Keywords

1 Introduction

All the graphs considered in the paper are finite, simple and undirected. Let \(G=(V(G),E(G))\) be a graph. For \(v\in V(G)\) and \(S\subseteq V(G)\), let \(N_G(v)\) denote the open neighborhood of v in G and \(\left\langle S \right\rangle \) denote the subgraph induced by S. Let \(\overline{G}\) denote the complement of a graph G. A proper k-coloring of a graph G is a function from V(G) into a set of k colors such that no two adjacent vertices receive the same color. The chromatic number of a graph G, denoted by \(\chi (G)\), is the least positive integer k such that there exists a proper k-coloring of G. A clique in a graph G is a complete subgraph of G. The clique number of G is the largest size of a clique in G and it is denoted by \(\omega (G)\). Let G be a graph with \(V(G)=\{u_1,u_2,\ldots , u_n\}\) and \(H_1,H_2,\ldots , H_n\) be pairwise disjoint graphs. The G-generalized join graph, denoted by \(G[H_1,H_2,\ldots , H_n]\), of \(H_1,H_2,\ldots , H_n\) is the graph obtained by replacing each vertex \(u_i\) of G by \(H_i\) and joining each vertex of \(H_i\) to each vertex of \(H_j\) by an edge if \(u_i\) is adjacent to \(u_j\) in G. If \(H_i\cong H\), for \(1\le i\le n\), then \(G[H_1,H_2,\ldots , H_n]\) becomes the standard lexicographic product G[H].

For a graph G, we define a relation \(\mathtt {\sim }_G\) on V(G) as follows: For any \(x, y \in V(G)\), define \(x \mathtt {\sim }_G\ y\) if and only if \(N_G(x) = N_G(y)\). Clearly, \(\mathtt {\sim }_G\) is an equivalence relation on V(G). Let [x] be the equivalence class which contains x and S be the set of all equivalence classes of this relation \(\mathtt {\sim }_G\). Based on this equivalence classes we define the reduced graph \(G_r\) of a graph G as follows. The reduced graph \(G_r\) of G (defined in [13]) is the graph with vertex set \(V(G_r)=S\) and two distinct vertices [x] and [y] are adjacent in \(G_r\) if and only if x and y are adjacent in G.

Note that if \(V(G_r)=\big \{[x_1],[x_2],\ldots , [x_k]\big \}\), then G is the \(G_r\)-generalized join of \(\left\langle [x_1]\right\rangle ,\left\langle [x_2]\right\rangle ,\ldots , \left\langle [x_k]\right\rangle \), that is, \(G=G_r\big [\left\langle [x_1]\right\rangle ,\left\langle [x_2]\right\rangle ,\ldots , \left\langle [x_k]\right\rangle \big ]\) and each \([x_i]\) is an independent subset of G (that is, \(\langle [x_i]\rangle \) has no edge). Clearly, \(G_r\) is isomorphic to an induced subgraph of G. It is easy to observe the following observation.

Observation 1. If \(G_r\) is the reduced graph of G with \(\omega (G_r)=\chi (G_r)\), then \(\chi (G)=\omega (G_r)\).

Let G be a graph with \(\omega (G)=k\), and let \(\varDelta _k(G)\) be the set of all the vertices of a graph G which lie in some clique of size k of G. A connected graph G is called a generalized complete k-partite graph (see [13]) if the vertex set V(G) of G is a disjoint union of A and H satisfying the following conditions:

  1. (1)

    \(A = \varDelta _k(G)\) and the subgraph induced by A is a complete k-partite graph with parts, say, \(A_i, i = 1, 2,\ldots , k\).

  2. (2)

    For any \(h\in H\) and \(i\in \{1,2,\ldots , k\}\), h is adjacent to some vertex of \(A_i\) if and only if h is adjacent to any vertex of \(A_i\). Set \(W(h) = \{1\le i \le k\ | N(h)\cap A_i\ne \emptyset \}\) for any \(h\in H\).

  3. (3)

    For any \(h_1, h_2 \in H, h_1\) is adjacent to \(h_2\) if and only if \(W(h_1)\cup W(h_2) = \{1,2,\ldots , k\}.\)

A graph G is called a compact graph (see [13]) if G contains no isolated vertices and for each pair xy of non-adjacent vertices of G, there is a vertex z in G with \(N(x)\cup N(y)\subseteq N(z)\). A graph G is said to be k-compact if it is compact and \(\omega (G)=k\).

Throughout this paper, rings are finite non-zero commutative rings with unity. Let R be a ring. A non-zero element x of R is said to be a zero-divisor if there exists a non-zero element y of R such that \(xy=0\). A non-zero element u of R is unit in R if there exists v in R such that \(uv=1\). For \(x\in R\), the annihilator of x is the set \(Ann(x)=\{y\in R\mid xy=0\}\). A ring R is said to be local if it has unique maximal ideal M. The nilradical of a ring R is the set \(J=\{x\in R : x^t=0,\) for some positive integer \(t\}\). The index of nilpotency of J is the least positive integer m for which \(J^m=\{0\}\), where \(J^m=JJ\ldots J\) (m-times). A ring R is said to be reduced if \(J=\{0\}\). A ring is said to be indecomposable if it can not be written as a direct product of two rings. Let \(\mathbb {Z}_n\) be the ring of integer modulo n.

