975
Views
3
CrossRef citations to date
0
Altmetric
Research Article

Commutative rings with one-absorbing factorization

, , & ORCID Icon
Pages 2689-2703 | Received 29 Jul 2020, Accepted 21 Jan 2021, Published online: 13 Apr 2021

Abstract

Let R be a commutative ring with nonzero identity. Yassine et al. defined the concept of 1-absorbing prime ideals as follows: a proper ideal I of R is said to be a 1-absorbing prime ideal if whenever xyzI for some nonunit elements x,y,zR, then either xyI or zI. We use the concept of 1-absorbing prime ideals to study those commutative rings in which every proper ideal is a product of 1-absorbing prime ideals (we call them OAF-rings). Any OAF-ring has dimension at most one and local OAF-domains (D, M) are atomic such that M2 is universal.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

Throughout this article, all rings are commutative with nonzero identity and all modules are unital. Let N denote the set of positive integers. For mN, let [1,m]={nN|1nm}. Let R be a ring. An ideal I of R is said to be proper if IR. The radical of I is denoted by I={xR|xnI for some nN}. We denote by Min(I) the set of minimal prime ideals over the ideal I. The concept of prime ideals plays an important role in ideal theory and there are many ways to generalize it.

In [Citation9], Badawi introduced and studied the concept of 2-absorbing ideals which is a generalization of prime ideals. An ideal I of R is a 2-absorbing ideal if whenever a,b,cR and abcI, then abI or acI or bcI. In this case, I=P is a prime ideal with P2I or I=P1P2 where P1, P2 are incomparable prime ideals with P1P2I, cf. [Citation9, Theorem 2.4]. In [Citation8], Anderson and Badawi introduced the concept of n-absorbing ideals as a generalization of prime ideals where n is a positive integer. An ideal I of R is called an n-absorbing ideal of R, if whenever a1,a2,,an+1R and i=1n+1aiI, then there are n of the ai’s whose product is in I. In this case, due to Choi and Walker [Citation13, Theorem 1], (I)nI.

In [Citation23], Mukhtar et al. studied the commutative rings whose ideals have a TA-factorization. A proper ideal is called a TA-ideal if it is a 2-absorbing ideal. By a TA-factorization of a proper ideal I, we mean an expression of I as a product i=1rJi of TA-ideals. Mukhtar et al. prove that any TAF-ring has dimension at most one and the local TAF-domains are atomic pseudo-valuations domains. Recently in [Citation1], Ahmed et al. studied commutative rings whose proper ideals have an n-absorbing factorization. Let I be a proper ideal of R. By an n-absorbing factorization of I we mean an expression of I as a product i=1rIi of proper n-absorbing ideals of R. Ahmed et al. called AF-dim(R) (absorbing factorization dimension) the minimum positive integer n such that every ideal of R has an n-absorbing factorization. If no such n exists, set AF-dim(R)=. An FAF-ring (finite absorbing factorization ring) is a ring such that AF-dim(R)<. Recall that a general ZPI-ring is a ring whose proper ideals can be written as a product of prime ideals. Therefore, AF-dim(R) measures, in some sense, how far R is from being a general ZPI-ring, cf. [Citation1, Proposition 3]. By dim(R), we denote the Krull dimension of R.

In [Citation25], Yassine et al. introduced the concept of a 1-absorbing prime ideal which is a generalization of a prime ideal. A proper ideal I of R is a 1-absorbing prime ideal (our abbreviation OA-ideal) if whenever we take nonunit elements a,b,cR with abcI, then abI or cI. In this case, I=P is a prime ideal, cf. [Citation25, Theorem 2.3]. And if R is a ring in which exists an OA-ideal that is not prime, then R is a local ring, that is a ring with one maximal ideal.

Let I be a proper ideal of R. By an OA-factorization of I, we mean an expression of I as a product i=1nJi of OA-ideals. The aim of this note is to study the commutative rings whose proper ideals (resp., proper principal ideals, resp., proper 2-generated ideals) have an OA-factorization.

We call R a 1-absorbing prime factorization ring (OAF-ring) if every proper ideal has an OA-factorization. An OAF-domain is a domain which is an OAF-ring. Our article consists of five sections (including the introduction).

In the next section, we characterize OA-ideals (Lemma 2.1) and we prove that if I is an OA-ideal, then I is a primary ideal. We also show that the OAF-ring property is stable under factor ring (resp., fraction ring) formation (Propositions 2.2 and 2.3). Furthermore, we investigate OAF-rings with respect to direct products (Corollary 2.5) and polynomial ring extensions (Corollary 2.6). We prove that the general ZPI-rings are exactly the arithmetical OAF-rings (Theorem 2.8).

The third section consists of a collection of preparational results which will be of major importance in the fourth section. For instance, we show that the Krull dimension of an OAF-ring is at most one (Theorem 3.5).

The fourth section contains the main results of our article. Among other results, we provide characterizations of OAF-rings (Theorem 4.2), rings whose proper principal ideals have an OA-factorization (Corollary 4.3) and rings whose proper (principal) ideals are OA-ideals (Proposition 4.5).

In the last section, we study the transfer of the various OA-factorization properties to the trivial ring extension.

2. Characterization of OA-ideals and simple facts

We start with a characterization of OA-ideals. Recall that a ring R is a Q-ring (cf. [Citation3]) if every proper ideal of R is a product of primary ideals.

Lemma 2.1.

Let R be a ring with Jacobson radical M and I be an ideal of R.

  1. If R is not local, then I is an OA-ideal if and only if I is a prime ideal.

  2. If R is local, then I is an OA-ideal if and only if I is a prime ideal or M2IM.

  3. Every OA-ideal is a primary TA-ideal. In particular, every OAF-ring is both a Q-ring and a TAF-ring.

Proof.

(1) This follows from [Citation25, Theorem 2.4].

(2) Let R be local. Then, M is the maximal ideal of R.

() Let I be an OA-ideal such that I is not a prime ideal. Since I is proper, we infer that IM. Since I is not prime, there are a,bMI such that abI. To prove that M2I, it suffices to show that xyI for all x,yM. Let x,yM. Then, xyabI. Since xy,a,bM,bI and I is an OA-ideal, it follows that xyaI. Again, since x,y,aM,aI and I is an OA-ideal, we have that xyI.

() Clearly, if I is a prime ideal, then I is an OA-ideal. Now let M2IM. Then, I is proper. Let a,b,cM be such that abcI. Then, abM2I. Therefore, I is an OA-ideal.

