777
Views
6
CrossRef citations to date
0
Altmetric
Articles

On strongly primary monoids and domains

&
Pages 4085-4099 | Received 06 Apr 2019, Accepted 07 Apr 2020, Published online: 04 May 2020

Abstract

A commutative integral domain is primary if and only if it is one-dimensional and local. A domain is strongly primary if and only if it is local and each nonzero principal ideal contains a power of the maximal ideal. Hence, one-dimensional local Mori domains are strongly primary. We prove among other results that if R is a domain such that the conductor (R:R̂) vanishes, then Λ(R) is finite; that is, there exists a positive integer k such that each nonzero nonunit of R is a product of at most k irreducible elements. Using this result, we obtain that every strongly primary domain is locally tame, and that a domain R is globally tame if and only if Λ(R)=. In particular, we answer Problem 38 of the open problem list by Cahen et al. in the affirmative. Many of our results are formulated for monoids.

2010 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

Factorization theory of integral domains studies factorizations of elements and ideals ([Citation1, Citation11, Citation18]). Some ideal theoretic conditions (such as the ascending chain condition on principal ideals) guarantee that every nonzero, nonunit element of a domain can be written as a product of atoms (irreducible elements). The goal is to describe the nonuniqueness of factorizations by arithmetical invariants and study their relationship with classical ring theoretical invariants. Class groups and the structure of certain localizations are such algebraic invariants which control element factorizations. In case of weakly Krull Mori domains, this interplay is described by the T-block monoid of the domain which is built by the v-class group and localizations at height-one primes containing the conductor ([Citation18, Theorem 3.7.1]). These localizations are one-dimensional and local, and this connection stimulated the interest of factorization theory in one-dimensional local domains.

To recall some arithmetical invariants, let R be an integral domain. If a=u1··uk is a factorization of an element aR into atoms u1,,uk, then k is a factorization length. The set L(a)N of all possible factorization lengths of a is called the set of lengths of a. The local tame degree t(u) of an atom uR is the smallest N with the following property: For any multiple a of u and any factorization a=v1··vn of a, there is a subproduct which is a multiple of u, say v1··vm, and a refactorization of this subproduct which contains u, say v1··vm=uu2··u such that max{,m}N. Local tameness is a central finiteness property which – in many settings – implies further arithmetical finiteness properties (see, e.g. [Citation16, Citation18, Section 4]).

Here are two classes of locally tame monoids:

  • Krull monoids with finite class group [Citation18, Theorem 3.4.10];

  • C-monoids [Citation18, Theorem 3.3.4].

Moreover, if R is a Mori domain with nonzero conductor f=(R:R̂) such that R/f and the class group C(R̂) are finite, then the multiplicative monoid R=R{0} is a C-monoid by [Citation18, Theorem 2.11.9], whence R is locally tame (compare with Example 3.17.4). Precise values for local and global tame degrees were studied for Krull monoids and numerical monoids ([Citation8, Citation9, Citation12, Citation31]). For computational aspects and for results in the noncommutative setting we refer to [Citation5, Citation13].

By a local ring, we mean a commutative ring with a unique maximal ideal, not necessarily Noetherian. It is well known that a domain R is one-dimensional and local if and only if its multiplicative monoid of nonzero elements is primary. A monoid H is strongly primary if each principal ideal of H contains a power of the maximal ideal. The multiplicative monoid of a one-dimensional local Mori domain is strongly primary, and this was the key property to prove local tameness for one-dimensional local Mori domains under a variety of additional assumptions ([Citation22, Theorem 3.5] and [Citation27, Theorem 3.5]). However, the general case remained open.

In the present paper, we prove that every strongly primary domain is locally tame and we provide a characterization of global tameness (Theorems 3.8 (b) and Citation3.Citation9.Citation2). In particular, all one-dimensional local Mori domains turn out to be locally tame and this answers Problem 38 in [Citation7] in the affirmative. Although our present approach is semigroup theoretical over large parts (Theorem 3.8 (a)), it also uses substantially the ring structure, and this is unavoidable since strongly primary Mori monoids need not be locally tame as shown in [Citation22, Proposition 3.7 and Example 3.8] (see Example 3.17.1).

2. Background on primary monoids and domains

We denote by N the set of positive integers and by N0=N{0} the set of non-negative integers. For integers a,bZ, we denote by [a,b]={xZ|axb} the discrete interval between a and b. Let L,LZ be subsets of the integers. Then L+L={a+b|aL,bL} denotes the sumset. A positive integer dN is a distance of L if there is some aL such that L[a,a+d]={a,a+d}. We denote by Δ(L)N the set of distances of L. Thus, Δ(L)= if and only if |L|1, and L is an arithmetical progression with difference d if and only if Δ(L){d}. Note that the empty set is considered an arithmetical progression.

By a monoid, we mean a commutative cancellative semigroup with identity and by a domain, we mean a commutative integral domain. Thus, if R is a domain, then R=R{0} is a monoid. All ideal theoretic and arithmetical concepts are introduced in a monoid setting, but they will be used both for monoids and domains.

Ideal theory of monoids and domains. Let H be a monoid and q(H) the quotient group of H. We denote by

  • H={xq(H)|there exists some NN such that xnH for all nN} the seminormal closure of H, by

  • H˜={xq(H)|there exists some NN such that xNH} the root closure of H, and by

  • Ĥ={xq(H)|there exists cH such that cxnH for all nN} the complete integral closure of H.

Then, we have (2.1) HHH˜Ĥq(H) ,(2.1) and all inclusions can be strict. We say that H is seminormal (root-closed, resp. completely integrally closed) if H=H (H=H˜, resp. H=Ĥ). Note that H is seminormal, H˜ is root-closed, but Ĥ need not be completely integrally closed.

Lemma 2.1.

Let H be a root-closed monoid and xq(H).

