1 Introduction

Throughout this paper, R is an associative ring with identity, and all the modules are unital right R-modules. Let M be an R-module. By \({\text {End}}_R(M)\) and E(M), we denote the endomorphism ring of M and the injective hull of M, respectively. The notation \(N \subseteq M\) means that N is a subset of M and \(N \le M\) means that N is a submodule of M. By \(\mathbb {Q}\) and \(\mathbb {Z}\), we denote the ring of rational numbers and integer numbers, respectively. For any prime number p, \(\mathbb {Z}(p^{\infty })\) denotes the Prüfer p-group.

A submodule K of M is called fully invariant if \(f(K)\subseteq K\) for all \(f \in {\text {End}}_R(M)\). In 1967, Clark introduced the concept of quasi-Baer rings. A ring R is called right quasi-Baer if the right annihilator of any right ideal of R is generated as a right ideal by an idempotent. Rizvi and Roman generalized the notion of right quasi-Baer rings to a module theoretic version. A module M is called quasi-Baer if the right annihilator in M of any two-sided ideal of \(End_R(M)\) is a direct summand of M. This notion was dualized by Amouzegar–Talebi in 2013 [1]. A module M is said to be quasi-dual Baer if for every fully invariant submodule N of M, there exists an idempotent e in S such that \(\{\phi \in S \mid Im\phi \subseteq N\} = eS\). In Sect. 2, we continue the study of quasi-dual Baer modules. We show that an R-module M is quasi-dual Baer if and only if \(\sum _{\varphi \in \mathfrak {I}} \varphi (M)\) is a direct summand of M for every left ideal \(\mathfrak {I}\) of \({\text {End}}_R(M)\). We provide examples to separate the properties of quasi-dual Baer modules and dual Baer modules. We prove that for a ring R, the R-module \(R_R\) is quasi-dual Baer if and only if R is a finite product of simple rings. We obtain conditions under which an indecomposable module is quasi-dual Baer. It is also shown that if M is a quasi-dual Baer module over a commutative ring R, then \(M\mathfrak {a}\) is a direct summand of M for any ideal \(\mathfrak {a}\) of R. We give a characterization of when a direct sum of a module \(M_1\) with \({\text {Rad}}(M_1)=M_1\) and a semisimple module \(M_2\) is quasi-dual Baer. It is also shown that over a commutative perfect ring, every quasi-dual Baer module is semisimple.

2 Some results on quasi-dual Baer modules

Let M be an R-module with \(S={\text {End}}_R(M)\). For a submodule N of M, the right ideal \(\{\phi \in S \mid Im\phi \subseteq N\}\) of S will be denoted by \(\mathfrak {D}(N)\). For a subset \(\mathfrak {X}\) of S and a submodule N of M, let \(\mathfrak {X}(N)\) denote the submodule \(\sum _{\varphi \in \mathfrak {X}} \varphi (N)\) of M. It is easily seen that if N is a fully invariant submodule of M, then \(\mathfrak {D}(N)\) is a two-sided ideal of S and if \(\mathfrak {I}\) is a two-sided ideal of S, then \(\mathfrak {I}(M)\) is a fully invariant submodule of M. Therefore, we have the following two maps:

\(\mathfrak {D}:\{\)fully invariant submodules of \(M\} \longrightarrow \{\)two-sided ideals of S} and

\(\mathfrak {T}:\{\)two-sided ideals of \(S\} \longrightarrow \{\)fully invariant submodules of M}

with \(\mathfrak {T}(\mathfrak {I})=\mathfrak {I}(M)\) for any two-sided ideal \(\mathfrak {I}\) of S.

Following [1, 4], the module M is called dual Baer (quasi-dual Baer) if for every submodule (fully invariant submodule) N of M, there exists an idempotent e in S such that \(\mathfrak {D}(N)=eS\). In this section, we continue the study of quasi-dual Baer modules. We begin with the following characterization of this class of modules.

Proposition 2.1

Let M be an R-module with \(S={\text {End}}_R(M)\). Then, the following statements are equivalent:

  1. (i)

    M is a quasi-dual Baer module;

  2. (ii)

    For any fully invariant submodule N of M, there exists a decomposition \(M=D_1 \oplus D_2\) with \(D_1 \subseteq N\) and \({\text {Hom}}_R(M, N \cap D_2)=0\).

Proof

(i) \(\Rightarrow \) (ii) Let N be a fully invariant submodule of M. Then \(\mathfrak {D}(N)=eS\) for some idempotent e of S. Set \(D_1=e(M)\) and \(D_2=(1-e)(M)\). Therefore, \(\mathfrak {T}(\mathfrak {D}(N))=D_1 \subseteq N\). Since \(M=D_1 \oplus D_2\), we have \(N=D_1 \oplus (N \cap D_2)\). Now, let h be an endomorphism of M such that \(h(M) \subseteq N \cap D_2\). Hence, \(h \in \mathfrak {D}(N)\) and so \(h=eg\) for some \(g \in S\). Thus, \(h(M) \subseteq e(M)\). But \(h(M) \subseteq (1-e)(M)\). Then, \(h=0\). It follows that \({\text {Hom}}_R(M, N \cap D_2)=0\).