(3) Let I be an OA-ideal. It is an immediate consequence of (1) and (2) that I is a primary ideal. Now let a,b,cR be such that abcI. We have to show that abI or acI or bcI.

First let a or b or c be a unit of R. Without restriction let a be unit of R. Since abcI, we infer that bcI.

Now let a, b, and c be nonunits. Then, abI or cI. If cI, then acI. The in particular statement is clear. □

Proposition 2.2.

Let R be an OAF-ring and I be a proper ideal of R. Then, R/I is an OAF-ring.

Proof.

Let J be a proper ideal of R which contains I. Let J=i=1mJi be an OA-factorization. Then, J/I=i=1m(Ji/I). It suffices to show that Ji/I is an OA-ideal for each i[1,m]. Let i[1,m] and let a,b,cR be such that a¯,b¯,c¯ are three nonunit elements of R/I and a¯b¯c¯Ji/I. Clearly, a, b, c are nonunit elements of R and abcJi. Since Ji is an OA-ideal of R, we get that abJi or cJi which implies that a¯b¯Ji/I or c¯Ji/I. Therefore, R/I is an OAF-ring. □

Proposition 2.3.

Let S be a multiplicatively closed subset of R0. If R is an OAF-ring, then S1R is an OAF-ring. In particular, RM is an OAF-ring for every maximal ideal M of R.

Proof.

Let J be a proper ideal of S1R. Then, J=S1I for some proper ideal I of R with IS=. Let I=i=1mIi be an OA-factorization. Then, J=i=1m(S1Ii) where each S1Ii which is proper is an OA-ideal by [Citation25, Theorem 2.18]. Thus, S1R is an OAF-ring. The in particular statement is clear. □

Let R be a ring. Then, R is said to be a π-ring if every proper principal ideal of R is a product of prime ideals. We say that R is a unique factorization ring (in the sense of Fletcher, cf. [Citation4]) if every proper principal ideal of R is a product of principal prime ideals. A unique factorization domain is an integral domain which is a unique factorization ring.

Remark 2.4.

Let R be a non local ring.

  1. R is a general ZPI-ring if and only if R is an OAF-ring.

  2. R is a π-ring if and only if each proper principal ideal of R has an OA-factorization.

  3. R is a unique factorization ring if and only if each proper principal ideal of R is a product of principal OA-ideals.

Proof.

This is an immediate consequence of Lemma 2.1(1). □

In the light of the above remark, we give the next result.

Corollary 2.5.

Let R1 and R2 be two rings and R=R1×R2 be their direct product. The following statements are equivalent.

  1. R is an OAF-ring.

  2. R is a general ZPI-ring.

  3. R1 and R2 are general ZPI-rings.

Proof.

This follows from Remark 2.4(1) and [Citation21, Exercise 6(g), page 223]. □

Let R be a ring. Then, R is called a von Neumann regular ring if for each xR there is some yR with x=x2y. The ring R is von Neumann regular if and only if R is a zero-dimensional reduced ring (see [Citation19, Theorem 3.1, page 10]).

Corollary 2.6.

Let R be a ring. The following statements are equivalent.

  1. R[X] is an OAF-ring.

  2. R is a Noetherian von Neumann regular ring.

  3. R is a finite direct product of fields.

Proof.

Observe that the polynomial ring R[X] is never local, since X and 1X are nonunit elements of R[X], but their sum is a unit. Consequently, R[X] is an OAF-ring if and only if R[X] is a general ZPI-ring by Remark 2.4(1). The rest is now an easy consequence of [Citation2, Theorem 6 and Corollary 6.1], [Citation21, Exercise 10, page 225] and Hilbert’s basis theorem. □

Let R be a ring and I be an ideal of R. Then, I is called divided if I is comparable to every ideal of R (or equivalently, I is comparable to every principal ideal of R).

Lemma 2.7.

Let R be a local ring with maximal ideal M such that M2 is divided. The following statements are equivalent.

  1. Each two principal OA-ideals which contain M2 are comparable.

  2. For each OA-ideal I of R, we have that I is a prime ideal or I=M2.

Proof.

(1) (2): Let I be an OA-ideal of R such that I is not a prime ideal of R. Then, M2IM by Lemma 2.1(2). Assume that M2I. Let xIM2 and let yMI. Then, x,yM2, and thus, M2xR,yR (since M2 is divided). It follows that xR and yR are (principal) OA-ideals of R by Lemma 2.1(2). Since yxR and xR and yR are comparable, we infer that xRyR. Consequently, there is some zM such that x = yz, and hence xM2, a contradiction. Therefore, I=M2.

(2) (1): This is obvious. □

Let R be a ring. An ideal I of R is called 2-generated if I=xR+yR for some (not necessarily distinct) x,yR. Note that every principal ideal of R is 2-generated. We say that R is a chained ring if each two ideals of R are comparable under inclusion. Moreover, R is said to be an arithmetical ring if RM is a chained ring for each maximal ideal M of R.

Theorem 2.8.

Let R be a ring. The following statements are equivalent.

  1. R is a general ZPI-ring

  2. R is an arithmetical OAF-ring.

  3. R is an arithmetical ring and each proper principal ideal of R has an OA-factorization.

Proof.

First, we show that if R is an arithmetical π-ring, then R is a general ZPI-ring. Let R be an arithmetical π-ring and let M be a maximal ideal of R. It is straightforward to show that RM is a π-ring. Moreover, RM is a chained ring, and hence every 2-generated ideal of RM is principal. Therefore, every proper 2-generated ideal of RM is a product of prime ideals of RM. Consequently, RM is a general ZPI-ring by [Citation22, Theorem 3.2]. This implies that dim(RM)1 by [Citation21, page 205]. We infer that dim(R)1, and thus R is a general ZPI-ring by [Citation16, Theorems 39.2, 46.7, and 46.11].

(1) (2) (3): This is obvious.

(3) (1): It is sufficient to show that R is a π-ring. If R is not local, then R is a π-ring by Remark 2.4(2). Therefore, we can assume that R is local with maximal ideal M. Since R is local, we have that R is a chained ring. Therefore, M2 is divided and each two OA-ideals of R are comparable. We infer by Lemma 2.7 that each OA-ideal of R is a product of prime ideals. Now it clearly follows that R is a π-ring. □

3. Preparational results

From Lemma 2.1(3), we have that |Min(I)|=1 for every OA-ideal I of R. In view of this remark, we obtain the following result.

Proposition 3.1.

Let R be a ring and I be a proper ideal of R. If I has an OA-factorization, then Min(I) is finite.

Proof.

Let I=i=1nIi be an OA-factorization. It follows that Min(I)i=1nMin(Ii), and thus |Min(I)|n.

