1 Introduction

Let the group \(G=AB\) be the product of two subgroups A and B, i.e. \(G=\{ab\mid a\in A, b\in B\}\). It was proved by N. Itô that the group G is metabelian if the subgroups A and B are abelian (see [1, Theorem 2.1.1]). This result laid the foundation for a systematic study of groups of the form \(G = AB\) with various conditions on the subgroups A and B. In particular, it follows directly from Itô’s result that every periodic group \(G=AB\) with abelian subgroups A and B is locally finite. It is also well-known that the group \(G=AB\) with cyclic subgroups A and B is supersoluble and abelian-by-finite ([1, Lemma 7.4.6]). Furthermore, a detailed description of the structure of the group \(G=AB\) with torsion-free locally cyclic subgroups A and B was obtained by the second author in [9].

The aim of this paper is to describe the structure of groups which are products of two locally cyclic subgroups in the periodic and the mixed case. Altogether this gives a complete answer to [1, Question 15].

Theorem 1.1

Let the periodic group \(G=AB\) be the product of two locally cyclic subgroups A and B. Then G contains uniquely determined locally cyclic normal subgroups S and T and a locally nilpotent subgroup \(H=A^*B^*\) with \(A^*\le A\) and \(B^*\le B\) such that

$$\begin{aligned} G=(S\times T)\rtimes H=(S\times A^*)(T\times B^*) \end{aligned}$$

where \(\pi (S)\cap \pi (A^*)=\pi (T)\cap \pi (B^*)= \emptyset \), \(S=[S,B^*],\) and \(T=[T,A^*]\).

We recall that a group G has finite Prüfer rank \(r=r(G)\) (special rank in the sense of Mal’cev in Russian terminology) if each finitely generated subgroup of G can be generated by r elements and r is the least positive integer with this property. Clearly a group is of rank 1 if and only if it is locally cyclic. It is also known that every finite p-group of the form \(G=AB\) with cyclic subgroups A and B has rank at most 2 for p odd [5, Satz 8] and at most 3 for \(p=2\) [6, Theorem 5.1]. The following consequence of Theorem 1.1 gives an exact upper bound for the Prüfer rank of the product \(G=AB\) of two periodic locally cyclic subgroups A and B.

Corollary 1.2

If \(G=AB\) is a periodic group with locally cyclic subgroups A and B, then the Prüfer rank of G does not exceed 3.

It should be noted that this result is also new for arbitrary finite groups of the form \(G=AB\) with cyclic subgroups A and B.

Finally, the following theorem extends a well-known result of B. Huppert on the supersolubility of finite groups which are products of pairwise permutable cyclic subgroups (see [5, Satz 34] or [4, Satz VI.10.3]). This gives, in particular, an affirmative answer to [2, Question 1].

Theorem 1.3

Let the group \(G=A_1A_2\cdots A_n\) be the product of finitely many pairwise permutable periodic locally cyclic subgroups \(A_1,\ldots ,A_n\). Then G is a periodic locally supersoluble group.

2 Preliminaries

In what follows \(G=AB\) is a group with locally cyclic subgroups A and B.

Lemma 2.1

If \(G=AB\) is an infinite p-group, then up to a permutation of the factors A and B the subgroup A is quasicyclic and one of the following statements hold:

  1. (1)

    \(G=A\times B\) with B cyclic or quasicyclic;

  2. (2)

    \(p=2\) and \(G=A\rtimes \langle b\rangle \) for some element \(b \in G\) with \(a^b=a^{-1}\) for all \(a\in A\);

  3. (3)

    \(p=2\) and \(G=A\langle b\rangle \) for some element \(b \in G\) with \(b^{2^n}=1\) for some \(n>1\), \(b^{2^{n-1}}\in A\) and \(a^b=a^{-1}\) for all \(a\in A\).

Proof

Clearly without loss of generality we may assume that the subgroup A is infinite and so quasicyclic. Then the subgroup B is either cyclic or quasicyclic. Since in the latter case the group G is abelian by [1, Lemma 7.4.4], the subgroup A is complemented in G and hence statement (1) holds.

