \(L\). It suffices to only consider these cases:

\(\pmb {x}\) conjugate to \(\pmb {d_\omega }\) in \(\pmb {L}\). There is a \(y \in L\) such that \(x = y d_\omega y^{-1}\). Since \(x \not\in Z\), we have \(d_\omega \not\in Z\), because otherwise we get the contradiction,

\begin{align*} x = y d_\omega y^{-1} = d_\omega \in Z. \end{align*}

Thus \(\omega \neq \pm 1\). Let \(A = C_G(x)\), since \(C_G(x) \in \mathfrak {M}\) by part (i). Observe that

\begin{align*} C_G(d_\omega ) & {\lt} C_L(d_\omega ) \tag {see proof of (i)} \\ & = D \tag {by Lemma \ref{6.4ii}} \\ & \cong F^*. \tag {by Lemma \ref{6.1b}} \end{align*}

Since \(A\) is conjugate to \(C_G(d_\omega )\) by Proposition 5.23, we have that \(A\) is isomorphic to a finite subgroup of \(F^*\) and by Lemma 6.5, \(A\) is cyclic. By Lagrange’s Theorem any finite subgroup of \(F^*\) has an order which divides \(p^m - 1\) for some \(m \in \mathbb {Z}^+\), and since \(p \nmid (p^m - 1)\), \(|A|\) is relatively prime to \(p\).

\(\pmb {x}\) conjugate to \(\pmb {\pm t_\lambda }\) in \(\pmb {L}\). Again let \(A = C_G(x) \in \mathfrak {M}\). \(A\) is conjugate to \(C_G({\pm t_\lambda })\) in \(L\) by Proposition 5.23. Since \(x \notin Z\), we have \(\lambda \neq 0\). Observe that

\begin{align*} C_G({\pm t_\lambda }) & {\lt} C_L({\pm t_\lambda }) \\ & = T \times Z \tag {by Lemma \ref{6.4i}} \\ & \cong F \times Z. \tag {by Lemma \ref{6.1b}} \end{align*}

So \(A\) is isomorphic to a finite subgroup of \(F \times Z\), call it \(Q \times Z\). Now \(A = Q \times Z \cong QZ\) by Corollary 3.21, which means that an arbitrary element of \(A\) is of the form \(q_1z_1\), where \(q_1 \in Q\), \(z_1 \in Z\).

\begin{align*} q_1z_1q_2z_2 & = q_2z_2 q_1z_1, \tag {$A \in \mathfrak {M}$} \\ q_1q_2z_1z_2 & = q_2q_1z_1z_2, \tag {$z_1$, $z_2 \in Z$} \\ q_1q_2z_1z_2(z_1z_2)^{-1} & = q_2q_1z_1z_2(z_1z_2)^{-1}, \\ q_1q_2 & = q_2q_1. \end{align*}

Thus \(Q\) is also abelian. Recall from the proof of Proposition 5.18(ii) that all non-trivial elements of \(T\) have order \(p\), so each non-trivial element of \(Q\) has order \(p\) which means that \(Q\) is elementary abelian. Thus \(Q\) has order \(p^m\), for some \(m \in \mathbb {Z}^+\).

Now let \(S\) be a Sylow \(p\)-subgroup containing \(Q\). We apply Lemma 6.4 to determine that \(p\) divides \(|Z(S)|\), moreover \(|Z(S)| \geq p\).

If \(p=2\), then \(Z=I_L\) by Lemma 5.16. So \(|Z| = 1\) and hence \(|Z(S)| \geq 2 {\gt} |Z|\).
If \(p {\gt} 2\), then \(Z = \langle - I_L \rangle \) also by Lemma 5.16. So \(|Z| = 2\) and again we get \(|Z(S)| {\gt} 2 = |Z|\).

So \(Z(S)\) must contain at least one element which is not in \(Z\), let \(y\) be one such element. Let \(s_1z_1\) be an arbitrary element of \(S \times Z\).

\begin{align*} (s_1z_1)y(s_1z_1)^{-1} & = (s_1z_1)y(z_1^{-1}s_1^{-1}) \\ & = s_1y(z_1z_1^{-1})s_1^{-1} \tag {since $y \in L$, $z_1 \in Z$} \\ & = y(s_1s_1^{-1}) \tag {since $s_1 \in S$, $y \in Z(S)$} \\ & = y \end{align*}

Thus \(s_1z_1 \in C_G(y)\) and since it was chosen arbitrarily, \(S \times Z \subset C_G(y)\). Also since \(y \in G \! \setminus \! Z\) we have \(C_G(y) \in \mathfrak {M}\) by part (i).

\begin{equation*} A = Q \times Z \subset S \times Z \subset C_G(y). \end{equation*}

Since \(A\) and \(C_G(y)\) are both in \(\mathfrak {M}\) it must be that \(A = C_G(y)\). This means \(Q = S\) and \(Q\) is a Sylow \(p\)-subgroup of G.

(iv) If \(|A| \leq 2\) then \(A=Z=G\). So \(A\) is trivially normal in \(G\) and \([N_G(A): A] = 1\).