For any ring R, in [6], Beck associated a simple graph with R whose vertices are the elements of R and any two distinct vertices x and y are adjacent if and only if \(xy=0\) in R. Beck conjectured that (see [6]) the chromatic number and clique number of this graph are the same and this was disproved by Anderson and Naseer in [2] (also, see [10]). It can be observed that for the graph associated with the ring, the vertex 0 is adjacent to every other vertex. Anderson and Livingston in [5] slightly modified the definition of the graph associated with a ring by considering the zero-divisors as the vertices and any two distinct vertices x and y are adjacent if and only if \(xy=0\) in R. They called this zero-divisor graph of the ring R and it is denoted by \(\varGamma (R)\). Zero-divisor graphs have been extensively studied in the past. This can be seen in [1, 3, 4, 11, 20].

The following definitions and results can be found in [4, 20]. For \(x,y\in R\), define \(x\mathtt {\sim }_R\ y\) if and only if \(Ann(x)=Ann(y)\). It is proved in [4] that the relation \(\mathtt {\sim }_R\) is an equivalence relation on R. For \(x\in R\), let \(D_x=\{r\in R\ |\ x\mathtt {\sim }_R\ r \}\) be the equivalence class of x. Let \(R_E=\{D_{x_1},D_{x_2},\ldots , D_{x_k}\}\) be the set of all equivalence classes of the relation \(\mathtt {\sim }_R\). The compressed zero-divisor graph \(\varGamma _E(R)\) of R (defined in [20]) is a simple graph with vertex set \(R_E\backslash \{D_0,D_1\}\) and two distinct vertices \(D_x\) and \(D_y\) are adjacent if and only if \(xy=0\). The following result can be found in [18].

Theorem 1

[18]. If R is a ring, then

  1. (i)

    \(\varGamma (R)\cong \varGamma _E(R)[\langle D_{x_1}\rangle ,\langle D_{x_2}\rangle ,\ldots , \langle D_{x_{k-2}}\rangle ]\), where \(D_{x_i}\ne D_0, D_1\), for \(1\le i\le k-2\),

  2. (ii)

    \(\left\langle D_{x_i}\right\rangle \) is complete if and only if \(x_i^2=0\), and

  3. (iii)

    \(\left\langle D_{x_i}\right\rangle \) is totally disconnected (that is, \(\left\langle D_{x_i}\right\rangle \) has no edge) if and only if \(x_i^2\ne 0\).

The following result is proved in [3].

Theorem 2

[3]. If R is a non-zero reduced ring, then there exists a positive integer k such that the compressed zero-divisor graph \(\varGamma _E(R) \cong \varGamma (\mathbb {Z}_2^k)\), where \(\mathbb {Z}_2^k=\mathbb {Z}_2\times \mathbb {Z}_2 \times \ldots \times \mathbb {Z}_2\) (k-times).

In [9], Hala and Jukl introduced the concept of the zero-divisor graph of a poset. Let \((P,\le )\) be a finite poset with the least element 0. For any \(a, b\in P\), denote \(L(a,b) =\{c\in P\ |\ c \le a\) and \(c\le b\}\). A non-zero element \(a\in P\) is said to be a zero-divisor if \(L(a,b) = \{0\}\) for some \(0\ne b\in P\). We say a non-zero element \(a\in P\) is an atom (primitive) if for any \(0\ne b\in P,\ b \le a\) implies \(a=b\). The zero-divisor graph \(\varGamma (P)\) of a poset P is a graph whose vertex set \(V(\varGamma (P))\) consists of the zero-divisors of P, in which a is adjacent to b if and only if \(L(a,b) =\{0\}\). It is shown in [9] that for any poset P, the clique number and the chromatic number of \(\varGamma (P)\) are the same.

By a semigroup, we mean a finite commutative semigroup with the zero element 0. A semigroup S is said to be reduced if for any \(a\in S\) and any positive integer n, \(a^n = 0\) implies \(a = 0\). A semigroup S is said to be idempotent (it is a so-called semilattice, see [13]) if for each \(a\in S\), \(a^2=a\).

We define a zero-divisor graph of a semigroup in a similar manner in the definition of zero-divisor graph of a ring.

Let \(R = \mathbb {Z}_2^k\). Clearly, it is a Boolean ring and it becomes a poset by defining \(a\le b\) iff \(ab=a\) for any \(a,b \in R\). Note that, the zero-divisor graphs of R as a ring (or a semigroup) and as a poset coincide. Let H be a subgraph of \(\varGamma (\mathbb {Z}_2^k)\). We say that H is minimal (see [13]) if H is an induced subgraph of \(\varGamma (\mathbb {Z}_2^k)\) which contains all the atoms of the poset \(\mathbb {Z}_2^k\), and we say H is minimal closed (see [13]) if H is minimal and \(V(H)\cup \{0\}\) is a sub-semigroup of \(\mathbb {Z}_2^k\). The following results can be found in [13].