1. If cH such that cxnH for some nN, then cxkH for every k[1,n].

2. We have xĤ if and only if there exists cH such that cxnH for infinitely many nN.

Proof.

1. Let cH and nN such that cxnH. If k[1,n], then (cxk)n=(cxn)kcnkH whence cxkH because H is root-closed.

2. If there is a cH such that cxnH for infinitely many nN, then 1. implies that cxkH for all kN whence xĤ.

Lemma 2.2.

Let H be a monoid.

1. Ĥ is root-closed.

2. If (H:Ĥ), then Ĥ is completely integrally closed.

3. An element xq(H) lies in Ĥ̂ if and only if there exists an element cH such that cxnĤ for infinitely many nN.

Proof.

1. Let xq(H) such that xeĤ for some eN. We have to show that xĤ. There is an element cH such that cxiH for every i[1,e] and such that c(xe)kH for all kN0. If nN, then n=ke+i, with kN0 and i[1,e], and c2xn=c(xe)k(cxi)H which implies that xĤ.

2. See [Citation18, Proposition Citation2.Citation3.Citation4].

3. Since Ĥ is root-closed by 1., the assertion follows from Lemma 2.1.2 (applied to the monoid Ĥ). □

Let H be a monoid. Then, H× denotes the group of units of H and Hred={aH×|aH} its associated reduced monoid. We also let m=HH× and n=ĤĤ×. Let X,Yq(H) be subsets. Then, X is an s-ideal of H if XH and XH = X. We set (X:Y)={zq(H)|zYX} and X1=(H:X). A divisorial ideal (v-ideal) is a set of the form (X1)1 for Xq(H). We denote by v-spec(H) the set of all prime v-ideals of H. The monoid H is a Mori monoid if it satisfies the ascending chain condition on divisorial ideals, and it is a Krull monoid if it is a completely integrally closed Mori monoid.

Arithmetic of monoids and domains. For any set P, let F(P) be the free abelian monoid with basis P. Let |·|:F(P)N0 denote the unique epimorphism satisfying |p|=1 for each pP, whence |·| is mapping each zF(P) onto its length. We denote by A(H) the set of atoms of H. Then, Z(H)=F(A(Hred)) is the factorization monoid of H and π:Z(H)Hred denotes the canonical epimorphism. For an element aH,

  • Z(a)=π1(aH×)Z(H) is the set of factorizations of a, and

  • L(a)={|z||zZ(a)}N0 is the set of lengths of a.

To define the distance of factorizations, consider two factorizations z,zZ(H). Then, we write z=u1··uv1··vmandz=u1··uw1··wn , where ,m,nN0 and all ui,vj,wkA(Hred) such that {v1,,vm}{w1,,wn}=. We call d(z,z)=max{m,n}N0 the distance between z and z. The function d:Z(H)×Z(H)N0 is a metric on Z(H). We say that H is

  • atomic if every nonunit can be written as a finite product of atoms, and

  • a BF-monoid (a bounded factorization monoid) if it is atomic and all sets of lengths are finite.

A monoid is a BF-monoid if and only if n0(HH×)n=, and Mori monoids are BF-monoids ([Citation18, Theorem 2.2.9]). For every kN, we set ρk(H)=k if H=H×, and otherwise we set ρk(H)=sup{supL(a)|aH,kL(a)}N{} .

Clearly, the sequence (ρk(H))k1 is increasing and, if ρk(H) is finite for some kN, then ρk(H) is the maximal length of a factorization of a product of k atoms. The elasticity of H, introduced by Valenza in [Citation34], is defined as ρ(H)=sup{m/n|m,nL,LL(H)}, and by [Citation18, Proposition 1.4.2] we have (2.2) ρ(H)=sup{m/n|m,nL,LL(H)}=limkρk(H)k.(2.2)

For a subset SH, the set of distances of S is defined by Δ(S)=aSΔ(L(a))and we setΛ(S)=sup{minL(a)|aS}N0{} .

For aH, let Λ(a)=Λ({a}). If a,bH, then L(a)+L(b)L(ab) whence minL(ab)minL(a)+minL(b) and thus (2.3) Λ(ab)Λ(a)+Λ(b) .(2.3)

The catenary degree c(a) of an element aH is the smallest NN0{} with the following property: If z,zZ(a) are two factorizations of a, then there are factorizations z=z0,z1,,zk=z of a such that d(zi1,zi)N for all i[1,k]. Then c(H)=sup{c(a)|aH}N0{} is the catenary degree of H. The monoid H is factorial if and only if c(H)=0, and if this does not hold, then (2.4) 2+supΔ(H)c(H)by[18,Theorem1.6.3].(2.4)

Primary monoids and domains. Let H be a monoid and m=HH×. Then, H is said to be

  • primary if HH× and for all a,bm there is an nN such that bnaH, and

  • strongly primary if HH× and for every am, there is an nN such that mnaH (we denote by M(a) the smallest nN having this property).

If H is strongly primary, then H is a primary BF-monoid ([Citation18, Lemma 2.7.7]). However, primary BF-monoids need not be strongly primary (Example 3.17.2).

Lemma 2.3.

[Citation15, Proposition 1] Let H be a primary monoid. Then H×=H×H=H˜×H=Ĥ×H.

Lemma 2.4.

Let H be a primary monoid.

1. If aH and xq(H), then there exists an NN such that aNxH, so anxH for all nN. Moreover, for n sufficiently large, we have both anxH and anHx.

2. If H is strongly primary and xq(H), then, for NN sufficiently large, mNxH and mNHx.

Proof.

1. Let x=bc1, where b,cH. We have HbHbc1=Hx. Since H is primary, Hb and so also Hx contain a power of a. Hence, also Hx1 contains a power of a. Thus for n sufficiently large we have both anxH and anHx.

2. Use a similar argument as in the previous item. □

If H is strongly primary and xq(H), we denote by M(x)N the smallest nN with mnxH (see Lemma 2.4.2).