Now assume that \(|A| {\gt} 2\). Since \(|A|\) is relatively prime to \(p\), we have that \(A\) is a cyclic group conjugate to a finite subgroup of \(D\) in \(L\) by the proof of part (iii), call this subgroup \({\widetilde{A}}\). Thus both \({\widetilde{A}}\) and \(D\) have orders greater than 2. Applying Proposition 5.22 we observe that

\begin{align} \label{norm1} N_L({\widetilde{A}}) = \langle D , w \rangle = N_L(D). \end{align}

Since \(A\) and \({\widetilde{A}}\) are conjugate in \(L\), there exists an element \(z \in L\) such that \(zAz^{-1} = {\widetilde{A}}\). This \(z\) determines an inner automorphism of \(L\) defined by

\begin{align*} i_z: L \longrightarrow L, \qquad \text{where} \quad i_z(t) = z t z^{-1} \quad \forall \; t \in L. \end{align*}

Let \(i_z(G) = {\widetilde{G}}\) denote the image of \(G\) under \(i_z\). Since \(A\) is a maximal abelain subgroup of \(G\) it’s a simple task to show that \({\widetilde{A}}\) is a maximal abelian subgroup of \({\widetilde{G}}\) and I will leave this to the reader to verify. We now show that \(i_z(N_G(A)) = N_{\widetilde{G}}({\widetilde{A}})\) . Take an arbitrary \(g \in N_G(A)\).

\begin{align*} (z g z^{-1}) {\widetilde{A}} (z g z^{-1})^{-1} & = z g (z^{-1} {\widetilde{A}} z) g^{-1} z^{-1} \\ & = z (g A g^{-1}) z^{-1} \tag {since $zAz^{-1} = {\widetilde{A}}$ } \\ & = z A z^{-1} \tag {since $g \in N_G(A)$} \\ & = {\widetilde{A}}. \end{align*}

So \(z g z^{-1} = i_z(g) \in N_{\widetilde{G}}({\widetilde{A}})\) and since it was chosen arbitrarily, \(i_z(N_G(A)) \subset N_{\widetilde{G}}({\widetilde{A}})\). Now take an arbitrary \(z h z^{-1} \in N_{\widetilde{G}}({\widetilde{A}})\).

\begin{align*} {\widetilde{A}} & = (z h z^{-1}) {\widetilde{A}} (z h z^{-1})^{-1} \\ & = z h (z^{-1} {\widetilde{A}} z) h^{-1} z^{-1} \\ & = z h A h^{-1} z^{-1}. \tag {since $A = z^{-1} {\widetilde{A}} z$} \end{align*}

Now multiplication on the left by \(z^{-1}\) and right by \(z\) gives:

\begin{align*} A = z^{-1} {\widetilde{A}} z = h A h^{-1}, \end{align*}

so \(h \in N_G(A)\). Furthermore, \(z h z^{-1}\) and indeed the whole of \(N_{\widetilde{G}}({\widetilde{A}})\) is contained in \(i_z(N_G(A))\). Thus \( i_z(N_G(A)) = N_{\widetilde{G}}({\widetilde{A}})\). In particular, we have,

\begin{align} \label{6.8iv1} [N_G(A): A] = [N_{\widetilde{G}}({\widetilde{A}}): {\widetilde{A}}]. \end{align}

Since \({\widetilde{G}} {\lt} L\), the normaliser of \({\widetilde{A}}\) in \({\widetilde{G}}\) is simply the normaliser of \({\widetilde{A}}\) in \(L\) restricted to \({\widetilde{G}}\), thus \(N_{\widetilde{G}}({\widetilde{A}}) {\lt} N_L({\widetilde{A}}) = N_L(D)\) by (2). Now since \(D \vartriangleleft N_L(D)\), the Second Isomorphism Theorem shows that,

\begin{align} \label{2iso} N_{\widetilde{G}}({\widetilde{A}})/( N_{\widetilde{G}}({\widetilde{A}}) \cap D) \; \cong \; DN_{\widetilde{G}}({\widetilde{A}}) / D. \end{align}


Clearly \({\widetilde{A}} \subset {\widetilde{G}} \cap D\). We show that this inclusion is infact an equality. Assume that there exists some \(d_\omega \in {\widetilde{G}} \cap D\) which is not in \({\widetilde{A}}\). The group \(\langle d_\omega , {\widetilde{A}} \rangle \) is thus an abelian subgroup of \({\widetilde{G}}\), strictly larger than \({\widetilde{A}}\) and contradicting the fact that \({\widetilde{A}}\) is maximal abelian in \({\widetilde{G}}\). Thus \({\widetilde{A}} = {\widetilde{G}} \cap D\). It is trivial to see that \({\widetilde{A}} \subset N_{\widetilde{G}}({\widetilde{A}}) \cap D\). Also \(N_{\widetilde{G}}({\widetilde{A}}) \cap D \subset {\widetilde{G}} \cap D = {\widetilde{A}}\). So,

\begin{align} \label{parti} {\widetilde{A}} = N_{\widetilde{G}}({\widetilde{A}}) \cap D. \end{align}

