734
Views
4
CrossRef citations to date
0
Altmetric
Research Article

On half-factoriality of transfer Krull monoids

, ORCID Icon, &
Pages 409-420 | Received 17 Jan 2020, Accepted 20 Jul 2020, Published online: 08 Aug 2020

Abstract

Let H be a transfer Krull monoid over a subset G0 of an abelian group G with finite exponent. Then every non-unit aH can be written as a finite product of atoms, say a=u1··uk. The set L(a) of all possible factorization lengths k is called the set of lengths of a, and H is said to be half-factorial if |L(a)|=1 for all aH. We show that, if aH is a non-unit and |L(a(3exp(G)3)/2)|=1, then the smallest divisor-closed submonoid of H containing a is half-factorial. In addition, we prove that, if G0 is finite and |L(gG0g2ord(g))|=1, then H is half-factorial.

MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

Let H be a monoid. If an element aH has a factorization a=u1··uk, where kN and u1,,uk are atoms of H, then k is called a factorization length of a, and the set L(a) of all possible k is referred to as the set of lengths of a. The monoid H is said to be half-factorial if |L(a)|=1 for every aH.

The study of half-factoriality was pioneered by Leonard Carlitz in the setting of algebraic number theory: he proved in [Citation4] that the ring of integers OK of a number field K is half-factorial if and only if the cardinality of its class group is either 1 or 2. After this, the concept of half-factoriality seemed to remain dormant for more than a decade until papers by Abraham Zaks [Citation32], Ladislav Skula [Citation28], and Jan Śliwa [Citation29] simultaneously appeared in 1976. In such papers, half-factoriality was studied in the context of Krull domains and c-monoids. Since then half-factoriality has been investigated in different classes of monoids (see [Citation3, Citation6, Citation17, Citation18]) and integral domains (see [Citation5, Citation7, Citation13, Citation19, Citation20, Citation23, Citation33]).

Given aH, let a={bH|b divides some power of a} be the smallest divisor-closed submonoid of H containing a. Then a is half-factorial if and only if |L(an)|=1 for all nN, and H is half-factorial if and only if c is half-factorial for every cH. It is thus natural to ask:

Does there exist an integer NN such that, if aH and |L(aN)|=1, then a is half-factorial? (Note that, if a is half-factorial for some aH, then of course |L(ak)|=1 for every kN.)

We answer this question affirmatively for transfer Krull monoids over finite abelian groups, and we study the smallest N having the above property (Theorems 1.1 and 1.2).

Transfer Krull monoids and transfer Krull domains are a recently introduced class of monoids and domains including, among others, all commutative Krull domains and wide classes of non-commutative Dedekind domains (see Section 2 and, for a survey, see [Citation10]).

Let H be a transfer Krull monoid over a subset G0 of an abelian group G. Then H is half-factorial if and only if the monoid B(G0) of zero-sum sequences over G0 is half-factorial (in this case, we also say that the set G0 is half-factorial). It is a standing conjecture that every abelian group has a half-factorial generating set, which implies that every abelian group can be realized as the class group of a half-factorial Dedekind domain [Citation11].

Suppose now that H is a commutative Krull monoid with class group G and that every class contains a prime divisor. It is a classic result that H is half-factorial if and only if |G|2, and it turns out that, also for |G|3, half-factorial subsets (and minimal non-half-factorial subsets) of the class group G play a crucial role in a variety of arithmetical questions (see [Citation12, Chapter 6.Citation7], [Citation15]). However, we are far away from a good understanding of half-factorial sets in finite abelian groups (see [Citation25] for a survey, and [Citation21, Citation22, Citation26]). To mention one open question, the maximal size of half-factorial subsets is unknown even for finite cyclic groups [Citation22]. Our results open the door to a computational approach to the study of half-factorial sets.

More in detail, denote by hf(H) the infimum of all NN with the following property:

If aH and |L(aN)|=1, then a is half-factorial.

(Here, as usual, we assume inf=.) We call hf(H) the half-factoriality index of H. If H is not half-factorial, then hf(H) is the infimum of all NN with the property that |L(aN)|2 for every aH such that a is not half-factorial. In particular, if G is an abelian group with |G|3, then hf(B(G)) is the infimum of all NN with the property that

For every sequence S over G, if |L(SN)|=1, then |L(Sk)|=1 for every kN.

Theorem 1.1.

Let H be a transfer Krull monoid over a finite subset G0 of an abelian group G with finite exponent. The following are equivalent.

  1. H is half-factorial.

  2. hf(H)=1.

  3. G0 is half-factorial.

  4. |L(gG0g2ord(g))|=1.

