657
Views
3
CrossRef citations to date
0
Altmetric
Original Articles

A characterization of finite abelian groups via sets of lengths in transfer Krull monoids

ORCID Icon
Pages 4021-4041 | Received 04 Dec 2017, Published online: 26 Feb 2018

ABSTRACT

Let H be a transfer Krull monoid over a finite abelian group G (for example, rings of integers, holomorphy rings in algebraic function fields, and regular congruence monoids in these domains). Then each nonunit aH can be written as a product of irreducible elements, say a=u1uk, and the number of factors k is called the length of the factorization. The set L(a) of all possible factorization lengths is the set of lengths of a. It is classical that the system (H) = {L(a)∣aH} of all sets of lengths depends only on the group G, and a standing conjecture states that conversely the system (H) is characteristic for the group G. Let H be a further transfer Krull monoid over a finite abelian group G and suppose that (H) = (H). We prove that, if GCnr with rn−3 or (rn−1≥2 and n is a prime power), then G and G are isomorphic.

2010 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction and main result

Let H be an atomic unit-cancellative monoid. Then each non-unit aH can be written as a product of atoms, and if a=u1uk with atoms u1,,uk of H, then k is called the length of the factorization. The set L(a) of all possible factorization lengths is the set of lengths of a, and (H) = {L(a)∣aH} is called the system of sets of lengths of H (for convenience we set L(a) = {0} if a is an invertible element of H). Under a variety of noetherian conditions on H (e.g., H is the monoid of nonzero elements of a commutative noetherian domain) all sets of lengths are finite. Sets of lengths (together with invariants controlling their structure, such as elasticities and sets of distances) are a well-studied means for describing the arithmetic structure of monoids.

Let H be a transfer Krull monoid over a finite abelian group G. Then, by definition, there is a weak transfer homomorphism 𝜃:H(G), where (G) denotes the monoid of zero-sum sequences over G, and hence (H) = ((G)). We use the abbreviation (G) = ((G)). By a result due to Carlitz in 1960, we know that H is half-factorial (i.e., |L| = 1 for all L(H)) if and only if |G|≤2. Suppose that |G|≥3. Then there is some aH with |L(a)|>1. If k,L(a) with k< and m, then L(am)⊃{km+ν(k)∣ν∈[0,m]} which shows that sets of lengths can become arbitrarily large. Note that the system of sets of lengths of H depends only on the class group G. The associated inverse question asks whether or not sets of lengths are characteristic for the class group. In fact, we have the following conjecture (it was first stated in [Citation6] and for a detailed description of the background of this problem, see [Citation6], [Citation7, Section 7.3], [Citation8, page 42], and [Citation24]).

Conjecture 1.1.

Let G be a finite abelian group with D(G)≥4. If G is an abelian group with (G) = (G), then G and G are isomorphic.

Note if D(G) = 3, then we have (C3)=(C2C2). The system of sets of lengths (G) is studied with methods from Additive Combinatorics. In particular, zero-sum theoretical invariants (such as the Davenport constant or the cross number) and the associated inverse problems play a crucial role. Most of these invariants are well-understood only in a very limited number of cases (e.g., for groups of rank two, the precise value of the Davenport constant D(G) is known and the associated inverse problem is solved; however, if n is not a prime power and r≥3, then the value of the Davenport constant D(Cnr) is unknown). Thus it is not surprising that most affirmative answers to the Characterization Problem so far have been restricted to those groups where we have a good understanding of the Davenport constant. These groups include elementary two-groups, cyclic groups, and groups of rank 2 (for recent progress we refer to [Citation1, Citation9]).

The first and so far only groups, for which the Characterization Problem was solved whereas the Davenport constant is unknown, are groups of the form Cnr, where r,n and 2r<n−2 (this is done by Geroldinger and Zhong [Citation12] and Zhong [Citation27]), which use a deep characterization of the structure of Δ*(G). In this paper, we go on to study groups of the form Cnr and obtain the following theorem.

Theorem 1.2.

Let H be a transfer Krull monoid over a finite abelian group G with D(G)≥4. Suppose GCnr with r,n and (H) = (H), where H is a further transfer Krull monoid over a finite abelian group G. Then

  1. If rn−3, then GG.

  2. If rn−1 and n is a prime power, then GG.

This is made possible by introducing new invariants ρ(G,d) and ρ*(G,d) which are only depending on (G) (see Definitions 3.1 and 3.3). In Section 2 we gather the required background both on transfer Krull monoids as well as on Additive Combinatorics. In Section 3, we provide a detailed study of ρ(G,d) and ρ*(G,d). The proof of Theorem 1.2 will be provided in Section 4. The final section is concluding remarks and conjectures.

Throughout the paper, let GCn1Cnr be a finite abelian group with D(G)≥4, where r,n1,,nr and 1<n1||nr.

2. Background on transfer Krull monoids and sets of lengths

Our notation and terminology are consistent with [Citation7, Citation9]. For convenience, we set min = 0. Let be the set of positive integers, let be the set of integers, let be the set of rational numbers, and let ℝ be the set of real numbers. For rational numbers a,b, we denote by [a,b] = {xaxb} the discrete, finite interval between a and b. If L is a subset, then Δ(L) denotes the set of (successive) distances of L (that is, dΔ(L) if and only if d = ba with a,bL distinct and [a,b]∩L = {a,b}) and ρ(L) = supL∕minL denotes its elasticity (for convenience, we set ρ({0}) = 1).

Let r and let (e1,,er) be an r-tuple of elements of G. Then (e1,,er) is said to be independent if ei≠0 for all i∈[1,r] and if for all (m1,,mr)r an equation m1e1++mrer=0 implies that miei=0 for all i∈[1,r]. Furthermore, (e1,,er) is said to be a basis of G if it is independent and G=e1er. For every n, we denote by Cn an additive cyclic group of order n. Since GCn1Cnr, r = r(G) is the rank of G and nr = exp(G) is the exponent of G.

2.1. Sets of lengths

By a monoid, we mean an associative semigroup with unit element, and if not stated otherwise we use multiplicative notation. Let H be a monoid with unit element 1 = 1HH. An element aH is said to be invertible (or an unit) if there exists an element aH such that aa=aa=1. The set of invertible elements of H will be denoted by H×, and we say that H is reduced if H× = {1}. The monoid H is said to be unit-cancellative if for any two elements a,uH, each of the equations au = a or ua = a implies that uH×. Clearly, every cancellative monoid is unit-cancellative.

Suppose that H is unit-cancellative. An element uH is said to be irreducible (or an atom) if uH× and for any two elements a,bH, u = ab implies that aH× or bH×. Let 𝒜(H) denote the set of atoms, and we say that H is atomic if every non-unit is a finite product of atoms. If H satisfies the ascending chain condition on principal left ideals and on principal right ideals, then H is atomic [Citation4, Theorem 2.6]. If aHH× and a=u1uk, where k and u1,,uk𝒜(H), then k is a factorization length of a, and

LH(a)=L(a)={kkis a factorization length ofa}
denotes the set of lengths of a. It is convenient to set L(a) = {0} for all aH×. The family
(H)={L(a)aH}
is called the system of sets of lengths of H, and
ρ(H)=sup{ρ(L)L(H)}1{}
denotes the elasticity of H. We call
Δ(H)=L(H)Δ(L)
the set of distances of H. Note that Δ(H) can be infinite, and by definition we have Δ(H) =  if and only if ρ(H) = 1. If Δ(H)≠, then we have minΔ(H) = gcdΔ(H) [Citation3, Proposition 2.9].

2.2. Monoids of zero-sum sequences

Let G0G be a non-empty subset. Then ⟨G0⟩ denotes the subgroup generated by G0. In Additive Combinatorics, a sequence (over G0) means a finite sequence of terms from G0 where repetition is allowed and the order of the elements is disregarded, and (as usual) we consider sequences as elements of the free abelian monoid with basis G0. Let

S=g1g=gG0gvg(S)(G0)
be a sequence over G0. We call
supp(S)={gGvg(S)>0}Gthesupportof S,|S|==gGvg(S)0thelengthof S,σ(S)=i=1githesumof S, and Σ(S)={iIgiI[1,]} thesetofsubsequencesumsof  S.

