958
Views
2
CrossRef citations to date
0
Altmetric
Research Article

On the arithmetic of stable domains

, & ORCID Icon
Pages 4763-4787 | Received 19 Dec 2020, Accepted 10 May 2021, Published online: 30 Jun 2021

Abstract

A commutative ring R is stable if every non-zero ideal I of R is projective over its ring of endomorphisms. Motivated by a paper of Bass in the 1960s, stable rings have received wide attention in the literature ever since then. Much is known on the algebraic structure of stable rings and on the relationship of stability with other algebraic properties such as divisoriality and the 2-generator property. In the present paper, we study the arithmetic of stable integral domains, with a focus on arithmetic properties of semigroups of ideals of stable orders in Dedekind domains.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

Motivated by a paper of Bass [Citation6], Lipman, Sally, and Vasconcelos [Citation36, Citation54] introduced the concept of stable ideals and stable rings, which has received wide attention in the literature ever since then. In the present paper, we restrict to integral domains. Let R be a commutative integral domain. A non-zero ideal IR is stable if it is projective over its ring of endomorphisms (equivalently, if it is invertible as an ideal of the overring (I:I) of R). The domain R is called (finitely) stable if every non-zero (finitely generated) ideal of R is stable. By definition, invertible ideals are stable and this implies that Dedekind domains are stable and Prüfer domains are finitely stable. On the other hand, stable domains need neither be noetherian, nor one-dimensional, nor integrally closed. For background on stable rings, their applications, and for results till 2000 we refer to the survey [Citation43] by Olberding. Since then stable rings and domains were studied in a series of papers by Bazzoni, Gabelli, Olberding, Roitman, Salce, and others [Citation13–15, Citation41, Citation44–47].

The goal of the present paper is to study the arithmetic of stable domains, by building on the existing algebraic results. Mori domains and Mori monoids play a central role in factorization theory of integral domains. Every Mori domain R is a BF-domain (this means that every non-zero non-unit element aR has a factorization into irreducible elements and the set L(a)N of all possible factorization lengths is finite). For every Mori domain R, the monoid Iv*(R) of v-invertible v-ideals is a Mori monoid. Our starting point is a recent result by Gabelli and Roitman [Citation14] stating that a domain is stable and Mori if and only if it is one-dimensional stable (Proposition 3.1). This implies that stable Mori domains with non-zero conductor to their complete integral closure are stable orders in Dedekind domains (Theorem 3.7), and these domains are in the center of our interest.

In Section 2 we put together some basics on monoids and domains. In Section 3 we first gather structural results on stable domains (Propositions 3.13.5). Then we apply them to domains that are of central interest in factorization theory, namely seminormal domains, weakly Krull domains, and Mori domains. In Section 4, we study semigroups of r-ideals in the setting of ideal systems of cancellative monoids. We derive structural algebraic results and use them to understand when such semigroups of r-ideals are half-factorial. Section 5 contains our main arithmetical results. The main purpose of Section 5 is to highlight the arithmetical advantages of stability in the context of orders in Dedekind domains. In particular, we show that a series of properties, valid in orders in quadratic number fields (which are stable), also hold true for general stable orders in Dedekind domains. The main result of Section 5 is Theorem 5.10. Among others, it states that the monoid of non-zero ideals and the monoid of invertible ideals of a stable order in a Dedekind domain are transfer Krull if and only if they are half-factorial (this means that they are transfer Krull only in the trivial case). This is in contrast to a recent result on Bass rings (which are stable). In [Citation5, Theorem 1.1], Baeth and Smertnig show that the monoid of isomorphism classes of finitely generated torsion-free modules over a Bass ring is always transfer Krull.

2. Background on monoids and domains

We denote by N the set of positive integers and by N0 we denote the set of non-negative integers. For rational numbers a,bQ,[a,b]={xZ|axb} is the discrete interval between a and b. For subsets A,BZ,A+B={a+b|aA,bB} denotes their sumset. The set of distances Δ(A) is the set of all dN for which there is aA such that A[a,a+d]={a,a+d}. If AN, then ρ(A)=sup(A)/min(A)Q1{} is the elasticity of A, and we set ρ({0})=1.

Let H be a multiplicatively written commutative semigroup with identity element. We denote by H× the group of invertible elements of H. We say that H is reduced if H×={1} and we denote by Hred={aH×|aH} the associated reduced semigroup of H. An element uH is said to be cancellative if au = bu implies a = b for all a,bH. The semigroup H is said to be

  • cancellative if all elements of H are cancellative;

  • unit-cancellative if a,uH and a = au implies that uH×.

Clearly, every cancellative monoid is unit-cancellative. We will study semigroups of ideals that are unit-cancellative but not necessarily cancellative.

Throughout, a monoid means a commutative unit-cancellative semigroup with identity element.

For a set P, we denote by F(P) the free abelian monoid with basis P. Elements aF(P) are written in the form a=pPpvp(a), where vp:F(P)N0 is the p-adic valuation. We denote by |a|=pPvp(a)N0 the length of a and by supp(a)={pP|vp(a)>0}P the support of a.

Let H be a monoid. A non-unit aH is said to be an atom (or irreducible) if a = bc with b,cH implies that bH× or cH×. We denote by A(H) the set of atoms of H and we say that H is atomic if every non-unit is a finite product of atoms. Two elements a,bH are called associated if a = bc for some cH×. If a=u1··ukH, where kN and u1,,ukA(H), then k is a factorization length and the set L(a)N of all factorization lengths of a is called the set of lengths of a. For convenience, we set L(a)={0} for aH×. Then L(H)={L(a)|aH} is the system of sets of lengths of H, Δ(H)=LL(H)Δ(L)Nis thesetofdistancesofH,andρ(H)=sup{ρ(L)|LL(H)}R1{}is theelasticityofH .

We say that the elasticity is accepted if ρ(H)=ρ(L) for some LL(H). The monoid H is

  • half-factorial if it is atomic and |L|=1 for all LL(H),

  • an FF-monoid if it is atomic and each element of H is divisible by only finitely many non-associated atoms, and

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

By definition, an atomic monoid is half-factorial if and only if Δ(H)= if and only if ρ(H)=1. FF-monoids are BF-monoids, BF-monoids satisfy the ACCP (ascending chain condition on principal ideals), and monoids satisfying the ACCP are atomic and Archimedean (i.e. n1anH= for all aHH×). If H is atomic but not half-factorial, then we have gcdΔ(H)=minΔ(H).

Suppose that H is cancellative, m=HH×, and let q(H) be the quotient group of H. We denote by

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

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

Then HHĤq(H), and we say that H is seminormal (resp. completely integrally closed) if H=H (resp. H=Ĥ). Let A,Bq(H) be subsets. We set (A:B)={zq(H)|zBA} and A1=(H:A). If AH, then A is a divisorial ideal (or a v-ideal) if A=Av:=(A1)1, and A is an s-ideal if A = AH. If pH is an s-ideal of H, then p is called a prime s-ideal of H if for all x,yH with xyp, it follows that xp or yp. For an s-ideal I of H let I={xH| there is nN such that xnI} denote the radical of I. The monoid H is said to be

  • Mori if it satisfies the ascending chain condition on divisorial ideals,

  • Krull if it is a completely integrally closed Mori monoid,

  • a G-monoid if the intersection of all non-empty prime s-ideals is non-empty,

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

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

  • finitely primary (of rank s and exponent α) if H is a submonoid of a factorial monoid F=F××F({p1,,ps}) such that mp1··psF and (p1··ps)αFH.

Finitely primary monoids and primary Mori monoids are strongly primary. (To see that the last statement is valid let H be a primary Mori monoid, m=HH× and am. Then aH=m=Ev for some finite set EaH. We infer that EnaH for some nN. Therefore, mn(En)vaH, and thus H is strongly primary.) Mori monoids and strongly primary monoids are BF-monoids.

By a domain, we mean a commutative ring with non-zero identity element and without non-zero zero-divisors. Let R be a domain. We denote by R=R{0} the multiplicative monoid of non-zero elements, by R× the group of units, by R¯ the integral closure of R, by R̂ the complete integral closure of R, by X(R) the set of non-zero minimal prime ideals of R, and by q(R) the quotient field of R. An ideal IR is called 2-generated if there are some a,bI such that I=aR+bR. We say that R is atomic (a BF-domain, an FF-domain, a Mori domain, a Krull domain, a G-domain, Archimedean, (strongly) primary, seminormal, completely integrally closed) if its monoid R has the respective property. By [Citation21, Proposition 2.10.7], R is primary if and only if R is one-dimensional and local. The domain R

  • has finite character if every non-zero element is contained in only finitely many maximal ideals.

  • is divisorial if every non-zero ideal is divisorial.

  • is h-local if R has finite character and every non-zero prime ideal of R is contained in a unique maximal ideal of R.

One-dimensional Mori domains have finite character by [Citation14, Lemma 3.11].

Let S be an integral domain such that RS is a subring. Then R is an order in S if q(R)=q(S) and S is a finitely generated R-module. Moreover, the following statements are equivalent if R is not a field [Citation21, Theorem 2.10.6]:

  • R is an order in a Dedekind domain.

  • R is one-dimensional noetherian and the integral closure R¯ of R is a finitely generated R-module.

The extension RS is quadratic if xyxR+yR+R for all x,yS; equivalently, every R-module between R and S is a ring. If RS is quadratic, then, for every xS, we have x2xR+R; that is, every xS is a root of a monic polynomial of degree at most 2 with coefficients in R. Thus every quadratic extension is an integral extension.

3. Stable domains

In this section, we first gather main properties of stable domains (Propositions 3.13.5). Then we analyze what consequences stability has on some key classes of domains studied in factorization theory, including seminormal domains, weakly Krull domains, G-domains, and Mori domains (Theorems 3.6 and 3.7).

Let R be a domain. A non-zero ideal IR is stable if it is invertible as an ideal of the overring (I:I) of R. The domain is called (finitely) stable if every non-zero (finitely generated) ideal of R is stable. Since invertible ideals are obviously stable, Dedekind domains are stable and Prüfer domains are finitely stable. Conversely, if R is completely integrally closed and stable, then R=(I:I) for every non-zero ideal IR, whence every non-zero ideal is invertible in R and R is a Dedekind domain. Recall that R is an almost Dedekind domain if Rm is a Dedekind domain for each mmax(R). Every almost Dedekind domain is a completely integrally closed Prüfer domain, and thus it is finitely stable. Nevertheless, R is a Dedekind domain if and only if R is a stable almost Dedekind domain. In particular, every almost Dedekind domain that is not a Dedekind domain is not stable. For an example of an almost Dedekind domain that is not a Dedekind domain we refer to [Citation37, Example 35, page 290]. We recall that stable domains need neither be noetherian, nor integrally closed, nor one-dimensional [Citation43, Sections 3 and 4], and we use without further mention that overrings of stable domains are stable [Citation44, Theorem 5.1].

Proposition 3.1.

Let R be a domain that is not a field. Then the following statements are equivalent.

