684
Views
0
CrossRef citations to date
0
Altmetric
Research Articles

On graphs with equal coprime index and clique number

, &
Pages 235-243 | Received 07 Feb 2022, Accepted 26 Jan 2023, Published online: 13 Jun 2023

Abstract

Recently, Katre et al. introduced the concept of the coprime index of a graph. They asked to characterize the graphs for which the coprime index is the same as the clique number. In this paper, we partially solve this problem. In fact, we prove that the clique number and the coprime index of a zero-divisor graph of an ordered set and the zero-divisor graph of a ring Zpn coincide. Also, it is proved that the annihilating ideal graphs, the co-annihilating ideal graphs and the comaximal ideal graphs of commutative rings can be realized as the zero-divisor graphs of specially constructed posets. Hence the coprime index and the clique number coincide for these graphs as well.

MATHEMATICS SUBJECT CLASSIFICATION:

1 Introduction

Graph labeling is a very active area of research. Many researchers studied several types of graph labelings. For an excellent survey on graph labelings, we refer to Gallian [Citation12]. Recently, Katre et al. [Citation16] studied another labeling known as coprime labeling.

Definition 1.1.

(Katre et al. [Citation16]) Let G=(V,E) be a graph of order n. An injection f:V{2,3,4,} is called a coprime labeling of G if for any two vertices u,vV, u and v are adjacent if and only if f(u) and f(v) are coprime. A graph that admits a coprime labeling is called a coprime graph. A prime number p is said to be used by the coprime labeling f if p divides f(v) for some vV. Let μ(G,f) be the number of primes used by the coprime labeling f. Then μ(G)=min{μ(G,f) : f isacoprimelabelingof G} is called the coprime index of G.

It is known that every graph G is a coprime graph (cf. [Citation16, Theorem 2.3]). Further, the coprime index of a graph G is nothing but the edge clique covering number of the complement of G (cf. [Citation16, Theorem 2.10]). Note that finding the edge clique covering number of a graph G is NP-complete; see [Citation13]. By an edge clique cover of a graph G, we mean a collection of cliques C1,C2,,Ck such that i=1kE(Ci)=E(G). The minimum cardinality of an edge clique cover of G is called the edge clique covering number of G and is denoted by θe(G). The best bound for θe(G) is due to Erdös, Goodman, and Posa [Citation11] which states that if G is any graph of order n, then θe(G) n2/4, and this bound is attained.

The edge clique covering number of a graph G is in fact the intersection number of G, that is, i(G)=θe(G); see Fred Roberts [Citation21, Theorem 5]. The intersection number of G, denoted by i(G), is the minimum cardinality of a set X such that G is the intersection graph of a family of subsets of X.

Looking at the relations between the coprime index and the edge clique covering number, it is clear that finding the coprime index of a graph is a tough problem.

The clique number ω(G) of a graph G is the size of the largest complete subgraph of G. The chromatic number χ(G) of a graph G is the minimum number of colors required to color the vertices of G such that no two adjacent vertices receive the same color.

The following are some observations.

Observation 1.2.

[16, Observations 2.6, 2.7, 2.8]

  1. μ(Kn)=n.

  2. If H is an induced subgraph of a graph G, then μ(H)μ(G).

  3. ω(G)μ(G).

In view of Observation 1.2 (3), Katre et al. [Citation16] raised the following problem.

Problem 1.3.

[16, Problem 2.9] Characterize graphs G for which ω(G)=μ(G).

In this paper, we partially solve this problem. In fact, we prove that the clique number and the coprime index of a zero-divisor graph of an ordered set and the zero-divisor graph of a ring Zpn coincide. Also, it is proved that the annihilating ideal graphs, the co-annihilating ideal graphs and the comaximal ideal graphs of commutative rings can be realized as the zero-divisor graphs of specially constructed posets. Hence the coprime index and the clique number coincide for these graphs as well.

In the first result, we prove that the coprime index of a graph and its reduced graph remain the same. The reduced graph usually has less number of vertices. So this technique is useful and effective.

Definition 1.4.

[Citation7, Definition 2.1] Let G be a simple graph. Consider an equivalence relation ∼ on G: uv if and only if N(u)=N(v), where N(x)={yV(G)xyisanedgeinG}. The reduced graph of G, denoted by Gred, is the simple graph whose vertex set is V(Gred)={[v]:vV(G)}, and two distinct equivalence classes [u] and [v] are adjacent in Gred if and only if u and v are adjacent in G.

A graph G and its reduced graph Gred are shown in . It is easily observed that Gred is isomorphic to an induced subgraph of G. Henceforth, we abuse the notation and will say that Gred is an induced subgraph of G. Clearly, if the graphs G and H are isomorphic, then μ(G)=μ(H).

Theorem 1.5.

Let G be a finite simple graph of order n and Gred be its reduced graph. Then μ(G)=μ(Gred).

Proof.

Since Gred is an induced subgraph of G, μ(Gred)μ(G). Hence, it is enough to prove that μ(G)μ(Gred). For this, let μ(Gred)=μ(Gred,f)=m, where f:V(Gred){2,3,4,} be a coprime labeling of Gred. Now, we define the coprime labeling f¯ of G using f. Let [x1],[x2],,[xk] be the k distinct vertices of Gred. Clearly, V(G)=i=1k.[xi]. Let [xi]={yij:j=1,2,,ki=|[xi]|}. Therefore, V(G)={yij:i=1,2,,k and j=1,2,,ki}.

