1 Introduction

Many algorithms in computational group theory depend on the existence of small bases. Here, a base of a permutation group G acting on a set \(\Omega \) is a subset \(\Delta \subseteq \Omega \) such that the pointwise stabilizer \(G_\Delta \) is trivial (i.e. if \(g\in G\) fixes every \(\delta \in \Delta \), then \(g=1\)). The aim of this short note is to introduce a generalized base without the presence of a group action. To this end, let us first consider a finite group G acting faithfully by automorphisms on a p-group P. If p does not divide |G|, then G always admits a base of size 2 by a theorem of Halasi–Podoski [5]. Now suppose that G is p-solvable, P is elementary abelian, and G acts completely reducibly on P. Then G has a base of size 3 (2 if \(p\ge 5\)) by Halasi–Maróti [4]. In those situations, we may form the semidirect product \(H:=P\rtimes G\). Now there exists \(\Delta \subseteq P\) such that \(|\Delta |\le 3\) and \(\mathrm {C}_H(\Delta )=\mathrm {C}_H(\langle \Delta \rangle )\le P\). This motivates the following definition.

Definition 1

Let G be a finite group with Sylow p-subgroup P. A subset \(\Delta \subseteq P\) is called a p-base of G if \(\mathrm {C}_G(\Delta )\) is p-nilpotent, i.e. \(\mathrm {C}_G(\Delta )\) has a normal p-complement.

Clearly, any generating set of P is a p-base of G since \(\mathrm {C}_G(P)=\mathrm {Z}(P)\times \mathrm {O}_{p'}(\mathrm {C}_G(P))\) (this observation is generalized in Lemma 7 below).

Our main theorem extends the work of Halasi–Maróti as follows.

Theorem 2

Every p-solvable group has a p-base of size 3 (2 if \(p\ge 5\)).

Although Halasi–Maróti’s theorem does not extend to non-p-solvable groups, the situation for p-bases seems more fortunate. For instance, if V is a finite vector space in characteristic p, then every base of \({\text {GL}}(V)\) (under the natural action) contains a basis of V, so its size is at least \(\dim V\). On the other hand, \(G={\text {AGL}}(V)=V\rtimes {\text {GL}}(V)\) possesses a p-base of size 2. To see this, let P be the Sylow p-subgroup of \({\text {GL}}(V)\) consisting of the upper unitriangular matrices. Let \(x\in P\) be a Jordan block of size \(\dim V\). Then \(\mathrm {C}_{{\text {GL}}(V)}(x)\le P\mathrm {Z}({\text {GL}}(V))\). For any \(y\in \mathrm {C}_V(x)\setminus \{1\}\), we obtain a p-base \(\Delta :=\{x,y\}\) such that \(\mathrm {C}_G(\Delta )\le VP\). We have even found a p-base consisting of commuting elements. After checking many more cases, we believe that the following might hold.

Conjecture 3

Every finite group has a (commutative) p-base of size 2 for every prime p.

The role of the number 2 in Conjecture 3 appears somewhat arbitrary at first. There is, however, an elementary dual theorem: A finite group is p-nilpotent if and only if every 2-generated subgroup is p-nilpotent. This can be deduced from the structure of minimal non-p-nilpotent groups (see [6, Satz IV.5.4]). It is a much deeper theorem of Thompson [8] that the same result holds when “p-nilpotent” is replaced by “solvable”. Similarly, 2-generated subgroups play a role in the Baer–Suzuki theorem and its variations.

Apart from Theorem 2 we give some more evidence of Conjecture 3.

Theorem 4

Let G be a finite group with Sylow p-subgroup P. Then Conjecture 3 holds for G in the following cases:

  1. (i)

    P is abelian.

  2. (ii)

    G is a symmetric group or an alternating group.

  3. (iii)

    G is a general linear group, a special linear group, or a projective special linear group.

  4. (iv)

    G is a sporadic simple group or an automorphism group thereof.

Our results on (almost) simple groups carry over to the corresponding quasisimple groups by Lemma 8 below. The notion of p-bases generalizes to blocks of finite groups and even to fusion systems.

