This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the cohomological triviality of the center of the Frattini subgroup

Jaime Calles Universidad Nacional Autónoma de México
Instituto de Matemáticas
Investigación Científica, C.U.
Coyoacán, Ciudad de México, CDMX 04510
Mexico.
calles@im.unam.mx
José Cantarero Centro de Investigación en Matemáticas, A.C., Unidad Mérida
Parque Científico y Tecnológico de Yucatán
Carretera Sierra Papacal–Chuburná Puerto Km 5.5
Sierra Papacal, Mérida, YUC 97302
Mexico.
cantarero@cimat.mx
Juan Omar Gómez Centro de Investigación en Matemáticas, A.C., Unidad Mérida
Parque Científico y Tecnológico de Yucatán
Carretera Sierra Papacal–Chuburná Puerto Km 5.5
Sierra Papacal, Mérida, YUC 97302
Mexico.
juan.gomez@cimat.mx
 and  Gustavo Ortega Universidad Nacional Autónoma de México
Facultad de Ciencias
Investigación Científica, C.U.
Coyoacán, Ciudad de México, CDMX 04510
Mexico.
gustavo.ortega@ciencias.unam.mx
Abstract.

We improve the existing lower bounds on the order of counterexamples to a conjecture by P. Schmid, determine some properties of the possible counterexamples of minimum order for each prime, and the isomorphism type of the center of the Frattini subgroup for the counterexamples of order 256256. We also show that nonabelian metacyclic pp-groups, nonabelian groups of maximal nilpotency class and 22-groups of coclass two satisfy the conjecture.

Key words and phrases:
Finite pp-groups, Tate cohomology, cohomologically trivial modules
2020 Mathematics Subject Classification:
20D15, 20J05

Introduction

In 1973, Ya. G. Berkovich conjectured that every finite nonabelian pp-group admits a noninner automorphism of order pp. This is Problem 4.13 from the 4th issue of The Kouravka Notebook [12]. The existence of noninner automorphisms of pp-power order had been shown previously in [10].

This problem can be attacked by cohomological methods. In [15], P. Schmid showed that if GG is a regular nonabelian finite pp-group and Φ(G)\Phi(G) is its Frattini subgroup, then Z(Φ(G))Z(\Phi(G)) is not cohomologically trivial over the Frattini quotient of GG, and in this case this implies the existence of a noninner automorphism of order pp that fixes the Frattini subgroup elementwise. The article conjectured that Z(Φ(G))Z(\Phi(G)) is not a cohomologically trivial module over the Frattini quotient of GG for any finite nonabelian pp-group GG.

This conjecture turned out to be false. In [3], A. Abdollahi found ten counterexamples of order 256256 using GAP. Hence the direction of this problem shifted towards proving the conjecture for several families of groups, such as semi-abelian groups [6] and groups with small nilpotency class ([1], [3], [5]). Nonabelian pp-groups are called S–groups if they satisfy the conjecture and NS–groups otherwise.

In this article we contribute to the understanding of this problem in two directions. First, we improve the existing lower bound on the order of possible NS–groups, giving now new bounds that depend on pp. We show that the minimum order of an NS–group is at least pp+6p^{p+6} and find additional properties that an NS–group of order pp+6p^{p+6} should satisfy. In the case of order 256256, we are able to identify the center of the Frattini subgroup up to isomorphism, by a deeper analysis of cohomologically trivial modules over pp-groups which are finite pp-groups. The following theorem combines the most important results of Sections 1 and 2.

Theorem.

If GG is an NS–group, then |G|pp+6|G|\geq p^{p+6}. Moreover, if |G|=pp+6|G|=p^{p+6}, then

  • (a)

    d(G)=2d(G)=2

  • (b)

    |Z(Φ(G))|=pp+2|Z(\Phi(G))|=p^{p+2}

  • (c)

    Z(G)Z(G) is cyclic of order pp.

  • (d)

    If p=2p=2, then Z(Φ(G))Z(\Phi(G)) is isomorphic to (/2)4({\mathbb{Z}}/2)^{4} or (/4)×(/2)2({\mathbb{Z}}/4)\times({\mathbb{Z}}/2)^{2}.

In fact, we are able to improve these lower bounds for NS–groups that are mm-generated with m>2m>2, but no such groups are known so far. The ten NS–groups known at the moment are all 22-generated. In the computations carried out in Section 2, we found the following result which may be of independent interest.

Theorem.

Let MM be a finite abelian pp-group with an effective action of an elementary abelian pp-group EE of rank at least two. If MM is cohomologically trivial, then

  • MM is not cyclic.

  • MM is not isomorphic to /pk×/p{\mathbb{Z}}/p^{k}\times{\mathbb{Z}}/p for any k2k\geq 2.

  • MM is not isomorphic to (/p)k({\mathbb{Z}}/p)^{k} unless kk is a multiple of prp^{r}, where rr is the rank of EE.

  • If p=2p=2, then MM is not isomorphic to /4×/4{\mathbb{Z}}/4\times{\mathbb{Z}}/4.

Finally, in Section 3 we determine some new families of S–groups, inspired by results on Berkovich’s conjecture for groups of small coclass and the method of determining the trace map for metacyclic 22-groups. The following theorem outlines the main results of that section.

Theorem.

Any group in one of the following families is an SS–group.

  • Nonabelian maximal nilpotency class pp-groups.

  • 22-groups of coclass two.

  • Nonabelian metacyclic pp-groups.

Given the close connection between Berkovich’s and Schmid’s conjectures and the recent results about Berkovich’s conjecture on groups of coclass at most three [14], it seems possible that all groups in this family are S–groups. However, our methods do not apply to pp-groups of coclass two for p2p\neq 2 nor to pp-groups of coclass three. And it is worth noting that Berkovich’s conjecture is still unsettled for 33-groups of coclass three.