  1. R is a one-dimensional stable domain.

  2. R is a finitely stable Mori domain.

  3. R is a stable Mori domain.

Proof.

This is due to Gabelli and Roitman. More precisely, the equivalence of (a) and (b) is proved in [Citation14, Theorem 4.8]. Clearly, (c) implies (b). If (a) and (b) hold, then (c) holds by [Citation14, Proposition 4.4]. □

Examples given by Olberding in [Citation41, Citation46] show that one-dimensional stable domains need not be noetherian. The ring Int(Z) of integer-valued polynomials is a two-dimensional completely integrally closed Prüfer domain and a BF-domain. Int(Z) is finitely stable (as it is Prüfer) but not stable (as it is not Dedekind). Thus in Statement (b), the property “Mori” cannot be replaced by “BF”. In Example 3.9.3 we show that “Mori” cannot be replaced by “BF” in Statement (c) even if R is a Prüfer domain. Next we consider the local case.

Corollary 3.2.

Let R be a local domain that is not a field.

  1. The following statements are equivalent.

    1. R is a one-dimensional stable domain.

    2. R is a primary stable domain.

    3. R is a stable Mori domain.

    4. R is a strongly primary stable domain.

If these conditions hold and (R:R¯)={0} (for example, see [Citation47, Theorem 2.13]), then R¯ is a discrete valuation domain.

  • 2. If R is one-dimensional, then the following statements are equivalent.

    1. R is stable.

    2. R is finitely stable with stable maximal ideal.

    3. R¯ is a quadratic extension of R and R¯ is a Dedekind domain with at most two maximal ideals.

  • 3. If R is finitely stable with stable maximal ideal m, then the following statements are equivalent.

    1. nNmn={0}.

    2. R is a BF-domain.

    3. R satisfies the ACCP.

    4. R is Archimedean.

Proof.

1. Since R is one-dimensional and local if and only if R is primary, Conditions (a) and (b) are equivalent. Conditions (a) and (c) are equivalent by Proposition 3.1. Obviously, Condition (d) implies Condition (b). Since primary Mori monoids are strongly primary by [Citation23, Lemma 3.1], Conditions (b) and (c) imply Condition (d). If (a)–(d) hold and (R:R¯)={0}, then R¯ is a discrete valuation domain by [Citation44, Corollary 4.17].

2. See [Citation47, Theorem 4.2].

3. (a) (b) This is an immediate consequence of [Citation21, Theorem 1.3.4].

(b) (c) This follows from [Citation21, Corollary 1.3.3].

(c) (d) This is clear (e.g. see page 2 of [Citation15]).

(d) (a) This follows from [Citation15, Proposition 2.12]. □

Let R be a domain. By Corollary 3.2.1, every strongly primary stable domain is Mori. This is not true for general strongly primary domains [Citation26, Section 3] and it is in strong contrast to other classes of strongly primary monoids [Citation19, Theorem 3.3]. By [Citation15, Example 5.17], there exists a stable two-dimensional Archimedean local integral domain. We infer by Corollary 3.2.3 that such a domain is a BF-domain. In particular, a local stable BF-domain need not satisfy the equivalent conditions of Corollary 3.2.1.

Note that if R is a local domain whose ideals are 2-generated, then R is finitely stable with stable maximal ideal (e.g. see Proposition 3.5.4) and the equivalent conditions in Corollary 3.2.3 are satisfied (since R is noetherian). Nevertheless, such a domain is (in general) neither half-factorial nor an FF-domain. In what follows, we present suitable counterexamples.

Let K be a quadratic number field with maximal order OK and p be a prime number such that p is split (i.e. pOK is the product of two distinct prime ideals of OK). (For instance, let K=Q(7) and p = 2.) Let O be the unique order in K with conductor pOK and let p be a maximal ideal of O that contains the conductor. Set S=Op. Then S is a local domain whose ideals are 2-generated and there are precisely two maximal ideals of S¯ that are lying over the maximal ideal of S. It follows from [Citation21, Theorem 3.1.5.2] that S is not half-factorial. (Note that S is a finitely primary monoid of rank two, and thus it has infinite elasticity by [Citation21, Theorem 3.1.5.2]. Therefore, it cannot be half-factorial.)

Let T=R+XC[[X]]. Then T is a local domain with maximal ideal XC[[X]] and every ideal of T is 2-generated by Corollary 3.2.2 and Proposition 3.5.4. Observe that T is not an FF-domain, since aX,a1XA(T) and X2=(aX)(a1X) for each aC{0}.

We do not know whether a local atomic finitely stable domain with stable maximal ideal satisfies the equivalent conditions in Corollary 3.2.3.

Proposition 3.3.

Let R be a domain.

  1. R is finitely stable if and only if RR¯ is a quadratic extension, R¯ is Prüfer, and there are at most two maximal ideals of R¯ lying over every maximal ideal of R.

  2. A semilocal Prüfer domain is stable if and only if it is strongly discrete.

  3. R is an integrally closed stable domain if and only if R is a strongly discrete Prüfer domain with finite character if and only if R is a generalized Dedekind domain with finite character.

  4. An integrally closed one-dimensional domain is stable if and only if it is Dedekind.

Proof.

Recall that a Prüfer domain R is strongly discrete provided that no non-zero prime ideal P of R satisfies P=P2.

  1. [Citation45, Corollary 5.11].

  2. See [Citation2, Proposition 2.10] and [Citation13, Proposition 2.5].

  3. It is an immediate consequence of [Citation40, Theorem 4.6] that R is an integrally closed stable domain if and only if R is a strongly discrete Prüfer domain with finite character. Moreover, it follows from [Citation13, Corollary 2.13] that R is integrally closed and stable if and only if R is a generalized Dedekind domain with finite character.

  4. Since a domain is Dedekind if and only if it is generalized Dedekind of dimension one [Citation12, Proposition 2.1], this follows from 3. □

Proposition 3.3.4 characterizes integrally closed stable domains, that are one-dimensional. However, there are, for every nN, n-dimensional local stable valuation domains ([Citation15, Example 5.11], and recall that valuation domains are integrally closed).

Lemma 3.4.

Let R be a local domain with maximal ideal m such that R is not a field.

  1. If R is noetherian, then R is divisorial if and only if R is one-dimensional and m1/R is a simple R-module.

  2. If R is seminormal and one-dimensional, then (R:R̂)m.

Proof.

1. This follows from [Citation7, Theorem A].

2. This is an immediate consequence of [Citation24, Lemma 3.3]. □

Proposition 3.5.

Let R be a domain.

  1. R is divisorial if and only if R is h-local and Rm is divisorial for every mmax(R).

  2. R is stable if and only if R is of finite character and Rm is stable for every mmax(R).

  3. R is a divisorial Mori domain if and only if R is of finite character and Rm is a divisorial Mori domain for every mmax(R).

  4. Every ideal of R is 2-generated if and only if R is a divisorial stable Mori domain. If R is a stable Mori domain with (R:R¯){0}, then R is divisorial and every ideal of R is 2-generated.

  5. Every ideal of R is 2-generated if and only if R is of finite character and for all mmax(R), every ideal of Rm is 2-generated.

Proof.

1. This follows from [Citation8, Proposition 5.4].

2. This follows from [Citation44, Theorem 3.3].

3. Without restriction assume that R is not a field. First let R be a divisorial Mori domain. It follows by 1. that R is of finite character and Rm is divisorial for all mmax(R). Clearly, Rm is a Mori domain for every mmax(R).

Now let R be of finite character and let Rm be a divisorial Mori domain for every mmax(R). We infer by [Citation53, Théorème 1] that R is a Mori domain. If mmax(R), then Rm is clearly noetherian, and hence Rm is one-dimensional by Lemma 3.4.1. Therefore, R is one-dimensional, and thus R is h-local. Therefore, R is divisorial by 1.

4. We infer by [Citation42, Theorems 3.1 and 3.12] that every ideal of R is 2-generated if and only if R is a noetherian stable divisorial domain. Clearly, R is noetherian and divisorial if and only if R is a divisorial Mori domain, and hence the first statement follows. If R is a stable Mori domain with (R:R¯){0}, then R is at most one-dimensional by Proposition 3.1, and thus every ideal of R is 2-generated by [Citation41, Proposition 4.5].

5. This is an immediate consequence of 2., 3. and 4. □

By Proposition 3.5.4, orders in quadratic number fields are stable because every ideal is 2-generated (for background on orders in quadratic number fields, we refer to [Citation33]). Much research was done to characterize domains, for which all ideals are 2-generated ([Citation8, Theorem 7.3], [Citation29, Theorem 17], [Citation39]). We continue with a characterization within the class of seminormal domains.

Theorem 3.6.

Let R be a seminormal domain. Then the following statements are equivalent.

  1. Every ideal of R is 2-generated.

  2. R is a divisorial Mori domain.

  3. R is a finitely stable Mori domain.

Proof.

Without restriction assume that R is not a field. Note that if R is of finite character, then R is a Mori domain if and only if Rm is a Mori domain for every mmax(R) [Citation53, Théorème 1]. We obtain by Proposition 3.5.3 that R is a divisorial Mori domain if and only if R is of finite character and Rm is a divisorial Mori domain for every mmax(R). Besides that we infer by Propositions 3.1 and 3.5.2 that R is a finitely stable Mori domain if and only if R is of finite character and Rm is a finitely stable Mori domain for every mmax(R). By using Proposition 3.5.5 and the fact that Rm is seminormal for every mmax(R), it suffices to prove the equivalence in the local case. Let R be local with maximal ideal m.

(a) (b) This follows from Proposition 3.5.4.

(b) (c) Observe that R is noetherian, and thus R is one-dimensional by Lemma 3.4.1. We infer that R¯ is a semilocal principal ideal domain, and thus R¯ is a finitely stable Mori domain. In particular, we can assume without restriction that R is not integrally closed. Since RR¯, it follows that (R:R¯)=(R:R̂)=m by Lemma 3.4.2. Since R is not integrally closed, we have that m is not invertible. Therefore, mm1m. Moreover, R¯m=R¯(R:R¯)R, and hence R¯m1. We infer that m1(m:m)R¯m1, and thus R¯=m1.

Consequently, R¯/R is a simple R-module by Lemma 3.4.1. In particular, RR¯ is a quadratic extension. Observe that lR(R¯/m)=lR(R¯/R)+lR(R/m)=2. Set k=|{qmax(R¯)|qR=m}|. Then k=|max(R¯)|. Assume that k3. There are some distinct q1,q2,q3max(R¯). Note that mq1q2q3q1q2q1R¯, and thus lR(R¯/m)3, a contradiction. We infer that k2. It follows from Corollary 3.2.2 that R is finitely stable.