Let the group G be non-abelian. Then the subgroup B is cyclic and so A as a quasicyclic p-subgroup of finite index in G must be normal and non-central in G. In particular, B induces on A a non-trivial cyclic p-group of automorphisms. On the other hand, since quasicyclic p-groups have no automorphisms of order \(p>2\), it follows that \(p=2\). But then \(B=\langle b\rangle \) with \(b^{2^n}=1\) for some \(n\ge 1\) and b induces on A an automorphism of order 2 that inverts the elements of A. In particular, if \(A\cap B=1\), we obtain statement (2). In the second case, \(A\cap B=\langle b^{2^{n-1}}\rangle \) and hence statement (3) holds, as claimed. \(\square \)

Corollary 2.2

If \(G=AB\) is a p-group and CD are subgroups of A and B, respectively, then \(CD=DC\) and so CD is a subgroup of G.

Proof

This is known if G is finite (see [5, Satz 3]), and follows from Lemma 2.1 in the general case. \(\square \)

Lemma 2.3

If the group \(G=AB\) is periodic and H is a finite subgroup of G, then H is contained in a finite subgroup E of G such that \(E=(A\cap E)(B\cap E)\). In particular, the group G is locally supersoluble.

Proof

Since the group H is finite, there exist finite subsets C of A and D of B such that H is contained in the set CD. Then the subgroups \(A_0=\langle C\rangle \), \(B_0=\langle D\rangle \), and \(\langle C,D\rangle \) are finite, because the group G is locally finite. Furthermore, it follows from [1, Lemma 1.2.3] that the normalizer \(N_G(\langle A_0, B_0\rangle )\) contains a finite subgroup E such that \(\langle A_0, B_0\rangle \le E=(A\cap E)(B\cap E)\). Since the subgroup E is supersoluble by [1, Lemma 7.4.6] and \(H\subseteq CD\subseteq \langle C,D\rangle = \langle A_0, B_0\rangle \), the lemma is proved. \(\square \)

As a direct consequence of this lemma, we have

Corollary 2.4

If the group \(G=AB\) is periodic, then there exists an ascending series of finite subgroups \(1=G_0<G_1<\cdots<G_n<\cdots G\) such that \(G_n=(A\cap G_n)(B\cap G_n)\) for each \(n>0\) and \(G=\bigcup _{n=1}^\infty G_n\).

If G is a periodic group and \(\pi \) is a set of primes, then a subgroup H of G is called a \(\pi \)-subgroup provided that all prime divisors of the order of any element of H are contained in \(\pi \). By a Sylow \(\pi \)-subgroup of G we simply mean a maximal \(\pi \)-subgroup \(G_{\pi }\) of G which will be denoted by \(G_p\) if \(\pi =\{p\}\).

Lemma 2.5

Let \(G=AB\) be a periodic group and \(\pi \) a set of primes. Then the following statements hold.

  1. (1)

    If \(A_{\pi }\) and \(B_{\pi }\) are Sylow \(\pi \)-subgroups of A and B, respectively, then \(G_{\pi }=A_{\pi }B_{\pi }\) is a Sylow \(\pi \)-subgroup of G and

    $$N_G(G_{\pi })=N_A(G_{\pi })N_B(G_{\pi }).$$
  2. (2)

    If pq are primes with \(p>q\), then a Sylow p-subgroup \(G_p\) is normalized by a Sylow q-subgroup \(G_q\). In particular, \(G_{\{p, q\}}=G_pG_q\) for any primes p and q.

Proof

(1) It follows from [1, Lemma 1.3.2] that in the notation of Corollary 2.4 for each \(n\ge 1\) the set \((A_{\pi }\cap G_n)(B_{\pi }\cap G_n)\) is a Hall \({\pi }\)-subgroup of \(G_n\). Therefore \(G_{\pi }=\bigcup _{n=1}^\infty (A_{\pi }\cap G_n)(B_{\pi }\cap G_n)\) is a Sylow \(\pi \)-subgroup of G. Since \(A_{\pi }=\bigcup _{n=1}^\infty (A_{\pi }\cap G_n)\) and \(B_{\pi }=\bigcup _{n=1}^\infty (B_{\pi }\cap G_n)\), this implies \(G_{\pi }=A_{\pi }B_{\pi }\). In addition, applying [1, Lemma 1.2.2], we have \(N_G(G_{\pi })= N_A(G_{\pi })N_B(G_{\pi })\).

