1 Introduction and Preliminaries

The notion of finite Goldie dimension (denoted by FGD) of a module was defined by Goldie [14] wherein, the key notions for the study of FGD are essential submodules, uniform submodules and complement of a submodule (see, [4, 5]). The dualization of this concept namely, finite spanning dimension (denoted by FSD) in modules over rings was defined by Fleury [13] with the notions such as superfluous submodules, hollow submodules and supplements. Later, these concepts were studied in [1, 2, 7, 18]. The idea of FSD was generalized to module over nearrings (known as N-groups) in [6, 11, 16, 20]. They introduced the notions such as superfluous ideal, hollow ideal and FSD in N-groups and proved the corresponding structure theorems. The motivation of this paper arises from a natural question that what if one substitutes an arbitrary ideal \(\varOmega \) in place of an N-group G, so that it generalises the existing study of these notions. The classes of N-groups with these new notions are different from the classes of N-groups studied in [6, 8, 11, 19, 20]. Eventually, in this paper we define \(\varOmega \)-superfluous ideal of an N-group G, where \(\varOmega \) is an ideal of G and obtain some connections to matrix nearrings, and some combinatorial aspects. In section 2, we define g-superfluous and g-supplement ideals of N-groups as a generalization of respective notions defined by [20]. We have constructed examples where G is non-abelian.

In Sect. 3, we introduce superfluous ideals of N-groups and prove some important properties and provide necessary examples. In Sect. 4, we consider the notion strictly superfluous in terms of N-subgroups and gave examples which indicate that the classes of N-groups with superfluous and strictly superfluous are different. Matrix nearrings over arbitrary nearrings were defined in [15] and studied in [23]. In Sect. 5, we introduce the superfluous and g-superfluous ideals in \(M_{n}(N)\)-group \(N^{n}\), and establish a one-one correspondence between superfluous and g-superfluous ideals of N (over itself) and those of \(M_{n}(N)\)-group \(N^{n}\). In Sect. 6, we introduce superfluous ideal graphs of nearrings and prove some properties with examples.

A (right) nearring is a set N together with two binary operations \(``+''\) and \(``\cdot ''\) such that \((N,+)\) is a group, \((N, \cdot )\) is a semigroup and right distributive law holds. In general, for some \(n\in N\), \(n\cdot 0 \ne 0\), and so we call N is zero-symmetric if \(n\cdot 0 =0\) for all \(n\in N\). A normal subgroup I of a nearring N is called an ideal of N (denoted by \(I\trianglelefteq N\)) if \(IN\subseteq I\) and \(a(b+i)-ab\in I\) for all \(a,b\in N\) and \(i\in I\). An additive group G is said to be an N-group if there exists a map \(N\times G \rightarrow G\) defined by \((n,g)\longmapsto ng\) satisfying \((n+n_{1})g=ng+n_{1}g\) and \((n\cdot n_{1})g=n(n_{1}g)\) for all \(n,n_{1}\in N\) and \(g\in G\). Throughout, we use G for an N-group. A subgroup H of G is said to be an N-subgroup (denoted as, \(H\le _{N} G\)) of G if \(NH\subseteq H\); and a normal subgroup I of G is called an ideal (denoted as, \(I\trianglelefteq _{N} G\)) of G if \(n(g+i)-ng\in I\), for all \(n\in N\), \(g\in G\) and \(i\in I\). An ideal S of G is said to be superfluous in G if \(S+K=G\) and K is an ideal of G, imply \(K=G\), and G is called hollow if every proper ideal of G is superfluous in G. For any two N-subgroups H and K of G, K is said to be a supplement for H if \(H+K=G\) and \(H+K^{'}\ne G\) for any proper ideal \(K^{'}\)of K. For any ideals I, J, K of N (or of G), if \(K\subseteq I\), then \(I\cap (J+K)=(I\cap J)+K\). We use \(I\oplus J\) to denote the direct sum of ideals I and J of G. We refer to Pilz [17] and Bhavanari and Kuncham [10] for fundamental literature in nearrings.

We consider simple and finite graphs, whose vertex set is V and edge set is E. A vertex v of a graph is called a universal vertex if degree of v= \(|V|-1\). If there exists a path between every two vertices of a graph, then the graph is connected otherwise the graph is disconnected. A graph whose vertex set is empty is called a null graph and a graph having atleast one vertex and empty edge set is called an empty graph.

2 Generalized supplements

The notion of superfluous submodule of module over a ring was studied by [3, 7]. We define generalized superfluous (briefly, g-superfluous) ideal of an N-group G as follows.

Definition 2.1

An ideal K of G is called g-superfluous if \(G=K+T\) and \(T\le _{e} G\), then \(T=G\). We denote this by \(K\ll _{gs} G\).

Remark 2.2

Every superfluous ideal of G is g-superfluous.

Example 2.3

Consider the nearring \(N=({\mathbb {Z}}_{2}\times {\mathbb {Z}}_{2}\times {\mathbb {Z}}_{2}, +, \cdot )\) with the notation given in page no. 420, (N) of [17]. That is, \((0,0,0)=0\), \((0,0,1)=1\) \((0,1,0)=2\) \((0,1,1)=3\), \((1,0,0)=4\), \((1,0,1)=5\), \((1,1,0)=6\), and \((1,1,1)=7.\) The multiplication table is given below. Let \(G=N\) (Table 1).

The proper ideals of G are \(I_{1}=\{0,2,4,6\}\) and \(I_{2}=\{0,1\}\). It can be seen that \(I_{1}\) and \(I_{2}\) are g-superfluous but not superfluous in G as \(I_{1}+I_{2}=G\)

Table 1 Multiplication \(\star \) on N

Proposition 2.4

Let K, J be ideals of G such that \(K\subseteq J\). If \(K\ll _{gs} J\), then \(K\ll _{gs} L\) for any ideal L of G with \(J\subseteq L\).

Proof

Let \(T\le _{e} L\) such that \(K+T= L\). We prove \(T=L\). Clearly \(T\subseteq L\). Since \(K+T=L\), we have \((K+T)\cap J=L\cap J\). Since \(K\subseteq J\) and modular law, we have \( K+(T\cap J)=J\). Since \(T\cap J\le _{e}J\) and \(K\ll _{gs} J\), we have \((T\cap J)=J\) and so \( T\subseteq J\). Since \(K\subseteq J\subseteq T\), we get \(L=K+T\subseteq J+T=T\). Therefore \(L=T\) and hence \(K\ll _{gs} L\). \(\square \)

Proposition 2.5

Let \(K_{1}\), \(K_{2}\), \(G_{1}\), \(G_{2}\) be ideals of G such that \(K_{1}\subseteq G_{1}\) and \(K_{2}\subseteq G_{2}\). If \(K_{1}\ll _{gs} G_{1}\) and \(K_{2}\ll _{gs} G_{2}\), then \(K_{1}+K_{2}\ll _{gs} G_{1}+G_{2}\).

Proof

Since \(K_{i}\ll _{gs} G_{i}\), by Proposition 2.4, we have \(K_{i}\ll _{gs} G_{1}+G_{2}\) for \(i=\{1,2\}\). Let \(T\le _{e} G_{1}+G_{2}\) be such that \(K_{1}+K_{2}+T=G_{1}+G_{2}\). Since \(T\le _{e} G_{1}+G_{2}\), we have \(T+K_{2}\le _{e} G_{1}+G_{2}\). Now \(K_{1}+(K_{2}+T)=G_{1}+G_{2}\) and \(K_{1}\ll _{gs} G_{1}+G_{2}\), imply that \(K_{2}+T=G_{1}+G_{2}\). Again, since \(K_{2}\ll _{gs} G_{1}+G_{2}\) and \(T\le _{e} G_{1}+G_{2}\) implies \(T=G_{1}+G_{2}\). \(\square \)

Note 2.6

Let X, K be ideals of G such that \(K\subseteq X\). If \(\frac{X}{K}\le _{e} \frac{G}{K}\), then \(X\le _{e} G\).

Proposition 2.7

Let U, V and K be ideals of G. If \(U\ll _{gs} V\), then \(\frac{U+K}{K}\ll _{gs}\frac{V+K}{K}\).

Proof

Let \(\frac{T}{K}\le _{e}\frac{V+K}{K}\) be such that \(\frac{U+K}{K}+\frac{T}{K}=\frac{V+K}{K}\). Then \(U+K+T=V+K\). Since \(K\subseteq T\), we get \(U+T=V+K\). Now \(U\ll _{gs} V\) implies that \(U\ll _{gs} V+K\), and \(\frac{T}{K}\le _{e} \frac{V+K}{K}\) implies \(T\le _{e} V+K\). Since \(U\ll _{gs} V+K\), \(T\le _{e} V+K\) and \(U+T=V+K\), we get \(T=V+K\), which implies \(\frac{T}{K}=\frac{V+K}{K}\). Therefore, \(\frac{U}{K}\ll _{gs} \frac{V+K}{K}\). \(\square \)

Proposition 2.8

Let J, K, L be ideals of G such that \(K\subseteq J\).

  1. 1.

    If \(J\ll _{gs} G\), then \(K\ll _{gs} G\) and \(\frac{J}{K}\ll _{gs} \frac{G}{K}\).

  2. 2.

    \(J+L\ll _{gs}G\) if and only if \(J\ll _{gs}G\) and \(L\ll _{gs} G\).

Proof

  1. 1.

    Suppose \(J\ll _{gs} G\). To prove \(K\ll _{gs} G\), let \(T\le _{e} G\) such that \(K+T=G\). Since \(K\subseteq J\), we get \(J+T=G\). Since \(J\ll _{gs} G\), we have \(T=G\), shows that \(K\ll _{gs} G\).

    Next we prove \(\frac{J}{K}\ll _{gs} \frac{G}{K}\). Let \(\frac{X}{K}\le _{e} \frac{G}{K}\) such that \(\frac{J}{K}+\frac{X}{K}=\frac{G}{K}\). Then \(\frac{J+X}{K}=\frac{G}{K}\), implies that \(J+X=G\). Since \(X\le _{e}G\), we get \(X=G\). Therefore \(\frac{X}{K}=\frac{G}{K}\).

  2. 2.

    Suppose \(J+L\ll _{gs} G\). To prove \(J\ll _{gs}G\) and \(L\ll _{gs} G\), let \(T\le _{e} G\) such that \(J+T=G\). Then \((J+L)+T=G\). Since \(J+L \ll _{gs} G\), we have \(T=G\). In a similar way, we get \(L\ll _{gs} G\).

    Conversely, suppose that \(J\ll _{gs} G\) and \(L\ll _{gs} G\). To prove \(J+L\ll _{gs} G\), let \(T\le _{e} G\) such that \((J+L)+T=G\). This means, \(J+(L+T)=G\). Since \(L+T\le _{e}G\) and \(J\ll _{gs} G\), it follows that \(L+T =G\). Again since \(L\ll _{gs}G\) and \(T\le _{e}G\), we get \(T=G\). Therefore \(J+L\ll _{gs} G\).

