Hostname: page-component-848d4c4894-ndmmz Total loading time: 0 Render date: 2024-05-06T20:41:18.269Z Has data issue: false hasContentIssue false

STRICT REGULARITY FOR $2$-COCYCLES OF FINITE GROUPS

Published online by Cambridge University Press:  24 March 2023

R. J. HIGGS*
Affiliation:
School of Mathematics and Statistics, University College Dublin, Belfield, Dublin 4, Ireland
Rights & Permissions [Opens in a new window]

Abstract

Let $\alpha $ be a complex-valued $2$-cocycle of a finite group $G.$ A new concept of strict $\alpha $-regularity is introduced and its basic properties are investigated. To illustrate the potential use of this concept, a new proof is offered to show that the number of orbits of G under its action on the set of complex-valued irreducible $\alpha _N$-characters of N equals the number of $\alpha $-regular conjugacy classes of G contained in $N,$ where N is a normal subgroup of $G.$

Type
Research Article
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), 2023. Published by Cambridge University Press on behalf of Australian Mathematical Publishing Association Inc.

1 Introduction

Throughout this paper, G will denote a finite group and it will be implicitly assumed that all projective representations affording projective characters are defined over the field of complex numbers $\mathbb {C}.$

Definition 1.1. A $2$ -cocycle of G over $\mathbb {C}$ is a function $\alpha : G\times G\rightarrow \mathbb {C}^*$ such that $\alpha (1, 1) = 1$ and $\alpha (x, y)\alpha (xy, z) = \alpha (x, yz)\alpha (y, z)$ for all x, y, $z\in G.$

The set of all such $2$ -cocycles of G form a group $Z^2(G, \mathbb {C}^*)$ under multiplication. Let $\delta : G\rightarrow \mathbb {C}^*$ be any function with $\delta (1) = 1.$ Then $t(\delta )(x, y) = \delta (x)\delta (y)/\delta (xy)$ for all $x, y\in G$ is a $2$ -cocycle of G, which is called a coboundary. Two $2$ -cocycles $\alpha $ and $\beta $ are cohomologous if there exists a coboundary $t(\delta )$ such that $\beta = t(\delta )\alpha .$ This defines an equivalence relation on $Z^2(G, \mathbb {C}^*)$ and the cohomology classes $[\alpha ]$ form a finite abelian group, called the Schur multiplier $M(G).$

Definition 1.2. Let $\alpha $ be a $2$ -cocycle of G.

  1. (a) Define $f_{\alpha }: G\times G\rightarrow \mathbb {C}^*$ by

    $$ \begin{align*}f_{\alpha}(g, x) = \frac{\alpha(g, x)\alpha(gx, g^{-1})}{\alpha(g, g^{-1})}.\end{align*} $$
  2. (b) For each $x\in G$ , define $\alpha _x: C_G(x)\rightarrow \mathbb {C}^*$ by $\alpha _x(g) = \alpha (g, x)/\alpha (x, g).$

These two functions arise naturally in the twisted group algebra $(\mathbb {C}(G))_{\alpha }$ in which $\bar {x}\bar {y} = \alpha (x, y)\overline {xy}$ for all $x, y\in G$ (see [Reference Karpilovsky4, page 66]). Here, $\bar {g}\bar {x}\bar {g}^{-1} = f_{\alpha }(g, x)\overline {gxg^{-1}}$ for $g, x\in G$ and $\bar {g}\bar {x}\bar {g}^{-1} = \alpha _x(g)\bar {x}$ if $g\in C_G(x).$ Also, if $\beta = t(\delta )\alpha $ , then $f_{\beta }(g, x) = (\delta (x)/\delta (gxg^{-1}))f_{\alpha }(g, x)$ for all $g, x\in G$ and consequently $\alpha _x = \beta _x.$