Notation and terminology. Given a finite group GG and a G{\mathbb{Z}}G–module MM, we denote by H^n(G;M)\hat{H}^{n}(G;M) the nnth Tate cohomology group of GG with coefficients in MM. It is convenient to introduce some terminology to describe the behaviour of nonabelian pp-groups with respect to the question raised by P. Schmid from the remark in page 3 of [15]. Namely, we say that a nonabelian pp-group GG is an S–group if

H^0(G/Φ(G);Z(Φ(G)))0\hat{H}^{0}(G/\Phi(G);Z(\Phi(G)))\neq 0

for the conjugation action of the Frattini quotient G/Φ(G)G/\Phi(G) on Z(Φ(G))Z(\Phi(G)). Otherwise, we say that GG is an NS–group. Note that by a theorem of Gaschütz and Uchida (see for instance Theorem 4 in [16]), a nonabelian pp-group GG is an NS–group if and only if Z(Φ(G))Z(\Phi(G)) is a cohomologically trivial G/Φ(G)G/\Phi(G)–module.

Given a Q{\mathbb{Z}}Q–module AA, we denote by [A,Q][A,Q] the subgroup of AA generated by elements of the form qaaq\cdot a-a. Since [A,Q][A,Q] is a QQ–submodule of AA, we can iterate this construction and define inductively

[A,Q,,Qn times]=[[A,Q,,Qn1 times],Q][A,\underbrace{Q,\ldots,Q}_{n\text{ times}}]=[[A,\underbrace{Q,\ldots,Q}_{n-1\text{ times}}],Q]

We denote by AQA^{Q} the submodule of fixed points and use AQ=A/[A,Q]A_{Q}=A/[A,Q] for the coinvariants. Note that our notation for invariants and coinvariants differs from the notation in [3], [15] and related articles. We refer to the map AQAQA_{Q}\to A^{Q} induced by aqQqaa\mapsto\sum_{q\in Q}qa as the trace. The rest of notation that we use is standard in group theory and group cohomology.

Acknowledgments. This project was seeded by the program “Research groups” that took place in the period January-June 2022 at CIMAT Mérida’s Algebraic Topology Seminar.

1. The conjecture for groups of small order

In this section we find a lower bound for NS–groups that depends on the minimum number of generators of the group. In particular, we show that Schmid’s conjecture holds for nonabelian pp-groups of order at most pp+5p^{p+5}.

We begin with the following proposition, which is implicit in the proof of Theorem 1.3 in [11].

Proposition 1.1.

Let GG be a finite nonabelian pp-group. If Φ(G)\Phi(G) is abelian, then GG is an SS–group.

Proof.

Assume that GG is an NS–group. Since H2(G/Φ(G);Φ(G))=0H^{2}(G/\Phi(G);\Phi(G))=0, the group GG is the semidirect product of G/Φ(G)G/\Phi(G) by Φ(G)\Phi(G). But the elements of Φ(G)\Phi(G) are non-generators, hence the quotient GG/Φ(G)G\to G/\Phi(G) is an isomorphism. Then Φ(G)\Phi(G) is trivial, which contradicts the fact that GG is nonabelian. ∎

Theorem 1.2.

If GG is an NS–group with d(G)=md(G)=m, then |Φ(G)|pm+p+2|\Phi(G)|\geq p^{m+p+2}.

Proof.

Let A=Z(Φ(G))A=Z(\Phi(G)) and Q=G/Φ(G)Q=G/\Phi(G). Since AA is a cohomologically trivial QQ–module,

|A|=|AQ||AQQ||[A,Q,Q]||A|=|A^{Q}|\cdot|A^{Q}\otimes Q|\cdot|[A,Q,Q]|

by Corollary 2.2 in [5]. Let A(n)=[A,Q,,Q]A(n)=[A,Q,\ldots,Q], where QQ appears nn times. Since GG is an NS–group, if follows by Theorem 1.3 from [5] that Z(Φ(G))Z(\Phi(G)) is not contained in Zp(G)Z_{p}(G). Therefore A(p)0A(p)\neq 0 and so |[A,Q,Q]|pp1|[A,Q,Q]|\geq p^{p-1}. On the other hand, the subgroup of QQ–fixed points of AA is not trivial, hence |AQ|p|A^{Q}|\geq p and |AQQ|pm|A^{Q}\otimes Q|\geq p^{m}. We obtain |Z(Φ(G))|pm+p|Z(\Phi(G))|\geq p^{m+p} and by Proposition 1.1, we have |Φ(G)|pm+p+2|\Phi(G)|\geq p^{m+p+2}. ∎

In particular, we get a lower bound on the order of NS–groups depending on the minimum number of generators.

Corollary 1.3.

If GG is an NS–group with d(G)=md(G)=m, then |G|p2m+p+2|G|\geq p^{2m+p+2}.

Corollary 1.4.

If GG is a pp-group with |Φ(G)|<pp+4|\Phi(G)|<p^{p+4} or |G|<pp+6|G|<p^{p+6}, then it is an S–group.

For p=2p=2, this result was shown in [3] using GAP and later in [11] mathematically. Note that ten 22-generated NS–groups of order 282^{8} were found in [3] using GAP. The following proposition shows that, should it exist, an NS–group of order pp+6p^{p+6} would behave similarly to those groups.

Proposition 1.5.

Let GG be an NS–group of order pp+6p^{p+6}. Then we have:

  • (a)

    d(G)=2d(G)=2.

  • (b)

    |Z(Φ(G))|=pp+2|Z(\Phi(G))|=p^{p+2}.

  • (c)

    Z(G)Z(G) is cyclic of order pp.

Proof.

The proof of Theorem 1.2 showed that |Z(Φ(G))|pp+d(G)|Z(\Phi(G))|\geq p^{p+d(G)}. If d(G)3d(G)\geq 3, we would have |Z(Φ(G))|p3+p|Z(\Phi(G))|\geq p^{3+p} and then [G:Φ(G)]p[G:\Phi(G)]\leq p. But then Φ(G)\Phi(G) would be abelian and by Proposition 1.1, GG would be an S–group. Thus d(G)=2d(G)=2 and therefore |Φ(G)|=pp+4|\Phi(G)|=p^{p+4}. The order of Z(Φ(G))Z(\Phi(G)) must be exactly pp+2p^{p+2}, otherwise Φ(G)\Phi(G) would be abelian.