\(\square \)

Definition 2.9

Let P and Q be N-subgroups of G. Q is said to be a g-supplement of P if \(G=P+Q\) and \(G=P+T\) with \(T\le _{e}Q\) implies that \(T=Q\).

An N-group G is called g-supplemented if every ideal of G has a g-supplement.

Remark 2.10

Every supplemented N-group is g-supplemented.

Example 2.11

Consider the nearring given in (K(139), page 418 of [17]). Let \(N=(D_{8}, +, \star )\), the dihedral group of order 8 and \(G=N\).

The subgroups of \((D_{8}, +)\) are \(H_{1}=\langle e\rangle \), \(H_{2}=\langle s\rangle \), \(H_{3}=\langle sr^{2}\rangle \), \(H_{4}=\langle r^{2}\rangle \), \(H_{5}=\langle sr^{3}\rangle \), \(H_{6}=\langle sr\rangle \), \(H_{7}=\langle \{s, r^{2}\}\rangle \), \(H_{8}=\langle r\rangle \) and \(H_{9}=\langle \{r^{2},sr^{3}\}\rangle \). The N-subgroups are \(H_{1}\), \(H_{2}\), \(H_{3}\), \(H_{4}\), \(H_{5}\), \(H_{7}\) and \(H_{9}\), and ideals of \(D_{8}\) are \(H_{4}\), \(H_{7}\) and \(H_{9}\). The ideals of \(H_{9}\)(when it is considered as an N-group) are \(H_{4}\) and \(H_{5}\) which are not essential in \(H_{9}\). Observe that \(H_{9}\) is not a supplement of \(H_{7}\) as there exists an ideal \(H_{5}\) of \(H_{9}\) such that \(H_{7}+H_{5}=D_{8}\). Furthermore, all ideals of \(H_{9}\) are not essential, we do not have any essential ideal I of \(H_{9}\) such that \(H_{7}+I=D_{8}\). Therefore \(H_{9}\) is a g-supplement of \(H_{7}\) (Table 2).

Table 2 Multiplication \(\star \) on N

Lemma 2.12

Let P, Q be ideals of G. Then Q is a g-supplement of P if and only if \(G=P+Q\) and \(P\cap Q \ll _{gs} Q\).

Proof

Suppose Q is a g-supplement of P in G. Then \(P+Q=G\) and \(P+Q^{'}\ne G\) for any essential ideal \(Q^{'}\) of Q. We prove \(P\cap Q \ll _{gs} Q\). Let \(T\le _{e} Q\) such that \((P\cap Q)+T=Q\). Then \(G=P+Q=P+(P\cap Q)+T=P+T\), as \((P\cap Q)\subseteq P\). Now \(G=P+T\) where \(T\le _{e} Q \). Since Q is a g-supplement of P, we get \(T=Q\). Therefore \(P\cap Q \ll _{gs} Q\).

Conversely, suppose that \(G=P+Q\) and \(P\cap Q \ll _{gs} Q\). To show Q is a g-supplement of P in G, let \(G=P+T\) for some essential ideal T of Q. Now, since \(T\subseteq Q\), by modular law we get \(Q=Q\cap G=Q\cap (P+T)=(Q\cap P)+T\). Since \((P\cap Q)\ll _{gs}Q\) and \(T\le _{e} Q\), we get \(T=Q\). Therefore Q is a g-supplement of P in G. \(\square \)

Proposition 2.13

[12]

  1. 1.

    Let G be an N-group and let I, J be the ideals of G with \(G=I\oplus J\). Then \(a+b=b+a\) for all \(a\in I\) and \(b\in J\).

  2. 2.

    If \(N=N_{0}\), \(n\in N\), \(a\in I\), \(b\in J\) and the sum \(I\oplus J\) is direct in G, then \(n(a+b)=na+nb\).

  3. 3.

    Let \(N=N_{0}\) and \(I\trianglelefteq _{N}G\) be a direct summand. Then each ideal of I is an ideal of G.

Lemma 2.14

Let A, B and C be ideals of G. Then

$$\begin{aligned} A\cap (B+C) \trianglelefteq _{N} B\cap (A+C) + C\cap (A+B). \end{aligned}$$

Proof

We have \(A\cap (B+C)\trianglelefteq _{N} G\). Let \(p\in A\cap (B+C)\). Then \(p\in A\) and \(p \in B+C\), which implies \(p=b+c \) for some \(b\in B\) and \(c\in C\). Now, \(b=p-c \in A+C\) and \(c= -b+p \in B+A= A+B\) and hence \(p=b+c \in B\cap (A+C) + C\cap (A+B)\). Therefore \(A\cap (B+C)\trianglelefteq _{N} G\), which is contained in \(B\cap (A+C) + C\cap (A+B)\) and hence \(A\cap (B+C)\trianglelefteq _{N} B\cap (A+C) + C\cap (A+B)\). \(\square \)

Lemma 2.15

Let N be zero-symmetric and \(G_{1}\), U be ideals of G and \(G_{1}\) be g-supplemented and a direct summand of G. If \(G_{1}+U\) has a g-supplement in G, then U has a g-supplement in G.

Proof

Let X be a g-supplement of \(G_{1}+U\) in G. Then by Lemma 2.12, \(G_{1}+U+X=G\) and \((G_{1}+U)\cap X\ll _{gs} X\). Since \(G_{1}\) is g-supplemented, \((U+X)\cap G_{1}\) has a g-supplement Y in \(G_{1}\). That is, \(G_{1}\cap (U+X)+Y=G_{1}\) and \(G_{1}\cap (U+X)\cap Y \ll _{gs} Y\), by Lemma 2.12. Since \(G_{1}\) is a direct summand, \(Y\trianglelefteq _{N} G\). This yield,

$$\begin{aligned} G=G_{1}\cap (U+X)+Y+(U+X)=U+X+Y \end{aligned}$$

and

$$\begin{aligned} U\cap (X+Y)&\trianglelefteq _{N} X\cap (U+Y)+ Y\cap (U+X)\\ {}&\trianglelefteq _{N} X\cap (G_{1}+U)+Y\cap G_{1}\cap (U+X) \\&\ll _{gs} X+Y . \end{aligned}$$

Hence, \(X+Y\) is a g-supplement of U in G. \(\square \)

Proposition 2.16

Let G be an N-group. Let K, U and V be ideals of G such that \(K\subseteq U\). Let V be a g-supplement of U in G. Then \(\frac{V+K}{K}\) is a g-supplement of \(\frac{U}{K}\).

Proof

Since V is a g-supplement of U in G, we have \(G=U+V\) and \(U\cap V\ll _{gs} V\) which implies \(\frac{U\cap V+K}{K} \ll _{gs} \frac{V+K}{K}\). Now \(\frac{G}{K}=\frac{U+V}{K}=\frac{U}{K}+\frac{V+K}{K}\). Also \(\frac{U}{K}\cap \frac{V+K}{K}=\frac{U\cap (V+K)}{K}=\frac{U\cap V +K}{K} \ll _{gs} \frac{V+K}{K}\). Therefore, \(\frac{V+K}{K}\) is a g-supplement of \(\frac{U}{K}\). \(\square \)

3 Superfluous ideals

Definition 3.1

Let \(\varOmega \trianglelefteq _{N} G \). An ideal (or N-subgroup) H of G is said to be \(\varOmega \)-superfluous in G if \(\varOmega \nsubseteq H\) and for any ideal L of G, \(\varOmega \subseteq L+H\) implies \(\varOmega \subseteq L\). We denote it by \(H \ll _{\varOmega } G\).

Note 3.2

If \(\varOmega =G\), then \(\varOmega \)-superfluous coincides with the notion of superfluous defined by [20]. In this case, we denote \(H\ll G\) whenever an ideal H is superfluous in G. Trivially, the ideal (0) is superfluous in G.

Example 3.3

Let \(N={\mathbb {Z}},\) the set of integers and \(G=({\mathbb {Z}}_{24}, +_{24})\). Then G is an N-group. Let \(\varOmega = 8{\mathbb {Z}}_{24}\). Then \(6{\mathbb {Z}}_{24}\), \(3{\mathbb {Z}}_{24}\), \(12{\mathbb {Z}}_{24}\) are \(\varOmega \)-superfluous, whereas \(3{\mathbb {Z}}_{24}\) is not superfluous in G, since \(3{\mathbb {Z}}_{24}+2{\mathbb {Z}}_{24}={\mathbb {Z}}_{24}\) but \(2{\mathbb {Z}}_{24} \ne {\mathbb {Z}}_{24}\).

Example 3.4

Let \(N={\mathbb {Z}}\) and \(G={\mathbb {Z}}_{12}\). Then G is an N-group. Let \(\varOmega = 4{\mathbb {Z}}_{12}\). Then \(6{\mathbb {Z}}_{12}\), \(3{\mathbb {Z}}_{12}\) are \(\varOmega \)-superfluous, whereas \(3{\mathbb {Z}}_{12}\) is not superfluous in G, since \(3{\mathbb {Z}}_{12}+2{\mathbb {Z}}_{12}={\mathbb {Z}}_{12}\) but \(2{\mathbb {Z}}_{12} \ne {\mathbb {Z}}_{12}\).

Example 3.5

Let \(N=\begin{pmatrix} 0 &{} {\mathbb {Z}}_{q^m} \\ 0 &{} 0 \end{pmatrix}\) and \(G=N\). Then the ideals and N-subgroups are \(H_{i} = \{\begin{pmatrix} 0 &{} q^{i}{\mathbb {Z}}_{q^m}\\ 0 &{} 0 \end{pmatrix}: 0\le i \le m \}\). Let \(\varOmega = H_{k}\). Then \(H_{j}\ll _{\varOmega } G\) for all \(j\le k\).

Example 3.6

Consider the N-group given in the Example 2.11.

  1. 1.

    \(H_{7}\ll _{H_{9}}G\), \(H_{4}\ll _{H_{9}} G\), \(H_{9}\ll _{H_{7}}G\), \(H_{4}\ll _{H_{7}}G\).

  2. 2.

    The N-subgroups \(H_{2}\), \(H_{3}\), \(H_{4}\) and \(H_{7}\) are \(\varOmega \)-superfluous in G with \(\varOmega =H_{9}\).

  3. 3.

    The N-subgroups \(H_{4}\), \(H_{5}\), \(H_{9}\) are \(\varOmega \)-superfluous in G with \(\varOmega =H_{7}\).

Proposition 3.7

Let \(\varOmega \) be an ideal and X be an ideal (or N-subgroup) of G. If X is \(\varOmega \)-superfluous in G, then \(X\cap Y\) is \(\varOmega \)-superfluous in G for any ideal (or N-subgroup) Y of G.