Lemma 2.5.

Let H be a primary monoid.

1. H=H˜. In particular, H is seminormal if and only if H is root-closed.

2. If H is seminormal, then m(H:Ĥ), so Ĥ=(m:m).

3. [Citation14, Theorem 4] H˜̂ is completely integrally closed. Thus the complete integral closure of a seminormal primary monoid is completely integrally closed.

Proof.

1. By (Equation2.1), we have HH˜ whence it remains to verify the reverse inclusion. Let xq(H), and kN such that xkH. By Lemma 2.4.1, there is an NN such that (xk)nxH for all nN. Thus xknH and xkn+1H for all nN, which implies that xmH for all m sufficiently large since kn and kn + 1 are coprime integers (explicitly, xmH for m(kn)2). Therefore xH.

2. Let xĤ, thus dxnH for some dH and all nN. Let aH. Since H is primary, we have akd1H for some kN, so akxnH for all nN. Thus, ak(xn)kH for all nN. Since H is root-closed by item 1., we obtain that axnH for all nN. Thus, aĤH, so mĤH. Since mn by Lemma 2.3, we infer that mĤm, that is, Ĥ(m:m), so Ĥ=(m:m). Cf. [Citation19, Proposition 4.8]. □

Monoid properties do not always carry over to integral domains. However, the domain R is seminormal (completely integrally closed, Mori, Krull, primary, strongly primary, atomic) if and only if its monoid R has the respective property. We consider, for example, the Mori property. By definition, the domain R is Mori if and only if it satisfies the ascending chain condition on integral divisorial ideals. If XR is a subset, then I=(R:(R:X)) is divisorial, and we have I=(R:(R:X)), and I=(R:(R:X)){0}, where Y=Y{0} for Yq(R). It follows that the domain R is Mori if and only if the monoid R is Mori.

Note that a domain R is primary if and only if R is one-dimensional and local ([Citation18, Proposition 2.10.7]). Every primary Mori monoid is strongly primary with v-spec(H)={,HH×} whence every one-dimensional local Mori domain is strongly primary ([Citation18, Proposition 2.10.7]. A survey on the ideal theory of Mori domains is given by Barucci in [Citation6]).

All finitely primary monoids (including all numerical monoids) are strongly primary ([Citation18, Section 2.7]). Examples of finitely primary domains which are not Mori can be found in [Citation26, Sections 3 and 4]. Moreover, [Citation26, Examples 4.6 and 4.7] are not multiplicative monoids of domains; Example 4.6 is Mori, while Example 4.7 is not. Puiseux monoids (these are additive submonoids of (Q0,+)), which are strongly primary, are discussed in [Citation16] and strongly primary monoids stemming from module theory can be found in [Citation4, Theorems 5.1 and Citation5.Citation3].

An example going back to Grams ([Citation2, Example 1.1] or [Citation18, Example 1.1.6]) exhibits an atomic one-dimensional local domain which is not a BF-domain whence it is neither strongly primary nor locally tame (see Definition 3.1 below).

3. On the arithmetic of strongly primary monoids and domains

We start with the concept of local tameness as given in [Citation18].

Definition 3.1.

Let H be an atomic monoid.

1. For an element aH, let ω(a) denote the smallest NN0{} with the following property: If nN and a1,,anH with a|a1··an, then there is a subset Ω[1,n] such that |Ω|Nanda|λΩaλ .

We set ω(H)=sup{ω(u)|uA(H)}.

2. For an element uA(Hred), let t(u) denote the smallest NN0{} with the following property: If aH with Z(a)uZ(H) and zZ(a), then there is a zZ(a)uZ(H) such that d(z,z)N.

3. H is said to be

(a) locally tame if t(u)< for all uA(Hred) and

(b) (globally) tame if t(H)=sup{t(u)|uA(Hred)}<.

If uA(H), we let t(u)=t(uH×). For a prime element uH, we have ω(u)=1 and t(u)=0, thus ω(H)=1 for a factorial monoid. If uA(H) is not prime, then ω(u)t(u) whence for a nonfactorial monoid we have ω(H)t(H). Moreover, H is globally tame if and only if ω(H)< [Citation23, Proposition 3.5]. Every Mori monoid satisfies ω(a)< for all aH ([Citation21, Theorem 4.2], [Citation10, Proposition 3.3]) but this need not hold true for the local tame degrees t(·). If H is an atomic monoid, since supL(a)ω(a) for all aH ([Citation21, Lemma 3.3]), the finiteness of the ω(a) values (hence, in particular, local tameness) implies that H is a BF-monoid.

We continue with a simple reformulation of local tameness which we use in the following. Clearly, for an atom uA(Hred), the local tame degree t(u) is the smallest NN0{} with the following property:

  • For every multiple aHred of u and any factorization z=v1··vn of a which does not contain u, there is a subproduct which is a multiple of u, say v1··vm, and a refactorization of this subproduct which contains u, say v1··vm=uu2··u such that max{,m}N.

Recall that a monoid is half-factorial if all the factorizations of an element in H are of the same length, equivalently, if the set of distances Δ(H) is empty.

Proposition 3.2.

Let H be a strongly primary monoid.

1. For every atom uH, we have ω(u)M(u). Thus ω(u)<.

2. Assume that Λ(H)<. Then ρΛ(H)(H)=, so ρk(H)= for all kΛ(H). In particular, H is not half-factorial.

3. For every atom uH we have t(u)ρω(u)(H). Hence, if ρω(u)(H)< for every atom uH, then H is locally tame.

Proof.

1. Since u divides each product of M(u) nonunits in H, we see that N=M(u) satisfies the property in Definition 3.1.1, and since ω(u) is the least N satisfying this property, it follows that ω(u)M(u).

2. For every nN, a product of n atoms is also a product of k atoms for some positive integer kΛ(H). Hence, ρΛ(H)(H)ρk(H)n. Thus, ρΛ(H)(H)=.

