Home Mathematics On modules related to McCoy modules
Article Open Access

On modules related to McCoy modules

  • Jongwook Baeck ORCID logo EMAIL logo
Published/Copyright: December 31, 2022

Abstract

In this paper, we first investigate the relationships between the McCoy module and related modules based on their relationships in rings. After that, we improve some properties of McCoy modules and introduce ZPZC modules which extend the notion of McCoy modules. We observe the structure of ZPZC modules providing a number of examples of problems that arise naturally in the process. Finally, answers to some open questions related to the ZPZC condition are provided.

MSC 2010: 13C13; 16D70; 16D80; 16U99

1 Introduction

All rings in this paper are associative with identity, and modules are unitary right modules unless we specify otherwise. Let R be a ring and M be the right R -module. We denote R [ x ] for the polynomial ring over R , and M [ x ] for the polynomial module over M .

For a commutative ring R , McCoy [1, Theorem 2] proved that if a polynomial f ( x ) is a zero-divisor in R [ x ] , then f ( x ) c = 0 for some nonzero c in R . This result does not hold for noncommutative rings according to [2]. Following [3], the rings which meet the above condition are called the right McCoy rings. The left McCoy rings are defined analogously, and McCoy rings are both left and right McCoy. There are several ways in which McCoy theorem can be generalized by weakening the commutativity condition. For instance, all reversible rings (i.e., a b = 0 implies b a = 0 for any a , b R ) are McCoy [4, Theorem 2]. As another class of examples, according to [5, Theorem 8.2], right duo rings are right McCoy, where a ring is right duo if every right ideal is also a left ideal. The left duo rings are defined symmetrically, and rings with both conditions are called the duo rings. There are other natural conditions that imply the McCoy property, not directly related to commutativity. For instance, reduced rings are reversible; therefore, they are McCoy rings. Furthermore, reduced rings are Armendariz rings. Here, a ring R is called Armendariz if whenever polynomials f ( x ) = i = 0 m a i x i and g ( x ) = j = 0 n b j x j in R [ x ] satisfy f ( x ) g ( x ) = 0 , we have a i b j = 0 for every i and j [3]. Thus, McCoy rings are shown to be a unifying generalization of reversible, Armendariz, and duo rings. Recently, the author introduced the class of ZPZC rings which generalizes the class of McCoy rings (ZPZC is short for “zero-divisor polynomials have zero-divisor coefficients”). An example of a ZPZC ring that is not a McCoy ring was provided in [6, Example 2.7] and a condition for the right ZPZC rings to be right McCoy is given in [6, Proposition 2.15].

However, it is natural to extend the notion of rings to one of the modules. For instance, Cui and Chen [7] introduced the class of McCoy modules, extending the notion of McCoy rings: m ( x ) g ( x ) = 0 , where m ( x ) = i = 0 p m i x i M [ x ] and g ( x ) = j = 0 q b j x j R [ x ] { 0 } , implies that there exists a nonzero element r R such that m ( x ) r = 0 . They established some properties of the class of McCoy modules and presented some equivalent conditions for McCoy modules. Not only the concepts of reduced, reversible, Armendariz, and right duo rings but also the concepts of rings related to them are also naturally extended to modules. For instance, the right R -module M is called:

  1. reduced if for any m M and any a R , m a = 0 implies m R M a = 0 [8],

  2. symmetric if m a b = 0 implies m b a = 0 , for any m M and a , b R [9],

  3. Armendariz if whenever polynomials m ( x ) = i = 0 p m i x i M [ x ] and g ( x ) = j = 0 q b j x j R [ x ] satisfy m ( x ) g ( x ) = 0 , we have m i b j = 0 for every i and j [10],

  4. duo if for any R -submodule N of M and any R -endomorphism f of M , f ( N ) N [11],

  5. semicommutative if for any m M and any a R , m a = 0 implies m R a = 0 [10].

We note that symmetric modules are the notion of module corresponding to reversible rings, and as stated in [11], the right R -module R , which is a duo, is a right duo ring. All Armendariz modules are obviously McCoy modules, and according to [8, Lemma 1.5], reduced modules are McCoy. It follows from [7, Proposition 2.4] that semicommutative modules over reduced rings are McCoy modules. Recall that a ring R is semicommutative if a b = 0 entails a R b = 0 for any a , b R . Generally, semicommutative rings are not McCoy, as stated by [4, Section 3]. For more details on the McCoy condition, the readers can refer to [4,5,7, 12,13,14, 15,16,17, 18,19,20].

The purposes of this paper are as follows:

  1. to complete the implication diagram among modules related to McCoy modules,

  2. to improve some known results of McCoy modules,

  3. to introduce ZPZC modules which extend McCoy modules,

  4. to answer the question [6, Question 2.8(2)] that is “is there a ring which is left McCoy and right ZPZC, but not right McCoy?”, suggested by the referee of [6].

Thus, the implication diagram and various related examples are provided in Section 2. Among other things, an unpredictable example for which duo modules may not be McCoy is constructed (Example 2.5). In Section 3, one improves some known properties of McCoy modules. Precisely, we prove that every right module over a right π -duo reduced ring is McCoy (Theorem 3.3), which is an extended result of [7, Proposition 2.4]. It is also proven that if a ring R is right uniform, then the direct sum of ZPZC modules as a right R -module is McCoy (Theorem 3.9), which makes [7, Proposition 2.9] a corollary. In Section 4, we observe some properties of ZPZC modules with several examples. In particular, we show that for every cyclic module M over a right duo ring R , T n ( M ) is a ZPZC right T n ( R ) -module (Theorem 4.12) from which one can find many examples of ZPZC modules. In Section 5, one can see some counterexamples for answers to questions of [6]. Especially, we construct a left McCoy and right ZPZC ring but not right McCoy (Example 5.1) for giving a negative answer of [6, Question 2.8(2)].

For m M , ann R ( m ) is used for the set of all r R such that m r = 0 . For fixed integer n 1 , M n ( A ) and T n ( A ) stand for the set of all n × n full matrices over a set A and the set of all n × n upper triangular matrices over a set A , respectively, and E i , j means the matrix unit which is the matrix with ( i , j ) -entry 1 and elsewhere 0.

2 McCoy and duo modules

We here study the relationships between McCoy and related modules, as stated in Section 1. We in advance show the diagram that contains all implications among related modules (with no other implications holding, except by transitivity). In [5, p. 615], one can see implications immediately in the case of M = R .

In the above diagram, we cannot find duo modules even though the right duo rings are both semicommutative and right McCoy. To explain the reason, we construct several examples in this section.

An R -module M is called abelian if for any m M and a R , any idempotent e R , we have m a e = m e a [22]. Recall that a ring R is abelian if every idempotent in R is central. Following [22, Lemma 2.7], semicommutative modules are abelian modules. An R -module M is called Gaussian [12] if for any m ( x ) = i = 0 p m i x i M [ x ] , g ( x ) = j = 0 q b j x j R [ x ] and m ( x ) g ( x ) = k = 0 p + q c k x k , we have i = 0 p j = 0 q m i R b j R = k = 0 p + q c k R . It was shown that Gaussian modules are Armendariz [12, p. 2594].

Proposition 2.1

If an R-module M is Gaussian, then M is a semicommutative module.

Proof

Let m M and a R such that m a = 0 . Since M is Gaussian, m R a m R a R = m a R = { 0 } . Therefore, M is semicommutative.□

Remark 2.2

  1. Note that R = Z 4 is a local ring with J ( R ) 2 = { 0 } , where J ( R ) is the Jacobson radical of R . Thus, the commutativity of R guarantees that R is Gaussian according to [24, Proposition 1.8]. But R is not reduced. Hence, Gaussian modules may not be reduced.

  2. As stated in [25, p. 2268], the polynomial ring over Z , Z [ x ] is not Gaussian since Z [ x ] is not Prüfer by [26, Theorem 7.7]. However, Z [ x ] is a reduced module over itself. Thus, reduced modules do not imply Gaussian in general.

  3. Referring to [25, Example 11], symmetric modules need not be Gaussian modules.

  4. There exists a duo module which is not Gaussian by (2).

According to [24, Lemma 1.4], every right Gaussian ring is a right duo ring. However, this implication does not hold in modules anymore.

Example 2.3

Let R = Z 2 and M = m 1 R m 2 R be the free R -module with basis { m 1 , m 2 } . One can easily check that M is Gaussian. Consider the R -module endomorphism h of M induced by h ( m 1 ) = m 2 and h ( m 2 ) = m 1 . Then, h ( m 1 R ) = m 2 R m 1 R . Thus, M is not duo.

Right duo rings are semicommutative rings and thus abelian rings. However, duo modules need not be abelian modules.

Example 2.4

Let E = Z 2 a , b be the free algebra with identity and noncommuting indeterminates a , b over Z 2 . Set R = E / I where I is the ideal of E generated by the relation b a = 1 . We denote r = r + I for simplicity. Now consider the cyclic free R -module L = m R and the R -submodule N = m a R of L . Then, the factor module M = L / N is isomorphic to R / a R as an R -module. Note that each element r R can be written in the following form:

r = α + g 0 ( a ) a + h ( b ) b + g 1 ( a ) a b + g 2 ( a ) a b 2 + ,

where α Z 2 and g 0 ( x ) , , h ( x ) Z 2 [ x ] . If h ( x ) = j = 0 t β j x j with β t 0 , then

r a t + 1 = α a t + 1 + g 0 ( a ) a t + 2 + j = 0 t 1 β j b j b a t + 1 + β t + .

Therefore, every R -submodule of R / a R contains 1. Thus, M R / a R is a simple R -module, which forces that M is a duo module. Finally, we claim that M is not abelian. Let e = a b in R . Then, e 2 = e in R and m ¯ e b = ( m ¯ a ) b 2 = 0 ¯ in M . But m ¯ b e = m ¯ ( b a ) b = m ¯ b 0 ¯ in M . Hence, M is not abelian, which implies that M is not semicommutative by [22, Lemma 2.7].

Camillo and Nielsen proved that all right duo rings are right McCoy in [5, Theorem 8.2]. However the next example shows that there is a duo module which is not McCoy.

Example 2.5

Let E = Z 2 a 0 , a 1 , a 2 , b 0 , b 1 , c be the free algebra with identity and noncommuting indeterminates as labeled above over Z 2 . Set R = E / I where I is the ideal of E generated by the following relations: for each 0 i 2 , 0 j , k 1 ,

a 0 b 0 = 0 , a 0 b 1 + a 1 b 0 = 0 , a 1 b 1 + a 2 b 0 = 0 , a 2 b 1 = 0 , a 0 a i = a 2 a i = b j a i = b j b k = c a i = c b j = 0 .

We denote r = r + I for simplicity. Applying Bergman’s diamond lemma [27], we can write each r R in the following unique reduced form:

r = α + f 0 ( a 1 ) a 0 + f 1 ( a 1 ) a 1 + f 2 ( a 1 ) a 2 + f 3 ( a 1 ) b 0 + f 4 ( a 1 ) b 1 + g ( c ) c + h 0 ( 1 ) ( a 1 ) a 0 c + h 1 ( 1 ) ( a 1 ) a 1 c + h 2 ( 1 ) ( a 1 ) a 2 c + h 3 ( 1 ) ( a 1 ) b 0 c + h 4 ( 1 ) ( a 1 ) b 1 c + h 0 ( 2 ) ( a 1 ) a 0 c 2 + h 1 ( 2 ) ( a 1 ) a 1 c 2 + h 2 ( 2 ) ( a 1 ) a 2 c 2 + h 3 ( 2 ) ( a 1 ) b 0 c 2 + h 4 ( 2 ) ( a 1 ) b 1 c 2 + ,

where α Z 2 and f 0 ( x ) , , f 4 ( x ) , g ( x ) , h 0 ( p ) ( x ) , , h 4 ( p ) ( x ) Z 2 [ x ] for each p N . Note that if r c c R for some r R , then r must be zero or identity.