Proof

Suppose X is \(\varOmega \)-superfluous in G. Let \(Y\trianglelefteq _{N} G\). Since \(\varOmega \nsubseteq X\), we have \(\varOmega \nsubseteq X\cap Y\). On a contrary, suppose \(X\cap Y\) is not superfluous in G. Then there exists a proper ideal K of G such that \(\varOmega \nsubseteq K\) and \(\varOmega \subseteq (X\cap Y)+K\). Now, since \(X\cap Y \subseteq X\) we get \(\varOmega \subseteq X+K\), a contradiction as \(X\ll _{\varOmega } G\). Therefore \(X\cap Y\ll _{\varOmega }G\). \(\square \)

Proposition 3.8

Let \(\varOmega \), K be ideals of G. If \(K \ll _{\varOmega } G\), then \(K\cap \varOmega \ll G\).

Proof

Let \(K \ll _{\varOmega } G\). To prove \(K \cap \varOmega \ll G\), let \(L \trianglelefteq _{N} G\) be such that \((K \cap \varOmega )+L=G\). Now \(\varOmega \subseteq (K \cap \varOmega )+L \subseteq K+L\). Now since \(K \ll _{\varOmega } G\), we have that \(\varOmega \subseteq L\). Also since \(K\cap \varOmega \subseteq \varOmega \subseteq L\), it follows that \(L=(K\cap \varOmega )+L=G\). Therefore, \(K\cap \varOmega \ll G\). \(\square \)

Proposition 3.9

Let N be zero-symmetric and \(\varOmega \trianglelefteq _{N}G\), which is a direct summand, and let \(P \trianglelefteq _{N} G\) contained in \(\varOmega \). Then \(P\ll _{\varOmega } G\) if and only if \(P\ll \varOmega \).

Proof

Suppose \(P\ll _{\varOmega } G\). Since \(P\trianglelefteq _{N}G\) and \(P\subseteq \varOmega \), we have \(P\trianglelefteq _{N}\varOmega \). To prove \(P\ll \varOmega \), let \(L \trianglelefteq _{N} \varOmega \) be such that \(P+L=\varOmega \). Since \(\varOmega \) is a direct summand, by Proposition 2.13(3), \(L\trianglelefteq _{N} G\). Now \(\varOmega \subseteq P+L\) and \(P \ll _{\varOmega } G\), we get \(\varOmega \subseteq L\). Since \(L\subseteq \varOmega \), it follows that \(L=\varOmega \).

Conversely, suppose that \(P\ll \varOmega \). Let \(L \trianglelefteq _{N} G\) be such that \(\varOmega \subseteq P+L\). Now \(\varOmega = (P+L)\cap \varOmega =P+ ( L \cap \varOmega )\), by modular law, and since \(P\ll \varOmega \), it follows that \(\varOmega =L\cap \varOmega \). Hence \(\varOmega \subseteq L\). \(\square \)

Proposition 3.10

Let \(K \trianglelefteq _{N}G\) and let P and \(\varOmega \) be ideals of G which are contained in K. If \(P \ll _{\varOmega } K\), then \(P \ll _{\varOmega } G\).

Proof

Suppose that \(P \ll _{\varOmega } K\). To prove \(P \ll _{\varOmega } G\), let \(L \trianglelefteq _{N} G\) be such that \(\varOmega \subseteq P+L\). Since \(\varOmega \subseteq K\) and by modular law, we have \(\varOmega \subseteq (P+L)\cap K=P+(L\cap K)\). Since \(L\cap K\trianglelefteq _{N} K\) and \(P \ll _{\varOmega } K\), we have \( \varOmega \subseteq (L\cap K)\), which implies \(\varOmega \subseteq L\). Hence \(P\ll _{\varOmega } G\). \(\square \)

Remark 3.11

It can be easily seen that the Propositions 3.7, 3.8, 3.9 and 3.10 hold for N-subgroups also.

Remark 3.12

The following proposition holds for ideals of G but not for N-subgroups, as sum of two N-subgroups need not be an N-subgroup.

Proposition 3.13

Let \(N_{1}\), \(N_{2}\), \(\varOmega \) be ideals of G. Then \(N_{1}\ll _{\varOmega } G\) and \(N_{2}\ll _{\varOmega } G\) if and only if \(N_{1}+N_{2}\ll _{\varOmega } G\).

Proof

Suppose that \(N_{1}\ll _{\varOmega } G\) and \(N_{2}\ll _{\varOmega } G\). Let \(L\trianglelefteq _{N} G\) be such that \(\varOmega \subseteq (N_{1}+N_{2})+L = N_{1}+(N_{2}+L)\). Since \(N_{1}\ll _{\varOmega } G\), we have \(\varOmega \subseteq N_{2}+L\), and again since \(N_{2}\ll _{\varOmega } G\), we get \(\varOmega \subseteq L\).

Conversely, suppose \(N_{1}+N_{2}\ll _{\varOmega } G\). Let \(L\trianglelefteq _{N} G\) be such that \(\varOmega \subseteq N_{1}+L \subseteq (N_{1}+N_{2})+L\). Now since \(N_{1}+N_{2}\ll _{\varOmega } G\), we get \(\varOmega \subseteq L\). Similar assertion proves \(N_{2}\ll _{\varOmega } G\). \(\square \)

Note 3.14

Let N be zero-symmetric and \(K_{1}\trianglelefteq _{N} G_{1}\trianglelefteq _{N} G\) and \(K_{2}\trianglelefteq _{N} G_{2}\trianglelefteq _{N} G\), \(\varOmega \trianglelefteq _{N} G\) such that \(G_{1}\oplus G_{2}=G\). Then \(K_{1}\ll _{\varOmega } G_{1}\) and \(K_{2}\ll _{\varOmega } G_{2}\) if and only if \(K_{1}+K_{2} \ll _{\varOmega } G_{1}+G_{2} \).

Proposition 3.15

Let \(\varOmega \), K, P be ideals of G such that \(K \subset \varOmega \), \(K\subseteq P\) and \(\varOmega \nsubseteq P\). Then \(P\ll _{\varOmega } G\) if and only if \(K\ll _{\varOmega } G\) and \(\frac{P}{K}\ll _{\frac{\varOmega }{K}} \frac{G}{K}\).

Proof

Suppose \(P\ll _{\varOmega } G\). To prove \(K\ll _{\varOmega } G\), let \(L \trianglelefteq _{N} G\) such that \(\varOmega \subseteq K+L\). Since \(K\subseteq P\), we get \(\varOmega \subseteq P+L\). Since \(P \ll _{\varOmega } G\), we have \(\varOmega \subseteq L\), and thus \(K\ll _{\varOmega } G\).

Now to prove \(\frac{P}{K}\ll _{\frac{\varOmega }{K}} \frac{G}{K}\), let \(\frac{L}{K} \trianglelefteq _{N} \frac{G}{K}\), where \(K\subseteq L\trianglelefteq _{N} G\) such that, \(\frac{\varOmega }{K} \subseteq \frac{P}{K} + \frac{L}{K} =\frac{(P+L)}{K}\). Then \(\varOmega \subseteq P+L\). Since \(P\ll _{\varOmega } G\), we get \(\varOmega \subseteq L\), which implies that \(\frac{\varOmega }{K} \subseteq \frac{L}{K}\). Hence, \(\frac{P}{K}\ll _{\frac{\varOmega }{K}} \frac{G}{K}\).

Conversely, suppose that \(K\ll _{\varOmega }G\) and \(\frac{P}{K}\ll _{\frac{\varOmega }{K}} \frac{G}{K}\). To prove \(P\ll _{\varOmega } G\), let \(L\trianglelefteq _{N} G\) such that \(\varOmega \subseteq P+L\). Then \(\frac{\varOmega }{K} \subseteq \frac{(P+L)}{K} = \frac{P}{K} + \frac{L+K}{K}\). Since \(\frac{P}{K}\ll _{\frac{\varOmega }{K}} \frac{G}{K}\), it follows that \(\frac{\varOmega }{K} \subseteq \frac{L+K}{K}\), which implies \(\varOmega \subseteq L+K\). Since \(K\ll _{\varOmega }G\), we get \(\varOmega \subseteq L\). Hence, \(P\ll _{\varOmega } G\). \(\square \)

Proposition 3.16

Let \(\{\varOmega _{i}\}_{i\in I}\) be a family of ideals of G and \(K\trianglelefteq _{N} G\). If for each \(i\in I\), \(K\ll _{\varOmega _{i}} G\), then \(K \ll _{\sum \limits _{i\in I}\varOmega _{i}} G\).

Proof

Suppose \(K\ll _{\varOmega _{i}} G\) for each \(i\in I\) and \(\sum \limits _{i\in I}\varOmega _{i} \subseteq K+L\) where \(L\trianglelefteq _{N} G\). Then since \(\varOmega _{i}\subseteq \sum \varOmega _{i} \subseteq K+L\) for each \(i\in I\) and \(K\ll _{\varOmega _{i}} G\), we have \(\varOmega _{i} \subseteq L\), which shows that \(\sum \varOmega _{i} \subseteq L\). Hence, \(K\ll _{\sum \limits _{i\in I}\varOmega _{i}} G\). \(\square \)

Corollary 3.17

Let \(K_{1}\) and \(K_{2}\) be ideals of G such that \(K_{1}\ll _{K_{2}}G\) and \(K_{2}\ll _{K_{1}}G\). Then \(K_{1}\cap K_{2} \ll _{K_{1}+K_{2}} G\).

Proof

First we show that \(K_{1}\cap K_{2}\ll _{K_{1}} G\). For this, let \(K_{1}\subseteq (K_{1}\cap K_{2})+X\), where X is an ideal of G. Now \(K_{1}\subseteq K_{2}+X\) and since \(K_{2}\ll _{K_{1}}G\) we get \(K_{1}\subseteq X\). Therefore, \(K_{1}\cap K_{2}\ll _{K_{1}} G\). In a similar way, we get \(K_{1}\cap K_{2}\ll _{K_{2}} G\). Hence, by Proposition 3.16, it follows that \(K_{1}\cap K_{2} \ll _{K_{1}+K_{2}} G\). \(\square \)

The converse of the Corollary 3.17 need not be true, as shown in the following example.

Example 3.18

Consider the N-group \({\mathbb {Z}}_{48}\) over \({\mathbb {Z}}\). Let \(K_{1}=8{\mathbb {Z}}_{48}\) and \(K_{2}=6{\mathbb {Z}}_{48}\). Then \(8{\mathbb {Z}}_{48}\cap 6{\mathbb {Z}}_{48} \ll _{8{\mathbb {Z}}_{48}+6{\mathbb {Z}}_{48}} {\mathbb {Z}}_{48}\), whereas \(8{\mathbb {Z}}_{48}\ll _{6{\mathbb {Z}}_{48}} {\mathbb {Z}}_{48}\) and \(6{\mathbb {Z}}_{48}\) is not \(8{\mathbb {Z}}_{48}\)-superfluous in \({\mathbb {Z}}_{48}\).

