Filters: Topological congruence relations on groups2000 Mathematics Subject Classification: 22A05 54A20 (22C05 20K45 54H11)

Gábor Lukács I gratefully acknowledge the generous financial support received from the Alexander von Humboldt Foundation, the Killam Trusts, and Dalhousie University that enabled me to do this research.

November 27, 2006

Abstract
A filter   on a group G   is a T   -filter if there is a Hausdorff group topology τ   on G   such that τ 0   . This notion can be specialized for sequences, in which case we say that { a n }   is a T   -sequence. In this paper, T   -filters and T   -sequences are studied. We characterize T   -filters in non-abelian groups, show that certain filters can be interpreted as topological extensions of the notion of kernel (i.e., normal subgroup, congruence relation), and provide several sufficient conditions for a sequence in an abelian group to be a T   -sequence. As an application, special sequences in the Prüfer groups Z ( p )   are investigated. We prove that for p 2   , there is a Hausdorff group topology τ   on Z ( p )   that is neither maximally nor minimally almost periodic—in other words, the von Neumann radical n ( Z ( p ) , τ )   is a non-trivial finite subgroup. In particular, n ( n ( Z ( p ) , τ ) ) n ( Z ( p ) , τ )   .

1 INTRODUCTION

Given a set X   , a point x 0 X   , and a filter   on X   such that { x 0 }   , one can easily construct a Hausdorff topology τ   on X   such that τ x 0   . This is, however, not the case for groups and group topologies, as Example  1.2 below demonstrates.
Following Zelenyuk and Protasov [?, who were the first to investigate this type of question, we say that a filter   on a group G   is a T   -filter if there is a Hausdorff group topology τ   on G   such that τ 0   . In this case, the group G   equipped with the finest group topology with this property is denoted by G { }   . This notion can be specialized for sequences, in which case we say that { a n }   is a T   -sequence, and write G { a n }   for G   equipped with the above mentioned finest group topology.
Example 1.1. For a prime number p   , let A = Z ( p )   be the Prüfer group. It can be seen as the subgroup of Q / Z   generated by the elements of p   -power order, or the group formed by all p n   th roots of unity in C   . For e n = 1 p n   , { e n }   is clearly a T   -sequence in A   , because e n 0   in the subgroup topology that A   inherits from Q / Z   .
Example 1.2. Keeping the notations of Example  1.1 , set b n = 1 p + 1 p n = e 1 + e n   . If b n τ 0   for some group topology τ   , then also e n 1 = p b n τ 0   , and therefore, e 1 = p b n + 1 b n τ 0   .
Hence, τ   cannot be Hausdorff, and so { b n }   is not a T   -sequence in A   .
Zelenyuk and Protasov characterized T   -filters and T   -sequences in abelian groups [?,Thm. 1,Thm. 2, and studied the topological properties of A { a n }   , where { a n }   is a T   -sequence in the abelian group A   . Their work is, undoubtedly, of a ground-breaking nature. Nevertheless, it may leave one with a certain feeling of incompleteness, for two reasons: First, the universal property that defines A { }   is mentioned only tangentially; second, in the characterization of T   -filters, the condition of commutativity seems to be rather artificial.
Every topological group G   admits a “largest” compact Hausdorff group b G   and a continuous homomorphism ρ G : G b G   such that every continuous homomorphism φ : G K   into a compact Hausdorff group K   factors uniquely through ρ G   :

diagram

The group b G   is called the Bohr-compactification of G   . The image ρ G ( G )   is dense in b G   .
The kernel of ρ G   is called the von Neumann radical of G   , and denoted by n ( G )   . One says that G   is maximally almost periodic if n ( G ) = 1   , and minimally almost periodic if n ( G ) = G   (cf. [?). Let A ^ = ( A , T )   be the Pontryagin dual of A   —in other words, the group of continuous characters of A   (i.e., continuous homomorphisms χ : A T   , where T = R / Z   ), equipped with the compact-open topology. It follows from the famous Peter-Weyl Theorem ([?,Thm. 33) that the Bohr-compactification of an abelian group A   can be quite easily computed: b A = A ^ d ^   , where A ^ d   stands for the group A ^   with the discrete topology. Thus, n ( A ) = χ A ^ ker χ   for every abelian topological group A   .
T   -sequences turn out to be a very useful tool for constructing “pathological” examples. For example, Zelenyuk and Protasov used T   -sequences to show (independently of Ajtai, Havas andKomlós [?) that every infinite abelian group admits a non-maximally almost periodic Hausdorff group topology (cf. [?,Thm. 16). There are plenty of examples of minimally (or maximally) almost periodic groups (cf. [?, [?). Nevertheless, it appears that no example is known for a Hausdorff topological group G   whose von Neumann radical n ( G )   is non-trivial and finite. This raises the following question:
Problem I. How to characterize Hausdorff topological groups G   such that n ( G )   is finite and non-trivial?
If n ( G )   is finite and non-trivial, then n ( n ( G ) ) = 1   , and thus n ( n ( G ) ) n ( G )   , which leads to a second problem:
Problem II. How to characterize Hausdorff topological groups G   such that n ( n ( G ) ) n ( G )   ?
Note that since n   is productive, one may focus on topologically indecomposable groups (i.e., that cannot be expressed as G 1 × G 2   for G 1 , G 2 { e }   ).
Our goal in this paper is two-fold, namely, to fill the void left by Zelenyuk and Protasov in the context of T   -filters on the one hand, and to present ample non-isomorphic algebraically directly indecomposable abelian Hausdorff topological groups whose von Neumann radical is finite and non-trivial on the other hand. The latter also shows that Problems  I and  II are meaningful.