Let L = y R be the cyclic free R -module. Set M = L / N , where N = { 0 , y c , y c 2 , } = y c R is the R -submodule of L generated by the relation y c = 0 . Then, we can write each m ¯ M uniquely in the following form:

m ¯ = y ¯ α + y ¯ f 0 ( a 1 ) a 0 + y ¯ f 1 ( a 1 ) a 1 + y ¯ f 2 ( a 1 ) a 2 + y ¯ f 3 ( a 1 ) b 0 + y ¯ f 4 ( a 1 ) b 1 + y ¯ h 0 ( 1 ) ( a 1 ) a 0 c + y ¯ h 1 ( 1 ) ( a 1 ) a 1 c + y ¯ h 2 ( 1 ) ( a 1 ) a 2 c + y ¯ h 3 ( 1 ) ( a 1 ) b 0 c + y ¯ h 4 ( 1 ) ( a 1 ) b 1 c + y ¯ h 0 ( 2 ) ( a 1 ) a 0 c 2 + y ¯ h 1 ( 2 ) ( a 1 ) a 1 c 2 + y ¯ h 2 ( 2 ) ( a 1 ) a 2 c 2 + y ¯ h 3 ( 2 ) ( a 1 ) b 0 c 2 + y ¯ h 4 ( 2 ) ( a 1 ) b 1 c 2 + .

First, we claim that an R -module M is not McCoy. Consider the following nonzero polynomials m ( x ) = y ¯ a 0 + y ¯ a 1 x + y ¯ a 2 x 2 M [ x ] and r ( x ) = b 0 + b 1 x R [ x ] . Then, m ( x ) r ( x ) = 0 by the first row in the relations. But there is no nonzero element r R such that y ¯ a 1 r = 0 by the relations. Thus, M is not McCoy.

Finally, we claim that M is a duo R -module. Referring to [11, Lemma 1.1], we will find all R -endomorphism of M . Since M is still a cycle R -module, for any R -endomorphism, we need to examine the image of y ¯ . Consider the correspondence h : M M defined by h ( y ¯ r ) = y ¯ a 0 r for any r R . Note that h ( y ¯ 0 ) = y ¯ a 0 0 = 0 ¯ y ¯ a 0 c = h ( y ¯ c ) despite 0 + N = y c + N . Therefore, h is not well-defined. From the fact that if r c c R for some r R , then r must be zero or identity in R , we can notice that if h is an R -endomorphism of M induced by h ( y ¯ ) = y ¯ r for some r R , then r must be zero or identity in R by the well-definedness. This means that if h is an R -endomorphism of M , then h must be either the zero or the identity endomorphism. Thus, M must be duo by [11, Lemma 1.1].

Remark 2.6

Based on Example 2.5, one may suspect that every duo module is cyclic. However, the example in [28, Section 3] eliminates the possibility of the suspicion.

Even though reversible (and so symmetric) rings are McCoy rings [4, Theorem 2], symmetric modules need not be McCoy modules. We construct a symmetric R -module that is not McCoy.

Example 2.7

Let E = Z 2 a 0 , a 1 be the free algebra with identity and indeterminates a 0 , a 1 over Z 2 . Set R = E / I where I is the ideal of E generated by the following relations:

a i a j = 0 ( 0 i , j 1 ) .

We simply identify r = r + I in R . Applying the diamond lemma [27], we can write each element r R in the following reduced form:

r = α + α 0 a 0 + α 1 a 1 ,

where α , α 0 , α 1 Z 2 . Now, let L = m 0 R m 1 R m 2 R be the free R -module with basis { m 0 , m 1 , m 2 } . Set M = L / N where N is the submodule of L generated by the following relations:

m 0 a 0 = 0 , m 0 a 1 + m 1 a 0 = 0 , m 1 a 1 + m 2 a 0 = 0 , m 2 a 1 = 0 .

We also denote m = m + N in M for simplicity. Then, we can write each element m M uniquely in the following form:

m = m 0 β 0 + m 1 β 1 + m 2 β 2 + m 1 γ 0 a 0 + m 1 γ 1 a 1 ,

where β 0 , β 1 , β 2 , γ 0 , γ 1 Z 2 . Since R is commutative, M is symmetric. Next, we claim that M is not McCoy. Consider the nonzero polynomials m ( x ) = m 0 + m 1 x + m 2 x 2 M [ x ] and r ( x ) = a 0 + a 1 x R [ x ] . Then, m ( x ) r ( x ) = 0 by the relations. It suffices to show that ann R ( m 1 ) = { 0 } . Suppose that m 1 r = 0 for some r R and write r = α + α 0 a 0 + α 1 a 1 . Then, m 1 r = m 1 ( α + α 0 a 0 + α 1 a 1 ) = 0 implies that α = α 0 = α 1 = 0 , indicating that r = 0 as required.

Recall that an R -module M is Dedekind finite (or directly finite) if M M N for some R -module N implies that N = { 0 } . From [29, Exercise 1.8], M is Dedekind finite if and only if the endomorphism ring E n d ( M ) is a Dedekind finite ring (i.e., for any f , g E n d ( M ) , f g = 1 implies g f = 1 ). Both right McCoy and abelian rings are Dedekind finite rings. However, the implications do not hold for modules.

Example 2.8

Let R = Z 2 and let M = m 1 R m 2 R be the infinite free module over R . Note that the R -module M is reduced and Gaussian. However, by [30, Exercise 4.3.3], M is injective. This implies that M is not a Dedekind finite module by [31, Proposition 5.7].

3 McCoy modules and ZPZC modules

In [6], the class of ZPZC rings that contains all McCoy rings is introduced. In this section, we extend the notion of (right) ZPZC rings to the one of the modules. Then, we provide new properties of McCoy modules by improving known results, and offer some conditions for which ZPZC modules are McCoy.

Definition 3.1

Let M be a right module over a ring R and M [ x ] be the corresponding polynomial module over R [ x ] . M is referred to as ZPZC if ann R [ x ] ( m ( x ) ) { 0 } implies ann R ( m i ) { 0 } for each m i , where m ( x ) = i = 0 p m i x i M [ x ] .

All McCoy modules are obviously ZPZC modules. But the converse does not hold by [6, Example 2.7]. Here, we construct a nontrivial ZPZC module which is not McCoy.

Example 3.2

Let E = Z 2 a 0 , a 1 , b 0 , b 1 , b 2 be the free algebra over Z 2 with identity and commuting indeterminates a 0 , a 1 , b 0 , b 1 , b 2 . Set R = E / I where I is the ideal of E generated by the following relations: for each i , j { 0 , 1 } and k , { 0 , 1 , 2 } ,

a i a j = a i b k = b k b = 0 .

We identify r = r + I for simple expression. Then, we can write each element r R uniquely in the following reduced form:

r = α + α 0 a 0 + α 1 a 1 + β 0 b 0 + β 1 b 1 + β 2 b 2 ,

where α , α 0 , α 1 , β 0 , β 1 , β 2 Z 2 . Now let L = m 0 R m 1 R n 0 R n 1 R n 2 R be the free R -module with basis { m 0 , m 1 , n 0 , n 1 , n 2 } . Set M = L / N where N is the R -submodule of L generated by the following relations:

m 0 a 0 = m 0 a 1 = m 0 b 0 = m 0 b 1 = m 0 b 2 = 0 , m 1 a 0 = m 1 a 1 = m 1 b 1 = 0 , m 1 b 2 + n 1 b 2 = 0 , n 0 a 0 = 0 , n 0 a 1 + n 1 a 0 = 0 , n 0 b 0 = n 0 b 1 = n 0 b 2 = 0 , n 1 a 1 + n 2 a 0 = 0 , n 1 b 0 = 0 , n 2 a 1 = n 2 b 0 = n 2 b 1 = n 2 b 2 = 0 .

For every m M , we also identify m = m + N for simple expression. Then, we can write each element m M uniquely in the following form:

m = m 0 γ 0 + m 1 γ 1 + m 1 γ 2 b 0 + n 0 δ 0 + n 1 δ 1 + n 2 δ 2 + n 1 δ 3 a 0 + n 1 δ 4 a 1 + n 1 δ 5 b 1 + n 1 δ 6 b 2 ,

where γ 0 , γ 1 , γ 2 , δ 0 , , δ 6 Z 2 . For any m M , we can easily check that m must be annihilated by one of the elements in { b 0 , b 1 , b 2 } . Thus, M is ZPZC. Finally, consider the nonzero polynomials m ( x ) = m 0 + m 1 x + n 0 x 2 + n 1 x 3 + n 2 x 3 M [ x ] and r ( x ) = a 0 + a 1 x R [ x ] . Then, m ( x ) r ( x ) = 0 . But since ann R ( m 1 ) ann R ( n 1 ) = { 0 } , we can conclude that M is not McCoy.

From Example 2.7, we obtain that a module M over a commutative ring may not be ZPZC. Thus, we will determine some condition that makes a module over a commutative ring implies ZPZC. A ring R is called right π -duo if for any a R , there is a positive integer n such that R a n a R (see [32]). Of course, the notion of right π -duo rings is an extended notion of semicommutative modules. Thus we extend [7, Proposition 2.4] by the same method.

Theorem 3.3

(cf. [7, Proposition 2.4]). If a ring R is reduced and right π -duo, then every right R-module M is McCoy.

Proof

Let m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) = j = 0 q r j x j R [ x ] \ { 0 } with m ( x ) r ( x ) = 0 and r 0 0 . Since R is right π -duo, r 1 r 0 n = r 0 r 1 for some n N and r 1 R . From m 0 r 1 + m 1 r 0 = 0 , we obtain that

m 0 r 1 r 0 n + m 1 r 0 n + 1 = m 0 r 0 r 1 + m 1 r 0 n + 1 = m 1 r 0 n + 1 = 0 .

Similarly, from r 2 r 0 n 2 + n = r 0 n + 1 r 2 and r 1 r 0 n 2 + n = r 0 n + 1 r 1 for some r 2 , r 1 R , we obtain

m 0 r 2 r 0 n 2 + n + m 1 r 1 r 0 n 2 + n + m 2 r 0 n 2 + n + 1 = m 0 r 0 n + 1 r 2 + m 1 r 0 n + 1 r 1 + m 2 r 0 n 2 + n + 1 = m 2 r 0 n 2 + n + 1 = 0 .

Continuing in this manner, we arrive at the conclusion that m ( x ) r 0 k = 0 for some k N .□

Remark 3.4

According to [32, Proposition 1.9(4)], right π -duo rings are abelian. From the fact that every reduced ring is abelian, the converse of Theorem 3.3 does not hold true in general by [5, Theorem 7.1]. Moreover, the hypothesis of Theorem 3.3 is not superfluous by the following examples.

  1. In Example 2.7, the R -module M is not McCoy where the ring R is right π -duo but not reduced.

  2. Let R = Z 2 a 0 , a 1 be the free algebra with identity and noncommuting indeterminates a 0 , a 1 over Z 2 . Clearly, R is a domain and so reduced. Note that R is not right π -duo. Let L = m 0 R m 1 R m 2 R be the free R -module with basis { m 0 , m 1 , m 2 } . Set M = L / N where N is the submodule of L generated by the following relations:

    m 0 a 0 = 0 , m 0 a 1 + m 1 a 0 = 0 , m 1 a 1 + m 2 a 0 = 0 , m 2 a 1 = 0 .