Let A=Z(Φ(G))A=Z(\Phi(G)) and Q=G/Φ(G)Q=G/\Phi(G). Note that AQ=Z(G)A^{Q}=Z(G). If |AQ|>p|A^{Q}|>p, then we would have |QAQ|p2|Q\otimes A^{Q}|\geq p^{2} and so

|A|=|AQ||AQQ||[A,Q,Q]|pp+3|A|=|A^{Q}|\cdot|A^{Q}\otimes Q|\cdot|[A,Q,Q]|\geq p^{p+3}

which contradicts part (b). Hence Z(G)Z(G) is cyclic of order pp. ∎

We are particularly interested in the NS–groups of order 282^{8}, and we see in this proposition that for such groups, the order of the center of the Frattini subgroup is 242^{4}. In the next section we will identify the isomorphism type of Z(Φ(G))Z(\Phi(G)) for these groups.

Remark 1.6.

NS–groups are natural choices when looking for counterexamples to Berkovich’s conjecture. However, one can find in GAP [9] that the ten NS–groups of order 282^{8} found in [3] do have noninner automorphisms of order 22. We include below a table with the number of order-two noninner automorphisms for each of these groups.

IdSmallGroup[256,i][256,i] 298298 299299 300300 301301 302302 303303 304304 305305 306306 307307
Order-22 noninner aut. 576576 576576 512512 512512 576576 576576 576576 704704 704704 576576

2. Cohomologically trivial faithful finite modules

The results in the previous section led to lower bounds on the order of the center of the Frattini subgroup of an NS–group. In this section we find further restrictions on the isomorphism type of the center of the Frattini subgroup. This is achieved thanks to the following lemma, which places a constraint on the action of the Frattini quotient.

Lemma 2.1.

Let GG be an NS–group. Then the action of G/Φ(G)G/\Phi(G) on Z(Φ(G))Z(\Phi(G)) is effective.

Proof.

Note that gΦ(G)g\Phi(G) acts trivially on Z(Φ(G))Z(\Phi(G)) if and only if gCG(Z(Φ(G))g\in C_{G}(Z(\Phi(G)). But CG(Z(Φ(G))=Φ(G)C_{G}(Z(\Phi(G))=\Phi(G) by the corollary on page 2 of [15], hence gΦ(G)=Φ(G)g\Phi(G)=\Phi(G) and so the action is effective. ∎

Now let GG be an NS–group. Since GG is not cyclic, we have d(G)=r2d(G)=r\geq 2. Then Z(Φ(G))Z(\Phi(G)) is an abelian pp-group with an effective action of (/p)r({\mathbb{Z}}/p)^{r}, which is cohomologically trivial. We will exploit this fact to rule out some possibilities for Z(Φ(G))Z(\Phi(G)). First we dismiss cyclic pp-groups.

Proposition 2.2.

Let MM be a cyclic pp-group with an effective action by automorphisms of an elementary abelian pp-group of rank at least two. Then MM is not cohomologically trivial.

Proof.

If pp is odd, the group Aut(/pk)\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k}) is cyclic of order pk1(p1)p^{k-1}(p-1), so it does not have any elementary abelian pp-subgroup of rank two. Therefore p=2p=2 and M/2kM\cong{\mathbb{Z}}/2^{k} for some kk. Since Aut(/2)\operatorname{Aut}\nolimits({\mathbb{Z}}/2) is trivial and Aut(/4)\operatorname{Aut}\nolimits({\mathbb{Z}}/4) is cyclic, we must have k3k\geq 3. In this case Aut(/2k)/2×/2k2\operatorname{Aut}\nolimits({\mathbb{Z}}/2^{k})\cong{\mathbb{Z}}/2\times{\mathbb{Z}}/2^{k-2} contains a unique elementary abelian 22-subgroup of rank two. Since this subgroup contains the automorphism that sends each element to its inverse, the trace map is trivial and therefore MM is not cohomologically trivial. ∎

Next we exclude certain elementary abelian pp-groups.

Proposition 2.3.

Let MM be an elementary abelian pp-group which is cohomologically trivial for the action by automorphisms of an elementary abelian pp-group of rank r2r\geq 2. Then the rank of MM is a multiple of prp^{r}.

Proof.

If an elementary abelian pp-group has an action by automorphism of E=(/p)rE=({\mathbb{Z}}/p)^{r}, then it can be regarded as an 𝔽pE{\mathbb{F}}_{p}E–module. By Theorem VI.8.5 in [8], it is cohomologically trivial if and only if it is free as a 𝔽pE{\mathbb{F}}_{p}E–module. The result follows since 𝔽pE(/p)pr{\mathbb{F}}_{p}E\cong({\mathbb{Z}}/p)^{p^{r}} as an abelian group. ∎

In order to eliminate groups of the form /pk×/p{\mathbb{Z}}/p^{k}\times{\mathbb{Z}}/p, we need the following auxiliary result.

Lemma 2.4.

Let k2k\geq 2. If there exists a nonnegative integer rr with 0r<2k0\leq r<2^{k} and r212k1mod 2kr^{2}-1\equiv 2^{k-1}\ \mathrm{mod}\ 2^{k}, then k4k\geq 4 and

r{2k2±1,32k2±1}r\in\{2^{k-2}\pm 1,3\cdot 2^{k-2}\pm 1\}
Proof.

Note that rr must be odd, say r=2n+1r=2n+1. Then r2=4n2+4n+1r^{2}=4n^{2}+4n+1, which is congruent to 11 modulo 88. Hence there are no solutions to the desired congruence when kk equals two or three. Hence assume k4k\geq 4 from now on. It is easy to check that 2k2±12^{k-2}\pm 1 and 32k2±13\cdot 2^{k-2}\pm 1 are solutions to the congruence. If rr is a solution, we must have (r1)(r+1)=2k1c(r-1)(r+1)=2^{k-1}c for some cc odd. Then r1=2amr-1=2^{a}m and r+1=2bnr+1=2^{b}n with mn=cmn=c and a+b=k1a+b=k-1. Then

2am+2=2bn2^{a}m+2=2^{b}n

Checking the parity we see that a=0a=0 if and only if b=0b=0, but this possibility is incompatible with k2k\geq 2. Therefore aa and bb must be positive and so

2a1m+1=2b1n2^{a-1}m+1=2^{b-1}n

Checking parity again, we find that either a=1a=1 or b=1b=1. If a=1a=1, then b=k2b=k-2 and so r=2k2n1r=2^{k-2}n-1. If b=1b=1, then a=k2a=k-2 and r=2k2m+1r=2^{k-2}m+1. Since r<2kr<2^{k}, the only possible values of nn and mm are one and three. ∎

For groups of the form /pk×/p{\mathbb{Z}}/p^{k}\times{\mathbb{Z}}/p, we will perform a delicate analysis of the elementary abelian pp-subgroups of their automorphism groups.