These aims are carried out according to the following structure: In section  2 , we characterize T   -filters in non-abelian groups (Theorem  2.2 ), and show that filters can be interpreted as a topological extension of the notion of kernel (i.e., normal subgroup, congruence relation). In section  3 , several results that provide sufficient conditions for a sequence in an abelian group to be a T   -sequence are presented (Theorem  3.2 ). Finally, in section  4 , special sequences of Prüfer groups are investigated (Theorem  4.2 ), and we prove that for p 2   , Z ( p )   admits a neither maximally nor minimally almost periodic Hausdorff group topology τ   (Theorem  4.4 ). In particular, n ( Z ( p ) , τ )   is finite, and n ( n ( Z ( p ) , τ ) ) n ( Z ( p ) , τ )   .

2 T-FILTERS IN NON-ABELIAN GROUPS

In this section, we develop the non-abelian analogue of one of Zelenyuk and Protasov's results [?,Thm. 1. Thus, group operations are written multiplicatively.
Let G   be a group. For a group topology τ   on G   , N τ   denotes the neighborhood filter of e   in τ   ; for X G   , put X ˙ = { F G | X F }   . For g G   and F G   , one sets F g = { g 1 x g | x F }   , and for a filter   on G   we put g = { F g | F }   and 1 = { F 1 | F }   . We say that   is a normal filter on G   if g   for every g G   ;   is symmetric if 1 =   and e F   for every F   . The normal closure of   is defined as = g G g   . For instance, N τ   is normal for every group topology τ   on G   . Our definition of normal closure might remind one more of the normal core. However, as the lemma below reveals, this is not the case.
Lemma 2.1. Let   be a filter on a group G   .
  • (a) For any group topology τ   on G   , τ e   if and only if τ e   .
  • (b)   is a T   -filter if and only if   is so.
  • (c)   is a T   -filter if and only if it contains a normal T   -filter.
  • (d) s = 1 e ˙   is symmetric, and for any group topology τ   on G   , τ e   if and only if s τ e   .
  • (e) If H   is a subgroup of G   , then ( H ˙ ) = X ˙   , where X = g G H g   ; in particular, the normal closure of H   is the smallest subgroup in ( H ˙ )   .
  • Proof. Suppose that τ e   . Then N τ   , and so N τ = N τ g g   for every g G   , and hence N τ   , which shows that τ e   . The converse is obvious. This shows (a); (b) follows from (a), and (c) follows from (b). Since H X g = X   , certainly X ˙ ( H ˙ )   . On the other hand, if F ( H ˙ )   , then H g F   for every g G   , and thus X F   . This shows (e), and since (d) is similar to (a), it completes the proof.
Next, we introduce the non-abelian analogue of the Zelenyuk-Protasov notation. For a countable family { F n } n = 1   of subsets of G   , we put
[ F 1 , , F n ] = σ S n F σ ( 1 ) F σ ( n ) , (1)
[ { F n } ] = n = 1 [ F 1 , , F n ] where S n stands for the symmetric group on n elements. For a filter on G , set = { [ { F n } ] | { F n } n = 1 } . (2)
Theorem  2.2 is the non-abelian analogue of [?,Thm. 1. Note that Lemma  2.1 enables us to confine our attention to normal symmetric filters in what follows.
Theorem 2.2. Let   be a symmetric normal filter on G   . Then:
  • (a)   is a base at e   for a group topology τ   on G   , and τ e   ;
  • (b) τ τ   for every group topology τ   on G   such that τ e   ; in other words, τ   is the finest group topology on G   with this property;
  • (c) A homomorphism f : G L   into a topological group L   is continuous with respect to τ   if and only if f ( ) e   in L   ;
  • (d) the following statements are equivalent:
    • (i)   is a T   -filter;
    • (ii) τ   is Hausdorff;
    • (iii) = { e }   ;
  • (e) For every M G   , τ / M = τ q ( )   , where q : G G / M   is the canonical projection;
  • (f) There is a smallest normal subgroup N G   such that the image ¯   of   in G / N   is a T   -filter.
In order to prove Theorem  2.2 , a technical lemma is required.
Lemma 2.3. Let { V i } i = 1   be a family of subsets of the group G   such that V i + 1 3 V i   . For every n N   and every σ S n   , one has V σ ( 1 ) V σ ( n ) V 1 3   .
  • Proof. We proceed by induction on n   . For n 3   , the statement is trivial. So, suppose that the statement is true for every product of less than n   sets. Set l = σ 1 ( 1 )   , order the numbers σ ( 1 ) , , σ ( l 1 )   by size as j 1 < < j l 1   , where τ S l 1   is such that j τ ( k ) = σ ( k )   . Since j i < j i + 1   , one has V j i + 1 3 V j i + 1 V j i   . Thus, V σ ( 1 ) V σ ( l 1 ) = V j τ ( 1 ) V j τ ( l 1 ) V j 1 3   by the inductive hypothesis ( l 1 < n   ). The index j 1   cannot be 1   , because σ ( l ) = 1   , so j 1 2   , and thus V j 1 3 V 1   . Therefore, V σ ( 1 ) V σ ( l 1 ) V 1   , and similarly, V σ ( l + 1 ) V σ ( n ) V 1   . Hence, ( V σ ( 1 ) V σ ( l 1 ) ) V σ ( l ) ( V σ ( l + 1 ) V σ ( n ) ) V 1 V 1 V 1 = V 1 3   , as desired.