Now $\alpha _x\in \operatorname {\mathrm {Lin}}(C_G(x))$ from [Reference Read6, Lemma 4.2], where $\operatorname {\mathrm {Lin}}(C_G(x))$ is the group of linear characters of $C_G(x).$ The kernel of $\alpha _x$ is the absolute centraliser $C_{\alpha }(x)$ of x with respect to $\alpha $ and $C_G(x)/C_{\alpha }(x)\cong \langle \alpha _x\rangle .$

Definition 1.3. Let $\alpha $ be a $2$ -cocycle of G. Then $x\in G$ is $\alpha $ -regular if $\alpha _x$ is the trivial character of $C_G(x)$ (or equivalently $C_{\alpha }(x) = C_G(x)).$

First, every element of G is $\alpha $ -regular if $[\alpha ]$ is trivial. Second, setting $y = 1$ and $z = 1$ in Definition 1.1 yields $\alpha (x, 1) = 1$ and similarly $\alpha (1, x) = 1$ for all $x\in G,$ and hence $1$ is always $\alpha $ -regular. Third, if $x\in G$ is $\alpha $ -regular, then it is $\alpha ^k$ -regular for any integer $k.$ Finally, if $x\in G$ is $\alpha $ -regular, then so too is any conjugate of x (see [Reference Karpilovsky4, Lemma 2.6.1]), so that one may refer to the $\alpha $ -regular conjugacy classes of $G.$

Now let $\operatorname {\mathrm {Proj}}(G, \alpha )$ denote the set of all irreducible $\alpha $ -characters of G (see [Reference Karpilovsky4, page 184]). Then $x\in G$ is $\alpha $ -regular if and only if $\xi (x)\not = 0$ for some $\xi \in \operatorname {\mathrm {Proj}}(G, \alpha )$ (see [Reference Karpilovsky5, Proposition 1.6.3]) and $\vert \!\operatorname {\mathrm {Proj}}(G, \alpha )\vert $ is the number of $\alpha $ -regular conjugacy classes of G (see [Reference Karpilovsky5, Theorem 1.3.6]).

Let N be a normal subgroup of $G.$ Then G acts on $\operatorname {\mathrm {Proj}}(N, \alpha _N)$ by

$$ \begin{align*}\zeta^g(x) = f_{\alpha}(g, x)\zeta(gxg^{-1})\end{align*} $$

for $\zeta \in \operatorname {\mathrm {Proj}}(N, \alpha _N), g\in G$ and all $x\in N$ . Clifford’s theorem for projective characters applies to this action (see [Reference Karpilovsky5, Theorem 2.2.1]).

A new concept of strict $\alpha ^d$ -regularity, which refines the notion of $\alpha ^d$ -regularity, will be defined and investigated in Section 2 for d a divisor of the order of $[\alpha ].$ This concept will be used in Section 3 to give an alternative proof that the number of orbits of G under its action on $\operatorname {\mathrm {Proj}}(N, \alpha _N),$ for N a normal subgroup of $G,$ is equal to the number of $\alpha $ -regular conjugacy classes of G contained in N from [Reference Higgs2, Lemma 3.1]. It is also easy to show that this result is independent of the choice of $2$ -cocycle from $[\alpha ].$ The result is well known when $\alpha $ is trivial (see [Reference Isaacs3, Corollary 6.33]); the method employed will be to apply this to the orbits of an $\alpha $ -covering group of G under its action on the irreducible characters and conjugacy classes of a normal subgroup, but to decompose these orbits into corresponding sets.