After suppressing the bar notation in M , we consider the nonzero polynomials m ( x ) = m 0 + m 1 x + m 2 x 2 M [ x ] and r ( x ) = a 0 + a 1 x R [ x ] . Then, m ( x ) r ( x ) = 0 but ann R ( m 1 ) = { 0 } . Hence, M is not McCoy.

A ring R is semiprime if the zero ideal is semiprime, and right Gaussian if R R is Gaussian. Recall that a ring R is called weakly right duo if for each a R , there exists a positive integer n such that a n R is a two-sided ideal of R (see [32] or [33]). According to [24, Lemma 1.4], every right Gaussian ring is right duo. It is clear that every right duo ring is weakly right duo, and every weakly right duo ring is right π -duo but not conversely in each case. Note that every semiprime right duo ring is reduced. Thus by Theorem 3.3 and [32, Proposition 2.2(1)], we obtain:

Corollary 3.5

Let J ( R ) be the Jacobson radical of a ring R and M be a right module over R. Then, M is McCoy if one of the following conditions holds:

  1. R is right π -duo with J ( R ) = { 0 } .

  2. R is weakly right duo with J ( R ) = { 0 } .

  3. R is semiprime and right duo.

  4. R is semiprime and right Gaussian.

  5. R is commutative with J ( R ) = { 0 } .

  6. R is commutative semiprime.

  7. R is a commutative domain.

Let R be a commutative ring and S a multiplicative subset of R . We say that an ideal A of R is S-principal if A s a R A for some s S and a A (see [34, Definition 1]). We say that R is an S-principal ideal ring if each ideal of R is S -principal. Clearly, every commutative principal ideal ring is an S -principal ideal ring for any S of R . According to [35, Corollary 2.2], every commutative domain D is always an S -principal ideal ring, where S = D { 0 } . Based on the fact that every module over a commutative domain or a commutative principal ideal ring is McCoy (see [7, Proposition 2.15]), one may suspect that every module over a commutative S -principal ideal ring is McCoy or ZPZC. But the following example eliminates the possibility of suspicion.

Example 3.6

Let E = Z 2 a 0 , a 1 , s be the free algebra with identity and commuting indeterminates a 0 , a 1 , s over Z 2 . Set R = E / I where I is the ideal of E generated by the relations:

a 0 2 = a 0 a 1 = a 1 2 = a 0 s = a 1 s = 0 , s 2 = s .

We denote r = r + I for simplicity. Then, we can write each r R in the following reduced form:

r = α + α 0 a 0 + α 1 a 1 + β s ,

where α , α 0 , α 1 , β Z 2 . One can easily show that if an ideal A of R is nonprincipal, then either A s { 0 } or A s s R A . Therefore, R is a commutative S -principal ideal ring, where S = { 1 , s } . Now let L = m 0 R m 1 R m 2 R be the free R -module with basis { m 0 , m 1 , m 2 } . Set M = L / N where N is the R -submodule of L generated by the following relations:

m 0 a 0 = 0 , m 0 a 1 + m 1 a 0 = 0 , m 1 a 1 + m 2 a 0 = 0 , m 2 a 1 = 0 .

Consider the following nonzero polynomials m ( x ) = m 0 ¯ + m 1 ¯ x + m 2 ¯ x 2 M [ x ] and r ( x ) = a 0 + a 1 x R [ x ] . Then, m ( x ) r ( x ) = 0 by the relations. However, ann R ( m 1 ¯ ) = { 0 } . Hence, M R is not ZPZC.

We consider some conditions under which ZPZC modules are McCoy. Recall that a ring R is called right chain [36] if the lattice of right ideals of R is linearly ordered. Right chain rings can also be referred to as right uniserial rings.

Proposition 3.7

(see Example 4.4(3)). If R is a right uniserial ring and M is a ZPZC right R-module, then M is McCoy.

Proof

Let m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) = j = 0 q r j x j R [ x ] \ { 0 } with m ( x ) r ( x ) = 0 . Then, the ZPZC condition guarantees that each ann R ( m i ) is nonzero. Since R is right uniserial, we can find a nonzero element c k such that c k ann R ( m k ) ann R ( m i ) for all i . Hence, m i c k = 0 for all i , which implies that M is McCoy.□

From the opposite ring R o p of the ring R in [4, Section 3], one can notice easily that the converse of Proposition 3.7 does not hold true in general. From Example 3.2, one can also notice that there exists a ZPZC R -module but not McCoy, when the ring R is not right uniserial. In fact, Proposition 3.7 was started from the suspicion whether every module over a right uniserial ring is McCoy or not. The suspicion is very reasonable because every right uniserial ring is McCoy [37, Corollary 6.3]. Thus, it is worthwhile exploring the suspicion. However, we failed to solve it and so leave it as an open question:

Question 3.8

Can we delete the condition that M is ZPZC in Proposition 3.7?

From [29, p. 84], A module is called uniform if the intersection of any two nonzero submodules is nonzero. Recall that a submodule N of a module M is essential if for every submodule H of M , H N = { 0 } implies that H = { 0 } . A module M is uniform if and only if every nonzero submodule of M is essential. A ring is right uniform if it is uniform as a right module over itself. According to [29, p. 506], a module M is finitely cogenerated if for any family of submodules { M k k K } in M , k K M k = { 0 } implies k K M k = { 0 } for some finite subset K of K . The following theorem makes [7, Proposition 2.9] as a corollary.

Theorem 3.9

(cf. [7, Proposition 2.9] and see Example 4.2). Let K be an index set and let M k be a ZPZC right R-module for each k K .

  1. If R is right uniform, then the direct sum of M k as a right R-module, M = k K M k is McCoy.

  2. For an infinite index set K , if R R is uniform and finitely cogenerated, then the direct product of M k as a right R-module, M = k K M k is McCoy.

Proof

(1) Consider nonzero polynomials m ( x ) = i = 0 p ( m k i ) x i M [ x ] and r ( x ) R [ x ] with m ( x ) r ( x ) = 0 . Fix k K and let m k ( x ) = i = 0 p m k i x i M k [ x ] where m k i is the k th component of the coefficient ( m k i ) of m ( x ) for each i . Note that the set K = { k K m k ( x ) 0 } is finite. From m k ( x ) r ( x ) = 0 in M k [ x ] and M k is ZPZC for each k , we obtain ann R ( m k i ) { 0 } for each i and k K . The uniform condition guarantees that there exists a nonzero c k K , 0 i p ann R ( m k i ) . Thus, m ( x ) c = 0 which shows that M is McCoy.

(2) Consider nonzero polynomials m ( x ) = i = 0 p ( m k i ) x i M [ x ] and r ( x ) R [ x ] with m ( x ) r ( x ) = 0 . Fix k K and let m k ( x ) = i = 0 p m k i x i M k [ x ] where m k i is the k th component of the coefficient ( m k i ) of m ( x ) for each i . Then, for any finite subset K of K , k K , 0 i p ann R ( m k i ) { 0 } by the uniform condition and ZPZC condition. Since R is finitely cogenerated, there exists a nonzero c k K , 0 i p ann R ( m k i ) { 0 } . Hence, c annihilates m ( x ) , completing the proof.□

Corollary 3.10

[6, Proposition 2.15] If R is a right uniform and right ZPZC ring, then R is necessarily right McCoy.

Remark 3.11

The converse of Theorem 3.9(1) does not be true (see Remark 5.2(5)). Additionally, for right ideals ann R ( m 1 ) and ann R ( n 1 ) of the ring R in Example 3.2, ann R ( m 1 ) ann R ( n 1 ) = { 0 } illustrates that R is not right uniform. Thus, we cannot drop the condition that R is right uniform in Theorem 3.9(1).

4 Properties of ZPZC modules

In this section, we focus on the structure of ZPZC modules. Various examples and practical properties of ZPZC modules are provided. First, we examine the direct products of ZPZC modules.

Proposition 4.1

Let R i be a ring and M i be an R i -module for each i Γ . Then, each M i is ZPZC if and only if the direct product of M i , i Γ M i is a ZPZC ( i Γ R i ) -module.

Proof

This proof is almost the same as the proof of [6, Proposition 2.6].□

The ZPZC condition does not pass to direct products of ZPZC R -modules as an R -module.

Example 4.2

(cf. [12, Example 3.3]). We use the ring R and module M in Example 2.7. Note that R is local with a unique maximal ideal J = { 0 , a 0 , a 1 , a 0 + a 1 } . Set S = R [ y , z ] / K , where K is the ideal of the polynomial ring R [ y , z ] generated by the following relations:

a 0 y = a 0 z = 0 , a 1 y = a 1 z = 0 , y 2 = y z = z 2 = 0 .

Then, M ¯ = M [ y , z ] / M [ y , z ] K is an S -module. Put M 1 = M ¯ / M ¯ y and M 2 = M ¯ / M ¯ z . Because y ann S ( M 1 ) and z ann S ( M 2 ) , M 1 and M 2 are McCoy (and so ZPZC) as S -modules. After suppressing the bar notation for simplicity, we have

( m ( x ) , m ( x ) ) r ( x ) = ( 0 , 0 ) ( M 1 [ x ] M 2 [ x ] ) S [ x ] ( M 1 M 2 ) [ x ] S [ x ] ,

where m ( x ) = m 0 + m 1 x + m 2 x 2 in M 1 [ x ] and M 2 [ x ] , and r ( x ) = a 0 + a 1 x S [ x ] . But, ann S ( ( m 1 , m 1 ) ) = { 0 } . Thus, M 1 M 2 is not a ZPZC S -module.

Next, we introduce some basic properties of ZPZC modules.

Proposition 4.3

Let R be a ring and M be a right R-module.

  1. (cf. [6, Example 2.11(1)]). Every submodule of a ZPZC module is ZPZC.

  2. (see Example 4.5). R is right ZPZC if and only if every cyclic free right R-module is ZPZC.

  3. (cf. [7, Proposition 2.3(2)]). M is ZPZC if and only if every finitely generated submodule of M is ZPZC.

Proof

(1) This is obvious.

(2) This follows from M = m R R as a right R -module.

(3) The forward direction is clear by (1) and so we only deal with the reverse direction. Let m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) R [ x ] \ { 0 } with m ( x ) r ( x ) = 0 . Consider the submodule N of M generated by all m i . Then, m ( x ) N [ x ] . By hypothesis, N is ZPZC. Therefore, for each m i , there exists a nonzero c i R such that m i c i = 0 , which shows that M is ZPZC.□

We exemplify the results of Proposition 4.3.

Example 4.4

(1) The author showed that the class of right ZPZC rings is not closed under subrings in [6, Example 2.11(1)]. However Proposition 4.3(1) illuminates that every right ideal of a right ZPZC ring must be right ZPZC (without identity).

(2) Let M = m T 2 ( Z 6 ) be the free right module over T 2 ( Z 6 ) . Since T 2 ( Z 6 ) is a right ZPZC ring but not right McCoy by [5, Proposition 10.2] and [6, Theorem 2.2], M is ZPZC but not McCoy by Proposition 4.3(2).

In view of Proposition 4.3(2), one may suspect that if M R is either cyclic or free over a right ZPZC ring R , then M is ZPZC. But the following example eliminates the possibility of each suspicion.

Example 4.5

(1) Let R = Z 2 a 0 , a 1 , a 2 , b 0 , b 1 be the free algebra with identity and noncommuting indeterminates a 0 , a 1 , a 2 , b 0 , b 1 over Z 2 . Obviously, R is a domain and so a right ZPZC ring. Let L = m R be the free R -module and N be an R -submodule of L generated by the following relations:

m a 0 b 0 = 0 , m a 0 b 1 + m a 1 b 0 = 0 , m a 1 b 1 + m a 2 b 0 = 0 , m a 2 b 1 = 0 .