Define a map f¯:V(G){2,3,4,} by f¯(yij)=(f([xi]))j. Clearly, f¯ is an injective map, as f is injective. It is easy to observe that yij and yrs are adjacent in G if and only if [xi] and [xr] are adjacent in Gred if and only if f([xi]) and f([xr]) are coprime (as f is a coprime labeling of Gred) if and only if (f([xi]))j and (f([xr]))s are coprime if and only if f¯(yij) and f¯(yrs) are coprime. This proves f¯ is a coprime labeling of G. Thus, μ(G)μ(Gred). Hence, μ(G)=μ(Gred). □

Proposition 1.6.

[Citation7, Proposision 2.4] G is a complete r-partite graph if and only if Gred is the complete graph Kr.

In [Citation16], Katre et al. observed that, μ(Kn)=n. In view of Theorem 1.5 and Proposition 1.6, we have the following corollary.

Corollary 1.7.

Let G be a complete r-partite graph. Then μ(G)=r.

In [Citation16], Katre et al. observed that the clique number ω(G) is less than or equal to the coprime index μ(G) of G. We sharpen this lower bound of the coprime index μ(G) in the following result.

Theorem 1.8.

Let G be a finite simple graph. Then ω(G)χ(G)μ(G).

Proof.

It is well known that ω(G)χ(G). We want to prove that χ(G)μ(G). Let μ(G)=μ(G,f)=m and let p1,p2,,pm be the m distinct primes used by the coprime labeling f of G. To prove χ(G)μ(G), it is enough to prove that χ(G)m. For this, define the set V1={vV : p1 divides f(v)}, and the set Vk={vV(j=1k1Vj) : pk divides f(v)} for k{2,3,m}. Clearly, V=i=1m.Vi. We assign m different colors to these Vi’s. Now, we prove that, this is a proper coloring of G. For this, if x and y are adjacent in G, then f(x) and f(y) are coprime (as f is a coprime labeling of G). This implies x and y cannot be contained in the same Vi. So x and y have different colors. Hence, this is a proper coloring with m colors. Thus, χ(G)m. This proves ω(G)χ(G)μ(G). □

Definition 1.9.

(Anderson and Livingston [Citation5]) Let R be a commutative ring with identity. The zero-divisor graph Γ(R) of R is a simple graph with the vertex set Z*(R)=Z(R){0}, the set of non-zero zero-divisors of R, and two distinct vertices x, y are adjacent if and only if xy = 0.

Notation 1.10.

(Pirzada et al. [Citation20]) Let p be a prime number. Consider the partition of the vertex set V(Γ(Zpn)) into V1,V2,,Vn1, where Vi={lipi:pli},1in1. Then |Vi|=(p1)pni1.

Rewrite the elements of Vi as Vi={xi1,xi2,,xiki}, where ki=|Vi|=(p1)pni1. Then the elements xi1j1 and xi2j2 are adjacent in Γ(Zpn) if and only if i1+i2n. Also, the elements xi1j1 and xi2j2 are adjacent in the complement graph Γc(Zpn) if and only if i1+i2<n. The zero-divisor graphs Γ(Zpn) and their complements Γc(Zpn) for n = 8 and n = 7 are shown in . In the figures, a loop at Vi denotes that any two vertices of Vi are adjacent, that is, the vertices of Vi induce a complete graph on |Vi| vertices and Vi is adjacent to Vj denotes that each vertex of Vi is adjacent to every vertex of Vj, that is, Vi+Vj is a graph join. The join G + H of two graphs G and H is a graph formed from disjoint copies of G and H by connecting each vertex of G to each vertex of H.

Fig. 1 Zero-divisor graphs and their complements.

Fig. 1 Zero-divisor graphs and their complements.

We illustrate the coprime labelings of Γ(Z24) and Γ(Z25) in . The primes used in the coprime labeling of Γ(Z24) are p21, p22 and p31. Hence, μ(Γ(Z24))=3=ω(Γ(Z24)). Also, the primes used in the coprime labeling of Γ(Z25) are p2,p31,p32 and p41. Therefore, μ(Γ(Z25))=4=ω(Γ(Z25)). The following result is essentially proved in [Citation8, Proposition 2.3] and [Citation20, Theorem 2].

Fig. 2 Coprime labeling of zero-divisor graphs.

Fig. 2 Coprime labeling of zero-divisor graphs.

Theorem 1.11.

([Citation8, Proposition 2.3], [Citation20, Theorem 2]) Let Γ(Zpn) be the zero-divisor graph of a ring Zpn, where p is a prime. Then χ(Γ(Zpn))=ω(Γ(Zpn))={pn21if n iseven;pn12if n isodd.

With this preparation, we prove our first main result.

Theorem 1.12.

Let Γ(Zpn) be the zero-divisor graph of a ring Zpn, where p is a prime. Then μ(Γ(Zpn))=χ(Γ(Zpn))=ω(Γ(Zpn))={pn21if n iseven;pn12if n isodd.

Proof.

Consider the partition Vi ; i=1,2,,n1 of V(Γ(Zpn)) as defined in Notation 1.10.

Let Vi={xi1,xi2,,xiki}, where ki=|Vi|=(p1)pni1. We have the following two main cases on the values of n.

Case A: Let n be an even number.