Theorem 3

[13]. Let G be a simple graph with \(\omega (G)=k\). Then the following statements are equivalent:

  1. (i)

    G is the zero-divisor graph of a poset.

  2. (ii)

    G is a k-compact graph.

  3. (iii)

    G is a generalized complete k-partite graph.

  4. (iv)

    The reduced graph \(G_r\) of G is isomorphic to a minimal subgraph of \(\varGamma (\mathbb {Z}_2^k)\).

Theorem 4

[13]. Let G be a simple graph with \(\omega (G)=k\). Then the following statements are equivalent:

  1. (i)

    G is the zero-divisor graph of a reduced semigroup with 0.

  2. (ii)

    G is a generalized complete k-partite graph such that for any non-adjacent vertices \(a, b\in V(G)\), there is a vertex \(c\in V(G)\) with \(W(c)= W(a)\cup W(b)\).

  3. (iii)

    The reduced graph \(G_r\) of G is isomorphic to a minimal closed subgraph of \(\varGamma (\mathbb {Z}_2^k)\).

  4. (iv)

    G is the zero-divisor graph of a semilattice (or equivalently, idempotent semigroup) with 0.

A graph G is perfect if \(\omega (H)=\chi (H)\) for every induced subgraph H of G.

The following result was proved by Lovasz, see [12].

Theorem 5

[12]. The complement of every perfect graph is perfect.

In [7], Berge conjectured the following and it was proved by Chudnovsky et al., see [8].

Theorem 6

(Strong Perfect Graph Theorem [8]). A graph G is perfect if and only if it does not contain an induced subgraph which is either an odd cycle of length at least 5 or the complement of such a cycle.

The paper mainly deals with the results on perfect graph using the Strong Perfect Graph Theorem. As a result, we deduced many known results in the literature. This is precisely as follows.

In Sect. 2, we prove that the G-generalized join of complete graphs and totally disconnected graphs is perfect if and only if G is perfect. As a consequence, we deduce the results proved in [14] and [17] and prove that the lexicographic product of a perfect graph and a complete graph and the lexicographic product of a perfect graph and a complement of a complete graph are perfect.

In Sect. 3, we characterize rings, posets and reduced semigroups whose zero-divisor graphs and ideal based zero-divisors are perfect. As a result, we characterize distributive lattices with 0, reduced semirings and boolean rings whose zero divisor graphs are perfect, which are proved in [15]. Further, we completely characterize rings the ideal based zero-divisor graph of the ring \(\mathbb {Z}_n\) is perfect.

2 When a G-generalized Join of Complete and Totally Disconnected Graphs is Perfect

In this section, we prove the following result on perfect graphs.

Theorem 7

If G is a graph with vertex set \(V(G)=\{v_1,v_2,\ldots , v_n\}\) and \(H_1, H_2,\)

\(\ldots , H_n\) are graphs such that each \(H_i\) is either complete or a totally disconnected graph, then G is perfect if and only if \(G[H_1,H_2,\ldots , H_n]\) is perfect.

Proof

Let \(G'=G[H_1,H_2,\ldots , H_n]\). It is enough to prove if G is perfect, then \(G'\) is perfect. Suppose \(G'\) is not perfect, then by Theorem 6, \(G'\) contains either an odd cycle of length at least 5 as an induced subgraph or the complement of an odd cycle of length at least 5 as an induced subgraph.

Case 1. \(G'\) contains an odd cycle \(C_{2k+1}\) as an induced subgraph, where \(k\ge 2\).

Let \(V(C_{2k+1})=\{x_0,x_1,\ldots , x_{2k}\}\) such that \(x_i\) is adjacent to \(x_{i+1}\) (where the addition in subscript is taken modulo \(2k+1\)) and \(x_i\) is not adjacent to \(x_j\), where \(j\ne i-1,i+1\). Suppose there exists \(1\le t\le n\) such that \(|V(C_{2k+1})\cap V(H_t)|\ge 2\).

First, if there exists \(0\le i\le 2k\) such that \(x_i,x_{i+1}\in V(H_t)\). Then \(H_t\) is complete and hence \(x_{i-1}\notin V(H_t)\) (otherwise, \(C_{2k+1}\) would not be induced in \(G'\)). Thus there exists \(1\le s\le n\) with \(s\ne t\) such that \(x_{i-1}\in V(H_s)\) and hence \(x_{i-1}\) is adjacent to \(x_{i+1}\), which is a contradiction.

Next, if there exist \(0\le i,j\le 2n\) such that \(j\ne i-1,i, i+1\) and \(x_i,x_j\in V(H_t)\). Then \(H_t\) has no edge in \(G'\) and \(x_{i+1},x_{i-1}\notin V(H_t)\). Suppose if \(j\ne i+2\), then there exists \(1\le s\le n\) such that \(s\ne t\) and \(x_{i+1}\in V(H_s)\) and hence \(x_j\) is adjacent to \(x_{i+1}\), (because of \(x_ix_{i+1}\in E(C_{2k+1})\)) which is a contradiction. Therefore, if \(j=i+2\), then there exists \(1\le s\le n\) such that \(s\ne t\) and \(x_{i-1}\in V(H_s)\) and therefore \(x_{i-1}\) is adjacent to \(x_j\), which is again a contradiction.