Then, M = L / N is a cyclic R -module. After suppressing the bar notation in M , consider the nonzero polynomials m ( x ) = m a 0 + m a 1 x + m a 2 x 2 M [ x ] and r ( x ) = b 0 + b 1 x R [ x ] . Then m ( x ) r ( x ) = 0 by the relations. But, ann R ( m a 1 ) = { 0 } . Thus, M is not a ZPZC R -module.

(2) Let R = T 2 ( Z 6 ) be the right ZPZC ring and M = m 1 R m 2 R be the free R -module with basis { m 1 , m 2 } . After suppressing the bar notation in Z 6 , consider the following nonzero polynomials:

m ( x ) = m 1 0 4 0 0 , m 2 0 3 0 0 + m 1 2 1 0 0 , m 2 3 1 0 0 x + m 1 5 0 0 0 , m 2 5 0 0 0 x 2 M [ x ] , r ( x ) = 0 1 0 0 + 0 0 0 1 x R [ x ] .

Then, m ( x ) r ( x ) = 0 . However, an easy calculation shows that if r R annihilates the middle coefficient of m ( x ) on the right side, then r = 0 . This means that the free R -module M is not ZPZC.

Remark 4.6

(1) (see Remark 4.23(3)). As well-known results (see [29, p. 21, Proposition 4.3, and p. 127]), the following irreversible implications hold:

free projective flat torsionfree.

Thus, Example 4.5(2) shows that a projective or flat or torsionfree module over a right ZPZC ring may not be ZPZC.

(2) Based on Example 2.8, one may suspect that every injective module over a right ZPZC ring is ZPZC. For a counterexample, let D be any division ring, R = T 2 ( D ) , and E = M 2 ( D ) . Then, D is certainly a semisimple ring and thus E R is injective by [29, Example 3.43], when R is a right ZPZC ring by [6, Corollary 2.3]. From [6, Remark 3.9], we can conclude that E R is not a ZPZC R -module.

(3) (cf. [7, Proposition 2.10]). In Example 4.5(2), each cyclic free R -module m 1 R and m 2 R is ZPZC by Proposition 4.3(2). Additionally, m 1 R is R -isomorphic to m 2 R . Thus, a direct sum of copies of a ZPZC module does not have to be ZPZC.

(4) (cf. [7, Proposition 2.15]). Note that T 2 ( Z 6 ) is a principal right ideal ring. Thus, Example 4.5(2) shows that the commutative condition must be required in [7, Proposition 2.15]. Additionally, T 2 ( Z 6 ) is also an I 2 ( S ) -principal right ideal ring for any multiplicative subset S of Z 6 (see [35, Theorem 2.14]). Hence, there is a free module that is not a ZPZC module over an S -principal right ideal ring.

Recall that a module is called Bézout if every finitely generated submodule is cyclic. We provide a sufficient condition under which a right module over a right ZPZC ring is ZPZC.

Proposition 4.7

(cf. [7, Proposition 2.7]). Let R be a right ZPZC ring. If M is a cyclic submodule of a flat right R-module, then M is ZPZC.

Proof

Let M = m R and consider m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) = j = 0 q r j x j R [ x ] { 0 } with m ( x ) r ( x ) = 0 . Then for each m i , we can write m i = m a i for some a i R . By [29, Theorem 4.24], the flat condition guarantees that a ( x ) r ( x ) = 0 where a ( x ) = i = 0 p a i x i R [ x ] . Hence, m i c i = m a i c i = 0 for some nonzero c i R and each i , as desired.□

The hypothesis “Bézout and flat” of the following corollary is not superfluous by Example 4.5.

Corollary 4.8

Let R be a right ZPZC ring. If a right R-module M is Bézout and flat, then M is ZPZC.

Generally, although some R -submodule is ZPZC of an R -module M , M may not be ZPZC. We find a type of R -submodule that the ZPZC condition can be lifted up to R -overmodule.

Proposition 4.9

(cf. [6, Proposition 2.12]). Let M be a right module over a right ZPZC ring R and e be a nontrivial central idempotent such that r e r 0 for all nonzero r R with 1 e r e . If M e R is a ZPZC e R -module and M ( 1 e ) R is a ZPZC ( 1 e ) R -module, then M R is ZPZC.

Proof

Let m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) = j = 0 q r j x j R [ x ] { 0 } with m ( x ) r ( x ) = 0 . Then, m ( x ) r ( x ) e = m e ( x ) e r ( x ) = 0 and m ( x ) r ( x ) ( 1 e ) = m ( 1 e ) ( x ) ( 1 e ) r ( x ) = 0 . Because one of e r ( x ) and ( 1 e ) r ( x ) is nonzero, there exists a nonzero element e c i e R or ( 1 e ) d i ( 1 e ) R such that m i e c i = 0 or m i ( 1 e ) d i = 0 for each i . Thus, ann R ( m i ) { 0 } for each i .□

Question 4.10

Is the converse of Proposition 4.9 true?

Next, we consider matrix modules. According to [7, Proposition 2.6], every cyclic module over a right duo ring is McCoy. Based on this fact, we can extend [6, Lemma 2.1 and Theorem 2.2] to cyclic modules by the same arguments.

Lemma 4.11

(cf. [6, Lemma 2.1]). Let R be a right duo ring and M = m R be a cyclic right R-module. If A = [ m i , j ] T n ( M ) and some ann R ( m i , i ) { 0 } for some i 1 , then ann T n ( R ) ( A ) { 0 } .

Proof

We first write m i , j = m r i , j for some r i , j R and each i , j . Without loss of generality, we assume that i 1 is the smallest index such that ann R ( m i , i ) { 0 } . Fix β i R \ { 0 } such that m i , i β i = 0 . Consider the matrix

B = ( β 1 ) E 1 , i + ( r 1 , 1 β 2 ) E 2 , i + ( r 2 , 2 r 1 , 1 β 3 ) E 3 , i + + ( r i 2 , i 2 r 1 , 1 β i 1 ) E i 1 , i + ( r i 1 , i 1 r 1 , 1 β i ) E i , i .

Then, the upper triangular matrix B is nonzero. We will choose the β ’s so that A B = 0 . Note that the equation A B = 0 is equivalent to the set of equations

0 = m 1 , 1 β 1 + m 1 , 2 ( r 1 , 1 β 2 ) + + m 1 , i ( r i 1 , i 1 r 1 , 1 β i ) , 0 = m 2 , 2 ( r 1 , 1 β 2 ) + m 2 , 3 ( r 2 , 2 r 1 , 1 β 3 ) + + m 2 , i ( r i 1 , i 1 r 1 , 1 β i ) , 0 = m i 1 , i 1 ( r i 2 , i 2 r 2 , 2 r 1 , 1 β i 1 ) + m i 1 , i ( r i 1 , i 1 r 2 , 2 r 1 , 1 β i ) , 0 = m i , i ( r i 1 , i 1 r 2 , 2 r 1 , 1 β i ) .

Now notice that ann r ( m i , i ) is a two-sided ideal of R since R is right duo. Therefore,

0 r i 1 , i 1 r 2 , 2 r 1 , 1 β i ann R ( m i , i )

from our choice β i . Thus, the last equation holds. Working upwards, we will recursively choose β i 1 , β i 2 , , β 1 R such that the equations hold. For each i 1 k 1 , we can find β k such that

( r k , k r 1 , 1 ) β k = ( r k , k + 1 ) ( r k , k r 1 , 1 ) β k + 1 + + ( r k , i ) ( r i 1 , i 1 r 1 , 1 ) β i = s = k + 1 i r k , s t = s 1 1 r t , t β s R ( r k , k r 1 , 1 ) R

from the fact that a R = R a R for every a R . Hence, 0 B ann T n ( R ) ( A ) .□

Theorem 4.12

(cf. [6, Theorem 2.2]). Fix a positive integer n. If R is a right duo ring and M is a cyclic right R-module, then T n ( M ) is a ZPZC right T n ( R ) -module.

Proof

Let A ( x ) T n ( M ) [ x ] and B ( x ) T n ( R ) [ x ] { 0 } such that A ( x ) B ( x ) = 0 . Since there is the natural identifications T n ( M ) [ x ] = T n ( M [ x ] ) and T n ( R ) [ x ] = T n ( R [ x ] ) , we can write A ( x ) = [ a i , j ( x ) ] and B ( x ) = [ b i , j ( x ) ] where a i , j ( x ) M [ x ] and b i , j ( x ) R [ x ] . Because B ( x ) 0 , we have b i , j ( x ) 0 for the maximal i , and then the maximal j . In particular, a i , i ( x ) b i , j ( x ) = 0 . Since R is right McCoy by [5, Theorem 8.2], we obtain that for each coefficient a i , i , k of a i , i ( x ) , ann R ( a i , i , k ) { 0 } . Now, the previous lemma applies to show that each right annihilator set of each coefficient of A ( x ) is nonzero in T n ( R ) .□

From Theorem 4.12, we obtain many examples of ZPZC modules. For instance, if R is a division ring, then R is clearly (right) duo and so (right) McCoy. Thus, a cyclic module over a division ring is McCoy by [7, Proposition 2.6]. This implies that if M is a cyclic module over a division ring R , then T n ( M ) is a ZPZC T n ( R ) -module. Further, from the diagrams in [5] and Section 2, we obtain the following:

Corollary 4.13

Fix an integer n 1 , and let M be a cyclic module over a ring R. Then, T n ( M ) is a ZPZC T n ( R ) -module if one of the following assertions holds true:

  1. R is (right) duo and M is McCoy.

  2. R is (right) Gaussian and M is Gaussian.

  3. R is commutative and M is Armendariz.

  4. (see Remark 5.4). R is a division ring and M is reduced.

In Theorem 4.12, the hypothesis that M is cyclic is not superfluous by the following example or Example 5.5.

Example 4.14

Let R = Z 6 be the right ZPZC ring and M = m 1 R m 2 R be the free R -module with basis { m 1 , m 2 } . After suppressing the bar notation in Z 6 , consider the following nonzero polynomials:

A ( x ) = 0 m 1 4 + m 2 3 0 0 + m 1 2 + m 2 3 m 1 + m 2 0 0 x + m 1 5 + m 2 5 0 0 0 x 2 T 2 ( M ) [ x ] , B ( x ) = 0 1 0 0 + 0 0 0 1 x T 2 ( R ) [ x ] .

Then, A ( x ) B ( x ) = 0 . However, an easy calculation shows that if C T 2 ( R ) annihilates the middle coefficient of A ( x ) on the right side, then C = 0 . Thus, the right T 2 ( R ) -module T 2 ( M ) is not ZPZC.

Given an R -module M and n 2 , we denote B n ( M ) = { [ m i j ] T n ( M ) m 11 = = m n n } and H n ( M ) = { [ m i j ] B n ( M ) m s t = m ( s + 1 ) ( t + 1 ) for s = 1 , , n 2 and t = 2 , , n 1 } . Then, B n ( M ) is a B n ( R ) -module and H n ( M ) is a H n ( R ) -module.

Proposition 4.15

(cf. [7, Proposition 2.12]). Fix n 1 . Then, the following statements are equivalent:

  1. M is a ZPZC R -module.

  2. B n ( M ) is a ZPZC B n ( R ) -module.

  3. H n ( M ) is a ZPZC H n ( R ) -module.

Proof

This proof is nearly identical to the proof of [6, Theorem 3.5].□

If we consider M n ( M ) , T n ( M ) , B n ( M ) , and H n ( M ) for n 2 as R -modules under the usual scalar multiplication, the above equivalences will be broken even when R is commutative.