2 Strictly $\alpha ^d$ -regular elements

Let $o(\phantom {.})$ denote the order of an element in a group. Then for $[\beta ]\in M(G)$ , there exists $\alpha \in [\beta ]$ such that $o(\alpha ) = o([\beta ])$ and $\alpha $ is a class-function cocycle, that is, the elements of $\operatorname {\mathrm {Proj}}(G, \alpha )$ are class functions (see [Reference Karpilovsky5, Corollary 4.1.6]). To avoid repetition throughout the rest of this paper, it will be assumed that $\alpha $ has these two properties with $n = o(\alpha ).$ A consequence of the second property is that $x\in G$ is $\alpha $ -regular if and only if $f_{\alpha }(g, x) = 1$ for all $g\in G$ (see [Reference Karpilovsky5, page 33]). The first property allows us to make the following definition in terms of $\alpha ^d$ rather than for the more clumsy $\beta \in [\alpha ]^d.$

Definition 2.1. Define $x\in G$ to be strictly $\alpha ^d$ -regular if d is the smallest integer with $1\leq d\leq n$ such that x is $\alpha ^d$ -regular.

Next suppose $o(\alpha ^d)= o(\alpha ^k) = m.$ If $\omega $ is a primitive mth root of unity, then there exists a field automorphism $\tau $ of $\mathbb {Q}(\omega )$ over $\mathbb Q$ such that $\tau (\alpha ^d) =\alpha ^k.$ Consequently, $x\in G$ is $\alpha ^d$ -regular if and only if it is $\alpha ^k$ -regular. Thus, $d\mid n$ in Definition 2.1.

Let $\pi (d)$ denote the set of prime numbers that divide d and let $d_p$ denote the pth part of d for any prime number $p.$

Lemma 2.2. We have $x\in G$ is strictly $\alpha ^d$ -regular if and only if either:

  1. (a) x is $\alpha ^d$ -regular but not $\alpha ^{d/p}$ -regular for each $p\in \pi (d);$ or

  2. (b) $o(\alpha _x) = d$ in $\operatorname {\mathrm {Lin}}(C_G(x)).$

Proof. For condition (a), if x is not $\alpha ^{d/p}$ -regular, then it is not $\alpha ^t$ -regular for all positive integers t with $t\mid d/p.$ For condition (b), observe that x is $\alpha ^d$ -regular if and only if $\alpha _x^d$ is trivial, that is, $o(\alpha _x)\mid d.$ Now for $d> 1$ , x is strictly $\alpha ^d$ -regular if and only if $o(\alpha _x)\mid d$ , but $\alpha _x^{d/p}\not =1$ for each prime $p\in \pi (d)$ from condition (a). The latter is true if and only if $d_p\mid o(\alpha _x)$ for each prime $p\in \pi (d),$ that is, if and only if $d\mid o(\alpha _x).$

An equivalent way of stating Lemma 2.2(b) is that $x\in G$ is strictly $\alpha ^d$ -regular if and only if $\vert C_G(x)/C_{\alpha }(x)\vert = d.$

Now by definition for each $x\in G$ , there exists a unique $d\mid n$ such that x is strictly $\alpha ^d$ -regular. Thus, the conjugacy classes of G are partitioned into strictly $\alpha ^d$ -regular conjugacy classes. So for $d\mid n$ and N a normal subgroup of $G,$ let $t_d$ be the number of strictly $\alpha ^d$ -regular conjugacy classes of G contained in $N.$ Thus, the number of $\alpha ^d$ -regular conjugacy classes of G contained in N is $\sum _{s|d} t_s;$ in particular, $\sum _{d\mid n} t_d = t(N),$ where $t(N)$ is the number of conjugacy classes of G contained in $N.$

The choice of $2$ -cocycle $\alpha $ allows the construction of an $\alpha $ -covering group H of G with the following three properties (see [Reference Karpilovsky4, Section 4.1]):

  1. (a) H has a cyclic subgroup $A\leq Z(H)\cap H'$ of order $n;$

  2. (b) there exists a conjugacy-preserving transversal (see below) $\{r(g): g\in G\}$ of A in H such that $\theta : H\rightarrow G$ defined by $\theta (r(g)a) = g$ for all $g\in G$ and all $a\in A$ is a homomorphism with kernel $A;$

  3. (c) there exists a faithful character $\lambda \in \operatorname {\mathrm {Lin}}(A)$ such that $\alpha (x, y) = \lambda (A(x, y))$ for all $x, y\in G,$ where $r(x)r(y) = A(x, y)r(xy).$

A conjugacy-preserving transversal means that $r(x)$ and $r(y)$ are conjugate in H if and only if x and y are conjugate in G (see [Reference Karpilovsky5, Lemma 4.1.1]).