Proposition 3.19

Let K and \(\varOmega \) be ideals of G such that \(\varOmega \nsubseteq K\). Let \(G^{'}\) be an N-group and \(f:G\rightarrow G^{'}\) be an epimorphism with \(f(\varOmega )\nsubseteq f(K)\). If \(K\ll _{\varOmega }G\), then \(f(K)\ll _{f(\varOmega )}G^{'}\). The converse holds if f is injective.

Proof

Suppose that \(K\ll _{\varOmega }G\). Since f is an epimorphism, we have \(f(K)\trianglelefteq _{N}G^{'}\) by Theorem 1.30 of [17]. Let \(X\trianglelefteq _{N} G^{'}\) be such that \(f(\varOmega )\subseteq f(K)+X\). Then \(\varOmega \subseteq K+f^{-1}(X)\). Since \(f^{-1}(X)\trianglelefteq _{N}G\) and \(K\ll _{\varOmega }G\), it follows that \(\varOmega \subseteq f^{-1}(X)\). Hence \(f(\varOmega )\subseteq X\).

Conversely, suppose that f is injective and \(f(K)\ll _{f(\varOmega )}G^{'}\). Let \(X\trianglelefteq _{N}G\) be such that \(\varOmega \subseteq K+X\). Then \(f(\varOmega )\subseteq f(K+X)= f(K)+f(X)\). Since \(f(K)\ll _{f(\varOmega )}G^{'}\), we have \(f(\varOmega )\subseteq f(X)\). Therefore, \(f^{-1}(f(\varOmega ))\subseteq f^{-1}(f(X))\). Now by 2.17 of [17], \(\varOmega +ker~f \subseteq X+ker~f\). As f is injective, we get \(\varOmega \subseteq X\). \(\square \)

Remark 3.20

Unlike in module over rings, the condition f is a homomorphism is not sufficient, as a homomorphic image of an ideal need not be an ideal. So we consider f to be an epimorphism. The following example justifies the condition f is a homomorphism is not sufficient.

Example 3.21

Consider the nearring given in the Example 3.6 and the ideals \(H_{9}=\{e, r^{2}, sr^{3}, sr\}\) and \(H_{7}=\{e, r^{2}, s, sr^{2}\}\) of G. Let f be an N-endomorphism of G defined by

$$\begin{aligned} f(g)=g\cdot sr~~~\text {for all}~~ g\in G. \end{aligned}$$

Then \(f(H_{9})=\{e,sr^{3}\}\) and \(f(H_{7})=\{e,r^{2}\}\). It can be seen that \(H_{7}\ll _{H_{9}}G\), but \(f(H_{7})\) is not \(f(H_{9})\) superfluous in G, since \(f(H_{9}) \ntrianglelefteq _{N} G\).

Definition 3.22

Let \(\varOmega \trianglelefteq _{N}G\). G is said to be \(\varOmega \)-hollow if every proper ideal of G which does not contain in \(\varOmega \) is \(\varOmega \)-superfluous in G.

Remark 3.23

  1. 1.

    Every hollow N-group is \(\varOmega \)-hollow with \(\varOmega =G\).

  2. 2.

    \(\varOmega \)-hollow need not be hollow and we justify this in the following example.

Example 3.24

Consider the Example 3.6 in which \(H_{4}\),  \(H_{7}\) are \(H_{9}\)-superfluous in G and \(H_{4}\),\(H_{9}\) are \(H_{7}\)-superfluous in G. Hence it is \(H_{7}\)-hollow as well as \(H_{9}\)-hollow. However, G is not hollow, since \(H_{7}\) is not superfluous in G as \(H_{7}+H_{9}=G\) but \(H_{9}\ne G\).

Definition 3.25

Let N be zero-symmetric, and let \(\varOmega \), H be ideals of G such that \(\varOmega \nsubseteq H\). An N-subgroup K of G is said to be an \(\varOmega \)-supplement of H if \(\varOmega \subseteq H+K\) and \(\varOmega \nsubseteq H+K^{'}\) for any ideal \(K^{'}\) of K.

Example 3.26

Consider the Example 2.11.

Let \(\varOmega =H_{7}\). Here \(H_{2}\) is an \(\varOmega \)-supplement of \(H_{4}\), but \(H_{2}\) is not a supplement of \(H_{4}\) as \(H_{2}+H_{4}\ne G\).

Example 3.27

\(N=D_{8}\) with the multiplication given in the Table 3. Let \(G=N\).

The ideals of G are \(I_{1}=G\), \(I_{2}=\{e, r^{2},r^{3}, r\}\), \(I_{3}=\{e, sr^{3}, r^{2}, sr\}\), \(I_{4}=\{e,sr^{2},s,r^{2}\}\), \(I_{5}=\{e,r^{2}\}\) and \(I_{6}=\{e\}\), and N-subgroups are \(I_{1}\), \(I_{2}\), \(I_{3}\), \(I_{4}\), \(I_{5}\), \(X_{1}=\{e,s\}\), \(X_{2}=\{e,sr^{2}\}\), \(X_{3}=\{e,sr\}\). Let \(\varOmega =I_{4}\). Here \(I_{3}\) is an \(\varOmega \)-supplement of \(I_{2}\), \(X_{1}\) is an \(\varOmega \)-supplement of \(I_{2}\), \(I_{3}\) and \(I_{5}\). Further, \(X_{1}\) is not a supplement of \(I_{5}\) as \(I_{5}+X_{1}\ne G\).

Table 3 Multiplication \(\star \) on N

Note 3.28

If \(\varOmega =G\), then \(\varOmega \)-supplement coincides with the supplement as defined by [20].

4 Strictly superfluous ideals

In case of N-groups, we have substructures namely N-subgroups and ideals, whereas in modules over rings, these concepts coincide. So we consider the notion strictly superfluous in terms of N-subgroups. We provide explicit examples which indicate that the classes superfluous and strictly superfluous are different.

Definition 4.1

An ideal H of G is called strictly superfluous in G (denoted by \(H\ll ^{s}G\)) if K is any N-subgroup of G such that \(H+K=G\), then \(K=G\).

Definition 4.2

Let G be an N-group and \(\varOmega \le _{N} G\). An ideal H of G is said to be strictly \(\varOmega \)-superfluous in G if for any N-subgroup L of G, \(\varOmega \subseteq L+H\) implies \(\varOmega \subseteq L\). We denote this by \(H\ll _{\varOmega }^{s} G\).

Example 4.3

Let \(N=\Bigg (\ \begin{pmatrix} {\mathbb {Z}}_{4} &{} 2{\mathbb {Z}}_{4} \\ 0 &{} {\mathbb {Z}}_{4} \end{pmatrix},~~ +, ~~\cdot \Bigg )\) where \({\mathbb {Z}}_{4}\) is the set of residue classes modulo 4 and \(G=N\).

N-subgroups of G are

$$\begin{aligned} H_{1}= & {} \begin{pmatrix} 0 &{} 0 \\ 0 &{} 0 \end{pmatrix}, H_{2}=\begin{pmatrix} 2{\mathbb {Z}}_{4} &{} 0 \\ 0 &{} 0 \end{pmatrix},\\ H_{3}= & {} \begin{pmatrix} 0 &{} 0 \\ 0 &{} 2{\mathbb {Z}}_{4} \end{pmatrix}, H_{4}=\begin{pmatrix} 0 &{} 2{\mathbb {Z}}_{4} \\ 0 &{} 0 \end{pmatrix}, H_{5}=\begin{pmatrix} 0 &{} 2{\mathbb {Z}}_{4} \\ 0 &{} 2{\mathbb {Z}}_{4} \end{pmatrix}, H_{6}=\begin{pmatrix} 0 &{} 2{\mathbb {Z}}_{4} \\ 0 &{} {\mathbb {Z}}_{4} \end{pmatrix},\\ H_{7}= & {} \begin{pmatrix} 2{\mathbb {Z}}_{4} &{} 2{\mathbb {Z}}_{4} \\ 0 &{} 0 \end{pmatrix}, H_{8}=\begin{pmatrix} 2{\mathbb {Z}}_{4} &{} 2{\mathbb {Z}}_{4} \\ 0 &{} 2{\mathbb {Z}}_{4} \end{pmatrix}, H_{9}=\begin{pmatrix} 2{\mathbb {Z}}_{4} &{} 2{\mathbb {Z}}_{4} \\ 0 &{} {\mathbb {Z}}_{4} \end{pmatrix}, H_{10}=\begin{pmatrix} {\mathbb {Z}}_{4} &{} 2{\mathbb {Z}}_{4} \\ 0 &{} 0 \end{pmatrix},\\ H_{11}= & {} \begin{pmatrix} {\mathbb {Z}}_{4} &{} 2{\mathbb {Z}}_{4} \\ 0 &{} 2{\mathbb {Z}}_{4} \end{pmatrix}, H_{12}=\begin{pmatrix} 2{\mathbb {Z}}_{4} &{} 0 \\ 0 &{} 2{\mathbb {Z}}_{4} \end{pmatrix}, H_{13}=\begin{pmatrix} {\mathbb {Z}}_{4} &{} 0 \\ 0 &{} 0 \end{pmatrix}, H_{14}=\begin{pmatrix} {\mathbb {Z}}_{4} &{} 2{\mathbb {Z}}_{4} \\ 0 &{} {\mathbb {Z}}_{4} \end{pmatrix}. \end{aligned}$$

Ideals are \(H_{1}, H_{2}, H_{3}, H_{4}, H_{5}, H_{6}, H_{7}, H_{8}, H_{9}, H_{10}, H_{11}, H_{12}, H_{14}\). Let \(\varOmega = H_{3}\). Then \(H_{10}\ll _{\varOmega }^{s} G\) but not strictly superfluous in G since \( H_{10}+ H_{11}=G\) and \( H_{11}\ne G\). \(H_{7}\) is not strictly \(H_{12}\)-superfluous in G as there exists \(H_{5}\) such that \(H_{12}\nsubseteq H_{5}\) and \(H_{12}\subseteq H_{7}+H_{5}\).

Example 4.4

Consider the N-group given in Example 3.6. Then \(H_{4}\ll _{H_{5}}^{s}G\), \(H_{7}\ll _{H_{5}}^{s}G\). Here \(H_{7}\) is not strictly superfluous, since \(H_{7}+H_{5}=G\) but \(H_{5}\ne G\). Also \(H_{7}\ll _{H_{9}} G\) but \(H_{7}\) is not strictly \(H_{9}\)-superfluous since there exists \(H_{5}\le _{N} G\) such that \(H_{9}\nsubseteq H_{5}\) but \(H_{9}\subseteq H_{7}+H_{5}\).

Proposition 4.5

Let \(\varOmega \le _{N}G\), \(K\trianglelefteq _{N}G\). If \(K\ll _{\varOmega }^{s}G\), then \(K\cap \varOmega \ll ^{s}G\).