Example 4.16

We use the ring R and module M in Example 3.2. Take m ( x ) = m 0 + m 1 x , n ( x ) = n 0 + n 1 x + n 2 x 2 in M [ x ] , and r ( x ) = a 0 + a 1 x R [ x ] . Then m ( x ) r ( x ) = 0 = n ( x ) r ( x ) . Now consider the nonzero polynomial H ( x ) = m 0 n 0 0 m 0 + m 1 n 1 0 m 1 x + 0 n 2 0 0 x 2 H 2 ( M ) [ x ] B 2 ( M ) [ x ] T 2 ( M ) [ x ] M 2 ( M ) [ x ] . Then H ( x ) r ( x ) = 0 . Since ann R ( m 1 ) ann R ( n 1 ) = { 0 } , m 1 n 1 0 m 1 r = 0 implies r = 0 . Thus, M 2 ( M ) , T 2 ( M ) , B 2 ( M ) , and H 2 ( M ) are not ZPZC R -modules.

Remark 4.17

If M is a McCoy R -module, then M n ( M ) , T n ( M ) , B n ( M ) , and H n ( M ) are McCoy as R -modules. Indeed, if m ( x ) r ( x ) = 0 and n ( x ) r ( x ) = 0 for some nonzero polynomials m ( x ) , n ( x ) M [ x ] and r ( x ) R [ x ] , then ( m ( x ) + n ( x ) x k ) r ( x ) = 0 , where k = deg ( m ( x ) ) + 1 . Therefore, there exists a nonzero r R such that m ( x ) r = n ( x ) r = 0 . Thus, m ( x ) n ( x ) 0 m ( x ) r ( x ) = 0 implies m ( x ) n ( x ) 0 m ( x ) r = 0 . From this idea, one can easily show that M n ( M ) , T n ( M ) , B n ( M ) and H n ( M ) are McCoy as R -modules for each positive integer n , when M is McCoy.

According to [29, Theorem 10.6], a multiplicative subset T of a ring R is a right denominator set if and only if the right ring of fractions with respect to T , R T 1 exists. A right R -module M is called T-torsionfree if for m M and t T , m t = 0 implies m = 0 . According to [38, Proposition 10.11(a)], every right R T 1 -module is T -torsionfree as a right R -module.

Theorem 4.18

Let T be a right denominator set in a ring R. If M is a right R T 1 -module, then M is a right ZPZC R-module if and only if M is a right ZPZC R T 1 -module.

Proof

( ) Let m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) = j = 0 q r j x j R T 1 [ x ] \ { 0 } with m ( x ) r ( x ) = 0 . Then, we can write that for each j , r j = b j t 1 , where t T . Take b ( x ) = j = 0 q b j x j R [ x ] \ { 0 } . From the fact that

m ( x ) r ( x ) t 1 = m ( x ) b 0 t + b 1 t x + + b q t x q t 1 = 0 ,

we obtain that m ( x ) b ( x ) = 0 . Since M R is ZPZC, there exists a nonzero c i R such that m i c i = 0 for each i . Therefore, m i c i 1 1 = 0 for each i . Hence, we can conclude that M is a ZPZC R T 1 -module.

( ) Let m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) = j = 0 q b j x j R [ x ] \ { 0 } with m ( x ) r ( x ) = 0 . If we let b ( x ) = j = 0 q b j 1 1 x j R T 1 [ x ] \ { 0 } , then m ( x ) b ( x ) = 0 . Since M R T 1 is ZPZC, there exists a nonzero c i = r i t i 1 R T 1 such that m i c i = 0 . From the fact that

( m i c i ) t i 1 = m i r i t i t i 1 = m i r i 1 = 0 ,

for each i , we obtain that m i r i = 0 for each i . Hence, we are done.□

According to [29, (10.17)], a ring R is right Ore if and only if the classical right quotient ring of R , Q c l r ( R ) exists.

Corollary 4.19

(cf. [7, Theorem 2.17]). Let R be a right Ore ring. If M is a Q c l r ( R ) -module. then M is a ZPZC R-module if and only if M is a ZPZC Q c l r ( R ) -module.

Following the same approach in the proof of Theorem 4.18, we partially extend [6, Theorem 3.12].

Corollary 4.20

If T is a right denominator set in a ring R, then R is right ZPZC if and only if R T 1 is right ZPZC.

The proof of the next theorem is almost the same as the proof of [6, Theorem 3.15], but we insert it for the sake of completeness.

Theorem 4.21

(cf. [15, Proposition 2.6(1)]). The class of ZPZC R -modules is closed under direct limits.

Proof

Let D = M u , α u , v be a direct system of right ZPZC R -modules M u for u U and R -module homomorphisms α u , v : M u M v for each u v and m M u , satisfying α u , v ( m ) = m , where U is a directed partially ordered set. Set M = lim M u to be the direct limit of D with ι u : M u M and ι v α u , v = ι u . Let a and b M . Then, a = ι u ( a u ) , b = ι v ( b v ) for some u , v U , and there is w U such that u w , v w . Define a + b = ι w ( α u w ( a u ) + α v w ( b v ) ) and a r = ι w ( α u w ( a u r ) ) , where r R . Under the above operations, M forms a right R -module. Now consider polynomials m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) = j = 0 q r j x j R [ x ] \ { 0 } with m ( x ) r ( x ) = 0 . Then, there exists w U such that m ( x ) M w [ x ] . Therefore, we get m ( x ) r ( x ) = 0 in M w [ x ] . Thus for each nonzero m i , there exists nonzero c i R such that m i c i = 0 in M w since M w is a ZPZC R -module. Hence, for each nonzero m i , m i c i = 0 in M , completing the proof.□

For an index set I of a chain A , a subset J of I is said to be dense in I if between any two distinct elements x , y I with x < y , there exists at least one element z J with x z y (see [39, 2.6.6]).

Corollary 4.22

(cf. [15, Proposition 2.6(2)]). Let A be a chain of R-modules M i and J be a subset of I. If M j is ZPZC for all j J and J is dense in I, then M = i I M i is a ZPZC R -module.

Proof

Note that M = i I M i is a direct limit of ZPZC R -modules. Thus, Theorem 4.21 applies.□

Remark 4.23

(1) In [6, Proposition 3.15], the condition “with injective maps” must be required. However, the condition does not be required in Theorem 4.21.

(2) By a similar approach as the proof of Theorem 4.21, the class of McCoy R -modules is closed under direct limits. Of course, the condition “with injective maps” does not be required.

(3) Due to [29, Theorem 4.34], a right R -module M is flat if any only if M is a direct limit of finitely generated free R -modules. Thus, if a right module M over a right McCoy ring is flat, then M is McCoy by (2). This fact also follows from [29, Theorem 4.24] and extends [7, Corollary 2.11] by weakening the projective condition to the flat condition.

In view of Remark 4.6(1) and Remark 4.23(3), it is of interest to provide an explicit example of a right module M over a right McCoy ring that is torsionfree but not McCoy. However, the author did not find any clue for this question so far. Thus, we leave it as an open question:

Question 4.24

Over a right McCoy ring, is there a torsionfree module but not McCoy?

We end this section with the following remarks.

Remark 4.25

In the diagram of Section 2, we can add the implication “McCoy ZPZC”. The other implications (except the transitivity) do not hold according to Examples 2.5 and 2.7.

Remark 4.26

Anderson and Chun [12] defined dual McCoy, content McCoy and dual content McCoy modules. We can extend these concepts to ZPZC modules. A right R -module M is called dual ZPZC if for nonzero polynomials m ( x ) M [ x ] and r ( x ) = j = 0 q r j x j R [ x ] , m ( x ) r ( x ) = 0 implies that there exists a nonzero n j M such that n j r j = 0 for each j . A right R -module M is said to be content ZPZC (resp., dual-content ZPZC), if for nonzero polynomials m ( x ) = i = 0 p m i x i M [ x ] and r ( x ) = j = 0 q r j x j R [ x ] , m ( x ) r ( x ) = 0 implies that there exists a nonzero c i C ( r ( x ) ) (resp., n j S ( m ( x ) ) ) such that m i c i = 0 for each i (resp., n j r j = 0 for each j ), where C ( r ( x ) ) is the right ideal of R generated by the coefficients of r ( x ) (resp., S ( m ( x ) ) is the R -submodule of M generated by the coefficients of m ( x ) ).

(1) Even though M is a ZPZC R -module, M need not be dual ZPZC when we consider the opposite ring of the ring in [4, Section 3]. Conversely, the ring in [4, Section 3] shows that dual ZPZC modules need not be ZPZC.

(2) ZPZC modules do not imply content ZPZC modules. We use once again the ring R and the R -module M in Example 3.2. Consider the nonzero polynomials n ( x ) = n 0 + n 1 x + n 2 x 2 M [ x ] and r ( x ) = a 0 + a 1 x R [ x ] with n ( x ) r ( x ) = 0 . Note that C ( r ( x ) ) = a 0 R + a 1 R = { 0 , a 0 , a 1 , a 0 + a 1 } and ann R ( n 1 ) = { 0 , b 0 } . Thus, the ZPZC module M is not content ZPZC.

(3) There is a ZPZC module but not dual content ZPZC. Let R = Z 2 b 0 , b 1 , b 2 , c be the free algebra with identity and noncommuting indeterminates b 0 , b 1 , b 2 , c over Z 2 . Then, it is clear that R is a domain. Let L = m 0 R m 1 R be the free R -module with basis { m 0 , m 1 } . Set M = L / N , where N is the R -submodule of L generated by the following relations: for each r R ,

m 0 b 0 = 0 , m 0 b 1 + m 1 b 0 = 0 , m 0 b 2 + m 1 b 1 = 0 , m 1 b 2 = 0 , m 0 r c = m 1 r c = 0 .

We simply denote m = m + N . One can easily notice that M is a McCoy R -module since m c = 0 for every m M . Now consider the following nonzero polynomials m ( x ) = m 0 + m 1 x M [ x ] and r ( x ) = b 0 + b 1 x + b 2 x 2 R [ x ] . Then, m ( x ) r ( x ) = 0 by the relations. But there is no nonzero element n M such that n b 1 = 0 . Thus, the ZPZC R -module M is not dual content ZPZC.

(4) To show that there is a dual content ZPZC module but not ZPZC, we use the ring R and the module M in Example 2.7. Consider nonzero polynomials u ( x ) = i = 0 p u i x i M [ x ] and v ( x ) = j = 0 q v j x j R [ x ] with u ( x ) v ( x ) = 0 . Assume that there exists a coefficient v of v ( x ) with 1 in its support and is minimally chosen. If the element m 0 appears in the support of u k and k is minimally chosen, then m 0 must appear in the support of the ( k + )-degree coefficient of u ( x ) v ( x ) , which is a contradiction. Therefore, m 0 must not appear in the support of any coefficient of u ( x ) . By the same reason, the elements m 1 and m 2 must not appear in the support of any coefficient of u ( x ) and thus the elements m 1 a 0 and m 1 a 1 must not appear in the support of any coefficient of u ( x ) . This means that u ( x ) = 0 which contradicts to the hypothesis that u ( x ) is a nonzero polynomial in M [ x ] . Hence, there is no coefficient of v ( x ) with 1 in its support. Now note that one of the following three elements m 0 a 1 = m 1 a 0 , m 1 a 1 = m 2 a 0 , and m 1 a 0 + m 1 a 1 must be in the submodule S ( u ( x ) ) of M . Since there is no coefficient of v ( x ) with 1 in its support, we obtain that m 0 a 1 v ( x ) = 0 or m 1 a 1 v ( x ) = 0 or ( m 0 a 1 + m 1 a 1 ) v ( x ) = 0 by the relations, which implies that M is a dual content McCoy module. Consequently, there exists a dual content ZPZC module but not ZPZC.