3. Let a be an element of H that is divisible by an atom u. A factorization of a has a subproduct that is divisible by u and of length kω(u). Hence, every factorization of this subproduct is of length ρk(H)ρω(u)(H). By the reformulation of Definition 3.1.1, we obtain that t(u)ρω(u)(H).

A monoid H is said to be a

  • valuation monoid if for all a,bH we have a|b or b|a.

  • discrete valuation monoid if Hred(N0,+).

Lemma 3.3.

Let H be a monoid with m=HH×.

  1. H is a primary valuation monoid if and only if H is a completely integrally closed valuation monoid if and only if Hred is isomorphic to a monoid of non-negative elements of a nonzero subgroup of (R,+).

  2. The following conditions are equivalent:

    1. H is a discrete valuation monoid.

    2. H is atomic and m is principal.

    3. H is a completely integrally closed strongly primary monoid.

  3. [Citation24, Theorem 16.4 (g)] A valuation monoid is discrete if and only if it is atomic if and only if it is strongly primary.

Proof.

For 1. and for the equivalence of the first two items of 2. we refer to [Citation24, Theorems 15.5 and Citation16.Citation4] and to [Citation14, Section 3]. To complete the proof of 2., we note that (a)(c), so it is enough to prove the implication (c)(b). Thus, it remains to show that m is principal. Let uA(H). Then, mM(u)uH. If M(u)=1, then m=uH and we are done. Assume that M(u)>1. Then, mM(u)uH, whence u1mM(u)m and u1mM(u)1(m:m). Since H is completely integrally closed, we have (m:m)H and thus mM(u)1uH, contradicting the minimality of M(u).

Lemma 3.4.

Let (H,m) be a primary monoid and n=ĤĤ×.

1. If every element of n has a power lying in m, then Ĥ is primary.

2. If (H:Ĥ), then Ĥ is primary if and only if every element of n has a power lying in m.

Proof.

1. Let x,yn. There exists an nN such that xn=mm. Since H is primary, by Lemma 2.4.1., there exists a kN such that mkyH. Thus, xnkyH, implying that Ĥ is primary.

2. Assume that Ĥ is primary. Let cm(H:Ĥ) and xn. As Ĥ is primary, and mn by Lemma 2.3, we obtain that for some integer nN we have xncĤm. We complete the proof using item 1. □

Lemma 3.5.

Let (H,m) be a strongly primary monoid and n=ĤĤ×.

1. If xq(H), then Λ(HxH)<M(x)<.

2. We have Λ(H)< if and only if there is a cm with Λ({cm|mN0})<.

3. If Λ(H)=, then every element of n has a power lying in m, so Ĥ is primary.

Proof.

1. If aHxH, then amM(x)xH. Thus, Λ(HxH)<M(x).

2. Let cm such that Λ({cm|mN0})<. Let aH. By [Citation18, Lemma 2.7.7.1], we have a|ck for some kN, so there is an nk in N0 such that a=cnb, where bH is not divisible by c. Now 1. implies that Λ(a)Λ(cn)+Λ(b)Λ({cm|mN0})+M(c) .

Thus, Λ(H)Λ({cm|mN0})+M(c), and the reverse implication is trivial.

3. Suppose there is an xn such that no power of x belongs to m, and we will prove that Λ(H)<. Let dm such that dxnH for all nN. We choose an element cm and assert that Λ(c)<M(d)+M(x), implying that Λ(H)<M(d)+M(x). If cdH, then Λ(c)<M(d) by 1. Suppose that cdH. Then, there exists an integer kN such that c(dxkHdxk+1H), because otherwise we would have c(1/x)nH for all nN implying that 1/xĤ. Thus, c=(dxk)b, where bHxH whence Λ(b)<M(x) by 1. If dxkdH, then xkHm, so x is invertible in H, a contradiction. This implies that dxkHdH whence Λ(dxk)<M(d) by 1. Thus, Λ(c)=Λ(dxkb)Λ(dxk)+Λ(b)<M(d)+M(x). By Lemma 3.4, Ĥ is primary. □

The converse of Lemma 3.5.3 is false even for domains (see Example 3.15 below and Lemma 3.4). Let (H,m) be a strongly primary monoid. If Ĥ is not primary (equivalently, if m is not the unique prime s-ideal of Ĥ), then Λ(H)<. In particular, if R is a strongly primary domain such that R̂ is not local, then Λ(R)<.

Lemma 3.6.

Let (H,m) be a strongly primary monoid such that Λ(H)=. Let m˜=H˜H˜× and n=ĤĤ×. Then

1. m˜=n ;

2. For every xn, we have xnm for all sufficiently large nN;

3. If (H:Ĥ), then Ĥ is a primary valuation monoid.

Proof.

1. By 2.1, we have H˜Ĥ, so by Lemma 2.3, we infer that m˜n. Since Λ(H)=, Lemma 3.5.3 implies that every element of n has a power lying in m, whence nH˜H˜×=m˜.

2. Let xn. Since Λ(H)=, Lemma 3.5.3 implies that there is a kN such that xkm. Since H is primary, there is a q0N such that xq0k+r=(xk)q0xrm for all r[0,k1]. If nN with nq0k, then n=qk+r, where qq0 and r[0,k1], and xn=xk(qq0)xq0k+rm.