Consider a map f:V(Γ(Zpn)){2,3,4,} as follows: f(xij)={pijfor n2in1,1jki;(l=1kn2pn2l)jfor i=n21,1jk(n21);[(l=1kn2pn2l)(l=1k(n2+1)p(n2+1)l)(l=1kni1p(ni1)l)]jfor 1i(n22),1jki,where pij are distinct primes. Now, we prove that f is a coprime labeling of Γ(Zpn). For this, let u,vV(Γ(Zpn)) be distinct vertices of Γ(Zpn). We study the following three sub-cases.

Sub-case A-I: If uVi1,vVi2 for some 1i1i2n21, then u=xi1j1, v=xi2j2 for some 1j1ki1,1j2ki2 (see Notation 1.10). Clearly, i1+i2<n. This implies that u and v are not adjacent in Γ(Zpn). Clearly, the primes pn2l divide both f(u)=f(xi1j1) and f(v)=f(xi2j2), for 1lkn2. Therefore, f(u) and f(v) are not coprime.

Sub-case A-II: If uVi1,vVi2 for some n2i1i2n1, then u=xi1j1, v=xi2j2 for some 1j1ki1,1j2ki2. Clearly, i1+i2n. This gives u and v are adjacent in Γ(Zpn). By the definition of f, f(u)=f(xi1j1)=pi1j1 and f(v)=f(xi2j2)=pi2j2. As u and v are distinct vertices, pi1j1pi2j2. This implies f(u) and f(v) are coprime.

Sub-case A-III: If uVi1,vVi2 for some 1i1n21,n2i2n1, then u=xi1j1,v=xi2j2 for some 1j1ki1,1j2ki2. In this case, u=xi1j1 and v=xi2j2 are adjacent if and only if i1+i2n. If i1+i2n, then i2ni1>ni11. By the definition of f, the primes pi2l does not divide f(xi1j1), for all 1lki2. In particular, the prime pi2j2 does not divide f(xi1j1). As f(xi2j2)=pi2j2, we have f(xi1j1) and f(xi2j2) are coprime. Hence, f(u) and f(v) are coprime. If i1+i2<n, then u and v are not adjacent and i2<ni1 (i2ni11). By the definition of f, pi2j2 divides both f(xi1j1) and f(xi2j2). Therefore, f(u) and f(v) are not coprime.

Thus, f is a coprime labeling of Γ(Zpn).

The number of distinct primes used in the coprime labeling f=|Vn2|+|Vn2+1|++|Vn1|=(p1)pnn21+(p1)pn(n2+1)1++(p1)pn(n1)1=(p1)pn21+(p1)pn22++(p1)=(p1){pn21+pn22++1}=pn21=ω(Γ(Zpn)). This implies μ(Γ(Zpn))ω(Γ(Zpn)). Therefore, by Theorem 1.8, μ(Γ(Zpn))=χ(Γ(Zpn))=ω(Γ(Zpn))=pn21.

Case B: Let n be an odd number.

Consider a map f:V(Γ(Zpn)){2,3,4,} as follows: f(xij)={pijfor n2+1in1,1jki;(pn2)jfor i=n2,1jkn2;[(pn2)(l=1k(n2+1)p(n2+1)l)(l=1kni1p(ni1)l)]jfor 1i(n21),1jki,where pij and pn2 are distinct primes. Now, we prove that f is a coprime labeling of Γ(Zpn). For this, let u,vV(Γ(Zpn)) be distinct vertices of Γ(Zpn). We study the following three sub-cases.

Sub-case B-I: If uVi1,vVi2 for some 1i1i2n2, then u=xi1j1,v=xi2j2 for some 1j1ki1,1j2ki2. Clearly, i1+i2<n. Hence u and v are not adjacent in Γ(Zpn). Clearly, the prime pn2 divides both f(u)=f(xi1j1) and f(v)=f(xi2j2). Therefore, f(u) and f(v) are not coprime.

Sub-case B-II: If uVi1,vVi2 for some n2+1i1i2n1, then u=xi1j1,v=xi2j2 for some 1j1ki1,1j2ki2. Clearly, i1+i2n. Hence u and v are adjacent in Γ(Zpn). By the definition of f, f(u)=f(xi1j1)=pi1j1 and f(v)=f(xi2j2)=pi2j2. As u and v are distinct vertices, pi1j1pi2j2. This implies f(u) and f(v) are coprime.

Sub-case B-III: If uVi1,vVi2 for some 1i1n2,n2+1i2n1, then u=xi1j1, v=xi2j2 for some 1j1ki1,1j2ki2. In this case, u=xi1j1 and v=xi2j2 are adjacent if and only if i1+i2n. If i1+i2n, then i2ni1>ni11. By the definition of f, the primes pi2l does not divide f(xi1j1), for all 1lki2. In particular, the prime pi2j2 does not divide f(xi1j1). As f(xi2j2)=pi2j2, we have f(xi1j1) and f(xi2j2) are coprime. Hence, f(u) and f(v) are coprime. If i1+i2<n, then u and v are not adjacent and i2<ni1 (i2ni11). By the definition of f, pi2j2 divides both f(xi1j1) and f(xi2j2). Therefore, f(u) and f(v) are not coprime.

Thus, f is a coprime labeling of Γ(Zpn).

The number of distinct primes used in the coprime labeling f=|Vn2+1|+|Vn2+2|++|Vn1|+1=(p1)pn(n2+1)1+(p1)pn(n2+2)1+ +(p1)pn(n1)1+1=(p1)pn(n12+1)1+(p1)pn(n12+2)1+ +(p1)pn(n1)1+1=(p1)p(n32)+(p1)p(n52)++(p1)+1=(p1){p(n32)+p(n52)++1}.