(ii) \(\Rightarrow \) (i) Let N be a fully invariant submodule of M. So there exists a decomposition \(M=D_1 \oplus D_2\) with \(D_1 \subseteq N\) and \({\text {Hom}}_R(M, N \cap D_2)=0\). Let e be an idempotent of S with \(D_1=e(M)\). Let us prove that \(\mathfrak {D}(N)=eS\). It is clear that \(eS \subseteq \mathfrak {D}(N)\). Now, let \(f \in \mathfrak {D}(N)\). Let \(\pi :N=D_1 \oplus (N \cap D_2) \rightarrow N \cap D_2\) be the projection map. Then, \(\pi f \in Hom_R(M, N \cap D_2)=0\). This implies that \(f(M) \subseteq e(M)\). Thus, \((1-e)f=0\). Since \(f=ef+(1-e)f\), we obtain \(f=ef \in eS\). Therefore, \(\mathfrak {D}(N)=eS\) and so M is quasi-dual Baer. \(\square \)

The following corollary follows directly from Proposition 2.1.

Corollary 2.2

Let M be an R-module. Assume that every fully invariant submodule of M is a direct summand of M. Then, M is quasi-dual Baer.

Example 2.3

  1. (i)

    Clearly, every dual Baer module is quasi-dual Baer. From [9, Proposition 4.16], it follows that every torsion-free injective module over a commutative domain is quasi-dual Baer.

  2. (ii)

    Let M be an R-module such that \({\text {Hom}}_R(M, N)=0\) for every fully invariant proper submodule N of M. Then, clearly M is quasi-dual Baer (see Proposition 2.1). For example, the \(\mathbb {Z}\)-modules \(\mathbb {Q}\) and \(\mathbb {Z}(p^{\infty })\) (where p is a prime number) are quasi-dual Baer modules.

  3. (iii)

    Recall that a nonzero ring R is called a simple ring if (0) and R are the only two-sided ideals in R. Let M be a module and assume that M has no nonzero proper fully invariant submodules. Then, M is quasi-dual Baer. For example, for any simple ring S, the right S-module \(S_S\) is quasi-dual Baer.

Next, we present another characterization of quasi-dual Baer modules.

Proposition 2.4

Let M be an R-module with \(S={\text {End}}_R(M)\). Then, the following are equivalent:

  1. (i)

    M is a quasi-dual Baer R-module;

  2. (ii)

    For every left ideal \(\mathfrak {I}\) of S, \(\mathfrak {I}(M)\) is a direct summand of M;

  3. (iii)

    For every two-sided ideal \(\mathfrak {I}\) of S, \(\mathfrak {I}(M)\) is a direct summand of M.

Proof

(i) \(\Rightarrow \) (ii) Let \(\mathfrak {I}\) be a left ideal of S. Put \(N=\mathfrak {I}(M)\). Clearly, N is a fully invariant submodule of M. By hypothesis, \(\mathfrak {D}(N)=eS\) for some idempotent \(e \in S\). Therefore, \(e(M) \subseteq N\). Now, let \(\varphi \in \mathfrak {I}\). Then \(Im \varphi \subseteq \mathfrak {I}(M)\). Hence, \(\varphi \in \mathfrak {D}(N)\). This implies that \(\varphi = e\psi \) for some \(\psi \in S\). Thus \(\varphi (M) \subseteq e(M)\). It follows that \(N=\mathfrak {I}(M) \subseteq e(M)\). Consequently, \(N=e(M)\) is a direct summand of M.

(ii) \(\Rightarrow \) (iii) This is obvious.

(iii) \(\Rightarrow \) (i) Let L be a fully invariant submodule of M. Note that \(\mathfrak {I}= \mathfrak {D}(L)\) is a two-sided ideal of S. By (iii), \(\mathfrak {I}(M)=e(M)\) for some \(e^2=e \in S\). Let \(\phi \in \mathfrak {I}\). Then, \(Im \phi \subseteq e(M)\). Hence, \((1_S-e)\phi = 0\). Since \(1_S=e+(1_S-e)\), we have \(\phi = e\phi \in eS\). Therefore, \(\mathfrak {I} \subseteq eS\). Now let \(\psi \in S\). Then, \(e\psi (M) \subseteq e(M) \subseteq \mathfrak {I}(M) \subseteq L\). So \(e\psi \in \mathfrak {I}\). This yields \(eS \subseteq \mathfrak {I}\). It follows that \(\mathfrak {D}(L)=eS\). This completes the proof. \(\square \)

As an application of Proposition 2.4, we provide another way of proving [1, Theorem 2.1].

Corollary 2.5

Let M be a quasi-dual Baer module with \(S={\text {End}}_R(M)\). Then every direct summand of M is quasi-dual Baer.