3. By Lemma 3.4, Ĥ is primary.

Assume to the contrary, that Ĥ is not a valuation monoid. Thus, there exists an element xq(H) such that x,x1Ĥ. If s(H:Ĥ), then snx,snx1H for all sufficiently large nN. Hence, there exists c(H:Ĥ) such that cx,cx1H. For each kN, let nkN be the smallest integer such that cnkxkH, and let n̂kN be the smallest integer such that cn̂kxkĤ. Thus, by definition, 1n̂knkk,n̂k+1n̂k+1. Also, cnkxk is not divisible by c in H: For nk = 1, this holds since xkH, and for nk>1, this follows from the minimality of nk. Since cĤH, we obtain that nkn̂k+1, whence n̂knkn̂k+1. As Ĥ is root-closed by Lemma 2.2.1, we obtain by Lemma 2.1.1 (applied to Ĥ) that the sequence n̂k is increasing. Thus, we infer that |nk+1nk||n̂k+2n̂k|2. Since (H:Ĥ), Lemma 2.2 (items 2 and 3) implies that Ĥ is completely integrally closed and that, for every mN,cmxkĤ for just finitely many k’s. This implies that limkn̂k= whence limknk=.

Proceeding in the same way with the element x1 as with the element x, we obtain a sequence nk having all the properties of the sequence (nk)k1 with respect to the element x1. Then for all kN, the element cnk+nk=(cnkxk)(cnkxk) is a product of two elements not divisible by c in H.

Let nN and let kN be maximal such that nk+nkn. Then nk+nknnk+1+nk+1nk+nk+4. This implies that cn=cnk+nkcf=(cnkxk)(cnkxk)cffor some f[0,4],

whence Λ(cn)Λ(c)+Λ(c)+Λ(cf). Thus, Lemma 3.5.2 implies that Λ(H)<, a contradiction. □

Theorem 3.7.

Let (R, M) be a strongly primary domain. If (R:R̂)=(0), then Λ(R)<.

Proof.

Assume to the contrary that Λ(R)=. Let n=R̂(R̂)×. Choose a nonzero element cM. Let n be a positive integer. Since (R:R̂)=(0), we infer that (cn1n)R̂R, whence there exists an element xn such that cn1xR.

Since Λ(R)=, Lemma 3.6.2 implies that xiM (equivalently, xiR) for all sufficiently large integers iN. Let iN be maximal such that cn1xiR. Since R is primary, there exists a maximal non-negative integer k such that cn1ckxiM. Set y=ckxi, so cn1yR. We have cnyjM for all jN and cn1yjM for all j > 1 since cn1xjm for all j > i. We have yeM for some positive integer e. Hence (1y)(1+y++ye1)=1yeR×.

Thus cn(1y)R, andcn1y=cn1+y++ye11yeR.

We see that cn(1y) and cn1y are not divisible by c in R. Thus, c2n is a product of two elements that are not divisible by c in R: c2n=(cn(1y)) (cn1y).

Hence, Λ(c2n)<M(c)+M(c). By Lemma 3.5.2 applied to c2 replacing c, we conclude that Λ(R)<, a contradiction. □

Theorem 3.7 is false for monoids by Example 3.17.1 below.

Theorem 3.8.

  1. Let (H,m) be a strongly primary monoid, and let f=(H:Ĥ). Then the first two conditions below are equivalent, and each of the first four conditions implies its successor. Moreover, if f, then all nine conditions are equivalent.

    1. H is globally tame.

    2. uA(H)uH.

    3. ρ(H)<.

    4. ρk(H)< for all kN.

    5. Λ(H)=.

    6. Ĥ is a primary valuation monoid.

    7. Ĥ is a valuation monoid.

    8. fmuA(H)uĤ.

    9. fm2uA(H)uH.

  2. Let (R, M) be a strongly primary domain, and f=(R:R̂). All the following nine conditions are equivalent.

    1. R is globally tame.

    2. uA(H)uH.

    3. ρ(R)<.

    4. ρk(R)< for all kN.

    5. Λ(R)=.

    6. R̂ is a primary valuation domain and f(0).

    7. R̂ is a valuation domain and f(0).

    8. (0)fmuA(R)uR̂.

    9. (0)fm2uA(R)uR.

Proof.

(a)

1. (1)(2)

By assumption ω(H)<. Let cm. Then for every aA(H), there is an naN such that a|cna whence a|cω(H). This implies that cω(H)uA(H)uH.

(2)(1) Since H is strongly primary, we have mkuA(H)uH for some positive integer k. Since for every atom u, we have ω(u)k, it follows that ω(H)<, so that H is globally tame.

Thus, the first two conditions are equivalent.

(1)(3) See [Citation18, Theorem 1.6.6].

(3)(4) See EquationEquation (2.2).

(4)(5) This follows from Proposition 3.2.2.

Now assume that f(0).

(5)(6) See Lemma 3.6.3.

(6)(7) Obvious.

(7)(8) Let u be an atom in H. Since ufm and fm is an s-ideal of the valuation monoid Ĥ, it follows that fmuĤ. Thus, the assertion follows.

(8)(9)

Let u be an atom in H. Since fmuĤ, we infer that f2mufĤuH. The assertion follows.

(9)(2) Obvious.

(b) By Theorem 3.7, condition (5) implies that f(0). Thus, each of the first five conditions implies that f(0). Obviously, each of the other four conditions implies that f(0). By item a., all the nine conditions are equivalent.

Theorem 3.9.

1. Let H be a strongly primary monoid.

  1. If Λ(H)<, then H is locally tame, but not globally tame.

  2. If Λ(H)= and (H:Ĥ), then H is globally tame.

  3. H is locally tame if either Λ(H)<, or (H:Ĥ).

2. Let R be a strongly primary domain.

  1. R is locally tame.

  2. R is globally tame if and only if Λ(R)=.

Proof.

1.

(a) Let uA(Hred), a be a multiple of u, and let a=v1vn be a product of atoms. There exists a subproduct of v1vn of length ω(u) that is divisible by u. This subproduct has a refactorization of the form ub where b is a product of at most Λ(H) atoms. Hence, t(u)max{ω(u),Λ(H)+1}<, so H is locally tame.

The monoid H is not globally tame by the implication (1)(5) of Theorem 3.8.

(b) If (Ĥ:H), then all nine conditions of Theorem 3.8 (a) are equivalent. In particular, if Λ(H)= (condition (5)), then H is globally tame (condition (1)).