Proposition 2.5.

Let k2k\geq 2. If an elementary abelian pp-group of rank at least two acts effectively on M=/pk×/pM={\mathbb{Z}}/p^{k}\times{\mathbb{Z}}/p by automorphisms, then MM is not cohomologically trivial.

Proof.

Let q:/pk/pq\colon{\mathbb{Z}}/p^{k}\to{\mathbb{Z}}/p be mod pp reduction and let j:/p/pkj\colon{\mathbb{Z}}/p\to{\mathbb{Z}}/p^{k} be the standard inclusion. We have qj=0qj=0 and jq=pk1jq=p^{k-1}, where we are denoting by nn the automorphism [x][nx][x]\mapsto[nx].

We first treat the case p=2p=2. Note that every automorphism of /2k×/2{\mathbb{Z}}/2^{k}\times{\mathbb{Z}}/2 has the form

A=(αmjnq1)A=\left(\begin{array}[]{cc}\alpha&mj\\ nq&1\end{array}\right)

where αAut(/2k)\alpha\in\operatorname{Aut}\nolimits({\mathbb{Z}}/2^{k}) and m,n{0,1}m,n\in\{0,1\}. We refer to α\alpha as the corner of the automorphism. Since αj=j\alpha j=j and qα=qq\alpha=q, we have

A2=(α2+2k1mn001)A^{2}=\left(\begin{array}[]{cc}\alpha^{2}+2^{k-1}mn&0\\ 0&1\end{array}\right)

Hence A2=IA^{2}=I if and only if α2+2k1mn=1\alpha^{2}+2^{k-1}mn=1. If α([1])=[r]\alpha([1])=[r] with 0r<2k0\leq r<2^{k}, this holds if and only if

r2+2k1mn1mod 2kr^{2}+2^{k-1}mn\equiv 1\ \mathrm{mod}\ 2^{k}

If mn=0mn=0, then r21mod 2kr^{2}\equiv 1\ \mathrm{mod}\ 2^{k} and it is well known that r{1,1,2k1+1,2k11}r\in\{1,-1,2^{k-1}+1,2^{k-1}-1\}. If mn=1mn=1, then m=n=1m=n=1 and

r212k1mod 2kr^{2}-1\equiv 2^{k-1}\ \mathrm{mod}\ 2^{k}

By Lemma 2.4, we obtain k4k\geq 4 and r{2k2±1,32k2±1}r\in\{2^{k-2}\pm 1,3\cdot 2^{k-2}\pm 1\}. Hence if k4k\geq 4, we have automorphisms of order 22 of the form

(rjq1)\left(\begin{array}[]{cc}r&j\\ q&1\end{array}\right)

We use the following homomorphism

t:Aut(/2k×/2)/2/2\displaystyle t\colon\operatorname{Aut}\nolimits({\mathbb{Z}}/2^{k}\times{\mathbb{Z}}/2)\to{\mathbb{Z}}/2\oplus{\mathbb{Z}}/2
(αmjnq1)([m],[n])\displaystyle\left(\begin{array}[]{cc}\alpha&mj\\ nq&1\end{array}\right)\mapsto([m],[n])

to distinguish different types of automorphisms. More precisely, we say that t(A)t(A) is the type of AA. We outline in the following table when these order-two automorphisms commute depending on their type

([0],[0]) ([1],[0]) ([0],[1]) ([1],[1])
([0],[0]) Yes Yes Yes Yes
([1],[0]) Yes Yes No No
([0],[1]) Yes No Yes No
([1],[1]) Yes No No Yes

Let us consider first the case of an efective action of (/2)2({\mathbb{Z}}/2)^{2}. The corresponding subgroup of automorphisms is generated by two order-two commuting automorphisms, say with corners rr and rr^{\prime}. From the table above and the fact that tt is a homomorphism, we obtain that the trace map is represented by the matrix

(1+r+r+rr000)=((1+r)(1+r)000)\left(\begin{array}[]{cc}1+r+r^{\prime}+rr^{\prime}&0\\ 0&0\end{array}\right)=\left(\begin{array}[]{cc}(1+r)(1+r^{\prime})&0\\ 0&0\end{array}\right)

The element ([0],[1])([0],[1]) is not in the image of the trace and it is fixed by order-two automorphisms of type ([0],[0])([0],[0]) and ([0],[1])([0],[1]), hence we only need to to check two cases.

Case 1: The subgroup is generated by an automorphism of type ([0],[0])([0],[0]) and an automorphism of type ([1],[0])([1],[0]). If the automorphism of type ([0],[0])([0],[0]) is 1-1, the trace map is trivial. If r=2k11r=2^{k-1}-1, then

(1+r)(1+r)=2k1(1+r)(1+r)(1+r^{\prime})=2^{k-1}(1+r^{\prime})

Since rr^{\prime} is odd, the trace map is also zero. It is also zero when r=1r^{\prime}=-1 or r=2k11r^{\prime}=2^{k-1}-1. When r=2k1+1r=2^{k-1}+1 and r{2k1+1,1}r^{\prime}\in\{2^{k-1}+1,1\}, then

(1+r)(1+r)=2(2k2+1)(1+r)(1+r)(1+r^{\prime})=2(2^{k-2}+1)(1+r^{\prime})

which is a multiple of 44. However, ([2],[0])([2],[0]) is a fixed point.

Case 2: The subgroup is generated by an automorphism of type ([0],[0])([0],[0]) and an automorphism of type ([1],[1])([1],[1]). This can only happen if k4k\geq 4 by Lemma 2.4. Again we can assume that r=2k1+1r=2^{k-1}+1, since the trace map vanishes otherwise. Then if r=2k2n+1r^{\prime}=2^{k-2}n+1, the element ([2],[1])([2],[1]) is fixed and if r=2k2n1r^{\prime}=2^{k-2}n-1, the element ([2k2],[1])([2^{k-2}],[1]) is fixed. None of these elements belong to the image of the trace.

Now consider an effective action of an elementary abelian 22-group of rank r>2r>2. If the corresponding subgroup of automorphisms is generated by elements of type ([0],[0])([0],[0]) and ([0],[1])([0],[1]), the trace map is zero. Otherwise we need to consider two cases again.