Observe also that,

\begin{align} \label{index1or2} DN_{\widetilde{G}}({\widetilde{A}}) = \{ D, \langle D, w \rangle \} \subset \langle D, w \rangle = N_L(D). \end{align}

Now we piece the preceding results together to give the desired result.

\begin{align*} N_{\widetilde{G}}({\widetilde{A}}) / {\widetilde{A}} \; & \cong \; N_{\widetilde{G}}({\widetilde{A}})/( N_{\widetilde{G}}({\widetilde{A}}) \cap D) \tag {by (\ref{parti})} \\ & \cong \; DN_{\widetilde{G}}({\widetilde{A}}) / D \tag {by (\ref{2iso})} \\ & \subset N_L(D) / D \tag {by (\ref{index1or2})} \\ & = \langle D, w \rangle / D \; \cong \; \mathbb {Z}_2. \end{align*}

We have shown that \(N_{\widetilde{G}}({\widetilde{A}}) / {\widetilde{A}}\) is isomorphic to a subset of \(\mathbb {Z}_2\). Thus by (3) we have established that,

\[ [N_G(A): A] = [N_{\widetilde{G}}({\widetilde{A}}): {\widetilde{A}}] \leq 2. \]

For the second part, if \([N_G(A): A] = 2\), then the above argument shows that \(N_{\widetilde{G}}({\widetilde{A}}) / {\widetilde{A}} \; \cong \; \mathbb {Z}_2\). Thus \(DN_{\widetilde{G}}({\widetilde{A}}) = N_L(D) = \langle D, w \rangle \). This means that \(N_{\widetilde{G}}({\widetilde{A}})\) contains some element \(wd_\omega \). In fact, since \(w d_\omega \not\in D\), we have \(w d_\omega \in N_{\widetilde{G}}({\widetilde{A}}) \! \setminus \! {\widetilde{A}}\). Take any element \(x \in A\). Since \({\widetilde{A}} = zAz^{-1}\), \(zxz^{-1} \in {\widetilde{A}}\), call it \(d_\sigma \). Let \(y = z^{-1}w d_\omega z\). Since \(wd_\omega \in N_{\widetilde{G}}({\widetilde{A}}) \! \setminus \! {\widetilde{A}}\) it follows that \(y \in N_G(A)\! \setminus \! A\). We show that this \(y\) inverts \(x\):

\begin{align*} yxy^{-1} & = (z^{-1}w d_\omega z)(z^{-1} d_\sigma z)(z^{-1}d^{-1}_\omega w^{-1} z) \\ & = z^{-1} w d_\omega d_\sigma d^{-1}_\omega w^{-1} z \\ & = z^{-1} w d_\sigma w^{-1} z \\ & = z^{-1} d^{-1}_\sigma z \tag {by Lemma \ref{6.1}} \\ & = x^{-1}. \end{align*}

(v) By part (iii), \(Q\) is conjugate to a finite subgroup of \(T\) in \(L\). In fact, without loss of generality we can assume that \(Q \subset T\), moreoever \(Q \subset T \cap G\). We show that this is in fact an equality by showing that the reverse inclusion also holds. Let \(t_\lambda \) be an arbitrary element of \(T \cap G\). Then \(\langle t_\lambda , Q \rangle \) is a \(p\)-group of \(G\) which must be equal to \(Q\) since it is a Sylow \(p\)-subgroup of \(G\). Thus \(t_\lambda \in Q\) and

\begin{align} \label{Q=TNG} Q = T \cap G. \end{align}

Since \(|Q| {\gt} 1\), Proposition 5.21 gives that \(N_G(Q) \subset N_L(Q) \subset H\). So \(N_G(Q) \subset H \cap G\). Now take an arbitrarily chosen \(d_\omega t_\lambda \in H \cap G\) and \(t_\mu \in Q\).

\begin{align*} (d_\omega t_\lambda ) t_\mu (d_\omega t_\lambda )^{-1} & = d_\omega ( t_\lambda t_\mu t_{-\lambda }) d^{-1}_\omega \\ & = d_\omega t_\mu d^{-1}_\omega \tag {by Lemma \ref{6.1}} \\ & = t_\sigma . \tag {where $\sigma = \mu \omega ^{-2}$, by Lemma \ref{6.1}} \end{align*}

Since it is a product of elements of \(G\), \(t_\sigma \in T \cap G = Q\) by (7). Thus \(d_\omega t_\lambda \in N_G(Q)\) and indeed the whole of \(H \cap G\) is contained in \(N_G(Q)\) and

\begin{align} \label{normQ=HNG} N_G(Q) = H \cap G. \end{align}

We now define a map \(\phi \) by,

\begin{align*} \phi : N_G(Q) \longrightarrow D, \qquad \text{where} \quad \! \phi (d_\omega t_\lambda ) = d_\omega \quad \forall \; d_\omega t_\lambda \in N_G(Q). \end{align*}