Hence \(|V(C_{2k+1})\cap V(H_i)|=1\), for \(0\le i\le 2k\) which implies that G contains an odd cycle of length at least 5 as an induced subgraph, which is a contradiction.

Case 2. \(G'\) contains a complement of an odd cycle of length at least 5 as an induced subgraph.

Let \(\overline{C_{2k+1}}\) be the complement of the odd cycle \(C_{2k+1}\) as an induced subgraph of \(G'\), where \(k\ge 2\) with \(V(\overline{C_{2k+1}})=\{x_0,x_1,\ldots , x_{2k}\}\) such that \(x_i\) is not adjacent to \(x_j\) for \(j=i-1,i+1\) and \(x_i\) is adjacent to \(x_j\), for \(j\ne i-1,i,i+1\) (where the addition in subscripts is taken modulo \(2k+1\)). Suppose there exists \(1\le t\le n\) such that \(|V(C_{2k+1})\cap V(H_t)|\ge 2\).

First, if there exists \(0\le i\le 2k\) such that \(x_i,x_{i+1}\in V(H_t)\). Then \(H_t\) has no edge and \(x_{i-1}\notin H_t\) and hence there exists \(1\le s\le n\) with \(s\ne t\) such that \(x_{i-1}\in V(H_s)\). But \(x_{i+1}\) is adjacent to \(x_{i-1}\) and hence \(x_i\) is adjacent to \(x_{i-1}\), which is a contradiction.

Next, if there exist \(0\le i,j\le 2n\) such that \(j\ne i-1,i, i+1\) and \(x_i,x_j\in V(H_t)\). Then \(H_t\) is complete and \(x_{i-1},x_{i+1}\notin V(H_t)\). Suppose if \(j\ne i+2\), then there exists \(1\le s\le n\) such that \(s\ne t\) and \(x_{i+1}\in V(H_s)\). But \(x_j\) is adjacent to \(x_{i+1}\) and therefore \(x_i\) is adjacent to \(x_{i+1}\), which is impossible. Hence, if \(j=i+2\), then there exists \(1\le s\le n\) such that \(s\ne t\) and \(x_{i-1}\in V(H_s)\) and therefore \(x_{i-1}\) is adjacent \(x_i\), which is again a contradiction.

Thus \(|V(\overline{C_{2k+1}})\cap V(H_i)|=1\), for \(0\le i\le 2k\), which implies that G contains a complement of an odd cycle of length at least 5 as an induced, which is a contradiction.

The following corollary is an immediate consequence of Theorem 7.

Corollary 1

If G is perfect and n is a positive integer, then \(G[K_n]\) and \(G[K_n^c]\) are perfect.

Proof

As \(G[K_n]\cong G[K_n,K_n,\ldots , K_n]\) and \(G[K_n^c]\cong G[K_n^c,K_n^c,\ldots , K_n^c]\), the result follows from Theorem 7.

The following result proved in [17] is deduced from Theorem 7.

Corollary 2

(Corollary 3.2, [17]). A graph G is perfect if and only if it’s reduced graph \(G_r\) is perfect.

The following relation is defined on a graph G in [14]. For \(x,y\in V(G)\), define \(x\approx y\) if and only if either \(x=y\) or \(xy\in E(G)\) and \(N(x)\backslash \{y\}=N(y)\backslash \{x\}\). Clearly, it is an equivalence relation. Let [x] be the equivalence class of x, and \(S=\{[x_1],[x_2],\ldots , [x_r]\}\) be the set of all equivalence classes of the relation \(\approx \). Based on these equivalence classes of the relation \(\approx \), we defined (This can be seen in [14]) the graph \(G_{red}\) with vertex set \(V(G_{red})=S\) and two distinct vertices [x] and [y] are adjacent in \(G_{red}\) if and only if x and y are adjacent in G. Clearly, for any graph G, \(G=G_{red}[\langle [x_1]\rangle ,\langle [x_2]\rangle ,\ldots , \langle [x_r]\rangle ]\) and \(\langle [x_i]\rangle \) is complete, for \(1\le i\le r\).

By Theorem 7, we deduce the following result proved in [14].

Corollary 3

(Theorem 4.4, [14]). A graph is perfect if and only if \(G_{red}\) is perfect.

3 Perfect Zero-Divisor Graph of a Ring

In this section, we ask the following interesting question. When does the zero-divisor graph of a ring R perfect? To answer this question, we provide a necessary and sufficient condition for which the zero-divisor graph of a ring is perfect.

Theorem 8

If R is a ring, then \(\varGamma (R)\) is perfect if and only if its compressed zero-divisor graph \(\varGamma _E(R)\) of R is perfect.

Proof

The result follows from Theorems 1 and 7.

Let \(R_1,R_2,\ldots , R_k\) be rings. For \(x_j\in R_1\times R_2\times \ldots \times R_k\), there exists a unique \(x_j(i)\in R_i\), for \(1\le i\le k\), such that \(x_j=(x_j(1),x_j(2),\ldots , x_j(k))\).