Definition 5

  • Let B be a p-block of a finite group G with defect group D. A subset \(\Delta \subseteq D\) is called base of B if B has a nilpotent Brauer correspondent in \(\mathrm {C}_G(\Delta )\) (see [1, Definition IV.5.38]).

  • Let \({\mathcal {F}}\) be a saturated fusion system on a finite p-group P. A subset \(\Delta \subseteq P\) is called base of \({\mathcal {F}}\) if there exists a morphism \(\varphi \) in \({\mathcal {F}}\) such that \(\varphi (\langle \Delta \rangle )\) is fully \({\mathcal {F}}\)-centralized and the centralizer fusion system \({\mathcal {C}}:=\mathrm {C}_{{\mathcal {F}}}(\varphi (\langle \Delta \rangle ))\) is trivial, i.e. \({\mathcal {C}}={\mathcal {F}}_{\mathrm {C}_P(\Delta )}(\mathrm {C}_P(\Delta ))\) (see [1, Definition I.5.3, Theorem I.5.5]).

By Brauer’s third main theorem, the bases of the principal p-block of G are the p-bases of G (see [1, Theorem IV.5.9]). Moreover, if \({\mathcal {F}}\) is the fusion system attached to an arbitrary block B, then the bases of B are the bases of \({\mathcal {F}}\) (see [1, Theorem IV.3.19]). By the existence of exotic fusion systems, the following conjecture strengthens Conjecture 3.

Conjecture 6

Every saturated fusion system has a base of size 2.

We show that Conjecture 6 holds for p-groups of order at most \(p^4\).

2 Results

Proof of Theorem 2

Let G be a p-solvable group with Sylow p-subgroup P. Let \(N:=\mathrm {O}_{p'}(G)\). For \(Q\subseteq P\), \(\mathrm {C}_G(Q)N/N\) is contained in \(\mathrm {C}_{G/N}(QN/Q)\). Hence, \(\mathrm {C}_G(Q)\) is p-nilpotent whenever \(\mathrm {C}_{G/N}(QN/Q)\) is p-nilpotent. Thus, we may assume that \(N=1\). Instead we consider \(N:=\mathrm {O}_p(G)\). Since G is p-solvable, \(N\ne 1\). We show by induction on |N| that there exists a p-base \(\Delta \subseteq N\) such that \(\mathrm {C}_G(\Delta )\le N\). By the Hall–Higman lemma (see [6, Hilfssatz VI.6.5]), \(\mathrm {C}_{G/N}(N/\Phi (N))=N/\Phi (N)\) where \(\Phi (N)\) denotes the Frattini subgroup of N. It follows that \(\mathrm {O}_{p'}(G/\Phi (N))=1\). Hence, by induction, we may assume that N is elementary abelian. Then \({\overline{G}}:=G/N\) acts faithfully on N and it suffices to find a p-base \(\Delta \subseteq N\) such that \(\mathrm {C}_{{\overline{G}}}(\Delta )=1\). Thus, we may assume that \(G=N\rtimes H\) where \(\mathrm {C}_G(N)=N\) and \(\mathrm {O}_p(H)=1\).

Note that \(\Phi (G)\le \mathrm {F}(G)=N\) where \(\mathrm {F}(G)\) is the Fitting subgroup of G. Since H is contained in a maximal subgroup of G, we even have \(\Phi (G)<N\). Let \(K\unlhd H\) be the kernel of the action of H on \(N/\Phi (G)\). By way of contradiction, suppose that \(K\ne 1\). Since K is p-solvable and \(\mathrm {O}_p(K)\le \mathrm {O}_p(H)=1\), also \(K_0:=\mathrm {O}_{p'}(K)\ne 1\). Now \(K_0\) acts coprimely on N and we obtain

$$\begin{aligned}N=[K_0,N]\mathrm {C}_N(K_0)=\Phi (G)\mathrm {C}_N(K_0)\end{aligned}$$