(2) If \({\pi }=\{p, q\}\) for some primes \(p>q\), then \(A_{\{p,q\}}=A_p\times A_q\), \(B_{\{p,q\}}=B_p\times B_q\), and \(G_{\{p,q\}}=(A_p\times A_q)(B_p\times B_q)\). As G and so its subgroup \(G_{\{p,q\}}\) is locally supersoluble by Lemma 2.3, the Sylow p-subgroup \(G_p=A_pB_p\) is normal in \(G_{\{p,q\}}\). Therefore, \(G_{\{p,q\}}=(A_p\times A_q)(B_p\times B_q)=G_pA_qB_q=G_pG_q\), as claimed. \(\square \)

A Sylow basis of a periodic group G is defined to be a complete set \(\mathbf{S} =\{G_p\}\) of Sylow p-subgroups of G, one for each prime p, such that \(G_pG_q=G_qG_p\) for all pairs pq of primes, and \(G_{\pi }=\langle G_p \mid p\in {\pi }\rangle \) is a Sylow \(\pi \)-subgroup of G for each set \(\pi \) of primes. As is well-known (see [3, Lemma 2.1]), every countable periodic locally soluble group possesses Sylow bases. The basis normalizer \(N_G(\mathbf{S})\) of a Sylow basis \(\mathbf{S}\) of G is by definition the intersection \(N_G(\mathbf{S})=\bigcap _p N_G(G_p)\) of the normalizers \(N_G(G_p)\) of the Sylow p-subgroups \(G_p\) of \(\mathbf{S}\) for all p.

Lemma 2.6

Let the group \(G=AB\) be periodic and \(G_p=A_pB_p\) for each prime p. Then \(\mathbf{S}=\{G_p\}\) is a Sylow basis of G. Moreover, if \(A^*=\bigcap _p N_A(G_p)\) and \(B^*=\bigcap _p N_B(G_p)\), then \(N_G(\mathbf{S})=A^*B^*\).

Proof

Indeed, by Lemma 2.5, the set \(\mathbf{S}=\{G_p\}\) forms a Sylow basis of G and \(N_G(G_p)= N_A(G_p)N_B(G_p)\) for every p by [1, Lemma 1.2.2]. Therefore, \(N_G(\mathbf{S})=\bigcap _p N_G(G_p)=\bigcap _p N_A(G_p)N_B(G_p)\) and it is easy to check that \(\bigcap _p N_A(G_p)N_B(G_p)=(\bigcap _p N_A(G_p))(\bigcap _p N_B(G_p))=A^*B^*\) (see [1, Lemma 1.1.2]). Therefore \(N_G(\mathbf{S})=A^*B^*\). \(\square \)

The following lemma is a direct consequence of a well-known result of L. Kovacs (see [7, Theorem 2]).

Lemma 2.7

Let G be a finite soluble group, \(\pi \) a set of primes and H a Hall \(\pi \)-subgroup of G. If for each \(p\in \pi \) the Prüfer rank of a Sylow p-subgroup of G does not exceed r, then H is a subgroup of rank at most \(r+1\).

Proof

Indeed, it is obvious that if K is a subgroup of H, then every Sylow subgroup of K is generated by r elements. Therefore, K can be generated by \(r+1\) elements by the result of Kovacs cited above. Thus every subgroup of H is generated by \(r+1\) elements and so H has rank at most \(r+1\). \(\square \)

3 Proof of Theorem 1.1

First of all, it follows from Lemma 2.1 that for each prime p every Sylow p-subgroup of \(G=AB\) satisfies the minimal condition for subgroups. Therefore, G satisfies the minimal condition for p-subgroups for all primes p. Since the group G is metabelian by Ito’s theorem, the locally nilpotent residual R of G is contained in its derived subgroup \(G'\) and so it is abelian. It was proved by Hartley [3, Theorem 1] that in this case \(G=R\rtimes H\), where H is any basis normalizer of G. In particular, by Lemma 2.6, we can take \(H=A^*B^*\).

It is easy to see that the subgroup H is locally nilpotent and contains the center Z(G) of G. Furthermore, \(G'=R\times H'\) and so \(H'\) is a normal subgroup of G. Since R is abelian and \(N_G(H)= N_R(H)\times H\), it follows that \(N_R(H)\le Z(G)\le H\). Therefore, \(N_R(H)=1\) and hence \(H=N_G(H)\). We now show that the subgroup \(H=A^*B^*\) commutes with both subgroups A and B.