Let R be a ring and I be an ideal of R. Then, I is called a multiplication ideal of R if for each ideal J of R with JI, there is some ideal L of R such that J = IL.

Lemma 3.2.

Let R be a local ring such that each proper principal ideal of R has an OA-factorization. Then, each nonmaximal minimal prime ideal of R is principal.

Proof.

Let P be a nonmaximal minimal prime ideal of R. By [Citation2, Theorem 1], it is sufficient to show that P is a multiplication ideal.

Let xP and let xR=i=1nIi be an OA-factorization. There is some j[1,n] such that IjP. By Lemma 2.1(2), we have that P = Ij, and hence xR = PJ for some ideal J of R. We infer that xR=P(xR:P).

Now let I be an ideal of R such that IP. Then, I=yIyR=yIP(yR:P)=PyI(yR:P), and thus, P is a multiplication ideal. □

The next result is a generalization of [Citation16, Theorem 46.8] and its proof is based on the proof of the same result.

Proposition 3.3.

Let R be a local ring with maximal ideal M such that dim(R)1 and every proper principal ideal of R has an OA-factorization. Then, R is an integral domain and if dim(R)2, then R is a unique factorization domain.

Proof.

Let N be the nilradical of R. It follows from Proposition 3.1 and Lemma 3.2 that Min(0) is finite and each PMin(0) is principal.

Claim: Every proper principal ideal of R/N has an OA-factorization. Let I be a proper principal ideal of R/N. Then, I=(xR+N)/N for some xM. Let xR=i=1nIi be an OA-factorization. We infer that I=(xR)/N=(i=1nIi)/N=i=1n(Ii/N). It suffices to show that Ii/N is an OA-ideal of R/N for each i[1,n]. Let i[1,n]. If Ii is a prime ideal of R, then NIi, and hence, Ii/N is a prime ideal of R/N. Now let Ii be not a prime ideal of R. By Lemma 2.1(2), we have that M2IiM. Note that R/N is local with maximal ideal M/N. Since (M/N)2=M2/NIi/NM/N, it follows by Lemma 2.1(2) that Ii/N is an OA-ideal of R/N. This proves the claim.

Case 1: R is one-dimensional. We prove that R is an integral domain. If every OA-ideal of R is a prime ideal, then R is π-ring, and hence, R is an integral domain by [Citation16, Theorem 46.8]. Now let not every OA-ideal of R be a prime ideal. It follows from Lemma 2.1(2) that M is not idempotent. Set L=M2QMin(0)Q. Next we prove that M2xR for each xRL. Let xRL. Without restriction let x be a nonunit. Note that xR cannot be a product of more than one OA-ideal, and hence, xR is an OA-ideal. By Lemma 2.1(2), we have that M2xR.

Now we show that PM2 for each PMin(0). Let PMin(0). Assume that PM2. Let wRP. Then, P+wRL by the prime avoidance lemma, and thus there is some v(P+wR)L. It follows that M2vRP+wR. Since P is a nonmaximal prime ideal, we have that R/P has no simple R/P-submodules, and hence yRP(P+yR)=P. (Note that if yRP(P+yR)=P, then yRP(P+yR)/P is a simple R/P-submodule of R/P.) This implies that M2yRP(P+yR)=P, and thus P = M, a contradiction.

Let QMin(0). By the prime avoidance lemma, there is some zML. We infer that QM2zR. Consequently, Q = zQ. Since Q is principal, it follows that Q=0 (e.g. by Nakayama’s lemma), and hence R is an integral domain.

Case 2: dim(R)2 and R is reduced. We show that R is a unique factorization domain. There is some nonmaximal nonminimal prime ideal Q of R. By the prime avoidance lemma, there is some xQPMin(0)P. Since R is reduced, we have that LMin(0)L=0. If yR is nonzero with xy = 0, then yL and xyL for some LMin(0), and hence xL, a contradiction. We infer that x is a regular element of R. Let xR=i=1nIi be an OA-factorization. Then, IjQ for some j[1,n]. Since x is regular, Ij is invertible, and hence, Ij is a regular principal ideal (because invertible ideals of a local ring are regular principal ideals). Since IjQ and Q=M, we have that Ij is a prime ideal by Lemma 2.1(2). Consequently, PIj for some PMin(0). Since Ij is regular, we infer that PIj, and hence P = PIj (since Ij is principal). It follows (e.g. from Nakayama’s lemma) that P=0 (since P is principal). We obtain that R is an integral domain.

To show that R is a unique factorization domain, it suffices to show by [Citation4, Theorem 2.6] that every nonzero prime ideal of R contains a nonzero principal prime ideal. Since dim(R)2 and R is local, we only need to show that every nonzero nonmaximal prime ideal of R contains a nonzero principal prime ideal. Let L be a nonzero nonmaximal prime ideal of R and let zL be nonzero. Let zR=k=1mJk be an OA-factorization. Then, JL for some [1,m]. Since R is an integral domain, zR is invertible, and hence, J is invertible. Therefore, J is nonzero and principal (since R is local). Since L=M, it follows from Lemma 2.1(2) that J is a prime ideal.

Case 3: dim(R)2. We have to show that R is a unique factorization domain. Note that R/N is a reduced local ring with maximal ideal M/N and dim(R/N)2. Moreover, each proper principal ideal of R/N has an OA-factorization by the claim. It follows by Case 2 that R/N is a unique factorization domain, and thus N is the unique minimal prime ideal of R. Since R/N is a unique factorization domain and dim(R/N)2, R/N possesses a nonzero nonmaximal principal prime ideal. We infer that there is some nonminimal nonmaximal prime ideal Q of R such that Q/N is a principal ideal of R/N. Consequently, there is some qQ such that Q=qR+N. Let qR=i=1nIi be an OA-factorization. Then, IjQ for some j[1,n]. Since Q=M, we infer by Lemma 2.1(2) that Ij is a prime ideal of R. Therefore, Q=qR+NIjQ, and hence Ij = Q.

Assume that Q=qR. Then, qR = QJ for some proper ideal J of R. It follows that qqR=(qR+N)JqJ+N, and thus q(1a)N for some aJ. Since a is a nonunit of R, we obtain that qN. This implies that Q=qR+N=N, a contradiction. We infer that Q = qR. Since NQ and N is a prime ideal of R, we have that N = NQ. Consequently, N=0 (e.g. by Nakayama’s lemma, since N is principal), and thus RR/N is a unique factorization domain. □

Proposition 3.4.