(c) (a) Note that R is a one-dimensional stable domain by Proposition 3.1. It follows from Lemma 3.4.2 that {0}=m(R:R̂)(R:R¯), and thus every ideal of R is 2-generated by Proposition 3.5.4. □

A domain R is said to be weakly Krull if R=pX(R)Rpand the intersection is of finite character, which means that {pX(R)|xRp×} is finite for all xR. Weakly Krull domains were introduced by Anderson, Anderson, Mott, and Zafrullah [Citation1, Citation3], and their multiplicative character was pointed out by Halter-Koch [Citation31, Chapter 22].

Theorem 3.7.

Let R be a domain with (R:R̂){0}, and suppose that R is either weakly Krull or Mori. Then R is stable if and only if every ideal of R is 2-generated. If this holds, then R is an order in a Dedekind domain.

Proof.

If every ideal of R is 2-generated, then R is stable by Proposition 3.5.4. Conversely, let R be stable.

Let us first suppose that R is weakly Krull. Then, for every pX(R),Rp is one-dimensional and stable, whence Mori by Proposition 3.1. Since R is weakly Krull, this implies that R is Mori by [Citation24, Lemma 5.1].

Thus R is Mori in both cases. Using Proposition 3.1 again we infer that R is one-dimensional. Therefore, R¯ is one-dimensional integrally closed and stable, whence R¯ is a Dedekind domain by Proposition 3.3.4. Since (R:R¯){0}, Proposition 3.5.4 implies that every ideal of R is 2-generated. □

Corollary 3.8.

Let R be a seminormal G-domain and suppose that R is either Mori or one-dimensional. Then R is stable if and only if every ideal of R is 2-generated. If this holds, then R is an order in a Dedekind domain.

Proof.

Since R is a seminormal G-domain, (R:R̂){0} by [Citation22, Proposition 4.8]. Thus the claim follows from Theorem 3.7. □

Example 3.9.

  1. There exist integrally closed one-dimensional local Mori domains which are neither valuation domains nor finitely stable. Let K be a field, Y an indeterminate over K, and X an indeterminate over K(Y). Then R=K+XK(Y)[[X]] is an integrally closed one-dimensional local Mori domain which is not completely integrally closed. Thus, R is not a valuation domain. By Proposition 3.3.4, it is not stable because it is not a Dedekind domain, and hence it is not finitely stable by Proposition 3.1.

  2. There exists a seminormal two-dimensional local stable domain. Let pZ be a prime and R=Z(p)+XR[[X]]. Since R[[X]] is a discrete valuation domain with maximal ideal m=XR[[X]] and also Z(p) is a discrete valuation domain with maximal ideal pZ(p); R is a local two-dimensional domain with maximal ideal n=pR (and {0}mn). Now R is stable as well by [Citation41, Theorem 2.6]. Thus R is not Mori by Proposition 3.1.

    Moreover, R is seminormal. Indeed, we know Z(p) is integrally closed in Q and RD=Q+XR[[X]]R[[X]]. Let tq(R)=R((X)) with t2,t3R. Then t2,t3R[[X]], and hence tR[[X]] (since R[[X]] is completely integrally closed). We infer that t0R and t02,t03Q. If t0=0, then tD. If t00, then t0=t03t02Q, that is tD. In any case, D is seminormal. Now clearly R is seminormal in D. Therefore, R is seminormal.

    • 3. There exists a two-dimensional stable Prüfer domain R which is a BF-domain, whence R is a finitely stable BF-domain that is not Mori (cf. Proposition 3.1). To see this we analyze an example given by Gabelli and Roitman. Let K be a field and let X and Y be independent indeterminates over K. Set S=K[Y]YK[Y] and let R=S1(K[{X(1X)nYn,Yn+1(1X)n|nN0}]). Set T=1XY. It is shown in [Citation15, Example 5.13] that R is a two-dimensional stable Prüfer domain that satisfies the ACCP. In particular, R is Archimedean. Moreover, it is shown in [Citation15, Example 5.13] that Y and T are algebraically independent over K and R=S1(K[{(1YT)Tn,YTn|nN0}]).

      Next we prove that S1(K[Y,T,T1])R̂. Observe that T=YTY, and hence T and T1 are elements of the quotient field of R. Since (1YT)TnR and Y(T1)nR for every nN0, we infer that {T,T1}R̂. Clearly, K[Y]RR̂, and thus K[Y,T,T1]R̂. Since S1={s1|sS}RR̂, this implies that S1(K[Y,T,T1])R̂.

      Since Y and T are algebraically independent over K, it follows that K[Y,T] is factorial. Note that K[Y,T,T1] is a quotient overring of K[Y,T], and hence K[Y,T,T1] is factorial. We infer that S1(K[Y,T,T1]) is factorial. Moreover, since RS1(K[Y,T,T1]) and S1(K[Y,T,T1]) is completely integrally closed, we have that R̂S1(K[Y,T,T1]). This implies that R̂=S1(K[Y,T,T1]) is factorial, and thus R̂ is a BF-domain. Since R is Archimedean, it follows that R̂×R=R×, and hence R is a BF-domain by [Citation21, Corollary 1.3.3].

4. Monoids of ideals and half-factoriality

In this section, we study, for a finitary ideal system r of a cancellative monoid H, algebraic and arithmetic properties of the semigroup Ir(H) of r-ideals and of the semigroup Ir*(R) of r-invertible r-ideals. A focus is on the question when these monoids of r-ideals are half-factorial (other arithmetical properties of Ir*(H), such as radical factoriality, were recently studied in [Citation48]). In Section 5, we apply these results to monoids of divisorial ideals and to monoids of usual ring ideals of Mori domains.

Let H be a cancellative monoid and K a quotient group of H. An ideal system on H is a map r:P(H)P(H) such that the following conditions are satisfied for all subsets X,YH and all cH.

  • XXr.

  • XYr implies XrYr.cH{c}r.cXr=(cX)r.

We refer to [Citation31, Citation32] for background on ideal systems. Let r be an ideal system on H. A subset IH is called an r-ideal if Ir = I. Furthermore, a subset JK is called a fractional r-ideal of H if there is some cH such that cJ is an r-ideal of H. We denote by Ir(H) the set of all non-empty r-ideals, and we define r-multiplication by setting I·rJ=(IJ)r for all I,JIr(H). Then Ir(H) together with r-multiplication is a reduced semigroup with identity element H. Let Fr(H) denote the semigroup of non-empty fractional r-ideals, Fr(H)× the group of r-invertible fractional r-ideals, and Ir*(H)=Fr×(H)Ir(H) the cancellative monoid of r-invertible r-ideals of H with r-multiplication. We denote by X(H) the set of all non-empty minimal prime s-ideals of H, by r-spec(H) the set of all prime r-ideals of H, and by r-max(H) the set of all maximal r-ideals of H. We say that r is finitary if Xr=Er, where the union is taken over all finite subsets EX. For a subset Xq(H), we set Xs=XH,Xv=(X1)1and Xt=EX,|E|<Ev .

We will use the s-system, the v-system, and the t-system. For every ideal system r, we have XrXv, and if r is finitary, then XrXt for all XH. We say that H has finite r-character if each xH is contained in only finitely many maximal r-ideals of H.

Let R be a domain with quotient field K and r an ideal system on R (clearly, R is a monoid and r restricts to an ideal system r on R whence for every subset IR we have Ir=(I)r{0}). We denote by Ir(R) the semigroup of non-zero r-ideals of R and Ir*(R)Ir(R) is the subsemigroup of r-invertible r-ideals of R. The usual ring ideals form an ideal system, called the d-system, and for these ideals, we omit all suffices (i.e. I(R)=Id(R) and I*(R)=Id*(R)). For the following equivalent statements, let r be an ideal system on R such that every r-ideal of R is an ideal of R. We say that R is a Cohen-Kaplansky domain if one of the following equivalent statements hold [Citation4, Theorem 4.3] and [Citation25, Proposition 4.5].

  1. R is atomic and has only finitely many atoms up to associates.

  2. Ir(R) is a finitely generated semigroup for some ideal system r on R.

  3. Ir*(R) is a finitely generated semigroup for some ideal system r on R.

  4. R¯ is a semilocal principal ideal domain, R¯/(R:R¯) is finite, and |max(R)|=|max(R¯)|.

Thus, Corollary 3.2.2 and Property (d) imply that a Cohen-Kaplansky domain R is stable if and only RR¯ is a quadratic extension.

Lemma 4.1.

Let H be a cancellative monoid and let r be a finitary ideal system on H such that nN0(mn)r= for every mr-max(H). Then Ir(H) is unit-cancellative and if H is of finite r-character, then Ir(H) is a BF-monoid.

Proof.

Let I,JIr(H) be such that (IJ)r=I. Assume that J is proper. Then Jm for some mr-max(H). It follows by induction that (IJn)r=I for all nN0, and hence InN0(Jn)rnN0(mn)r. Therefore, I=, a contradiction. Consequently, Ir(H) is unit-cancellative.

Now let H be of finite r-character. We have to show that Ir(H) is a BF-monoid.

First we show that Ir(H) is atomic. Since Ir(H) is unit-cancellative it remains to show by [Citation11, Lemma 3.1(1)] that Ir(H) satisfies the ACCP. Assume that Ir(H) does not satisfy the ACCP. Then there is a sequence (Ii)i=0 of elements of Ir(H) such that IiIr(H)Ii+1Ir(H) for all iN0. Consequently, there is some sequence (Ji)i=0 of proper elements of Ir(H) such that Ii=(Ii+1Ji)r for all iN0. Note that I0Ji for all iN0. Since {mr-max(H)|I0m} is finite, there is some mr-max(H) such that {iN0|Jim} is infinite. By restricting to a suitable subsequence of (Ii)iN0, we can therefore assume that Jim for all iN0. Note that I0=(Ini=0n1Ji)r for every nN0, and thus I0(i=0n1Ji)r(mn)r for every nN0. This implies that I0nN0(mn)r, and thus I0=, a contradiction.

Finally, we prove that L(N) is finite for each NIr(H). Let NIr(H) and set M={mr-max(H)|Nm}. Observe that M is finite. For each mM set gm=max{N|N(m)r}. It is sufficient to show that nmMgm for each nL(N). Let nL(N). Clearly, there is a finite sequence (Ai)i=1n of atoms of Ir(H) such that N=(i=1nAi)r. Since [1,n]=mM{i[1,n]|Aim}, we infer that nmM|{i[1,n]|Aim}|mMgm.

Let H be a cancellative monoid and r a finitary ideal system on H. Observe that if H is strictly r-noetherian (for the definition of strictly r-noetherian monoids, we refer to [Citation31, 8.4 Definition, page 87]), then it follows from [Citation31, 9.1 Theorem, page 94] that nN0(mn)r= for every mr-max(H). Also note that if H is a Mori monoid and r-max(H)=X(H), then H is of finite r-character (this is an easy consequence of [Citation21, Theorem 2.2.5.1]).

Proposition 4.2.

Let H be a finitely primary monoid of rank one, m=HH×,q=ĤĤ×, and let r be a finitary ideal system on H.

  • 1. The following statements are equivalent.

  •  (a) H is half-factorial.

  •  (b) uĤ=vĤ for all u,vA(H).