Proof

Let N be a direct summand of M. Let \(\pi :M \rightarrow N\) be the projection map and \(\mu :N \rightarrow M\) be the inclusion map. For every \(f \in T={\text {End}}_R(N)\), let \(\overline{f}=\mu f \pi \in S\). Let I be a left ideal of T. Set \(\overline{I}=\{\overline{f} \mid f \in I\}\) and \(\mathfrak {I}=S\overline{I}\). By Proposition 2.4, \(\mathfrak {I}(M)=\mathfrak {I}(N)\) is a direct summand of M. Let \(e=\mu \pi \in S\). Then, \(e=e^2\) and \(N=e(M)\). Since \(\mathfrak {I}(M)\) is a fully invariant submodule of M, we have \(\mathfrak {I}(M)=e(\mathfrak {I}(M)) \oplus (1-e)(\mathfrak {I}(M))\). Therefore \(e(\mathfrak {I}(M)) = e(\mathfrak {I}(N))\) is a direct summand of N. Since \(e^2=e\), we have \(e(x)=x\) for any \(x \in N\). Thus \(\overline{I}(N) = e(\overline{I}(N)) \subseteq e(S\overline{I}(N))\). Now let \(g \in S\) and \(f \in I\). Then, \(eg\overline{f}(N) =\mu \pi g\mu f \pi (N) = \mu (\pi g\mu f) \pi (N)\). Since I is a left ideal of T, it follows that \(\pi g\mu f \in I\). This yields \(eg\overline{f}(N) = \overline{\pi g\mu f}(N) \subseteq \overline{I}(N)\). Hence, \(e(S\overline{I}(N)) \subseteq \overline{I}(N)\). Thus, \(\overline{I}(N) = e(\mathfrak {I}(N))\). Moreover, it is easily seen that \(I(N) = \overline{I}(N)\). It follows that \(I(N) = e(\mathfrak {I}(N))\) is a direct summand of N. Therefore, N is a quasi-dual Baer module by Proposition 2.4. \(\square \)

Corollary 2.6

Let M be a quasi-dual Baer module and let N be a fully invariant submodule of M. Then, the following are equivalent:

  1. (i)

    N is a direct summand of M;

  2. (ii)

    \(\mathfrak {I}(M) = N\), where \(\mathfrak {I} = \mathfrak {D}(N)\).

Proof

(i) \(\Rightarrow \) (ii) Let \(\pi :M \rightarrow N\) be the projection map and \(\mu :N \rightarrow M\) be the inclusion map. Then \(\mu \pi \in \mathfrak {D}(N)\) and \(\mu \pi (M)=N\). Hence \(\mathfrak {I}(M) = N\).

(ii) \(\Rightarrow \) (i) This follows directly from Proposition 2.4. \(\square \)

Next, we present more examples of quasi-dual Baer modules.

Example 2.7

  1. (i)

    By Proposition 2.4, the notion of being quasi-dual Baer only involves two-sided ideals of the endomorphism ring. It follows that any R-module M for which \({\text {End}}_R(M)\) is a simple ring is quasi-dual Baer.

  2. (ii)

    To construct an explicit example, let S be a simple ring and let \(_SN_S\) be an S-S-bimodule. As in [7], consider the generalized matrix ring \(R = \left( \begin{array}{rr} S &{} N \\ N&{} S \\ \end{array} \right) \) with multiplication given by

    $$\left( \begin{array}{rr} a &{} x \\ y&{} b \\ \end{array} \right) \left( \begin{array}{rr} a^{\prime } \quad &{} x^{\prime } \\ y^{\prime }\quad &{} b^{\prime } \\ \end{array} \right) = \left( \begin{array}{cc} aa^{\prime }\quad &{} ax^{\prime }+xb^{\prime } \\ ya^{\prime }+by^{\prime }\quad &{} bb^{\prime } \\ \end{array} \right) $$

    for all \(a, a^{\prime }, b, b^{\prime } \in S\) and \(x, x^{\prime }, y, y^{\prime } \in N\). Then, \(M=N \oplus S\) is a right R-module by the following action:

    $$(n,s)\left( \begin{array}{rr} a &{} x \\ y&{} b \\ \end{array} \right) =(na+sy, sb).$$

    It is shown in [7, p. 1278] that the rings \({\text {End}}_R(M)\) and S are isomorphic. Therefore M is a quasi-dual Baer R-module.

Next, we compare quasi-dual Baer modules and dual Baer modules. Recall that a module M is called a duo module provided every submodule of M is fully invariant. A ring R is called a right duo ring if the right R-module \(R_R\) is a duo module; equivalently, every right ideal of R is a two-sided ideal.

Remark 2.8

Let M be an R-module such that \(S={\text {End}}_R(M)\) is a right duo ring (e.g., S is a commutative ring). From Proposition 2.4 and [4, Theorem 2.1], it follows that M is a quasi-dual Baer module if and only if M is a dual Baer module. In particular, if M is an indecomposable injective R-module over a commutative noetherian ring R, then M is quasi-dual Baer if and only if M is dual Baer (see [8, Proposition 5.13]).

In [1], there are no examples showing that the class of quasi-dual Baer modules contains properly that of dual Baer modules. In the next example, we provide some quasi-dual Baer modules which are not dual Baer.

Example 2.9

  1. (i)

    Let R be a ring. It is shown in [4, Corollary 2.9] that the R-module \(R_R\) is dual Baer if and only if R is a semisimple ring. It follows that for any simple ring R which is not semisimple, the R-module \(R_R\) is quasi-dual Baer but \(R_R\) is not dual Baer (see Example 2.3(iii)). For example, we can take the following rings:

  2. (a)

    The differential polynomial ring \(R=k[x; \delta ]\), where k is a simple ring of characteristic 0 and \(\delta \) is a non-inner derivation on k (see [5, Corollary 3.16]).