(c)This follows immediately from the previous two items.

2. By Theorem 3.7, if (R:R̂)=(0), then Λ(R)<. Hence, item 2. (for domains) follows from item 1.(for monoids).

In the next corollary, we answer in the positive Problem 38 in [Citation7].

Corollary 3.10.

A one-dimensional local Mori domain R is locally tame. Moreover, R is globally tame if and only if Λ(R)=.

Proof.

A one-dimensional local Mori domain is strongly primary. Thus, the corollary follows from Theorem 3.9.2. □

In the next proposition, we deal with two significant special cases of strongly primary monoids.

Proposition 3.11.

Let (H,m) be a strongly primary monoid.

1. Let H be seminormal. Then (H:Ĥ)m, so H is locally tame, and all conditions of Theorem 3.8 (a) are equivalent. If Ĥ is Krull, then H is Mori.

2. Let H be Mori with (H:Ĥ). Then Ĥ is Krull, Ĥred is finitely generated, and all conditions of Theorem 3.8 (a) are equivalent to Ĥ being a discrete valuation monoid.

Proof.

1. We have (H:Ĥ)m by Lemma 2.5.2. Thus, H is locally tame (Theorem 3.9), and the conditions of Theorem 3.8 (a) are equivalent. If Ĥ is Krull, then H is Mori by [Citation32, Lemma 2.6].

2. Suppose that H is Mori and (H:Ĥ). Then, Ĥ is Krull and Ĥred is finitely generated by [Citation18, Theorem 2.7.9]. Since (H:Ĥ), all conditions of Theorem 3.8 (a) are equivalent to Ĥ being a valuation monoid (condition (7)). Since Krull monoids are atomic, Lemma 3.3.3 implies that Krull valuation monoids are discrete. □

Proposition 3.12.

Let H be a strongly primary monoid that satisfies one of the following two properties.

1. H is not locally tame.

2. Λ(H)= and (H:Ĥ)=.

Then H is not the multiplicative monoid of a domain.

Proof.

For (1) see Theorem 3.9 (b). For (2), see Theorem 3.8 (b). □

For a strongly primary Mori monoid that satisfies both conditions of Proposition 3.12, see Example 3.17.1 below or [Citation16, Proposition 3.12].

Remark 3.13.

Let R be a strongly primary domain. Then, Theorems 3.7 and Citation3.Citation8 show that either (Λ(R)< and ρ(R)=) or (ρ(R)< and Λ(R)=). This was proved for one-dimensional local noetherian domains in [Citation27, Corollary 3.7], and it was assumed as an additional abstract property for strongly primary monoids in the study of weakly Krull domains in [Citation28, Corollary 4.11].

We present some examples related to the complete integral closure of a strongly primary domain. Let R be a strongly primary domain such that (R:R̂)(0). By Theorem 3.8 (b), if R̂ is a valuation domain, then R̂ is primary. The converse is false as shown in Example 3.15 below. Moreover, R̂ is strongly primary if and only if R̂ is a discrete valuation domain by Lemma 3.3.3. In Example 3.16, R̂ is a valuation domain, but it is not discrete. On the positive side, R̂ is a discrete valuation domain if R is Mori by Proposition 3.11.2. Clearly, the domain R̂ need not be primary. In particular, if R is a one-dimensional local noetherian domain, then R̂, which is equal to the integral closure of R, is not necessarily a local ring.

For Example 3.15 below, we need Proposition 3.14.

Proposition 3.14.

1. Let T be a primary monoid. There exists a strongly primary submonoid (H,m) of T such that T=Ĥ and (H:T) if and only if T is completely integrally closed. Moreover, in this case, the monoid (H,m) can be chosen such that m is a principal s-ideal of T, whence T=(m:m) and m(H:T). The conditions of Theorem 3.8 (a) are satisfied if and only if T is a discrete valuation monoid, and just in this case we may choose H = T.

2. Let T be a primary domain. There exists a strongly primary subring (R, M) of T such that T=R̂ and (R:T)(0) if and only if T is completely integrally closed. Moreover, in this case, the ring (R, M) can be chosen such that T=R̂ and the ideal MT of T is generated by two elements. If furthermore, T contains a field, then we may choose R such that M is a principal ideal of T, whence T=(M:M) and M(R:T). The conditions of Theorem 3.8 (b) are satisfied if and only if T is a discrete valuation domain, and just in this case we may choose R = T.

Proof.

By Lemma 2.2.2, if H is a monoid such that (H:Ĥ), then Ĥ is completely integrally closed. Thus, in both statements, we just have to prove the converse. We start with the proof of the second statement.

2. (i) Let n=TT×, and let c be an element of n. Let H=cT{1}, and let m=cT. Thus, (H,m) is a submonoid of (T,n). Since T is primary, if xH, then, cnxT for some integer nN. Thus, mn+1cn+1TxcT=cm. It follows that H is a strongly primary monoid. Also (H:T)m. Hence TĤ. Since T is completely integrally closed, we obtain that T=Ĥ. For the last sentence see Lemma 3.3.3.

2. (ii) The domain T is local since it is primary. Let N be the maximal ideal of T. Let c be a nonzero element of N, and let F be the prime field contained in the quotient field of T. Set A=(F+cT)T=(FT)+cT, thus A is a subring of T. Let P=AN=(FN)+cT, so P is a prime ideal of A. Set R = AP, and M = PAP. Thus, (R, M) is a local subring of (T, N).

If T contains a field (e.g. if T has finite characteristic), then P=cT=PAP=M, thus M is a principal ideal of T. Otherwise, T has zero characteristic, and we may identify F=Q. Thus, FT is a local subring of Q, that is, a localization of Z at a nonzero prime ideal. Hence, FT is a discrete valuation domain. Let d be a generator of the maximal ideal of FT. Thus, MT=cT+dT.