  • Proof of Theorem  2.2 . (a) Let [ { F n } ] , [ { T n } ]   and g G   .
    1. Since   is a filter, F n T n   for every n N   . Thus, [ { F n T n } ]   , and evidently, [ { F n T n } ] [ { F n } ] [ { T n } ]   .
    2. Since   is normal, F n g   for every n N   . Hence, g 1 [ { F n } ] g = [ { F n g } ]   .
    3. Since   is symmetric, F n 1   for every n N   , and so [ { F n } ] 1 = [ { F n 1 } ]   .
    4. Obviously, [ { F 2 n 1 } ] , [ { F 2 n } ]   , and [ { F 2 n 1 } ] [ { F 2 n } ] [ { F n } ]   .
    Therefore,   is a base at e   for a group topology; τ e   follows from   .
    (b) Let U N τ   . Since τ   is a group topology, there exists V 1 N τ   such that V 1 3 U   .
    Inductively, one can find V n N τ   such that V n + 1 3 V n   for every n N   . Since τ e   , there exists F n   such that F n V n   for every n N   . By Lemma  2.3 ,
    F σ ( 1 ) F σ ( n ) V σ ( 1 ) V σ ( n ) V 1 3 U (3)
    for every σ S n   and every n N   , and therefore [ { F n } ] U   . Hence, τ τ   .
    (c) Put τ f = { f 1 ( O ) | O open in L }   , the initial group topology that the homomorphism f   induces on G   . If f ( ) e   in L   , then τ f e   . Therefore, by virtue of (a), τ f τ   , and hence f   is continuous. The converse is trivial.
    (d) (i)   (ii): If   is a T   -filter, then there is a group topology τ   on G   such that τ e   . By (b), τ τ   , and thus τ   is Hausdorff. Conversely, if τ   is Hausdorff, then   is a T   -filter by (a). The equivalence (ii)   (iii) is obvious, because   is the base at e   for the group topology τ   .
    (e) On the one hand, by (b), τ / M τ q ( )   because q ( ) τ / M e   . On the other hand, by (c), the projection ( G , τ ) ( G / M , τ q ( ) )   is continuous, and therefore τ q ( ) τ / M   . (f) Put N   to be the closure of { e }   in τ   ; equivalently, N =   ; being the closure of a normal subgroup, N G   . Suppose that M G   is such that the image q ( )   of   under the projection q : G G / M   is a T   -filter. Then τ q ( )   is Hausdorff, and by (c), q   is continuous with respect to τ   .
    Therefore, q ( N ) = q ( { e } ¯ τ ) q ( { e } ) ¯ τ q ( ) = { e } ¯ τ q ( ) = { e }   , and hence N M   . On the other hand, ( G / N , τ / N )   is the maximal Hausdorff quotient of ( G , τ )   , and by (e), τ / N = τ ¯   .
    This shows that ¯   is a T   -filter (by (d)).
Zelenyuk and Protasov defined the group A { } = ( A , τ )   only when A   is an abelian group and   is a T   -filter on it. In our view, this definition ignores the essence hidden underneath. For a filter   on G   , we set G { }   to be the group ( G / N , τ / N )   , where = s   .
A homomorphism h : G L   of abstract groups factors through the quotient G / N   if and only if N ker h   . Furthermore, by the First Isomorphisms Theorem, G / ker h = im h   . While the first statement remains true in the category of topological groups, this is not the case for the second one. Because of the additional structure that topology constitutes, there are many continuous bijective homomorphisms that are not isomorphisms (i.e., their inverses are not continuous). In the next theorem, we propose an alternative to the well-known notion of kernel for the category of Hausdorff topological groups.
Theorem 2.4. Let f : G ( L , τ )   be a homomorphism of G   into a Hausdorff group ( L , τ )   .
Put Ker f = { f 1 ( U ) | U N τ }   .
  • (a) Given a filter   on G   , the following statements are equivalent:
    • (i) there is a (unique) continuous homomorphism f ¯ : G { } ( L , τ )   such that f = f ¯ p   , where p : G G / N   is the canonical projection;
    • (ii) f ( ) τ e   ;
    • (iii) Ker f   ;
  • (b) In the category of Hausdorff topological groups, G { Ker f } = im f   holds.
We refer to part (a) of Theorem  2.4 as the universal property of G { }   . Observe that if ( L , τ )   is discrete, then Ker f = N ˙   , where N = ker f   .
  • Proof. (a) (i)   (ii): If f ¯   is continuous, then f : ( G , τ s ) ( L , τ )   , being the composite of continuous maps, is continuous. Thus, by Theorem  2.2 (c), f ( s ) τ e   , and therefore (ii) follows, because f ( s ) f ( )   .
    (ii)   (iii): (ii) means that for every U N τ   , there is F   such that f ( F ) U   . In other words, F f 1 ( U )   , which proves (iii).
    (iii)   (i): Since Ker f   is symmetric and normal, (iii) implies Ker f s   . Thus, by Theorem  2.2 (c), f : ( G , τ s ) ( L , τ )   is continuous. Certainly, G / ker f = im f   as abstract groups, and so τ   induces a Hausdorff group topology on the quotient G / ker f   . Thus, the image of s   in G / ker f   is a T   -filter. Therefore, by Theorem  2.2 (f), N s ker f   , and hence f   factors through f ¯ : G { } ( L , τ )   .