We observe that in general if H is half-factorial, then hf(H)=1. But if H is a transfer Krull monoid over a subset of a torsion free group, then hf(H)=1 does not imply that H is half-factorial (see Example 2.4.1). Furthermore, for every nN, there exists a Krull monoid H with finite class group such that hf(H)=n (see Example 2.4.2).

Theorem 1.2.

Let H be a transfer Krull monoid over an abelian group G.

  1. hf(H)< if and only if exp(G)<.

  2. If exp(G)< and |G|3, then exp(G)hf(H)32(exp(G)1).

  3. If G is cyclic or exp(G)6, then hf(H)=exp(G).

We postpone the proofs of Theorems 1.1 and 1.2 to Section 3.

2. Preliminaries

Our notation and terminology are consistent with [Citation12]. Let N be the set of positive integers, let N0=N{0}, and let Q be the set of rational numbers. For integers a,bZ, we denote by [a,b]={xZ|axb} the discrete, finite interval between a and b.

Atomic monoids. By a monoid, we mean a semigroup with identity, and if not stated otherwise we use multiplicative notation. Let H be a monoid with identity 1=1HH. 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 of H. 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 [Citation9, Theorem 2.6]. If aHH× and a=u1uk, where kN and u1,,ukA(H), then k is a factorization length of a, and LH(a)=L(a)={kN|k is a factorization length of a} denotes the set of lengths of a. It is convenient to set L(a)={0} for all aH×.

Let H and B be atomic monoids. The homomorphism θ:HB is called a weak transfer homomorphism if it satisfies the following two properties.

(T1) B=B×θ(H)B× and θ1(B×)=H×.

(WT2) If aH, nN, v1,,vnA(B) and θ(a)=v1··vn, then there exist u1,,unA(H) and a permutation τSn such that a=u1··un and θ(ui)B×vτ(i)B× for each i[1,n].

A transfer Krull monoid is a monoid H having a weak transfer homomorphism θ:HB(G0), where B(G0) is the monoid of zero-sum sequences over a subset G0 of an abelian group G. If H is a commutative Krull monoid with class group G and G0G is the set of classes containing prime divisors, then there is a weak transfer homomorphism θ:HB(G0). Beyond that, there are wide classes of non-commutative Dedekind domains having a weak transfer homomorphism to a monoid of zero-sum sequences ([Citation31, Theorem 1.1], [Citation30, Theorem 4.4]). We refer to [Citation10, Citation16] for surveys on transfer Krull monoids. If θ:HB(G0) is a weak transfer homomorphism, then sets of lengths in H and in B(G0) coincide [Citation2, Lemma 2.7] and thus the statements of Theorems 1.1 and 1.2 can be proved in the setting of monoids of zero-sum sequences.

Monoids of zero-sum sequences. Let G be an abelian group and 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 unordered sequence of terms from G0 where repetition is allowed, and (as usual) we consider sequences as elements of the free abelian monoid with basis G0. Let S=g1··g=gG0gvg(S)F(G0) be a sequence over G0. We call supp(S)={gG|vg(S)>0}G the support of S,|S|==gGvg(S)N0 the length of S,σ(S)=i=1gi the sum of S, and Σ(S)={iIgi|I[1,]} the set of subsequence sums of S.

The sequence S is said to be

  • zero-sum free if 0Σ(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 B(G0)={SF(G0)|σ(S)=0}F(G0) is a submonoid, and the set of minimal zero-sum sequences is the set of atoms of B(G0). For any arithmetical invariant *(H) defined for a monoid H, we write *(G0) instead of *(B(G0)). In particular, A(G0)=A(B(G0)) is the set of atoms of B(G0) and hf(G0)=hf(B(G0)).

Let G be an abelian group. We denote by exp(G) the exponent of G which is the least common multiple of the orders of all elements of G. If there is no least common multiple, the exponent is taken to be infinity. Let rN and let (e1,,er) be an r-tuple of elements of G. Then (e1,,er) is said to be independent if ei0 for all i[1,r] and if for all (m1,,mr)Zr an equation m1e1++mrer=0 implies that miei=0 for all i[1,r]. Suppose G is finite. The r-tuple (e1,,er) is said to be a basis of G if it is independent and G=e1er. For every nN, 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.

Let G0G be a non-empty subset. For a sequence S=g1··gF(G0), we call k(S)=i=1l1ord(gi)Q0the cross number of S, and K(G0)=max{k(S)|SA(G0)}the cross number of G0. For the relevance of cross numbers in the theory of non-unique factorizations, see [Citation22, Citation24, Citation27] and [Citation12, Chapter 6].