Next we determine the kernel of \(\phi \).

\begin{align*} ker(\phi ) & = \{ d_\omega t_\lambda \in N_G(Q) : \phi (d_\omega t_\lambda ) = I_G \} \\ & = N_G(Q) \cap T \\ & = H \cap G \cap T \tag {by (\ref{normQ=HNG})} \\ & = T \cap G = Q. \tag {by (\ref{Q=TNG})} \end{align*}

We show that \(\phi \) is a group homomorphism. Take \(d_\omega t_\lambda \), \(d_\rho t_\mu \) from \( N_G(Q)\).

\begin{align*} \phi (d_\omega t_\lambda d_\rho t_\mu ) & = \phi (d_\omega d_\rho t_\sigma t_\mu ) \tag {where $\sigma = \lambda \rho ^2$, by Lemma \ref{6.1}} \\ & = d_\omega d_\rho \\ & = \phi (d_\omega t_\lambda ) \phi (d_\rho t_\mu ). \end{align*}

Thus by the First Isomorphism Theorem,

\begin{align} \label{6.8viso} N_G(Q) / Q & \cong \phi (N_G(Q)), \end{align}

Since \(N_G(Q)\) is a finite group, it’s image under \(\phi \) is thus a finite subgroup of \(D\). Furthermore, since \(D \cong F^*\) (by Lemma ??), \(\phi (N_G(Q))\) is a cyclic group whose order divides \(p^m-1\) and is therefore relatively prime to \(p\), and by 9, so too is \(N_G(Q) / Q\).

Let \(r\) be the order of \(N_G(Q) / Q\). Since it is cyclic, \(N_G(Q)/Q\) is generated by a single element, namely a coset of \(Q\) in \(N_G(Q)\), call it \(kQ\). So \(|kQ| = r\). Observe that,

\begin{align*} (kQ)^r & = Q, \\ k^rQ & = Q, \\ k^r & \in Q. \end{align*}

Since \(Q\) is elementary abelian, each of it’s non-trivial elements has order \(p\), so \(k\) has order \(r\) or \(rp\). In either case, since gcd\((r,p)=1\), the order of \(k^p\) is \(r\). Let \(K = \langle k^p \rangle \). Now \(|K| = r\) and

\begin{align*} |N_G(Q)| & = r|Q| \\ & = |K||Q| \\ & = |QK|. \tag {since $Q \cap K = I_G$} \end{align*}

Thus,

\begin{align} \label{QK} N_G(Q) & = QK. \end{align}

Now assume \(|K| {\gt} |Z|\). Since \(K\) is abelian, it must be contained in some maximal abelian group \(A \in \mathfrak {M}\). By part (iii), \(A\) must also be a cyclic group whose order is relatively prime to \(p\).

Since \(A\) is conjugate in \(L\) to a subgroup of \(D\), each non-central element of \(A\) has exactly 2 fixed points on the projective line \(\mathscr {L}\) by Proposition 5.28. Let \(A = \langle x \rangle \) and let \(P_1\) and \(P_2\) be the points fixed by \(x\). We show by induction on \(n\) that \(x^n\) also fixes \(P_1\) and \(P_2\), for all \(n \in \mathbb {Z^+}\). We do this by assuming first that \(x^{n-1}\) fixes \(P_i\).

\begin{align*} x^n P_i = x(x^{n-1} P_i) = x (P_i) = P_i. \end{align*}

The importance of this is that since each element of \(A\) can be expressed as some power of \(x\), they must have the same two fixed points, namely \(P_1\) and \(P_2\). In other words,

\begin{align} \label{stab} A \subset S_L(P_i), \qquad (\text{$i$ = 1 or 2}) \end{align}

By Proposition 5.28(ii), each element of \(T\) has a common fixed point \(P\) and Stab\((P) = H\). Since \(K \subset H\), each element in \(K\) fixes \(P\). Also, since \(K \subset A\), this \(P\) must be equal to either \(P_1\) or \(P_2\). Therefore by (11), \(A \subset \text{Stab}(P) = H\). We arrive at the following result:

\begin{align*} A & \subset H \cap G \\ & = N_G(Q) \tag {by (\ref{normQ=HNG})} \\ & = QK. \tag {by (\ref{QK})} \end{align*}

Furthermore, we get,

\begin{align*} A & = QK \cap A \\ & = QK \cap AK \tag {$K \subset A$ so $A = AK$} \\ & = (Q \cap A)K \\ & = K \tag {$Q \cap A = I_G$} \end{align*}

Thus \(K \in \mathfrak {M}\).

For the duration of this paper, unless otherwise stated, \(Q\) will denote a Sylow \(p\)-subgroup of \(G\) and \(K\) will be as described above.

6.3 Conjugacy of Maximal Abelian Subgroups

Definition 6.7
#

The set \(\mathcal{C}_i = \{ x A_i x^{-1} : x \in G \} \) is called the conjugacy class of \(A_i \in \mathfrak {M}\).