+1=pn121+1=pn12=ω(Γ(Zpn)) This implies that μ(Γ(Zpn))ω(Γ(Zpn)). Therefore, by Theorem 1.8, μ(Γ(Zpn))=χ(Γ(Zpn))=ω(Γ(Zpn))=pn12. □

Definition 1.13.

[Citation1, Definition 2.1] An independent set in a graph G is a subset I of the vertex set V(G) of G such that no two vertices of I are adjacent. The independence number of G, denoted by α(G), is defined as the cardinality of a maximum independent set of G.

Theorem 1.14.

[Citation1, Theorem 2] Let Γ(Zpn) be the zero-divisor graph of a ring Zpn, where p is a prime. Then α(Γ(Zpn))={pn1pn2+1if n is even;pn12(pn121)if n is odd.

The details of the proof of Theorem 1.15 can be obtained on similar lines as that of Theorem 1.11. Hence we skipped the detailed proof. The coprime labeling is given.

Theorem 1.15.

Let Γc(Zpn) be the complement of the zero-divisor graph of a ring Zpn, where p is a prime. Then μ(Γc(Zpn))=χ(Γc(Zpn))=ω(Γc(Zpn))={pn1pn2+1if n is even;pn12(pn121)if n is odd.

Proof.

Case A: Let n be an even number.

Consider a map f:V(Γc(Zpn)){2,3,4,} as follows: f(xij)={pijfor 1i(n21),1jki;(pn2)jfor i=n2,1jkn2;[(pn2)(l=1k(n21)p(n21)l)(l=1knip(ni)l)]jfor n2+1in1,1jki,where pij and pn2 are distinct primes. Then, f is a coprime labeling of Γc(Zpn).

The number of distinct primes used in the coprime labeling f=|V1|+|V2|++|Vn21|+1=(p1)pn2+(p1)pn3++(p1)pn2+1=(p1){pn2+pn3++pn2}+1=pn1pn2+1=α(Γ(Zpn))=ω(Γc(Zpn)).

Case B: Let n be an odd number.

Consider a map f:V(Γc(Zpn)){2,3,4,} as follows: f(xij)={pijfor 1in2,1jki;(l=1kn2pn2l)jfor i=n2+1,1jk(n2+1);[(l=1kn2pn2l)(l=1k(n21)p(n21)l)(l=1knip(ni)l)]jforn2+2in1,1jkiwhere pij are distinct primes. Then, f is a coprime labeling of Γc(Zpn).

The number of distinct primes used in the coprime labeling f=|V1|+|V2|++|Vn2|=(p1)pn2+(p1)pn3++(p1)pnn21+1=(p1){pn2+pn3++pn(n12)1}=(p1){pn2+pn3++pn12}=pn1pn12=pn12{pn121}=α(Γ(Zpn))=ω(Γc(Zpn)).

Remark 1.16.

(1) Note that ω(Γ(Zn))=μ(Γ(Zn)) may not be true, if npn. In fact, it is not true for n=12=22·3. Consider the ring Z12. The zero-divisor graph Γ(Z12) and its reduced graph Γred(Z12) are shown in . Clearly, χ(Γ(Z12))=ω(Γ(Z12))=2. Define the function f:V(Γred(Z12)){2,3,} such that f({3¯,9¯})=p1p3, f({4¯,8¯})=p2, f({6¯})=p1, f({2¯,10¯})=p2p3, where p1,p2,p3 are distinct primes. It is easy to verify that f is a coprime labeling of Γred(Z12) and μ(Γred(Z12))=3. Hence, by Theorem 1.5, μ(Γ(Z12))=μ(Γred(Z12))=3. Thus, ω(Γ(Z12))μ(Γ(Z12)).

Fig. 3 Zero-divisor graph of Z12 and its reduced graph.

Fig. 3 Zero-divisor graph of Z12 and its reduced graph.

(2) Also, ω(Γc(Zn))=μ(Γc(Zn)) may not be true, if npn. Consider the same example. Clearly, ω(Γc(Z12))=4. It is easy to check that θ1((Γ(Z12))=8. Therefore, μ(Γc(Z12))=8. Thus, ω(Γc(Z12))μ(Γc(Z12)).

(3) Using similar techniques used in Theorems 1.12 and 1.15, it can be proved that the coprime index and the clique number coincide for the zero-divisor graph and the complement of the zero-divisor graph of a finite commutative special principal ideal ring.

We need the following definitions given in [Citation10].

Let P be a poset. Given any AP, the upper cone of A is given by Au={bP : ab for every aA} and the lower cone of A is given by Al={bP : ba for every aA}. If aP, then the sets {a}u and {a}l will be denoted by a u and al, respectively. By Aul, we mean {Au}l. Dually, we have the notion of Alu.

Suppose that P is a poset with 0. If AP, then the annihilator of A is given by A={bP:{a,b}l={0}  for all  aA}, and if A={a}, then we write a=A. An element aP is an atom if a > 0 and {bP : 0<b<a}=, and P is called atomic if for every bP{0}, there exists an atom aP such that ab.

The direct product of posets P1,,Pn is the poset P=i=1nPi with defined such that ab in P if and only if aibi (in Pi) for every i{1,,n}, where a=(a1,a2,,an), b=(b1,b2,,bn). For any Ai=1nPi, note that Au={bi=1nPi : aibi for every aA and i{1,,n}}. Similarly, Al={bi=1nPi : biai for every aA and i{1,,n}}.