The set G0 is called

  • half-factorial if the monoid B(G0) is half-factorial;

  • non-half-factorial if the monoid B(G0) is not half-factorial;

  • minimal non-half-factorial if G0 is not half-factorial but all its proper subsets are;

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

The following simple result [Citation12, Proposition 6.7.3] will be used throughout the article without further mention.

Lemma 2.1.

Let G be a finite abelian group and G0G a subset. Then the following statements are equivalent.

  1. G0 is half-factorial.

  2. k(U)=1 for every UA(G0).

  3. L(B)={k(B)} for every BB(G0).

Lemma 2.2.

Let G be a finite group, let G0G be a subset, let S be a zero-sum sequence over G0, and let A be a minimal zero-sum sequence over G0.

  1. If k(A)1, then |L(Aexp(G))|2.

  2. If there exists a zero-sum subsequence T of S such that |L(T)|2, then |L(S)|2.

  3. If k(A)<1 and k(A) is minimal over all minimal zero-sum sequences over G0, then |L(Aord(g)vg(A))|2,for all gsupp(A).

Proof.

1. Suppose k(A)1 and let A=g1··g, where N and g1,,gG0. Then Aexp(G)=(g1ord(g1))exp(G)ord(g1)··(gord(g))exp(G)ord(g), which implies that {exp(G),i=1exp(G)ord(gi)}={exp(G),exp(G)k(A)}L(Aexp(G)). It follows by k(A)1 that |L(Aexp(G))|2.

2. Suppose T is a zero-sum subsequence of S with |L(T)|2. It follows by L(S)L(T)+L(ST1) that |L(S)||L(T)|2.

3. Suppose k(A)<1 and k(A) is minimal over all minimal zero-sum sequences over G0. Let gsupp(A). Then there exist sN and minimal zero-sum sequences W1,,Ws such that Aord(g)vg(A)=gord(g)·W1··Ws. Since k(Aord(g)vg(A))=ord(g)vg(A)k(A)=1+i=1sk(Wi)>(1+s)k(A), we have ord(g)vg(A)s+1 and hence |L(Aord(g)vg(A))|2.

For commutative and finitely generated monoids, we have the following result.

Proposition 2.3.

Let H be a commutative unit-cancellative monoid. If Hred is finitely generated, then hf(H) is finite.

Proof.

We may assume that H is reduced and not half-factorial. Suppose H is finitely generated and suppose A(H)={u1,,un}, where nN. Set A0={iIui|I[1,n]}. Then A0 is finite and hence there exists MN such that |L(a0M)|2 for all a0A0 with a0 not half-factorial. Let aHH× such that a is not half-factorial. It suffices to show that |L(aM)|2. Suppose a=u1k1··unkn, where k1,,knN0. Set I0={i[1,n]|ki1} and a0=iIui. Then a0 divides a and a0=a is not half-factorial, whence |L(a0M)|2 and |L(aM)|2.

If G0 is a finite subset of an abelian group, then B(G0) is finitely generated [Citation12, Theorem 3.4.2] and thus hf(G0)<. We refer to [Citation8, Sections 3.2 and 3.3] and [Citation14] for semigroups of ideals and semigroups of modules that are finitely generated unit-cancellative but not necessarily cancellative.

Example 2.4.

The following examples will help us to illustrate some important points.

  1. Let (e1, e2) be a basis of Z2 and let G0={e1,e1,e2,e2,e1+e2,e1e2}. Then A(G0)={e1(e1),e2(e2),(e1+e2)(e1e2),e1e2(e1e2),(e1)(e2)(e1+e2)}. Since e1(e1)·e2(e2)·(e1+e2)(e1e2)=e1e2(e1e2)·(e1)(e2)(e1+e2), we obtain G0 is not half-factorial. Furthermore, we have G1 is half-factorial for every non-empty proper subset G1G0. Let AB(G0). If supp(A)=G0, then |L(A)|2 and A=B(G0) is not half-factorial. If supp(A)G0, then A=B(supp(A)) is half-factorial and |L(A)|=1. Therefore hf(G0)=1.

  2. Let G be a cyclic group with order n and let gG with ord(g)=n, where nN3. Set G0={g,g}. Then G0 is not half-factorial. Let A0=g(g). Then A0 is not half-factorial and |L(A0n1)|=1, whence hf(G0)n. Let AB(G0) such that A is not half-factorial. Then supp(A)=G0 and A0 divides A, whence |L(An)|2. Therefore hf(G0)=n. Let GC22 and let (e1, e2) be a basis of G. Set G1={e1,e2,e1+e2}. Then G1 is not half-factorial. Let A1=e1e2(e1+e2). Then A1 is not half-factorial and |L(A1)|=1, whence hf(G1)2. Let AB(G1) such that A is not half-factorial. Then supp(A)=G1 and A1 divides A, whence |L(A2)|2. Therefore hf(G1)=2.