The sequence S is said to be

  • zero-sum free if 0∉Σ(S),

  • a zero-sum sequence if σ(S) = 0,

  • a minimal zero-sum sequence if it is a nontrivial zero-sum sequence and every proper subsequence is zero-sum free.

The set of zero-sum sequences (G0)={S(G0)σ(S)=0}(G0) is a submonoid, and the set of minimal zero-sum sequences is the set of atoms of (G0). For any arithmetical invariant ∗(H) defined for a monoid H, we write ∗(G0) instead of ∗((G0)). In particular, 𝒜(G0)=𝒜((G0)) is the set of atoms of (G0), (G0)=(B(G0)) is the system of sets of lengths of (G0), and so on. We denote by

D(G0)=max{|S|S𝒜(G0)}
the Davenport constant of G0.

2.3. Transfer Krull monoids

Let H be a atomic unit-cancellative monoid.

  1. We say a monoid homomorphism 𝜃:HB to an atomic unit-cancellative monoid B is a weak transfer homomorphism if it has the following two properties:

    • B=B×𝜃(H)B× and 𝜃1(B×)=H×.

    • If aH, n, v1,,vn𝒜(B) and 𝜃(a)=v1vn, then there exist u1,,un𝒜(H) and a permutation τ𝔖n such that a=u1un and 𝜃(ui)B×vτ(i)B× for each i∈[1,n].

    Let 𝜃:HB be a weak transfer homomorphism between atomic unit-cancellative monoids. It follows that for all aH, we have LH(a)=LB(𝜃(a)) and hence (H) = (B).

  2. We say H is a transfer Krull monoid if one of the following equivalent conditions holds:

    • There exists a weak transfer homomorphism 𝜃:H* for a commutative Krull monoid * (i.e., * is commutative, cancellative, completely integrally closed, and v-noetherian).

    • There exists a weak transfer homomorphism 𝜃:H(G0) for a subset G0 of an abelian group.

    If the second condition holds, then we say H is a transfer Krull monoid over G0. If G0 is finite, then H is said to be a transfer Krull monoid of finite type.

  3. We say a domain R is a transfer Krull domain (of finite type) if its monoid of cancelative elements R is a transfer Krull monoid (of finite type).

In particular, commutative Krull monoids are transfer Krull monoids. Rings of integers, holomorphy rings in algebraic function fields, and regular congruence monoids in these domains are commutative Krull monoids with finite class group such that every class contains a prime divisor [Citation7, Section 2.11 and Examples 7.4.2]. Monoid domains and power series domains that are Krull are discussed in [Citation17, Citation2], and note that every class of a Krull monoid domain contains a prime divisor. Thus all these commutative Krull monoids are transfer Krull monoids over a finite abelian group.

However, a transfer Krull monoid need neither be commutative nor v-noetherian nor completely integrally closed. To give a noncommutative example, let 𝒪 be a holomorphy ring in a global field K, A a central simple algebra over K, and H a classical maximal 𝒪-order of A such that every stably free left R-ideal is free. Then H is a transfer Krull monoid over a ray class group of 𝒪 [Citation25, Theorem 1.1]. Let R be a bounded HNP (hereditary noetherian prime) ring. If every stably free left R-ideal is free, then its multiplicative monoid of cancelative elements is a transfer Krull monoid [Citation26, Theorem 4.4]. A class of commutative weakly Krull domains which are transfer Krull but not Krull will be given in [Citation10, Theorem 5.8]. Extended lists of commutative Krull monoids and of transfer Krull monoids, which are not commutative Krull, are given in [Citation6].

Let G0G be a non-empty subset. For a sequence S=g1g(G0), we call

k(S)=i=1l1ord(gi)0thecrossnumberofS, and K(G0)=max{k(S)S𝒜(G0)}thecrossnumberofG0.

They were introduced by U. Krause in 1984 (see [Citation19]) and were studied under various aspects. For the relevance with the theory of non-unique factorizations, see [Citation21, Citation20, Citation22, Citation23] and [Citation7, Chapter 6].

Suppose GCq1Cqr, where r* is the total rank of G and q1,,qr are prime powers, and set

K(G)=1exp(G)+i=1rqi1qi.

It is easy to see that K*(G)≤K(G) and there is known no group for which inequality holds. For further progress on K(G), we refer to [Citation5, Citation15, Citation16, Citation18].

Lemma 2.1.

If G is a p-group, then K(G) = K*(G)<r(G).

Proof.

See [Citation7, Theorem 5.5.9].

A subset G0G is called half-factorial if Δ(G0) = . Otherwise, G0 is called non-half-factorial. Furthermore, the set G0 is called

  • minimal non-half-factorial if it is non-half-factorial and every proper subset G1G0 is half-factorial.

  • an LCN-set if k(A)≥1 for all A𝒜(G0).

We collect some easy or well known results which will be used throughout the manuscript without further mention.

Lemma 2.2.

Let G0G be a non-empty subset. Then

  1. G0 is half-factorial if and only if k(A) = 1 for all A𝒜(G0).

  2. If G0 is an LCN-set, then minΔ(G0)|G0|2.

  3. (G0) has accepted elasticity.

  4. ρ(G)=D(G)2.

  5. Δ(G) is an interval with minΔ(G) = 1.

  6. If B(G), then ρ(L(Bk))≥ρ(L(B)) for every k.

  7. If A𝒜(G), then {exp(G),exp(G)k(A)}⊂L(Aexp(G)).

Proof.

1. follows from [Citation7, Proposition 6.7.3] and 2. from [Citation7, Lemma 6.8.6].

2. see [Citation7, Proposition 6.7.3, Lemma 6.8.6].

3. follows from [Citation7, Theorem 3.1.4] and 4. from [Citation7, Section 6.3].

4. See [Citation7, Theorem 3.1.4, Section 6.3].

5. See [Citation11].

6. Let B(G) and k. Then maxL(Bk)≥kmaxL(B) and minL(Bk)≤kminL(B). It follows that ρ(L(Bk))=maxL(Bk)minL(Bk)kmaxL(B)kminL(B)=ρ(L(B)).

7. Let A𝒜(G) and suppose A=g1g, where and g1,,gG. Then

Aexp(G)=i=1(giord(gi))exp(G)ord(gi).

Since A, g1ord(g1),,gord(g) are atoms, we obtain {exp(G),exp(G)k(A)}⊂L(Aexp(G)).

We need the following lemma.

Lemma 2.3.

Let e1,,erG be independent elements with the same order n, where r, n≥2. Then

  1. Δ({e1++er,e1,,er})={r1}.

  2. If nr+1, then

    Δ({(e1++er),e1,,er})={|nr1|}.

  3. If n = r+1, then

    minΔ({(e1++er),e1++er,e1,,er})=r1.

Proof.

1. See [Citation7, Proposition 6.8.2].

2. See [Citation7, Proposition 4.1.2.5].

3. Let g=e1++er and G0={g,g,e1,,er}. Let B(G0) and assume

B=U1Uk=V1V, where k,N and U1,,Uk,V1,,V𝒜(G0).

Let

I1={i[1,k]vg(Ui)>0 and vg(Ui)=0},I2={i[1,k]Ui=g(g)}J1={j[1,]vg(Vj)>0 and vg(Vj)=0},J2={j[1,]Vj=g(g)}.

Then for every iI1, k(Ui)=1+nvg(Ui)n(n2) and for every jJ1, k(Vj)=1+nvg(Vj)n(n2).

Note that for every i[1,k](I1I2) and every j[1,](J1J2), k(Ui)=k(Vj)=1. Therefore

k(B)=k|I1||I2|+iI1(1+(nvg(Ui))(n2)n)+iI22n=k+(n2)|I1|n2nvg(B)
and
k(B)=|J1||J2|+jJ1(1+(nvg(Vj))(n2)n)+jJ22n=+(n2)|J1|n2nvg(B).