  •  (c) uĤ=q for all uA(H).

  • 2. The following statements are equivalent.

  •  (a) Ir(H) is half-factorial.

  •  (b) AĤ=BĤ for all A,BA(Ir(H)).

  •  (c) AĤ=q for all AA(Ir(H)).

  •  (d) If kN and AiA(Ir(H)) for every i[1,k], then i=1kAi(mk+1)r.

  •  (e) H is half-factorial and for every nonprincipal AA(Ir(H)) it follows that A(m2)r.

Proof.

Since H is finitely primary of rank one, there is some qq such that q=qĤ.

1.(a) 1.(b) Let u,vA(H). There are some k,N such that uĤ=qkĤ and vĤ=qĤ. It follows that uĤ=vkĤ, and hence u=vkε for some εĤ×. Moreover, there is some a(H:Ĥ). Since LH(aεn)[0,vq(a)] for every nN0, there are some n1,n2N0 such that n1<n2 and LH(aεn1)=LH(aεn2). Set b=aεn1 and set n=n2n1. Then nN,b,bεnH and LH(b)=LH(bεn). There is some hN0 such that LH(b)={h}. Note that unb=vnkbεn, and hence {n+h}=LH(unb)=LH(vnkbεn)={nk+h}. We infer that =k, and hence uĤ=qkĤ=qĤ=vĤ.

1.(b) 1.(c) Since (H:Ĥ)=, there is some mN such that qm,qm+1H. There is some N such that uĤ=qĤ for all uA(H). There are some a,bN such that qm is a product of a atoms of H and qm+1 is a product of b atoms of H. We infer that qmĤ=qaĤ and qm+1Ĥ=qbĤ. This implies that b=m+1=a+1, and hence =1 and uĤ=q.

1.(c) 1.(a) Let k,N0, let uiA(H) for every i[1,k] and let vjA(H) for every j[1,] be such that i=1kui=j=1vj. Then qk=i=1kuiĤ=j=1vjĤ=q. Consequently, k=.

2. Note that H is strongly primary and r-max(H)={m}. Therefore, nN0(mn)r=. We infer by Lemma 4.1 that Ir(H) is a unit-cancellative atomic monoid. Since H is r-local, we have that A(Ir*(H))={uH|uA(H)}. Moreover, Ir*(H) is a divisor-closed submonoid of Ir(H). Therefore, {uH|uA(H)}A(Ir(H)). Note that if I is a non-empty s-ideal of H, then IĤ=IrĤ (since IĤ=qmĤ for some mN0, it follows that qmĤ=IĤIrĤItĤ(IĤ)t=(qmĤ)t=qm(Ĥ)t=qmĤ).

2.(a) 2.(b) Let A,BA(Ir(H)). There are some k,N such that AĤ=qk and BĤ=q. This implies that AĤ=BkĤ, and hence (A(H:Ĥ))r=(Bk(H:Ĥ))r. Since (H:Ĥ) is a non-empty r-ideal of H, there is some mN0 such that L((H:Ĥ))={m}. Therefore, {+m}=L((A(H:Ĥ))r)=L((Bk(H:Ĥ))r)={k+m}, and thus =k. We infer that AĤ=qk=q=BĤ.

2.(b) 2.(c) Since (H:Ĥ)=, there is some mN such that qm,qm+1H. There is some N such that AĤ=qĤ for all AA(Ir(H)). Since Ir(H) is atomic, there are some a,bN such that qmH is an r-product of a atoms of Ir(H) and qm+1H is an r-product of b atoms of Ir(H). This implies that qmĤ=qaĤ and qm+1Ĥ=qbĤ. Therefore, b=m+1=a+1, and hence =1 and AĤ=q.

2.(c) 2.(a) Let k,N0, let AiA(Ir(H)) for every i[1,k] and let BjA(Ir(H)) for every j[1,] be such that (i=1kAi)r=(j=1Bj)r. Then qk=(i=1kAi)rĤ=(j=1Bj)rĤ=q. Therefore, k=.

2.(c) 2.(d) Let kN and let AiA(Ir(H)) for every i[1,k]. Assume that i=1kAi(mk+1)r. Note that mA(Ir(H)) (since Ir(H) is unit-cancellative). Therefore, AiĤ=mĤ=q for all i[1,k], and thus qk=(i=1kAi)rĤ(mk+1)rĤ=qk+1, a contradiction.

2.(d) 2.(e) It remains to show that H is half-factorial. Let k,N, let uiA(H) for every i[1,k] and let vjA(H) for every j[1,] be such that i=1kui=j=1vj. Observe that uiH,vjHA(Ir(H)) for all i[1,k] and j[1,]. We infer that i=1kui(m+1)r and j=1vj(mk+1)r. Therefore, k<+1 and <k+1, and hence k=.

2.(e) 2.(c) Let AA(Ir(H)).

Case 1. A is principal. Then A = uH for some uA(H). By 1. we have that AĤ=uĤ=q.

Case 2. A is not principal. Then A(m2)r, and hence there is some vA(m2)r. Observe that vA(H). It follows from 1. that q=vĤAĤq, and thus AĤ=q.

Observe that some of the semigroups (e.g. Ir(H)) in the following result may not always be unit-cancellative. In that case, we apply the original definitions for being an atom or being half-factorial to commutative semigroups with identity (which are not necessarily unit-cancellative).

Proposition 4.3.

Let H be a cancellative monoid and r be a finitary ideal system on H such that H is of finite r-character and r-max(H)=X(H).

  1. Ir(H)mX(H)Irm(Hm) and Ir*(H)mX(H)Irm*(Hm).

  2. Ir(H) is half-factorial if and only if Irm(Hm) is half-factorial for every mX(H) and Ir*(H) is half-factorial if and only if Hm is half-factorial for every mX(H).

  3. If AA(Ir(H)), then AX(H).

  4. For every mX(H) we have that A(Irm(Hm))={Am|AA(Ir(H)),Am}.

Proof.

Claim: For every IIr(H) it follows that I=(qX(H)(IqH))r.

Proof of the claim: Let IIr(H). Since H is of finite r-character, it follows that IqH=H for all but finitely many qX(H). Note that if qX(H) and Iq, then IqH is a q-primary r-ideal of H, and (IqH)q=Iq. Therefore, ((qX(H)(IqH))r)m=(qX(H)(IqH)m)rm=Im. Consequently, I=(qX(H)(IqH))r.

  1. Let f:Ir(H)mX(H)Irm(Hm) be defined by f(I)=(Im)mX(H) for every IIr(H). Since H is of finite r-character it is clear that f is well-defined. It is straightforward to show that f is a monoid homomorphism. If I,JIr(H) are such that Im=Jm for all mX(H), then I=mrmax(H)Im=mrmax(H)Jm=J. Therefore, f is injective. It remains to show that f is surjective. Let (Im)mX(H)mX(H)Irm*(Hm). Set I=(mX(H)(ImH))r. Then IIr(H) and (IqH)q=Iq for every qX(H). Therefore, f is surjective. If IIr*(H), then ImIrm*(Hm) for every mX(H), and thus f|Ir*(H):Ir*(H)mX(H)Irm*(Hm) is a monoid isomorphism.

  2. It is an immediate consequence of 1. that Ir(H) is half-factorial if and only if Irm(Hm) is half-factorial for every mX(H) and Ir*(H) is half-factorial if and only if Irm*(Hm) is half-factorial for every mX(H). Note that if mX(H), then Hm is rm-local, and hence Irm*(Hm)={xHm|xHm}. Clearly, {xHm|xHm}(Hm)red is half-factorial if and only if Hm is half-factorial.

  3. Let AA(Ir(H)). Then Am for some mX(H). Set J=(qX(H){m}(AqH))r. We infer by the claim that A=(J(AmH))r. Since AmH is a proper r-ideal of H this implies that A=AmH. Since Am is mm-primary, we have that AmH is m-primary, and thus A=m.

  4. Let mX(H). First let BA(Irm(Hm)). Set A=BH. Then A is a proper r-ideal of H and B=Am. It remains to show that AIr(H). Let I,JIr(H) be such that A=(IJ)r. Then B=(ImJm)rm, and hence Im=Hm or Jm=Hm. Without restriction let Im=Hm. Then Im. Since A is m-primary and AI, this implies that I = H.

    Now let BA(Ir(H)) be such that Bm. Let I,JIrm(Hm) be such that Bm=(IJ)rm. It is straightforward to check r-locally that B=((IH)(JH))r. Note that IH,JHIr(H), and hence IH=H or JH=H. Without restriction let IH=H. Consequently, I=Hm.

Theorem 4.4.

Let H be a cancellative monoid and let r be a finitary ideal system on H such that H is of finite r-character and Hm is finitely primary for every mr-max(H). Then Ir(H) is half-factorial if and only if Ir*(H) is half-factorial and for every AA(Ir(H))Ir*(H) we have that A((A)2)r.

Proof.

First let Ir(H) be half-factorial. Since Ir*(H) is a divisor-closed submonoid of Ir(H) we have that Ir*(H) is half-factorial. Let mr-max(H). It follows by Proposition 4.3.2 that Irm(Hm) and Hm are half-factorial. Therefore, Hm is finitely primary of rank one by [Citation21, Theorem 3.1.5]. We infer by Proposition 4.2.2 that for every nonprincipal BA(Irm(Hm)) we have that B(mm2)rm. We infer by Proposition 4.3.2 that Ir*(H) is half-factorial. Let AA(Ir(H))Ir*(H). Then Ar-max(H) by Proposition 4.3.3. Without restriction let A=m. It follows by Proposition 4.3 that AmA(Irm(Hm)). If Am is a principal ideal of Hm, then A is r-locally principal, and since H is of finite r-character, A is r-invertible, a contradiction. Therefore, Am is not a principal ideal of Hm and Am(mm2)rm. Since A and (m2)r are m-primary this implies that A(m2)r.

Now let Ir*(H) be half-factorial and let for every AA(Ir(H))Ir*(H),A((A)2)r. Let mr-max(H). It follows from Proposition 4.3.2 that Hm is half-factorial. Consequently, Hm is finitely primary of rank one. Let BA(Irm(Hm)) be not principal. Then B=Am for some AA(Ir(H)) with Am by Proposition 4.3.4. It follows from Proposition 4.3.3 that A=m. Obviously, A is not r-invertible. Therefore, A(m2)r. Since A and (m2)r are m-primary we have that B(mm2)rm. We infer by Proposition 4.2.2 that Irm(Hm) is half-factorial. □

Corollary 4.5.

Let H be a cancellative monoid and let r be a finitary ideal system on H such that H is of finite r-character and Hm is finitely primary and m2 is contained in some proper r-invertible r-ideal of H for every mr-max(H). Then Ir(H) is half-factorial if and only if Ir*(H) is half-factorial.

Proof.

By Theorem 4.4 it is sufficient to show that for every AA(Ir(H))Ir*(H), we have that A((A)2)r. Let AA(Ir(H))Ir*(H). Assume that A((A)2)r. There is some mr-max(H) such that Am. We infer that m2I for some proper IIr*(H). Since Am, it follows that A((A)2)r(m2)rI, and thus A=IIr*(H), a contradiction. □

Lemma 4.6.