as is well-known. Both \(\Phi (G)\) and \(\mathrm {C}_N(K_0)H\) lie in a maximal subgroup M of G. But then \(G=NH=\Phi (G)\mathrm {C}_N(K_0)H\le M\), a contradiction. Therefore, H acts faithfully on \(N/\Phi (G)\) and we may assume that \(\Phi (G)=1\). Then there exist maximal subgroups \(M_1,\ldots ,M_n\) of G such that \(N_i:=M_i\cap N<N\) for \(i=1,\ldots ,n\) and \(\bigcap _{i=1}^nN_i=1\). Since \(G=M_iN\), the quotients \(N/N_i\) are simple \(\mathbb {F}_pH\)-modules and N embeds into \(N/N_1\times \cdots \times N/N_n\). Hence, the action of H on N is faithful and completely reducible. Now, by the main result of Halasi–Maróti [4], there exists a p-base with the desired properties. \(\square \)

Next we work towards Theorem 4.

Lemma 7

Let P be a Sylow p-subgroup of G. Let \(Q\unlhd P\) such that \(\mathrm {C}_P(Q)\le Q\). Then every generating set of Q is a p-base of G.

Proof

Since \(P\in {\text {Syl}}_p(\mathrm {N}_G(Q))\), we have \(\mathrm {Z}(Q)=\mathrm {C}_P(Q)\in {\text {Syl}}_p(\mathrm {C}_G(Q))\) and therefore \(\mathrm {C}_G(Q)=\mathrm {Z}(Q)\times \mathrm {O}_{p'}(\mathrm {C}_G(Q))\) by the Schur–Zassenhaus theorem.\(\square \)

Lemma 8

Let \(\Delta \) be a p-base of G and let \(N\le \mathrm {Z}(G)\). Then \({\overline{\Delta }}:=\{xN:x\in \Delta \}\) is a p-base of G/N.

Proof

Let \(gN\in \mathrm {C}_{G/N}({\overline{\Delta }})\). Then g normalizes the nilpotent group \(\langle \Delta \rangle N\). Hence, g acts on the unique Sylow p-subgroup P of \(\langle \Delta \rangle N\). Since g centralizes

$$\begin{aligned}\langle {\overline{\Delta }}\rangle =\langle \Delta \rangle N/N=PN/N\cong P/P\cap N\end{aligned}$$

and \(P\cap N\le N\le \mathrm {Z}(G)\), g induces a p-element in \({\text {Aut}}(P)\) and also in \({\text {Aut}}(\langle \Delta \rangle N)\). Consequently, there exists a p-subgroup \(Q\le \mathrm {N}_G(\langle \Delta \rangle N)\) such that \(\mathrm {C}_{G/N}({\overline{\Delta }})=Q\mathrm {C}_G(\Delta N)/N=Q\mathrm {C}_G(\Delta )/N\). Since \(\mathrm {C}_G(\Delta )\) is p-nilpotent, so is \(Q\mathrm {C}_G(\Delta )\) and the claim follows. \(\square \)

The following implies the first part of Theorem 4.

Proposition 9

Let P be a Sylow p-subgroup of G with nilpotency class c. Then G has a p-base of size 2c.

Proof

The \(p'\)-group \(\mathrm {N}_G(\mathrm {Z}(P))/\mathrm {C}_G(\mathrm {Z}(P))\) acts faithfully on \(\mathrm {Z}(P)\). By Halasi–Podoski [5], there exists \(\Delta _0=\{x,y\}\subseteq \mathrm {Z}(P)\) such that \(\mathrm {N}_H(\mathrm {Z}(P))\le \mathrm {C}_H(\mathrm {Z}(P))\) where \(H:=\mathrm {C}_G(\Delta _0)\). If \(c=1\), then \(P=\mathrm {Z}(P)\) is abelian and Burnside’s transfer theorem implies that H is p-nilpotent. Hence, let \(c>1\). By a well-known fusion argument of Burnside, elements of \(\mathrm {Z}(P)\) are conjugate in H if and only if they are conjugate in \(\mathrm {N}_H(\mathrm {Z}(P))\). Consequently, all elements of \(\mathrm {Z}(P)\) are isolated in our situation. By the \(\mathrm {Z}^*\)-theorem (assuming the classification of finite simple groups), we obtain

$$\begin{aligned}\mathrm {Z}(H/\mathrm {O}_{p'}(H))=\mathrm {Z}(P)\mathrm {O}_{p'}(H)/\mathrm {O}_{p'}(H).\end{aligned}$$