Note that there are several rings satisfying Beck’s conjecture; see [2, 4, 6, 9, 10, 20]. One of them is a finite reduced ring. Using Observation 1, we give a shorter proof of this result as follows.

Corollary 4

[6, 20]. If R is a non-zero reduced ring, then \(\chi (\varGamma (R))=\omega (\varGamma (R))\).

Proof

By Observation 1 and Theorem 2, it is enough to prove \(\omega (\varGamma (\mathbb {Z}_2^k))=\chi (\varGamma (\mathbb {Z}_2^k))\). Clearly \(\{e_i\ |\ 1\le i\le k\}\), where \(e_i=(0,\ldots ,0,1,0,\ldots , 0),\) induces a clique. Color first \(e_i\) by i, for \(1\le i\le k\).

For any \( x=(x(1),x(2),\ldots , x(k))\in V(\varGamma (\mathbb {Z}_2^k))\backslash \{e_i\ |\ 1\le i\le k\},\) there exists a least j with \(1\le j\le k\), such that \(x(i)=0\) for \(1\le i\le j-1\) and \(x(j)=1\). Color x by j, then the resulting coloring is a proper k-coloring of \(\varGamma (\mathbb {Z}_2^k)\).

The following result gives a necessary condition for a product of rings whose zero-divisor graphs are perfect.

Theorem 9

Let \(R=R_1\times R_2\times \ldots \times R_k\), where \(R_i\)’s are indecomposable rings. If \(\varGamma (R)\) is perfect, then \(k\le 4\).

Proof

Suppose \(k\ge 5\). Then the set of vertices \(\{(1,1,0,0,0,0,\ldots ,0), (0,0,1,1,0,0, \ldots ,0),\) \((1,0,0,0,1,0,\ldots ,0),(0,1,0,1,0,0,\ldots ,0),(0,0,1,0,1,0,\ldots ,0)\}\) forms an induced cycle of length 5. By Theorem 6, we get a contradiction.

Next, let us prove the following result.

Theorem 10

If \(R=\mathbb {Z}_2^4\ (=\mathbb {Z}_2\times \mathbb {Z}_2\times \mathbb {Z}_2\times \mathbb {Z}_2)\), then \(\varGamma (R)\) is perfect.

Proof

Suppose \(\varGamma (R)\) is not perfect. Then, by Theorem 6, we consider the following cases.

Case 1. \(\varGamma (R)\) contains an odd cycle of length at least 5 as an induced subgraph.

Let \(C_{2r+1}\) be an induced cycle in \(\varGamma (R)\) of length \(2r+1\) with the vertex set \(\{x_0,x_1,\ldots , x_{2r}\}\), where \(r\ge 2\). If exactly one co-ordinate of \(x_i\) is non-zero, for \(0\le i\le 2r\), then \(2r+1\le 4\), a contradiction. Therefore there exists an \(x_i\) containing at least two non-zero co-ordinates. WLOG, \(x_i=(1,1,x_i(3),x_i(4))\), for some \(i,\ 0\le i\le 2r\). Then the \(1^{st}\) two coordinates of \(x_{i-1}, x_{i+1}\) are zeros, that is, \(x_{i-1}(1)=x_{i-1}(2)=x_{i+1}(1)=x_{i+1}(2)=0\). Since \(x_{i-1}\) and \(x_{i+1}\) are not adjacent, either the third coordinate or forth coordinate of \(x_{i-1}\) and \(x_{i+1}\) are non-zero. WLOG, \(x_{i-1}(3)=x_{i+1}(3)=1\). If \(x_{i-1}(4)=1\), then \(x_{i+1}(4)=0\), as \(x_{i-1}\ne x_{i+1}\) and hence \(x_{i-1}=(0,0,1,1)\) and \(x_{i+1}=(0,0,1,0)\). Since \(x_{i-2}\) is adjacent to \(x_{i-1}\), we have \(x_{i-2}=(x_{i-2}(1),x_{i-2}(2), 0,0)\). Thus \(x_{i-2}\) is adjacent to \(x_{i+1}\), which is a contradiction. Hence \(x_{i-1}(4)=0\), which implies that \(x_{i+1}(4)=1\) and thus \(x_{i+1}=(0,0,1,1)\) and \(x_{i-1}=(0,0,1,0)\). Since \(x_{i+2}\) is adjacent \(x_{i+1}\), we have \(x_{i+2}=(x_{i+2}(1),x_{i+2}(2),0,0)\) and hence \(x_{i+2}\) is adjacent to \(x_{i-1}\), which is a contradiction.

Case 2. \(\varGamma (R)\) contains the complement of an odd cycle of length at least 5 as an induced subgraph.