  3. (b)

    The Weyl algebra \(R=A_n(k_0)\), where \(k_0\) is a simple ring of characteristic 0 and n is a positive integer (see [5, Corollary 3.17]).

  4. (c)

    The ring of skew Laurent polynomials \(R=k[x, x^{-1}; \sigma ]\), where k is a field with an automorphism \(\sigma \) of infinite order (see [5, Example 1.8 and Corollary 3.19]).

  5. (d)

    See also other examples in [5, p. 40].

  6. (ii)

    Let S be a simple ring and let \(_SN_S\) be an S-S-bimodule. As in Example 2.7(ii), consider the ring \(R = \left( \begin{array}{rr} S &{} N \\ N&{} S \\ \end{array} \right) \) and the right R-module \(M=N \oplus S\). Assume that S is a domain that is not a division ring (we can take \(S=A_n(k)\), the n-th Weyl algebra over a field k of characteristic zero). Let \(t \in S\) be a nonzero non-invertible element. Consider the map \(f_t:M \rightarrow M\) defined by \(f_t((n, s))=(tn, ts)\) for all \((n, s) \in M\). Then \(f_t\) is an R-endomorphism of M. As S is a domain, we have \(tS \not = S\) (see, e.g., [5, Chapter 1, §1, p. 22, Exercise 1.4]). It follows that \(f_t\) is not an epimorphism. Note that the rings \({\text {End}}_R(M)\) and S are isomorphic (see [7, p. 1278]). Then \({\text {End}}_R(M)\) is a domain. This implies that M is an indecomposable R-module. Using [4, Corollary 2.2], we conclude that M is not a dual Baer R-module. However, M is a quasi-dual Baer R-module by Proposition 2.4.

  7. (iii)

    Let S be a simple Ore domain which is not a division ring. Consider the ring \(R=\left( \begin{array}{cc} S &{} K/S \\ O&{} S \\ \end{array} \right) \), where K is the classical right ring of quotients of S. Take the right R-module \(M=\left( \begin{array}{cc} S &{} K/S \\ 0&{} 0 \\ \end{array} \right) \). By similar arguments as in (ii), we prove that M is a quasi-dual Baer module which is not dual Baer. We include this proof for the convenience of the reader: By [6, Example 12], we have \({\text {End}}_R(M) \cong S\). So \({\text {End}}_R(M)\) is a simple domain. Hence M is an indecomposable R-module which is quasi-dual Baer.

On the other hand, S contains a nonzero non-invertible element t as S is not a division ring. Set \(r = \left( \begin{array}{rr} t &{} 0 \\ 0&{} 0 \\ \end{array} \right) \) and consider the R-endomorphism \(f_t\) of M defined by \(f_t(x)=rx\) for all \(x \in M\). It is easily seen that

$$f_t(M) = \left\{ \left( \begin{array}{rr} ts &{} \overline{tk} \\ 0&{} 0 \\ \end{array} \right) \mid (s, k) \in S \times K \right\} .$$

Since t is non-invertible in S and S is a domain, we have \(ts \not = 1\) for every \(s \in S\) (see [5, Chapter 1, §1, p. 22, Exercise 1.4]). Hence \(f_t\) is not an epimorphism. As M is indecomposable, M is not a dual Baer R-module by [4, Corollary 2.2].

It is clear that for a duo module M, M is quasi-dual Baer if and only if M is a dual Baer module. In particular, for a right duo ring R (e.g., R is a commutative ring), the R-module \(R_R\) is quasi-dual Baer if and only if \(R_R\) is a dual Baer R-module if and only if R is a semisimple ring by [4, Corollary 2.9]. Next, we characterize the class of rings R for which the right R-module \(R_R\) is quasi-dual Baer. A ring R is called a right quasi-dual Baer ring if the right R-module \(R_R\) is a quasi-dual Baer R-module. Left quasi-dual Baer rings are defined similarly.

Proposition 2.10

The following are equivalent for a ring R:

  1. (i)

    R is a right quasi-dual Baer ring;

  2. (ii)

    For every left ideal I of R, the right ideal IR is a direct summand of the right R-module \(R_R\);

  3. (iii)

    For every two-sided ideal I of R, the right ideal I is a direct summand of the right R-module \(R_R\);

  4. (iv)

    R is a finite product of simple rings.

Proof

Given \(a \in R\), let \(\varphi _a\) denote the R-endomorphism of the R-module \(R_R\) defined by \(\varphi _a(x)=ax\) for all \(x \in R\).

(i) \(\Rightarrow \) (ii) Let I be a left ideal of R. It is not hard to see that the set \(\mathfrak {I} = \{\varphi _a \mid a \in I\}\) is a left ideal of the ring \(S={\text {End}}_R(R_R)\). By Proposition 2.4, the right ideal \(\mathfrak {I}(R_R) = \sum _{a \in I} aR = IR\) is a direct summand of \(R_R\).

(ii) \(\Rightarrow \) (iii) is immediate.

(iii) \(\Rightarrow \) (iv) We claim that R is a semiprime ring. Suppose, contrary to our claim, that R contains a nonzero two-sided ideal I with \(I^2=0\). By assumption, there exists a right ideal K of R such that \(I \oplus K = R\). This implies that \(I^2 \oplus KI = I\). Hence \(KI=I \subseteq K\). Therefore, \(I=0\), a contradiction. Applying [3, Lemma 3.1], we deduce that every two-sided ideal of R is generated by a central idempotent. From [10, 3.5], it follows that R is a finite product of simple rings.