The group \({\overline{H}}:=H/\mathrm {Z}(P)\mathrm {O}_{p'}(H)\) has Sylow p-subgroup \({\overline{P}}\cong P/\mathrm {Z}(P)\) of nilpotency class \(c-1\). By induction on c, there exists a p-base \(\overline{\Delta _1}\subseteq {\overline{P}}\) of \({\overline{H}}\) of size \(2(c-1)\). We may choose \(\Delta _1\subseteq P\) such that \(\overline{\Delta _1}=\{{\overline{x}}:x\in \Delta _1\}\). Since \(\overline{\mathrm {C}_H(\Delta _1)}\le \mathrm {C}_{{\overline{H}}}(\overline{\Delta _1})\) is p-nilpotent, so is

$$\begin{aligned}\bigl (\mathrm {C}_H(\Delta _1)\mathrm {Z}(P)\mathrm {O}_{p'}(H)/\mathrm {O}_{p'}(H)\bigr )/\mathrm {Z}(H/\mathrm {O}_{p'}(H)).\end{aligned}$$

It follows that \(\mathrm {C}_H(\Delta _1)\mathrm {Z}(P)\mathrm {O}_{p'}(H)/\mathrm {O}_{p'}(H)\) and \(\mathrm {C}_H(\Delta _1)=\mathrm {C}_G(\Delta _0\cup \Delta _1)\) are p-nilpotent as well. Hence, \(\Delta :=\Delta _0\cup \Delta _1\) is a p-base of G of size (at most) 2c.

\(\square \)

Proposition 10

The symmetric and alternating groups \(S_n\) and \(A_n\) have commutative p-bases of size 2 for every prime p.

Proof

Let \(n=\sum _{i=0}^k a_ip^i\) be the p-adic expansion of n. Suppose first that \(G=S_n\). Let

$$\begin{aligned}x=\prod _{i=0}^k\prod _{j=1}^{a_i}x_{ij}\in G\end{aligned}$$

be a product of disjoint cycles \(x_{ij}\) where \(x_{ij}\) has length \(p^i\) for \(j=1,\ldots ,a_i\). Then x is a p-element and

$$\begin{aligned}\mathrm {C}_G(x)\cong \prod _{i=0}^kC_{p^i}\wr S_{a_i}.\end{aligned}$$

Since \(a_i<p\), \(P:=\langle x_{ij}:i=0,\ldots ,k,j=1,\ldots ,a_i\rangle \) is an abelian Sylow p-subgroup of \(\mathrm {C}_G(x)\). Let \(y:=\prod _{i=0}^k\prod _{j=1}^{a_i}x_{ij}^{j}\in P\). It is easy to see that \(\Delta :=\{x,y\}\) is a commutative p-base of G with \(\mathrm {C}_G(\Delta )=P\).

Now let \(G=A_n\). If \(p>2\), then xy lie in \(A_n\) as constructed above and the claim follows from \(\mathrm {C}_{A_n}(\Delta )\le \mathrm {C}_{S_n}(\Delta )\). Hence, let \(p=2\). If \(\sum _{i=1}^ka_i\equiv 0\pmod {2}\), then we still have \(x\in A_n\) and \(\mathrm {C}_G(x)=\langle x_{ij}:i,j\rangle \) is already a 2-group. Thus, we have a 2-base of size 1 in this case. In the remaining case, let \(m\ge 1\) be minimal such that \(a_m=1\). We adjust our definition of x by replacing \(x_{m1}\) with a disjoint product of two cycles of length \(2^{m-1}\). Then \(x\in A_n\) and \(\mathrm {C}_G(x)\) is a 2-group or a direct product of a 2-group and \(S_3\) (the latter case happens if and only if \(m=1=a_0\)). We clearly find a 2-element \(y\in \mathrm {C}_G(x)\) such that \(\mathrm {C}_G(x,y)\) is a 2-group. \(\square \)

The following elementary facts are well-known, but we provide proofs for the convenience of the reader.