Let \(\overline{C_{2r+1}}\) be an induced subgraph of \(\varGamma (R)\) with vertex set \(\{x_0,x_1,\ldots , x_{2r}\}\), where \(r\ge 2\). If no \(x_i\) contains exactly two coordinates that are non-zeros, then there exists \(j,\ 1\le j\le k\) such that \(x_j\) contains exactly three that coordinates that are non-zero (otherwise \(2r+1\le 4\)), which is impossible. Thus there exists \(i,\ 1\le i\le k\) such that \(x_i\) contains exactly two coordinates that are non-zeros. WLOG, \(x_i=(1,1,x_i(3),x_i(4))\). Since \(x_i\) is adjacent to \(2r-2\) vertices in \(\overline{C_{2r+1}}\), namely \(x_{i+2},x_{i+3},\ldots , x_{i+2r-1}\) (where the addition in subscripts taken modulo \(2r+1\)), we have the 1\(^{st}\) two coordinates of \(x_{i+2},x_{i+3},\ldots , x_{i+2r-1}\) are zero’s and hence \(x_{i+2},x_{i+3},\ldots , x_{i+2r-1}\in \{(0,0,1,1),(0,0,1,0),(0,0,0,1)\}\). Thus \(2r-2\le 3\), which implies \(2r+1\le 6\). As it is an odd number and \(r\ge 2\), we have \(2r+1=5\). Therefore \(\overline{C_5}\cong C_5\). By Case 1, which is impossible.

The following result in [14] is a consequence of Theorems 9 and 10.

Corollary 5

[14]. If \(R=\mathbb {Z}_2^k\), then \(\varGamma (R)\) is perfect if and only if \(k\le 4\).

Proof

By Theorems 9 and 10, it is enough to prove that \(\varGamma (R)\) is perfect if \(k\le 3\). In this case we have \(|V(\varGamma (R))|\le 6\), and hence \(\varGamma (R)\) does not contain a cycle of length 5 as an induced subgraph of \(\varGamma (R)\) and, thus the result follows.

It is well-known that any finite non-zero reduced commutative ring R is isomorphic to a finite direct product of finite fields, say \(\mathbb {F}_{p_1^{\alpha _1}},\mathbb {F}_{p_2^{\alpha _2}}, \ldots , \mathbb {F}_{p_{\ell }^{\alpha _\ell }}\), where \(p_i\)’s are prime numbers and \(\alpha _i\)’s are positive integers, that is \(R\cong \mathbb {F}_{p_1^{\alpha _1}}\times \mathbb {F}_{p_2^{\alpha _2}}\times \ldots \times \mathbb {F}_{p_{\ell }^{\alpha _\ell }}\).

By Theorem 2, the compressed zero-divisor graph of a reduced ring R is isomorphic to the zero-divisor graph of \(\mathbb {Z}_2^k\), for some \(k\ge 1\), that is \(\varGamma _E(R)\cong \varGamma (\mathbb {Z}_2^k)\). So, the following result is a consequence of Theorem 9 and Corollary 5.

Theorem 11

If \(R\cong \mathbb {F}_{p_1^{\alpha _1}}\times \mathbb {F}_{p_2^{\alpha _2}}\times \ldots \times \mathbb {F}_{p_{\ell }^{\alpha _\ell }}\) is a non-zero reduced ring, where \(\mathbb {F}_{p_i^{\alpha _i}}\)’s are finite fields, then \(\varGamma (R)\) is perfect if and only if \(\ell \le 4\).

Proof

The first part is clear from Theorem 9. For the second part, let us assume that \(\ell \le 4\). Then \(\omega (\varGamma (R))\le 4\). By the above discussion, \(\varGamma _E(R)\cong \varGamma (\mathbb {Z}_2^k)\) for some \(k\ge 1\). Suppose \(k\ge 5\), then \(\varGamma (\mathbb {Z}_2^k)\) contains a clique \(\langle \{e_i: 1\le i\le k\}\rangle \) of size at least 5 (where \(e_i\)’s are defined in Corollary 4) and hence \(\omega (\varGamma (R))\ge 5\), which is impossible. Thus \(k\le 4\) and therefore, by Corollary 5 \(\varGamma (\mathbb {Z}_2^k)\) is perfect, and hence \(\varGamma (R)\) is perfect by Theorem 8.

The following result in [15] is an immediate consequence of Corollary 5, because every finite Boolean ring R is isomorphic to \(\mathbb {Z}_2^k\), for some \(k\ge 1\).

Corollary 6

[15]. Let R be a finite Boolean ring. Then the following are equivalent,

  1. (1)

    \(\varGamma (R)\) is perfect.

  2. (2)

    \(\varGamma (R)\) does not contain \(K_5\) as a subgraph.

  3. (3)

    \(|R|\le 2^4\).

3.1 Perfect Ideal Based Zero-Divisor Graph of Rings

In this subsection, we characterize rings whose ideal based zero-divisor graphs are perfect. In particular, under what values of n, the ideal based zero divisor graph of the ring \(\mathbb {Z}_n\) of integers modulo n is perfect.

The following observation is observed in [16] and [21].