Let R be a local ring with maximal ideal M such that each proper 2-generated ideal of R has an OA-factorization. Then, dim(R)2 and each nonmaximal prime ideal of R is principal.

Proof.

First we show that dim(RP)1 for each nonmaximal prime ideal P of R. Let P be a nonmaximal prime ideal and let I be a proper 2-generated ideal of RP. Observe that I = JP for some 2-generated ideal J of R with JP. Let J=i=1nJi be an OA-factorization. Then, I=JP=i=1n(Ji)P=i=1,JiPn(Ji)P. If i[1,n] is such that JiP, then Ji is a prime ideal of R by Lemma 2.1(2), and thus (Ji)P is a prime ideal of RP. We infer that I is a product of prime ideals of RP. It follows from [Citation22, Theorem 3.2], that RP is a general ZPI-ring. It is an easy consequence of [Citation21, page 205] that dim(RP)1.

This implies that dim(R)2. It remains to show that every nonmaximal prime ideal of R is principal. Without restriction let dim(R)1. It follows from Proposition 3.3 that R is either a one-dimensional domain or a two-dimensional unique factorization domain. In any case, we have that each nonmaximal prime ideal of R is principal. □

In the next result, we will prove a generalization of the fact that every OAF-ring has Krull dimension at most one.

Theorem 3.5.

Let R be a ring such that every proper 2-generated ideal of R has an OA-factorization. Then, dim(R)1.

Proof.

If every OA-ideal of R is a prime ideal, then R is a general ZPI-ring by [Citation22, Theorem 3.2], and hence, dim(R)1 by [Citation21, page 205]. Now let not every OA-ideal of R be a prime ideal. We infer by Lemma 2.1 that R is local and the maximal ideal of R is not idempotent. Let M be the maximal ideal of R. It suffices to show that if Q is a nonmaximal prime ideal of R, then Q=0. Let Q be a nonmaximal prime ideal of R.

Assume that QM2. Since dim(R)2 by Proposition 3.4, there is some prime ideal P of R such that QP and dim(R/P)=1. Next we show that M2P+yR for each yRP. Let yRP and set J=P+yR. Without restriction let JM. Note that J is 2-generated by Proposition 3.4. Since JM2, J cannot be a product of more than one OA-ideal, and thus, J is an OA-ideal of R. Since PJM, we have that J is not a prime ideal of R, and thus M2J by Lemma 2.1(2). Moreover, R/P is an integral domain that is not a field. Consequently, R/P does not have any simple R/P-submodules, which implies that P=xRP(P+xR). (Observe that if xRP(P+xR)=P, then xRP(P+xR)/P is a simple R/P-submodule of R/P.) Therefore, M2xRP(P+xR)=P, and hence P = M, a contradiction. We infer that QM2.

There is some zMM2 (since M is not idempotent). Since zR is a product of OA-ideals, we have that zR is an OA-ideal of R. As shown before, LM2 for each nonmaximal prime ideal L of R, and thus zR is not a nonmaximal prime ideal. Consequently, QM2zR by Lemma 2.1(2), and hence Q = zQ. Since Q is principal by Proposition 3.4, it follows (e.g. by Nakayama’s lemma) that Q=0.

Lemma 3.6.

Let D be a local domain with maximal ideal M. Then, each proper principal ideal of D has an OA-factorization if and only if D is atomic and each irreducible element generates an OA-ideal. If these equivalent conditions are satisfied, then nNPn=0 for each height-one prime ideal P of D.

Proof.

() Let each proper principal ideal of D have an OA-factorization. If D is a unique factorization domain, then D is atomic and each irreducible element generates a prime ideal. Now let D be not a unique factorization domain. Then, dim(D)=1 by Proposition 3.3.

Assume that M2 is principal. Then, M is invertible, and hence M is principal (since D is local). Note that D is a DVR (since dim(D)=1), and hence, D is a unique factorization domain, a contradiction.

We infer that M2 is not principal. We show that D is atomic. Let yD be a nonzero nonunit. Then, yD=i=1nIi for some principal OA-ideals Ii. There are nonzero nonunits xiD such that y=i=1nxi and Ij = xjD for each j[1,n]. Let i[1,n]. If Ii is a prime ideal, then xi is a prime element, and thus xi is irreducible. Now let Ii not be a prime ideal. It follows from Lemma 2.1(2) that M2Ii. Since M2 is not principal, we have that xiM2. Therefore, xi is irreducible.

Finally, let zD be irreducible. Then, zD=j=1mJj for some principal OA-ideals Jj. Since zD is maximal among the proper principal ideals of D, we obtain that zD = Jj for some j[1,n].

() Let D be atomic such that each irreducible element generates an OA-ideal. Let I be a proper principal ideal of D. Without restriction let I be nonzero. Then, I = xD for some nonzero nonunit xD. Observe that x=i=1nxi for some irreducible elements xiD. It follows that i=1nxiD is an OA-factorization of I.

Now let the equivalent conditions be satisfied and let P be a height-one prime ideal of D. First let P=M. Then, D is a unique factorization domain by Proposition 3.3, and hence, P is principal. Therefore, nNPn is a prime ideal of D by [Citation5, Theorem 2.2(1)]. Since nNPnP, we infer that nNPn=0.

Now let P = M. Assume that nNMn=0 and let xnNMn be nonzero. Then, xD is a product of m OA-ideals of D for some positive integer m. We infer by Lemma 2.1(2) that M2mxD, and hence, M2mxDM4mM2m. This implies that xD=M2m=M4m=x2D, and thus, x is a unit of D, a contradiction. Therefore, nNMn=0.

Lemma 3.7.

Let R be a local ring with maximal ideal M such that M2 is divided and such that either M is nilpotent or R is an integral domain with nNMn=0. Then, R is an OAF-ring and every proper principal ideal of R is a product of principal OA-ideals.

Proof.

If M is idempotent, then M=0, and hence, R is a field and both statements are clearly satisfied. Now let M be not idempotent. There is some xMM2. In what follows, we freely use the fact that if N is an ideal of R and zR such that NzR, then N=z(N:zR), and hence, N = zJ for some ideal J of R.

Next we prove that M2=xM and xR is an OA-ideal of R. Since xM2 and M2 is divided, we have that M2xRM. Therefore, xR is an OA-ideal by Lemma 2.1(2). Since M2xR, there is some proper ideal J of R with M2=xJ, and thus M2xM. Obviously, xMM2, and hence M2=xM.