Lemma 11

Let p be a prime and let q be a prime power such that \(p\not \mid q\). Let \(e\mid p-1\) be the multiplicative order of q modulo p. Let \(p^s\) be the p-part of \(q^e-1\). Then for every \(n\ge 1\), the polynomial \(X^{p^n}-1\) decomposes as

$$\begin{aligned}X^{p^n}-1=(X-1)\prod _{k=1}^{(p^s-1)/e}\gamma _{0,k}\prod _{i=1}^{n-s}\prod _{k=1}^{\varphi (p^s)/e}\gamma _{i,k}\end{aligned}$$

where the \(\gamma _{i,k}\) are pairwise coprime polynomials in \(\mathbb {F}_q[X]\) of degree \(ep^i\) for \(i=0,\ldots ,n-s\).

Proof

Let \(\zeta \) be a primitive root of \(X^{p^n}-1\) in some finite field extension of \(\mathbb {F}_q\). Then

$$\begin{aligned}X^{p^n}-1=\prod _{k=0}^{p^n-1}(X-\zeta ^k).\end{aligned}$$

Recall that \(\mathbb {F}_q\) is the fixed field under the Frobenius automorphism \(c\mapsto c^q\). Hence, the irreducible divisors of \(X^{p^n}-1\) in \(\mathbb {F}_q[X]\) correspond to the orbits of \(\langle q+p^n\mathbb {Z}\rangle \) on \(\mathbb {Z}/p^n\mathbb {Z}\) via multiplication. The trivial orbit corresponds to \(X-1\). For \(i=1,\ldots ,s\), the order of q modulo \(p^i\) is e by the definition of s. This yields \((p^s-1)/e\) non-trivial orbits of length e in \(p^{n-s}\mathbb {Z}/p^n\mathbb {Z}\). The corresponding irreducible factors are denoted by \(\gamma _{0,k}\) for \(k=1,\ldots ,(p^s-1)/e\).

For \(i\ge 1\), the order of q modulo \(p^{s+i}\) divides \(ep^i\) (it can be smaller if \(p=2\) and \(s=1\)). We partition \((p^{n-s-i}\mathbb {Z}/p^n\mathbb {Z})^\times \) into \(\varphi (p^{s+i})/(ep^i)=\varphi (p^s)/e\) unions of orbits under \(\langle q+p^n\mathbb {Z}\rangle \) such that each union has size \(ep^i\). The corresponding polynomials \(\gamma _{i,1},\ldots ,\gamma _{i,\varphi (p^s)/e}\) are pairwise coprime (but not necessarily irreducible). \(\square \)

Lemma 12

Let A be an \(n\times n\)-matrix over an arbitrary field F such that the minimal polynomial of A has degree n. Then every matrix commuting with A is a polynomial in A.

Proof

By hypothesis, A is similar to a companion matrix. Hence, there exists a vector \(v\in F^n\) such that \(\{v,Av,\ldots ,A^{n-1}v\}\) is a basis of \(F^n\). Let \(B\in F^{n\times n}\) such that \(AB=BA\). There exist \(a_0,\ldots ,a_{n-1}\in F\) such that \(Bv=a_0v+\cdots +a_{n-1}A^{n-1}v\). Set \(\gamma :=a_0+a_1X+\cdots +a_{n-1}X^{n-1}\). Then

$$\begin{aligned}BA^iv=A^iBv=a_0A^iv+\cdots +a_{n-1}A^{n-1}A^iv=\gamma (A)A^iv\end{aligned}$$

for \(i=0,\ldots ,n-1\). Since \(\{v,Av,\ldots ,A^{n-1}v\}\) is a basis, we obtain \(B=\gamma (A)\) as desired. \(\square \)

Proposition 13

The groups \({\text {GL}}(n,q)\), \({\text {SL}}(n,q)\), and \({\text {PSL}}(n,q)\) possess commutative p-bases of size 2 for every prime p.