It follows that k=(n2)(|J1||I1|) and hence n−2 | minΔ(G0). By 1., we obtain n−2 = r−1 = minΔ(G0).

If d and ,M0, then a finite subset L is called an almost arithmetical progression (AAP for short) with difference d, length , and bound M if

(1) L=y+(L{0,d,,d}L)y+d(1)
where y, L⊂[−M,−1], and Ld+[1,M].

Lemma 2.4.

There exist constants M1,M2 such that for every A(G) with Δ(supp(A))≠, the set L(AM1) is an AAP with difference minΔ(supp(A)), length at least 1, and bound M2.

Proof.

See [Citation7, Theorem 4.3.6].

Next, we recall the definition of the invariants Δ*(G) and Δ1(G) (see [Citation7, Definition 4.3.12]) in the Characterization Problem.

Let

Δ(G)={minΔ(G0)G0G is a subset with Δ(G0)}.

We define

m(G)=max{minΔ(G0)G0Gis an LCN-set withΔ(G0)},
and we denote by Δ1(G) the set of all d with the following property:
  • For every k, there exists some L(G) which is an AAP with difference d and length k.

Lemma 2.5.

Let k be maximal such that G has a subgroup isomorphic to Cexp(G)k. Then

  1. Δ(G)Δ1(G){dΔ(G)there exists dΔ(G) such that d|d}.

  2. maxΔ(G)=maxΔ1(G)=max{exp(G)2,r(G)1}.

  3. m(G)max{r(G)1,exp(G)21}.

  4. {1,r(G)1,exp(G)2}Δ(G)Δ1(G)[1,max{r(G)1,exp(G)21}][max{1,exp(G)k1,exp(G)2}].

  5. Δ1(G) is an interval if and only if Δ*(G) is an interval if and only if r(G)+k≥exp(G)−1 or GC2r(G)+2r(G).

Proof.

1. Follows from [Citation7, Corollary 4.3.16]

2. From [Citation14, Theorem 1.1].

3. See [Citation27, Proposition 3.7].

4. Follows from 1. and [Citation27, Theorem 1.1].

5. If Δ1(G) is an interval, then exp(G)k2max{r(G)1,exp(G)21} by 4 which implies that Δ*(G) is an interval by Zhong [Citation27, Theorem 1.1.2]. If Δ*(G) is an interval, then Δ1(G) is an interval by 1.. It follows by Zhong [Citation27, Theorem 1.1.2] that Δ*(G) is an interval if and only if r(G)+k≥exp(G)−1 or GC2r(G)+2r(G).

Lemma 2.6.

Let G0G with minΔ(G0)≥⌊exp(G)∕2⌋. Then

  1. Let A𝒜(G0) with k(A)<1. Then k(A)=exp(G)minΔ(G0)exp(G) and ord(h) = exp(G), vh(A) = 1 for all hsupp(A).

  2. Let A𝒜(G0) with k(A)≥1. Then supp(A) is an LCN-set and

    minΔ(supp(A))|supp(A)|2.

  3. If G0 is a minimal non-half-factorial LCN-set with minΔ(G0)=maxΔ(G), then |G0| = r(G)+1 and for every hG0, we have r(⟨G0∖{h}⟩) = r(G).

  4. Let B(G) with ρ(L(B)) = ρ(G). Suppose B=U1Uk=V1V, where k = minL(B),  = maxL(B), and U1,,Uk,V1,V are atoms. Then k(Ui)≥1 for all i∈[1,k] and k(Vj)≤1 for all j∈[1,].

Proof.

1. Since {exp(G),exp(G)k(A)}⊂L(Aexp(G)), we have

minΔ(supp(A))|exp(G)exp(G)k(A)
and hence
minΔ(supp(A))exp(G)exp(G)k(A)exp(G)2.

Since minΔ(G0) | minΔ(supp(A)) and minΔ(G0)≥⌊exp(G)∕2⌋, it follows that

12(exp(G)2)<exp(G)2minΔ(G0)minΔ(supp(A))exp(G)exp(G)k(A)exp(G)2.

Therefore minΔ(G0) = minΔ(supp(A)) = exp(G)−exp(G)k(A) and hence k(A)=exp(G)minΔ(G0)exp(G).

Let gsupp(A) such that ord(g)vg(A)=min{ord(h)vh(A)hsupp(A)}. Then Aord(g)vg(A)=gord(g)B1, where B1(supp(A)). If there exists an atom A1 with k(A1)<1 such that A1 divides B1, then k(A1)=exp(G)minΔ(G0)exp(G)=k(A). Therefore 1+maxL(B1)<ord(g)vg(A) and hence exp(G)2minΔ(G0)ord(g)vg(A)2. It follows by the minimality of ord(g)vg(A) that ord(h) = exp(G) and vh(A) = 1 for all hsupp(A).

2. Assume to the contrary that supp(A) is not an LCN-set. Then there exists A1𝒜(supp(A)) with k(A1)<1. It follows by 1 that vh(A1)=1 for all hsupp(A1) which implies that A1 | A, a contradiction. Thus supp(A) is an LCN-set and hence minΔ(supp(A))≤|supp(A)|−2.

3. By Geroldinger and Zhong [Citation14, Lemma 4.2], we have |G0| = r(G)+1 and r(⟨G0⟩) = r(G). If h∈⟨G0∖{h}⟩, then r(⟨G0∖{h}⟩) = r(G). Otherwise let d = min{kkh∈⟨G0h⟩} and hence (G0∖{h})∪{dh} is also a minimal non-half-factorial LCN-set with minΔ(G0)=maxΔ(G) by Geroldinger and Halter-Koch [Citation7, Lemma 6.7.10]. It follows that r(⟨G0∖{h}⟩) = r(G).

4. Assume to the contrary that there exists i∈[1,k], say i = 1, such that k(U1)<1. Then exp(G)k(U1)<exp(G). Since exp(G)k(U1)L(U1exp(G)), we infer

minL(Bexp(G))exp(G)k(U1)+exp(G)(k1)<exp(G)k.

It follows by maxL(Bexp(G))≥exp(G) that ρ(G)=ρ(L(B))ρ(L(Bexp(G)))>k=ρ(L(B)), a contradiction.

Assume to the contrary that there exists j∈[1,], say j = 1, such that k(V1)>1. Then exp(G)k(U1)>exp(G). Since exp(G)k(U1)L(U1exp(G)), we infer

maxL(Bexp(G))exp(G)k(U1)+exp(G)(1)>exp(G).

It follows by minL(Bexp(G))≤exp(G)k that ρ(G)=ρ(L(B))ρ(L(Bexp(G)))>k=ρ(L(B)), a contradiction.

3. The invariants ρ(G,d), ρ*(G,d), and K(G,d)

First, we introduce new invariants which play a crucial role in the proof of Theorem 1.2 (see Proposition 3.5).

Definition 3.1.

Let dΔ1(G) and k. We define

ρ(G,d,k)=sup{ρ(Lk)Lk(G) is an AAP with difference d and length at leastk}1.

Then (ρ(G,d,k))k=1 is a decreasing sequence of positive real numbers and hence converges. We denote by ρ(G,d) the limit of (ρ(G,d,k))k=1.

It follows by Geroldinger and Zhong [Citation13, Theorem 3.5] that ρ(G,1) = ρ(G) if and only if G is not a cyclic group of order 4,6 or 10.

Lemma 3.2.

Let G0G be a subset with Δ(G0)≠. For every B(G0) with minΔ(G0)∈Δ(L(B)), we have ρ(G,minΔ(G0))≥ρ(L(B)).

Proof.

Let B(G0) with minΔ(G0)∈Δ(L(B)). By definition, L(B) is an AAP with difference minΔ(G0) and length at least 1. Therefore for every k, L(Bk) is an AAP with difference minΔ(G0) and length at least k. Thus for every k, ρ(G,minΔ(G0),k)ρ(L(Bk))ρ(L(B)) by Lemma 2.2.6. Therefore ρ(G,minΔ(G0))≥ρ(L(B)).