(iv) \(\Rightarrow \) (i) Let I be a fully invariant submodule of the R-module \(R_R\). Then, I is a two-sided ideal of R. By [10, 3.5], I is a direct summand of \(R_R\). The result follows from Corollary 2.2. \(\square \)

Applying Proposition 2.10, we obtain the following result which shows that the quasi-dual Baer property is left–right symmetric for any ring.

Corollary 2.11

A ring R is right quasi-dual Baer if and only if it is left quasi-dual Baer.

In the next proposition, we provide a characterization of indecomposable quasi-dual Baer modules over any ring R.

Proposition 2.12

Let M be an indecomposable R-module. Then, the following are equivalent:

  1. (i)

    M is quasi-dual Baer;

  2. (ii)

    For every fully invariant proper submodule N of M, \({\text {Hom}}_R(M, N)=0\).

Proof

(i) \(\Rightarrow \) (ii) This follows from Proposition 2.1.

(ii) \(\Rightarrow \) (i) This follows from the fact that \(\mathfrak {D}(M)={\text {End}}_R(M)\) and \(\mathfrak {D}(N)=0\) for every fully invariant proper submodule N of M. \(\square \)

Corollary 2.13

Let R be a local ring with maximal right ideal m. Assume that \(M=E(R/m)\) is a quasi-dual Baer module. Then R/m is injective (i.e., R is a division ring) or \({\text {Rad}}(M)=M\).

Proof

Put \(S=R/m\). Assume that S is an injective R-module. Then R is a right V-ring. This yields \(m=0\) and hence R is a division ring. Now assume that \(E(S) \not = S\). Suppose that M has a maximal submodule L. As R is a local ring, there exists an R-isomorphism \(\lambda :M/L \rightarrow S\). Let \(\pi :M \rightarrow M/L\) denote the natural epimorphism. Then, \(\lambda \pi \in {\text {Hom}}_R(M, S)\) and \(\lambda \pi \not = 0\). Therefore, \({\text {Hom}}_R(M, S) \not = 0\). From Proposition 2.12, it follows that S is not fully invariant in M, a contradiction. Consequently, \({\text {Rad}}(M)=M\). \(\square \)

A ring R is called right Bass if \({\text {Rad}}(M) \not = M\) for every nonzero right R-module M. It is well known that every right perfect ring is right Bass (see [2, Remark 28.5(3)]).

Corollary 2.14

Let R be a local ring with maximal right ideal m and \(S=R/m\). Assume that \({\text {Rad}}(E(S)) \not = E(S)\) (e.g., R is a local right perfect ring). Then, the following are equivalent:

  1. (i)

    E(S) is quasi-dual Baer;

  2. (ii)

    R is a division ring.

Proof

This follows from Corollary 2.13. \(\square \)

Next, we deal with some special cases of direct sums of quasi-dual Baer modules.

Proposition 2.15

Let an R-module \(M=M_1 \oplus M_2\) be a direct sum of submodules \(M_1\) and \(M_2\) such that \({\text {Rad}}(M_1)=M_1\) and \(M_2\) is semisimple. Let \(S={\text {End}}_R(M)\). Then, the following are equivalent:

  1. (i)

    M is quasi-dual Baer;

  2. (ii)

    \(M_1\) is quasi-dual Baer and \({\mathfrak {I}}(M_2) \cap M_1 \subseteq {\mathfrak {I}}(M_1)\) for every left ideal \(\mathfrak {I}\) of S;

  3. (iii)

    \(M_1\) is quasi-dual Baer and \({\mathfrak {I}}(M_2) \cap M_1 \subseteq {\mathfrak {I}}(M_1)\) for every two-sided ideal \(\mathfrak {I}\) of S.

Proof

  1. (i)

    \(\Rightarrow \) (ii) Clearly, \(M_1\) is a quasi-dual Baer module by Corollary 2.5. Now let \(\mathfrak {I}\) be a left ideal of S. Therefore, \(\mathfrak {I}(M)\) is a fully invariant submodule of M. Then \(\mathfrak {I}(M) = (\mathfrak {I}(M) \cap M_1) \oplus (\mathfrak {I}(M)\cap M_2)\). Note that \(\mathfrak {I}(M) = K_1 + K_2\), where \(K_1={\mathfrak {I}}(M_1)\) and \(K_2={\mathfrak {I}}(M_2)\). Moreover, \(M_1={\text {Rad}}(M)\) is a fully invariant submodule of M. Therefore, \(K_1 \subseteq M_1\). By modularity, we have \(\mathfrak {I}(M) \cap M_1 = K_1 + (K_2\cap M_1)\). This yields \(\mathfrak {I}(M) = (K_1 + (K_2\cap M_1)) \oplus (\mathfrak {I}(M)\cap M_2)\). Since \(K_2\cap M_1\) is semisimple, there exists a semisimple submodule N of \(K_2\cap M_1\) such that \(K_1 + (K_2\cap M_1) = K_1 \oplus N\). Therefore, \(\mathfrak {I}(M)=(K_1 \oplus N) \oplus (\mathfrak {I}(M)\cap M_2)\). As M is quasi-dual Baer, \(\mathfrak {I}(M)\) is a direct summand of M by Proposition 2.4. This implies that \(K_1 \oplus N\) is a direct summand of \(M_1\). Thus \({\text {Rad}}(N)=N\) as \({\text {Rad}}(M_1)=M_1\). But N is semisimple. Then \(N=0\). It follows that \(\mathfrak {I}(M)=K_1 \oplus (\mathfrak {I}(M)\cap M_2)\). Hence, \(K_1+(K_2\cap M_1) = K_1\). We thus get \(K_2 \cap M_1 \subseteq K_1\).