Now we show that R is an OAF-ring. Let I be a proper ideal of R. First let I=0. If M is nilpotent, then I is obviously a product of OA-ideals. If R is an integral domain, then I is an OA-ideal. Now let I be nonzero. In any case, there is a largest positive integer n such that IMn. Observe that IMn=xn1Mxn1R. Consequently, I=xn1L=(xR)n1L for some proper ideal L of R. Assume that LM2. Note that LM2=xMxR. This implies that L = xA for some proper ideal A of R, and hence I=xnAxnM=Mn+1, a contradiction. We infer that M2L (since M2 is divided). It follows from Lemma 2.1(2) that L is an OA-ideal. In any case, I is a product of OA-ideals.

Finally, we prove that every proper principal ideal of R is a product of principal OA-ideals. Let yM. First let y = 0. If M is nilpotent, then xk=0 for some kN, and thus yR=(xR)k is a product of principal OA-ideals. If R is an integral domain, then yR is a principal OA-ideal. Now let y be nonzero. There is some greatest N such that yM. Therefore, y=x1z for some zM. If zM2, then z = xv for some vM, and hence y=xvM+1, a contradiction. We infer that zM2, and thus M2zRM. It follows from Lemma 2.1(2) that zR is an OA-ideal of R. Consequently, yR=(xR)1(zR) is a product of principal OA-ideals. □

4. Characterization of OAF-rings and related concepts

First, we recall several definitions and discuss the factorization theoretical properties of local one-dimensional OAF-domains. Let D be an integral domain with quotient field K. Then, D̂={xK| there is some nonzero cD such that cxnD for all nN} is called the complete integral closure of D. Let (D:D̂)={xD|xD̂D} be the conductor of D in D̂. The domain D is called completely integrally closed if D=D̂ and D is said to be seminormal if for all xK such that x2,x3D, it follows that xD. Note that every completely integrally closed domain is seminormal. We say that D is a finitely primary domain of rank one if D is a local one-dimensional domain such that D̂ is a DVR and (D:D̂)=0. For each subset, XK let X1={xK|xXD} and Xv=(X1)1. An ideal I of D is called divisorial if Iv = I. Moreover, D is called a Mori domain if D satisfies the ascending chain condition on divisorial ideals. It is well known that every unique factorization domain and every Noetherian domain is a Mori domain (see [Citation14, Corollary 2.3.13] and [Citation11, page 57]). We say that D is half-factorial if D is atomic and each two factorizations of each nonzero element of D into irreducible elements are of the same length. Finally, D is called a C-domain if the monoid of nonzero elements of D (i.e. D0) is a C-monoid. For the precise definition of C-monoids, we refer to [Citation14, Definition 2.9.5].

Let D be a local domain with quotient field K and maximal ideal M. Set (M:M)={xK|xMM}. Then, (M:M) is called the ring of multipliers of M. Moreover, M2 is said to be universal if M2uD for each irreducible element uD.

Theorem 4.1.

Let D be a local domain with maximal ideal M such that D is not a field. The following statements are equivalent.

  1. D is an OAF-domain.

  2. D is a TAF-domain.

  3. D is one-dimensional and every proper principal ideal has an OA-factorization.

  4. D is one-dimensional and atomic and every irreducible element generates an OA-ideal.

  5. D is atomic such that M2 is universal.

  6. (M:M) is a DVR with maximal ideal M.

  7. D is a seminormal finitely primary domain of rank one.

If these equivalent conditions are satisfied, then D is a half-factorial C-domain and a Mori domain.

Proof.

(1) (2): This follows from Lemma 2.1(3).

(1) (3): By Theorem 3.5, D is one-dimensional. The rest of assertion (3) is clear.

(2) (5) (6): This follows from [Citation23, Theorem 4.3].

(3) (4): This is an immediate consequence of Lemma 3.6.

(4) (5): Let yD be an irreducible element. Since yD is an OA-ideal and yD=M, we deduce from Lemma 2.1(2) that M2yD. Hence, M2 is universal.

(5) + (6) (1): It follows from [Citation6, Theorem 5.1] that M2 is comparable to every principal ideal of D, and thus M2 is divided. Since (M:M) is a DVR with maximal ideal M, we have that nNMn=0. Consequently, D is an OAF-domain by Lemma 3.7.

(5) + (6) (7): First we show that D is finitely primary of rank one. Let P be a nonzero prime ideal of D. Then, P contains an irreducible element yD, and hence, M2yDP. Therefore, P = M, and thus, D is one-dimensional. It remains to show that D̂ is a DVR and (D:D̂)=0. Since (M:M) is a DVR, we have that (M:M) is completely integrally closed. Observe that D(M:M)D̂, and hence D̂(M:M)̂=(M:M). Therefore, D̂=(M:M) is a DVR. Since MD̂=M(M:M)MD and M=0, we infer that (D:D̂)=0.

Next we show that D is seminormal. Let V be the group of units of D̂. Let K be the field of quotients of D and let xK be such that x2,x3D. Then, x2,x3D̂. Since D̂ is a DVR, D̂ is seminormal, and thus xD̂. In particular, xM or xV. If xM, then xD. Now let xV. Note that VD is the group of units of D (by [Citation24, Corollary 1.4] and [Citation12, Proposition 2.1]), and thus x2 and x3 are units of D. Therefore, x=x2x3 is a unit of D, and hence xD.

(7) (6): By [Citation15, Lemma 3.3.3], we have that M is the maximal ideal of D̂. If xD̂, then xMM (since M is an ideal of D̂). It is straightforward to show that (M:M)D̂. We infer that (M:M)=D̂ is a DVR.

Now let the equivalent statements of Theorem 4.1 be satisfied. It remains to show that D is a half-factorial C-domain and a Mori domain. It follows from [Citation6, Theorem 6.2] that D is a half-factorial domain. Obviously, V is a subgroup of finite index of V and VMD̂M=(M:M)MM. It follows from [Citation18, Corollary 2.8] and [Citation14, Corollary 2.9.8] that D is a C-domain. Moreover, D is a Mori domain by [Citation18, Proposition 2.5.1]. □

We want to point out that a local one-dimensional OAF-domain need not be Noetherian. Let KL be a field extension such that [L:K]= and let D=K+XL[[X]]. Then, D is a local one-dimensional domain with maximal ideal M=XL[[X]] and (M:M)=L[[X]] is a DVR with maximal ideal M. Consequently, D is an OAF-domain by Theorem 4.1. Since [L:K]=, it follows that D is not Noetherian.

An integral domain D is called a Cohen-Kaplansky domain if D is atomic and D has only finitely many irreducible elements up to associates. It follows from [Citation6, Example 6.7] that there exists a local half-factorial Cohen-Kaplansky domain with maximal ideal M for which M2 is not universal. We infer by Theorem 4.1 that the aforementioned domain is not an OAF-domain.