Definition 3.3.

Let dΔ1(G). We define

ρ(G,d)=max{ρ(G0)G0G is a non-half-factorial subset with d divides minΔ(G0)}, andK(G,d)=max{K(G0)G0G is a non-half-factorial subset with d divides minΔ(G0)}.

Note that ρ*(G,1) = ρ(G) and K(G,1) = K(G). For every dΔ1(G), there always exists G0G with G0 non-half-factorial such that d | minΔ(G0) by Lemma 2.5.1. Since ρ(G0)>1 and K(G0)≥1, we have ρ*(G,d)>1 and K(G,d)≥1. Furthermore, there exist G1,G2G with d | minΔ(G1) and d | minΔ(G2) such that ρ(G,d)=ρ(G1) and K(G,d) = K(G2).

Lemma 3.4.

Let dΔ1(G). Then

  1. ρ*(G,d)≥ρ(G,d).

  2. ρ(G,d)ρ(G,k0d) for some k0 with k0dΔ1(G).

  3. ρ(G,d)=max{ρ(G,kd)k and kdΔ1(G)}.

Proof.

1. By the definition of Δ1(G) and dΔ1(G), for every k, we let Bk(G) be such that L(Bk) is an AAP with difference d and length at least k. Let

k=min{minLkLk(G) is an AAP with difference d and length at least k}
and hence limkk= by ρ(G) is finite.

By Lemma 2.4, there exists a constant M such that minΔ(supp(B))∈Δ(L(BM)) for all B(G) with Δ(supp(B))≠. Since limkk=, we let k be large enough such that minL(Bk)≥M|𝒜(G)| and assume

Bk=U1t1Usts=V1VmaxL(Bk),
where s,t1,,ts, t1++ts=minL(Bk), U1,,Us are pair-wise distinct atoms, and V1,,VmaxL(Bk) are atoms.

Since minL(Bk)≥M|𝒜(G)|, we infer there exists i∈[1,s], say i = 1, such that t1M. Set I = {i∈[1,s]∣tiM} and G0=supp(iIUi). It follows by the choice of M that

minΔ(G0)Δ(L(iIUiM)).

Since L(Bk) is an AAP with difference d and iIUiM) is a subsequence of Bk, we obtain

d|gcdΔ(L(Bk))|gcdΔ(L(iIUiM)).

Therefore d | minΔ(G0) and hence ρ(G0)ρ(G,d). Since |iIUiti|M|𝒜(G)|D(G), there exists J⊂[1,maxL(Bk)] with |J|≥maxL(Bk)−M|𝒜(G)|D(G) such that jJVj divides iIUiti. It follows by minL(iIUiti)=iIti that

maxL(Bk)M|𝒜(G)|D(G)|J|maxL(iIUiti)(iIti)ρ(G0)minL(Bk)ρ(G,d).

Therefore

ρ(L(Bk))ρ(G,d)+M|𝒜(G)|D(G)minL(Bk)ρ(G,d)+M|𝒜(G)|D(G)k.

By definition, we infer ρ(G,d,k)ρ(G,d)+M|𝒜(G)|D(G)k which implies that ρ(G,d)≤ρ*(G,d).

2. Let G0G with d | minΔ(G0) and ρ(G0)=ρ(G,d). Then there exists B(G0) such that ρ(L(B)) = ρ*(G,d). Since supp(B)⊂G0, we infer minΔ(G0) divides minΔ(supp(B)) and hence d divides minΔ(supp(B)).

By Lemma 2.4, there exists a constant M such that minΔ(supp(B))∈Δ(L(BM)). It follows by Lemma 3.2 and Lemma 2.2.6 that

ρ(G,minΔ(supp(B)))ρ(L(BM))ρ(L(B))=ρ(G,d).

Thus the assertion follows by d divides minΔ(supp(B)).

3. For every k such that kdΔ1(G), we have ρ(G,kd)ρ(G,kd)ρ(G,d) by 1.. Therefore ρ(G,d)max{ρ(G,kd)k and kdΔ1(G)}. It follows by 2. that ρ(G,d)=max{ρ(G,kd)k and kdΔ1(G)}.

Proposition 3.5.

Suppose (G) = (G) for some finite abelian group G and let dΔ1(G). Then dΔ1(G), ρ(G,d) = ρ(G,d), and ρ(G,d)=ρ(G,d).

Proof.

Since (G) = (G), it follows by definition that Δ1(G)=Δ1(G) and ρ(G,kd) = ρ(G,kd) for every k such that kdΔ1(G). By Lemma 3.4.3, we obtain ρ(G,d)=ρ(G,d).

Lemma 3.6.

Let dΔ1(G). Then ρ*(G,d)≥K(G,d). In particular, if d∈[1,r−1], then ρ(G,d)K(G,d)1+(n11)dn11+d2.

Proof.

Suppose G0G with d dividing minΔ(G0) and K(G0) = K(G,d). Then there exists A𝒜(G0) such that k(A) = K(G,d)≥1. Since

{exp(G),exp(G)k(A)}L(Aexp(G)),
we have
ρ(G,d)ρ(G0)ρ(L(Aexp(G)))exp(G)k(A)exp(G)=k(A)=K(G,d).

In particular, if d∈[1,r−1], we let e1,,ed+1 be independent elements of order n1. Set e0=e1++ed+1 and G0={e0,e1,,ed+1}. Then minΔ(G0) = d by Lemma 2.3.1.

Since A=e0e1n11ed+1n11 is an atom with k(A)=1+(n11)dn1, it follows that

K(G,d)K(G0)k(A)1+(n11)dn11+d2.

Lemma 3.7.

Suppose rnr2+1. Let G0G be a subset with d | minΔ(G0), where dΔ1(G) satisfies r1dnr2. If A𝒜(G0) with k(A)>1, then k(A)=1+sdnrd for some s.

In particular,

  1. K(G,r1)=1+s(r1)n1 for some sn1(ui=1u1piki), where n1=p1k1puku with u,k1,,ku and p1,,pu are pairwise distinct primes.

  2. if nr is a prime power, then K(G,r1)=1+(n11)(r1)n1<r.

  3. if n1 is not a prime power, then K(G,r−1)>r.

Proof.

Let t = rd and we start with the following claim.

Claim A.

Let B(G0) with vg(B)0(modnt) for each gsupp(B) and supp(B) is an LCN-set. Then there exists B0(G0) with B0 is a product of atoms having cross number 1 and vg(B0)0(modnt) for each gsupp(B0) such that BB0 is a product of atoms having cross number 1.

Proof of Claim A.

Assume to the contrary that there exists a B(G0) with vg(B)0(modnt) for each gsupp(B) and supp(B) is an LCN-set, such that the assertion does not hold. Suppose |supp(B)| is minimal in all the counterexamples.

Set G1 = supp(B). If for all gG1, ord(g)∣vg(B), then B is a product of atoms having cross number 1, a contradiction. Therefore there exits g0G1 such that ord(g0)vg0(B). Since vg(B)0(modnt) for each gsupp(B), we infer vg0(B)g0{ntggG1{g0}} and ntnr.

Let nrnt=q1s1qvsv, where v,s1,,sv and q1,,qv are pairwise distinct primes. Let i∈[1,v] and Hi={nrntqisintggG1{g0}}. Then

nrntqisivg0(B)g0Hi
and Hi is an qi-group of rank r(Hi)≤rt = d which implies that there exists EG1{g0} with |E|≤r(Hi)≤d such that
nrntqisivg0(B)g0{nrntqisintggE}.

Therefore there exists Bi(G1) such that

|supp(Bi)|d+1,vg0(Bi)=nrntqisivg0(B), and nt|vg(Bi) for every gG1{g0}.

Since supp(Bi)G1 is an LCN-set, we have

minΔ(supp(Bi))|supp(Bi)|2d1.

Note that d|minΔ(G0)|minΔ(supp(Bi)). We infer minΔ(supp(Bi)) = 0 and hence supp(Bi) is half-factorial. Therefore Bi is a product of atoms having cross number 1.