Indeed, put \(S=R\cap \langle A,B^*\rangle \) and \(T=R\cap \langle A^*,B\rangle \). It is clear that S and T are normal subgroups of G, \(\langle A,B^*\rangle =S\rtimes H\), and \(\langle A^*,B\rangle =T\rtimes H\). On the other hand, as \(G=AB\), we have also \(\langle A,B^*\rangle =AB_1\) and \(\langle A^*,B\rangle =A_1B\) for some subgroups \(A_1\) and \(B_1\) such that \(A^*\le A_1\le A\) and \(B^*\le B_1\le B\). From here, we deduce \(AB_1\cap A_1B=A_1B_1=(S\cap T)\rtimes H\). Moreover, passing to the factor group \(G/H'\), we may restrict ourselves to the case when the subgroup \(H=A^*B^*\) is abelian. Then the subgroups \(A^*\) and \(B^*\) centralize S and T, respectively, and so the subgroup H centralizes the intersection \(S\cap T\). Since \(H=N_G(H)\), this implies \(S\cap T=1\). Thus \(A_1B_1=H=A^*B^*\) and hence \(\langle A,B^*\rangle =AH=AB^*\) and \(\langle A^*,B\rangle =BH=A^*B\), as asserted.

Further, taking into account the equalities \(AB^*=S\rtimes H\) and \(H=A^*B^*\), we conclude that the subgroup \(A^*\) centralizes S, because \([A^*,S]\le H'\cap S=1\). Since in this case the normalizer \(N_S(B^*)\) is contained in \(N_G(H)=H\), we have \(N_S(B^*)=1\). Therefore, every element \(b\in B^*\) induces on S an automorphism leaving only the identity element fixed. But then every element of S can be written in the form \(b^{-1}s^{-1}bs\) with \(s\in S\) and hence \(S=[B^*,S]\). Similarly, using the equality \(A^*B=T\rtimes H\), we derive \(T=[A^*,T]\).

Finally, we put \(A_0=A\cap BS\) and \(B_0=AT\cap B\). Clearly from the equalities \(G=AB\), \(AB^*=S\rtimes H\), and \(A^*B=T\rtimes H\), it follows that \(G=S\rtimes A^*B=T\rtimes AB^*\), \(A=A^*\times A_0\), and \(B=B^*\times B_0\). Therefore, \(S\rtimes B=A_0B\) and \(T\rtimes A=AB_0\). Furthermore, if \(S_p\) is a Sylow p-subgroup of S, then \(S_p\rtimes B=(A_0\cap S_pB)B\) and if \(S_p\ne 1\), then \(A_0\cap S_pB\ne 1\). Since the subgroup \(A_0\) is locally cyclic, this implies that S is locally cyclic and \(\pi (S)\) is contained in \(\pi (A_0)\). Moreover, as \(A^*\) and \(A_0\) are subgroups of the locally cyclic subgroup A, it also follows that \(\pi (A^*)\cap \pi (A_0)= \emptyset \). Similarly, using the equality \(T\rtimes A=AB_0\), we obtain \(\pi (T)=\pi (B_0)\) and \(\pi (B^*)\cap \pi (B_0)=\emptyset \). \(\square \)

Proof of Corollary 1.2

By Corollary 2.4, we may restrict ourselves to the case in which the group \(G=AB\) is finite. By Theorem 1.1, G then contains cyclic normal subgroups S and T and a nilpotent subgroup \(H=A^*B^*\) with \(A^*\le A\) and \(B^*\le B\) such that

$$\begin{aligned} G=(S\times T)\rtimes H=(S\times A^*)(T\times B^*), \end{aligned}$$

where \(\pi (S)\cap \pi (A^*)=\pi (T)\cap \pi (B^*)= \emptyset \), \(S=[S,B^*]\), and \(T=[T,A^*]\). In particular, if for some prime p the subgroup H contains a non-cyclic Sylow p-subgroup P, then S and T are \(p'\)-subgroups of G.