Theorem 4.2.

Let R be a ring with Jacobson radical M. The following statements are equivalent.

  1. R is an OAF-ring.

  2. Each proper 2-generated ideal of R has an OA-factorization.

  3. dim(R)1 and each proper principal ideal has an OA-factorization.

  4. R satisfies one of the following conditions.

    1. R is a general ZPI-ring.

    2. R is a local domain, M2 is divided and nNMn=0.

    3. R is local, M2 is divided and M is nilpotent.

Proof.

(1) (2): This is obvious.

(2) (3): This is an immediate consequence of Theorem 3.5.

(3) (4): First let each OA-ideal of R be a prime ideal. Then, R is a π-ring. By [Citation16, Theorems 39.2, 46.7, and 46.11], R is a general ZPI-ring. Now let there be an OA-ideal of R which is not a prime ideal. It follows from Lemma 2.1 that R is local with maximal ideal M and M is not idempotent. Note that if xMM2, then xR cannot be a product of more than one OA-ideal, and hence xR is an OA-ideal.

Case 1: R is zero-dimensional. Let xMM2. Then, xR is an OA-ideal. We infer by Lemma 2.1(2) that M2xR. Consequently, M2 is divided. It follows from Lemma 2.1 that M2I for each OA-ideal I of R. Since 0 is a product of OA-ideals, we have that 0 contains a power of M. This implies that M is nilpotent.

Case 2: R is one-dimensional. It follows from Proposition 3.3 that R is an integral domain, and hence nNMn=0 by Lemma 3.6. It remains to show that M2 is divided. Let xRM2. Without restriction let x be a nonunit. Then, xR is an OA-ideal. By Lemma 2.1(2), we have that M2xR.

(4) (1): Clearly, every general ZPI-ring is an OAF-ring. The rest follows from Lemma 3.7. □

Corollary 4.3.

Let R be a ring with Jacobson radical M. The following statements are equivalent.

  1. Each proper principal ideal of R has an OA-factorization.

  2. R is a π-ring or an OAF-ring.

  3. R satisfies one of the following conditions.

    1. R is a π-ring.

    2. R is a local domain, M2 is divided andnNMn=0.

    3. R is local, M2 is divided and M is nilpotent.

Proof.

(1) (2): If R is not local, then R is a π-ring by Remark 2.4(2). Now let R be local. If dim(R)2, then R is a unique factorization domain by Proposition 3.3, and hence, R is a π-ring. If dim(R)1, then R is an OAF-ring by Theorem 4.2.

(2) (1): This is obvious.

(2) (3): This is an immediate consequence of Theorem 4.2 and the fact that every general ZPI-ring is a π-ring. □

Corollary 4.4.

Let R be a ring with Jacobson radical M. The following statements are equivalent.

  1. Each proper principal ideal of R is a product of principal OA-ideals.

  2. R satisfies one of the following conditions.

    1. R is a unique factorization ring.

    2. R is a local domain, M2 is divided and nNMn=0.

    3. R is local, M2 is divided and M is nilpotent.

Proof.

(1) (2): If R is not local, then R is a unique factorization ring by Remark 2.4(3). If R is local, then the statement follows from Corollary 4.3 and the fact that every local π-ring is a unique factorization ring ([Citation4, Corollary 2.2]).

(2) (1): Obviously, if R is a unique factorization ring, then each proper principal ideal of R is a product of principal OA-ideals. The rest is an immediate consequence of Lemma 3.7. □

In Lemma 2.1, we saw that if R is a local ring with maximal ideal M and I is an ideal of R such that M2I, then I is an OA-ideal of R. Now we will give a characterization of the rings for which every proper (principal) ideal is an OA-ideal.

Proposition 4.5.

Let R be a ring with Jacobson radical M. The following statements are equivalent.

  1. Every proper ideal of R is an OA-ideal.

  2. Every proper principal ideal of R is an OA-ideal.

  3. R is local and M2=0.

Proof.

(1) (2): This is obvious.

(2) (3): Assume that R is not local. Then, every proper principal ideal of R is a prime ideal by Lemma 2.1(1). Consequently, R is an integral domain. If xR is a nonunit, then x2R is a prime ideal, and hence x2R=xR and x = 0. Therefore, R is a field, a contradiction. This implies that R is local with maximal ideal M. We infer by Lemma 2.1(2) that 0 is a prime ideal or M2=0.

Assume that M2=0. Then, R is an integral domain and there is some nonzero xM2. It follow from Lemma 2.1(2) that x2R is a prime ideal or M2x2R. If x2R is a prime ideal, then x2R=xR. If M2x2R, then M2x2RxRM2, and thus x2R=xR. In any case, we have that x2R=xR, and hence, x is a unit (since x is regular), a contradiction.

(3) (1): This is an immediate consequence of Lemma 2.1(2). □

5. OA-factorization properties and trivial ring extensions

Let A be a ring and E be an A-module. Then, AE, the trivial (ring) extension of A by E, is the ring whose additive structure is that of the external direct sum AE and whose multiplication is defined by (a,e)(b,f)=(ab,af+be) for all a,bA and all e,fE. (This construction is also known by other terminology and other notation, such as the idealization A(+)E.) The basic properties of trivial ring extensions are summarized in the textbooks [Citation17, Citation19]. Trivial ring extensions have been studied or generalized extensively, often because of their usefulness in constructing new classes of examples of rings satisfying various properties (cf. [Citation7, Citation10, Citation20]). We say that E is divisible if E = aE for each regular element aA.

We start with the following lemma.

Lemma 5.1.

Let A be a ring, I be an ideal of A and E be an A-module. Let R=AE be the trivial ring extension of A by E.

  1. IE is an OA-ideal of R if and only if I is an OA-ideal of A.

  2. Assume that A contains a nonunit regular element and E is a divisible A-module. Then, the OA-ideals of R have the form LE where L is an OA-ideal of A.

Proof.

(1) This follows immediately from [Citation25, Theorem 2.20].

(2) Let J be an OA-ideal of R. Our aim is to show that 0EJ. Let eE and let aA be a nonunit regular element. Then, e = af for some fE and thus (a,0)(0,f)(0,e)=(0,0)J. Since J is an OA-ideal, we conclude that (a,0)(0,f)=(0,e)J or (0,e)J which implies that 0EJ. Therefore, J=LE with L={bA|(b,g)J for some gE} and L is an ideal of A by [Citation7, Theorems 3.1 and 3.3(1)]. Now the result follows from (1). □

Corollary 5.2.