  2. (ii)

    \(\Rightarrow \) (iii) This is clear.

  3. (iii)

    \(\Rightarrow \) (i) Let \(\mathfrak {I}\) be a nonzero two-sided ideal of \(S={\text {End}}_R(M)\). Let \(\pi _1:M \rightarrow M_1\) be the projection map and \(\mu _1: M_1 \rightarrow M\) be the inclusion map. Consider the subset \(\mathfrak {I}_1=\{\pi _1f\mu _1 \mid f\in \mathfrak {I}\}\) of \(S_1={\text {End}}_R(M_1)\). It is easily seen that \(\mathfrak {I}_1\) is a left ideal of \(S_1\). Since \(M_1\) is quasi-dual Baer, \(N_1={\mathfrak {I}}(M_1)=\mathfrak {I}_1(M_1)\) is a direct summand of \(M_1\) by Proposition 2.4. Since \(\mathfrak {I}(M)\) is a fully invariant submodule of M, we have \(\mathfrak {I}(M) = (\mathfrak {I}(M) \cap M_1) \oplus (\mathfrak {I}(M) \cap M_2)\). As \(M_2\) is semisimple, there exists a semisimple submodule \(L \le {\mathfrak {I}}(M_2)\) such that \(\mathfrak {I}(M) = N_1 + {\mathfrak {I}}(M_2) = N_1 \oplus L\). By modularity, we have \(\mathfrak {I}(M) \cap M_1 = N_1 \oplus (L \cap M_1)\). By hypothesis, \(L \cap M_1 \subseteq N_1\). Thus \(L \cap M_1 = 0\). Therefore, \(\mathfrak {I}(M) \cap M_1 = N_1\) is a direct summand of \(M_1\). Since \(\mathfrak {I}(M) \cap M_2\) is a direct summand of \(M_2\), \(\mathfrak {I}(M)\) is a direct summand of M. Using again Proposition 2.4, it follows that M is quasi-dual Baer. \(\square \)

The next result provides a necessary condition for a direct sum of a module \(M_1\) with \({\text {Rad}}(M_1)=M_1\) and a semisimple module \(M_2\) to be quasi-dual Baer.

Corollary 2.16

Let an R-module \(M=M_1 \oplus M_2\) be a direct sum of submodules \(M_1\) and \(M_2\) such that \({\text {Rad}}(M_1)=M_1\) and \(M_2\) is semisimple. Assume that M is quasi-dual Baer. Then, \(Im \varphi \) is not fully invariant in \(M_1\) for all \(0 \not = \varphi \in Hom_R(M_2, M_1)\).

Proof

Suppose to the contrary that there exists a nonzero \(\varphi \in {\text {Hom}}_R(M_2, M_1)\) such that \(Im \varphi \) is fully invariant in \(M_1\). Consider the endomorphism \(\psi \) of M defined by \(\psi (x_1)=0\) for every \(x_1 \in M_1\) and \(\psi (x_2)=\varphi (x_2)\) for every \(x_2 \in M_2\). Let \(S={\text {End}}_R(M)\) and consider the left ideal \(\mathfrak {I}=S\psi \) of S. By Proposition 2.15, we have \({\mathfrak {I}}(M_2) \cap M_1 \subseteq {\mathfrak {I}}(M_1)\). Since \(\varphi (M_2)\) is fully invariant in \(M_1\) and \(M_1={\text {Rad}}(M)\) is fully invariant in M, it follows that \(\varphi (M_2)=\psi (M)\) is fully invariant in M. Note that \({\mathfrak {I}}(M_2) = \varphi (M_2)\) and \({\mathfrak {I}}(M_1) = 0\). So \(\varphi (M_2) \cap M_1 = 0\); that is \(\varphi (M_2) = 0\). This contradicts the choice of \(\varphi \). \(\square \)

Example 2.17

Let L be a simple R-module such that \({\text {Rad}}(E(L))=E(L)\) (e.g., we can take a simple module L over a Dedekind domain R). Consider the R-module \(M=E(L) \oplus L\) and let \(\mu :L \rightarrow E(L)\) be the inclusion map. Note that \(Im \mu \) is fully invariant in E(L). By Corollary 2.16, M is not quasi-dual Baer.

The next corollary is a direct consequence of Proposition 2.15.

Corollary 2.18

Let an R-module \(M=M_1 \oplus M_2\) be a direct sum of submodules \(M_1\) and \(M_2\) such that \({\text {Rad}}(M_1)=M_1\), \(M_2\) is semisimple and \({\text {Hom}}_R(M_2, M_1)=0\). Then, the following are equivalent:

  1. (i)

    M is quasi-dual Baer;

  2. (ii)

    \(M_1\) is quasi-dual Baer.