Since gcd({vg0(Bi)i[1,v]})=vg0(B), there exist x1,,xv such that

vg0(BB1x1Bvxv)0(modord(g0)).

Therefore BB1x1Bvxv=g0yord(g0)C for some y and C(G1{g0}). Note nt|vg(Bi) for every i∈[1,v] and every gG1{g0}. Thus nt|vg(C) for each gsupp(C). Since |supp(C)|<|supp(B)|, it follows by the minimality of |supp(B)| that there exists C0(supp(G0)) satisfying C0 is a product of atoms having cross number 1 and nt|vg(C0) for each gsupp(C0), such that CC0 is a product of atoms having cross number 1. Let B0=C0B1x1Bt1xt1. Then BB0 is a product of atoms having cross number 1, a contradiction to our assumption.

Let A𝒜(G0) be with k(A)>1. Then Lemma 2.6.2 implies that supp(A) is an LCN-set. Set B=Ant and hence Claim A implies that there exist atoms W1,,W𝒜(G0) having cross number 1, where 0, such that AntW1W is a product of atoms having cross number 1. Therefore {ntk(A)+,nt+}L(AntW1W). Since d | minΔ(G0), we infer d|(ntk(A)nt). It follows that k(A)=1+sdnt for some s.

Now we begin to prove the “in-particular’’ parts.

1. For every j∈[1,u] and every m, we denote by ∥mj the least positive residue of m modulo pjkj, that is, mj[1,pjkj] and mjm(modpjkj).

By definition of K(G,r−1), we have K(G,r1)=1+sdn1 for some s. Let H be a subgroup of G with

HCn1rCp1k1r1Cpukur1Cn1
and let (e1,1,,e1,r1,,eu,1,,eu,r1,e) be a basis of H with ord(e) = n1 and ord(ej,i)=pjkj for all i∈[1,r−1], j∈[1,u]. Set e0=e+j=1ui=1r1ej,i and G2={e1,1,,e1,r1,,eu,1,,eu,r1,e,e0}. For every W𝒜(G2), we have
W=e0ve0(W)en1ve0(W)i[1,r1],j[1,u]ej,ipjkjve0(W)j
and
k(W)=1+j=1upjkjve0(W)jpjkj(r1).

We suppose that U1U1=V1V2 with 21=minΔ(G2), where 1,2 and U1,,U1, V1,,V2𝒜(G2). Then

k(U1U1)=1+x=11j=1upjkjve0(Ux)jpjkj(r1)=2+y=12j=1upjkjve0(Vy)jpjkj(r1).

Since ve0(U1U1)=ve0(V1V2), we infer

x=11ve0(Ux)jy=12ve0(Vy)j(modpjkj) for each j[1,u].

Therefore

x=11pjkjve0(Ux)jpjkjy=12pjkjve0(Vy)jpjkj=12x=11ve0(Ux)jy=12ve0(Vy)jpjkj
for each j∈[1,u], whence
x=11j=1upjkjve0(Ux)jpjkjy=12j=1upjkjve0(Vy)jpjkj.

Since minΔ(G2)=21, we infer r−1 | minΔ(G2) which implies that K(G,r−1)≥K(G2). Let W1𝒜(G2) be the atom with ve0(W1)=1. Then k(W1)=1+(ui=1u1piki)(r1).

Since K(G,r1)=1+s(r1)n1k(W1)=1+(ui=1t1piki)(r1) for some s, it follows that sn1(ui=1u1piki).

2. If nr is a prime power, then K(G)<r by Lemma 2.1. It follows by 1. that r>K(G,r1)=1+s(r1)n11+(n11)(r1)n1 for some s. Therefore K(G,r1)=1+(n11)(r1)n1.

3. If n1 is not a prime power, then u≥2 and ui=1u1piki>uu21. It follows by 1. that K(G,r−1)>1+r−1 = r. □

Proposition 3.8.

Let s be maximal such that G has a subgroup isomorphic to Cnrs. Suppose dΔ1(G) with dmax{nr2,r+12}.

  1. If dr, then ρ(G,d)=nrnrd.

  2. If dnr−1, then ρ*(G,d) = K(G,d).

  3. Suppose r = nr−1≥3. If n1=nr=pk for some prime p and k, then K(G,r1)<ρ(G,r1)=1+(r+1)(r1)r+2<r. Otherwise, K(G,r−1) = ρ*(G,r−1).

Proof.

1. Since dmax{nr2,r+12} and dr, it follows by Lemma 2.5(items 3 and 4) that d>m(G) and d[nrs1,nr2].

Let e1,,enrd1 be independent elements with order nr and let e0=e1enrd1. Then minΔ({e0,e1,,enrd1})=|nr(nrd1)1|=d by Lemma 2.3.2. Since A=e0e1enrd1 is an atom, we infer L(Anr)={nrd,nr}. Therefore

ρ(G,d)ρ({e0,e1,,enrd1})ρ(L(Anr))=nrnrd.

Let G0G with d | minΔ(G0) be such that ρ(G0)=ρ(G,d). Then there exists B(G0) with ρ(L(B)) = ρ*(G,d). Set

B=U1Uk=V1V,
where k = minL(B),  = maxL(B), and U1,,Uk,V1,,V𝒜(G0) are atoms.

If there exists i∈[1,k] such that k(Ui)>1, then Lemma 2.6.2 implies that supp(Ui) is an LCN-set and hence minΔ(supp(Ui))≤m(G). Since

d|minΔ(G0)|minΔ(supp(Ui)),
we get a contradiction to d>m(G). Therefore k(Ui)≤1 for all i∈[1,k].

If there exists i∈[1,] such that k(Vi)<1, then Lemma 2.6.1 implies that k(Vi)=nrdnr. Therefore k(Vi)nrdnr for all i∈[1,]. It follows that

ki=1kk(Ui)=k(B)=i=1k(Vi)nrdnr.

Then ρ(G,d)=ρ(L(B))=knrnrd and hence ρ(G,d)=nrnrd.

2. Let G0G be such that d | minΔ(G0) and ρ(G0)=ρ(G,d). If there exists an atom A𝒜(G0) such that k(A)<1, then Lemma 2.6.1 implies that k(A)=nrdnr1nr, a contradiction to |A|≥2. Thus G0 is an LCN-set. Let B(G0) such that ρ(L(B))=ρ(G0)=ρ(G,d). Then

minL(B)K(G,d)k(B)maxL(B)
which implies that ρ*(G,d) = ρ(L(B))≤K(G,d).

Let G0G be such that d | minΔ(G0) and K(G0) = K(G,d). Then there exists an atom A𝒜(G0) such that k(A) = K(G,d)≥1. Since {nr,nrk(A)}L(Anr), we infer

ρ(G,d)ρ(G0)ρ(L(Anr))k(A)=K(G,d)
and hence ρ*(G,d) = K(G,d).

3. Let r = nr−1≥3 and we proceed to prove the following claim.

Claim B.

Suppose ρ*(G,r−1)>K(G,r−1) and let G0G be such that (r−1) | minΔ(G0) and ρ(G0)>K(G,r−1).

  1. There exists gG0 with ord(g) = nr such that −gG0.

  2. Let G2G0{g,g} with |G2| = r. If there exists a∈[1,nr−1] such that ag∈⟨G2⟩, then g=hG2h, GCnrr, and G2 is a basis of G.

  3. GCnrr, G0={e1,,er,g,g}, where g=e1++er and (e1,,er) is a basis of G, and ρ(G0)=1+nr(r1)nr+1. In particular, ρ(G,r1)=1+nr(r1)nr+1.

Proof of Claim B.