    (b) Without loss of generality, we may assume that f   is surjective. By (a) applied to = Ker f   , f ~ : G { Ker f } ( L , τ )   is continuous. Since τ   is Hausdorff and Ker f   is a base at e   for the initial group topology on G   induced by f   , N Ker f = Ker f = ker f   , and thus f ~   is a bijection. For every [ { U n } ] Ker f   , U 1 [ { U n } ]   holds, therefore f ~   is open. Hence, f ~   is a homeomorphism.
We conclude this section with problems arising in the context of non-abelian groups whose solution for abelian groups is known, at least in some special cases.
Problem III. Is it possible to characterize T   -sequences in a non-abelian group in a more simple way than Theorem  2.2 ?
For a sequence { a n }   in an abelian group A   , Zelenyuk and Protasov gave an elegant criterion for being a T   -sequence (see Theorem  3.1 below).
Problem IV. Suppose G 1 G 2   , G 2   is a Hausdorff topological group, and b   in G 2   , where b   is such that G 1 b = { e }   . Under which additional conditions can one conclude that   is a T   -filter?
No additional conditions are required if the groups are abelian and   is generated by a sequence (cf. [?,Thm. 3). Furthermore, for an abelian group A   and a T   -sequence { a n }   in A   , the group A { a n }   is complete (cf. [?,Thm. 8), which motivates the following problem:
Problem V. How to characterize filters   in a group G   such that G { }   is (Raĭkov-)complete?

3 T-SEQUENCES IN ABELIAN GROUPS

In this section, A   is an abelian group and a ̲ = { a k } A   is a sequence in A   . In what follows, we provide several sufficient conditions for { a k }   to be a T   -sequence in A   . For l , m N   , one puts
A ( l , m ) a ̲ = { m 1 a k 1 + + m h a k h | m k 1 < < k h , m i Z \ { 0 } , | m i | l } . (4)
The Zelenyuk-Protasov criterion for T   -sequences states:
Theorem 3.1. ([?,Theorem 2) A sequence { a k }   in an abelian group A   is a T   -sequence if and only if for every l N   and g 0   , there exists m N   such that g A ( l , m ) a ̲   .
Put A [ n ] = { a A : n a = 0 }   for every n N   . One says that A   is almost torsion-free if A [ n ]   is finite for every n N   (cf. [?).
Theorem 3.2. Let A   be an abelian group, and let { a n } A   be a sequence such that t k : = o ( a k )   is finite for every k N   . Consider the following statements:
  • (i)
    lim k t k gcd ( t k , lcm ( t 1 , , t k 1 ) ) = . (5)
  • (ii) For every l N   ,
    lim m inf { max 1 i l t k i gcd ( t k i , lcm ( t k 1 , , t k i 1 , t k i + 1 , , t k l ) ) | k ̲ N m < l } = , (6)
    where N m < l = { k ̲ = ( k 1 , , k l ) N l | m k 1 < < k l }   .
  • (iii) For every l N   ,
    lim m inf { max 1 i l o ( a k i + A k ̲ i ) | k ̲ N m < l } = , (7)
    where A k ̲ i = a k 1 , , a k i 1 , a k i + 1 , , a k h   .
  • (iv) For every l , n N   , there exists m 0 N   such that A [ n ] A ( l , m ) a ̲ = { 0 }   for every m m 0   .
  • (v) { a n }   is a T   -sequence.
One has (i)   (ii)   (iii)   (iv)   (v), and if A   is almost torsion-free, then (v)   (iv).
  • Proof. (i)   (ii) is obvious.
    (ii)   (iii): Clearly, the order of a k i   in ( a k i + A k ̲ i ) / A k ̲ i   is equal to its order modulo a k i A k ̲ i   , and | a k i A k ̲ i |   divides both t k i   and exp ( A k ̲ i )   . The exponent exp ( A k ̲ i )   , in turn, divides d = lcm ( t k 1 , , t k i 1 , t k i + 1 , , t k l )   , because A k ̲ i   is generated by elements of orders t k 1 , , t k i 1 , t k i + 1 , , t k l   . Therefore, | a k i A k ̲ i |   divides their greatest common divisor of t k   and d   . Hence,
    t k i gcd ( t k i , lcm ( t k 1 , , t k i 1 , t k i + 1 , , t k l ) ) | | a k i | | a k i A k ̲ i | = o ( a k i + A k ̲ i ) . (8)
    (iii)   (iv): Given l , n N   , let m 0 N   be such that n l < max 1 i h o ( a k i + A k ̲ i )   for every 1 h l   and every k ̲ N m 0 < h   . (By ( 7 ), such m 0   exists.) Let g = m 1 a k 1 + + m h a k h A ( l , m ) a ̲   be a non-zero element, where m 0 m k 1 < < k h   , m i Z \ { 0 }   , and | m i | l   . It follows from the last two conditions that h l   . So, there exists 1 i h   such that n l < o ( a k i + A k ̲ i )   , and thus n < o ( m i a k i + A k ̲ i )   . To complete the proof, note that g m i a k i + A k ̲ i   , and therefore o ( m i a k i + A k ̲ i ) | o ( g )   . Hence, n < o ( g )   , and so g A [ n ]   , as desired.
    (iv)   (v): Let g A   be a non-zero element. If the order of g   is infinite, then g A ( l , 1 ) a ̲   for every l N   , and so suppose that n : = o ( g )   is finite. By (iii), for every l N   there exists m 0 ( l )   such that A [ n ] A ( l , m 0 ( l ) ) a ̲ = { 0 }   . In particular, g A ( l , m 0 ( l ) ) a ̲   for every l   .