5 Three solved problems on ZPZC rings

We devote this section to providing answers to some questions raised in [6], which are related to ZPZC rings. First, a ring that is left McCoy and right ZPZC, but not right McCoy is constructed for giving a negative answer of [6, Question 2.8(2)].

Example 5.1

Let E = Z 2 a 0 , a 1 , a 2 , b 0 , b 1 , c be the free algebra with identity and six noncommuting indeterminates a 0 , a 1 , a 2 , b 0 , b 1 , c over Z 2 . Set R = E / I where I is the ideal of E generated by the following relations:

a 0 b 0 = 0 , a 0 b 1 = a 1 b 0 , a 1 b 1 = 0 , a 2 b i = 0 ( 0 i 1 ) , a i a j = a i c = 0 ( 0 i 1 , 0 j 2 ) , a 2 a j = a j ( 0 j 2 ) , a 2 c = c a 2 = c , b i a j = b i b k = b i c = c a i = c b i = c 2 = 0 ( 0 i , k 1 , 0 j 2 ) .

One can check, via the diamond lemma [27], that these relations form a reduced system. We identify a = a + I in R for simplicity. Then, we can write each element r R uniquely in the following form:

r = α + α 0 a 0 + α 1 a 1 + α 2 a 2 + ( β 0 + β 1 a 1 ) b 0 + β 2 b 1 + γ c ,

where α , α 0 , α 1 , α 2 , β 0 , β 1 , β 2 , γ Z 2 .

Claim 1. The ring R is not right McCoy.

Proof

Let a ( x ) = a 0 + a 1 x + a 2 x 2 and b ( x ) = b 0 + b 1 x in R [ x ] . Then, a ( x ) b ( x ) = 0 by the first row of the relations. It suffices to show that if a ( x ) r = 0 for some r R , then r = 0 . Write r in the unique form. From a 0 r = 0 , we obtain α = β 2 = 0 . Similarly, a 1 r = 0 implies β 0 = 0 . Finally, since α = β 0 = β 2 = 0 , we have that a 2 r = 0 implies α 0 = α 1 = α 2 = β 1 = γ = 0 . Thus, r = 0 , as desired.□

Claim 2. The ring R is right ZPZC.

Proof

Let p ( x ) = i = 0 m p i x i and q ( x ) = j = 0 n q j x j be nonzero polynomials with p ( x ) q ( x ) = 0 in R [ x ] . First, observe that if a nonzero element r in the unique form has zero constant term, then r can be annihilated by one of the following elements in { b 0 , b 1 , b 0 + b 1 } on the right side. Thus, we may assume that there exists a coefficient of p ( x ) with 1 in its support. Let k be the smallest index such that p k has a nonzero constant term. Now we will use a similar argument as in the proof of [5, Proposition 3.2]. Assume that q has a nonzero constant term in its support and is minimally chosen. If we compute the ( k + ) -degree coefficient of p ( x ) q ( x ) , then, by the minimality of k and , we can see that the ( k + ) -degree coefficient of p ( x ) q ( x ) must have a nonzero constant term. This contradicts to the fact that p ( x ) q ( x ) = 0 . Thus, every nonzero coefficient of q ( x ) must have zero constant term. Under the situation that the nonzero constant element 1 appears in the support of p k with the smallest index k , and 1 does not appear in the support of any coefficient of q ( x ) , we have the following two cases.

Case 1. a 2 is not in the support of p k .

Suppose that there exists a coefficient q of q ( x ) with a 0 in its support and is minimally chosen. If a 2 appears in the support of some coefficient among { p 0 , , p k 1 } , then we can find the smallest index 0 s k 1 such that a 2 is in the support of p s and a 2 is not in the support of any p 0 , , p s 1 . Since 1 does not appear in the support of any p 0 , , p k 1 and a 2 does not appear in the support of any p 0 , , p s 1 , we obtain that the nonzero element a 2 a 0 = a 0 must be in the support of the ( s + )-degree coefficient of p ( x ) q ( x ) . This is a contradiction. Therefore, a 2 must not appear in the support of any coefficient among { p 0 , , p k 1 } . This implies that a 0 must be in the support of the ( k + )-degree coefficient of p ( x ) q ( x ) . This is also a contradiction. Thus, a 0 cannot be in the support of any coefficients of q ( x ) . Similarly, a 1 , a 2 , b 0 , b 1 , c and so a 1 b 0 do not appear in the support of any coefficients of q ( x ) . Thus, q ( x ) = 0 , which is a contradiction.

Case 2. a 2 is in the support of p k .

Since 1 and a 2 are in the support of p k , we easily see that p k c = 0 by the relations.

Hence, we can reach that if there exists the smallest index k such that 1 is in the support of p k , then 1 must not be in the support of any coefficient of q ( x ) and a 2 must be in the support of p k . Under this result, we finally assume that there exists another coefficient of p ( x ) with 1 in its support among { p k + 1 , , p m } . Let u be the smallest index such that p u has a nonzero constant term for k + 1 u m . If a 2 is not in the support of p u , then we can obtain a contradiction by the same argument of Case 1. Thus, a 2 must be in the support of p u . Repeating this argument, we obtain that for each coefficient p i of p ( x ) , if 1 appears in the support of p i , then a 2 must be also in the support of p i . Combining the above results, we can conclude that for each i ,

p i b 0 = 0 or p i b 1 = 0 or p i ( b 0 + b 1 ) = 0 , if 1 does not appear in the support of p i p i c = 0 , if 1 appears in the support of p i ,

which confirms that the ring R is right ZPZC.□

Claim 3. The ring R is left McCoy.

Proof

Let p ( x ) = i = 0 m p i x i and q ( x ) = j = 0 n q j x j be nonzero polynomials with p ( x ) q ( x ) = 0 in R [ x ] . If q ( x ) has no coefficient with 1 in its support, then b 0 q ( x ) = 0 . Thus, we may assume that there exists a coefficient with 1 in its support. Let k be the smallest index such that q k has this property. Then, we can show that each coefficient p i of p ( x ) must have zero constant term. Following a similar approach in Case 1 of Claim 2, we observe that a 2 must appear in the support of q k . Following the same method as the last paragraph of Claim 2, we can also observe that a 2 must be in the support of q v for each coefficient q v with 1 of q ( x ) . Furthermore, if there is a coefficient q s of q ( x ) such that a 2 is in the support of q s with zero constant term (under the case that a 2 is in the support of q v for each coefficient q v with 1 of q ( x ) ), we can also obtain a contradiction. Thus, we can reach that a 2 must be in the support of each coefficient with 1 of q ( x ) , and a 2 cannot be in the support of each coefficient with zero constant term of q ( x ) . Hence, c q ( x ) = 0 , leading to the conclusion that R is left McCoy.□

Capitalizing on Example 5.1, we have some remarks.

Remark 5.2

  1. From Example 5.1, one may ask that if an ( S , T ) -bimodule M is McCoy as an S -module, then M is ZPZC as a T -module. As mentioned in [6, Example 2.9 and Remark 2.10(1)], the ring R in [4, Section 3] provides a negative answer.

  2. If we think M = R R in Example 5.1, then Example 5.1 shows that there exists a dual McCoy and ZPZC module but not McCoy. Further, note that M is both content and dual content ZPZC.

  3. Referring [13], we note that the ring R in Example 5.1 is left outer McCoy. Of course, the ring R is left Camillo. But there is no nonzero element r R such that r f ( x ) = r g ( x ) = 0 , for some nonzero polynomials f ( x ) , g ( x ) with f ( x ) g ( x ) = 0 in R [ x ] . To show this, consider the two nonzero polynomials p ( x ) = ( 1 + a 2 ) and q ( x ) = a 2 with p ( x ) q ( x ) = 0 . An easy calculation shows that if r p ( x ) = r q ( x ) = 0 , then r = 0 .

  4. The ring R in Example 5.1 is not right duo since R b 0 = { 0 , b 0 , a 1 b 0 , b 0 + a 1 b 0 } b 0 R = { 0 , b 0 } (or by [5, Proposition 8.2]). Thus, there exists a right ZPZC ring which is neither right McCoy nor right duo.

  5. For the ring R in Example 5.1, R o p is not right uniform since a 0 R o p b 0 R o p = { 0 } . Thus, the converse of Theorem 3.9(1) does not be true in general by Example 5.1.

In [6, Questions 2.5 and 3.10], the author asked whether T n ( R ) is a right ZPZC ring or not, where R is one of a domain, a reduced, a right ZPZC ring. We provide answers to these questions negatively.

Example 5.3

Let E = Z 2 a 0 , a 1 , a 2 , b 1 , c 0 , c 1 be the free algebra with identity and noncommuting indeterminates a 0 , a 1 , a 2 , b 1 , c 0 , c 1 over Z 2 . Set R = E / I where I is the ideal of E generated by the following relations:

a 0 c 1 = b 1 c 0 , a 1 c 1 = b 1 c 1 + a 2 c 0 .

We simply denote a = a + I in R . Note that R is a domain (and thus it is both reduced and ZPZC). To show that T 2 ( R ) is not right ZPZC, consider the following nonzero polynomials in T 2 ( R ) [ x ] :

A ( x ) = a 0 0 0 0 + a 1 b 1 0 0 x + a 2 a 2 0 0 x 2 and B ( x ) = 0 0 0 c 0 + 0 c 1 0 c 1 x .

Then, A ( x ) B ( x ) = 0 by the relations. Suppose that a 1 b 1 0 0 p q 0 r = 0 0 0 0 in T 2 ( R ) . Then, we obtain p = q = r = 0 in R . Thus, a 1 b 1 0 0 is not a left zero-divisor in T 2 ( R ) , as desired.

Remark 5.4

Let M be an R -module. By Example 5.3, T n ( M ) need not be a ZPZC T n ( R ) -module even when M is a reduced module over a domain R .

Every division ring is duo. Thus, if R is a division ring, then, T n ( R ) is a right ZPZC ring by [6, Theorem 2.2]. Based on this fact, one may suspect that if M is a module over a division ring R , then T n ( M ) is a ZPZC T n ( R ) -module for any positive integer n . But the following example eliminates the possibility of the suspicion.

Example 5.5

We use the ring R and module M in Example 2.3. Then, M is a ZPZC R -module. Consider the following nonzero polynomials

A ( x ) = m 2 0 0 0 + m 1 m 2 0 0 x + m 1 + m 2 m 1 + m 2 0 0 x 2 T 2 ( M ) [ x ] , B ( x ) = 0 ¯ 0 ¯ 0 ¯ 1 ¯ + 0 ¯ 1 ¯ 0 ¯ 1 ¯ x T 2 ( R ) [ x ] .

Then, A ( x ) B ( x ) = 0 . Suppose that m 1 m 2 0 0 p ¯ q ¯ 0 ¯ r ¯ = 0 0 0 0 in T 2 ( M ) . Then, we obtain p ¯ = q ¯ = r ¯ = 0 ¯ in R . Thus, ann T 2 ( R ) m 1 m 2 0 0 = 0 ¯ 0 ¯ 0 ¯ 0 ¯ , as needed.

Remark 5.6

(1) Let S 2 ( R ) be the set of all 2 × 2 scalar matrices over R in Example 5.5. Then, S 2 ( R ) is a subring of T 2 ( R ) with the same identity. Clearly, T 2 ( M ) is a right S 2 ( R ) -module. Since R is (ring) isomorphic to S 2 ( R ) , T 2 ( M ) is a McCoy S 2 ( R ) -module. Thus, we obtain that a McCoy R 1 -module may not be ZPZC R 2 -module, where R 1 is a subring of a ring R 2 with the same identity.