  3. Let H be a bifurcus monoid (i.e., 2L(a) for all aH(H×A(H))). For examples, see [Citation1, Examples 2.1 and 2.2]. Since for every aHH×, we have {2,3}L(a3), it follows that hf(H)3 and hf(H) is the minimal integer tN such that |L(at)|2 for all aHH×. Therefore hf(H)=3 if and only if there exists a0A(H) such that L(a02)={2}.

  4. Let HF=F××[p1,,ps] be a non-half-factorial finitely primary monoid of rank s and exponent α (see [Citation12, Definition 2.9.1]). For every a=ϵp1t1pstsF, we define ||a||=t1++ts, where t1,,tsN0 and ϵF×. Let aHH×. Since H is primary, we have H=a is not half-factorial. Thus hf(H) is the minimal integer tN such that |L(at)|2 for all aHH×. Suppose a0H with ||a0||=min{||a||:aHH×}. Then a0A(H) and L(a02)={2}, whence hf(H)3.

If HH×=(p1ps)αF and s2, then H is bifurcus and hence hf(H)=3. Suppose s = 1 and HH×=(p1)αF. Let b=ϵpβH. Then p3α divides b4. It follows by p3α=(pα)3=pα+1p2α1 that |L(b4)|2, whence hf(H)4. If 3β4α, then p3α divides b3 and hence |L(b3)|2. If (HTML translation failed), then b is an atom and b3=ϵ3p2α1p3β(2α1), whence |L(b3)|2. If 3β=4α1, then L(b3)={3}. Put all together, if α1 mod 3, then hf(H)=4. Otherwise hf(H)=3.

3. Proof of main theorem

Proposition 3.1.

Let G0G be a non-half-factorial subset and let S be a zero-sum sequence over G0 with supp(S)=G0.

  1. If G0 is an LCN-set, then |L(gG0gord(g))|2.

  2. If |G0|=2, then |L(gG0gord(g))|2.

  3. If G0 is a minimal non-half-factorial subset, then |L(Sexp(G))|2.

  4. If |{gG0|ord(g)/vg(S)=exp(G)}|1, then |L(Sexp(G))|2.

Proof.

1. Suppose G0 is an LCN-set. Since G0 is not half-factorial, there exists a minimal zero-sum sequence T over G0 such that k(T)>1. Note that T is a subsequence of gG0gord(g). Then there exist W1,,WlA(G0) such that gG0gord(g)=T·W1··Wl. Thus k(gG0gord(g))=|G0|=k(T)+i=1lk(Wi)>1+l. The assertion follows by {|G0|,1+l}L(gG0gord(g)).

2. Suppose |G0|=2 and let G0={g1,g2}. If G0 is an LCN-set, the assertion follows by 1. Suppose there exists a minimal zero-sum sequence T over G0 with k(T)<1. Let T0=g1l1·g2l2 be the minimal zero-sum sequence over G0 such that k(T0) is minimal. If min{ord(g1)l1,ord(g2)l2}2, say ord(g1)l12 then T02=g1ord(g1)·W, where W is an onempty zerosum sequence. Thus k(W)=2k(T0)1<k(T0), a contradiction to the minimality of k(T0). Therefore min{ord(g1)l1,ord(g2)l2}>2 and hence g1ord(g1)·g2ord(g2)=T02·V, where V is nonempty zerosum sequence. It follows that |L(g1ord(g1)·g2ord(g2))|2.

3. Suppose that G0 is a minimal non-half-factorial set. If S has a minimal zero-sum subsequence A with k(A)1, then the assertion follows by Lemma 2.2. If G0 is an LCN-set, then the assertion follows from 1 and Lemma 2.2.2. Therefore we can suppose L(S)={k(S)} and suppose there exists a minimal zero-sum sequence T over G0 with k(T)<1.

Let T0=i=1|G0|gili be the minimal zero-sum sequence over G0 such that k(T0) is minimal. The minimality of G0 implies that li1 for all i[1,|G0|]. After renumbering if necessary, we let ord(g1)l1=min{ord(gi)li|i[1,|G0|]}. By Lemma 2.2.3, |L(T0ord(g1)l1)|2. If T0ord(g1)l1 divides Sexp(G), then the assertion follows by Lemma 2.2.2. Suppose T0ord(g1)l1Sexp(G). Let I={i[1,|G0|]|ord(g1)l1li>exp(G)vgi(S)}. Thus for each iI, we have 2ord(gi)>liord(gi)liliord(g1)l1>exp(G)vgi(S)exp(G), which implies that ord(gi)=exp(G), vgi(S)=1, and ord(g1)l1>ord(gi)li=exp(G)li.