    (v)   (iv): Given l , n N   , and suppose that A [ n ] = { 0 , g 1 , , g j }   is finite. For each g i   , pick m i ( l ) N   such that g i A ( l , m i ( l ) ) a ̲   , and put m 0 ( l ) = max m i ( l )   . Clearly, one has A [ n ] A ( l , m ) a ̲ = { 0 }   for every m m 0 ( l )   , as desired.
Remark 3.3. In Theorem  3.2 , (iv) does not imply (iii). Indeed, although (iii) fails for the sequence { e n }   from Example  1.1 , it is a T   -sequence in Z ( p )   .
Corollary 3.4. Let A   be an abelian group, and let { a n } A   be a sequence such that t k : = o ( a k )   is finite for every k N   .
  • (a) If the t k   are pairwise coprime, then { a k }   is a T   -sequence.
  • (b) If t k | t k + 1   and lim k t k + 1 t k =   , then { a k }   is a T   -sequence.
  • Proof. If the t k   are pairwise coprime, then the expression in ( 5 ) is equal to t k   and t k   . If t k | t k + 1   , then the expression in ( 5 ) is precisely t k + 1 t k   . In both cases, the statement follows from Theorem  3.2 (i).

4 SPECIAL SEQUENCES IN THE PR¨FER GROUPS

In this section, we present method for constructing a Hausdorff abelian topological groups A   such that n ( A )   is finite and non-trivial. An implicit yet rather thick hint for the construction of a group with these properties appears in [?, in the proof of Corollary 4.9 and the paragraph following it. It was Dikranjan who pointed out to the author that considering a suitable T   -sequence in a Prüfer group (and the maximal Hausdorff group topology thus obtained) would lead to the desired example. Prüfer groups are distinguished by the property of having only finite proper subgroups, which implies that these subgroups are closed in any Hausdorff group topology. This property makes Prüfer groups particularly suitable candidates for the aforesaid construction, because if n ( Z ( p ) { d n } )   is a non-trivial proper subgroup, then it must be finite. Therefore, in this section, we study certain sequences in the Prüfer groups Z ( p )   , and construct a T   -sequence { d n }   such that Z ( p ) { d n }   is neither maximally nor minimally almost periodic. A second important property that Prüfer groups, being p   -groups, have is that for every a , b Z ( p )   ,
o ( a ) o ( b ) o ( a + b ) = max { o ( a ) , o ( b ) } . (9)
Recall that if A   is a subgroup of an abelian Hausdorff topological group S   , and { a k } A   is so that a k b   in S   , where b A = { 0 }   , then { a k }   is a T   -sequence in A   (cf. [?,Thm. 3). The setting of this result is so that the sequence a n   converges to an external element (namely, b A   ) in some group topology. In contrast, in this section, we investigate sequences in Z ( p )   that converge to a non-zero (internal) element of Z ( p )   in the “usual” topology, that is, the one inherited from Q / Z   .
We start off with an immediate consequence of Corollary  3.4 .
Lemma 4.1. Let { a k }   be a sequence in Z ( p )   , and suppose that o ( a k ) = p n k   . If n k + 1 n k   , then { a k }   is a T   -sequence.
  • Proof. One has o ( a k + 1 ) o ( a k ) = p n k + 1 p n k = p n k + 1 n k   . Therefore, the statement follows from Corollary  3.4 (b).
Example  1.1 shows that the converse of Lemma  4.1 does not hold in general. Nevertheless, for some special sequences, the condition of n k + 1 n k   turns out to be necessary for being a T   -sequence, as Theorem  4.2 below reveals.
We proceed by introducing some terminology. A topological group G   is precompact if for every neighborhood U   of the identity in G   there is a finite subset F G   such that G = F U   .
Following [?, we say that a filter   (a sequence { a n }   ) on a group G   is a T B   -filter ( T B   -sequence) if there is a precompact Hausdorff group topology τ   on G   such that τ 0   ( a n τ 0   ).
An abelian group is precompact if and only if it carries the initial topology induced by its group of continuous characters. Thus, if { a k }   is a sequence in an abelian group A   , then by the universal property of A { a k }   , a character χ : A T   is continuous on A { a k }   if and only if χ ( a k ) 0   . Therefore, for H = { χ hom Z ( A , T ) | χ ( a k ) 0 }   , the closure of { 0 }   in the initial topology induced by H   is n ( A { a k } ) = χ H ker χ   . Hence, { a k }   is a T B   -sequence if and only if H   separates the points of A   . (Observe that H = A { a k } ^ d   .)
Theorem 4.2. Let x Z ( p )   be a non-zero element, { n k } N   an increasing sequence of positive integers, and set
a k = x + e n k = x + 1 p n k Z ( p ) . (10)
  • (a) { a k }   is a T   -sequence if and only if n k + 1 n k   .
  • (b) { a k }   is a T   -sequence if and only if it is a T B   -sequence. In this case, Z ( p ) { a k }   is maximally almost periodic, and it has c   many faithful characters (in particular, | Z ( p ) { a k } ^ | = c   ).
Since every T B   -filter is a T   -filter, but the converse is not true in general, (b) of the Theorem  4.2 is a non-trivial result. Its proof, however, requires a technical lemma. Note that the group of all characters of Z ( p )   is isomorphic to the group Z p   of the p   -adic integers. In other words, Z p = hom Z ( Z ( p ) , T )   .