We conclude this section by providing conditions under which a direct sum of quasi-dual Baer modules is quasi-dual Baer.

Proposition 2.19

Let \(M=\oplus _{\lambda \in \Lambda } M_{\lambda }\) such that each \(M_{\lambda }\) \((\) \(\lambda \in \Lambda \) \()\) is a fully invariant submodule of M which is a quasi-dual Baer R-module. Then M is quasi-dual Baer.

Proof

Let \(S={\text {End}}_R(M)\) and for each \(\lambda \in \Lambda \), let \(S_\lambda ={\text {End}}_R(M_\lambda )\). Let \(\mathfrak {I}\) be a left ideal of the ring S. Since \(M_{\lambda }\) is fully invariant in M for all \(\lambda \in \Lambda \), we obtain \(\mathfrak {I}(M)=\oplus _{\lambda \in \Lambda } \mathfrak {I}(M_{\lambda })\). For each \(\lambda \in \Lambda \) and \(f \in S\), let \(f_\lambda \) denote the endomorphism of \(M_{\lambda }\) defined by \(f_\lambda (x)=f(x)\) for all \(x \in M_{\lambda }\). It is not hard to see that \(\mathfrak {I}_\lambda = \{f_\lambda \mid f \in \mathfrak {I}\}\) is a left ideal of \(S_\lambda \). Since each \(M_{\lambda }\) \((\) \(\lambda \in \Lambda \) \()\) is quasi-dual Baer, it follows that \(\mathfrak {I}_\lambda (M_{\lambda })\) is a direct summand of \(M_{\lambda }\) for every \(\lambda \in \Lambda \) (see Proposition 2.4). Therefore \(\mathfrak {I}(M)\) is a direct summand of M. Hence, by Proposition 2.4, we conclude that M is a quasi-dual Baer module. \(\square \)

3 Modules over commutative rings

This section is devoted to the study of some properties of quasi-dual Baer modules over commutative rings. Throughout this section, R will denote a commutative ring, unless otherwise stated.

Proposition 3.1

Let R be a commutative ring and let M be a cyclic R-module. Then M is quasi-dual Baer if and only if M is semisimple.

Proof

Without loss of generality, we can assume that \(M=R/I\) for some ideal I of R. It is easily seen that the module \(M_R\) is quasi-dual Baer if and only if the module \(M_{R/I}\) is quasi-dual Baer. The result follows from Proposition 2.10. \(\square \)

A module M is said to have the strong summand sum property (SSSP for short) if the sum of any family of direct summands of M is a direct summand of M. The class of dual Baer modules was characterized in [4, Theorem 2.1] as that of modules M having the SSSP such that for every \(\varphi : M \rightarrow M\), \(Im \varphi \) is a direct summand of M. To obtain a partial analogue of this characterization for quasi-dual Baer modules, Amouzegar and Talebi [1] introduced the FI-strong summand sum property (FI-SSSP for short). A module M is said to have the FI-SSSP if the sum of any family of fully invariant direct summands of M is a direct summand of M. The next result is taken from [1, Lemmas 2.1 and 2.2].

Lemma 3.2

Let R be a ring (not necessarily commutative) and let M be a quasi-dual Baer module. Then:

  1. (1)

    M has the FI-SSSP, and

  2. (2)

    for every endomorphism \(\varphi \) of M such that \(Im \varphi \) is fully invariant in M, \(Im \varphi \) is a direct summand of M.

Proposition 3.3

Let M be a module over a commutative ring R. If M is quasi-dual Baer, then \(M\mathfrak {a}\) is a direct summand of M for any ideal \(\mathfrak {a}\) of R.

Proof

Let \(\mathfrak {a}\) be an ideal of R and let \(r \in \mathfrak {a}\). It is clear that \(Mr=Im f_r\) where \(f_r\) is the R-endomorphism of M defined by \(f_r(x)=xr\) for every \(x \in M\). Moreover, it is easily seen that Mr is a fully invariant submodule of M. Now using Lemma 3.2, it follows that \(Im f_r\) is a direct summand of M for all \(r \in \mathfrak {a}\). Note that M has the FI-SSSP (see Lemma 3.2). Therefore, \(M\mathfrak {a} = \sum _{r \in \mathfrak {a}}Mr = \sum _{r \in \mathfrak {a}} Im f_r\) is a direct summand of M. \(\square \)

The converse of Proposition 3.3 is not true, in general, as shown in the following example. This example shows also that the class of quasi-dual Baer modules is not closed under direct products.

Example 3.4

Let \(\mathbf {P}\) denote the set of all prime numbers. Consider the \(\mathbb {Z}\)-module \(M=\prod _{p \in \mathbf {P}} \mathbb {Z}/p\mathbb {Z}\). Then the torsion submodule of M is \(T(M) = \oplus _{p \in \mathbf {P}} \mathbb {Z}/p\mathbb {Z}\). For each \(p \in \mathbf {P}\), let \(f_p\) denote the endomorphism of M defined by \((x_p)_{p \in \mathbf {P}} \mapsto (y_p)_{p \in \mathbf {P}}\) such that \(y_q=0\) for every \(q \not = p\) and \(y_p=x_p\). Let \(\mathfrak {I}\) denote the left ideal of \(S={\text {End}}_{\mathbb {Z}}(M)\) generated by all the \(f_p\) (\(p \in \mathbf {P}\)). Then \(\mathfrak {I}(M)=T(M)\) is not a direct summand of M. Therefore M is not quasi-dual Baer by Proposition 2.4. On the other hand, it is easily seen that \(M(n\mathbb {Z})=Mn\) is a direct summand of M for every integer n.