  1. (i)

    If I is an ideal of R and \(x_1+I, x_2+I,\ldots , x_k+I\) are the distinct co-sets of I, which are zero-divisors of the quotient ring \(\frac{R}{I}\), then \(\varGamma _I(R)\) is a \(\varGamma (\frac{R}{I})\)-generalized join of \(\langle x_1+I \rangle , \langle x_2+I \rangle , \ldots , \langle x_k+I \rangle \), that is,

    $$\varGamma _I(R)=\varGamma \Big (\frac{R}{I}\Big )\big [\langle x_1+I \rangle , \langle x_2+I \rangle , \ldots , \langle x_k+I \rangle \big ],$$
  2. (ii)

    \(\langle x_i+I\rangle \) is a complete subgraph of \(\varGamma _I(R)\) if and only if \(x_i^2\in I\),

  3. (iii)

    \(\langle x_i+I\rangle \) is a totally disconnected subgraph of \(\varGamma _I(R)\) if and only if \(x_i^2\notin I\).

Hence, by Theorems 7 and 8, we have

Theorem 12

Let I be an ideal of R, then the following are equivalent,

  1. (i)

    \(\varGamma _I(R)\) is perfect;

  2. (ii)

    \(\varGamma (\frac{R}{I})\) is perfect;

  3. (iii)

    \(\varGamma _E(\frac{R}{I})\) is perfect.

We recall the following result proved in [19].

Theorem 13

[19]. The zero divisor graph \(\varGamma (\mathbb {Z}_n)\) of a ring \(\mathbb {Z}_n\) is perfect if and only if \(n=p^a , p^aq^b, p^aqr\), or pqrs, where pqr and s are distinct primes and a and b are positive integers.

It is well known that if I is an ideal of \(\mathbb {Z}_n\) generated by m, then \(\frac{\mathbb {Z}_n}{I}\cong \mathbb {Z}_m\). So, we have

Corollary 7

If I is an ideal of \(\mathbb {Z}_n\) generated by m, then \(\varGamma _I(\mathbb {Z}_n)\) is perfect if and only if \(m=p^a , p^aq^b, p^aqr\), or pqrs, where pqr and s are distinct primes and a and b are positive integers.

Proof

By Theorems 12 and 13, \(\varGamma _I(\mathbb {Z}_n)\) is perfect if and only if \(\varGamma (\mathbb {Z}_m)\) is perfect if and only if \(m=p^a , p^aq^b, p^aqr\), or pqrs.

3.2 Zero-Divisor Graph of Rings, Reduced Semigroups and Posets

In [13], it is shown that the chromatic number is equal to the clique number of zero-divisor graphs of poset, reduced semiring with 0 and reduced semigroup with 0. So it is interesting to consider the following problem.

Problem. Characterize the posets, reduced rings and reduced semigroups whose zero-divisor graphs are perfect.

Now we characterize posets whose zero-divisor graphs are perfect using Theorem 3.

Theorem 14

Let G be a zero-divisor graph of a poset with 0 and \(\omega (G)=k\). Then the following are equivalent,

  1. (i)

    G is perfect.

  2. (ii)

    The reduced graph \(G_r\) of G is perfect.

  3. (iii)

    The reduced graph \(H_r\) of H (where H is in the Definition of generalized complete k-partite graph) is perfect.

Proof

  1. (i)

    \(\Leftrightarrow \ (ii)\) It follows from Corollary 2.

  2. (ii)

    \(\Rightarrow \ (iii)\) It follows from the definition of perfect.

  3. (iii)

    \(\Rightarrow \ (ii)\) Suppose \(G_r\) is not perfect graph, then by the Theorem 6, \(G_r\) contains an odd cycle of length at least 5 as an induced subgraph or the complement of an odd cycle of length at least 5 as an induced subgraph.

Let \(e_1,e_2,\ldots , e_k\) be the atoms of G.

Case 1. \(G_r\) contains an odd cycle of length at least 5 as an induced subgraph.

Let \(C_{2s+1}\) be an odd cycle of \(G_r\) as an induced subgraph with vertex set \(V(C_{2s+1})=\{a_0,a_1\ldots , a_{2s}\}\), where \(s\ge 2\). Then \(V(C_{2s+1})\) is not a subset of \(V(H_r)\). As the atoms forms a clique, we have \(|V(C_{2s+1})\cap \{e_1,e_2,\ldots , e_k\}|\le 2\). First if \(|V(C_{2s+1})\cap \{e_1,e_2,\ldots , e_k\}|= 2\), then there exist \(i,j\in \{1,2,\ldots , k\}\) and \(\ell \in \{0,1,2,\ldots , 2s\}\) such that \(a_{\ell }=e_i\) and \(a_{\ell +1}=e_j\). Since \(a_{\ell +2}\) and \(a_{\ell +3}\) are not adjacent to \(a_{\ell }=e_i\), we have \(i\notin W(a_{\ell +2})\cup W(a_{\ell +3})\) and hence \(W(a_{\ell +2})\cup W(a_{\ell +3})\ne \{1,2,\ldots , k\}\), which is a contradiction to the definition of generalized complete k-partite graph. Next if \(|V(C_{2s+1})\cap \{e_1,e_2,\ldots , e_k\}|=1\), then we get a contradiction in a similar way as above. Thus \(V(C_{2s+1})\cap \{e_1,e_2,\ldots , e_k\}=\emptyset \) and hence \(C_{2s+1}\) is an induced odd cycle of \(H_r\), which is a contradiction.

Case 2. \(G_r\) contains the complement of an odd cycle of length at least 5 as an induced subgraph.