Definition 6.8
#

Let \(A_i^*\) be the non-central part of \(A_i \in \mathfrak {M}\), let \(\mathfrak {M}^*\) be the set of all \(A_i^*\) and let \(\mathcal{C}_i^*\) be the conjugacy class of \(A_i^*\).

For some \(A_i \in \mathfrak {M}\) and \(A_i^* \in \mathfrak {M}^*\) let,

\begin{align*} C_i = \bigcup \limits _{x \in G} x A_i x^{-1}, \quad \text{and} \quad C_i^* = \bigcup \limits _{x \in G} x A_i^* x^{-1}. \end{align*}

In other words, \(C_i\) denotes the set of elements of \(G\) which belong to some element of \(\mathcal{C}_i\). It’s evident that \(C_i^* = C_i \setminus Z\) and that there is a \(C_i\) corresponding to each \(\mathcal{C}_i\). Clearly we have the relation,

\begin{align} \label{orderorder} |C_i^*| = |A_i^*||\mathcal{C}_i^*|. \end{align}
Theorem 6.9

Let \(G\) be a finite subgroup of \(L\) and \(S\) be a subset of \(\mathfrak {M}^*\) containing exactly one element from each of its conjugacy classes.

(i) The set of \(C_i^*\) form a partition of \(G \! \setminus \! Z\). That is,

\begin{align*} G \! \setminus \! Z = \bigcup \limits _{A_i^* \in S} C_i^*, \qquad \text{and} \qquad C_i^* \cap C_j^* = \varnothing , \qquad \forall \; i \neq j. \end{align*}

(ii)   \(|\mathcal{C}_i^*| = |\mathcal{C}_i|\).

(iii)   \(|\mathcal{C}_i| = [G : N_G(A_i)]\).

(iv)

\[ |G \! \setminus \! Z| = \sum _{A_i^* \in S} |A_i^*| [G:N_G(A_i)]. \]
Proof

(i) Define a relation \(\sim \) on \(\mathfrak {M}^*\) as follows:

\begin{align*} A_i^* \sim A_j^* \quad \text{if} \quad A_i^* = xA_j^*x^{-1} \quad \text{for some} \quad x \in G. \end{align*}

If we choose \(x \in A_i^*\), then clearly \(A_i^* = A_i^*xx^{-1} = xA_i^*x^{-1}\), thus \(A_i^* \sim A_i^*\) and \(\sim \) is reflexive.

If \(A_i^* \sim A_j^*\), then \(\exists \; x \in G\) such that,

\begin{align*} A_i^*= xA_j^*x^{-1} \iff x^{-1}A_i^*x = A_j^* \iff A_j^* = yA_i^*y^{-1} \quad \text{for} \; y = x^{-1} \in G. \end{align*}

Thus \(A_j^* \sim A_i^*\) and \(\sim \) is symmetric.

If \(A_i^* \sim A_j^*\) and \(A_j^* \sim A_k^*\), then \(\exists \; x, y \in G\) such that,

\begin{align*} A_i^* = xA_j^*x^{-1} \; \text{and} \; A_j^* = yA_k^*y^{-1} \Rightarrow A_i^* = xyA_k^*y^{-1}x^{-1} = (xy)A_k^*(xy)^{-1}. \end{align*}

Thus \(A_i^* \sim A_k^*\) (since \(xy \in G\)), which shows that \(\sim \) is transitive and moreover an equivalence relation on \(\mathfrak {M}^*\).

The equivalence class of \(A_i^*\) in \(\mathfrak {M}^*\) therefore coincides with the set \(\mathcal{C}_i^* = \{ xA_i^*x^{-1} : x \in G \} \). Furthermore, this tells us that each \(A_i^*\) belongs to exactly one conjugacy class. Thus the conjugacy classes \(\mathcal{C}_i^*\) form a partition of \(\mathfrak {M}^*\),

\begin{align*} \mathfrak {M}^* = \bigcup \limits _{A_i^* \in S} \mathcal{C}_i^*, \qquad \text{and} \qquad \mathcal{C}_i^* \cap \mathcal{C}_j^* = \varnothing , \qquad \forall \; i \neq j. \end{align*}

Since the set of \(\mathcal{C}_i^*\) are pairwise disjoint, it follows that the set of \(C_i^*\) are also pairwise disjoint and we get the desired result,

\begin{align*} G \! \setminus \! Z = \bigcup \limits _{A_i^* \in S} C_i^*, \qquad \text{and} \qquad C_i^* \cap C_j^* = \varnothing , \qquad \forall \; i \neq j. \end{align*}

(ii) Let \(x A_i x^{-1} \in \mathcal{C}_i\) and \(x A_i^* x^{-1} \in \mathcal{C}_i^*\). Since \(x A_i x^{-1} \! \setminus \! Z = x A_i^* x^{-1}\), it is quite clear that,

\begin{align*} x A_i x^{-1} \in \mathcal{C}_i \iff x A_i^* x^{-1} \in \mathcal{C}_i^*. \end{align*}

Thus \(|\mathcal{C}_i^*| = |\mathcal{C}_i|\) as desired.