Lemma 4.3. Suppose that n k + 1 n k   . For χ = n = 0 α n p n Z p   ( 0 α n p 1   ) and γ ( 0 , 1 )   , χ ( e n k ) γ   if and only if
r k : = l = n k n k + 1 1 α l p l n k p n k + 1 n k γ . (11)
  • Proof. One has
    χ ( e n k + 1 ) = l = 0 n k + 1 1 α l p l p n k + 1 = l = 0 n k 1 α l p l p n k + 1 + l = n k n k + 1 1 α l p l p n k + 1 = χ ( e n k ) p n k + 1 n k + r k , (12)
    and thus lim k χ ( e n k ) = lim k r k   in T   (by the equality of limits we mean that one exists if and only if the other does, and in that case they are equal), because n k + 1 n k   . Since γ 0   , small enough neighborhoods of γ   in T   and ( 0 , 1 )   are the same, and therefore lim k χ ( e n k ) = lim k r k   in ( 0 , 1 )   .
  • Proof of Theorem  4.2 . (a) Since { n k }   is increasing, one has n k   . Thus, p n k > o ( x )   for k   large enough, and so o ( a k ) = p n k   except for maybe a finite number of k   (by ( 9 )). Hence, a k   is a T   -sequence by Lemma  4.1 .
    Conversely, let p n 0 = o ( x )   , and assume that n k + 1 n k ⟶̸   . Then o ( p n 0 1 x ) = p   ,
    p n 0 1 a k = p n 0 1 x + e n k n 0 + 1 , (13)
    and the differences ( n k + 1 n 0 + 1 ) ( n k n 0 + 1 ) = n k + 1 n k ⟶̸   . Thus, it suffices to show that p n 0 1 a k   is not a T   -sequence. Therefore, without loss of generality, we may assume that o ( x ) = p   from the outset. Since n k + 1 n k ⟶̸   , there exists a number d   and a subsequence k r   of k   such that n k r + 1 n k r d   for every r   . If a k 0   in a group topology τ   on Z ( p )   , then in particular, p a n k r + 1 = e n k r + 1 1 0   , and so for every 1 i d   , e n k r + 1 i 0   . Thus, the sequence b n   defined as e n k 1 + 1 d , e n k 1 + 1 d + 1 , , e n k 1 + 1 1 , e n k 2 + 1 d , e n k 2 + 1 d + 1 , , e n k 2 + 1 1 ,   also converges to 0   in τ   . One has n k r + 1 d n k r n k r + 1 1   , and therefore e n k r   is a subsequence of b n   , and hence e n k r 0   in τ   . Since a k r = x + e n k r   , this shows that τ   is not Hausdorff. (b) If { a k }   is a T B   -sequence, then clearly it is a T   -sequence. Conversely, suppose that { a k }   is a T   -sequence. In order to show that { a n }   is a T B   -sequence, we find a faithful continuous character of Z ( p ) { a k }   , in other words, χ Z p   such that χ ( a k ) 0   and ker χ = { 0 }   . Let p n 0 = o ( x )   . For χ = n = 0 α n p n Z p   , a character of Z ( p )   , if
    α 0 = 1 , α 1 = = α n 0 1 = 0 , (14)
    then χ   acts on the subgroup x   as the identity, where Z ( p )   is viewed as a subgroup of T   . Thus, χ ( e 1 ) = e 1 0   and χ ( x ) = x 0   , and in particular, χ   is faithful. By Lemma  4.3 , χ ( a k ) 0   (i.e., χ   is continuous on Z ( p ) { a k }   ) if and only if
    r k = l = n k n k + 1 1 α l p l n k p n k + 1 n k x . (15)
    Conditions ( 14 ) and ( 15 ) are satisfied (simultaneously) by continuum many elements in Z p   , which completes the proof.
We proceed by presenting the construction of a non-minimally almost periodic non-maximally almost periodic Hausdorff group topology on the group Z ( p )   for p 2   . Our technique makes substantial use of the assumption that p 2   ; nevertheless, we conjecture that a similar construction is available for p = 2   .
Theorem 4.4. Let p   be a prime number such that p 2   , x Z ( p )   be a non-zero element with p n 0 = o ( x )   , and put
b n = x + e n 3 n 2 + + e n 3 2 n + e n 3 n + e n 3 = x + 1 p n 3 n 2 + + 1 p n 3 2 n + 1 p n 3 n + 1 p n 3 . (16)
Consider the sequence d n   defined as b 1 , e 1 , b 2 , e 2 , b 3 , e 3 ,   . Then:
  • (a) { d n }   is a T   -sequence in Z ( p )   ;
  • (b) the underlying group of Z ( p ) { d n } ^   is p n 0 Z Z p = hom Z ( Z ( p ) , T )   ;
  • (c) n ( Z ( p ) { d n } ) = x   .
In particular, Z ( p ) { d n }   is neither maximally almost periodic nor minimally almost periodic,and n ( Z ( p ) { d n } )   is finite.
Corollary 4.5. Let p   be a prime number such that p 2   , and put
b n = e 1 + e n 3 n 2 + + e n 3 2 n + e n 3 n + e n 3 = 1 p + 1 p n 3 n 2 + + 1 p n 3 2 n + 1 p n 3 n + 1 p n 3 . (17)
Consider the sequence d n   defined as b 1 , e 1 , b 2 , e 2 , b 3 , e 3 ,   . Then:
  • (a) { d n }   is a T   -sequence in Z ( p )   ;
  • (b) the underlying group of Z ( p ) { d n } ^   is p Z Z p = hom Z ( Z ( p ) , T )   ;
  • (c) n ( Z ( p ) { d n } ) = 1 p   .
In particular, Z ( p ) { d n }   is neither maximally almost periodic nor minimally almost periodic, and n ( Z ( p ) { d n } )   is finite.