Let P=i=1nPi,(n2), where Pi’s are finite bounded posets such that Z(Pi)={0},i. Since Pi’s are finite bounded poset, Pi’s are atomic. Note that Z(Pi)=0 if and only if Pi has the unique atom. Further, assume that qiPi is the unique atom Pi for every i. It is not difficult to prove that (q1,0,,0),(0,q2,0,,0), ,(0,,0,qn) are the only n atoms of P.

A poset P is called bounded if P has both the least element 0 and the greatest element 1.

Let P be a poset with 0. Define a zero-divisor of P to be any element of the set Z(P)={aP | there exists bP{0} such that {a,b}l={0}}. As in [Citation19], the zero-divisor graph of P is the graph G(P) whose vertices are the elements of Z(P){0} such that two vertices a and b are adjacent if and only if {a,b}l={0}.

We set D=PZ(P). The elements dD are the dense elements of P. The zero-divisor graph G*(P) of P with vertex set is P{0,1} and a and b are adjacent if and only if {a,b}l={0}.

Throughout, P denotes a poset with 0 and qi, i{1,2,,n} are all atoms of P, where n1.

Afkhami et al. [Citation3] partitioned the set P{0} as follows.

Let 1i1<i2<<ikn. The notation Pi1i2ik stands for the set: {xP : x(s=1k{qis}u)\(j{1,2,,n}\{i1,i2,,ik}{qj}u)}.

In [Citation3], the following observations are proved.

  1. Let {i1,,ik} and {j1,,jk} be the index sets of Pi1i2ik and Pj1j2jk, respectively. Then (Pi1i2ik)(Pj1j2jk)=, if {i1,,ik}{j1,,jk}.

  2. P\{0}=1i1<i2<<iknk=1,n.Pi1i2ik.

From the above two properties, it is clear that the sets Pi1i2ik forms a partition of P{0}. In fact, one can see that a relation on P{0} defined as follows: xy if and only if x,yPi1i2ik for some partition Pi1i2ik of P{0} is an equivalence relation.

The following discussion can be found in [Citation17].

The set of equivalence classes under of P{0} will be denoted by [P]={ Pi1i2ik:{i1,i2,ik}{1,2,,n} andPi1i2ik  }. Now, we set [P]=[P]P0. We define relation on [P] as follows. Pi1i2ikPj1j2jm if and only if ba, for some aPi1i2ik and for some bPj1j2jm, where Pi1i2ik, Pj1j2jm[P] and P0Pi1i2ik for all {i1,i2,,ik}{1,2,,n}. It is not very difficult to prove that ([P], ) is a poset. The least element of ([P],) is P0, and if P has the greatest element 1, then the greatest element of the poset ([P],) is P12n. Sometimes, we abuse the notation and write Pi1i2ik as PI, where I={i1,i2,,ik}

Since [P] is a poset with the least element [0]=P0, and except [d] (where dD), every element of [P] is a zero-divisor. Note that D may be empty. Hence the zero-divisor graphs G([P]) and G*([P]) of the poset [P] are same, that is, G([P])=G*([P]). Hence afterwards, we write G*([P]) for the zero-divisor graph of [P]. Clearly, a and b are adjacent in G(P) if and only if [a] and [b] are adjacent in G*([P]). More about the inter-relationship between the properties of G(P) and G*([P]) are mentioned in Lemma 1.17.

The following statements are essentially proved in [Citation17] and will be used frequently in the sequel.

Lemma 1.17.

[17, Lemma 2.11] The following statements are true.

  1. If q1,q2,,qn are distinct atoms of P, then P1,,Pn are distinct atoms of [P].

  2. The elements Pi1i2ikPj1j2jm in [P] if and only if {i1,i2,,ik}{j1,j2,,jm}.

  3. The elements Pi1i2ik and Pj1j2jm are adjacent in G*([P]) if and only if {i1,i2,,ik}{j1,j2,,jm}=. Further, aV(G*(P)) if and only if [a]V(G*([P])).

  4. Let Pi1i2ikV(G*([P])). Then for any x,yPi1i2ik,{x,y}l{0} in P. Hence, the vertices of Pi1i2ik forms an independent set in G*(P). Further, if {Pi1i2ik,Pj1j2jm}l={[0]} in [P], then for any xPi1i2ik and for any yPj1j2jm,{x,y}l={0} in P. In particular, Pi1i2ik and Pj1j2jm are adjacent in G*([P]) with |Pi1i2ik|=r,|Pj1j2jm|=s, then the vertices of Pi1i2ik and Pj1j2jm forms an induced complete bipartite subgraph Kr,s of G*(P).

Theorem 1.18.

(Joshi [Citation15, Corollary 2.11]) Let G(P) be the zero-divisor graph of an atomic poset P. Then χ(G(P))=ω(G(P))=n, where n is the finite number of atoms of P.

Theorem 1.19.

Let P be a finite poset with 0 and G(P) be its zero-divisor graph. Then μ(G(P))=χ(G(P))=ω(G(P))= Number of atoms of P.

Proof.