Proof

Let q be a prime power. By Lemma 8, it suffices to consider \({\text {GL}}(n,q)\) and \({\text {SL}}(n,q)\). Suppose first that \(p\mid q\). Let \(x\in G:={\text {GL}}(n,q)\) be a Jordan block of size \(n\times n\) with eigenvalue 1. Then x is a p-element since \(x^{p^n}-1=(x-1)^{p^n}=0\). Moreover, \(\mathrm {C}_G(x)\) consists of polynomials in x by Lemma 12. In particular, \(\mathrm {C}_G(x)\) is abelian and therefore p-nilpotent. Hence, we found a p-base of size 1. Since \((q-1,p)=1\), this is also a p-base of \({\text {SL}}(n,q)\).

Now let \(p\not \mid q\). We “linearize” the argument from Proposition 10. Let e and s be as in Lemma 11. Let \(0\le a_0\le e-1\) be such that \(n\equiv a_0\pmod {e}\). Let

$$\begin{aligned}\frac{n-a_0}{e}=\sum _{i=0}^{r}a_{i+1}p^i\end{aligned}$$

be the p-adic expansion. Let \(M_i\in {\text {GL}}(ep^i,q)\) be the companion matrix of the polynomial \(\gamma _{i,1}\) from Lemma 11 for \(i=0,\ldots ,r\). Let \(G_i:={\text {GL}}(ea_{i+1}p^i,q)\) and \(x_i:={\text {diag}}(M_i,\ldots ,M_i)\in G_i\). Then the minimal polynomial of

$$\begin{aligned}x:={\text {diag}}(1_{a_0},x_0,\ldots ,x_r)\in G\end{aligned}$$

divides \(X^{p^{r+s}}-1\) by Lemma 11. In particular, x is a p-element. Since the \(\gamma _{i,1}\) are pairwise coprime, it follows that

$$\begin{aligned}\mathrm {C}_G(x)={\text {GL}}(a_0,q)\times \prod _{i=0}^r\mathrm {C}_{G_i}(x_i).\end{aligned}$$

Since \(a_0<e\), \({\text {GL}}(a_0,q)\) is a \(p'\)-group. By Lemma 12, every matrix commuting with \(M_i\) is a polynomial in \(M_i\). Hence, the elements of \(\mathrm {C}_{G_i}(x_i)\) have the form \(A=(A_{kl})_{1\le k,l\le a_{i+1}}\) where each block \(A_{kl}\) is a polynomial in \(M_i\). We define

$$\begin{aligned}y_i:={\text {diag}}(M_i,M_i^2,\ldots , M_i^{a_{i+1}})\in \mathrm {C}_{G_i}(x_i)\end{aligned}$$

and \(y:={\text {diag}}(1_{a_0},y_0,\ldots ,y_r)\in \mathrm {C}_G(x)\). Let \(A=(A_{kl})\in \mathrm {C}_{G_i}(x_i,y_i)\). We want to show that \(A_{kl}=0\) for \(k\ne l\). To this end, we may assume that \(k<l\) and \(A_{kl}=\rho (M_i)\) where \(\rho \in \mathbb {F}_q[X]\) with \(\deg (\rho )<\deg (\gamma _{i,1})=ep^i\). Since \(A\in \mathrm {C}_{G_i}(x_i,y_i)\), we have \(M_i^kA_{kl}=M_i^lA_{kl}\) and \((M^{l-k}-1)A_{kl}=0\). It follows that the minimal polynomial \(\gamma _{i,1}\) of \(M_i\) divides \((X^{l-k}-1)\rho \). By way of contradiction, we assume that \(\rho \ne 0\). Then \(\gamma _{i,1}\) divides \(X^{l-k}-1\) and \(X^{p^{r+s}}-1\). However, \(l-k\le a_{i+1}<p\) and \(\gamma _{i1}\) must divide \(X-1\). This contradicts the definition of \(\gamma _{i,1}\) in Lemma 11. Hence, \(A_{kl}=0\) for \(k\ne l\). We have shown that the elements of \(\mathrm {C}_G(x,y)\) have the form

$$\begin{aligned}L\oplus \bigoplus _{i=0}^r\bigoplus _{j=1}^{a_{i+1}}L_{ij}\end{aligned}$$

where \(L\in {\text {GL}}(a_0,q)\) and each \(L_{ij}\) is a polynomial in \(M_i\). In particular, \(\mathrm {C}_G(x,y)\) is a direct product of a \(p'\)-group and an abelian group. Consequently, \(\mathrm {C}_G(x,y)\) is p-nilpotent.