(iii) Now we define a map \(\phi \) by:

\begin{align*} \phi : \mathcal{C}_i & \longrightarrow G / N_G(A_i), \\ \phi (xA_ix^{-1}) & = xN_G(A_i). \tag {$\forall \; x \in G, \; A_i \in \mathfrak {M}$} \end{align*}

Clearly \(\phi \) is trivially surjective. We now show that it is both well-defined and injective.

\begin{align*} xN_G(A_i) = yN_G(A_i) & \iff y^{-1}xN_G(A_i) = N_G(A_i) \\ & \iff y^{-1}x \in N_G(A_i) \\ & \iff (y^{-1}x)A_i(y^{-1}x)^{-1} = A_i \\ & \iff y^{-1}xA_ix^{-1}y = A_i \\ & \iff xA_ix^{-1} = yA_iy^{-1}. \end{align*}

Hence \(\phi \) is well-defined and injective. This shows that \(\phi \) is a bijection proving that \(|\mathcal{C}_i| = [G:N_G(A_i)]\). This is a crucial result which shows that the number of maximal abelian subgroups conjugate to \(A_i\) is equal to the index of the normaliser of \(A_i\) in \(G\).

(iv) This follows directly from parts (i), (ii) and (iii) and 12.

\begin{align*} G \! \setminus \! Z & = \bigcup \limits _{A_i^* \in S} C_i^*, \qquad \text{and} \qquad C_i^* \cap C_j^* = \varnothing , \qquad \forall \; i \neq j, \\ |G \! \setminus \! Z| & = \sum _{A_i^* \in S} |C_i^*| = \sum _{A_i^* \in S} |A_i^*||\mathcal{C}_i^*| = \sum _{A_i^* \in S} |A_i^*||\mathcal{C}_i| \\ & = \sum _{A_i^* \in S} |A_i^*| [G:N_G(A_i)]. \end{align*}

This theorem proves that the non-central parts of the maximal abelian subgroups form a partition of the non-central part of \(G\). This will serve as a powerful tool in decomposing \(G\) and counting its elements.

6.4 Constructing The Class Equation

It is necessary to prove the following 2 short lemmas before we proceed further.

Lemma 6.10

\(N_G(A) =N_G(A^*)\).

Proof

(iii) Let \(x \in N_G(A^*)\). Take an arbitary \(a \in A = A^* \cup Z\). If \(a \in A^*\), then since \(x \in N_G(A^*)\), we have \(xax^{-1} \in A^* \subset A\). If \(a \in Z\), then \(xzx^{-1} = zxx^{-1} = z \in A\). Therefore \(x\) is in the normaliser of \(A\) and \(N_G(A^*) \subset N_G(A)\).

Conversely, take \(y \in N_G(A)\) and \(a \in A^*\). \(yay^{-1} \in A = A^* \cup Z\). If \(yay^{-1} \in Z\), then

\begin{align*} yay^{-1} & = z, \tag {some $z \in Z$} \\ a & = y^{-1}zy = y^{-1}yz = z \not\in A^*. \end{align*}

This contradicts the fact that \(a \in A^*\). Therefore \(yay^{-1} \in A^*\) and \(y \in N_G(A^*)\). Since \(y\) was chosen arbitrarily we get \(N_G(A) \subset N_G(A^*)\) and hence \(N_G(A) =N_G(A^*)\).

Lemma 6.11

\(N_G(Q \times Z) = N_G(Q)\).

Proof

If \(p= 2\) then \(Z = I_G\) and the result is trivial. Now assume \(p \neq 2\). Thus \(|Z| = 2\). Let \(x\) and \(q_1\) be arbitrarily chosen elements of \(N_G(Q)\) and \(Q\) respectively.

\begin{align*} xq_1x^{-1} & = q_2, \tag {for some $q_2 \in Q$} \\ xq_1x^{-1}z_1 & = q_2z_1, \\ xq_1z_1x^{-1} & = q_2z_1 \in Q \times Z. \end{align*}

Thus any element \(x\) which is in \(N_G(Q)\) is also in \(N_G(Q \times Z)\) so we have \(N_G(Q) \subset N_G(Q \times Z)\).

Let \(q_1 z_1\) be an arbitrarily chosen element of \(Q \times Z\) such that \(q_1 \in Q\) and \(z_1 \in Z\). Now let \(y\) be an arbitrarily chosen element of \(N_G(Q \times Z)\).

\begin{align*} y q_1 z_1 y^{-1} = q_2 z_2 \in Q \times Z. \qquad (\text{where $q_2 \in Q$ and $z_2 \in Z$}) \end{align*}

Consider now the order of \(q_1z_1\) in \(G\). Since \(p \neq 2\), \(Q \cap Z = I_G\) and \(|q_1 z_1| = |q_1| |z_1|\). Note that \(q_1 z_1\) and \(q_2 z_2\) are conjugate in \(G\), and thus their orders are equal. This means that \(|z_1| = |z_2|\), because otherwise 2 would divide one of them and not the other. Thus \(z_1 = z_2\) and,

\begin{align*} y q_1z_1 y^{-1} & = q_2z_2 = q_2z_1 \\ y q_1 y^{-1} z_1 & = q_2z_1, \\ y q_1 y^{-1} & = q_2 \in Q \end{align*}

Hence \(y \in N_G(Q)\). Furthermore, since \(y\) was chosen arbitrarily, any element which is in \(N_G(Q \times Z)\) is also in \(N_G(Q)\), so \(N_G(Q \times Z) = N_G(Q)\) as desired.