Let q1,q2,,qn be the all n atoms of P. Hence, [P] will also have exactly n atoms P1,P2,,Pn. Then V(G([P]))={PI:()I{1,2,n},PI}. Since the zero-divisor graph G([P]) is nothing but the (G(P))red and χ(G([P]))=n. Therefore, to prove χ(G(P))=μ(G(P)), it is enough to prove that μ(G([P]))=n. Define a map f:V(G([P])){2,3,4,} by f(PI)=iIpi, where p1,p2,,pn are distinct primes. We show that f is a coprime labeling of G([P]). By Lemma 1.17(3), PI and PJ are adjacent in G([P]) if and only if IJ= if and only if iIpi and jJpj are coprime if and only if f(PI) and f(PJ) are coprime. This proves f is a coprime labeling of G([P]) and μ(G([P]))=n, i.e., μ(G([P]))=n=μ(G(P)), by Theorem 1.5. The result follows from Theorem 1.18. □

Remark 1.20.

Note that μ(Gc(P))=ω(Gc(P)) may not be true. Consider the Boolean lattice L=23 given in . It is easy to check that ω(Gc(L))=3, however, μ(Gc(L))=4.

Fig. 4 Coprime labeling of the complement of zero-divisor graph.

Fig. 4 Coprime labeling of the complement of zero-divisor graph.

Definition 1.21.

(West [Citation23]) The disjoint union of graphs is an operation that combines two or more graphs to form a larger graph. It is analogous to the disjoint union of sets and is constructed by making the vertex set of the result be the disjoint union of the vertex sets of the given graphs and by making the edge set of the result be the disjoint union of the edge sets of the given graphs. Any disjoint union of two or more nonempty graphs is necessarily disconnected. If G and H are two graphs, then GH denotes their disjoint union.

Remark 1.22.

It is easy to observe that for any graph G, μ(GIm)=μ(G), where Im is an independent set with m vertices. In particular, μ(G*(P))=μ(G(P)Im)=μ(G(P)).

2 Applications

In this section, we show that various graphs associated with commutative rings with identity can be studied via zero-divisor graphs of specially constructed posets.

I. Zero-divisor graphs of finite reduced commutative rings

In [Citation10], it is observed that if R is a reduced Artinian ring with exactly k prime ideals, then there exist fields F1,,Fk such that RF1××Fk. Further, it is proved that the ring-theoretic zero-divisor graph Γ(R)=Γ(i=1kFi) equals the poset-theoretic zero-divisor graph of R (R treated as a poset under the partial order given in [10, Lemma 3.3]), that is, Γ(i=1kFi)=G(i=1kFi) (cf. [Citation18, Remark 3.4]).

Corollary 2.1. Let R be a finite reduced commutative ring with identity and Γ(R) be its zero-divisor graph. Then μ(Γ(R))=χ(Γ(R))=ω(Γ(R))= Number of prime ideals of R.

II. Annihilating ideal graphs, Co-annihilating ideal graphs and Comaximal ideal graphs of commutative rings

Let R be a commutative ring with identity that is not an integral domain. An ideal I of a ring R is called an annihilating ideal if there exists rR{0} such that Ir=(0).

Definition 2.2.

(Visweswaran and Patel [Citation22]) Let R be a commutative ring with identity. Associate a simple undirected graph with R, called as the annihilating ideal graph AG(R) of R, where the vertices of AG(R) are nonzero annihilating ideals of R and two vertices I and J are adjacent if and only if I + J is also an annihilating ideal of R.

The following is the modified definition of an annihilating ideal graph. The modified annihilating ideal graph AG*(R) be a simple undirected graph with vertex set V(AG*(R)) is set of all nonzero proper ideals of R and two distinct vertices I, J are joined if and only if I + J is also an annihilating ideal of R. By [Citation2, Lemma 2.1], IJ is an edge of AG(R) if and only if Ann(I) Ann(J)(0).

Let Id(R) be the set of all ideals of R. Consider an equivalence relation on Id(R) such that IJ if and only if Ann(I)=Ann(J). The equivalence class of an ideal I is the set {JId(R) : IJ}, denoted by [I]. The set of all equivalence classes of Id(R) is { [I] : IId(R)  }, denoted by [Id(R)]. It is easy to observe that [(0)]={(0)} and Ann(I)=(0), if I[R].

Now, we define a relation on [Id(R)] such that [I][J] if and only if Ann(I)Ann(J). It is relatively easy to prove that is a partial order on [Id(R)]. Thus, ([Id(R)],) is a poset. Since Ann((0))=R and Ann(R)=(0), then [(0)] is the 1 and [R] is the 0 of ([Id(R)],).

We construct a poset P from the poset ([Id(R)],). The elements [(0)] and [R] of the poset ([Id(R)],) are denoted by (0) and R respectively, in the poset P. For any element [I]{[(0)],[R]} of ([Id(R)],), we replace the element [I] of ([Id(R)],) by the chain of elements of the equivalence class [I] in P whose length is equal to |[I]| in some pre-determined well-order. Note that the element [R] of ([Id(R)],) is replaced by the one element chain, namely R, in P irrespective of the cardinality of [R]. Clearly, every element I{0,1} is an annihilating ideal of R. We call this poset the corresponding poset of the ring R.

We illustrate the above discussion with an example. Let R=F2[x,y](x,y)2, where F2 is the field of two elements. We can check that, Id(R)={(0),(x),(y),(x+y),(x,y),R}. Also, [(x)]=[(y)]=[(x+y)]=[(x,y)]. This gives, [Id(R)]={[(0)],[(x)],[R]}. The Hasse diagrams of the ideal lattice (Id(R),),([Id(R)],) and the constructed poset P are shown in .