Since \(P=A_pB_p\) with \(A_p=A\cap P\) and \(B_p=B\cap P\), both subgroups \(A_p\) and \(B_p\) are non-trivial and so \(p\notin \pi (S)\cup \pi (T)\). Therefore, if \(G_p\) is a non-cyclic Sylow p-subgroup of G, \(S_p=G_p\cap S\), and \(T_p=G_p\cap T\), then up to conjugation \(G_p\) coincides with one of the following subgroups of G: \(P=A_pB_p\), \(T_p\rtimes A_p \), \(S_p\rtimes B_p\), and \(S_p\rtimes T_p\). In particular, the Sylow p-subgroups of G have rank at most 2 for \(p>2\) (see [4, Satz III.11.5]) and at most 3 for \(p=2\) (see [6, Theorem 5.1]). We now show that G is in fact a group of rank at most 3.

Indeed, suppose the contrary and let the group G contain a subgroup K whose minimal number of generators d(K) is at least 4. Since the Sylow subgroups of odd orders in K have rank at most 2 by what was noted above, each Sylow 2-subgroup Q of K must have rank 3 by Lemma 2.7. It is clear that \(Q=K\cap P\) for a Sylow 2-subgroup \(P=A_2B_2\) of G. As the group G is metabelian, the derived subgroup \(P'\) is abelian and normal in G. Moreover, \(P'\) has rank at most 2 by [6, Theorems 4.2 and 4.3(e)].

Put \(N=P'\cap Q\). As \(Q=K\cap P\), we have \(N=K\cap P'\) and so N is an abelian normal subgroup of K with rank at most 2. In addition, \(N\ne 1\), because otherwise the subgroup Q is embedded in the factor group \(P/P'\) whose rank is equal to 2. On the other hand, since the factor group \(Q/N=Q/P'\cap Q\) is isomorphic to the factor group \(QP'/P'\le P/P'\), it is abelian of rank at most 2. Therefore, the Sylow 2-subgroups of the factor group K/N have rank at most 2 and hence K/N has rank at most 3 by Lemma 2.7. In particular, \(d(K/N)<d(K)=4\) and so N is not contained in the Frattini subgroup \(\Phi (K)\) of K, because otherwise \(d(K/N)=d(K)\). Clearly, passing to the factor group \(K/\Phi (K)\), we may assume that \(\Phi (K)=1\). Then the normal subgroup N is complemented in K and so in Q by [4, Hilfsatz 3.4.4]. Moreover, since \(\Phi (N)=1\) by [4, Hilfsatz 3.3.b], the subgroup N is elementary abelian of order at most 4.

Let L be a complement to N in K and \(M=Q\cap L\). Then M is a Sylow 2-subgroup of L and \(Q=N\times M\), so that M is abelian with \(d(M)\le 2\). Since the subgroup K is supersoluble, its maximal subgroup U of odd order is normal in K and centralizes N. Therefore, \(K=U\times Q=(U\times N)\rtimes M\) and hence the centralizer \(C_N(M)\) is a non-trivial central subgroup of K. As \(\Phi (K)=1\), the subgroup \(C_N(M)\) is complemented in K and thus in Q. From this, it follows that \(Q=M\times N\) is abelian and N is a central subgroup of K. Therefore, \(K=(U\rtimes M)\times N\) and the subgroup \(U\rtimes M\) is three-generated by Lemma 2.7. This means that there exist elements uvw of U and xyz of M such that \(U\rtimes M=\langle ux, vy, wz\rangle \). Then M modulo U is generated by xyz. In particular, if \(d(M)=1\), without loss of generality we may assume that \(M=\langle x\rangle \) and \(y=z=1\), so that \(U\rtimes M=\langle ux, v, w\rangle \). In the case \(d(M)=2\) we can take \(M=\langle x, y\rangle \) and \(z=1\). Then \(U\rtimes M=\langle ux, vy, w\rangle \).

Finally, since \(Q=M\times N\) is a 2-subgroup of rank 3 as noted above, only two cases are possible: either \(M=\langle x\rangle \) and \(N=\langle a, b\rangle \) has order 4 or \(M=\langle x, y\rangle \) with \(x\ne 1\ne y\) and \(N=\langle a\rangle \) is of order 2. Therefore, \(K=\langle ux, av, bw \rangle \) in the first case and \(K=\langle ux, vy, aw \rangle \) in the second case. In both cases \(d(K) < 4\) and this contradiction completes the proof. \(\square \)

4 Products of a periodic and a torsion-free locally cyclic group

Recall that a group G has finite torsion-free rank if it has a series of finite length whose factors are either periodic or infinite cyclic. The number \(r_0(G)\) of infinite cyclic factors in such a series is an invariant of G called its torsion-free rank. In this section, we describe the structure of the group \( G = AB \) with locally cyclic subgroups A and B, the first of which is periodic and the other non-trivial torsion-free. Clearly \(r_0(B)=1\) and we note first that \(r_0(G)=1\).