By Lemma 2.5.2, we infer that minΔ(G0)=r1=nr2=maxΔ(G).

a. If G0 is an LCN-set, then for every B(G0), we have

minL(B)K(G,r1)k(B)maxL(B)
which implies that ρ(L(B))≤K(G,r−1). Therefore ρ(G0)≤K(G,r−1), a contradiction. Thus there exists A𝒜(G0) such that k(A)<1. Lemma 2.6.1 implies that A = g(−g) for some gG0 with ord(g) = nr. Hence {g,−g}⊂G0.

b. Let EG2 be minimal such that there exists a∈[1,nr−1] such that ag∈⟨E⟩ and let dg[1,nr1] be minimal such that dgg∈⟨E⟩. Then there exists an atom V𝒜(E∪{g}) with vg(A)=dg and |supp(V)|≤r+1. Let V=gdgT, where T(E). Then

V(g)nr=(g(g))dg((g)nrdgT), where L((g)nrdgT)[1,nrdg].

Note that for each L((g)nrdgT), we have r−1 | dg+−2. Therefore L((g)nrdgT)={nrdg} or (dg = 1 and L((g)nr1T)={1}). We distinguish two cases.

Suppose L((g)nrdgT)={nrdg}. Let T=T1Tnrdg such that (−g)Ti, i[1,nrdg], are atoms, where T1,,Tnrdg(E). Thus −g∈⟨E⟩ which implies that dg = 1 by the minimality of dg. The minimality of E implies that supp(Ti) = E for each i∈[1,nr−1]. Then for every hE, vh(V)nr1. Therefore ord(h) = nr and vh(V)=nr1. It follows that

T1==Tnr1=hEh.

If k((−g)T1)<1, then Lemma 2.6.1 implies that T1 = g, a contradiction. Therefore

k((g)T1)=1nr+hE1ord(h)=1+|E|nr1.

Since |E|≤|G2|≤r and r = nr−1, we have |E| = |G2| = r. Let G2={e1,er}. Then V=ge1nr1ernr1 implies that (e1,,er) is a basis of G, GCnrr, and g=e1++er.

Suppose dg = 1 and L((g)nr1T)={1}. Note that V = gT and hence |T|≥2. We infer k((g)nr1T)>1. It follows by Lemma 2.6.2 that {−g}∪E is an LCN-set and minΔ({−g}∪E)≤|E|−1. Since (r−1) | minΔ({−g}∪E) and |E|≤|G2| = r, we have

|E|=|G2|=r and minΔ({g}G2)=r1.

Let E1{g}G2 be a minimal non-half-factorial LCN-set. Then

minΔ(E1)=minΔ({g}G2)=r1=maxΔ(G).

It follows by Geroldinger and Zhong [Citation14, Lemma 4.2] that |E1| = r+1 and hence E1={g}G2. Therefore

{g}G2 is a minimal non-half-factorial LCN-set.

Let G2={e1,,er} and assume

V=ge1k1erkr,V1=(g)nr1e1k1erkr,
where ki[1,ord(ei)1] for each i∈[1,r].

If k(V)>1, then Lemma 2.6.2 and [Citation14, Lemma 4.5] imply

{g}G2 is a minimal non-half-factorial LCN-set with minΔ({g}G2)=maxΔ(G)=r1.

By the minimality of E = G2, we have for every m∈[1,nr−1] and every hG2, mg∉⟨G2∖{h}⟩. Note that nr≥4. Let I={i[1,r]kiord(ei)2}. Then [Citation14, Lemma 4.4.1] implies that

W1=g2iIei2kiord(ei)i[1,r]Iei2ki
and
W2=(g)nr2iIei2kiord(ei)i[1,r]Iei2ki
are both atoms. Since V2=iIeiord(ei)W1 and V12=(g)nriIeiord(ei)W2, we infer r−1 | |I|−1 and r−1 | |I|. Hence r = 2, a contradiction to r≥3.

Thus k(V) = 1 which implies that k1==kr=1 and ord(ei)=nr for each i∈[1,r]. It follows by k(V1)>1 and [Citation14, Lemma 4.3.2] that (g)e1nr1ernr1 is also an atom. Therefore e1,,er are independent and g=e1++er. Since r(G) = r and exp(G) = nr, we infer (e1,,er) is a basis of G and GCnrr.

c. If for all W𝒜(G0), k(W)≤1, then for every B(G0), we have

minL(B)k(B)maxL(B)2nr
which implies that ρ(G0)nr2. It follows by Lemma 3.6 that K(G,r1)1+r12=nr2, a contradiction to ρ(G0)>K(G,r−1).

Let W𝒜(G0) with k(W)>1. Then supp(W) is an LCN-set with minΔ(supp(W)) = r−1 by Lemma 2.6.2. Let G1supp(W) be a minimal non-half-factorial subset. Then minΔ(G1) = r−1 and hence |G1| = r+1 by Geroldinger and Zhong [Citation14, Lemma 4.2.1]. Since {g,−g}⊄supp(W), we choose hG1 such that {g,−g}∩(G1∖{h}) = . Lemma 2.6.3 implies r(⟨G1∖{h}⟩) = r. Thus there exists a∈[1,nr−1] such that ag∈⟨G1∖{h}⟩. Since |G1∖{h}| = r, it follows by a and b that GCnrr and there exists a basis (e1,,er) of G such that g=e1++er and {e1,,er,g,g}G0.

Assume to the contrary that there exists h0G0{e1,,er,g,g}. After renumbering if necessary, we may assume that h0=k1e1++ktet, where t∈[1,r], ki[1,nr1] for each i∈[1,t] and k1=min{k1,kt}. Thus k1g{h0,e2,,er}. It follows by b that g=h0+e2++er and hence h0=e1, a contradiction. Therefore G0={e1,,er,g,g}.

We only need to prove ρ(G0)=1+nr(r1)nr+1 which immediately implies that ρ(G,r1)=1+nr(r1)nr+1.

Since (ge1nr1ernr1)nr(g)nr=(g(g))nr(e1nr)nr1(ernr)nr1, we obtain

ρ(G0)nr+r(nr1)nr+1=1+nr(r1)nr+1.

Let B(G0) such that ρ(L(B)) = ρ(G0) and assume that

B=U1Uk=V1V, and k=ρ(G0),
where k, and U1,,Uk,V1,,V are atoms.

Note that 𝒜(G0)=𝒜(G0{g}){g(g),(g)e1er,(g)nr}. By Lemma 2.6.4, we have k(Ui)≥1 for all i∈[1,k] and k(Vj)≤1 for all j∈[1,]. Since ((g)e1er)nr=(g)nri[1,r]einr and ρ(Bnr)=ρ(G0), substituting B by Bnr, if necessary, we can assume that Ui(g)e1er, Vj(g)e1er for each i∈[1,k] and each j∈[1,].

Since there must exist j0∈[1,] such that Vj0=g(g), there must exists i0∈[1,k] such that Ui0=(g)nr which implies that Vj(g)nr for all j∈[1,]. If there exists i1∈[1,k] such that Ui1=gnr, then maxL(B)=nr+maxL(B(Ui0Ui1)1) which implies that

ρ(L(B))=nr+maxL(B(Ui0Ui1)1)2+minL(B(Ui0Ui1)1)=ρ(G).

Therefore

ρ(L(B))=nr2=ρ(L(B(Ui0Ui1)1)).

It follows by Lemma 3.6 that K(G,r1)1+r12=nr2, a contradiction to ρ(G0)>K(G,r−1). If there exists j1∈[1,] such that Vj1=gnr, then maxL(B(g)nr)=nr+1 which implies that

ρ(L(B(g)nr))nr+1k>k=ρ(G0),
a contradiction.

To sum up, we obtain

{Uii[1,k]}(A(G0{g}){gnr}){(g)nr}
and
{Vjj[1,]}𝒜(G0{g,g}){(g)g}.

Let I1={i[1,k]Ui=(g)nr}, I2={i[1,k]vg(Ui)1}, and J = {j∈[1,]∣Vj = g(−g)}. Then

nr|I1|=|J|=vg(B)=vg(B)=iI2vg(Ui)|I2|
and
2nr|J|+|J|=k(B)=iI2k(Ui)+k|I2|=iI2(1+nrvg(Ui)nr(nr2))+k|I2|=k+(nr2)|I2|nr2nriI2vg(Ui)=k+(nr2)|I2||J|(nr2)nr.