Fig. 5 The ideal lattice and the corresponding poset.

Fig. 5 The ideal lattice and the corresponding poset.

From the construction mentioned above, the greatest element 1 and the least element 0 of a poset P are the ideals (0) and R, respectively. Let I,JP. Clearly, I and J are comparable in P if and only if either I and J in the same equivalence class, or [I] and [J] are comparable in ([Id(R)],).

Remark 2.3.

(1) We observe that P is a meet-semi lattice. For this, let I,JP. Since [I+J]=inf{[I],[J]}, we have ([Id(R)],) is a meet-semi lattice. Further, the poset P is constructed from ([Id(R)],) by replacing the elements of ([Id(R)],) by the chains. Hence P is a meet-semi lattice.

(2) Let R be an Artinian ring. Hence Ann(I)=(0) for every proper ideal I of R. Then the number of atoms of P is equal to the number of maximal ideals of R. For this, we prove that [Id(R)] has |Max(R)| number of atoms given by {[M] : MMax(R)}. Let Mi,MjMax(R). Then Ann(Mi+Mj)=Ann(R)=(0)=Ann(Mi)Ann(Mj). Hence Ann(Mi)Ann(Mj) and Ann(Mj)Ann(Mi). Therefore, [Mi][Mj] and [Mj][Mi]. Also, for any (R)IId(R),IM for some MMax(R). Therefore, Ann(M)Ann(I). Hence [R]=[M][I]. This proves {[M] : MMax(R)} are all distinct atoms of [Id(R)]. Thus, by the construction of P, the number of atoms of P is equal to the number of maximal ideals of R.

The elements I and J are adjacent in G*(P) if and only if {I,J}l={0} for I,JV(G*(P)). By Remark 2.3, I and J are adjacent in G*(P) if and only if Ann(I)Ann(J)=Ann(R)=(0). Note that, we can easily verify AG(R)=G*c(P) for the above example. With this preparation, we prove that the complement of an annihilating ideal graph of R is nothing but the zero-divisor graph of P.

Lemma 2.4.

Let R be a commutative ring with identity that is not an integral domain, and let P be the corresponding poset of R. Then AGc(R)=G*(P).

Proof.

First we prove that the vertex set of AGc(R) and G*(P) are equal. For this, let IV(AGc(R)). Then I is an annihilating ideal of R, and I is an element of P. Clearly, I[(0)] and I[R]. This gives that I{0,1}. Therefore IV(G*(P)). This proves V(AGc(R))V(G*(P)).

Now, assume that IV(G*(P)). Clearly, I{0,1}. By the construction of P, we have [I][(0)] and [I][R]. Thus Ann(I)Ann(R) and Ann(I)Ann((0)). This gives Ann(R)=(0)Ann(I)R=Ann((0)). As already observed I{0,1} is an annihilating ideal of R, IV(AGc(R)). This proves V(G*(P))V(AGc(R)). Thus, V(AGc(R))=V(G*(P)).

Now, we prove that I and J are adjacent in AGc(R) if and only if I and J is adjacent in G*(P). For this, suppose that I and J are adjacent in AGc(R). By (1), IJ is an edge of AGc(R) if and only if Ann(I)Ann(J)=(0)=Ann(I+J). This implies inf{I,J}=IJ=R, where R is the zero of the meet-semi lattice P. Thus, I and J are adjacent in G*(P).

Assume that I and J are adjacent in G*(P). This implies Ann(I)Ann(J)=(0). Thus, I and J are adjacent in AGc(R). This proves AGc(R)=G*(P). □

Definition 2.5.

(Akbari et al. [Citation4]) The co-annihilating-ideal graph of R, denoted by CAG*(R), is a graph whose vertex set is the set of all nonzero proper ideals of R and two distinct vertices I and J are adjacent whenever Ann (I) Ann (J)=(0).

Definition 2.6.

(Ye and Wu [Citation24]) Let R be a commutative ring with identity. We associate a simple undirected graph with R, called the comaximal ideal graph CG(R) of R, where the vertices of CG(R) are the proper ideals of R that are not contained in the Jacobson radical J(R) of R and two vertices I and J are adjacent if and only if I + J = R.

The following is the modified definition of the comaximal ideal graph. The modified comaximal ideal graph CG*(R) is the graph with the vertex set as the nonzero proper ideals of R and two vertices I and J are adjacent if and only if I and J are comaximal. Clearly, one can see that the comaximal ideal graph CG(R) is the subgraph of the modified comaximal ideal graph CG*(R). Moreover, CG*(R)=CG(R)Im, where m=|J(R)|1.

Corollary 2.7.

[Citation4, Corollary 1.2] If R is an Artinian ring, then CG*(R)=CG(R)Im=CAG*(R)=AG*c(R)=AGc(R), where m=|J(R)|1.

By Theorem 1.19, Remarks 1.22, 2.3(2), Lemma 2.4, Corollary 2.7, we have:

Corollary 2.8.

Let R be an Artinian ring with finitely many ideals. Then μ(CG(R))=μ(CAG*(R))=μ(AG*c(R))=|Max(R)|.