Let x be a nonzero element of M, since T is primary, and MT is a finitely generated ideal of T, we have (MT)kcxTxR for some positive integer k. Hene, MkxR, so (R, M) is strongly primary. Since (0)cT(R:R̂), we infer that TR̂. Since T is completely integrally closed, we obtain that T=R̂.

For the last sentence of item 2. see Lemma 3.3.3.

1. The first statement follows from the second one. Indeed, we may use the monoid H=R, so Ĥ=T, where R and T are the domains in 2. □

Example 3.15.

  1. There is a strongly primary monoid (H,m) such that Ĥ is primary completely integrally closed, but not a valuation monoid. Moreover, m is a principal s-ideal of Ĥ, so Ĥ=(m:m), abd m(H:Ĥ). Thus, none of the conditions of Theorem 3.8 (a) holds, in particular Λ(H)<.

  2. For any field k, there is a strongly primary domain (R, M) containing k such that R̂ is a primary completely integrally closed domain, but not a valuation domain. Moreover, M is a principal ideal of R̂, so R̂=(M:M), and M(R:R̂). Thus, none of the conditions of Theorem 3.8 (b) holds, in particular Λ(R)<.

Proof.

First we prove the existence of a domain as in item 2. There exists a primary completely integrally closed domain T containing k that is not a valuation domain (we refer to [Citation29, Citation33] and [Citation30], and briefly sketch the idea. Indeed, the field K of Puiseux series over the algebraic closure k¯ of k is algebraically closed and it has a discrete valuation that vanishes on k¯ with value group Q. As follows from the cited papers, this fact implies the existence of a primary completely integrally closed domain containing k¯). By Proposition 3.14.2., there exists a subring R of T that is strongly primary and such that T=R̂ and M is a principal ideal of T. Since R̂ is not a valuation domain, by Theorem 3.8 (b), none of the nine conditions of this theorem are satisfied, in particular, Λ(R)<.

As for item 1., we define H=R. so Ĥ=T, where R is the domain in item 2. □

Example 3.16.

There is a strongly primary domain (R, M) such that R̂ is a primary valuation domain, but not strongly primary and (R:R̂)=M is a principal ideal of R̂. Thus, all the conditions of Theorem 3.8 (b) are satisfied, in particular, Λ(R)=.

Proof.

Let F be a field, A=F[Xq|q rational,q1], and let P be the maximal ideal of A generated by the set {Xq|q rational ,q1}. We set R = AP and M = PRP. Clearly, each nonzero element rR is of the form r=vXq, where q1 is rational and v is a unit in R, whence Mq+1rR. Thus, R is strongly primary. It is easy to show that R̂ is equal to BQ, where B=F[Xq|q rational,q>0] and Q is the maximal ideal of B generated by the set {Xq|q rational ,q>0}. Each nonzero element of R̂ is of the form uXq, where q > 0 is rational and u is a unit in R̂. Clearly, R̂ is a valuation domain and (R:R̂)=M=XR̂.

As mentioned in Section 1, Krull monoids with finite class group and C-monoids are locally tame. Furthermore, finitely generated monoids are locally tame ([Citation18, Theorem 3.1.4]) and hence, the same is true for Cohen-Kaplansky domains and their monoids of invertible ideals ([Citation3, Theorem 4.3]). On the other hand, examples of monoids or domains, that are not locally tame, are rare in the literature. Thus, we end this section with a brief overview.

Example 3.17.

1. In contrast to Theorems 3.7 and Citation3.Citation8 (b), there is a strongly primary monoid H with the following properties (see [Citation22, Proposition 3.7 and Example 3.8]):

  1. H is Mori with (H:Ĥ)= and Ĥ is a discrete valuation monoid,

  2. ρ(H)=Λ(H)=c(H)= and H is not locally tame.

Moreover, H is a submonoid of a one-dimensional local noetherian domain, although H is not the multiplicative monoid of a domain since H is not locally tame.

2. For every αRQ, the additive monoid Hα={(x,y)N2|y<αx}{(0,0)}(N02,+) is a root-closed primary BF-monoid which is neither Mori, nor strongly primary, nor locally tame ([Citation21, Example 4.7]).

3. The additive monoid H={(a,b,c)N03|a>0orb=c}(N03,+) is Mori with catenary degree c(H)=3 (whence, by (2.4), all sets of lengths are arithmetical progressions with difference 1) but it is not locally tame ([Citation25, Example 1]).

4. The domain R=Q[X2,X3] is one-dimensional noetherian, R¯=Q[X] is factorial, and (R:R¯)=(X2). Nevertheless, R fails to be locally tame ([Citation20, Example 6.11]).

5. Let H be a Krull monoid with infinite cyclic class group. Then, H is locally tame if and only if its catenary degree is finite if and only if its set of distances is finite ([Citation17, Theorem 4.2]).

6. Let H be a Krull monoid with class group G such that every class contains a prime divisor. Then, H is locally tame if and only if H is globally tame if and only if G is finite ([Citation21, Theorem 4.4]).

Funding

This work was supported by the Austrian Science Fund FWF, Project Number P28864-N35.