Therefore k|I1|+|I2|nr+1nr|I2| and k=1+|I2|k(nr2)1+nr(nr2)nr+1. It follows that ρ(G0)=1+nr(r1)nr+1.

We distinguish two cases to finish the proof.

Suppose that GCpkpk1 for some prime p and k with pk≥4. Then K(G,pk2)=1+(pk1)(pk2)pk by Lemma 3.7.2. Let (e1,,epk1) be a basis of G and G0={e1,,epk1,g,g}, where g=i[1,pk1]ei. By Lemma 2.3.3, we have pk2=minΔ(G0). Since

(ge1pk1erpk1)pk(g)pk=(g(g))pk(e1pk)pk1(erpk)pk1,
we obtain
ρ(G,pk2)ρ(G0)pk+(pk1)rpk+1=1+pk(pk2)pk+1>K(G,pk2).

It follows by Claim B that

K(G,pk2)<ρ(G,pk2)=1+pk(pk2)pk+1<pk1.

Suppose that GCpkpk1 for any prime p and any k. Assume to the contrary that K(G,r−1)<ρ*(G,r−1). Then Claim B implies that GCnrr and ρ(G,r−1)<r. Since nr is not prime power, it follows by Lemma 3.7.3 that K(G,r−1)>r, a contradiction. □

4. Proof of main theorems

From now on, we assume G is a further finite abelian group with GCn1Cnr, where r,n1,,nr and 1<n1||nr. For convenience, we collect some necessary results which will be used all through the following two sections without further mention.

Lemma 4.1.

Suppose (G) = (G) and dΔ1(G). Then

  1. If G is isomorphic to a subgroup of G, then GG.

  2. max{r(G)1,exp(G)2}=max{r(G)1,exp(G)2}.

  3. dΔ1(G) and ρ(G,d)=ρ(G,d).

Proof.

1. By Geroldinger and Halter-Koch [Citation7, Proposition 7.3.1.3], we have D(G) = D(G). It follows from [Citation7, Proposition 5.1.3.2 and 5.1.11.1] that GG.

2. follows from [Citation7, Corollary 4.3.16] and [Citation14, Theorem 1.1.3].

3. See Proposition 3.5.

Theorem 4.2.

Suppose (G) = (G). Then

  1. If rnr−1 and n1≠2, then r(G)=r(G)exp(G)1.

  2. If rnr−1, n1≠2, and nr is a prime power , then r(G) = r(G) and n1=n1.

  3. If rnr−3, then exp(G) = exp(G) and r(G)nr3.

  4. If nr2+1rnr3, then exp(G) = exp(G) and r(G) = r(G).

  5. If rnr−3 and Δ*(G) is an interval, then exp(G) = exp(G) and r(G) = r(G).

Proof.

1. Assume to the contrary that r(G)≠r(G) = r. Since r1=max{r(G)1,exp(G)2}=max{r(G)1,exp(G)2}, it follows that r(G)1<exp(G)2=r1. Let d = r−1. Then dΔ1(G)=Δ1(G). Since ρ(G,d)1+(n11)dn1 by Lemma 3.6 and ρ(G,d)=exp(G)2 by Proposition 3.8.1, it follows that ρ(G,d)=exp(G)2=ρ(G,d)1+(n11)(exp(G)2)n1, a contradiction to n1≠2. Thus r1=r(G)1=max{r(G)1,exp(G)2} and hence r(G)exp(G)1.

2. Note that rnr−1≥2. If r = 2, then GC3C3. Therefore D(G) = D(G) = 5 and max{32,21}=max{exp(G)2,r(G)1}. It follows that GG. Thus we can assume r≥3.

By 1., we have r(G)=r(G)exp(G)1. Since nr is a prime power and nr≥3, it follows by Lemma 3.7.2 that K(G,r1)=1+nr(r1)nr<r. Then r≥3 and Propositions 3.8.2 and 3.8.2 imply that ρ*(G,r−1)<r and hence ρ(G,r1)=ρ(G,r1)<r. Therefore

1+(n11)(r1)n1K(G,r1)ρ(G,r1)<r
by Lemma 3.6. It follows by Lemma 3.7.1 that K(G,r1)=1+(n11)(r1)n1.

If K(G,r−1) = ρ*(G,r−1) and K(G,r1)=ρ(G,r1), then 1+(nr1)(r1)nr=1+(n11)(r1)n1 which infers n1=nr=n1.

If K(G,r−1)<ρ*(G,r−1) and K(G,r1)=ρ(G,r1), then GCnrr, r = nr−1, and ρ(G,r1)=1+(nr)(r1)nr+1 by Proposition 3.8.3. Therefore 1+(nr)(r1)nr+1=1+(n11)(r1)n1 which infers r+2=nr+1=n1exp(G), a contradiction.

If K(G,r−1) = ρ*(G,r−1) and K(G,r1)<ρ(G,r1), then GCexp(G)r, r = exp(G)−1, and ρ(G,r1)=1+exp(G)(r1)exp(G)+1 by Proposition 3.8.3. Therefore 1+(n11)(r1)n1=1+(exp(G))(r1)exp(G)+1 which infers nrn1=exp(G)+1=r+2, a contradiction.

If K(G,r−1)<ρ*(G,r−1) and K(G,r1)<ρ(G,r1), then GCr+1rG by Proposition 3.8.3.

3. Note that nr≥4. Assume to the contrary that r(G)nr2. Let d=nr3[1,r(G)1]. Therefore dΔ1(G)=Δ1(G) and Lemma 3.6 implies that ρ(G,d)1+d2. Since d≥max{r,⌊nr∕2⌋}, Proposition 3.8.1 implies that ρ(G,d)=nr3<1+d2, a contradiction to ρ(G,d)=ρ(G,d).

Thus r(G)nr3. Since nr2=max{r(G)1,exp(G)2}=max{r(G)1,exp(G)2}, it follows that nr2=exp(G)2.

4. By 3, exp(G) = exp(G) and r(G)exp(G)3. Assume to the contrary that r(G)≠r(G).

Suppose that r(G)>r(G). Choose d = r−1. Therefore dΔ1(G)=Δ1(G) and Lemma 3.6 implies that ρ(G,d)1+r12. Since dmax{r(G),exp(G)2}, Proposition 3.8.1 implies that ρ(G,d)=exp(G)exp(G)dnr4<1+r12, a contradiction to ρ(G,d)=ρ(G,d).

Suppose that r(G)<r(G). Choose d = r∈[1,r(G)−1]. Then dΔ1(G)=Δ1(G) and Lemma 3.6 implies that ρ(G,d)1+r2. Since d≥max{r,⌊nr∕2⌋}, Proposition 3.8.1 implies that ρ(G,d)=nrnrr=1+rnrr<1+r2, a contradiction to ρ(G,d)=ρ(G,d).

5. By 3, we have r(G)+3exp(G)=nr and by 4, we can assume that rnr2 and r(G)nr2. Since Δ*(G) is an interval, we obtain that Δ(G)=Δ1(G)=Δ1(G)=Δ(G) is an interval by Lemma 2.5.5. Let k be maximal such that there exists a subgroup H of G with HCnrk and let k be maximal such that there exists a subgroup H of G with HCnrk. Then 2rr+knr−2 and 2r(G)r(G)+knr2 by Lemma 2.5.5.

Assume to the contrary that rr(G) and by symmetry, we can assume r<r(G). Thus nr21r<r(G)nr2 which implies that nr is even, nr = 2r+2, and r(G) = r+1. Since Δ*(G) and Δ(G) are intervals, it follows by Lemma 2.5.5 that GCnrr and kr. Then G is a subgroup of G which implies that GG by Lemma 4.1.1, a contradiction.

Proof of Theorem 1.2.

By definition of transfer Krull monoids, it follows that (G)=(H)=(H)=(G).

1. Let rn−3. If Δ*(G) is not an interval, then [Citation27, Theorems 1.1 and 1.2] implies that GG.

Suppose Δ*(G) is an interval. Then Theorem 4.2.5 implies that n = exp(G) and r = r(G). Therefore G is isomorphic to a subgroup of G which implies that GG by Lemma 4.1.1.