We now start to count the elements of the seperate components of \(G\) and use the preceeding 2 theorems to construct what will be an invaluable formula in determining the structure of \(G\), something we will call the Maximal Abelian Subgroup Class Equation of \(G\).

First we spilt \(\mathfrak {M}\) into the conjugacy classes of it’s elements. Theorem 6.6(iii) tells us that every maximal abelian subgroup is either a cyclic subgroup whose order is relatively prime to \(p\) or of the form \(Q \times Z\) where \(Q\) is a Sylow \(p\)-subgroup. Let \(\mathcal{C}_1, \mathcal{C}_2,...,\mathcal{C}_s, \mathcal{C}_{s+1},..., \mathcal{C}_{s+t}\) (where \(s, t \in \mathbb {Z}^+\)) denote the conjugacy classes of the cyclic subgroups whose order is relatively prime to \(p\). Recall that part (iv) of Theorem 6.6 tells us that \([N_G(A): A] = 1\) or 2. Let \(A_i\) be a representative from each \(\mathcal{C}_i\) such that,

\begin{align*} [N_G(A_i) : A_i] & = 1, \tag {for $i \leq s$} \\[2mm] [N_G(A_i) : A_i] & = 2. \tag {for $s {\lt} i \leq s+t$}, \end{align*}

Now let \(Q_1\) and \(Q_2\) be any two Sylow \(p\)-subgroups of \(G\). By the Second Sylow Theorem, \(Q_1\) and \(Q_2\) are conjugate to each other in \(G\). That is, there exists a \(g \in G\) such that \(gQ_1g^{-1} = Q_2\).

\begin{align*} gQ_1g^{-1} = Q_2 & \iff gQ_1g^{-1}Z = Q_2Z \\ & \iff gQ_1Zg^{-1} = Q_2Z \\ & \iff g(Q_1 \times Z)g^{-1} = (Q_2 \times Z). \tag {by Corollary \ref{directproductZ}} \end{align*}

So \(Q_1 \times Z\) and \(Q_2 \times Z\) belong to the same conjugacy class, furthermore there is thus only 1 conjugacy class of elements of this form in \(\mathfrak {M}\). Let \(\mathcal{C}_{Q \times Z}\) denote this conjugacy class and let \(Q \times Z\) be a representative from it. The following diagram provides a visual representation of \(G\) divided into it’s maximal abelian subgroups.

We can reformulate the counting formula in Theorem 6.9(iv) using the notation we have introduced to show that it agrees with the intuitive approach that Fig 1 suggests.

\begin{align*} |G \! \setminus \! Z| = \sum _{A_i^* \in S} |A_i^*| [G:N_G(A_i)] = \sum _{A_i^* \in S} |C_i^*| = |C_{Q \times Z}^*| + \sum _{i=1}^{s+t} |C_i^*|. \end{align*}

We are now able to begin to evaluate \(G\). Firstly, let \(|Z| = e\) and \(|G| = eg\). We know well by now that \(e = 1\) or 2 depending on whether \(p\) equals 2 or not, and by Lagrange’s Theorem, the order of a subgroup divides the order of the group, so \(e\) divides \(|G|\) since \(Z {\lt} G\).

We consider the cyclic case first. Again, by Lagrange’s Theorem, since \(Z\) is a subgroup of each \(A_i\), \(e\) divides \(|A_i|\). So set \(|A_i| = eg_i\). Since \(Z \notin \mathfrak {M}\), each \(A_i\) is therefore strictly larger than \(Z\) and so each \(g_i\) is an integer greater than or equal to 2.

To determine the order of each \(C_i\), we return to the set \(\mathfrak {M}^*\). The size of one representative of each class is,

\begin{align*} |A_i^*| = |A_i \! \setminus \! Z| = eg_i-e = e(g_i-1). \end{align*}

The number of \(A_i^*\) in each conjugacy class \(\mathcal{C}_i\) for \(i \leq s\) is thus,

\begin{align*} |\mathcal{C}_i^*| = |\mathcal{C}_i| = [G:N_G(A_i)] = \frac{|G|}{|A_i|} = \frac{eg}{eg_i} = \frac{g}{g_i}. \end{align*}


Therefore the total number of elements of \(G\) in the noncentral part of \(C_i\) for \(i \leq s\) is,

\begin{align} \label{classeq1of3} \sum _{i=1}^{s} |C_i^*| = \sum _{i=1}^{s} |A_i^*| |\mathcal{C}_i^*| = \sum _{i=1}^{s} \frac{eg(g_i-1)}{g_i}. \end{align}