Let i0I such that li0=max{li|iI}. Therefore for every j[1,|G0|]I, we have ljexp(G)vgj(S)ord(g1)l1exp(G)vgj(S)exp(G)li0=li0vgj(S). Note that for every iI, we have lili0=li0vgi(S). It follows by vgi0(T0)=li0=li0vgi0(S)=vgi0(Sli0) that Sli0=T0·W, where W is a zerosum sequence over G0{gi0}. By the minimality of G0, we have G0{gi0} is half-factorial which implies that k(W)N. Therefore k(T0)=li0k(S)k(W) is an integer, a contradiction to k(T0)<1.

4. Let G1={gG0|ord(g)=exp(G)vg(S)}. Suppose G0G1 is not half-factorial. If G0G1 is an LCN-set, then the assertion follows by Proposition 3.1.1 and Lemma 2.2.2. Otherwise there exists a minimal zero-sum sequence A over G0G1 such that k(A)<1. We may assume that k(A) is minimal over all minimal zero-sum sequences over G0G1 and that min{ord(g)vg(A)|gsupp(A)}=ord(g0)vg0(A) for some g0supp(A)G0G1. Thus by Lemma 2.2.3, we have |L(Aord(g0)vg0(A))|2. The definition of G1 implies that Aord(g0)vg0(A) divides Sexp(G) and hence the assertion follows.

Suppose G0G1 is half-factorial. Then G1 is non-empty and hence G1={g0} for some g0G0. If G0 is an LCN-set, then the assertion follows by Proposition 3.1.1 and Lemma 2.2.2. Otherwise there exists a minimal zero-sum sequence A over G0 such that k(A)<1. We may assume that k(A) is minimal over all minimal zero-sum sequences over G0 and that min{ord(g)vg(A)|gsupp(A)}=ord(g1)vg1(A) for some g1supp(A)G0. Thus by Lemma 2.2.3, we have |L(Aord(g1)vg1(A))|2. For every gG0G1, we obtain vg(A)ord(g1)vg1(A)vg(A)ord(g)vg(A)<2ord(g)exp(G)vg(S). If vg0(A)ord(g1)vg1(A)ord(g0)=exp(G), then Aord(g1)vg1(A) divides Sexp(G) and hence |L(Sexp(G))|2.

Otherwise for every gGG1, we have exp(G)vg0(A)<ord(g1)vg1(A)ord(g)vg(A)exp(G)vg(S)2vg(A)exp(G)vg(S)vg(A). Therefore vg(A)<vg0(A)vg(S) for all gG0G1 which implies that A divides Svg0(A). Thus there exists a zero-sum sequence W over G0G1 such that Svg0(A)=A·W. Since G0G1 is half-factorial, we obtain k(A)=vg0(A)k(S)k(W) is an integer, a contradiction to k(A)<1.

Proof of Theorem 1.1.

By the definition of transfer Krull monoid, it suffices to prove the assertions for H=B(G0) and hence H is half-factorial if and only if G0 is half-factorial. If G0 is half-factorial, it is easy to see that hf(G0)=1 and |L(gG0g2ord(g))|=1. Therefore we only need to show that (b) implies (c) and that (d) implies (c).

(b) (c) Suppose hf(G0)=1 and assume to the contrary that G0 is not half-factorial. Then there exists AA(G0) such that k(A)1, whence supp(A) is not half-factorial. Therefore hf(supp(A))2, a contradiction.

(d) (c) Suppose |L(gG0g2ord(g))|=1 and assume to the contrary that G0 is not half-factorial. If G0 is an LCN set, then Proposition 3.1.1 implies that |L(gG0gord(g))|2, a contradiction. Thus there exists an atom AA(G0) with k(A)<1 and we may assume that k(A) is minimal over all atoms of B(G0). Let g0supp(A). Then by Lemma 2.2.3, we have |L(Aord(g0)vg0(A))|2, a contradiction to Aord(g0)vg0(A) | gG0g2ord(g).

Proof of Theorem 1.2.

By the definition of transfer Krull monoid, it suffices to prove all assertions for H=B(G).