Now let \(G:={\text {SL}}(n,q)\). If \(p\not \mid q-1\), then the p-base of \({\text {GL}}(n,q)\) constructed above already lies in G. Thus, we may assume that \(p\mid q-1\). Then \(e=1\) and \(a_0=0\) with the notation above. We now have the polynomials \(\gamma _{i,k}\) with \(i=0,\ldots ,r\) and \(k=1,\ldots ,p-1\le \varphi (p^s)\) at our disposal. Let \(M_{i,k}\) be the companion matrix of \(\gamma _{i,k}\). Define

$$\begin{aligned}x_i:={\text {diag}}(M_{i,1},\ldots ,M_{i,a_{i+1}})\end{aligned}$$

for \(i=0,\ldots ,r\). Then the minimal polynomial of \(x:={\text {diag}}(x_0,\ldots ,x_r)\in {\text {GL}}(n,q)\) has degree n and therefore \(\mathrm {C}_{{\text {GL}}(n,q)}(x)\) is abelian by Lemma 12. Let \(i\ge 0\) be minimal such that \(a_{i+1}>0\). We replace the block \(M_{i,1}\) of x by the companion matrix of \(X^{p^i}-1\). Then, by Lemma 11, the minimal polynomial of x still has degree n. Moreover, x has at least one block B of size \(1\times 1\). We may modify B such that \(\det (x)=1\). After doing so, it may happen that B occurs twice in x. In this case, \(\mathrm {C}_G(x)\le {\text {GL}}(2,q)\times H\) where H is abelian. Then the matrix

$$\begin{aligned}y:={\left\{ \begin{array}{ll} \begin{pmatrix}0&{}-1\\ 1&{}0\end{pmatrix}\oplus 1_{n-2}&{}\text {if }p=2,\\ {\text {diag}}(M_{0,1},M_{0,1}^{-1},1_{n-2})&{}\text {if }p>2 \end{array}\right. } \end{aligned}$$

lies in \(\mathrm {C}_G(x)\) and \(\mathrm {C}_G(x,y)\) is abelian. Hence, \(\{x,y\}\) is a p-base of G. \(\square \)

Proposition 13 can probably be generalized to classical groups. The next result completes the proof of Theorem 4.

Proposition 14

Let S be a sporadic simple group and \(G\in \{S,S.2\}\). Then G has a commutative p-base of size 2 for every prime p.

Proof

If \(p^4\) does not divide |G|, then the claim follows from Lemma 7. So we may assume that \(p^4\) divides |G|. From the character tables in the Atlas [2], we often find p-elements \(x\in G\) such that \(\mathrm {C}_G(x)\) is already a p-group. In this case, we found a p-base of size 1 and we are done. If G admits a permutation representation of “moderate” degree (including \(Co_1\)), then the claim can be shown directly in GAP [3]. In the remaining cases, we use the Atlas to find p-elements with small centralizers:

  • \(G=Ly\), \(p=2\): There exists an involution \(x\in G\) such that \(\mathrm {C}_G(x)=2.A_{11}\). By the proof of Proposition 10, there exists \(y\in A_{11}\) such that \(\mathrm {C}_{A_{11}}(y)\) is a 2-group. We identify y with a preimage in \(\mathrm {C}_G(x)\). Then \(\mathrm {C}_G(x,y)\) is a 2-group.

  • \(G=Ly\), \(p=3\): Here we find \(x\in G\) of order 3 such that \(\mathrm {C}_G(x)=3.McL\). Since McL contains a 3-element y such that \(\mathrm {C}_{McL}(y)\) is a 3-group, the claim follows.

  • \(G=Th\), \(p=2\): There exists an involution \(x\in G\) such that \(\mathrm {C}_G(x)=2^{1+8}_+.A_9\). As before, we find \(y\in \mathrm {C}_G(x)\) such that \(\mathrm {C}_G(x,y)\) is a 2-group.