Let \(\overline{C_{2s+1}}\) be the complement of the odd cycle \(C_{2s+1}\) in \(G_r\) with vertex set \(V(\overline{C_{2s+1}})=\{a_0,a_1,\ldots , a_{2s}\}\), where \(s\ge 2\). If \(V(\overline{C_{2s+1}})\cap \{e_1,e_2,\ldots , e_k\}\ne \emptyset \), then there exists \(i\in \{1,2,\ldots , k\}\) such that \(e_i=a_{\ell }\), for some \(\ell \in \{0,1,2,\ldots , 2s\}\). Then \(a_{\ell -1},a_{\ell +1}\notin \{e_1,e_2,\ldots , e_k\}\) and they are not adjacent to \(e_i\) and hence \(i\notin W(a_{\ell -1})\cup W(a_{\ell +1})\), which is impossible. Thus, \(V(\overline{C_{2s+1}})\cap \{e_1,e_2,\ldots , e_k\}= \emptyset \) and therefore \(\overline{C_{2s+1}}\) lies in \(H_r\), which is a contradiction.

Next, we present equivalent conditions for a zero-divisor graph of a reduced semigroup to be perfect using Theorem 4.

Theorem 15

Let G be a zero-divisor graph of a reduced semigroup with \(\omega (G)=k\). Then the following are equivalent,

  1. (i)

    G is perfect.

  2. (ii)

    The reduced graph \(G_r\) of G is perfect.

  3. (iii)

    The reduced graph \(H_r\) of H (where H is given in the definition of generalized complete k-partite graph) is perfect.

Proof

The proof is similar to that of Theorem 14.

A lattice \(L=(L,\wedge , \vee )\) with 0 is distributive if for \(x,y,z\in L\), \(x\wedge (y\vee z)=(x\wedge y) \vee (x\wedge z)\). As every lattice is a poset, we have the following result proved in [15].

Corollary 8

[15]. Let L be a distributive lattice with 0. Then the following are equivalent,

  1. (i)

    \(\varGamma (L)\) is perfect.

  2. (ii)

    \(\varGamma (L)\) contains no induced cycle of length 5.

  3. (iii)

    \(\omega (\varGamma (L))\le 4\), (equivalently, the number of atoms of \(\varGamma (L)\) is at most 4).

Proof

(i) \(\Rightarrow \) (ii) It is trivial from the definition of perfect graph.

(ii) \(\Rightarrow \) (iii) If \(\left\langle \{a_1,a_2,\ldots , a_s\}\right\rangle \) is a clique in \(\varGamma (L)\), where \(s\ge 5\), then the subgraph induced by \(\{a_1\vee a_2, a_3\vee a_4, a_1\vee a_5, a_2\vee a_3,a_4\vee a_5\}\) is an induced cycle of length 5 (as L is distributive) which is a contradiction.

(iii) \(\Rightarrow \) (i) Suppose \(\varGamma (L)\) is not perfect. Then by Theorem 14, the reduced subgraph \(H_r\) of H, defined in Theorem 14, is not perfect. By Theorem 6, \(H_r\) contains an odd cycle of length at least 5 as an induced subgraph or its complement of an odd cycle of length at least 5 as an induced subgraph. If \(H_r\) contains an induced odd cycle \(C_{2s+1}\) with vertex set \(V(C_{2s+1})=\{a_1,a_2,\ldots , a_{2s+1}\}\), where \(s\ge 2\). Then \(a_i\wedge a_{i+1}=0\), for \(1\le i\le 2s\), \(a_{2s+1}\wedge a_1=0\) and \(a_i\wedge a_j\ne 0\), for \(j\ne i-1,i,i+1\) and hence the subgraph induced by \(\{a_1\wedge a_3, a_1\wedge a_4, a_2\wedge a_4, a_2\wedge a_5, a_3\wedge a_{2s+1}\}\) is a clique in \(\varGamma (L)\) of size 5, which is a contradiction. Similarly if \(H_r\) contains the complement \(\overline{C_{2s+1}}\) of an induced odd cycle \(C_{2s+1}\) with vertex set \(V(\overline{C_{2s+1}})=\{a_1,a_2,\ldots , a_{2s+1}\}\), where \(s\ge 2\), then the subgraph induced by \(\{a_1\wedge a_2, a_2\wedge a_3, a_3\wedge a_4, a_4\wedge a_5, a_5\wedge a_1\}\) is a clique in \(\varGamma (L)\) of size 5, which is again a contradiction.

As every semiring is a semigroup and by Theorem 15, we have the following result proved in [15].

Corollary 9

[15]. Let R be a reduced semiring with 0. Then the following are equivalent,

  1. (i)

    \(\varGamma (R)\) is perfect.

  2. (ii)

    \(\varGamma (R)\) contains no induced cycle of length 5.

  3. (iii)

    \(\omega (\varGamma (R))\le 4\), (equivalently, the number of atoms of \(\varGamma (R)\) is at most 4).

Proof

The proof is similar to that of Corollary 8 by replacing \(\vee \) and \(\wedge \) by addition and multiplication, respectively.