Let L be a finite field, let KL be a subfield, let X be an indeterminate over L and let R=K+XL[[X]]. Then R is a local Cohen-Kaplansky domain with maximal ideal XL[[X]] and R is divisorial if and only if [L:K]2.

Proof.

It is an immediate consequence of [Citation4, Corollary 7.2] that R is a local Cohen-Kaplansky domain with maximal ideal XL[[X]]. Set m=XL[[X]]. Without restriction let K=L. Then m1=(m:m)=L[[X]]. Since R is a local one-dimensional noetherian domain we have by [Citation38, Theorem 3.8] that R is divisorial if and only if L[[X]] is a 2-generated R-module. For hL[[X]] let h0 denote the constant term of h.

If L[[X]] is a 2-generated R-module, then L[[X]]=f,gR, whence L=f0,g0K, and so [L:K]=2. Conversely, let [L:K]=2. Then L=1,aK for some aL. Observe that L[[X]]=1,aR.

Example 4.7.

Let L be a finite field, let KL be a subfield, let nN2 and let R=K+XnL[[X]]. Then R is a local Cohen-Kaplansky domain with maximal ideal XnL[[X]], R is not half-factorial and the square of the maximal ideal of R is contained in a proper principal ideal of R.

Proof.

By [Citation4, Corollary 7.2] we have that R is a local Cohen-Kaplansky domain with maximal ideal XnL[[X]] such that R is not half-factorial. Set m=XnL[[X]]. Then m2=X2nL[[X]]XnR and XnR is a proper principal ideal of R. □

5. Arithmetic of stable orders in Dedekind domains

In this section, we derive the main arithmetical results of the paper. For monoids of ideals of stable Mori domains, we study the catenary degree, the monotone catenary degree and we establish characterizations when these monoids are half-factorial and when they are transfer Krull. We demonstrate in remarks and examples that none of the main statements in Theorems 5.9 and 5.10 hold true without the stability assumption.

We need the concepts of catenary degrees, transfer homomorphisms, and transfer Krull monoids. Let H be an atomic monoid. The free abelian monoid Z(H)=F(A(Hred)) denotes the factorization monoid of H and π:Z(H)Hred the canonical epimorphism. For every element aH, Z(a)=π1(aH×) is the set of factorizations of a. Note that L(a)={|z||zZ(a)}N0 is the set of lengths of a. Suppose that H is atomic. If z,zZ(H) are two factorizations, say z=u1··uv1··vmandz=u1··uw1··wn, where ,m,nN0 and all ui,vj,wkA(Hred) such that vjwk for all j[1,m] and all k[1,n], then d(z,z)=max{m,n} is the distance between z and z.

Let aH and NN0{}. A finite sequence z0,,zkZ(a) is called a (monotone) N-chain of factorizations of a if d(zi1,zi)N for all i[1,k] (and |z0||zk| or |z0||zk|). We denote by c(a) (or by cmon(a) resp.) the smallest NN0{} such that any two factorizations z, zZ(a) can be concatenated by an N-chain (or by a monotone N-chain resp.). Then c(H)=sup{c(b)|bH}N0{}andcmon(H)=sup{cmon(b)|bH}N0{}  denote the catenary degree and the monotone catenary degree of H. By definition, we have c(H)cmon(H), and H is factorial if and only if c(H)=0. If H is cancellative but not factorial, then, by [Citation21, Theorem 1.6.3], (5.1) 2+supΔ(H)c(H)cmon(H) ,(5.1) whence c(H)2 implies that H is half-factorial and that 2=c(H)=cmon(H). Let (5.2) HF=F××F({p1,,ps})(5.2) be a finitely primary monoid of rank sN and exponent αN. Then, by [Citation21, Theorem 3.1.5], we have (5.3) If s=1, then ρ(H)2α1 and c(H)3α1.(5.3) (5.4) If s2, then ρ(H)= and c(H)2α+1.(5.4)

A monoid homomorphism θ:HB between monoids is said to be a transfer homomorphism if the following two properties are satisfied.

(T 1) B=θ(H)B× and θ1(B×)=H×.

(T 2) If uH, b, cB and θ(u)=bc, then there exist v, wH such that u = vw, θ(v)bB× and θ(w)cB×.

A monoid H is said to be a transfer Krull monoid if it allows a transfer homomorphism θ to a Krull monoid B. Since the identity map is a transfer homomorphism, Krull monoids are transfer Krull, but transfer Krull monoids need neither be commutative (though here we restrict to the commutative setting), nor Mori, nor completely integrally closed. The arithmetic of Krull monoids is best understood (compared with various other classes of monoids and domains), and a transfer homomorphism allows to pull back arithmetical properties of the Krull monoid B to the original monoid H. We refer to the surveys [Citation18, Citation28] for examples and basic properties of transfer Krull monoids.

All Dedekind domains are transfer Krull and stable. However, there are orders in Dedekind domains that are transfer Krull but not stable (Remark 5.15) and there are orders that are stable but not transfer Krull (all orders in quadratic number fields are stable but not all of them are transfer Krull). Half-factorial monoids are trivial examples of transfer Krull monoids (if H is half-factorial, then θ:H(N0,+), defined by θ(u)=1 for all uA(H) and θ(ε)=0 for all εH×, is a transfer homomorphism). Thus a result (as given in Theorems 5.1 and 5.9), stating that monoids of a given type are transfer Krull if and only if they are half-factorial, means that their arithmetic is different from the arithmetic of Krull monoids and equal only in the trivial case. For recent work on the half-factoriality of transfer Krull monoids we refer to [Citation16].

We start with a result on the finiteness of the catenary degree of weakly Krull Mori domains.

Theorem 5.1.

Let R be a weakly Krull Mori domain.

  1. For every pX(R),c(Rp)<, and ρ(Rp)< if and only if (Rp:Rp̂){0} and Rp̂ is local.

  2. c(Iv(R))=sup{c(Ivp(Rp))|pX(R)} and c(Iv*(R))=sup{c(Ivp*(Rp))|pX(R)}.

  3. If (R:R̂){0}, then c(Iv*(R))c(Iv(R))<.

  4. Iv*(R) is a Mori monoid and it is half-factorial if and only if it is transfer Krull.

Proof.

Since R is a weakly Krull Mori domain, we have t-spec(R)=X(R) by [Citation31, Theorem 24.5]. Thus all assumptions of Proposition 4.3 are satisfied.

  1. Let pX(R). Since Rp is a one-dimensional local Mori domain, it is strongly primary and hence locally tame by [Citation26, Theorem 3.9]. Thus its catenary degree is finite by [Citation19, Theorem 4.1]. If (Rp:Rp̂)={0}, then ρ(Rp)= by [Citation26, Theorem 3.7]. Suppose that (Rp:Rp̂){0}. Then R is finitely primary by [Citation21, Proposition 2.10.7], whence the claim on the elasticity follows from Equation(5.3) and Equation(5.4).

  2. Since the catenary degree of a coproduct equals the supremum of the individual catenary degrees [Citation21, Proposition 1.6.8], the assertion follows from Proposition 4.3.1.

  3. Since Iv*(R) is a divisor-closed submonoid of Iv(R), the inequality between their catenary degrees holds. If (R:R̂){0}, then almost all Rp are discrete valuation domains whence their catenary degree is finite. Thus the claim follows from 2. and from Proposition 4.3.1.

  4. See [Citation28, Proposition 7.3]. □

There are primary Mori monoids H with c(H)= [Citation23, Proposition 3.7], in contrast to the domain case as given in Theorem 5.1.1.

Let H be a finitely primary monoid of rank sN such that there exist some exponent αN of H and some system {pi|i[1,s]} of representatives of the prime elements of Ĥ with the following property: for all i[1,s] and for all aĤ with vpi(a)α we have piaH if and only if aH. Then H is said to be

  • strongly ring-like if Ĥ×/H× is finite and {(vpi(a))i=1s|aHH×}Ns has a smallest element with respect to the partial order.

The concept of strongly ring-like monoids was introduced by Hassler [Citation35], and the question which one-dimensional local domains are strongly ring-like was studied in [Citation25, Section 5].

A numerical monoid is a submonoid of (N0,+) with finite complement, whence numerical monoids are finitely primary of rank one. Conversely, if HF=F××F({p}) is finitely primary of rank one, then its value monoid vp(H)={vp(a)|aH}N0 is a numerical monoid.

Proposition 5.2.

Let R be a local stable Mori domain with (R:R̂){0}. Then R is finitely primary of rank s2 and it is strongly ring-like. If s = 2, then ρ(R)= and if s = 1 and X(R̂)={p}, then the elasticity ρ(R) is accepted with ρ(R)=maxvp(R)/minvp(R).

Proof.

By Corollary 3.2.1, R is one-dimensional. By [Citation21, Proposition 2.10.7], one-dimensional local Mori domains with non-zero conductor are finitely primary of rank |X(R̂)|. By Corollary 3.2.2, R¯ is a Dedekind domain with at most two maximal ideals, whence s=|X(R̂)|2. Since (R:R¯){0}, every ideal of R is 2-generated by Proposition 3.5.4, whence R is noetherian. If m is the maximal ideal of R, then R̂=R¯ and |max(R¯)|2|R/m|, whence R is strongly ring-like by [Citation25, Corollary 5.7]. If s = 2, then ρ(R)= by (5.4). Suppose that s = 1. Since R is strongly ring-like, R̂×/R× is finite and thus the elasticity is accepted and has the asserted value by [Citation27, Lemma 4.1]. □

Let R be a one-dimensional local Mori domain with (R:R̂){0}. If R is stable, then, by Proposition 5.2, we have |X(R̂)|2. Example 5.5 shows that the converse does not hold in general. Example 5.4 and Proposition 5.7.1 show that also for stable domains the exponent of R can be arbitrarily large. We start with a lemma.

Lemma 5.3.

Let R be a Mori domain and a G-domain and let I be a divisorial stable ideal of R. Then I2=xI for some xI.

Proof.

Since every overring of a G-domain is a G-domain, (I:I) is a G-domain. Since I is divisorial and R is a Mori domain, we have that (I:I) is a Mori domain. Therefore, spec((I:I)) is finite by [Citation21, Theorem 2.7.9], and hence (I:I) is semilocal. Consequently, I=x(I:I) for some xI, and thus I2=xI.

Example 5.4

(Stable orders in number fields). 1. Let K=Q(d) be a quadratic number field, where dZ{0,1} is squarefree, and let ω={d, if d2,3mod41+d2, if d1mod4 .