1. Suppose exp(G)<. If |G|3, then 2 implies that hf(G)<. If |G|2, then B(G) is half-factorial and hence hf(G)=1.

Suppose exp(G)=. If there exists an element gG with ord(g)=, then An=((n+1)g)(ng)(g) is an atom for every nN. Since {(n+1)g,ng,g} is not half-factorial and |L(Ann)|=1 for every n2, we obtain that hf(G)n for every n2, that is, hf(G)=. Otherwise G is torsion. Then there exists a sequence (gi)i=1 with giG and limiord(gi)=. It follows by 1 that hf(G)hf(gi)ord(gi) for all iN, that is, hf(G)=.

2. If G is an elementary 2-group and e1, e2 are two independent elements, then {e1,e2,e1+e2} is not a half-factorial set and |L(e1e2(e1+e2))|=1 which implies that hf(G)2=exp(G). Otherwise there exists an element gG with ord(g)=exp(G)3. Since {g,g} is not half-factorial and |L(gord(g)1(g)ord(g)1)|=1, we obtain hf(G)ord(g)=exp(G).

Let S be a zero-sum sequence over G such that supp(S) is not half-factorial. In order to prove hf(G)3exp(G)32, we show that |L(S3exp(G)32)|2. Set G0=supp(S). If G0 is an LCN-set, the assertion follows by Proposition 3.1.1. Suppose there exists an atom AA(G0) with k(A)<1. Let A0A(supp(S)) be such that k(A0) is minimal over all minimal zero-sum sequences over G0 and set A0=g1l1··gyly, where y,l1,lyN and g1,,gysupp(S) are pairwise distinct elements. If there exists j[1,y] such that 2ljord(gj), then gjord(gi) divides A02 and hence A02=gjord(gj)·W for some non-empty sequence WB(supp(S)). Thus k(W)=2k(A0)1<k(A0), a contradiction to the minimality of k(A0). Therefore 2liord(gi)1 for all i[1,y]. After renumbering if necessary, we assume ord(g1)l1=min{ord(gi)li|i[1,y]}. Then liord(g1)l1liord(gi)liliord(gi)+li1liord(gi)+ord(gi)1213exp(G)32, which implies that A0ord(g1)l1 divides S3exp(G)32. The assertion follows by Lemma 2.2.3.

3(a). Suppose that G is cyclic and that gG with ord(g)=|G|3. We will show that hf(G)=exp(G).

Let S be a zero-sum sequence over G such that supp(G) is not half-factorial. It suffices to show that |L(Sexp(G))|2. If |{gsupp(S)|ord(g)=|G|vg(S)}|1, then the assertion follows from Proposition 3.1.4. Suppose |{gsupp(S)|ord(g)=|G|vg(S)}|2. Then there exist distinct g1,g2supp(S) such that ord(g1)=ord(g2)=|G|. We may assume that g1 = kg2 for some kN2 with gcd(k,|G|)=1. It follows by k(g1|G|k·g2)<1 that G0={g1,g2}={g1,kg1} is not half-factorial. By Proposition 3.1.2, we obtain that |L(Sexp(G))|2.

3(b). Suppose G is a finite abelian group with exp(G)6. We need to prove that hf(G)=exp(G). Let S be a zero-sum sequence over G such that supp(G) is not half-factorial. It suffices to show that |L(Sexp(G))|2.

If supp(S) is an LCN-set, the assertion follows by Proposition 3.1.1. Thus there is a minimal zero-sum sequence W over supp(S) such that k(W)<1. By Proposition 3.1.2 and Lemma 2.2.2, we have |supp(W)|3. Suppose W | S. Since k(W)<1, it follows by Lemma 2.2 that |L(Wexp(G))|2 and hence |L(Sexp(G))|2. Therefore we may assume that WS, whence |W||supp(W)|+14. It follows that 6exp(G)|W|k(W)>|W|4.

We distinguish two cases according to exp(G){5,6}.

Case 1. exp(G)=5.

Then, GC5r and for all WA(supp(S)) with k(W)<1, we have that WS, |supp(W)|=3,  and  |W|=4.

Let W0 be an atom over supp(S) with k(W0)<1. Then W0=g12g2g3andS=Tg1g2g3, where g1,g2,g3supp(S) are pairwise distinct and TF(supp(S){g1}) with σ(T)=g1.

We may assume T is zero-sum free. Otherwise T=T0T, where T0 is a zero-sum sequence and T is zero-sum free. We can replace S by Tg1g2g3, since |L((Tg1g2g3)5)|2 implies that |L(S5)|2. Therefore S is a product of at most three atoms and every term of S has order 5.

Assume to the contrary that |L(S5)|=1, that is, L(S5)={|T|+3}. Since g15g25g35=W02(g1g23g33) is a zero-sum subsequence of S5, we obtain that g1g23g33 is an atom. Note that (g12g2g3)2S=(g14T)(g1g23g33) is a zerosum subsequence of S5.