Case 1: The subgroup is generated by automorphisms of type ([0],[0])([0],[0]) and type ([1],[0])([1],[0]). If any of the generating automorphisms has corner r=1r=-1 or r=2k11r=2^{k-1}-1, the trace map is trivial. So we can assume that the group is generated by the automorphism of type ([0],[0])([0],[0]) with corner r=2k1+1r=2^{k-1}+1 and the automorphisms of type ([1],[0])([1],[0]) with corners 2k1+12^{k-1}+1 and 11. In this case we have

(1+r)(1+r)(1+r′′)=(2k1+2)(2k1+2)28mod 2k(1+r)(1+r^{\prime})(1+r^{\prime\prime})=(2^{k-1}+2)(2^{k-1}+2)2\equiv 8\ \mathrm{mod}\ 2^{k}

Hence ([2],[0])([2],[0]) is a fixed point which does not belong to the image of the trace.

Case 2: The subgroup is generated by automorphisms of type ([0],[0])([0],[0]) and type ([1],[1])([1],[1]). We can again assume that there is only one automorphism of type ([0],[0])([0],[0]) with corner r=2k1+1r=2^{k-1}+1. If there is at least one automorphism of type ([1],[1])([1],[1]) with corner r=2k2n1r^{\prime}=2^{k-2}n-1, then

(1+r)(1+r)(1+r′′)=(2k1+2)2k2n(1+r)0mod 2k(1+r)(1+r^{\prime})(1+r^{\prime\prime})=(2^{k-1}+2)2^{k-2}n(1+r^{\prime})\equiv 0\ \mathrm{mod}\ 2^{k}

and the trace map is zero. Hence we can assume that the group is generated by the automorphism of type ([0],[0])([0],[0]) and corner r=2k1+1r=2^{k-1}+1 and the automorphisms of type ([1],[1])([1],[1]) with corners r=2k2+1r^{\prime}=2^{k-2}+1 and r′′=32k2+1r^{\prime\prime}=3\cdot 2^{k-2}+1. In this case the element ([2],[1])([2],[1]) is fixed and not in the image of the trace map.

We now treat the case p2p\neq 2, where every automorphism of /pk×/p{\mathbb{Z}}/p^{k}\times{\mathbb{Z}}/p has the form

A=(αmjnqβ)A=\left(\begin{array}[]{cc}\alpha&mj\\ nq&\beta\end{array}\right)

with αAut(/pk)\alpha\in\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k}), βAut(/p)\beta\in\operatorname{Aut}\nolimits({\mathbb{Z}}/p) and m,n{0,,p1}m,n\in\{0,\ldots,p-1\}. Recall that Aut(/pk)/pk1(p1)\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k})\cong{\mathbb{Z}}/p^{k-1}(p-1). Since the homomorphism

Aut(/pk)(/p)×\displaystyle\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k})\to({\mathbb{Z}}/p)^{\times}
α[α(1)]\displaystyle\alpha\mapsto[\alpha(1)]

is surjective, a pp-Sylow of Aut(/pk)\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k}) is given by

{αAut(/pk)α(1)1modp}\{\alpha\in\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k})\mid\alpha(1)\equiv 1\ \mathrm{mod}\ p\}

Therefore a pp-Sylow of Aut(/pk×/p)\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k}\times{\mathbb{Z}}/p) is given by

S={(αmjnq1)Aut(/pk×/p)|α(1)1modp}S=\left\{\left(\begin{array}[]{cc}\alpha&mj\\ nq&1\end{array}\right)\in\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k}\times{\mathbb{Z}}/p)\>\Bigg{|}\>\alpha(1)\equiv 1\ \mathrm{mod}\ p\right\}

Note that these automorphisms α\alpha in the first entry of elements of SS fix [pk1][p^{k-1}]. Hence αj\alpha j and qα=qq\alpha=q. Given such an automorphism AA in SS, by induction

Ar=(αr+r(r1)2mnpk1rmjrnq1)A^{r}=\left(\begin{array}[]{cc}\alpha^{r}+\frac{r(r-1)}{2}mnp^{k-1}&rmj\\ &\\ rnq&1\end{array}\right)

and in particular

Ap=(αp001)A^{p}=\left(\begin{array}[]{cc}\alpha^{p}&0\\ 0&1\end{array}\right)

We see then that Ap=1A^{p}=1 if and only if αp=1\alpha^{p}=1. Since Aut(/pk)\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k}) has a unique subgroup of order pp, this happens if and only if α([1])=[apk1+1]\alpha([1])=[ap^{k-1}+1] for some 0ap10\leq a\leq p-1. For the corresponding AA we compute

I+A++Ap1={(p000) if p5(3+3k1mn000) if p=3I+A+\ldots+A^{p-1}=\left\{\begin{array}[]{ll}\left(\begin{array}[]{cc}p&0\\ 0&0\end{array}\right)&\text{ if }p\geq 5\\ \\ \left(\begin{array}[]{cc}3+3^{k-1}mn&0\\ 0&0\end{array}\right)&\text{ if }p=3\end{array}\right.

Any elementary abelian pp-subgroup of Aut(/pk×/p)\operatorname{Aut}\nolimits({\mathbb{Z}}/p^{k}\times{\mathbb{Z}}/p) is conjugate to a subgroup of SS. By the computation above, if the rank of this subgroup is at least two, then the image of the trace map is contained in p2/pk×{0}p^{2}{\mathbb{Z}}/p^{k}\times\{0\}. But ([p],[0])([p],[0]) is always a fixed point, hence the module is not cohomologically trivial. ∎

We arrive at the main result of this section by analyzing /4×/4{\mathbb{Z}}/4\times{\mathbb{Z}}/4.

Theorem 2.6.

Let MM be a finite abelian 22-group of order at most 242^{4} which is cohomologically trivial for the effective action of an elementary abelian 22-group of rank at least two. Then MM is isomorphic to (/2)4({\mathbb{Z}}/2)^{4} or /4×(/2)2{\mathbb{Z}}/4\times({\mathbb{Z}}/2)^{2}.

Proof.

By Propositions 2.2, 2.3 and 2.5, it suffices to show that /4×/4{\mathbb{Z}}/4\times{\mathbb{Z}}/4 can not be a cohomologically trivial effective EE–module for any elementary abelian 22-group EE of rank at least two.