Lemma 4.1

Let \(G=AB\) be a group with subgroups A and B such that A is periodic abelian and B is non-trivial torsion-free locally cyclic. Then \(r_0(G)=1\).

Proof

It was proved by Zaitsev [11, Theorem 3.7] (see also [1, Lemma 7.1.2]) that there exists a non-trivial normal subgroup of G contained in A or B. Therefore G has the normal series \(A_0< A_0B_0 < G\) in which \(A_0\) is the core of A in G and \(B_0\) is the core of B in G modulo \(A_0\). As is easily seen, the factors \(A_0\) and \(G/A_0B_0\) are periodic and the factor group \(A_0B_0/A_0\) is isomorphic to \(B_0\). Thus \(r_0(G)=r_0(B)=1\), as claimed. \(\square \)

The following lemma is a consequence of the well-known theorem of I. Schur on the finiteness of the derived subgroup of a group that is finite over its center (see [8, Corollary to Theorem 4.12]).

Lemma 4.2

If a group G contains a central subgroup Z such that the factor group G/Z is locally finite, then the derived subgroup of G is locally finite.

Theorem 4.3

Let the group \(G=AB\) be the product of two locally cyclic subgroups A and B such that A is periodic and B is non-trivial torsion-free. Then one of the following statements holds.

  1. (1)

    The subgroup A is normal in G and so \(G=A\rtimes B\);

  2. (2)

    \(A=A_1\langle a\rangle \) with \(a^2\in A_1\), the subgroup \(A_1\) is normal in G and \(G=(A_1\rtimes B)\langle a\rangle \) with \(b^a=b^{-1}\phi (b)\) for all \(b\in B\), where \(\phi : B \rightarrow A_1\) is a derivation of B into \(A_1\).

Proof

It is easy to see that each periodic normal subgroup H of G is contained in A, because \(AH=A(AH\cap B)\) and \(AH\cap B=1\). Therefore the core \(A_1=\cap _{g\in G}A^g\) of A in G is the maximal periodic normal subgroup of G.

Assume first that \(A_1=1\) and let \(B_1\) be the core of B in G. Then \(B_1\ne 1\) by the theorem of Zaitsev noted above and so the factor group \(G/B_1\) is periodic, because it is the product of two periodic subgroups \(AB_1/B_1\) and \(B/B_1\). Moreover, since the centralizer \(C_G(B_1)\) of \(B_1\) in G contains B, the group G induces on \(B_1\) a periodic group of automorphisms which is isomorphic to the factor group \(A/C_A(B_1)\). As is well-known, a periodic group of automorphisms of any locally cyclic torsion-free group is of order 2. Therefore the order of \(A/C_A(B_1)\) does not exceed 2 and hence either \(A=C_A(B_1)\) or \(A=C_A(B_1)\langle a\rangle \) with \(a\in A\) and \(a^2\in C_A(B_1)\).

On the other hand, since the centralizer \(C_G(B_1)=C_A(B_1)B\) is normal in G and periodic over \(B_1\), its derived subgroup \(C_G(B_1)'\) is periodic by Lemma 4.2 and normal in G. Therefore \(C_G(B_1)'\le A_1=1\) and hence \(C_G(B_1)=C_A(B_1)\times B\). But then again \(C_A(B_1)\) is normal in G and so \(C_A(B_1)=1\). Thus in the case \(A_1=1\) we have either \(A=1\) and \(G=B\) or \(A=\langle a\rangle \) with \(a^2=1\) and \(G=B\rtimes \langle a\rangle \) with \(b^a=b^{-1}\) for all \(b\in B\).