Suppose S is an atom. Then L(g14T)={2} and hence |g14T|2×4=8, that is, |T|4. It follows by {5}=L(S5)={|T|+3} that |T|=2, a contradiction.

Suppose S is a product of two atoms. Then L(g14T)={3} and hence |g14T|3×4=12, that is, |T|8. It follows by {10}=L(S5)={|T|+3} that |T|=7, a contradiction.

Suppose S is a product of three atoms. Then L(g14T)={4} which implies that T=T1T2T3T4 such that g1Ti is zero-sum for all i[1,4]. Since g1Ti | S, we obtain k(g1Ti)1 and hence |g1Ti|5. Therefore |g14T|4×5=20, that is, |T|16. It follows by {15}=L(S5)={|T|+3} that |T|=12, a contradiction.

Case 2. exp(G)=6.

Let W be an atom over supp(S) with k(W)<1. If |W|=4, then |supp(W)|=3 and hence W must be of the form W=g12g2g3 where g1,g2,g3supp(S) are pairwise distinct. Since (g16)(g26)(g36)=W3(g23g33), we obtain that |L(g16g26g36)|2. It follows from the fact that g16g26g36 divides Sexp(G) that |L(Sexp(G))|2.

If |W|=5, then |supp(W)|=3 or 4 and hence W can be written in one of the following ways.

  1. W=g13g2g3, where g1,g2,g3supp(S) are pairwise distinct.

  2. W=g12g22g3, where g1,g2,g3supp(S) are pairwise distinct.

  3. W=g12g2g3g4, where g1,g2,g3,g4supp(S) are pairwise distinct.

Suppose (i) holds. Then 0=2σ(W)=6g1+2g2+2g3=2g2+2g3. Since (g16)(g26)(g36)=W2(g22g32)2, we obtain that |L(g16g26g36)|2. It follows from the fact that g16g26g36 divides Sexp(G) that |L(Sexp(G))|2.

Suppose (ii) holds. Then 0=3σ(W)=6g1+6g2+3g3=3g3. Thus ord(g3)=2 and hence k(W)1/2+4/6>1, a contradiction.

Suppose (iii) holds. Then 0=3σ(W)=6g1+3g3+3g3+3g4=3g2+3g3+3g4. Therefore W0=g23g33g43 is zero-sum. If W0=g16g26g36g46(W)3 is not a minimal zero-sum sequence, then |L(g16g26g36g46)|2 and hence |L(Sexp(G))|2. If W0 is minimal zero-sum, then g16g26g36g46=(g16)W02 implies that |L(g16g26g36g46)|2 and hence |L(Sexp(G))|2.

Acknowledgments

We thank the referee for very careful reading and for providing valuable suggestions.

Additional information

Funding

This work has been supported in part by the National Science Found of China with grant no. 11671218 and by the Austrian Science Fund with grant no. P33499.