Let us regard the elements of Aut(/4×/4)\operatorname{Aut}\nolimits({\mathbb{Z}}/4\times{\mathbb{Z}}/4) as elements of M2×2(/4)M_{2\times 2}({\mathbb{Z}}/4). Consider the subgroup 2(/4×/4)/2×/22({\mathbb{Z}}/4\times{\mathbb{Z}}/4)\cong{\mathbb{Z}}/2\times{\mathbb{Z}}/2. The restriction

res:Aut(/4×/4)Aut(/2×/2)\operatorname{res}\nolimits\colon\operatorname{Aut}\nolimits({\mathbb{Z}}/4\times{\mathbb{Z}}/4)\to\operatorname{Aut}\nolimits({\mathbb{Z}}/2\times{\mathbb{Z}}/2)

has kernel

K={I+2BBM2×2(/4)}(/2)4K=\{I+2B\mid B\in M_{2\times 2}({\mathbb{Z}}/4)\}\cong({\mathbb{Z}}/2)^{4}

It fits into a short exact sequence

1(/2)4Aut(/4×/4)Σ311\to({\mathbb{Z}}/2)^{4}\to\operatorname{Aut}\nolimits({\mathbb{Z}}/4\times{\mathbb{Z}}/4)\to\Sigma_{3}\to 1

Let AA be the subgroup of Aut(/2×/2)\operatorname{Aut}\nolimits({\mathbb{Z}}/2\times{\mathbb{Z}}/2) generated by the automorphism σ\sigma that permutes the coordinates, and let S=res1(A)S=\operatorname{res}\nolimits^{-1}(A). The subgroup SS is a 22-Sylow of Aut(/4×/4)\operatorname{Aut}\nolimits({\mathbb{Z}}/4\times{\mathbb{Z}}/4) and it is a semidirect product (/2)4/2({\mathbb{Z}}/2)^{4}\rtimes{\mathbb{Z}}/2. Hence we may assume that EE is a subgroup of SS. An element I+2BI+2B commutes with (I+2B)σ(I+2B^{\prime})\sigma if and only if 2B2B has the form

(abba)\left(\begin{array}[]{cc}a&b\\ b&a\end{array}\right)

Similarly, the element (I+2B)σ(I+2B^{\prime})\sigma has order two if and only if 2B2B^{\prime} has the same form.

Assume first EE is generated by elements I+2BI+2B and I+2BI+2B^{\prime}. Then the trace map is given by

I+(I+2B)+(I+2B)+(I+2B+2B)=0I+(I+2B)+(I+2B^{\prime})+(I+2B+2B^{\prime})=0

and therefore MM is not cohomologically trivial. If EE is generated by elements I+2BI+2B and (I+2B)σ(I+2B^{\prime})\sigma, the trace map has the form

τ\displaystyle\tau =I+(I+2B)+(I+2B)σ+(I+2B+2B)σ\displaystyle=I+(I+2B)+(I+2B^{\prime})\sigma+(I+2B+2B^{\prime})\sigma
=2I+2B+2σ+2Bσ\displaystyle=2I+2B+2\sigma+2B\sigma
=2(I+B)(1+σ)\displaystyle=2(I+B)(1+\sigma)
=(2+a+b2+a+b2+a+b2+a+b)\displaystyle=\left(\begin{array}[]{cc}2+a+b&2+a+b\\ 2+a+b&2+a+b\end{array}\right)

If a+b=2a+b=2, the trace map is trivial and MM is not cohomologically trivial. If a+b=0a+b=0, the image of τ\tau is the subgroup generated by (2,2)(2,2). Note that in this case a=b=2a=b=2 and so

I+2B=(3223)I+2B=\left(\begin{array}[]{cc}3&2\\ 2&3\end{array}\right)

On the other hand,

(I+2B)σ=(d11+c1+cd1)(I+2B^{\prime})\sigma=\left(\begin{array}[]{cc}d-1&1+c\\ 1+c&d-1\end{array}\right)

for certain elements cc, d{0,2}d\in\{0,2\}. If d1=1+cd-1=1+c, then (1,3)(1,3) is fixed by both matrices. If d11+cd-1\neq 1+c, then (1,1)(1,1) is fixed by both matrices. In any case, MM is not cohomologically trivial.

Finally, if the rank of EE is greater than two, it must have an elementary abelian 22-subgroup that is contained in KK and therefore the trace map is trivial. ∎

Recall from the previous section that for an NS–group GG of order 282^{8}, the center of its Frattini subgroup must have order 242^{4}. The following corollary is immediate from the previous theorem.

Corollary 2.7.

If GG is an NS–group of order 282^{8}, then Z(Φ(G))Z(\Phi(G)) is isomorphic to /4×(/2)2{\mathbb{Z}}/4\times({\mathbb{Z}}/2)^{2} or (/2)4({\mathbb{Z}}/2)^{4}.

When we check the list of ten NS–groups of order 282^{8} using GAP, we see that the center of the Frattini subgroup is isomorphic to (/2)4({\mathbb{Z}}/2)^{4} for six of them and /4×(/2)2{\mathbb{Z}}/4\times({\mathbb{Z}}/2)^{2} for the rest, so this result is the best possible.

Corollary 2.8.

If GG is an NS–group of order 282^{8}, then Z(Φ(G))/Z(G)(/2)3Z(\Phi(G))/Z(G)\cong({\mathbb{Z}}/2)^{3}.

Proof.

The center of an NS–group GG of order 282^{8} is isomorphic to /2{\mathbb{Z}}/2 by Proposition 1.5. Since /2×{0}{\mathbb{Z}}/2\times\{0\} is a characteristic subgroup of /4×/2{\mathbb{Z}}/4\times{\mathbb{Z}}/2, we see that if Z(Φ(G))/4×/2Z(\Phi(G))\cong{\mathbb{Z}}/4\times{\mathbb{Z}}/2, then the inclusion of Z(G)Z(G) in Z(Φ(G))Z(\Phi(G)) corresponds to the inclusion of /2×{0}{\mathbb{Z}}/2\times\{0\} in /4×/2{\mathbb{Z}}/4\times{\mathbb{Z}}/2. ∎

3. Metacyclic groups and groups of small coclass

Motivated by previous works on Berkovich’s conjecture for groups of small coclass, in this section we study Schmid’s conjecture for 22-groups of coclass at most two and nonabelian metacyclic pp-groups. The case of groups of coclass one follows easily from a result in [3].