Proof

Let \(K \ll _{\varOmega }^{s} G\). To prove \(K \cap \varOmega \ll ^{s} G\), let \(L \le _{N} G\) be such that \((K \cap \varOmega )+L=G\). Now \(\varOmega \subseteq (K \cap \varOmega )+L \subseteq K+L\). Since \(K \ll _{\varOmega }^{s} G\), we have that \(\varOmega \subseteq L\). Also since \(K\cap \varOmega \subseteq \varOmega \subseteq L\), it follows that \(L=(K\cap \varOmega )+L=G\). Therefore, \(K\cap \varOmega \ll ^{s} G\). \(\square \)

Proposition 4.6

Let \(P\trianglelefteq _{N} G\) and K, \(\varOmega \) be N-subgroups of G such that P and \(\varOmega \) are contained in K. Then \(P \ll _{\varOmega }^{s} K\) implies \(P\ll _{\varOmega }^{s} G\).

Proof

Suppose that \(P\ll _{\varOmega }^{s} K\). To prove \(P\ll _{\varOmega }^{s} G\), let \(L\le _{N}G\) be such that \(\varOmega \subseteq P+L\). Since \(\varOmega \subseteq K\) and by modular law, we get \(\varOmega \subseteq (P+L)\cap K=P+(L\cap K)\). Since \(L\cap K\le _{N}K\) and \(P\ll _{\varOmega }^{s} K\), we conclude that \(\varOmega \subseteq L\cap K \subseteq L\). Therefore \(P\ll _{\varOmega }^{s} G\). \(\square \)

The other implication follows when \(K=\varOmega \).

Proposition 4.7

Let \(P \trianglelefteq _{N} G\) and \(\varOmega \le _{N}G\) such that \(P\subset \varOmega \). Then \(P\ll _{\varOmega }^{s} G\) if and only if \(P\ll ^{s} \varOmega \).

Proof

Suppose \(P\ll _{\varOmega }^{s} G\). To prove \(P\ll ^{s} \varOmega \), let \(L \le _{N} \varOmega \) such that \(P+L=\varOmega \). Now \(\varOmega \subseteq P+L\) and \(P \ll _{\varOmega }^{s} G\), we get \(\varOmega \subseteq L\). Since \(L\subseteq \varOmega \), it follows that \(L=\varOmega \). \(\square \)

Proposition 4.8

Let \(N_{1}\), \(N_{2}\) be ideals of G. Let \(\varOmega \le _{N}G\) such that \(\varOmega \nsubseteq N_{1}\), \(\varOmega \nsubseteq N_{2}\). Then \(N_{1}\ll _{\varOmega }^{s} G\) and \(N_{2}\ll _{\varOmega }^{s} G\) if and only if \(N_{1}+N_{2}\ll _{\varOmega }^{s} G\).

Proof

The proof is similar to the proof of Proposition 3.13. \(\square \)

Proposition 4.9

Let N be zero-symmetric and \(\varOmega \le _{N}G\). Let K, P be ideals of G such that \(K\subseteq P\), \(K\subset \varOmega \) and \(\varOmega \nsubseteq P\). Then \(P\ll _{\varOmega }^{s} G\) if and only if \(K\ll _{\varOmega }^{s} G\) and \(\frac{P}{K}\ll _{\frac{\varOmega }{K}}^{s} \frac{G}{K}\).

Proof

The proof is similar to the proof of Proposition 3.15. \(\square \)

In Proposition 4.10 and 4.11, we assume N to be zero-symmetric, so that every ideal can also be considered as an N-group.

Proposition 4.10

Let N be zero-symmetric, \(\varOmega \) be an N-subgroup of G, and \(\{\Theta _{j}\}_{j\in J}\) be a family of ideals of G. If \(K\trianglelefteq _{N}G\) such that \(K\ll _{\varOmega }^{s}G\) and \(K\ll _{\Theta _{j}}^{s}G\) for all \(j\in J\), then \(K\ll _{ \varOmega +\sum _{j} \Theta _{j}}^{s} G\).

Proof

Let \(K\ll _{\varOmega }^{s}G\) and \(K\ll _{\Theta _{j}}^{s}G\) for all \(i\in I\) \(j\in J\). Let \(L\le _{N}G\) be such that \( \varOmega +\sum _{j} \Theta _{j} \subseteq K+L\). Now \(\varOmega \subseteq \varOmega +\sum _{j} \Theta _{j} \subseteq K+L\). Since \(K\ll _{\varOmega }^{s}G\), we get \(\varOmega \subseteq L\). Now \(\Theta _{j} \subseteq \varOmega +\sum _{j} \Theta _{j} \subseteq K+L\). Since \(K\ll _{\Theta _{j}}^{s}G\), we get \(\Theta _{j}\subseteq L\). Therefore \( \varOmega +\sum _{j} \Theta _{j}\subseteq L\). Hence \(K\ll _{ \varOmega +\sum _{j} \Theta _{j}}^{s}G\). \(\square \)

Proposition 4.11

Let N be zero-symmetric and \(K_{1}\), \(K_{2}\) be ideals of G. If \(K_{1}\ll _{K_{2}}^{s}G\) and \(K_{2}\ll _{K_{1}}^{s}G\), then \(K_{1}\cap K_{2} \ll _{K_{1}+K_{2}}^{s}G\).

Proof

Suppose \(K_{1}\ll _{K_{2}}^{s}G\) and \(K_{2}\ll _{K_{1}}^{s}G\). First we show that \(K_{1}\cap K_{2}\ll _{K_{1}}^{s} G\). For this, let \(K_{1}\subseteq (K_{1}\cap K_{2})+X\), where \(X\le _{N}G\). Then \(K_{1}\subseteq K_{2}+X\) and since \(K_{2}\ll _{K_{1}}^{s}G\) we get \(K_{1}\subseteq X\). Therefore \(K_{1}\cap K_{2}\ll _{K_{1}}^{s} G\). In a similar way, we get \(K_{1}\cap K_{2}\ll _{K_{2}}^{s} G\). Hence, by Proposition 4.10, \(K_{1}\cap K_{2} \ll _{K_{1}+K_{2}}^{s} G\). \(\square \)

Definition 4.12

Let \(G_{1}\) and \(G_{2}\) be N-groups and \(\varOmega \le _{N}G\). An N-epimorphism \(f:G_{1}\rightarrow G_{2}\) is called strictly \(\varOmega \)-superfluous if \(ker~f \ll _{\varOmega }^{s} G_{1}\).

Lemma 4.13

Let \(K\trianglelefteq _{N}G\) and \(\varOmega \le _{N} G\) be such that \(\varOmega \nsubseteq K\). Then \(K\ll _{\varOmega }^{s}G\) if and only if the natural map \(f:G\rightarrow \frac{G}{K}\) is strictly \(\varOmega \)-superfluous.

Proof

Since \(ker~f=\{g\in G:f(g)=0\in \frac{G}{K}\}=K\), the proof is clear. \(\square \)

Lemma 4.14

Let \(K\trianglelefteq _{N}G\) and \(\varOmega \le _{N} G\) be such that \(\varOmega \nsubseteq K\). Then \(K\ll _{\varOmega }^{s}G\) if and only if for every N-group \(G_{1}\) and N-homomorphism \(h:G_{1}\rightarrow G\) with \(\varOmega \subseteq K+Im~h\), \(\varOmega \subseteq Im~h\).

Proof

Suppose \(K\ll _{\varOmega }^{s}G\). Let \(G_{1}\) be an N-group and \(h:G_{1}\rightarrow G\) be an N-homomorphism with \(\varOmega \subseteq K+Im~h\). Since Im h is an N-subgroup of G and \(K\ll _{\varOmega }^{s}G\), we have \(\varOmega \subseteq Im~h\).

Conversely, suppose that \(\varOmega \subseteq K+X\) where \(X\le _{N} G\). Let \(i:X\rightarrow G\) be an inclusion map. Clearly i is an N-homomorphism, and so by hypothesis, we can conclude that \(\varOmega \subseteq X\). Therefore, \(K\ll _{\varOmega }^{s}G\). \(\square \)

Lemma 4.15

Let \(\varOmega \) be an N-subgroup and K be an ideal of G. Let \(G^{'}\) be an N-group and \(f:G\rightarrow G^{'}\) be an N-epimorphism such that \(f(\varOmega )\nsubseteq f(K)\). If \(K\ll _{\varOmega }^{s}G\), then \(f(K)\ll _{f(\varOmega )}^{s}G\). The converse holds if f is injective.

Proof

Suppose \(K\ll _{\varOmega }^{s}G\). Since f is an epimorphism, we have \(f(K)\trianglelefteq _{N}G^{'}\). Let \(X\le _{N} G^{'}\) be such that \(f(\varOmega )\subseteq f(K)+X\). Then \(\varOmega \subseteq K+f^{-1}(X)\). Since \(f^{-1}(X)\le _{N}G\) and \(K\ll _{\varOmega }^{s}G\), we have \(\varOmega \subseteq f^{-1}(X)\). Hence \(f(\varOmega )\subseteq X\).

Conversely, suppose that \(f(K)\ll _{f(\varOmega )}^{s}G^{'}\). Let \(X\le _{N}G\) be such that \(\varOmega \subseteq K+X\). Then \(f(\varOmega )\subseteq f(K+X)=f(K)+f(X)\). Since \(f(K)\ll _{f(\varOmega )}^{s}G^{'}\), we have \(f(\varOmega )\subseteq f(X)\). Therefore, \(f^{-1}(f(\varOmega ))\subseteq f^{-1}(f(X))\) which implies \(\varOmega +ker~f \subseteq X+ker~f\). Since f is injective, we get \(\varOmega \subseteq X\). \(\square \)

Example 4.16

Consider the Example 3.21. Then it can be seen that

\(H_{9}\ll _{H_{7}}^{s}G\), but \(f(H_{9})\) is not strictly \(f(H_{7})\) superfluous in G, since

\(f(H_{9})\) is not an ideal of G.

Definition 4.17

Let N be zero-symmetric nearring. Let \(\varOmega \le _{N} G\) and \(H\trianglelefteq _{N}G\) be such that \(\varOmega \nsubseteq H\). An N-subgroup K of G is said to be a strictly \(\varOmega \)-supplement of H if \(\varOmega \subseteq H+K\) and \(\varOmega \nsubseteq H+K^{'}\) for any ideal \(K^{'}\) of K.

The following remark is a straightforward observation.

Remark 4.18

  1. 1.

    If N is zero-symmetric and \(\varOmega =G\), then every strictly \(\varOmega \)-supplement is a supplement (defined by [20]).

  2. 2.

    Let N be zero-symmetric. Let \(H\trianglelefteq _{N} G\) be such that \(\varOmega \nsubseteq H\). Then every \(\varOmega \)-supplement of H is a strictly \(\varOmega \)-supplement of H.

  3. 3.