References

  • Anderson, D. D., eds. (1997). Factorization in integral domains. Lecturer Notes in Pure & Applied Mathematics, Vol. 189. New York: Marcel Dekker.
  • Anderson, D. D., Anderson, D. F., Zafrullah, M. (1990). Factorization in integral domains. J. Pure Appl. Algebra 69(1):1–19. DOI: 10.1016/0022-4049(90)90074-R.
  • Anderson, D. D., Mott, J.L. (1992). Cohen-Kaplansky domains: Integral domains with a finite number of irreducible elements. J. Algebra 148(1):17–41. DOI: 10.1016/0021-8693(92)90234-D.
  • Baeth, N. R., Geroldinger, A., Grynkiewicz, D. J., Smertnig, D. (2015). A semigroup-theoretical view of direct-sum decompositions and associated combinatorial problems. J. Algebra Appl. 14(02):1550016–1550060. DOI: 10.1142/S0219498815500164.
  • 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.
  • Barucci, V. (2000). Mori Domains, Non-Noetherian Commutative Ring Theory, Mathematics and Its Applications, Vol. 520. Dordrecht: Kluwer Academic Publishers, pp. 57–73.
  • Cahen, P.-J., Fontana, M., Frisch, S., Glaz S. (2014). Open problems in commutative ring theory. In: Fontana, M., Frisch, S., Glaz, S., eds. Commutative Algebra. New York, NY: Springer, pp. 353–375.
  • Chapman, S. T., Corrales, M., Miller, A., Miller, C., Patel, D. (2014). The catenary and tame degrees on a numerical monoid are eventually periodic. J. Aust. Math. Soc. 97(3):289–300. DOI: 10.1017/S1446788714000330.
  • Chapman, S. T., García-Sánchez, P. A., Llena, D. (2009). The catenary and tame degree of numerical monoids. Forum Math. 21(1):117–129. DOI: 10.1515/FORUM.2009.006.
  • 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.
  • Fontana, M., Houston, E., Lucas, T. (2013). Factoring ideals in integral domains. Lecture Notes of the Unione Matematica Italiana, Vol. 14. Berlin: Springer.
  • Gao, W., Geroldinger, A., Schmid, W. A. (2015). Local and global tameness in Krull monoids. Commun. Algebra 43(1):262–296. DOI: 10.1080/00927872.2014.897585.
  • García-Sánchez, P. A. (2016). An overview of the computational aspects of nonunique factorization invariants. In: Chapman, S., Fontana, M., Geroldinger, A., Olberding, B., eds. Multiplicative Ideal Theory and Factorization Theory. Cham: Springer, pp. 159–181.
  • Geroldinger, A. (1993). The complete integral closure of monoids and domains. Pure Math. Appl. 4(2):147–165.
  • Geroldinger, A. (1996). On the structure and arithmetic of finitely primary monoids. Czech. Math. J. 46(4):677–695. http://eudml.org/doc/30338
  • Geroldinger, A., Gotti, F., Tringali, S. (2019). On strongly primary monoids with a focus on Puiseux monoids. https://arxiv.org/abs/1910.10270
  • Geroldinger, A., Grynkiewicz, D. J., Schaeffer, G. J., Schmid, W. A. (2010). On the arithmetic of Krull monoids with infinite cyclic class group. J. Pure Appl. Algebra 214(12):2219–2250. DOI: 10.1016/j.jpaa.2010.02.024.
  • Geroldinger, A., Halter-Koch, F. (2006). Non-Unique Factorizations. Algebraic, Combinatorial and Analytic Theory. Pure and Applied Mathematic, vol. 278. Boca Raton, FL: Chapman & Hall/CRC.
  • Geroldinger, A., Halter-Koch, F., Hassler, W., Kainrath, F. (2003). Finitary monoids. Semigroup Forum 67(1):1–21. DOI: 10.1007/s002330010141.
  • Geroldinger, A., Hassler, W. (2008). Arithmetic of Mori domains and monoids. J. Algebra 319(8):3419–3463. DOI: 10.1016/j.jalgebra.2007.11.025.
  • Geroldinger, A., Hassler, W. (2008). Local tameness of v-noetherian monoids. J. Pure Appl. Algebra 212(6):1509–1524.
  • Geroldinger, A., Hassler, W., Lettl, G. (2007). On the arithmetic of strongly primary monoids. Semigroup Forum 75(3):567–587. DOI: 10.1007/s00233-007-0721-y.
  • Geroldinger, A., Kainrath, F. (2010). On the arithmetic of tame monoids with applications to Krull monoids and Mori domains. J. Pure Appl. Algebra 214(12):2199–2218. DOI: 10.1016/j.jpaa.2010.02.023.
  • Halter-Koch, F. (1998). Ideal Systems: An Introduction to Multiplicative Ideal Theory. New York, NY: Marcel Dekker.
  • Halter-Koch, F. (2008). The tame degree and related invariants of non-unique factorizations. Acta Math. Univ. Ostraviensis 16(1):57–68.
  • Halter-Koch, F., Hassler, W., Kainrath, F. (2004). Remarks on the multiplicative structure of certain one-dimensional integral domains. In: Facchini, A., Houston, E., Salce, L., eds. Rings, Modules, Algebras, and Abelian Groups. Lecturer Notes in Pure and Applied Mathematics, Vol. 236. New York, NY: Marcel Dekker, pp. 321–331.
  • Hassler, W. (2002). Arithmetical properties of one-dimensional, analytically ramified local domains. J. Algebra 250(2):517–532. DOI: 10.1006/jabr.2001.9092.
  • Hassler, W. (2004). Arithmetic of weakly Krull domains. Commun. Algebra 32(3):955–968.
  • Nagata, M. (1952). On Krull’s conjecture concerning valuation rings. Nagoya Math. J. 4:29–33. DOI: 10.1017/S0027763000022959.
  • Nagata, M. (1955). Correction to my paper on Krull’s conjecture concerning valuation overrings. Nagoya Math. J. 9:209–212. DOI: 10.1017/S0027763000023436.
  • Omidali, M. (2012). The catenary and tame degree of numerical monoids generated by generalized arithmetic sequences. Forum Math. 24(3):627–640. DOI: 10.1515/form.2011.078.
  • Reinhart, A. (2013). On integral domains that are C-monoids. Houston J. Math. 39(4):1095–1116.
  • Ribenboim, P. (1956). Sur une note de nagata relative à un problème de krull. Math. Z. 64(1):159–168. (French). DOI: 10.1007/BF01166564.
  • Valenza, R. J. (1990). Elasticity of factorizations in number fields. J. Num. Theory 36(2):212–218. DOI: 10.1016/0022-314X(90)90074-2.