Let R=Z+pnωZ, where pN is a prime number and nN. Since every ideal of R is 2-generated, R is a stable order in the Dedekind domain R¯=Z+ωZ. Then m=pZ+pnωZX(R) and Rm is a one-dimensional local stable domain with non-zero conductor. By Corollary 3.2, Rm is Mori, whence it is finitely primary of rank s=|{qX(R¯)|qR=m}|2. Moreover, if αN is the exponent of Rm, then αmax{vq((R:R¯))|qX(R¯),qR=m} and since (R:R¯)=pnR¯, we obtain that αnmax{vq(pR¯)|qX(R¯),qR=m}n.

2. Let K be an algebraic number field, OK its ring of integers, and ROK an order. If the discriminant Δ(R)Z is not divisible by the fourth power of a prime, then R is stable by a result of Greither [Citation30, Theorem 3.6]. In particular, if aN is squarefree with 3a and R=Z[a3]Q(a3), then Δ(R)=27a2 is not divisible by a fourth power of a prime (for more on R and OK in the case of pure cubic fields, we refer to [Citation34, Theorem 3.1.9]).

Next we discuss the catenary degree of finitely primary monoids, which has received a lot of attention in the literature. Let HF be a finitely primary monoid of rank s and exponent α, with all notation as in Equation(5.2). Then the catenary degree is bounded above by 3α1 in case s = 1 and by 2α+1 otherwise. These bounds can be attained, but the catenary degree can also be much smaller. Indeed, as shown in Example 5.4.2, for every nN there is an order R in a quadratic number field whose localization Rp at a maximal ideal p is finitely primary of exponent greater than or equal to the given n but the catenary degree c(Rp) is bounded by 5 [Citation9, Theorem 1.1]. Let HF=F××F({p}) be finitely primary of rank one, suppose that its value monoid vp(H)={vp(a)|aH}=d1,,ds, with sN,1<d1<<ds, and gcd(d1,,ds)=1. The catenary degree of numerical monoids has been studied a lot in recent literature (see [Citation17, Citation49–52], for a sample). By (5.1), we have 2+maxΔ(H)c(H). There are also results for minΔ(H). Indeed, by [Citation27, Lemma 4.1], we have gcd(didi1|i[2,s])|minΔ(H) and if |F×/H×|=1, then gcd(didi1|i[2,s])=minΔ(H) .

We continue with examples of numerical semigroup rings and numerical power series rings. Let K be a field and HN0 be a numerical monoid. Then K[H]=K[Xh|hH]K[X]andK[[H]]=K[[Xh|hH]]K[[X]] denote the numerical semigroup ring and the numerical power series ring. Since H is finitely generated, K[H] is a one-dimensional noetherian domain with integral closure K[X]. The power series ring K[[H]] is a one-dimensional local noetherian domain with integral closure K[[X]], and its value monoid vX(K[[H]]) is equal to H.

Example 5.5.

Let K be a field and HN0 be a numerical monoid distinct from N0. Then H is not half-factorial, whence Equation(5.1) implies that c(H)3. If min(H{0})3, then X2XK[[H]]+K[[H]], whence RR¯ is not a quadratic extension and K[[H]] is not stable by Corollary 3.2.2.

  1. Let H=e,e+1,,2e1=Ne{0} with eN2 and R=K[[H]]. By [Citation21, Special case 3.1, page 216], we have c(H)=c(R)=3. Indeed, by [Citation53, Theorem 5.6], there is a transfer homomorphism θ:RH.

  2. Let K be finite, H=e,e+1,,2e1=Ne{0} with eN2, and R=K[H]. We set ρ=X+XeR̂R̂/XeR̂ and G=K[ρ]×/K×. Then, by [Citation21, Special Case 3.2, page 216], we have |G|=|K|e1 and c(R)c(B(G)), where B(G) is the monoid of zero-sum sequences over G. Since c(B(G))max{exp(G),1+r(G)} by [Citation21, Theorem 6.4.2], the catenary degree of R grows with |G|.

Lemma 5.6.

Let R be an order in a Dedekind domain such that R is a maximal proper subring of R¯. Then we have

  1. Every maximal ideal of R is stable.

  2. R is stable if and only if R¯/R is a simple R-module.

Proof.

1. Let mX(R). Then (m:m) is an intermediate ring of R and R¯, and hence (m:m){R,R¯}. If (m:m)=R¯, then m is clearly an invertible ideal of (m:m), since R¯ is a Dedekind domain. Now let (m:m)=R. Since m is divisorial, we have that R=(m:m)=((R:m1):m)=(R:mm1)=(mm1)1, and thus m is v-invertible. Consequently, m is invertible.

2. First let R be stable. Then RR¯ is a quadratic extension by Proposition 3.3.1. Let N be an R-submodule of R¯ with RN. Then N is an intermediate ring of R and R¯. Consequently, N{R,R¯}, and hence R¯/R is a simple R-module.

Conversely, let R¯/R be a simple R-module. Obviously, RR¯ is a quadratic extension. Since R¯/R and R/(R:R¯) are isomorphic as R-modules, we have that (R:R¯)X(R). Let mX(R). If m=(R:R¯), then Rm is a discrete valuation domain, and hence there is precisely one maximal ideal of R¯ lying over m. Now let m=(R:R¯) and set k=|{qX(R¯)|qR=m}|. Assume that k3. Then there are some distinct q1,q2,q3X(R¯) such that q1R=q2R=q3R. Therefore, mq1q2q3q1q2q1R¯, and thus lR(R¯/m)3. On the other hand lR(R¯/m)=lR(R¯/R)+lR(R/m)=2, a contradiction. Consequently, k2 and thus R is finitely stable by Proposition 3.3.1.. Since R is noetherian, we have that R is stable. □

In the next proposition, we study the catenary degree of finitely primary monoids stemming from one-dimensional local stable domains. We establish an upper bound for their catenary degree in case when RR¯ is a maximal proper subring.

  1. Let H be a finitely primary monoid of rank one. In general, the map θ:Hvp(H),avp(a), need not be a transfer homomorphism. (Example: If H=[ε1p,ε2p,p2]F××F({p}) with εiεj1 for all i,j[1,2]).

  2. Let us consider the following example: let H be a reduced finitely primary monoid of rank one, say HF××F({p}) .

Suppose that H is generated by the following k + 1 elements, where k is even: ε1p,,εkp,p2 , where ε1··εk is a minimal product-one sequence in the group F×.

Then (ε1p)··(εkp)=p2··p2(k/2 times) is a minimal relation of atoms of H, whence c(H)k.

Proposition 5.7.

Let R be a local stable order in a Dedekind domain. Define R0=R.

  1. Suppose that R0R1Rn=R¯ where Ri=(mi1:mi1) and Ri1 is local with maximal ideal mi1 for all i[1,n], and X(Rn)={P1,P2}. Write m0=R1m0 for some m0m0{0}. Then (R:R¯)=m0nR¯=P1nP2n, and R is a finitely primary monoid of rank at most two and exponent n.

  2. If RR¯ is a maximal proper subring, then c(R)5.

Proof.

1. All domains R0,,Rn1 are local with maximal ideals mi such that mi=m0Ri+1=m0Ri+1 by [Citation44, Proposition 4.2] and also the Jacobson radical of Rn, Jn=P1P2=P1P2=mn1 and for some k > 0, Jnk=(P1P2)k=mn1km0 by [Citation44, Corollary 4.4]. Therefore, mn1kRnm0R, i.e. (R:Rn)=mn1k=P1kP2k and since mn1=m0Rn,(R:Rn)=m0kRn. Also m0nRn=m0n1m0Rn=m0n1mn1m0n1Rn1m0R1=mR and (R:Rn)=m0nRn. Therefore, (R:R¯)=m0nR¯=P1nP2n and R is a finitely primary monoid of rank two and exponent n.

2. Let m denote the maximal ideal of R. Since |max(R¯)|2 by Proposition 3.3.1, we distinguish two cases.

First, suppose that R¯ is local with maximal ideal P. Then by [Citation44, Proposition 4.2(i)], P2m, which implies that P2R¯mR and hence (R:R¯)=Pk with k{1,2}. Thus R is finitely primary of rank one and exponent two, whence c(R)5 by Equation(5.3).

Second, suppose that max(R¯)={P1,P2}. Then 1. shows that (R:R¯)=P1P2. Thus R is finitely primary of rank two and exponent one, whence c(R)3 by Equation(5.4). □

For an atomic monoid H, we set k(H)=sup{min(L{2})|2LL(H),|L|>1} .

Then k(H)=0 if and only if L(uv)={2} for all u,vA(H), and k(H)3 otherwise. If H is not half-factorial, then (5.5) k(H)2+supΔ(H) .(5.5)

The question of whether equality holds was studied a lot. Among others, equality holds for large classes of Krull domains [Citation20, Corollary 4.5], for numerical monoids H with |A(H)|=2, but not for all finitely primary monoids.

Proposition 5.8.

Let R be a local domain with maximal ideal m such that R is not a field and nN0mn={0}, let xm be such that m2=xm and let U = xR.

  1. I(R) is a reduced atomic monoid, U is a cancellative atom of I(R) and for every II(R){R} there are nN0 and JA(I(R)) such that I=UnJ.

2. k(I(R))=2+supΔ(I(R)) and k(I*(R))=2+supΔ(I*(R)).

Proof.

1. It follows from Lemma 4.1 that I(R) is an atomic monoid. Since R is not a field, we have that U is a non-zero proper ideal of R. If I and J are non-zero ideals of R such that UI = UJ, then xI = xJ, and hence I = J. Therefore, U is cancellative. Assume that U is not an atom of I(R). Then there are some proper A,BI(R) such that U = AB. We infer that xRm2=xm. Consequently, x = xu for some um, and thus 1=um, a contradiction. This implies that U is an atom of I(R). Now let I be a non-zero proper ideal of R. Then Im, and since nN0mn={0}, there is some mN such that Imm and Imm+1. We infer that Ixm1m. Set n=m1. Then nN0 and there is some proper JI(R) such that I=xnJ=UnJ. Assume that J is not an atom of I(R). Then there are some non-zero proper ideals A and B of R with J = AB, and thus Jm2=xm. Therefore, I=xnJxmm=mm+1, a contradiction. It follows that JA(I(R)).

2. This follows from 1. and from [Citation9, Proposition 4.1]. □

Theorem 5.9.

Let R be a one-dimensional Mori domain such that for every mX(R),m is stable and (Rm:Rm̂)={0}.

  1. The following statements are equivalent.

    1. I(R) is transfer Krull.

    2. I*(R) is transfer Krull.

    3. I*(R) is half-factorial.

    4. I(R) is half-factorial.

    5. c(I(R))2.

    6. c(I*(R))2.

      If these conditions hold, then the map π:X(R̂)X(R), defined by PPR, is bijective.

  • 2. k(I(R))=2+supΔ(I(R)) and k(I*(R))=2+supΔ(I*(R)).

Proof.

1. Suppose that Condition (c) holds. By Proposition 4.3.2, I*(R) is half-factorial if and only if Rp is half-factorial for every pX(R). Thus the map π is bijective by Theorem 5.1.1.

(a) (b) I*(R)I(R) is a divisor-closed submonoid, and divisor-closed submonoids of transfer Krull monoids are transfer Krull.