The number of \(A_i^*\) in each conjugacy class \(\mathcal{C}_i\) for \(s {\lt} i \leq s+t\) is thus,

\begin{align*} |\mathcal{C}_i^*| = |\mathcal{C}_i| = [G:N_G(A_i)] = \frac{|G|}{2|A_i|} = \frac{eg}{2eg_i} = \frac{g}{2g_i}. \end{align*}


Therefore the total number of elements of \(G\) in the noncentral part of \(C_i\) for \(s {\lt} i \leq s+t\) is,

\begin{align} \label{classeq2of3} \sum _{i=s+1}^{s+t} |C_i^*| = \sum _{i=s+1}^{s+t} |A_i^*| |\mathcal{C}_i^*| = \sum _{i=s+1}^{s+t} \frac{eg(g_i-1)}{2g_i}. \end{align}

We next determine the order of \(C_{Q \times Z}\). Let \(|Q| = q\). If \(p \nmid |G|\) then \(q=1\) and if \(p = 0\), then we consider a Sylow \(p\)-subgroup to simply be \(I_G\). So \(q\) is always at least 1. Since \(Z {\lt} K\), we can let \(|K| = ek\). Observe that if \(K \in \mathfrak {M}\), then by Theorem 6.6(v), \(K = A_i\) for some \(0 {\lt} i \leq t\) and \(k = g_i\). Recall that \(N_G(Q) = QK\) and so,

\begin{align*} |N_G(Q \times Z)^*| & = |N_G(Q \times Z)| \tag {by Lemma \ref{unsureifneeded}} \\ & = |N_G(Q)| \tag {by Lemma \ref{unsure}} \\ & = |QK| = eqk. \end{align*}

Again we count the size and number of these maximal abelian groups.

\begin{align*} |(Q \times Z)^*| = |QZ| - |Z| = e(q-1). \end{align*}

Since there is only one conjugacy class of \(Q \times Z\), the number of \((Q \times Z)^*\) in \(\mathfrak {M}^*\) is thus,

\begin{align*} |\mathcal{C}_{Q \times Z}^*| = |\mathcal{C}_{Q \times Z}| = [G: N_G(Q \times Z)] = \frac{|G|}{|N_G(Q \times Z)^*|} = \frac{eg}{eqk} = \frac{g}{qk}. \end{align*}

Therefore the total number of elements of \(G\) in the noncentral parts of each \(Q \times Z\) is,

\begin{align} \label{classeq3of3} |C_{Q \times Z}^*| = |(Q \times Z)^*| |\mathcal{C}_{Q \times Z}^*| = \frac{eg(q-1)}{qk}. \end{align}

We now sum together (13), (14) and (15) to create the Maximal Abelian Subgroup Class Equation of \(G\).

\begin{align} \label{classeq} |G \! \setminus \! Z| & = |C_{Q \times Z}^*| + \sum _{i=1}^{s+t} |C_i^*|, \nonumber \\ |G \! \setminus \! Z| & = |(Q \times Z)^*| |\mathcal{C}_{Q \times Z}^*| + \sum _{i=1}^{s} |A_i^*| |\mathcal{C}_i^*| + \sum _{i=s+1}^{s+t} |A_i^*| |\mathcal{C}_i^*|, \nonumber \\ eg - e & = \frac{eg(q-1)}{qk} + \sum _{i=1}^{s} \frac{eg(g_i-1)}{g_i} + \sum _{i=s+1}^{s+t} \frac{eg(g_i-1)}{2g_i}, \nonumber \\ 1 & = \frac{1}{g} + \frac{q-1}{qk} + \sum _{i=1}^{s} \frac{g_i-1}{g_i} + \sum _{i=s+1}^{s+t} \frac{g_i-1}{2g_i}. \end{align}

Since \(g,k,q \in \mathbb {Z}^+\) this implies that,

\begin{align*} \frac{1}{g} {\gt} 0 \quad \text{and} \quad \frac{q-1}{qk} \geq 0. \end{align*}

Also, since \(g_i \geq 2\) for \(1 \leq i \leq s + t\), we have,

\begin{align*} \frac{g_i-1}{g_i} \geq \frac{1}{2}, \quad \sum _{i=1}^{s} \frac{g_i-1}{g_i} \geq \frac{s}{2} \quad \text{and} \quad \sum _{i=s+1}^{s+t} \frac{g_i-1}{2g_i} \geq \frac{t}{4}. \end{align*}

Thus we can find a lower bound for (16) which limits the possible number of conjugacy classes somewhat,

\begin{align*} 1 {\gt} \frac{s}{2} + \frac{t}{4}. \end{align*}

There are only 6 possible different pairs of values which \(s\) and \(t\) can take:

Case

I

II

III

IV

V

VI

\(s\)

1

1

0

0

0

0

\(t\)

0

1

0

1

2

3

Each case will be examined individually in the next chapter.