Let A be an integral domain that is not a field, E be a divisible A-module and R=AE. Then, the OA-ideals of R have the form IE where I is an OA-ideal of A.

Next, we study the transfer of the OAF-ring property to the trivial ring extension.

Theorem 5.3.

Let A be a ring with Jacobson radical M, E be an A-module and R=AE.

  1. R is an OAF-ring if and only if one of the following conditions is satisfied.

    1. A is a general ZPI-ring, E is cyclic and the annihilator of E is a (possibly empty) product of idempotent maximal ideals of A.

    2. A is local, M2 is divided, E=0 and either M is nilpotent or A is a domain with nNMn=0.

    3. A is local, M2=0, ME = aE for each nonzero aM and ME = Mx for each xEME.

    In particular, if R is an OAF-ring, then A is an OAF-ring.

  2. Every proper ideal of R is an OA-ideal if and only if A is local, M2=0 and ME=0.

Proof.

(1) () First let R be an OAF-ring. By Theorem 4.2, it follows that (a) R is a general ZPI-ring or (b) R is local with maximal ideal N, N2 is divided and (N is nilpotent or R is a domain such that nNNn=0). If R is a general ZPI-ring, then condition (A) is satisfied by [Citation7, Theorem 4.10].

From now on let R be local with maximal ideal N such that N2 is divided. Observe that A is local with maximal ideal M and N=ME by [Citation7, Theorem 3.2(1)]. If R is a domain such that nNNn=0, then E=0 (for if zE is nonzero, then (0,z) is a nonzero zero-divisor of R), and hence AR is a domain, M2 is divided and nNMn=0.

Now let N be nilpotent. If E=0, then AR, and thus M2 is divided and M is nilpotent. From now on let E be nonzero. There is some kN such that Nk=0. Note that N2=M2ME and Nk=MkMk1E, and thus Mk=0. Since N2 is divided, we have that 0EN2 or N20E. If 0EN2, then E = ME, and hence E=MkE=0, a contradiction. Therefore, N20E, which implies that M2=0.

Let aM be nonzero. Then, (a,0)N2, and hence N2(a,0)R=aAaE. Consequently, MEaE, and thus ME = aE. Finally, let xEME. Then, (0,x)N2. We infer that N2(0,x)R=0Ax. This implies that MEAx. If MEMx, then bxME for some unit bA, and hence xME, a contradiction. It follows that MEMx, which clearly implies that ME = Mx.

() Next we prove the converse. If condition (A) is satisfied, then R is a general ZPI-ring by [Citation7, Theorem 4.10], and thus R is an OAF-ring. If condition (B) is satisfied, then A is an OAF-ring by Theorem 4.2, and hence RA is an OAF-ring. Now let condition (C) be satisfied. Set N=ME. Then, R is local with maximal ideal N by [Citation7, Theorem 3.2(1)]. By Theorem 4.2, it suffices to show that N is nilpotent and N2 is divided. Since M2=0, we obtain that N3=M3M2E=0, and thus N is nilpotent. It remains to show that N2(a,x)R for each (a,x)RN2. Let aA and xE be such that (a,x)N2. Since N2=0ME, we have to show that 0ME(a,x)R. If a is a unit of A, then (a, x) is a unit of R by [Citation7, Theorem 3.7] and the statement is clearly true. Let z0ME. Then, z=(0,y) for some yME.

Case 1: a is a nonzero nonunit. Since ME = aE, there is some vE such that y = av. Observe that z=(0,av)=(a,x)(0,v)(a,x)R.

Case 2: a = 0. Then, xEME (since (a,x)N2). Since ME = Mx, there is some bM such that y = bx. It follows that z=(0,bx)=(a,x)(b,0)(a,x)R.

The in particular statement now follows from Theorem 4.2.

(2) First let every proper ideal of R be an OA-ideal. By Proposition 4.5, we have that R is local with maximal ideal N and N2=0. It follows that A is local with maximal ideal M and N=ME by [Citation7, Theorem 3.2(1)]. Moreover, 0=N2=M2ME, and hence M2=0 and ME=0.

Conversely, let A be local, M2=0 and ME=0. Set N=ME. Then, R is local with maximal ideal N by [Citation7, Theorem 3.2(1)] and N2=M2ME=0. We infer by Proposition 4.5 that each proper ideal of R is an OA-ideal. □

Corollary 5.4.

Let A be an integral domain, E be a nonzero A-module and R=AE. The following statements are equivalent.

  1. R is an OAF-ring.

  2. A is a field.

  3. Every proper ideal of R is an OA-ideal.

Proof.

(1) (2): It follows from Theorem 5.3(1) that A is a general ZPI-ring and the annihilator of E is a product of idempotent maximal ideals of A or that A is local with maximal ideal M such that M2=0.

First, let A be a general ZPI-ring such that the annihilator of E is a product of idempotent maximal ideals of A. Note that A is a Dedekind domain, and thus, the only proper idempotent ideal of A is the zero ideal. Since E is nonzero, the annihilator of E is a proper ideal of A, and hence A must possess an idempotent maximal ideal. We infer that the zero ideal is a maximal ideal of A, and thus A is a field.

Now let A be local with maximal ideal M such that M2=0. Since A is an integral domain, it follows that M=0, and hence A is a field.

(2) (3): Set M=0. Then, A is local with maximal ideal M, M2=0, and ME=0. Now the statement follows from Theorem 5.3(2).

(3) (1): This is obvious. □

Remark 5.5.

In general, if A is an OAF-ring and E is an A-module, then AE need not be an OAF-ring. Indeed, let A be an OAF-domain that is not a field and let E be a nonzero A-module. By Corollary 5.4, AE is not an OAF-ring.

Corollary 5.6.

Let A be a local ring with maximal ideal M and E be a nonzero A-module such that ME=0. Set R=AE. The following statements are equivalent.

  1. R is an OAF-ring.

  2. M2=0.

  3. Every proper ideal of R is an OA-ideal.

Proof.

(1) (2): Assume that M2=0. By Theorem 5.3(1), A is a local general ZPI-ring and M is idempotent (since the annihilator of E is a nonempty product of idempotent maximal ideals of A and M is the only maximal ideal of A). We infer by [Citation21, Corollary 9.11] that A is a Dedekind domain or each proper ideal of A is a power of M (because local rings are indecomposable). If A is a Dedekind domain, then clearly M2=M=0 (since M is idempotent and a Dedekind domain has no nonzero proper idempotent ideals). Moreover, if every proper ideal of A is a power of M, then again M2=M=0 (since M is idempotent). In any case, we obtain that M2=0, a contradiction.