    If N is zero-symmetric and \(\varOmega =G\), then every strictly \(\varOmega \)-supplement is a supplement.

5 Superfluous ideals of \(M_{n}(N)\)-group \(N^{n}\)

For a zero-symmetric right nearring N with 1, let \(N^{n}\) be the direct sum of n copies of \((N,+)\). The elements of \(N^{n}\) are column vectors and written as \(\left( r_1,\cdots ,r_n\right) \). The symbols \(i_{j}\) and \(\pi _{j}\) respectively, denote the \(i^{th}\) coordinate injective and \(j^{th}\) coordinate projective maps.

For an element \(a\in N\), \(i_{i}(a)=(0,\cdots ,\underbrace{a}_{i^{th}},\cdots ,0)\), and \(\pi _{j}(a_1,\cdots ,a_n)=a_{j}\), for any \((a_1,\cdots ,a_n)\in N^n\). The nearring of \(n\times n\) matrices over N, denoted by \(M_n(N)\), is defined to be the subnearring of \(M(N^{n})\), generated by the set of functions {\(f_{ij}^{a}:N^{n}\rightarrow N^{n}~|~a\in N,1\le i,j\le n\)} where \(f_{ij}^{a}\left( k_1,\cdots ,k_n\right) :=\left( l_1,l_2,\cdots ,l_n\right) \) with \(l_i=ak_j\) and \(l_p=0\) if \(p\ne i\). Clearly, \(f_{ij}^{a}=i_{i}f^{a}\pi _{j}\), where \(f^{a}(x)=ax\), for all \(a,x\in N\). If N happens to be a ring, then \(f_{ij}^{a}\) corresponds to the \(n\times n\)-matrix with a in position (ij) and zeros elsewhere.

Notation 5.1

([9], Notation 1.1)

For any ideal \({\mathcal {A}}\) of \(M_n(N)\)-group \(N^n\), we write

$$\begin{aligned} {\mathcal {A}}_{**}=\{a\in N:a=\pi _{j}A,~\text{ for } \text{ some }~A\in {\mathcal {A}},1\le j\le n\}, { \text{ an } \text{ ideal } \text{ of } _{N}{N}}. \end{aligned}$$

We denote \(M_n(N)\) for a matrix nearring, \(N^n\) for an \(M_n(N)\)-group \(N^n\). We refer to Meldrum & Van der Walt [15] for preliminary results on matrix nearrings.

From [10], for any \(s\in G\), the ideal generated by s is denoted by \(\langle s \rangle \) and defined as, \(\langle s \rangle =\displaystyle \bigcup _{i=1}^{\infty }U_{i+1}\), where \(U_{i+1}=U_{i}^{*}\cup U_{i}^{0}\cup U_{i}^{+}\) with \(U_{0}=\{s\}\), and \(U_{i}^{*}=\{g+y-g: g\in G, y\in U_{i}\},\) \(U_{i}^{0}=\{p-q: p,q\in U_{i}\}\cup \{p+q:p,q\in U_{i}\},\) \(U_{i}^{+}=\{n(g+a)-ng: n\in N, g\in G, a\in U_{i}\}.\)

Theorem 5.2

(Theorem 1.4 of [9]) Suppose \(A\subseteq N\).

  1. 1.

    If \(A^{n}\) is an ideal of \({}_{M_{n}(N)}{N^{n}}\), then \(A=(A^{n})_{\star \star }\).

  2. 2.

    If A is an ideal of \({}_{N}{N}\) if and only if \(A^{n}\) is an ideal of \({}_{M_{n}(N)}{N^{n}}\).

  3. 3.

    If A is an ideal of \({}_{N}{N}\), then \(A=(A^{n})_{\star \star }\).

Lemma 5.3

(Lemma 1.5 of [9])

  1. 1.

    If \({\mathcal {I}}\) is an ideal of \({}_{M_{n}(N)}{N^{n}}\), then \(({\mathcal {I}}_{\star \star })^{n}={\mathcal {I}}\).

  2. 2.

    Every ideal \({\mathcal {I}}\) of \({}_{M_{n}(N)}{N^{n}}\) is of the form \(K^{n}\) for some ideal K of \({}_{N}{N}\).

Note 5.4

(Note 1.7(iii) of [9]) Let A be an ideal of \({}_{N}^{N}\). Then \(A\le _{e} {}_{N}{N}\) if and only if \(A^{n}\le _{e} {}_{M_{n}(N)}{N^{n}}\).

Theorem 5.5

(Theorem 1.9 [9]) If \(l \in N\), then \(\langle l\rangle ^{n}=\langle (l,0,\cdots ,0)\rangle \).

Lemma 5.6

If I and J are ideals of N, then \((I+J)^{n}=I^{n}+J^{n}\).

Proof

Clearly, \(I\subseteq I+J\) and \(J\subseteq I+J\) which implies \(I^{n}\subseteq (I+J)^{n}\) and \(J^{n}\subseteq (I+J)^{n}\) and so \(I^{n}+J^{n}\subseteq (I+J)^{n}\). To prove the other part, let \((x_{1},x_{2},\cdots ,x_{n})\in (I+J)^{n}\). Then \(x_{i}\in I+J\) for every \(1\le i\le n\) which implies \(x_{i}=a_{i}+b_{i}\), where \(a_{i}\in I\) and \(b_{i}\in J\).

Now,

$$\begin{aligned} (x_{1},x_{2},\cdots ,x_{n})&=(a_{1}+b_{1}, a_{2}+b_{2},\cdots ,a_{n}+b_{n}) \\&=(a_{1},a_{2},\cdots ,a_{n})+(b_{1},b_{2},\cdots ,b_{n})\\&\in I^{n}+J^{n} \end{aligned}$$

Therefore, \((I+J)^{n}\subseteq I^{n}+J^{n}\). Hence, \((I+J)^{n}=I^{n}+J^{n}\). \(\square \)

Lemma 5.7

\(I+J=G\) if and only if \((I+J)^{n}=G^{n}\) if and only if \(I^{n}+J^{n}=G^{n}\).

Definition 5.8

An ideal \({\mathcal {A}}\) of \(M_{n}(N)\)-group \(N^{n}\) is said to be superfluous if for any ideal \({\mathcal {K}}\) of \(N^{n}\), \({\mathcal {A}}+{\mathcal {K}}=N^{n}\) implies \({\mathcal {K}}=N^{n}\).

Lemma 5.9

Let B be an ideal of \({}_{N}{N}\). If \(B\ll {}_{N}{N}\), then \(B^{n}\ll {}_{M_{n}(N)}{N^{n}}\).

Proof

Let \({\mathcal {A}}\trianglelefteq {}_{M_{n}(N)}{N^{n}}\) such that \(B^{n}+{\mathcal {A}} =N^{n}\). To show \({\mathcal {A}}=N^{n}\). Since \({\mathcal {A}}\trianglelefteq {}_{M_{n}(N)}{N^{n}}\), by Lemma 5.3, we have \({\mathcal {A}}=({\mathcal {A}}_{\star \star })^{n}\), which implies \(B^{n}+({\mathcal {A}}_{\star \star })^{n}=N^{n}\). Now using Lemma 5.6, we get \((B+{\mathcal {A}}_{\star \star })^{n}=N^{n}\). Therefore, by Lemma 5.7, \(B+{\mathcal {A}}_{\star \star }=N\). Since, \(B\ll {}_{N}{N}\), we get \({\mathcal {A}}_{\star \star }=N\). Hence, \({\mathcal {A}}=({\mathcal {A}}_{\star \star })^{n}=N^{n}\). \(\square \)

Lemma 5.10

If \({\mathcal {A}}\ll {}_{M_{n}(N)}{N^{n}}\), then \({\mathcal {A}}_{\star \star }\ll {}_{N}{N}\).

Proof

Let \(B\trianglelefteq {}_{N}{N}\) such that \({\mathcal {A}}_{\star \star }+B=N\). By Lemma 5.7, we have \(({\mathcal {A}}_{\star \star }+B)^{n}=N^{n}\). By Lemma 5.6, we have \(({\mathcal {A}}_{\star \star })^{n}+B^n=N^{n}\) which implies \({\mathcal {A}}+B^{n}=N^{n}\). Since \(B^{n}\trianglelefteq {}_{M_{n}(N)}{N^{n}}\) and \({\mathcal {A}}\ll {}_{M_{n}(N)}{N^{n}}\), we have \(B^{n}=N^{n}\). Let \(n\in N\). Then \((n,0,\cdots ,0)\in N^{n}=B^{n}\). Therefore, \(n\in (B^{n})_{\star \star }=B\) (by Theorem 5.2(3)). Therefore, \(B=N\). \(\square \)

Theorem 5.11

There is a one-one correspondence between the set of superfluous ideals of \({}_{N}{N}\) and those of \(M_{n}(N)\)-group \(N^{n}\).

Proof

Let \(P=\{A\trianglelefteq {}_{N}{N}: A\ll {}_{N}{N} \}\). \(Q=\{{\mathcal {A}}\trianglelefteq {}_{M_{n}(N)}{N^{n}}:{\mathcal {A}}\ll {}_{M_{n}(N)}{N^{n}} \}\). Define \(\Phi :P\rightarrow Q\) by \(\Phi (A)=A^{n}\). Then by Lemma 5.9, \(A^{n}\ll {}_{M_{n}(N)}{N^{n}}\). Define \(\Psi :Q\rightarrow P\) by \(\Psi ({\mathcal {A}})={\mathcal {A}}_{\star \star }\). By Lemma 5.10, \({\mathcal {A}}_{\star \star }\ll {}_{N}{N}\). Now \((\Psi \circ \Phi )(A)=\Psi (\Phi (A))=\Psi (A^{n})=(A^{n})_{\star \star }=A\). Therefore, \((\Psi \circ \Phi )=Id_{P}\). Also, \((\Phi \circ \Psi )({\mathcal {A}})=\Phi (\Psi ({\mathcal {A}}))=\Phi ({\mathcal {A}}_{\star \star })=({\mathcal {A}}_{\star \star })^{n}={\mathcal {A}}\), and hence \((\Phi \circ \Psi )=Id_{Q}\). \(\square \)

Definition 5.12

An ideal \({\mathcal {K}}\) of \(M_{n}(N)\)-group \(N^{n}\) is said to be g-superfluous if for any ideal \({\mathcal {A}}\) of \(N^{n}\), \({\mathcal {K}}+{\mathcal {A}}=N^{n}\) and \({\mathcal {A}}\le _{e} N^{n}\) implies \({\mathcal {K}}=N^{n}\).

Lemma 5.13

Let I be an ideal of \({}_{N}{N}\). If \(I\ll _{gs} {}_{N}{N}\), then \(I^{n}\ll _{gs}{}_{{}_{M_{n}(N)}}{N^{n}}\).