It is easy to see that $\theta (C_H(r(x))) = C_{\alpha }(x)$ for $x\in G$ and $\theta (C_H(r(x)A)) = C_G(x).$ Thus, working in H, we see that x is strictly $\alpha ^d$ -regular if and only if the cyclic group $C_H(r(x)A)/C_H(r(x))$ has order $d.$

Proposition 2.3. Let H be an $\alpha $ -covering group of $G.$ Then $x\in G$ is strictly $\alpha ^d$ -regular if and only if either:

  1. (a) $r(x)\langle z^m\rangle $ are the conjugates of $r(x)$ in $r(x)A,$ where $\langle z\rangle = A$ and $dm = n$ ; or

  2. (b) $\{r(x)z^i: i = 1,\ldots , m\}$ is a maximal set of conjugacy class representatives of H in $r(x)A.$

Proof. Define $k_{r(x)} : C_H(r(x)A)\rightarrow A$ by $k_{r(x)}(h) = hr(x)h^{-1}(r(x))^{-1}.$ Then $k_{r(x)}$ is a homomorphism with kernel $C_H(r(x))$ , since $\lambda (k_{r(x)}) = \alpha _x.$ Now let z be a generator of A. Then $r(x)z^i$ and $r(x)z^j$ are conjugate if and only if $z^{j-i}\in \operatorname {\mathrm {Im}}(k_{r(x)}),$ that is, if and only if $z^i\operatorname {\mathrm {Im}}(k_{r(x)}) = z^j\operatorname {\mathrm {Im}}(k_{r(x)}).$

Now x is strictly $\alpha ^d$ -regular if and only if $\operatorname {\mathrm {Im}}(k_{r(x)}) = \langle z^m\rangle ,$ that is, if and only if the cosets of $\operatorname {\mathrm {Im}}(k_{r(x)})$ in A are $z^i\langle z^m\rangle $ for $i =1,\ldots , m.$

3 Counting orbits of projective characters

Let N be a subgroup of $G.$ Let H be an $\alpha $ -covering group of G and, using the notation of Section 2, let M be the subgroup of H containing A such that $\theta (M) = N.$ Finally, for any integer k, let $\operatorname {\mathrm {Irr}}(M\vert \lambda ^k) = \{\chi \in \operatorname {\mathrm {Irr}}(M): \chi _A = \chi (1)\lambda ^k\},$ where $\operatorname {\mathrm {Irr}}(M)$ is the set of irreducible characters of $M.$ Then the mapping from $\operatorname {\mathrm {Proj}}(N, \alpha _N^k)$ to $\operatorname {\mathrm {Irr}}(M\vert \lambda ^k), \zeta \mapsto \chi $ is a bijection, where $\zeta (x) = \chi (r(x))$ for all $x\in N$ (see [Reference Karpilovsky4, pages 134–135] or [Reference Karpilovsky5, Corollary 4.1.3]). Now suppose N is normal in G, then it is easy to check that $\zeta ^g = \chi ^{r(g)}$ for all $g\in G$ and hence the orbit length of $\zeta $ under the action of G equals that of $\chi $ under the action of $H.$ By definition, for each $x\in G$ , there exists a unique $d\mid n$ such that x is strictly $\alpha ^d$ -regular. Thus, the conjugacy classes of H are partitioned according to $\vert C_H(r(x)A)/C_H(r(x)\vert $ for $r(x)a,$ where $x\in G$ and $a\in A.$ However, if x is a strictly $\alpha ^d$ -regular conjugacy class representative of $G,$ then $n/d$ corresponding conjugacy class representatives of H are obtained as detailed in Proposition 2.3. So the number of conjugacy classes of H in M corresponding to the number of $\alpha ^d$ -regular conjugacy classes of G contained in N is $\sum _{s|d} (n/s)t_s;$ in particular, $\sum _{d|n} (n/d)t_d = t(M),$ where $t(M)$ is the number of conjugacy classes of H contained in $M.$