In order to prove Theorem  4.4 , several auxiliary results of a technical nature are required.
Until the end of this section, we assume that p 2   . Each element y Z ( p )   admits many representations of the form y = σ n e n   , where σ n Z   (only finitely many of the σ n   are non-zero), and so we say that it is the canonical form of y   if | σ n | p 1 2   for every n N   ; in this case, we put Λ ( y ) = { n N | σ n 0 }   and λ ( y ) = | Λ ( y ) |   .
Lemma 4.6. Let y = σ n e n Z ( p )   . Then:
  • (a) y   admits a canonical form y = σ n e n   , and | σ n | | σ n |   ;
  • (b) the canonical form is unique, and so Λ   is well-defined.
    Furthermore,
  • (c) λ ( z ) l   for every z Z ( p ) ( l , 1 ) e ̲   and l N   .
  • Proof. (a) Let N   be the largest index such that σ N 0   . We proceed by induction on N   . If N = 1   , then y = σ 1 c 1   . Thus, if σ 1 = σ 1 + m p   is a division with residue in Z   , and σ 1   is chosen to have the smallest possible absolute value, then | σ 1 | p 1 2   , and
    y = ( σ 1 + p m ) e 1 = σ 1 e 1 + m p e 1 = σ 1 e 1 . (18)
    In particular, | σ 1 | | σ 1 |   . Suppose now that the statement holds for all elements with representation with maximal non-zero index less than N   . If σ N = σ N + k p   is a division with residue in Z   , and σ N   is chosen to have the smallest possible absolute value, then | σ N | p 1 2   , and
    y n = 1 N 2 σ n e n ( σ N 1 + k ) e N 1 = k e N 1 + σ N e N = k e N 1 + ( σ N + k p ) e N = σ N e N . (19)
    The element z = n = 1 N 2 σ n e n + ( σ N 1 + k ) e N 1   satisfies the inductive hypothesis, so z = n = 1 N 1 σ n e n   , where | σ n | p 1 2   and n = 1 N 1 | σ n | n = 1 N 1 | σ n | + | k |   . Therefore, y = σ n c n   , | σ n | p 1 2   , and | σ n | | σ n |   , because | σ N | + | k | | σ N |   .
    (b) Suppose that σ n e n = υ n e n   are two distinct canonical representations of the same element. Then ( σ n υ n ) e n = 0   , and | σ n υ n | p 1   . Let N   be the largest index such that σ N υ N   . (Since all coefficients are zero, except for a finite number of indices, such N   exists.) This means that 0 < | σ N υ N | p 1   , and o ( ( σ N υ N ) e N ) = p N   . Therefore, by ( 9 ), one has o ( ( σ n υ n ) e n ) = p N   , because o ( n < N ( σ n υ n ) e n ) p N 1   . This is a contradiction, and therefore σ n = υ n   for every n N   .
    (c) Let z = μ 1 e n 1 + + μ h e n h   , where | μ i | l   and n 1 < n 2 < < n h   . By (a), z   admits a canonical form z = μ n e n   , and | μ n | | μ i | l   . Therefore, μ n 0   only for at most l   many indices.
Lemma 4.7. Let m Z \ { 0 }   , and put l = log p | m |   . If n > l   , then Λ ( m e n ) { n l , , n 1 , n }   and 1 λ ( m e n )   .
  • Proof. It follows from n > l   that p n > | m |   , and so m e n 0   . Thus, 1 λ ( m e n )   . To show the first statement, expand m = μ 0 + μ 1 p + + μ l p l   , where μ i Z   and | μ i | p 1 2   . Then
    m e n = μ 0 e n + μ 1 e n 1 + + μ l e n l (20)
    is in canonical form, and therefore Λ ( m e n ) { n l , , n 1 , n }   , as desired.
Lemma 4.8. Let y , z Z ( p )   such that λ ( y ) > λ ( z )   , and suppose that Λ ( y ) = { k 1 , , k g }   where k 1 < < k g   and g = λ ( y )   . Then o ( y + z ) p k g λ ( z )   .
  • Proof. Let y = ν n e n   and z = μ n e n   in canonical form. Then y + z = ( ν n + μ n ) e n   , and | ν n + μ n | p 1   . Clearly, o ( y + z ) = p N   for N   the largest index such that ν N + μ N 0   . By the definition of N   , μ n = ν n   for every n > N   . In particular, μ k i 0   for every i   such that k i > N   .
    Thus, there are at most λ ( z )   many i   such that k i > N   , and therefore N k g λ ( z )   .
Remark 4.9. If y 1 , y 2 Z ( p )   and Λ ( y 1 ) Λ ( y 2 ) =   , then Λ ( y 1 + y 2 ) = Λ ( y 1 ) Λ ( y 2 )   and λ ( y 1 + y 2 ) = λ ( y 1 ) + λ ( y 2 )   .
Proposition 4.10. Let y = ν 1 e n 1 + ν f e n f   , where n 1 < < n f   and ν i 0   . Put l i = log p | ν i |   , and suppose that n i < n i + 1 l i + 1   for each 1 i f   . Then:
  • (a) f λ ( y )   ;
  • (b) if z Z ( p )   is such that λ ( z ) < λ ( y )   , then o ( y + z ) p n f λ ( z ) l f λ ( z )   .