(b) (c) Since R is a one-dimensional Mori domain, we have that Iv*(R)=I*(R), and thus the assertion follows from Theorem 5.1.4.

(c) (d) Since R is a one-dimensional Mori domain, we have that R is of finite character. Furthermore, if mX(R), then Rm is a one-dimensional local Mori domain with non-zero conductor, and hence Rm is finitely primary. By Corollary 4.5 it remains to show that for every mX(R),m2 is contained in a proper invertible ideal of R. Let mX(R). Since R is a Mori domain and m(m:m2)=(m:m), we infer that mm(mm:mm2)=(mm:mm), i.e. mm is a stable ideal of Rm. Clearly, mm is a divisorial ideal of Rm. It follows from Lemma 5.3 that mm2=xmm for some xmm. Observe that m2=mm2RxRmR. Moreover, xRmR is t-finitely generated and locally principal and xRmRm, and thus xRmR is a proper invertible ideal of R.

(d) (a) All half-factorial monoids are transfer Krull.

(d) (e) Let mX(R). By Proposition 4.3.2 we have that I(Rm) is half-factorial. Note that Rm is a Mori domain and a G-domain. Since m is stable and R is a Mori domain, we have that mm is a stable ideal of Rm. Clearly, mm is a divisorial ideal of Rm. Therefore, mm2=xmm for some xmm by Lemma 5.3. Since Rm is a one-dimensional local Mori domain, it follows that nN0mmn={0}. We infer by Proposition 5.8 and [Citation9, Proposition 4.1.4] that c(I(Rm))2. Therefore, c(I(R))2 by Theorem 5.1.

(e) (f) This is obvious, since I*(R) is a divisor-closed submonoid of I(R).

(f) (c) Since I*(R) is cancellative, this follows from (5.1).

2. If (Hi)iI is a family of atomic monoids, then supΔ(iIHi)=sup{supΔ(Hi)|iI}andk(iIHi)=sup{k(Hi)|iI} .

Thus the claim follows from Propositions 4.3 and 5.8.2. □

Let R be as in Theorem 5.9. Clearly, we have k(I*(R))k(I(R)), but in general, we do not have equality.

By Theorem 3.7, stable domains with non-zero conductor, that are Mori or weakly Krull, are already orders in Dedekind domains. Thus our next result is formulated for stable orders in Dedekind domains. Its first part generalizes a result valid for orders in quadratic number fields [Citation9, Theorem 1.1]. Note, if R is a semilocal domain, then Pic(R)=0. This means that every invertible ideal is principal, whence I*(R)={aR|aR}(R)red. If R is not semilocal, then the statements for I*(R) need not hold for R. If R is any order in an algebraic number field, then c(R)c(B(Pic(R))) [Citation21, Sections 3.4 and 3.7]. Moreover, R can be transfer Krull without being half-factorial [Citation24, Theorems 5.8 and 6.2].

Theorem 5.10.

Let R be a stable order in a Dedekind domain.

  1. The following statements are equivalent.

    1. I(R) is transfer Krull.

    2. I*(R) is transfer Krull.

    3. I*(R) is half-factorial.

    4. I(R) is half-factorial.

    5. c(I(R))2.

    6. c(I*(R))2.

  2. cmon(I*(R))<.

  3. I*(R) has finite elasticity if and only if π:X(R¯)X(R) is bijective. If this holds, then the elasticity is accepted.

Proof.

1. This is an immediate consequence of Theorem 5.9.

2. Since R is an order in a Dedekind domain, R is a weakly Krull Mori domain with non-zero conductor. By Proposition 5.2, the localizations Rp are strongly ring-like of rank at most two. Thus I*(R) has finite monotone catenary degree by [Citation25, Theorem 5.13].

3. By [Citation21, Proposition 1.4.5] and by Proposition 4.3, we have ρ(I*(R))=sup{ρ(Rp)|pX(R)} .

Thus the assertion follows from Proposition 5.2. □

In Remark 5.11 we briefly discuss further arithmetical properties, which follow from the ones given in Theorem 5.10. Then we work out, in a series of remarks, that none of the statements in Theorem 5.10 holds true in general without the stability assumption.

Remark 5.11

(Structure of sets of lengths and of their unions).

  1. (Structure of sets of lengths) If R is an order in a Dedekind domain, then sets of lengths of I*(R) are well-structured. They are almost arithmetical multiprogressions with global bounds for all parameters [Citation21, Section 4.7]. This holds without the stability assumption.

  2. (Structure of unions of sets of lengths) Let H be an atomic monoid. For every kN, Uk(H)=kLL(H)LN

is the union of sets of lengths containing k. The structure theorem for unions of sets of lengths states that there is a bound M such that almost all sets Uk(H)[minUk(H)+M,maxUk(H)M] are arithmetical progressions with difference minΔ(H). Now every atomic monoid with accepted elasticity satisfies this structure theorem for unions of sets of lengths, and the initial parts Uk(H)[minUk(H),minUk(H)+M] and the end parts [maxUk(H)M,maxUk(H)] fulfill a periodicity property (we refer to recent work of Tringali [Citation55, Theorem 1.2].

Remark 5.12

(On catenary degrees). Example 5.5.2 offers examples of non-stable orders in Dedekind domains whose catenary degree is arbitrarily large. Furthermore, there are finitely primary monoids with arbitrarily large catenary degree (see the discussion after Lemma 5.6). Non-stable local orders in Dedekind domains may have infinite monotone catenary degree [Citation35, Examples 6.3 and 6.5].

Remark 5.13

(Seminormal orders). We compare the arithmetic of stable orders with the arithmetic of seminormal orders in Dedekind domains. Note that stable orders need not be seminormal (all orders in quadratic number fields are stable but not all are seminormal [Citation10]) and seminormal orders need not be stable (see the example given in Remark 5.15).

Let R be a seminormal order in a Dedekind domain and let π:X(R¯)X(R) be defined by π(P)=PR for all PX(R¯). If π is bijective, then c(I*(R))=2. If π is not bijective, then c(I*(R))=3 and cmon(I*(R)){3,5} [Citation24, Theorem 5.8]. Furthermore, I*(R) is half-factorial if and only if π is bijective. For stable orders, only one implication is true (see Theorem 5.9).

Remark 5.14

(Half-factoriality of I*(R) does not imply half-factoriality of I(R)).

The Statements 1.(c) and 1.(d) of Theorem 5.10 need not be equivalent for divisorial orders in Dedekind domains. We construct a local divisorial order R in a Dedekind domain such that I*(R) is half-factorial, and yet I(R) is not half-factorial.

Let L be the field with 16 elements, let KL be the field with 2 elements, let yL be such that y4=1+y and let V=(1,y,y2)K. Let X be an indeterminate over L and let R=K+VX+X2L[[X]]. We assert that R is a local divisorial half-factorial Cohen-Kaplansky domain such that I(R) is not half-factorial.

Proof.

By [Citation4, Example 6.7] we have that R is a half-factorial local Cohen-Kaplansky domain, (1,y,y2,y3) is a K-basis of L and L={ab|a,bV}. Let m=VX+X2L[[X]] and let I=yX2,(1+y3)X2R. Then m is the maximal ideal of R. Note that m1=(m:m)={fL[[X]]|f(V+XL[[X]])V+XL[[X]]}={fL[[X]]|f0VV}=K+XL[[X]]=1,y3XR, and hence R is divisorial by [Citation38, Theorem 3.8]. By Proposition 4.2.2 it is sufficient to show that IA(I(R)) and Im2. Observe that m2=X2L[[X]],m3=X3L[[X]] and I={0,y,1+y3,1+y+y3}X2+X3L[[X]]. Therefore, m3Im2. For every ideal E of R let S(E)={aV|aX+zE for some zm2} and T(E)={aL|aX2+zE for some zm3}. Note that S(E) is a K-subspace of V and T(E) is a K-subspace of L.

Claim: If A and B are proper ideals of R, then T(AB)=(S(A)S(B))K.

Let A and B be proper ideals of R. First let aT(AB). Then aX2+zAB for some zm3. Therefore, aX2+z=i=1nfigi for some nN,fiA and giB for every i[1,n]. Since A,Bm, there are some ai,biV and zi,vim2 for every i[1,n] such that fi=aiX+zi and gi=biX+vi for every i[1,n]. Consequently, aiS(A) and biS(B) for all i[1,n]. Moreover, aX2+z=(i=1naibi)X2+i=1n(aiviX+biziX+zivi). Since i=1n(aiviX+biziX+zivi)m3 this implies that a=i=1naibi(S(A)S(B))K.

Now let a(S(A)S(B))K. Then a=i=1naibi with nN and aiS(A) and biS(B) for every i[1,n]. There are some zi,vim2 for every i[1,n] such that aiX+ziA and biX+viB for every i[1,n]. Therefore, aX2+i=1n(aiviX+biziX+zivi)=i=1n(aiX+zi)(biX+vi)AB. Since i=1n(aiviX+biziX+zivi)m3, we have that aT(AB). This proves the claim.

Assume that IA(I(R)). Then there are proper ideals A and B of R such that I = AB. It follows by the claim that {0,y,1+y3,1+y+y3}=T(I)=(S(A)S(B))K. Clearly, dimK(S(A)),dimK(S(B))>0. If dimK(S(A))=dimK(S(B))=1, then |(S(A)S(B))K|=2, a contradiction. Therefore, dimK(S(A))2 or dimK(S(B))2. Without restriction let dimK(S(A))2. There are some non-zero aS(B) and some two-dimensional K-subspace W of S(A). We infer that (S(A)S(B))KaW and 4=|(S(A)S(B))K||aW|=|W|=4, and thus {0,y,1+y3,1+y+y3}=aW. Clearly, a{1,y,1+y,y2,1+y2,y+y2,1+y+y2}. To obtain a contradiction it is sufficient to show that WV.

Case 1: a = 1. Then W={0,y,1+y3,1+y+y3}V.

Case 2: a = y. Then W=(1+y3){0,y,1+y3,1+y+y3}={0,1,1+y2+y3,y2+y3}V.

Case 3: a=1+y. Then W=(y+y2+y3){0,y,1+y3,1+y+y3}={0,1+y+y2+y3,1+y+y2,y3}V.

Case 4: a=y2. Then W=(1+y2+y3){0,y,1+y3,1+y+y3}={0,1+y3,1+y+y2+y3,y+y2}V.

Case 5: a=1+y2. Then W=(1+y+y3){0,y,1+y3,1+y+y3}={0,1+y2,y2+y3,1+y3}V.

Case 6: a=y+y2. Then W=(1+y+y2){0,y,1+y3,1+y+y3}={0,y+y2+y3,y+y3,y2}V.

Case 7: a=1+y+y2. Then W=(y+y2){0,y,1+y3,1+y+y3}={0,y2+y3,1+y,1+y+y2+y3}V.