Proof

Let \(I\ll _{gs} {}_{N}{N}\). To show \(I^{n}\ll _{gs}{}_{M_{n}(N)}{N^{n}}\), let \({\mathcal {K}}\) be an ideal of \({}_{M_{n}(N)}{N^{n}}\) such that \(I^{n}+{\mathcal {K}}={}_{M_{n}(N)}{N^{n}}\) and \({\mathcal {K}}\le _{e} {}_{M_{n}(N)}{N^{n}}\). Since \({\mathcal {K}}\trianglelefteq {}_{M_{n}(N)}{N^{n}}\), by Lemma 5.3(2), we have \({\mathcal {K}}=A^{n}\) for some ideal A of \({}_{N}{N}\). Since \({\mathcal {K}}=A^{n}\le _{e} {}_{M_{n}(N)}{N^{n}}\), by Note 5.4, we have \(A\le _{e}{}_{N}{N}\). Now, \(I^{n}+{\mathcal {K}}=I^{n}+A^{n}=(I+A)^{n}=N^{n}\) which implies \(I+A=N\). Since, \(I\ll _{gs} {}_{N}{N}\), we get \(A=N\). Therefore, \({\mathcal {K}}=A^{n}=N^{n}\). Hence, \(I^{n}\ll _{gs} {}_{M_{n}(N)}{N^{n}}\). \(\square \)

Lemma 5.14

If \({\mathcal {A}}\ll _{gs} {}_{M_{n}(N)}{N^{n}}\), then \({\mathcal {A}}_{\star \star }\ll _{gs} {}_{N}{N}\).

Proof

Let \({\mathcal {A}}\ll _{gs} {}_{M_{n}(N)}{N^{n}}\). To show \({\mathcal {A}}_{\star \star }\ll _{gs} {}_{N}{N}\), let \(B\le _{e}{}_{N}{N}\) such that \({\mathcal {A}}_{\star \star }+B={}_{N}{N}\). Since \(B\le _{e}{}_{N}{N}\), by Note 5.4, we have \(B^{n}\le _{e} {}_{M_{n}(N)}{N^{n}}\). Now, \({\mathcal {A}}_{\star \star }+B=N\) implies \(({\mathcal {A}}_{\star \star }+B)^{n}=N^{n}\). By Lemma 5.6, we get \(({\mathcal {A}}_{\star \star })^{n}+B^{n}=N^{n}\). Therefore, \({\mathcal {A}}+B^{n}=N^{n}\). Since \({\mathcal {A}}\ll _{gs} N^{n}\), we get \(B^{n}=N^{n}\). Now, by Theorem 5.2(3), we get \(B=(B^{n})_{\star \star }=(N^{n})_{\star \star }=N\). Therefore, \({mathcal{A}}_{\star \star }\ll _{gs} {}_{N}{N}\). \(\square \)

Theorem 5.15

There is a one-one correspondence between the set of g-superfluous ideals of \({}_{N}^{N}\) and those of \(M_{n}(N)\)-group \(N^{n}\).

Proof

Let \(P=\{A\trianglelefteq {}_{N}{N}: A\ll _{gs}{}_{N}{N}\}\). \(Q=\{{\mathcal {A}}\trianglelefteq {}_{M_{n}(N)}{N^{n}}:{\mathcal {A}}\ll _{gs}{}_{M_{n}(N)}{N^{n}} \}\). Define \(\Phi :P\rightarrow Q\) by \(\Phi (A)=A^{n}\). Then by Lemma 5.13, \(A^{n}\ll _{gs}{}_{M_{n}(N)}{N^{n}}\). Define \(\Psi :Q\rightarrow P\) by \(\Psi ({\mathcal {A}})={\mathcal {A}}_{\star \star }\). By Lemma 5.14, \({\mathcal {A}}_{\star \star }\ll _{gs}{}_{N}{N}\). Now \((\Psi \circ \Phi )(A)=\Psi (\Phi (A))=\Psi (A^{n})=(A^{n})_{\star \star }=A\). \((\Phi \circ \Psi )({\mathcal {A}})=\Phi (\Psi ({\mathcal {A}}))=\Phi ({\mathcal {A}}_{\star \star })=({\mathcal {A}}_{\star \star })^{n}={\mathcal {A}}\). Therefore, \((\Psi \circ \Phi )=Id_{P}\) and \((\Phi \circ \Psi )=Id_{Q}\). \(\square \)

Definition 5.16

An element \(s\in G\) is called hollow if \(\langle s\rangle \) is a hollow ideal of G. In this case we call s as an h-element of G.

Example 5.17

  1. 1.

    Let \(N=({\mathbb {Z}}_{12}, +_{{\mathbb {Z}}_{12}}, \cdot _{{\mathbb {Z}}_{12}})\) and \(G=N\). Then \(\langle 3\rangle \) is hollow. Therefore, 3 is a hollow element.

  2. 2.

    Let \(N=(D_{8},+,\cdot )\) given in Example 2.11 and \(G=N\). Then \(\langle r^{2}\rangle \) is hollow. Therefore, \(r^{2}\) is a hollow element.

Proposition 5.18

s is a hollow element of \({}_{N}{N}\) if and only if \((s,0,0,\cdots ,0)\) is a hollow element in \(M_{n}(N)\)-group \(N^{n}\).

Proof

Suppose s is a hollow element then \(\langle s\rangle \) is a hollow ideal. To show \(\langle (s,0,\cdots ,0)\rangle \) is a hollow ideal of \(M_{n}(N)\)-group \(N^{n}\), let \({\mathcal {I}}\), \({\mathcal {J}}\) be ideals of \(\langle (s,0,\cdots ,0)\rangle \) such that \({\mathcal {I}}+{\mathcal {J}}=\langle (s,0,\cdots ,0)\rangle \). Then by Lemma 5.3(1), we have \({\mathcal {I}}=({\mathcal {I}}_{\star \star })^{n}\), \({\mathcal {J}}=({\mathcal {J}}_{\star \star })^{n}\), which implies \(({\mathcal {I}}_{\star \star })^{n}+({\mathcal {J}}_{\star \star })^{n}=\langle (s,0,\cdots ,0)\rangle \). Using Lemma 5.6 and by Theorem 5.5, we get \(({\mathcal {I}}_{\star \star }+{\mathcal {J}}_{\star \star })^{n}=\langle (s,0,\cdots ,0)\rangle =\langle s\rangle ^{n}\) and so \({\mathcal {I}}_{\star \star }+{\mathcal {J}}_{\star \star }=\langle s\rangle \). Since, \(\langle s\rangle \) is hollow, we get either \({\mathcal {I}}_{\star \star }=\langle s\rangle \) or \({\mathcal {J}}_{\star \star }=\langle s\rangle \).

Therefore,

$$\begin{aligned} \mathcal {{\mathcal {I}}}=({\mathcal {I}}_{\star \star })^{n}=\langle s\rangle ^{n}=\langle (s,0,\cdots ,0)\rangle \end{aligned}$$

or

$$\begin{aligned} {\mathcal {J}}=({\mathcal {J}}_{\star \star })^{n}=\langle s\rangle ^{n}=\langle (s,0,\cdots ,0)\rangle . \end{aligned}$$

Conversely, suppose \((s,0,\cdots ,0)\) is hollow in \(N^{n}\). Then \(\langle (s,0,\cdots ,0)\rangle \) is a hollow ideal of \({}_{M_{n}(N)}^{N^{n}}\), which implies \(\langle s\rangle ^{n}\) is a hollow ideal of \({}_{M_{n}(N)}^{N^{n}}\). To show \(\langle s\rangle \) is hollow in N, let I and J be two ideals of N contained in \(\langle s\rangle \) such that \(I+J=\langle s\rangle \). Now, \((I+J)^{n}=\langle s\rangle ^{n}\). Therefore \(I^{n}+J^{n}=\langle s\rangle ^{n}\). Since \(\langle s\rangle ^{n}\) is hollow, we have \(I^{n}=\langle s\rangle ^{n}\) or \(J^{n}=\langle s\rangle ^{n}\), and hence, \(I=\langle s\rangle \) or \(J=\langle s\rangle \). \(\square \)

Definition 5.19

\(X=\{x_{1},x_{2},\cdots ,x_{n}\}\subseteq G\) is said to be a spanning set for G if \(\sum _{x_{i}\in X} \langle x_{i}\rangle =G\). If \(\{x_{i}: 1\le i\le n\}\) is a spanning set in G, then we say the elements \(x_{i},~ 1\le i\le n\) are spanning elements in G.

Theorem 5.20

\(\{x_{i}: 1\le i\le n\}\) is a spanning set in \({}_{N}{N}\) if and only if \(\{(x_{i},0,\cdots ,0): 1\le i\le n\}\) is a spanning set in \(M_{n}(N)\)-group \(N^{n}\).

Proof

Suppose \(\{x_{i}: 1\le i\le n\}\) is a spanning set in \({}_{N}{N}\). Then

$$\begin{aligned} \sum _{1\le i\le n} \langle x_{i}\rangle =N&\Leftrightarrow \langle x_{1}\rangle +\langle x_{2}\rangle +\cdots +\langle x_{n}\rangle =N\\&\Leftrightarrow (\langle x_{1}\rangle +\langle x_{2}\rangle +\cdots +\langle x_{n}\rangle )^{n}=N^{n}\\&\Leftrightarrow \langle x_{1}\rangle ^{n}+\langle x_{2}\rangle ^{n}+\cdots +\langle x_{n}\rangle ^{n}=N^{n}\\&\Leftrightarrow \langle (x_{1},\cdots ,0)\rangle +\langle (x_{2},0,\cdots ,0)\rangle +\cdots +\langle (x_{n},0,\cdots ,0)\rangle =N^{n}\\&\Leftrightarrow \sum _{1\le i\le n}\langle (x_{i},0,\cdots ,0)\rangle =N^{n}. \end{aligned}$$

Therefore \(\{(x_{i},0,\cdots ,0): 1\le i\le n\}\) is a spanning set in \(M_{n}(N)\)-group \(N^{n}\). \(\square \)

Definition 5.21

A subset X of G is said to be a h-spanning set if every element of X is a h-element and X is a spanning set.

Theorem 5.22

Suppose \(x_{1}, x_{2},\cdots ,x_{n}\in N\). Then \(\{x_{i}:1\le i\le n\}\) is a h-spanning set in N if and only if \(\{(x_{i},0,\cdots ,0):1\le i\le n\}\) is a h-spanning set in \(M_{n}(N)\)-group \(N^{n}\).