  • Proof. (a) By Lemma  4.7 , Λ ( ν i e n i ) { n i l i , , n i }   , and since n i 1 < n i l i   , the sets Λ ( ν i e n i )   are pairwise disjoint. Therefore, by Remark  4.9 , λ ( y ) = λ ( ν 1 e n 1 ) + + λ ( ν f e n f ) f   , and
    Λ ( y ) i = 1 f { n i l i , , n i } = { n 1 l 1 , , n 1 , , n i l i , , n i , , n f l f , , n f } . (21)
    (b) By Lemma  4.8 , o ( y + z ) p k λ ( y ) λ ( z )   , where Λ ( y ) = { k 1 , , k g }   (increasingly ordered).
    Since Λ ( ν i e n i )   is non-empty for each i   , it follows from ( 21 ) that k λ ( y ) λ ( z ) n f λ ( z ) l f λ ( z )   .
Corollary 4.11. Let l N   , z Z ( p ) ( l , 1 ) e ̲   , and y = e n 1 + + e n f   such that n 1 < < n f   , l < f   , and n i < n i + 1 l   . Then o ( μ y + z ) p n f l l p n 1 l   for every μ Z   such that 0 < | μ | l   .
  • Proof. Since | μ | l   , μ y = μ e n 1 + + μ e n f   satisfies the conditions of Proposition  4.10 (because log p | μ | l   ), and thus, l < f λ ( ν y )   . On the other hand, by Lemma  4.6 (c), λ ( z ) l   , and therefore o ( ν y + z ) p n f λ ( z ) l p n f l l   pursuant to Proposition  4.10 (b).
  • Proof of Theorem  4.4 . To shorten notations, put A = Z ( p )   .
    (a) In order to prove that { d n }   is a T   -sequence, we show that (iv) of Theorem  3.2 holds. For n   large enough, o ( b n ) = p n 3   , and so by Lemma  4.1 , { b n }   is a T   -sequence; { e n }   is evidently a T   -sequence (cf. Example  1.1 ). Thus, by Theorem  3.2 , there exists m 0   such that
    A [ n ] A ( l , m ) b ̲ = A [ n ] A ( l , m ) e ̲ = { 0 } (22)
    for every m m 0   (because A   is almost torsion-free). Without loss of generality, we may assume that m 0 > l + n + n 0   . Observe that
    A ( l , 2 m ) d ̲ A ( l , m ) b ̲ A ( l , m ) e ̲ ( A ( l , m ) b ̲ \ { 0 } + A ( l , m ) e ̲ \ { 0 } ) , (23)
    and therefore it suffices to show that A ( l , m ) b ̲ \ { 0 } + A ( l , m ) e ̲ \ { 0 }   contains no element of A [ n ]   for every m m 0   . Let z A ( l , m ) e ̲ \ { 0 }   and w = m 1 b n 1 + + m h b n h A ( l , m ) b ̲ \ { 0 }   where m n 1 < < n h   and 0 < | m i | l   . Put y = e n h 3 n h 2 + + e n h 3 n h + e n h 3   . The number of summands in y   is n h + 1   , and the differences between the indices of the terms is n h   . By the construction, n h m m 0 > l   . Thus, the conditions of Corollary  4.11 are satisfied, and since | m h | l   , we get o ( m h y + z ) p n h 3 n h 2 l > p ( n h 1 ) 3   . Therefore, o ( m h x ) o ( m h y + z )   (because o ( m h x ) p n 0 p m 0 1 p ( n h 1 ) 3   ), and  9 
    o ( m h b n h + z ) = o ( m h x + m h y + z ) = ( ) max { o ( m h x ) , o ( m h y + z ) } > p ( n h 1 ) 3 . (24)
    One has
     24  9 
    o ( w m h b n h ) o ( b n h 1 ) = p n h 1 3 p ( n h 1 ) 3 < ( ) o ( m h b n h + z ) , (25)
    and hence o ( w + z ) = o ( ( w m h b n h ) + ( z + m h b n h ) ) (26)
    = ( ) max { o ( w m h b n h ) , o ( z + m h b n h ) } > p ( n h 1 ) 3 > p ( m 0 1 ) 3 > n . (27)
    (b) As noted earlier, a character χ hom Z ( Z ( p ) , T )   is continuous on Z ( p ) { d n }   if and only if χ ( d n ) 0   (by the universal property)—in other words, χ ( b n ) 0   and χ ( e n ) 0   . The latter is equivalent to χ   having the form of m χ 1   , where χ 1   is the natural embedding of Z ( p )   into T   and m Z   (cf. [?,Example 6, [?,3.3). Since
    0 1 p n 3 n 2 + + 1 p n 3 2 n + 1 p n 3 n + 1 p n 3 n + 1 p n 3 n 2 0 , (28)
    one has χ 1 ( b n ) x   , and consequently χ ( b n ) = m χ 1 ( b n ) 0   if and only if m x = 0   (i.e., x ker χ   ).
    This means that χ = m χ 1   if and only if o ( x ) = p n 0 | m   , as desired.
    (c) We have already seen that x ker χ   for every continuous character of Z ( p ) { d k }   . On the other hand, n ( Z ( p ) { d k } ) ker p n 0 χ 1 = x   .
ACKNOWLEDGMENTS I am deeply indebted to Professor Dikran Dikranjan, who introduced me to this topic and spent so much time corresponding with me about it. Without his attention and encouragement, I would have struggled to carry out this research.
I wish to thank Professor Horst Herrlich, Professor Salvador Hernández and Dr. Gavin Seal for the valuable discussions and helpful suggestions that were of great assistance in writing this paper.
Department of Mathematics and Statistics Dalhousie University Halifax, B3H 3J5, Nova Scotia Canada e-mail: lukacs@mathstat.yorku.ca