Finally, returning now to the general case, we derive that either \(G=A\rtimes B\) or \(G=(A_1\times B)\langle a\rangle )\) with \(b^a=\phi (b)b^{-1}\) for every \(b\in B\) and some element \(\phi (b)\in A_1\). Moreover, since \(\phi (bc)(bc)^{-1}=(bc)^a =b^ac^a=(\phi (b)b^{-1})(\phi (c)c^{-1})=(\phi (b)\phi (c)^b)(bc)^{-1}\), it follows that \(\phi (bc)=\phi (b)\phi (c)^b\) for any \(b,c\in B\). The latter means in particular that the mapping \(\phi : B \rightarrow A_1\) is a derivation of B into \(A_1\), as claimed. \(\square \)

5 Products of finitely many periodic locally cyclic groups

A well-known theorem of Huppert cited in the introduction says that every finite group of the form \(G=A_1A_2\cdots A_n\) with pairwise permuting cyclic subgroups \(A_i\) for \(1\le i\le n\) is supersoluble. This result was later extended to products of pairwise permutable locally cyclic Chernikov groups by Tomkinson [10]. He proved that in this case \(G=A_1A_2,\ldots ,A_n\) is a locally supersoluble Chernikov group. In this section, we generalize this result to products of arbitrary periodic locally cyclic groups. Recall that a group is said to be hyperabelian (respectively, hypercyclic) if it has an ascending series of normal subgroups with abelian (respectively cyclic) factors.

Lemma 5.1

Let \(G=A_1A_2\cdots A_n\) be the product of pairwise permutable periodic locally cyclic subgroups \(A_i\). If the set \(\pi = \bigcup _{i=1}^n\pi (A_i)\) is finite, p is the largest prime in \(\pi \), \(P_i\) is the Sylow p-subgroup of \(A_i\), and \(Q_i\) is the p-complement to \(P_i\) in \(A_i\) for each \(1\le i\le n\), then G is a \(\pi \)-group, \(P=P_1P_2\cdots P_n\) is a normal Sylow p-subgroup of G, and \(Q=Q_1Q_2\cdots Q_n\) is a p-complement to P in G.

Proof

Since each of the \(A_i\) is a subgroup of Prüfer rank 1, the group \(G=A_1A_2\cdots A_n\) is hyperabelian of finite Prüfer rank by [2, Theorem 3.1]. Therefore, arguing by induction on n and applying [1, Corollary 3.2.7], and [2, Lemma 3.2], we derive that G is a \(\pi \)-group, \(P=P_1P_2\cdots P_n\) is a Sylow p-subgroup of G and \(Q=Q_1Q_2\cdots Q_n\) is a complement to P in G. Moreover, taking into account that the subgroups \(A_iA_j\) are locally supersoluble by Lemma 2.3, we conclude that \({P_i}^{A_j}\le P\) for all ij and so P is a normal subgroup of G. \(\square \)

Lemma 5.2

Let \(G=A_1A_2\cdots A_n\) be the product of pairwise permutable locally cyclic subgroups \(A_i\). If the group G is periodic and the set \(\pi (G)\) is finite, then G is locally supersoluble.

Proof

Since \(\pi (G)\) is finite, every locally cyclic subgroup \(A_i\) is a Chernikov group, i.e. a finite extension of a direct product of finitely many quasicyclic subgroups. Therefore G is a locally supersoluble Chernikov group by [10, Theorem B]. \(\square \)

Proof of Theorem 1.3

Let \(G=A_1A_2\cdots A_n\) be the product of pairwise permutable periodic locally cyclic subgroups \(A_i\). Then G is a periodic group by Lemma 5.1. If the set \(\pi (G)\) is finite, then the group G is locally supersoluble hypercyclic by Lemma 5.2. In the other case the set \(\pi (G)\) is infinite and thus it can be presented as a union \(\pi (G)=\bigcup _{i=1}^\infty \pi _{i}\) of finite subsets \(\pi _{i}\) such that \(\pi _{i}\subset \pi _{i+1}\) for all \(i\ge 1\). Let \(P_{ij}\) be the Sylow \(\pi _{i}\)-subgroup of \(A_j\) for \(1\le j\le n\) and \(G_i=P_{i1}P_{i2}\cdots P_{in}\). Then \(G_i\) is a Sylow \(\pi _{i}\)-subgroup of G by Lemma 5.1 which is locally supersoluble as a group for each \(i\ge 1\) by Lemma 5.2. Since \(G=\bigcup _{i=1}^\infty G_i\), the group G is also locally supersoluble, as claimed. \(\square \)