Corollary 3.5

Let M be a nonzero projective module over a commutative ring R. If M is a quasi-dual Baer R-module, then \({\text {Rad}}(M)=0\).

Proof

Assume that M is quasi-dual Baer. Let \(J={\text {Rad}}(R)\). Since M is projective, we have \({\text {Rad}}(M)=MJ\) by [2, Proposition 17.10]. From Proposition 3.3, it follows that \({\text {Rad}}(M)\) is a direct summand of M. So there exists a submodule N of M such that \(M={\text {Rad}}(M) \oplus N\). Clearly, we have \({\text {Rad}}(N)=0\). Therefore, \({\text {Rad}}({\text {Rad}}(M))={\text {Rad}}(M)\). Thus, \({\text {Rad}}(M)=0\) by [2, Proposition 17.14]. \(\square \)

Remark 3.6

Let R be a commutative ring and let M be an indecomposable quasi-dual Baer R-module. Using Proposition 3.3 and applying similar arguments as in the proof of [9, Proposition 4.4], we conclude that either \({\text {Rad}}(M)=M\) or M is a simple R-module. But this is not true, in general, if the ring R is not commutative. Indeed, we can consider the example given in [9, Example 4.10 and Remark 4.11].

Corollary 3.7

Let M be an indecomposable \(\mathbb {Z}\)-module. Then, the following are equivalent:

  1. (i)

    M is quasi-dual Baer;

  2. (ii)

    M is dual Baer;

  3. (iii)

    \(M \cong \mathbb {Q}\) or \(M \cong \mathbb {Z}(p^{\infty })\) or \(M \cong \mathbb {Z}/p\mathbb {Z}\), where p is a prime number.

Proof

(i) \(\Rightarrow \) (iii) Note that for every \(\mathbb {Z}\)-module N with \(Rad(N)=N\), N is an injective module. Now, this implication follows easily from Remark 3.6.

(iii) \(\Rightarrow \) (ii) This follows from [4, Corollary 3.5].

(ii) \(\Rightarrow \) (i) This is immediate. \(\square \)

Recall that a ring R is called semilocal if \(R/{\text {Rad}}(R)\) is a semisimple ring. The next result characterizes quasi-dual Baer modules over commutative semilocal rings.

Theorem 3.8

Let M be a nonzero module over a commutative semilocal ring R and let \(S={\text {End}}_R(M)\). Then, the following conditions are equivalent:

  1. (i)

    M is a quasi-dual Baer R-module;

  2. (ii)

    (a) \(M=M_1 \oplus M_2\) is a direct sum of submodules \(M_1\) and \(M_2\) such that \({\text {Rad}}(M_1)=M_1\) is quasi-dual Baer and \(M_2\) is semisimple, and

  3. (iii)

    (b) \({\mathfrak {I}}(M_2) \cap M_1 \subseteq {\mathfrak {I}}(M_1)\) for every two-sided ideal \(\mathfrak {I}\) of S.

Proof

(i)\(\Rightarrow \) (ii) Let \(J={\text {Rad}}(R)\). By [2, Corollary 15.18], we have \({\text {Rad}}(M)=MJ\). Applying Proposition 3.3, we deduce that \({\text {Rad}}(M)\) is a direct summand of M. Set \(M_1={\text {Rad}}(M)\) and let \(M_2\) be a submodule of M such that \(M=M_1 \oplus M_2\). Clearly, \({\text {Rad}}(M_2) = M_2J=0\) (see [2, Corollary 15.18]). Therefore \(M_2\) is an R/J-module. Since R/J is a semisimple ring, it follows that \(M_2\) is a semisimple R-module. Note that \(M_1\) is quasi-dual Baer by Corollary 2.5. This proves (a). Note that (b) follows directly from Proposition 2.15.

(ii) \(\Rightarrow \) (i) By Proposition 2.15. \(\square \)

In the following result, we describe the structure of quasi-dual Baer modules over commutative perfect rings.

Corollary 3.9

Let M be a nonzero module over a commutative perfect ring R. Then the following are equivalent:

  1. (i)

    M is a quasi-dual Baer module;

  2. (ii)

    M is a dual Baer module;

  3. (iii)

    M is a semisimple module.

Proof

(i) \(\Rightarrow \) (iii) Since R is a semilocal ring (see [2, Theorem 28.4]), there exists a semisimple submodule N of M such that \(M={\text {Rad}}(M) \oplus N\) by Theorem 3.8. But \({\text {Rad}}(M) \ll M\) by [2, Remark 28.5(3)]. Then, \({\text {Rad}}(M)=0\). It follows that \(M=N\) is a semisimple R-module.

(iii) \(\Rightarrow \) (ii) \(\Rightarrow \) (i) These implications are obvious. \(\square \)