Lemma 3.1. Let N be a normal subgroup of G and suppose that $o(\alpha ^d) = o(\alpha ^k).$ Let $\sigma $ be a field automorphism of $\mathbb {C}$ that extends $\tau $ , as described in Section 2, so that $\sigma (\alpha ^d) = \alpha ^k.$ Then $\zeta ^g = \zeta '$ if and only if $\sigma (\zeta )^{g} = \sigma (\zeta ')$ for $g\in G$ and $\zeta \in \operatorname {\mathrm {Proj}}(N, \alpha _N^d).$

Proof. If $\zeta \in \operatorname {\mathrm {Proj}}(N, \alpha _N^d),$ then $\sigma (\zeta )\in \operatorname {\mathrm {Proj}}(N, \sigma (\alpha _N^d)).$ Now

$$ \begin{align*}\sigma(\zeta)^g(x) = f_{\sigma(\alpha)}(g, x)\sigma(\zeta(gxg^{-1})) = \sigma(f_{\alpha}(g, x)\zeta(gxg^{-1}))\end{align*} $$

for all $x\in N.$

Lemma 3.1 sets up a one-to-one correspondence between the orbits of G under its action on $\operatorname {\mathrm {Proj}}(N, \alpha _N^d)$ and those under its action on $\operatorname {\mathrm {Proj}}(N, \alpha _N^k)$ in which orbit lengths are preserved. We next just restate Lemma 3.1 for an $\alpha $ -covering group H of $G.$

Corollary 3.2. Suppose that $o(\lambda ^d) = o(\lambda ^k)$ in $\langle \lambda \rangle = \operatorname {\mathrm {Lin}}(A).$ Let $\sigma $ be as in Lemma 3.1, so that $\sigma (\lambda ^d)=\lambda ^k.$ Then $\chi ^h = \chi '$ if and only if $\sigma (\chi )^h = \sigma (\chi ')$ for $h\in H$ and $\chi \in \operatorname {\mathrm {Irr}}(M\vert \lambda ^d).$

Let $\phi $ denote Euler’s totient function. We use the well-known result from number theory that $\sum _{d\mid n}\phi (d) = \sum _{d\mid n}\phi (n/d) = n.$

Theorem 3.3. Let N be a normal subgroup of $G.$ Then the number of orbits of G under its action on $\operatorname {\mathrm {Proj}}(N, \alpha _N)$ is equal to the number of $\alpha $ -regular conjugacy classes of G contained in $N.$

Proof. Proceeding by induction, we count the number of $\alpha ^d$ -regular conjugacy classes of G contained in N. First, if $d = n,$ then, as previously stated, the number of conjugacy classes of G contained in N is equal to the number of orbits of G under its action on $\operatorname {\mathrm {Irr}}(N).$ So assume by induction that the number of orbits of G under its action on $\operatorname {\mathrm {Proj}}(N, \alpha _N^d)$ is equal to the number of $\alpha ^d$ -regular conjugacy classes of G contained in N for each $d\mid n$ with $d\not = 1.$ Let H be an $\alpha $ -covering group of G and let M denote the subgroup of H containing A such that $\theta (M) = N.$