(2) We consider the rings R 1 = Z [ x ] Z [ x ] 0 Z , R 2 = T 2 ( Z [ x ] ) , and the nonzero polynomials

F ( y ) = 0 x 0 0 + x 1 0 0 y + 1 0 0 0 y 2 , G ( y ) = 0 1 0 0 + 0 0 0 1 y R 1 [ y ]

in [6, Example 2.11(1)]. Then, R 1 is a subring of a right ZPZC ring R 2 with the same identity. Since the middle coefficient of F ( y ) is a left zero-divisor in R 2 but not in R 1 , M = R 2 is a ZPZC R 2 -module but not a ZPZC R 1 -module.


Dedicated to Ilwon Kang for his retirement from Kyung Hee University.


Acknowledgments

The author is very thankful to the anonymous referees for the many helpful suggestions and valuable comments. The author also wishes to express his sincere appreciation to Professor Nam Kyun Kim and to Professor Pace P. Nielsen for the idea which improved the quality of this article and to Professor Ilwon Kang for his encouraging comments and supports which were helpful to construct Example 5.1 for seven years.

  1. Funding information: This research was supported by Basic Science Research Program through the National Research Foundation of Korea(NRF) funded by the Ministry of Education (NRF-2021R1l1A1A01041451).

  2. Conflict of interest: The author states no conflict of interest.

  3. Data availability statement: Data sharing is not applicable to this article as no datasets were generated or analyzed during the current study.

References

[1] N. H. McCoy, Remarks on divisors of zero, Amer. Math. Monthly 49 (1942), no. 5, 286–295, https://doi.org/10.1080/00029890.1942.11991226. 10.1080/00029890.1942.11991226Search in Google Scholar

[2] L. Weiner, Concerning a theorem of McCoy: Solution to problem 4419, Amer. Math. Monthly 59 (1952), no. 5, 336–337, DOI: https://doi.org/10.2307/2307526. 10.2307/2307526Search in Google Scholar

[3] M. B. Rege and S. Chhawchharia, Armendariz rings, Proc. Japan Acad. Ser. A Math. Sci. 73 (1997), no. 1, 14–17, DOI: https://doi.org/10.3792/pjaa.73.14. 10.3792/pjaa.73.14Search in Google Scholar

[4] P. P. Nielsen, Semi-commutativity and the McCoy condition, J. Algebra 298 (2006), no. 1, 134–141, DOI: https://doi.org/10.1016/j.jalgebra.2005.10.008. 10.1016/j.jalgebra.2005.10.008Search in Google Scholar

[5] V. Camillo and P. P. Nielsen, McCoy rings and zero-divisors, J. Pure Appl. Algebra 212 (2008), no. 3, 599–615, DOI: https://doi.org/10.1016/j.jpaa.2007.06.010. 10.1016/j.jpaa.2007.06.010Search in Google Scholar

[6] J. Baeck, The rings where zero-divisor polynomials have zero-divisor coefficients, Rocky Mountain J. Math. 51 (2021), no. 3, 771–785, DOI: http://doi.org/10.1216/rmj.2021.51.771. 10.1216/rmj.2021.51.771Search in Google Scholar

[7] J. Cui and J. Chen, On McCoy modules, Bull. Korean Math. Soc. 48 (2011), no. 1, 23–33, DOI: https://doi.org/10.4134/BKMS.2011.48.1.023. 10.4134/BKMS.2011.48.1.023Search in Google Scholar

[8] T. K. Lee and Y. Zhou, Reduced modules, in: Rings, Modules, Algebras, and Abelian Groups, Lecture Notes in Pure Appllied Mathematics, Vol. 236, Dekker, New York, 2004, pp. 365–377. 10.1201/9780824750817.ch29Search in Google Scholar

[9] J. Lambek, On the representations of modules by sheaves of factor modules, Canad. Math. Bull. 14 (1971), no. 3, 359–368, DOI: https://doi.org/10.4153/CMB-1971-065-1. 10.4153/CMB-1971-065-1Search in Google Scholar

[10] A. M. Buhphang and M. B. Rege, Semi-commutative modules and Armendariz modules, Arab. J. Math. Sci. 8 (2002), no. 1, 53–65. Search in Google Scholar

[11] A. Ç. Özcan, A. Harmanci, and P. F. Smith, Duo modules, Glasg. Math. J. 48 (2006), no. 3, 533–545, DOI: https://doi.org/10.1017/S0017089506003260. 10.1017/S0017089506003260Search in Google Scholar

[12] D. D. Anderson and S. Chun, McCoy modules and related modules over commutative rings, Comm. Algebra 45 (2017), no. 6, 2593–2601, DOI: https://doi.org/10.1080/00927872.2016.1233218. 10.1080/00927872.2016.1233218Search in Google Scholar

[13] J. Baeck, N. K. Kim, Y. Lee, and P. P. Nielsen, Zero-divisor placement, a condition of Camillo, and the McCoy property, J. Pure Appl. Algebra 224 (2020), no. 12, 106432, DOI: https://doi.org/10.1016/j.jpaa.2020.106432. 10.1016/j.jpaa.2020.106432Search in Google Scholar

[14] V. Camillo, T. K. Kwak, and Y. Lee, On a generalization of McCoy rings, J. Korean Math. Soc. 50 (2013), no. 5, 959–972, DOI: https://doi.org/10.4134/JKMS.2013.50.5.959. 10.4134/JKMS.2013.50.5.959Search in Google Scholar

[15] C. Y. Hong, Y. C. Jeon, N. K. Kim, and Y. Lee, The McCoy condition on noncommutative rings, Comm. Algebra 39 (2011), no. 5, 1809–1825, DOI: https://doi.org/10.1080/00927872.2010.480952. 10.1080/00927872.2010.480952Search in Google Scholar

[16] M. T. Koşan, Extensions of rings having McCoy condition, Canad. Math. Bull. 52 (2009), no. 2, 267–272, DOI: https://doi.org/10.4153/CMB-2009-029-5. 10.4153/CMB-2009-029-5Search in Google Scholar

[17] Z. Lei, J. Chen, and Z. Ying, A question on McCoy rings, Bull. Aust. Math. Soc. 76 (2007), no. 1, 137–141, DOI: https://doi.org/10.1017/S0004972700039526. 10.1017/S0004972700039526Search in Google Scholar

[18] R. Mazurek and M. Ziembowski, On right McCoy rings and right McCoy rings relative to u.p.-monoids, Commun. Contemp. Math. 17 (2015), no. 4, 15500496, DOI: https://doi.org/10.1142/S0219199715500492. 10.1142/S0219199715500492Search in Google Scholar

[19] M. Zahiri, A. Moussavi, and R. Mohammadi, On a skew McCoy ring, Comm. Algebra 47 (2019), no. 10, 4061–4065, DOI: https://doi.org/10.1080/00927872.2019.1576188. 10.1080/00927872.2019.1576188Search in Google Scholar

[20] R. Zhao and Z. Liu, Extensions of McCoy rings, Algebra Colloq. 16 (2009), no. 3, 495–502, DOI: https://doi.org/10.1142/S1005386709000467. 10.1142/S1005386709000467Search in Google Scholar

[21] N. J. Groenewald and D. Ssevviiri, Classical completely prime submodules, Hacet. J. Math. Stat. 45 (2016), no. 3, 717–729, DOI: http://doi.org/10.15672/HJMS.20164513096. 10.15672/HJMS.20164513096Search in Google Scholar

[22] N. Agayev, G. Güngöroğlu, A. Harmanci, and S. Halicioğlu, Abelian modules, Acta Math. Univ. Comenian. 78 (2009), no. 2, 235–244. Search in Google Scholar

[23] M. B. Rege and A. M. Buhphang, On reduced modules and rings, Int. Electron. J. Algebra 3 (2008), 58–74. Search in Google Scholar

[24] R. Mazurek and M. Ziembowski, Right Gaussian rings and skew power series rings, J. Algebra 330 (2011), no. 1, 130–146, DOI: https://doi.org/10.1016/j.jalgebra.2010.11.014. 10.1016/j.jalgebra.2010.11.014Search in Google Scholar

[25] D. D. Anderson and V. Camillo, Armendariz rings and Gaussian rings, Comm. Algebra 26 (1998), no. 7, 2265–2272, DOI: https://doi.org/10.1080/00927879808826274. 10.1080/00927879808826274Search in Google Scholar

[26] W. Narkiewicz, Polynomial Mappings, Springer, Berlin, 1995. 10.1007/BFb0076894Search in Google Scholar

[27] G. M. Bergman, The diamond lemma for ring theory, Adv. Math. 29 (1978), no. 2, 178–218, DOI: https://doi.org/10.1016/0001-8708(78)90010-5. 10.1016/0001-8708(78)90010-5Search in Google Scholar

[28] J. Castro Pérez, M. Medina Bárcenas, J. RíosMontes, and A. Zaldívar Corichi, On semiprime Goldie modules, Comm. Algebra 44 (2016), no. 11, 4749–4768, DOI: https://doi.org/10.1080/00927872.2015.1113290. 10.1080/00927872.2015.1113290Search in Google Scholar

[29] T. Y. Lam, Lectures on modules and rings, Graduate Texts in Mathematics vol. 189, Springer-Verlag, New York, 1999. 10.1007/978-1-4612-0525-8Search in Google Scholar

[30] T. W. Hungerford, Algebra, Springer, New York, 1974. 10.1007/978-1-4612-6101-8_4Search in Google Scholar

[31] K. R. Goodearl, Von Neumann Regular Rings, Pitman, London-San Francisco-Melbourne, 1979. Search in Google Scholar

[32] N. K. Kim, T. K. Kwak, and Y. Lee, On a generalization of right duo rings, Bull. Korean Math. Soc. 53 (2016), no. 3, 925–942, DOI: https://doi.org/10.4134/BKMS.b150441. 10.4134/BKMS.b150441Search in Google Scholar

[33] X. Yao, Weakly right duo rings, Pure Appl. Math. Sci. 21 (1985), 19–24. Search in Google Scholar

[34] D. D. Anderson and T. Dumitrescu, S-Noetherian rings, Comm. Algebra 30 (2002), no. 9, 4407–4416, DOI: http://doi.org/10.1081/AGB-120013328. 10.1081/AGB-120013328Search in Google Scholar

[35] J. Baeck, On S-principal right ideal rings, AIMS Math. 7 (2022), no. 7, 12106–12122, http://doi.org/10.3934/math.2022673. Search in Google Scholar

[36] C. Faith, Rings and things and a fine array of twentieth century associative algebra, Mathematical Surveys and Monographs, Vol. 65, American Mathematical Society, Providence, 1999. 10.1090/surv/065Search in Google Scholar

[37] G. Marks, R. Mazurek, and M. Ziembowski, A unified approach to various generalizations of Armendariz rings, Bull. Aust. Math. Soc. 81 (2010), no. 3, 361–397, DOI: https://doi.org/10.1017/S0004972709001178. 10.1017/S0004972709001178Search in Google Scholar

[38] K. R. Goodearl and R. B. Warfield Jr., An introduction to noncommutative Noetherian rings, London Mathematical Society Student Texts, Vol. 61, 2nd ed., Cambridge University Press, Cambridge, 2004. 10.1017/CBO9780511841699Search in Google Scholar

[39] R. Fraïssé, Theory of relations, Stud. Logic Found. Math, Vol. 145, North-Holland, Amsterdam, 2000. Search in Google Scholar

Received: 2022-07-04
Revised: 2022-11-25
Accepted: 2022-12-08
Published Online: 2022-12-31