  • \(G=M\), \(p=5\): There exists a 5-element \(x\in G\) such that \(\mathrm {C}_G(x)=C_5\times HN\). Since there is also a 5-element \(y\in HN\) such that \(\mathrm {C}_{HN}(y)\) is a 5-group, the claim follows.

  • \(G=M\), \(p=7\): In this case there exists a radical subgroup \(Q\le G\) such that \(\mathrm {C}_G(Q)=Q\cong C_7\times C_7\) by Wilson [9, Theorem 7] (this group was missing in the list of local subgroups in the Atlas). Any generating set of Q of size 2 is a desired p-base of G.

  • \(G=HN.2\), \(p=3\): There exists an element \(x\in G\) of order 9 such that \(|\mathrm {C}_G(x)|=54\). Clearly, we find \(y\in \mathrm {C}_G(x)\) such that \(\mathrm {C}_G(x,y)\) is 3-nilpotent. \(\square \)

Finally, we consider a special case of Conjecture 6.

Proposition 15

Let \({\mathcal {F}}\) be a saturated fusion system on a p-group P of order at most \(p^4\). Then \({\mathcal {F}}\) has a base of size 2.

Proof

Recall that \(A:={\text {Out}}_{{\mathcal {F}}}(P)\) is a \(p'\)-group and there is a well-defined action of A on P by the Schur–Zassenhaus theorem. If \({\mathcal {F}}\) is the fusion system of the group \(P\rtimes A\), then the claim follows from Halasi–Podoski [5] as before. We may therefore assume that P contains an \({\mathcal {F}}\)-essential subgroup. In particular, P is non-abelian. Let \(Q<P\) be a maximal subgroup of P containing \(\mathrm {Z}(P)\). The fusion system \(\mathrm {C}_{{\mathcal {F}}}(Q)\) on \(\mathrm {C}_P(Q)=\mathrm {Z}(Q)\) is trivial by definition. Hence, we are done whenever Q is generated by two elements.

It remains to deal with the case where \(|P|=p^4\) and all maximal subgroups containing \(\mathrm {Z}(P)\) are elementary abelian of rank 3. Since two such maximal subgroups intersect in \(\mathrm {Z}(P)\), we obtain that \(|\mathrm {Z}(P)|=p^2\) and \(|P'|=p\) by [7, Lemma 1.9] for instance. By the first part of the proof, we may choose an \({\mathcal {F}}\)-essential subgroup Q such that \(\mathrm {Z}(P)<Q<P\). Let \(A:={\text {Aut}}_{{\mathcal {F}}}(Q)\). Since Q is essential, P/Q is a non-normal Sylow p-subgroup of A (see [1, Proposition I.2.5]). Moreover, \([P,Q]=P'\) has order p. By [7, Lemma 1.11], there exists an A-invariant decomposition

$$\begin{aligned}Q=\langle x,y\rangle \times \langle z\rangle .\end{aligned}$$

We may choose those elements such that \(\Delta :=\{xz,y\}\nsubseteq \mathrm {Z}(P)\). Then \(\mathrm {C}_P(\Delta )=Q\) and \(\mathrm {C}_A(\Delta )=1\). Let \(\varphi :S\rightarrow T\) be a morphism in \({\mathcal {C}}:=\mathrm {C}_{{\mathcal {F}}}(\Delta )\) where \(S,T\le Q\). Then \(\varphi \) extends to a morphism \({\hat{\varphi }}:S\langle \Delta \rangle \rightarrow T\langle \Delta \rangle \) in \({\mathcal {F}}\) such that \({\hat{\varphi }}(x)=x\) for all \(x\in \langle \Delta \rangle \). Hence, if \(S\le \langle \Delta \rangle \), then \(\varphi ={\text {id}}\). Otherwise, \(S\langle \Delta \rangle =Q\) and \(\hat{\varphi } \in \mathrm {C}_A(\Delta )=1\) since morphisms are always injective. In any case, \({\mathcal {C}}\) is the trivial fusion system and \(\Delta \) is a base of \({\mathcal {F}}\). \(\square \)