Proof

\(\{x_{i}:1\le i\le n\}\) is a h-spanning set.

$$\begin{aligned} &\Leftrightarrow x_{i},~1\le i\le n \text { are } h\text {-elements and } \sum _{1\le i\le n}\langle x_{i}\rangle =N\\&\Leftrightarrow (x_{i},0,\cdots ,0), ~1\le i\le n \text { are} ~h\text {-elements in } {}_{M_{n}(N)}{N^{n}}\\ {}&~~~~ \text { and } \sum _{1\le i\le n}\langle (x_{i},0,\cdots ,0)\rangle =N^{n}. \end{aligned}$$

Therefore \(\{(x_{i},0,\cdots ,0):1\le i\le n\}\) is a h-spanning set in \(M_{n}(N)\)-group \(N^{n}\). \(\square \)

6 Superfluous ideal graph of a nearring

The authors [22] studied graphs with respect to superfluous elements in a lattice, and in [21] the authors studied the graphs with respect to the dual aspects such as essential elements, complements, etc. Lattice aspects of modules over rings are well-known due to [3, 7]. In this section, we define the superfluous ideal graph of a nearring and study some of its properties.

Definition 6.1

Let N be a nearring. An ideal I of N is said to be superfluous if for any ideal J of N, \(I+J=N\) implies \(J=N\).

Definition 6.2

The superfluous ideal graph of N, denoted by \(S_{N}(G)\), is a graph having set of all non-zero proper ideals of N as vertices and two vertices I and J are adjacent if \(I\cap J\ll N\).

Example 6.3

  1. 1.

    If N is simple, then \(S_{N}(G)\) is a null graph.

  2. 2.

    Suppose N is a finitely generated nearring which contains only one non-zero maximal ideal, then every proper ideal of N is superfluous. The vertices of \(S_{N}(G)\) are the non-zero proper ideals of N. Since every proper ideal of N is superfluous in N, we have \(I\cap J\ll G\) for all proper ideals I, J of N. Therefore \(S_{N}(G)\) is a complete graph.

    For example, let \(N=(Z_{p^{n}}, +_{p^{n}}, \cdot )\) where p is prime. Then possible ideals are of the form \(\langle p^{i}\rangle ,~~i\in \{0,1,\cdots ,n-1\}\). If N is simple, then \(S_{N}(G)\) is a null graph. If N is not simple, then N has only one non-zero maximal ideal of the form \(\langle p^{k}\rangle \) for some \(0\le k\le n-1\). Hence, N is a local nearring. In this case we get a complete graph.

    Consider \(R=(Z_{p^{n}}, +_{p^{n}}, \cdot _{p^{n}})\) where addition and multiplication are modulo \(p^{n}\). Then R is a ring. In this case, the superfluous ideal graph is a complete graph with \((n-1)\) vertices.

Example 6.4

Let \(N=({\mathbb {Z}}_{6}, +_{{\mathbb {Z}}_{6}}, \cdot _{{\mathbb {Z}}_{6}})\). Then \(V(S_{{\mathbb {Z}}_{6}}(G))=\{2{\mathbb {Z}}_{6}, 3{\mathbb {Z}}_{6}\}\). Now \(2{\mathbb {Z}}_{6}\cap 3{\mathbb {Z}}_{6} =(0)\ll N\). The graph \(S_{{\mathbb {Z}}_{6}}(G)\) is shown in Fig. 1.

Fig. 1
figure 1

\(S_{\mathbb {Z}_{6}}(G)\)

Fig. 2
figure 2

\(S_{\mathbb {Z}_{12}}(G)\)

Example 6.5

Let \(N=({\mathbb {Z}}_{12}, +_{{\mathbb {Z}}_{12}}, \cdot _{{\mathbb {Z}}_{12}})\). Non-zero proper ideals of N are \(2{\mathbb {Z}}_{12}\), \(3{\mathbb {Z}}_{12}\), \(4{\mathbb {Z}}_{12}\), \(6{\mathbb {Z}}_{12}\) and \(6{\mathbb {Z}}_{12}\) is superfluous in \({\mathbb {Z}}_{12}\). Then the corresponding superfluous ideal graph is given in Fig. 2.

Example 6.6

Let \(N=({\mathbb {Z}}_{2}\times {\mathbb {Z}}_{2}, +_{{\mathbb {Z}}_{2}}, \cdot _{{\mathbb {Z}}_{2}})\) where addition and multiplication are carried out component-wise. All non-zero proper ideals are of N are \((0)\times {\mathbb {Z}}_{2}\) and \({\mathbb {Z}}_{2}\times (0)\) and \(((0)\times {\mathbb {Z}}_{2}) \cap ({\mathbb {Z}}_{2}\times (0))=(0)\ll N\). Therefore, the superfluous ideal graph is given in Fig. 3.

Example 6.7

Let \(N=({\mathbb {Z}}_{4}\times {\mathbb {Z}}_{2}, +, \cdot )\) where addition and multiplication are carried out component-wise with the first component modulo 4 and the second component modulo 2. Then the nontrivial ideals are \(I_{1}=\{(0,0),(1,0),(2,0),(3,0)\}\), \(I_{2}=\{(0,0),(2,0),(0,1),(2,1)\}\), \(I_{3}=\{(0,0),(0,1)\}\), \(I_{4}=\{(0,0),(2,0)\}\) and \(I_{4}\) is a superfluous ideal. The corresponding superfluous ideal graph is given in Fig. 4.

Example 6.8

Consider the nearring given in the Example 2.11. The ideals of N are \(H_{4}\), \(H_{9}\), \(H_{7}\) and it can be seen that \(H_{4}\) is superfluous in N. We have \(H_{9}\cap H_{7}=H_{4}\), \(H_{9}\cap H_{4}=H_{4}\) and \(H_{7}\cap H_{4}=H_{4}\). Hence, we get a complete graph given in Fig. 5.

Proposition 6.9

Every non-zero superfluous ideal of N is a universal vertex in \(S_{N}(G)\).

Proof

Let X be a non-zero superfluous ideal of N. To prove \(XY \in E\) for every \(Y \in V\). Let \(Y \in V\). By Lemma 3.7, \(X\cap Y \ll N\) which implies \(XY \in E\). Since Y is arbitrary, X is a universal vertex. \(\square \)

The converse of the Proposition 6.9 need not be true. We justify this in the following example.

Example 6.10

In Example 6.5, \(N=({\mathbb {Z}}_{12}, +_{{\mathbb {Z}}_{12}}, \cdot _{{\mathbb {Z}}_{12}})\). Then \(6{\mathbb {Z}}_{12}\) is a non-zero superfluous ideal which is a universal vertex in the corresponding superfluous ideal graph given in Fig. 2. The vertex \(3{\mathbb {Z}}_{12}\) is universal but it is not superfluous, as \(3{\mathbb {Z}}_{12}+2{\mathbb {Z}}_{12}={\mathbb {Z}}_{12}\), and \(2{\mathbb {Z}}_{12} \ne {\mathbb {Z}}_{12}\).

Proposition 6.11

The subgraph \(S_{N}[min(N)]\) induced by min(N) is a clique, where min(N) is the set of minimal ideals of N.

Fig. 3
figure 3

\(S_{\mathbb {Z}_{2}\times \mathbb {Z}_{2}}(G)\)

Fig. 4
figure 4

\(S_{\mathbb {Z}_{4}\times \mathbb {Z}_{2}}(G)\)

Fig. 5
figure 5

Superfluous ideal graph of the nearring in Example 2.11

Proof

Case 1: Suppose N has exactly one minimal ideal. Then we get a clique \(K_{1}\).

Case 2: Suppose N has more than one minimal ideal. Let \(M_{1}\), \(M_{2}\) be two arbitrary minimal ideals of N. We prove that \(M_{1}M_{2}\in E(S_{N}(G))\). Since, \(M_{1}\) and \(M_{2}\) both are minimal \(M_{1}\cap M_{2}=(0)\), which is a superfluous ideal of N, which implies \(M_{1}M_{2}\in E(S_{N}(G))\). Since \(M_{1}\) and \(M_{2}\) are arbitrary, we conclude that there exists an edge between any two minimal ideals. Therefore \(S_{N}[min(N)]\) is a clique (Fig. 3). \(\square \)

Proposition 6.12

\(S_{N}(G)\) is an empty graph if and only if N has exactly one non-zero proper ideal.

Proof

If N has exactly one non-zero proper ideal then \(S_{N}(G)=K_{1}\). Conversely, suppose \(S_{N}(G)\) is an empty graph. We prove that N has exactly one non-zero proper ideal. First we prove that N has exactly one minimal ideal. Suppose on a contrary, N has two minimal ideals \(M_{1}\) and \(M_{2}\). Then by Proposition 6.11, \(M_{1}\) and \(M_{2}\) are adjacent in \(S_{N}(G)\), a contradiction since \(S_{N}(G)\) is an empty graph. Therefore N has a unique minimal ideal say, M. So every non-zero ideal of N different from M contains M. Therefore M is superfluous. We claim that M is the only unique proper ideal of N. On a contrary, suppose \(I\ne M\) be a non-zero proper ideal of N. Then \(M\subseteq I\), \(M\cap I =M\), which is superfluous in N, and we get \(MI\in E(S_{N}(G))\), a contradiction, since \(S_{N}(G)\) is an empty graph. Therefore M is the unique non-zero proper ideal of N (Fig. 4). \(\square \)

Definition 6.13

Let I be an ideal of N. The dual annihilator of I, denoted as \(ann_{d}(I)\) is the intersection of all ideals J of N such that \(I+J=N\). That is, \(ann_{d}(I)= \bigcap \limits _{J\trianglelefteq _{N} N,~ I+J=N}J\) (Fig. 5).

Example 6.14

  1. 1.

    In the nearring given in Example 2.11, the ideals of N are \(H_{7}\), \(H_{9}\), \(H_{5}\) and \(\{e\}\). Therefore \(ann_{d}(H_{7})=\cap \{H_{9}, N\}=H_{9}\).

  2. 2.

    In the nearring N given in Example 3.27, the ideals of N are N, \(I_{2}\), \(I_{3}\), \(I_{4}\), \(I_{5}\) and \(\{e\}\). We have \(I_{2}+I_{3}=N\) and \(I_{2}+I_{4}=N\). Therefore \(ann_{d}(I_{2})= \bigcap \{I_{3},I_{4}\}=I_{5}\).

Proposition 1

Let I be any arbitrary ideal of N. Then \(I\cap (ann_{d}(I))\ll N\).

Proof

Let \(K\trianglelefteq N\) such that \(I\cap (ann_{d}(I))+K=N\). Since \(I\cap (ann_{d}(I)) \subseteq I\), we have \(I+K=N\), which implies \(ann_{d}(I)\subseteq K\) and so \(I\cap (ann_{d}(I))\subseteq K\). Now \(K= K+I\cap (ann_{d}(I))=N\). Therefore, \(I\cap (ann_{d}(I))\ll N\). \(\square \)

7 Conclusion

We have defined the notions superfluous, strictly superfluous (with respect to an ideal \(\varOmega \)), generalised superfluous, generalised suppements in N-groups. We have proved some properties and exhibited examples which are different from the existing notions. We have defined graph on superfluous ideals of a nearring, and gave some properties. The concepts can be extended to study various finite spanning dimension aspects and related chain conditions in N-groups and those of \(M_{n}(N)\)-group \(N^{n}\).