© 2022 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. A random von Neumann theorem for uniformly distributed sequences of partitions
  3. Note on structural properties of graphs
  4. Mean-field formulation for mean-variance asset-liability management with cash flow under an uncertain exit time
  5. The family of random attractors for nonautonomous stochastic higher-order Kirchhoff equations with variable coefficients
  6. The intersection graph of graded submodules of a graded module
  7. Isoperimetric and Brunn-Minkowski inequalities for the (p, q)-mixed geominimal surface areas
  8. On second-order fuzzy discrete population model
  9. On certain functional equation in prime rings
  10. General complex Lp projection bodies and complex Lp mixed projection bodies
  11. Some results on the total proper k-connection number
  12. The stability with general decay rate of hybrid stochastic fractional differential equations driven by Lévy noise with impulsive effects
  13. Well posedness of magnetohydrodynamic equations in 3D mixed-norm Lebesgue space
  14. Strong convergence of a self-adaptive inertial Tseng's extragradient method for pseudomonotone variational inequalities and fixed point problems
  15. Generic uniqueness of saddle point for two-person zero-sum differential games
  16. Relational representations of algebraic lattices and their applications
  17. Explicit construction of mock modular forms from weakly holomorphic Hecke eigenforms
  18. The equivalent condition of G-asymptotic tracking property and G-Lipschitz tracking property
  19. Arithmetic convolution sums derived from eta quotients related to divisors of 6
  20. Dynamical behaviors of a k-order fuzzy difference equation
  21. The transfer ideal under the action of orthogonal group in modular case
  22. The multinomial convolution sum of a generalized divisor function
  23. Extensions of Gronwall-Bellman type integral inequalities with two independent variables
  24. Unicity of meromorphic functions concerning differences and small functions
  25. Solutions to problems about potentially Ks,t-bigraphic pair
  26. Monotonicity of solutions for fractional p-equations with a gradient term
  27. Data smoothing with applications to edge detection
  28. An ℋ-tensor-based criteria for testing the positive definiteness of multivariate homogeneous forms
  29. Characterizations of *-antiderivable mappings on operator algebras
  30. Initial-boundary value problem of fifth-order Korteweg-de Vries equation posed on half line with nonlinear boundary values
  31. On a more accurate half-discrete Hilbert-type inequality involving hyperbolic functions
  32. On split twisted inner derivation triple systems with no restrictions on their 0-root spaces
  33. Geometry of conformal η-Ricci solitons and conformal η-Ricci almost solitons on paracontact geometry
  34. Bifurcation and chaos in a discrete predator-prey system of Leslie type with Michaelis-Menten prey harvesting
  35. A posteriori error estimates of characteristic mixed finite elements for convection-diffusion control problems
  36. Dynamical analysis of a Lotka Volterra commensalism model with additive Allee effect
  37. An efficient finite element method based on dimension reduction scheme for a fourth-order Steklov eigenvalue problem
  38. Connectivity with respect to α-discrete closure operators
  39. Khasminskii-type theorem for a class of stochastic functional differential equations
  40. On some new Hermite-Hadamard and Ostrowski type inequalities for s-convex functions in (p, q)-calculus with applications
  41. New properties for the Ramanujan R-function
  42. Shooting method in the application of boundary value problems for differential equations with sign-changing weight function
  43. Ground state solution for some new Kirchhoff-type equations with Hartree-type nonlinearities and critical or supercritical growth
  44. Existence and uniqueness of solutions for the stochastic Volterra-Levin equation with variable delays
  45. Ambrosetti-Prodi-type results for a class of difference equations with nonlinearities indefinite in sign
  46. Research of cooperation strategy of government-enterprise digital transformation based on differential game
  47. Malmquist-type theorems on some complex differential-difference equations
  48. Disjoint diskcyclicity of weighted shifts
  49. Construction of special soliton solutions to the stochastic Riccati equation
  50. Remarks on the generalized interpolative contractions and some fixed-point theorems with application
  51. Analysis of a deteriorating system with delayed repair and unreliable repair equipment
  52. On the critical fractional Schrödinger-Kirchhoff-Poisson equations with electromagnetic fields
  53. The exact solutions of generalized Davey-Stewartson equations with arbitrary power nonlinearities using the dynamical system and the first integral methods
  54. Regularity of models associated with Markov jump processes
  55. Multiplicity solutions for a class of p-Laplacian fractional differential equations via variational methods
  56. Minimal period problem for second-order Hamiltonian systems with asymptotically linear nonlinearities
  57. Convergence rate of the modified Levenberg-Marquardt method under Hölderian local error bound
  58. Non-binary quantum codes from constacyclic codes over 𝔽q[u1, u2,…,uk]/⟨ui3 = ui, uiuj = ujui
  59. On the general position number of two classes of graphs
  60. A posteriori regularization method for the two-dimensional inverse heat conduction problem
  61. Orbital stability and Zhukovskiǐ quasi-stability in impulsive dynamical systems
  62. Approximations related to the complete p-elliptic integrals
  63. A note on commutators of strongly singular Calderón-Zygmund operators
  64. Generalized Munn rings
  65. Double domination in maximal outerplanar graphs
  66. Existence and uniqueness of solutions to the norm minimum problem on digraphs
  67. On the p-integrable trajectories of the nonlinear control system described by the Urysohn-type integral equation
  68. Robust estimation for varying coefficient partially functional linear regression models based on exponential squared loss function
  69. Hessian equations of Krylov type on compact Hermitian manifolds
  70. Class fields generated by coordinates of elliptic curves
  71. The lattice of (2, 1)-congruences on a left restriction semigroup
  72. A numerical solution of problem for essentially loaded differential equations with an integro-multipoint condition
  73. On stochastic accelerated gradient with convergence rate
  74. Displacement structure of the DMP inverse
  75. Dependence of eigenvalues of Sturm-Liouville problems on time scales with eigenparameter-dependent boundary conditions
  76. Existence of positive solutions of discrete third-order three-point BVP with sign-changing Green's function
  77. Some new fixed point theorems for nonexpansive-type mappings in geodesic spaces
  78. Generalized 4-connectivity of hierarchical star networks
  79. Spectra and reticulation of semihoops
  80. Stein-Weiss inequality for local mixed radial-angular Morrey spaces
  81. Eigenvalues of transition weight matrix for a family of weighted networks
  82. A modified Tikhonov regularization for unknown source in space fractional diffusion equation
  83. Modular forms of half-integral weight on Γ0(4) with few nonvanishing coefficients modulo
  84. Some estimates for commutators of bilinear pseudo-differential operators
  85. Extension of isometries in real Hilbert spaces
  86. Existence of positive periodic solutions for first-order nonlinear differential equations with multiple time-varying delays
  87. B-Fredholm elements in primitive C*-algebras
  88. Unique solvability for an inverse problem of a nonlinear parabolic PDE with nonlocal integral overdetermination condition
  89. An algebraic semigroup method for discovering maximal frequent itemsets
  90. Class-preserving Coleman automorphisms of some classes of finite groups
  91. Exponential stability of traveling waves for a nonlocal dispersal SIR model with delay
  92. Existence and multiplicity of solutions for second-order Dirichlet problems with nonlinear impulses
  93. The transitivity of primary conjugacy in regular ω-semigroups
  94. Stability estimation of some Markov controlled processes
  95. On nonnil-coherent modules and nonnil-Noetherian modules
  96. N-Tuples of weighted noncommutative Orlicz space and some geometrical properties
  97. The dimension-free estimate for the truncated maximal operator
  98. A human error risk priority number calculation methodology using fuzzy and TOPSIS grey
  99. Compact mappings and s-mappings at subsets
  100. The structural properties of the Gompertz-two-parameter-Lindley distribution and associated inference
  101. A monotone iteration for a nonlinear Euler-Bernoulli beam equation with indefinite weight and Neumann boundary conditions
  102. Delta waves of the isentropic relativistic Euler system coupled with an advection equation for Chaplygin gas
  103. Multiplicity and minimality of periodic solutions to fourth-order super-quadratic difference systems
  104. On the reciprocal sum of the fourth power of Fibonacci numbers
  105. Averaging principle for two-time-scale stochastic differential equations with correlated noise
  106. Phragmén-Lindelöf alternative results and structural stability for Brinkman fluid in porous media in a semi-infinite cylinder
  107. Study on r-truncated degenerate Stirling numbers of the second kind
  108. On 7-valent symmetric graphs of order 2pq and 11-valent symmetric graphs of order 4pq
  109. Some new characterizations of finite p-nilpotent groups
  110. A Billingsley type theorem for Bowen topological entropy of nonautonomous dynamical systems
  111. F4 and PSp (8, ℂ)-Higgs pairs understood as fixed points of the moduli space of E6-Higgs bundles over a compact Riemann surface
  112. On modules related to McCoy modules
  113. On generalized extragradient implicit method for systems of variational inequalities with constraints of variational inclusion and fixed point problems
  114. Solvability for a nonlocal dispersal model governed by time and space integrals
  115. Finite groups whose maximal subgroups of even order are MSN-groups
  116. Symmetric results of a Hénon-type elliptic system with coupled linear part
  117. On the connection between Sp-almost periodic functions defined on time scales and ℝ
  118. On a class of Harada rings
  119. On regular subgroup functors of finite groups
  120. Fast iterative solutions of Riccati and Lyapunov equations
  121. Weak measure expansivity of C2 dynamics
  122. Admissible congruences on type B semigroups
  123. Generalized fractional Hermite-Hadamard type inclusions for co-ordinated convex interval-valued functions
  124. Inverse eigenvalue problems for rank one perturbations of the Sturm-Liouville operator
  125. Data transmission mechanism of vehicle networking based on fuzzy comprehensive evaluation
  126. Dual uniformities in function spaces over uniform continuity
  127. Review Article
  128. On Hahn-Banach theorem and some of its applications
  129. Rapid Communication
  130. Discussion of foundation of mathematics and quantum theory
  131. Special Issue on Boundary Value Problems and their Applications on Biosciences and Engineering (Part II)
  132. A study of minimax shrinkage estimators dominating the James-Stein estimator under the balanced loss function
  133. Representations by degenerate Daehee polynomials
  134. Multilevel MC method for weak approximation of stochastic differential equation with the exact coupling scheme
  135. Multiple periodic solutions for discrete boundary value problem involving the mean curvature operator
  136. Special Issue on Evolution Equations, Theory and Applications (Part II)
  137. Coupled measure of noncompactness and functional integral equations
  138. Existence results for neutral evolution equations with nonlocal conditions and delay via fractional operator
  139. Global weak solution of 3D-NSE with exponential damping
  140. Special Issue on Fractional Problems with Variable-Order or Variable Exponents (Part I)
  141. Ground state solutions of nonlinear Schrödinger equations involving the fractional p-Laplacian and potential wells
  142. A class of p1(x, ⋅) & p2(x, ⋅)-fractional Kirchhoff-type problem with variable s(x, ⋅)-order and without the Ambrosetti-Rabinowitz condition in ℝN
  143. Jensen-type inequalities for m-convex functions
  144. Special Issue on Problems, Methods and Applications of Nonlinear Analysis (Part III)
  145. The influence of the noise on the exact solutions of a Kuramoto-Sivashinsky equation
  146. Basic inequalities for statistical submanifolds in Golden-like statistical manifolds
  147. Global existence and blow up of the solution for nonlinear Klein-Gordon equation with variable coefficient nonlinear source term
  148. Hopf bifurcation and Turing instability in a diffusive predator-prey model with hunting cooperation
  149. Efficient fixed-point iteration for generalized nonexpansive mappings and its stability in Banach spaces
Downloaded on 7.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/math-2022-0545/html
Scroll to top button