Proposition 3.1.

Let GG be a nonabelian pp-group of maximal nilpotency class. Then GG is an S–group.

Proof.

If the order of GG is p3p^{3}, then the nilpotency class of GG is two, and it follows by Theorem 3.6 in [2]. Assume now that |G|>p3|G|>p^{3}. Since GG has maximal nilpotency class, we have that Z2(G)/Z(G)Z_{2}(G)/Z(G) is cyclic and then

d(G)d(Z(G))d(Z2(G)Z(G))d(G)d(Z(G))\neq d\left(\frac{Z_{2}(G)}{Z(G)}\right)

since d(G)2d(G)\geq 2. We conclude that GG is an S–group by Lemma 3.8 in [3]. ∎

For the next result, we were inspired by the article [4], where Berkovich’s conjecture was established for pp-groups of coclass two.

Proposition 3.2.

Let GG be a finite 22-group of coclass two. Then GG is an S–group.

Proof.

Suppose that GG is an NS–group of order 2n2^{n}. We may assume n4n\geq 4, otherwise the result holds by Theorem 1.2 in [5]. By Lemma 3.8 in [3], we have that d(Z2(G)/Z(G))=d(G)d(Z(G))d(Z_{2}(G)/Z(G))=d(G)d(Z(G)). Since GG is of coclass two, the order of Z2(G)/Z(G)Z_{2}(G)/Z(G) is at most four, hence d(Z2(G)/Z(G))2d(Z_{2}(G)/Z(G))\leq 2. The group GG is not cyclic, thus d(G)=2=d(Z2(G)/Z(G))d(G)=2=d(Z_{2}(G)/Z(G)) and therefore Z(G)Z(G) is cyclic of order 22, the group Z2(G)/Z(G)Z_{2}(G)/Z(G) is elementary abelian of rank two and Z2(G)Z_{2}(G) is noncyclic of order 232^{3}. Therefore Zn2(G)=Φ(G)Z_{n-2}(G)=\Phi(G).

The elements of Z2(G)Z_{2}(G) commute with commutators. Since Z2(G)/Z(G)Z_{2}(G)/Z(G) is elementary abelian, we have x2Z(G)x^{2}\in Z(G) whenever xZ2(G)x\in Z_{2}(G). Then for all gGg\in G, we have

1=[x2,g]=[x,g]2=[x,g2]1=[x^{2},g]=[x,g]^{2}=[x,g^{2}]

and therefore the elements of Z2(G)Z_{2}(G) commute with the elements of Φ(G)\Phi(G). Thus Z2(G)Z(Φ(G))Z_{2}(G)\leq Z(\Phi(G)). Under these circumstances, the case p=2p=2 of the proof of Theorem 3.1 in [4] shows that [G,G][G,G] is cyclic. Note that |Φ(G)/[G,G]|=|Zn2(G)/[G,G]|2|\Phi(G)/[G,G]|=|Z_{n-2}(G)/[G,G]|\leq 2. By Proposition 1.1 we deduce that this order must be 22, hence [G,G][G,G] is a maximal subgroup of Φ(G)\Phi(G). By the classification of 22-groups with a cyclic maximal subgroup (see Result 5.3.4 in [13], for instance), the Frattini subgroup Φ(G)\Phi(G) must be abelian or have cyclic center. Hence the result follows by Propositions 1.1 and 2.2. ∎

Remark 3.3.

By Theorem 1.1 in [4], any finite pp-group GG of coclass 22 has a noninner automorphism of order 22 that fixes every element of the center. However, this is not enough to guarantee that GG is an S–group.

Remark 3.4.

Using GAP, one can see that a group GG of order 282^{8} is an NS–group if and only if its nilpotency class is four and Z(Φ(G))/Z(G)(/2)3Z(\Phi(G))/Z(G)\cong({\mathbb{Z}}/2)^{3}. By Theorem 1.2 in [5] and the results in this article, we can say that an NS–group GG of order 282^{8} has nilpotency class four or five and must satisfy Z(Φ(G))/Z(G)(/2)3Z(\Phi(G))/Z(G)\cong({\mathbb{Z}}/2)^{3}. Only the case of groups of order 282^{8} with coclass three needs to be dismissed to achieve one of the implications in this equivalence.

We end this section with an analysis of metacyclic 22-groups.

Proposition 3.5.

Let GG be a nonabelian metacyclic 22-group. Then GG is an S–group.

Proof.

Note that if GG has a cyclic maximal subgroup, then Φ(G)\Phi(G) is abelian, hence GG is an S–group by Proposition 1.1. Now assume that GG is an NS–group, in particular it does not have a maximal cyclic subgroup. By Theorem 2.2 in [17], the group GG is generated by two elements aa, bb such that [G,G][G,G] is generated by a power of aa.

Let us denote A=Z(Φ(G))A=Z(\Phi(G)) and Q=G/Φ(G)Q=G/\Phi(G). Consider the cyclic subgroup HgH_{g} of QQ generated by a nontrivial element gΦ(G)g\Phi(G). The trace map for the action of HgH_{g} on AA is given by

τg(x)=xxg=xg1xg=x2[x,g]=x2am\tau_{g}(x)=xx^{g}=xg^{-1}xg=x^{2}[x,g]=x^{2}a^{m}

for some mm, where the last equality holds because [G,G][G,G] is generated by a power of aa. Since GG is an NS–group, the image of τg\tau_{g} equals AHgA^{H_{g}} and by Proposition 1 in [15], we have CQ(AHg)=HgC_{Q}(A^{H_{g}})=H_{g}. In particular aΦ(G)CQ(AHa)a\Phi(G)\in C_{Q}(A^{H_{a}}) and therefore aa commutes with x2x^{2} for all xAx\in A. Then aa commutes with the image of τb\tau_{b}, hence aΦ(G)a\Phi(G) belongs to CQ(AHb)=HbC_{Q}(A^{H_{b}})=H_{b}, which is a contradiction. Thus GG is an S–group. ∎

Note that for p>2p>2, nonabelian metacyclic pp-groups are regular groups (by Theorem 9.8(a) and Theorem 9.11 from [7]), hence S–groups by the main theorem in [15]. Therefore this proposition settles the conjecture for all nonabelian metacyclic pp-groups.