2. If n = 2, then G is an elementary two-group and the assertion follows by Geroldinger and Halter-Koch [Citation7, Theorem 7.3.3]. We assume n≥3 is a prime power. Then Theorem 4.2.2 implies that r = r(G) and n=n1. Therefore G is isomorphic to a subgroup of G which implies that GG by Lemma 4.1.1.

5. Concluding remarks and conjectures

Throughout this section, let GCnr with D(G)≥4 and n=p1k1puku, where u,k1,,ku and p1,,pu are pair-wise distinct primes.

Conjecture 5.1.

Suppose rn−1. Then the following hold:

  1. K(G,r1)=1+s(r1)n for some s with gcd(s,n) = 1.

  2. K(G,r1)=1+(i=1upiki1piki)(r1).

It is easy to see that C2 implies C1. By Lemma 3.7.1, we have K(G,r1)=1+s(r1)n for some s and if n is a prime power, then C2 holds.

Proposition 5.2.

Let G be a finite abelian group with (G) = (G). If rn−1 and C1 holds, then GG.

Proof.

If n is a prime power, then the assertion follows by Theorem 1.2.

Suppose n is not a prime power. Then rn−1≥5. Let (G) = (G) and let GCn1Cnr with r,n1,,nr and 1<n1||nr. Then Theorem 4.2.1 implies that r=rnr1. By Lemma 3.7.1 and Proposition 3.8(items 2. and 3.), we have ρ(G,r1)=K(G,r1)=1+s(r1)n for some s. If ρ(G,r1)=K(G,r1), then Lemma 3.7.1 implies that ρ(G,r1)=1+s(r1)n1 for some s. Thus ρ(G,r1)=ρ(G,r1) implies that sn=sn1 and hence n|n1 by gcd(s,n) = 1. Therefore G is isomorphic to a subgroup of G which implies that GG by Lemma 4.1.1.

Suppose ρ(G,r1)K(G,r1). Then Lemma 3.6 implies that ρ(G,r1)>K(G,r1) and hence Proposition 3.8(items 2. and 3.) implies that nr=r+1 and ρ(G,r1)=1+(r+1)(r1)r+2. Thus ρ(G,r1)=ρ(G,r1) implies that sn=r+1r+2. Since gcd(s,n) = 1, we have n = r+2, a contradiction.

Recall that K(G)K(G) for all finite abelian group G and there is known no group G with K(G)<K(G).

Proposition 5.3.

Let G be a finite abelian group with (G) = (G). If r≥max{(u−1)n+1,n}≥3 and K(G) = K*(G), then GG.

Proof.

If u = 1, then the assertion follows by Theorem 1.2.2. Suppose u≥2. Note K(G)=K(G)=1n+r(i=1upiki1piki)=1+sn(r1)+i=1upiki1pikin1n, where s=n(i=1upiki1piki). Since r≥(ω(n)−1)n+1, we have i=1upiki1pikin1n<u1r1n. It follows by Lemma 3.7.1 that 1+sn(r1)K(G,r1)=1+sn(r1)1+s+1n(r1), where s. Therefore 1+sn(r1)=K(G,r1). Since s=n(i=1upiki1piki), we have gcd(s,n) = 1. The assertion follows by Proposition 5.2.

References

  • Baginski, P., Geroldinger, A., Grynkiewicz, D., Philipp, A. (2013). Products of two atoms in Krull monoids and arithmetical characterizations of class groups. Eur. J. Combin. 34:1244–1268. Special Issue in memory of Yahya Ould Hamidoune.
  • Chang, G. W. (2011). Every divisor class of Krull monoid domains contains a prime ideal. J. Algebra 336:370–377.
  • Fan, Y., Geroldinger, A., Kainrath, F., Tringali, S. (2017). Arithmetic of commutative semigroups with a focus on semigroups of ideals and modules. J. Algebra Appl. 16(11):42 p.
  • Fan, Y., Tringali, S. Power monoids: a bridge between factorization theory and arithmetic combinatorics. Arxiv:1701.09152.
  • Gao, W., Wang, L. (2012). On the maximal cross number of unique factorization zero-sum sequences over a finite abelian group. Integers 12, Paper A14:8 p.
  • Geroldinger, A. (2016). Sets of lengths. Am. Math. Monthly 123:960–988.
  • Geroldinger, A., Halter-Koch, F. (2006). Non-Unique Factorizations. Algebraic, Combinatorial and Analytic Theory. Pure and Applied Mathematics, Vol. 278. Boca Raton, FL: Chapman & Hall/CRC.
  • Geroldinger, A., Ruzsa, I. (2009). Combinatorial Number Theory and Additive Group Theory. Advanced Courses in Mathematics - CRM Barcelona. Basel, Switzerland: Birkhäuser.
  • Geroldinger, A., Schmid, W. A. A characterization of class groups via sets of lengths. ArXiv:1503.04679.
  • Geroldinger, A., Schmid, W. A., Zhong, Q. (2017). Systems of sets of lengths: transfer Krull monoids versus weakly Krull monoids. In: Fontana, M., Frisch, S., Glaz, S., Francesca Tartarone, F., Zanardo, P., eds. Rings, Polynomials, and Modules. Springer.
  • Geroldinger, A., Yuan, P. (2012). The set of distances in Krull monoids. Bull. London Math. Soc. 44:1203–1208.
  • Geroldinger, A., Zhong, Q. (2017). A characterization of class groups via sets of lengths II. J. Théor. Nombres Bordx. 29:327–346.
  • Geroldinger, A., Zhong, Q. (2018). Long sets of lengths with maximal elasticity. Canad. J. Math., to appear. DOI: 10.4153/CJM-2017-043-..
  • Geroldinger, A., Zhong, Q. (2016). The set of minimal distances in Krull monoids. Acta Arith. 173:97–120.
  • He, X. (2014). Cross number invariants of finite abelian groups. J. Number Theory 136:100–117.
  • Kim, B. (2015). The cross number of minimal zero-sum sequences over finite abelian groups. J. Number Theory 157:99–122.
  • Kim, H., Park, Y. S. (2001). Krull domains of generalized power series. J. Algebra 237:292–301.
  • Kriz, D. (2013). On a conjecture concerning the maximal cross number of unique factorization indexed sequences. J. Number Theory 133:3033–3056.
  • Krause, U., Zahlten, C. (1991). Arithmetic in Krull monoids and the cross number of divisor class groups. Mitt. Math. Ges. Hamb. 12:681–696.
  • Plagne, A., Schmid, W. A. (2005). On the maximal cardinality of half-factorial sets in cyclic groups. Math. Ann. 333:759–785.
  • Plagne, A., Schmid, W. A. (2018). On congruence half-factorial Krull monoids with cyclic class group, submitted.
  • Schmid, W. A. (2005). Differences in sets of lengths of Krull monoids with finite class group. J. Thé Nombres Bordx. 17:323–345.
  • Schmid, W. A. (2009). Arithmetical characterization of class groups of the form ℤ∕nℤ⊕ℤ∕nℤ via the system of sets of lengths. Abh. Math. Semin. Univ. Hamb. 79:25–35.
  • Schmid, W. A. (2009). Characterization of class groups of Krull monoids via their systems of sets of lengths : a status report, number Theory and applications. In: Adhikari, S. D., Ramakrishnan, B., eds. Proceedings of the International Conferences on Number Theory and Cryptography. New Delhi, India: Hindustan Book Agency, pp. 189–212.
  • Smertnig, D. (2013). Sets of lengths in maximal orders in central simple algebras. J. Algebra 390:1–43.
  • Smertnig, D. Factorizations in bounded hereditary noetherian prime rings. Proc. Edinburgh Math. Soc., to appear. ArXiv:1605.09274.
  • Zhong, Q. (2017). Sets of minimal distances and characterizations of class groups of Krull monoids. Ramanujan J., to appear. DOI:10.1007/s11139-016-9873-..