Remark 5.15

(Transfer Krull does not imply stability). If R is an order in a Dedekind domain with (R:R¯)max(R) and R¯=RR¯×, then R is transfer Krull by [Citation21, Proposition 3.7.5]. We provide an example showing that such an order need not be stable.

We construct a seminormal one-dimensional local noetherian domain R such that R¯=RR¯×,(R:R¯)max(R¯),R¯ is local, and R has ideals which are not 2-generated. Thus, Corollary 3.8 implies that R is not stable.

Let KL be a field extension with 3[L:K]<, X be an indeterminate over L, and R=K+XL[[X]]. Observe that R̂=L[[X]] is a completely integrally closed one-dimensional noetherian domain. Let BL be a K-basis of L. Then R̂=BR, and hence R̂ is a finitely generated R-module. Since R̂ is noetherian, it follows from the Theorem of Eakin-Nagata that R is noetherian, and hence R¯=R̂.

Since R¯ is one-dimensional local and RR¯ is an integral extension, we have that R itself is local and one-dimensional. Moreover, spec(R)={{0},XL[[X]]} and R is transfer Krull by [Citation24, Theorem 5.8].

Now if xq(R) with x2,x3R, then x2,x3R̂, and hence xR̂ (since R̂ is completely integrally closed), so x0L and x02,x03K (whence x0 is the constant term of x), and thus if x0=0, then xR and if x00, then x0=x03x02K, hence xR. Therefore, R is seminormal.

Note that (R:R¯)=(R:R̂)=XL[[X]]{0} and (R:R¯)max(R¯). It is clear that RR¯×R¯. Let yR¯=L[[X]]. We have that yy0XL[[X]] (where y0 is the constant term of y). If yR, then clearly yRR¯×. Now let yR. Then y0K, and hence y0=0. Observe that (yy0)y01XL[[X]] and y0L×R¯×. Therefore, y=(1+(yy0)y01)y0RR¯×.

Assume to the contrary, that XL[[X]] is 2-generated. Therefore, there exist x,yR with XL[[X]]=x,yR. Let x1, y1 be the linear coefficients of x respectively y. Then L=x1,y1K and hence [L:K]2, a contradiction.

Acknowledgment

We would like to thank the anonymous referee for many valuable suggestions and comments which improved the quality of this paper.

Additional information

Funding

This work was supported by the Austrian Science Fund FWF, Project Numbers J4023-N35, W1230, and P33499.

References

  • Anderson, D. D., Anderson, D. F., Zafrullah, M. (1992). Atomic domains in which almost all atoms are prime. Commun. Algebra 20(5):1447–1462. DOI: 10.1080/00927879208824413.
  • Anderson, D. D., Huckaba, J. A., Papick, I. J. (1987). A note on stable domains. Houston J. Math. 13(1):13–17.
  • Anderson, D. D., Mott, J., Zafrullah, M. (1992). Finite character representations for integral domains. Boll. Unione Mat. Ital. 6:613–630.
  • 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., Smertnig, D. Lattices over Bass rings and graph agglomerations. Algebr. Represent. Theory DOI: 10.1007/s10468-021-10040-2.
  • Bass, H. (1963). On the ubiquity of Gorenstein rings. Math. Z. 82(1):8–28. DOI: 10.1007/BF01112819.
  • Bazzoni, S. (2000). Divisorial domains. Forum Math. 12:397–419.
  • Bazzoni, S., Salce, L. (1996). Warfield domains. J. Algebra 185(3):836–868. DOI: 10.1006/jabr.1996.0353.
  • Brantner, J., Geroldinger, A., Reinhart, A. (2020). On monoids of ideals of orders in quadratic number fields. In: Facchini, A., Fontana, M., Geroldinger, A., Olberding, B. Advances in rings, modules, and factorizations, Vol. 321. Switzerland: Springer, pp. 11–54.
  • Dobbs, D. E., Fontana, M. (1987). Seminormal rings generated by algebraic integers. Mathematika 34(2):141–154. DOI: 10.1112/S0025579300013395.
  • 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.
  • Gabelli, S. (2006). Generalized Dedekind domains. In: Brewer, J. W., Glaz, S., Heinzer, W., Olberding, B. Multiplicative Ideal Theory in Commutative Algebra. New York, NY: Springer, pp. 189–206.
  • Gabelli, S. (2014). Ten problems on stability of domains. In: Fontana, M., Frisch, S., Glaz, S. Commutative Algebra. New York, NY: Springer, pp. 175–193.
  • Gabelli, S., Roitman, M. (2019). On finitely stable domains, I. J. Commut. Algebra 11(1):49–67. DOI: 10.1216/JCA-2019-11-1-49.
  • Gabelli, S., Roitman, M. (2020). On finitely stable domains II. J. Commut. Algebra 12:179–198.
  • Gao, W., Liu, C., Tringali, S., Zhong, Q. (2021). On half-factoriality of transfer Krull monoids. Commun. Algebra 49(1):409–420. DOI: 10.1080/00927872.2020.1800720.
  • García-Sánchez, P. A., O’Neill, C., Webb, G. (2019). The computation of factorization invariants for affine semigroups. J. Algebra Appl. 18(1):1950019–1950021. DOI: 10.1142/S0219498819500191.
  • Geroldinger, A. (2016). Sets of lengths. Amer. Math. Monthly 123:960–988.
  • Geroldinger, A., Gotti, F., Tringali, S. (2021). On strongly primary monoids, with a focus on Puiseux monoids. J. Algebra 567:310–345. DOI: 10.1016/j.jalgebra.2020.09.019.
  • Geroldinger, A., Grynkiewicz, D. J., Schmid, W. A. (2011). The catenary degree of Krull monoids I. J. Théor. Nombres Bordeaux. 23(1):137–169. DOI: 10.5802/jtnb.754.
  • 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 Press.
  • 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., 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., 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(3):419–457. DOI: 10.1142/S0218196719500097.
  • Geroldinger, A., Roitman, M. (2020). On strongly primary monoids and domains. Commun. Algebra 48(9):4085–4099. DOI: 10.1080/00927872.2020.1755678.
  • Geroldinger, A., Zhong, Q. (2018). Long sets of lengths with maximal elasticity. Can. J. Math. 70(6):1284–1318. DOI: 10.4153/CJM-2017-043-4.
  • Geroldinger, A., Zhong, Q. (2020). Factorization theory in commutative monoids. Semigroup Forum 100:22–51.
  • Goeters, H. P. (1998). Noetherian stable domains J. Algebra 202(1):343–356. DOI: 10.1006/jabr.1997.7305.
  • Greither, C. (1982). On the two generator problem for the ideals of a one-dimensional ring. J. Pure Appl. Algebra 24(3):265–276. DOI: 10.1016/0022-4049(82)90044-5.
  • Halter-Koch, F. (1998). Ideal Systems. An Introduction to Multiplicative Ideal Theory. New York, NY: Marcel Dekker.
  • Halter-Koch, F. (2011). Multiplicative ideal theory in the context of commutative monoids. In: Fontana, M., Kabbaj, S.-E., Olberding, B., Swanson, I. Commutative Algebra: Noetherian and Non-Noetherian Perspectives. New York, NY: Springer, pp. 203–231.
  • Halter-Koch, F. (2013). Quadratic Irrationals, Pure and Applied Mathematics. Vol. 306. Boca Raton, FL: Chapman & Hall/CRC Press.
  • Halter-Koch, F. (2020). An Invitation to Algebraic Numbers and Algebraic Functions. Pure and Applied Mathematics. Boca Raton, FL: CRC Press.
  • Hassler, W. (2009). Properties of factorizations with successive lengths in one-dimensional local domains. J. Commut. Algebra 1(2):237–268. DOI: 10.1216/JCA-2009-1-2-237.
  • Lipman, J. (1971). Stable ideals and Arf rings. Amer. J. Math. 93(3):649–685. DOI: 10.2307/2373463.
  • Loper, K. A. (2006). Almost Dedekind Domains Which Are Not Dedekind. In: Brewer, J. W., Glaz, S., Heinzer, W., Olberding, B. Multiplicative Ideal Theory in Commutative Algebra. Springer, pp. 279–292.
  • Matlis, E. (1968). Reflexive domains. J. Algebra. 8(1):1–33. DOI: 10.1016/0021-8693(68)90031-8.
  • Matlis, E. (1970). The two-generator problem for ideals. Michigan Math. J. 17:257–265.
  • Olberding, B. (1998). Globalizing local properties of Prüfer domains. J. Algebra 205(2):480–504. DOI: 10.1006/jabr.1997.7406.
  • Olberding, B. (2001). On the classification of stable domains. J. Algebra 243:177–197.
  • Olberding, B. (2001a). Stability, duality, 2-generated ideals and a canonical decomposition of modules. Rend. Sem. Mat. Univ. Padova 106:261–290.
  • Olberding, B. (2001b). Stability of ideals and its applications. In: Anderson, D. D., Papick, I. J., eds. Ideal Theoretic Methods in Commutative Algebra. Lecture Notes in Pure and Applied Mathematics (D.D. Anderson and I.J. Papick, eds.), Vol. 220. New York, NY: Marcel Dekker, pp. 319–341.
  • Olberding, B. (2002). On the structure of stable domains. Commun. Algebra 30:877–895.
  • Olberding, B. (2014). Finitely Stable Rings. Commutative Algebra. New York: Springer, pp. 269–291.
  • Olberding, B. (2014). One-dimensional bad Noetherian domains. Trans. Amer. Math. Soc. 366(8):4067–4095.
  • Olberding, B. (2016). One-dimensional stable rings. J. Algebra 456:93–122.
  • Olberding, B., Reinhart, A. (2020). Radical factorization in finitary ideal systems. Commun. Algebra 48(1):228–253. DOI: 10.1080/00927872.2019.1640237.
  • Omidali, M. (2012). The catenary and tame degree of numerical monoids generated by generalized arithmetic sequences. Forum Math. 24:627–640.
  • O’Neill, C., Pelayo, R. (2017). Factorization invariants in numerical monoids. In: Harrington, H. A., Omar, M., Wright, M., eds. Algebraic and Geometric Methods in Discrete Mathematics, Contemporary Mathematics, Vol. 685. Providence, RI: American Mathematical Society, pp. 231–249.
  • O’Neill, C., Pelayo, R. (2018). Realisable sets of catenary degrees of numerical monoids. Bull. Australian Math. Soc. 97:240–245.
  • Philipp, A. (2015). A characterization of arithmetical invariants by the monoid of relations II: The monotone catenary degree and applications to semigroup rings. Semigroup Forum 90:220–250.
  • Querré, J. (1976). Intersections d’anneaux intègres. J. Algebra 43(1):55–60.
  • Sally, J. D., Vasconcelos, W. V. (1973). Stable rings and a problem of Bass. Bull. Amer. Math. Soc. 79(3):574–576. DOI: 10.1090/S0002-9904-1973-13209-4.
  • Tringali, S. (2019). Structural properties of subadditive families with applications to factorization theory. Isr. J. Math. 234(1):1–35. DOI: 10.1007/s11856-019-1922-2.