Theorem 3.6.

Nonabelian metacyclic pp-groups are S–groups.

We now give an example of a nonabelian metacyclic 22-group for which the criteria to be an S–group developed in this paper and other articles do not apply, to emphasize that Proposition 3.5 is finding new S–groups.

The automorphism φ\varphi of /32{\mathbb{Z}}/32 which sends [x][x] to [11x][11x] has order eight. Consider the semidirect product G=/32/16G={\mathbb{Z}}/32\rtimes{\mathbb{Z}}/16 where the standard generator of /16{\mathbb{Z}}/16 acts via φ\varphi. This is certainly a nonabelian metacyclic 22-group. Let a=([1],[0])a=([1],[0]) and b=([0],[1])b=([0],[1]). Then

(ab)16=([1+11++1115],[0])=([0],[0])(ab)^{16}=([1+11+\ldots+11^{15}],[0])=([0],[0])

where the last equality holds because

1+11++1115=2(1+11++117)=11815=0mod 321+11+\ldots+11^{15}=2(1+11+\ldots+11^{7})=\frac{11^{8}-1}{5}=0\ \mathrm{mod}\ 32

since φ\varphi has order eight. However, a16b16=a16=([16],[0])(ab)16a^{16}b^{16}=a^{16}=([16],[0])\neq(ab)^{16}, hence GG is not semi-abelian. Therefore Theorem 1.1 from [6] does not apply to GG. The action of 2/322{\mathbb{Z}}/32 preserves the subgroup 2/162{\mathbb{Z}}/16 and the subgroup generated by both is a normal subgroup NN of GG. It is easy to check that G/NG/N is elementary abelian of rank two, hence N=Φ(G)N=\Phi(G). But Φ(G)\Phi(G) is not abelian since the action of 2/162{\mathbb{Z}}/16 only fixes 4/324{\mathbb{Z}}/32. Therefore Proposition 1.1 does not apply. Note also that |G|=29|G|=2^{9} and |Φ(G)|=27|\Phi(G)|=2^{7}, so Corollary 1.4 does not apply either. Finally, since GG is a semidirect product, [G,G][G,G] is generated by elements of the form aφ(a)1a\varphi(a)^{-1} with a/32a\in{\mathbb{Z}}/32. Then [G,G]=2/32[G,G]=2{\mathbb{Z}}/32. We find the rest of terms in the lower central series in the same way:

{1}<16/32<8/32<4/32<2/32<G\{1\}<16{\mathbb{Z}}/32<8{\mathbb{Z}}/32<4{\mathbb{Z}}/32<2{\mathbb{Z}}/32<G

hence the nilpotency class of GG equals five and its coclass equals four. Therefore Theorem 1.3 in [5] and Propositions 3.1 and 3.2 do not apply to GG. On the other hand, groups of order 292^{9} can be shown to be S–groups using GAP (see [5]), but this has not been proved mathematically yet.

References

  • [1] A. Abdollahi, Finite pp-groups of class 2 have noninner automorphisms of order pp, J. Algebra 312 (2007), no. 2, 876–879. MR 2333188
  • [2] Alireza Abdollahi, Powerful pp-groups have non-inner automorphisms of order pp and some cohomology, J. Algebra 323 (2010), no. 3, 779–789. MR 2574864
  • [3] by same author, Cohomologically trivial modules over finite groups of prime power order, J. Algebra 342 (2011), 154–160. MR 2824534
  • [4] Alireza Abdollahi, Seyed Mohsen Ghoraishi, Yassine Guerboussa, Miloud Reguiat, and Bettina Wilkens, Noninner automorphisms of order pp for finite pp-groups of coclass 2, J. Group Theory 17 (2014), no. 2, 267–272. MR 3176660
  • [5] Alireza Abdollahi, Maria Guedri, and Yassine Guerboussa, Non-triviality of Tate cohomology for certain classes of finite pp-groups, Comm. Algebra 45 (2017), no. 12, 5188–5192. MR 3691357
  • [6] Mohammed T. Benmoussa and Yassine Guerboussa, Some properties of semi-abelian pp-groups, Bull. Aust. Math. Soc. 91 (2015), no. 1, 86–91. MR 3294262
  • [7] Yakov Berkovich, Groups of prime power order. Vol. 1, De Gruyter Expositions in Mathematics, vol. 46, Walter de Gruyter GmbH & Co. KG, Berlin, 2008, With a foreword by Zvonimir Janko. MR 2464640
  • [8] Kenneth S. Brown, Cohomology of groups, Graduate Texts in Mathematics, vol. 87, Springer-Verlag, New York-Berlin, 1982. MR 672956
  • [9] The GAP Group, GAP – Groups, Algorithms, and Programming, Version 4.12.1, 2022.
  • [10] Wolfgang Gaschütz, Nichtabelsche pp-Gruppen besitzen äussere pp-Automorphismen, J. Algebra 4 (1966), 1–2. MR 193144
  • [11] S. Mohsen Ghoraishi, On noninner automorphisms of finite pp-groups that fix the Frattini subgroup elementwise, J. Algebra Appl. 17 (2018), no. 7, 1850137, 8. MR 3813710
  • [12] E. I. Khukhro and V. D. Mazurov, Unsolved problems in group theory. the kourovka notebook, 2022.
  • [13] Derek J. S. Robinson, A course in the theory of groups, second ed., Graduate Texts in Mathematics, vol. 80, Springer-Verlag, New York, 1996. MR 1357169
  • [14] Marco Ruscitti, Leire Legarreta, and Manoj K. Yadav, Non-inner automorphisms of order pp in finite pp-groups of coclass 3, Monatsh. Math. 183 (2017), no. 4, 679–697. MR 3669785
  • [15] Peter Schmid, A cohomological property of regular pp-groups, Math. Z. 175 (1980), no. 1, 1–3. MR 595626
  • [16] Kôji Uchida, On Tannaka’s conjecture on the cohomologically trivial modules, Proc. Japan Acad. 41 (1965), 249–253. MR 183759
  • [17] Mingyao Xu and Qinhai Zhang, A classification of metacyclic 2-groups, Algebra Colloq. 13 (2006), no. 1, 25–34. MR 2188472