Now for $d\mid n$ and $d\not = 1$ , G has $\sum _{s\mid d} t_s$ orbits under its action on $\operatorname {\mathrm {Proj}}(N, \alpha _N^d).$ Thus, H has the same number of orbits under its action on $\operatorname {\mathrm {Irr}}(M\vert \lambda ^d).$ Now $o(\lambda ^k) = o(\lambda ^d)$ for $\phi (n/d)$ values of k with $1\leq k\leq n.$ Thus, using Corollary 3.2, the total number of orbits of H under its actions on $\operatorname {\mathrm {Irr}}(M\vert \lambda ^c),$ for the $n - \phi (n)$ values of c with $1\leq c\leq n$ that are not relatively prime to $n,$ is

$$ \begin{align*} \sum_{\substack{d\mid n\\ d\not=1}}\phi\bigg(\frac{n}{d}\bigg)\bigg(\sum_{s\mid d} t_s\bigg) &= \sum_{s\mid d} t_s\bigg(\sum_{\substack{d\mid n\\ d\not=1}}\phi\bigg(\frac{n}{d}\bigg)\bigg) \\ &= \sum_{s\mid n} t_s\bigg(\sum_{\substack{r\mid (n/s)\\ (r, s)\not=(1,1)}}\phi\bigg(\frac{n/s}{r}\bigg)\bigg) \\ &= t_1(n-\phi(n))+ \sum_{\substack{s\mid n\\ s\not=1}}t_s\frac{n}{s}. \end{align*} $$

The total number of orbits of H under its action on $\operatorname {\mathrm {Irr}}(M)$ is $t(M),$ so the total number of orbits of H under its actions on $\operatorname {\mathrm {Irr}}(M\vert \lambda ^c),$ for the $\phi (n)$ values of c with $1\leq c\leq n$ that are relatively prime to $n,$ is

$$ \begin{align*}t(M) - t_1(n-\phi(n)) - \sum_{\substack{s\mid n\\ s\not=1}} t_s\frac{n}{s} = t_1\phi(n).\end{align*} $$

Hence, the number of orbits of H under its action on $\operatorname {\mathrm {Irr}}(M\vert \lambda )$ (and the number of orbits of G under its action on $\operatorname {\mathrm {Proj}}(N, \alpha _N)$ ) is $t_1,$ as required.

Suppose that $\beta = t(\delta )\alpha .$ Then from [Reference Higgs1, Lemma 1.4], we see that $\operatorname {\mathrm {Proj}}(N, \beta _N) = \{\delta _N\zeta : \zeta \in \operatorname {\mathrm {Proj}}(N, \alpha _N)\}$ and, for $g\in G$ , $\zeta ^g = \zeta '$ if and only if $(\delta _N\zeta )^g = \delta _N\zeta '$ for $\zeta \in \operatorname {\mathrm {Proj}}(N, \alpha _N).$ In particular, this establishes a one-to-one correspondence between the orbits of G under its action on $\operatorname {\mathrm {Proj}}(N, \beta _N)$ and those under its action on $\operatorname {\mathrm {Proj}}(N, \alpha _N)$ in which orbit lengths are preserved. So from this and Lemma 3.1, the result of Theorem 3.3 is independent of the choice of $2$ -cocycle from $[\alpha ]^c$ for c relatively prime to $n.$

References

Higgs, R. J., ‘Projective characters of degree one and the inflation-restriction sequence’, J. Aust. Math. Soc. Ser. A 46(2) (1989), 272280.CrossRefGoogle Scholar
Higgs, R. J., ‘Projective representations of abelian groups’, J. Algebra 242(2) (2001), 769781.CrossRefGoogle Scholar
Isaacs, I. M., Character Theory of Finite Groups, Pure and Applied Mathematics, 69 (Academic Press [Harcourt Brace Jovanovich, Publishers], New York–London, 1976).Google Scholar
Karpilovsky, G., Group representations, Vol. 2, North-Holland Mathematics Studies, 177 (North-Holland Publishing Co., Amsterdam, 1993).Google Scholar
Karpilovsky, G., Group Representations, Vol. 3, North-Holland Mathematics Studies, 180 (North-Holland Publishing Co., Amsterdam, 1994).Google Scholar
Read, E. W., ‘On the centre of a representation group’, J. Lond. Math. Soc. (2) 16(1) (1977), 4350.CrossRefGoogle Scholar