References

  • Adams, D., Ardila, R., Hannasch, D., Kosh, A., McCarthy, H., Ponomarenko, V., Rosenbaum, R. (2009). Bifurcus semigroups and rings. Involve. 2(3):351–356. DOI: 10.2140/involve.2009.2.351.
  • Baeth, N. R., Smertnig, D. (2015). Factorization theory: from commutative to noncommutative settings. J. Algebra 441:475–551. DOI: 10.1016/j.jalgebra.2015.06.007.
  • Baginski, P., Kravitz, R. (2010). A new characterization of half-factorial Krull monoids. J. Algebra Appl. 09(05):825–837. DOI: 10.1142/S0219498810004269.
  • Carlitz, L. (1960). A characterization of algebraic number fields with class number two. Proc. Am. Math. Soc. 11:291–392.
  • Chapman, S. T., Coykendall, J. (2000). Half-factorial domains, a survey. In: Non-Noetherian Commutative Ring Theory, Mathematics and Its Applications, Vol. 520, Boston, MA: Kluwer Academic Publishers, pp. 97–115.
  • Chapman, S. T., Krause, U., Oeljeklaus, E. (2000). Monoids determined by a homogeneous linear diophantine equation and the half-factorial property. J. Pure Appl. Algebra 151(2):107–133. DOI: 10.1016/S0022-4049(99)00062-6.
  • Coykendall, J. (2005). Extensions of half-factorial domains: A survey. In: Arithmetical Properties of Commutative Rings and Monoids, Lect. Notes Pure Appl. Math., Vol. 241. Boca Raton, FL: Chapman & Hall/CRC, pp. 46–70.
  • 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(12):1750234–1750242. DOI: 10.1142/S0219498817502346.
  • Fan, Y., Tringali, S. (2018). Power monoids: A bridge between factorization theory and arithmetic combinatorics. J. Algebra 512:252–294. DOI: 10.1016/j.jalgebra.2018.07.010.
  • Geroldinger, A. (2016). Sets of lengths. Amer. Math. Mon. 123:960–988.
  • Geroldinger, A., Göbel, R. (2003). Half-factorial subsets in infinite abelian groups. Houst. J. Math. 29:841–858.
  • Geroldinger, A., Halter-Koch, F. (2006). Non-unique factorizations. In: Algebraic, Combinatorial and Analytic Theory, Pure and Applied Mathematics, Vol. 278, Boca Raton, FL: Chapman & Hall/CRC.
  • Geroldinger, A., Kainrath, F., Reinhart, A. (2015). Arithmetic of seminormal weakly Krull monoids and domains. J. Algebra 444:201–245. DOI: 10.1016/j.jalgebra.2015.07.026.
  • Geroldinger, A., Reinhart, A. (2019). The monotone catenary degree of monoids of ideals. Int. J. Algebra Comput. 29(03):419–457. DOI: 10.1142/S0218196719500097.
  • Geroldinger, A., Zhong, Q. (2016). The set of minimal distances in Krull monoids. Acta Arith. 173:1–120. DOI: 10.4064/aa7906-1-2016.
  • Geroldinger, A., Zhong, Q. (2020). Factorization theory in commutative monoids. Semigroup Forum. 100(1):22–51. DOI: 10.1007/s00233-019-10079-0.
  • Gotti, F. (2020). Geometric and combinatorial aspects of submonoids of a finite-rank free commutative monoid. Linear Algebra Appl. 604:146–186. DOI: 10.1016/j.laa.2020.06.009.
  • Kainrath, F., Lettl, G. (2000). Geometric notes on monoids. Semigroup Forum. 61(2):298–302. DOI: 10.1007/PL00006026.
  • Malcolmson, P., Okoh, F. (2009). Power series extensions of half-factorial domains. J. Pure Appl. Algebra 213(4):493–495. DOI: 10.1016/j.jpaa.2008.07.014.
  • Malcolmson, P., Okoh, F. (2016). Half-factorial subrings of factorial domains. J. Pure Appl. Algebra 220(3):877–891. DOI: 10.1016/j.jpaa.2015.06.011.
  • Plagne, A., Schmid, W. A. (2005). On large half-factorial sets in elementary p-groups: maximal cardinality and structural characterization. Isr. J. Math. 145(1):285–310. DOI: 10.1007/BF02786695.
  • Plagne, A., Schmid, W. A. (2005). On the maximal cardinality of half-factorial sets in cyclic groups. Math. Ann. 333(4):759–785. DOI: 10.1007/s00208-005-0690-y.
  • Roitman, M. (2011). A quasi-local half-factorial domain with an atomic non-half-factorial integral closure. J. Commut. Algebra 3(3):431–438. DOI: 10.1216/JCA-2011-3-3-431.
  • Schmid, W. A. (2005). Differences in sets of lengths of Krull monoids with finite class group. J. Théor. Nombres Bordeaux. 17(1):323–345. DOI: 10.5802/jtnb.493.
  • Schmid, W. A. (2005). Half-factorial sets in finite abelian groups: a survey. Grazer Math. Ber. 348:41–64.
  • Schmid, W. A. (2006). Half-factorial sets in elementary p-groups. Far East J. Math. Sci. 22:75–114.
  • 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.
  • Skula, L. (1976). On c-semigroups. Acta Arith. 31(3):247–257. DOI: 10.4064/aa-31-3-247-257.
  • Śliwa, J. (1976). Factorizations of distinct lengths in algebraic number fields. Acta Arith. 31(4):399–417. DOI: 10.4064/aa-31-4-399-417.
  • Smertnig, D. (2019). Factorizations in bounded hereditary noetherian prime rings. Proc. Edinb. Math. Soc. 62(2):395–442. DOI: 10.1017/S0013091518000305.
  • Smertnig, D. (2013). Sets of lengths in maximal orders in central simple algebras. J. Algebra 390:1–43. DOI: 10.1016/j.jalgebra.2013.05.016.
  • Zaks, A. (1976). Half factorial domains. Bull. Amer. Math. Soc. 82(5):721–723. DOI: 10.1090/S0002-9904-1976-14130-4.
  • Zaks, A. (1980). Half-factorial-domains. Isr. J. Math. 37(4):281–302. DOI: 10.1007/BF02788927.