(1) (2) (3): This follows from Theorem 5.3. □

Example 5.7.

Let A be a local principal ideal ring with maximal ideal M such that A is not a field and M2=0 (e.g. A=Z/4Z). Set R=AA. Then, R is an OAF-ring, and yet not every proper ideal of R is an OA-ideal.

Proof.

Since M=0, it follows from Theorem 5.3(2) that not every proper ideal of R is an OA-ideal. By Theorem 5.3(1), it remains to show that M = aA for each nonzero aM and M = Mx for each xAM. Note that M = zA for some zM. If aM is nonzero, then a = zb for some bA. Clearly, bM, and thus b is a unit of A, which clearly implies that M = zA = aA. Finally, if xAM, then x is a unit of A, and thus M = Mx. □

Remark 5.8.

Let A be a ring with Jacobson radical M, E be an A-module and R=AE. Then, each proper principal ideal of R has an OA-factorization if and only if one of the following conditions is satisfied.

  1. A is a π-ring, E is cyclic and the annihilator of E is a ( possibly empty) product of idempotent maximal ideals of A.

  2. A is local, M2 is divided, E=0 and either M is nilpotent or A is a domain with nNMn=0.

  3. A is local, M2=0, ME = aE for each nonzero aM and ME = Mx for each xEME.

Proof.

This can be proved along similar lines as Theorem 5.3(1) by using Corollary 4.3 and [Citation7, Theorems 3.2(1) and 4.10]. □

Acknowledgments

We want to thank the referee for many helpful suggestions and comments which improved the quality of this paper. The first three authors dedicate this work to their Professor Faycal Lamrini for his retirement.

Additional information

Funding

The fourth-named author was supported by the Austrian Science Fund FWF, Project Number J4023-N35.

References

  • Ahmed, M. T., Dumitrescu, T., Khadam, A. (2020). Commutative rings with absorbing factorization. Commun. Algebra 48(12):5067–5075. DOI: 10.1080/00927872.2020.1778714.
  • Anderson, D. D. (1976). Multiplication ideals, multiplication rings, and the ring R(X). Can. J. Math. 28(4):760–768. DOI: 10.4153/CJM-1976-072-1.
  • Anderson, D. D., Mahaney, L. A. (1987). Commutative rings in which every ideal is a product of primary ideals. J. Algebra 106(2):528–535. DOI: 10.1016/0021-8693(87)90014-7.
  • Anderson, D. D., Markanda, R. (1985). Unique factorization rings with zero divisors. Houston J. Math. 11(1):15–30.
  • Anderson, D. D., Matijevic, J., Nichols, W. (1976). The Krull intersection theorem. Pacific J. Math. 66(1):15–22. DOI: 10.2140/pjm.1976.66.15.
  • 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.
  • Anderson, D. D., Winders, M. (2009). Idealization of a module. J. Commut. Algebra 1(1):3–56. DOI: 10.1216/JCA-2009-1-1-3.
  • Anderson, D. F., Badawi, A. (2011). On n-absorbing ideals of commutative rings. Commun. Algebra 39(5):1646–1672. DOI: 10.1080/00927871003738998.
  • Badawi, A. (2007). On 2-absorbing ideals of commutative rings. Bull. Austral. Math. Soc. 75(3):417–429. DOI: 10.1017/S0004972700039344.
  • Bakkari, C., Kabbaj, S., Mahdou, N. (2010). Trivial extensions defined by Prüfer conditions. J. Pure Appl. Algebra 214(1):53–60. DOI: 10.1016/j.jpaa.2009.04.011.
  • Barucci, V. (2000). Mori domains. Non-Noetherian commutative ring theory. In: Mathematical Application. Vol. 520. Dortrecht: Kluwer Academics Publishers, pp. 57–73. DOI: 10.1007/978-1-4757-3180-4_3.
  • Barucci, V., Dobbs, D. E. (1984). On chain conditions in integral domains. Can. Math. Bull. 27(3):351–359. DOI: 10.4153/CMB-1984-053-1.
  • Choi, S. H., Walker, A. (2020). The radical of an n-absorbing ideal. J. Commut. Algebra 12(2):171–177. DOI: 10.1216/jca.2020.12.171.
  • Geroldinger, A., Halter-Koch, F. (2006). Non-Unique Factorizations. Algebraic, Combinatorial and Analytic Theory (Pure and Applied Mathematics). Boca Raton, FL: Chapman & Hall/CRC.
  • Geroldinger, A., Kainrath, F., Reinhart, A. (2015). Arithmetic of seminormal weakly Krull monoids and domains. J. Algebra 444:201–245. DOI: 10.1016/j.jalgebra.2015.07.026.
  • Gilmer, R. (1992). Multiplicative Ideal Theory (90 Queen's Papers). Kingston: Queen's University.
  • Glaz, S. (1989). Commutative Coherent Rings (Lecture Notes in Mathematics). Berlin: Springer.
  • Halter-Koch, F., Hassler, W., Kainrath, F. (2004). Remarks on the multiplicative structure of certain one-dimensional integral domains. In: Rings, Modules, Algebras, and Abelian Groups. Lecture Notes in Pure and Applied Mathematics. Vol. 236. New York: Marcel Dekker, pp. 321–331.
  • Huckaba, J. A. (1988). Commutative Rings with Zero Divisors. New York: Marcel Dekker.
  • Kabbaj, S., Mahdou, N. (2004). Trivial extensions defined by coherent-like conditions. Commun. Algebra 32(10):3937–3953. DOI: 10.1081/AGB-200027791.
  • Larsen, M., McCarthy, P. (1971). Multiplicative Theory of Ideals (Pure and Applied Mathematics). New York: Academic Press.
  • Levitz, K. B. (1972). A characterization of general Z.P.I.-rings. II. Pacific J. Math. 42(1):147–151. DOI: 10.2140/pjm.1972.42.147.
  • Mukhtar, M., Ahmed, M. T., Dumitrescu, T. (2018). Commutative rings with two-absorbing factorization. Commun. Algebra 46(3):970–978. DOI: 10.1080/00927872.2017.1332202.
  • Ohm, J. (1966). Some counterexamples related to integral closure in D[[x]]. Trans. Am. Math. Soc. 122(2):321–333. DOI: 10.2307/1994550.
  • Yassine, A., Nikmehr, M. J., Nikandish, R. (2020). On 1-absorbing prime ideals of commutative rings. J. Algebra Appl. DOI: 10.1142/S0219498821501759.