III. Intersection graphs of ideals of Artinian principal ideal rings

Let R be a commutative ring with identity. Then the intersection graph of ideals IG(R) of R is the graph whose vertices are the nonzero proper ideals of R such that distinct vertices I and J are adjacent if and only if IJ{0}; see [Citation9]. The vertex set of the zero-divisor graph G*(Id(R)) of a lattice Id(R) is Id(R){0Id(R),1Id(R)}, that is, the set of nonzero proper ideals of R. Therefore V(IG(R))=V(G*(Id(R)))=V(G*c(Id(R))). The ideals I and J are adjacent in G*(Id(R)) if and only if IJ=0Id(R), that is, the ideals I and J are adjacent in G*(Id(R)) if and only if IJ=(0). The ideals I and J are adjacent in G*c(Id(R)) if and only if IJ(0). Hence the following result.

Lemma 2.9.

Let R be a commutative ring with identity. Then IG(R)=G*c(Id(R)).

The following discussion can be found in [Citation10].

Let R be a commutative ring with identity. Recall that R is a special principal ideal ring (SPIR for brevity) if R is a local Artinian principal ideal ring (cf. [Citation14]). If R is an SPIR with maximal ideal M, then there exists nN such that Mn={0},Mn1{0}, and if I is an ideal of R, then I=Mi for some i{0,1,,n} ([14, Proposition 4]). In this case, M is nilpotent with an index of nilpotency equal to n, and the lattice of ideals of R is isomorphic to the chain C of length n (chain of n + 1 elements). By [14, Lemma 10], R is an Artinian principal ideal ring if and only if there exist SPIRs R1,,Rn such that RR1××Rn (it is also a straightforward consequence of the structure theorem of Artinian rings in [Citation6, Theorem 8.7]). Thus, Id(R) of an Artinian principal ideal ring is product of chains C=i=1nCi (Ci is chain of length ni), where RR1××Rn, nilpotency index of maximal ideal Mi of Ri is ni. By Lemma 2.9 and from the above discussion, we have the following corollary.

Corollary 2.10.

Let R be an Artinian principal ideal ring. Then μ(IGc(R))=|Max(R)|.

Acknowledgments

The authors are grateful to the referees for their suggestions which improved the presentation of the paper.

Additional information

Funding

The second author is financially supported by the Council of Scientific and Industrial Research(CSIR), New Delhi, via Senior Research Fellowship Award Letter No. 09/137(0620)/2019-EMR-I.

References

  • Abd Al Jawad, E. E., Al-Ezeh, H. (2008). Domination and independence numbers of Γ(Zn). Int. Math. Forum 3(11): 503–511.
  • Aghapouramin, V., Nikmehr, M. J. (2018). Perfectness of a graph associated with annihilating ideals of a ring. Discrete Math. Algorithms Appl. 10(4): 1850047–11 pages).
  • Afkhami, M., Barati, Z., Khashyarmanesh, K. (2012). Planar zero divisor graphs of partially ordered sets. Acta Math. Hungarica 137(1–2): 27–35.
  • Akbari, S., Alilou, A., Amjadi, J., heikholeslami, S. M. (2017). The co-annihilating-ideal graphs of commutative rings. Can. Math. Bull. 60(1): 3–11.
  • Anderson, D. F., Livingston, P. S. (1999). The zero-divisor graph of a commutative ring. J. Algebra 217: 434–447.
  • Atiyah, M. F., MacDonald, I. G. (1969). Introduction to CommutativeAlgebra. Reading, MA: Addison-Wesley.
  • Bagheri, S., Nabaei, F., Rezaeii, R., Samei, K. (2018). Reduction graph and its application on algebraic graphs. Rocky Mountain J. Math. 48(3): 729–751.
  • Beck, I. (1988). Coloring of a commutative ring. J. Algebra 116:208–226.
  • Chakrabarty, I., Ghosh, S., Mukherjee, T. K., Sen, M. K. (2009). Intersection graphs of ideals of rings. Discrete Math. 309(17): 5381–5392.
  • Devhare, S., Joshi, V., LaGrange, J. D. (2019). Eulerian and Hamiltonian complements of zero-divisor graphs of pseudocomplemented posets. Palestine J. Math. 8(2): 30–39.
  • Erdös, P., Goodman, A., Posa, L. (1966). The representation of a graph by set intersections. Can. J. Math. 18: 106–112.
  • Gallian, J. A. (2015). A dynamic survey of graph labelings. Electron. J. Comb. # DS6.
  • Holyer, I. (1981). The NP-completeness of some edge-partition problems. SIAM J. Comput. 10(4): 713–717.
  • Hungerford, T. W. (1968). On the structure of principal ideal rings. Pac. J. Math. 25(3): 543–547.
  • Joshi, V. (2012). Zero divisor graph of a poset with respect to an ideal. Order 29: 499–506.
  • Katre, S. A., Yahyaei, L., Arumugam, S. (2017). Coprime index of a graphs. Electron. Notes Discrete Math. 60(6): 77–82.
  • Khandekar, N., Joshi, V. Coloring of zero-divisor graphs of posets and applications to graphs associated with algebraic structures, (Preprint).
  • LaGrange, J. D., Roy, K. A. (2013). Poset graphs and the lattice of graph annihilators. Discrete Math. 313(10): 1053–1062.
  • Lu, D., Wu, T. (2010). The zero-divisor graphs of partially orderedsets and an application to semigroups. Graphs Comb. 26: 793–804.
  • Pirzada, S., Aijaz, M., Imran Bhat, M. (2020). On the zero divisor graphs of Zn. Afr. Mat. 31(3–4): 727–737.
  • Roberts, F. S. (1985). Applications of edge coverings by cliques. Discrete Applied Math. 10(1): 93–109.
  • Visweswaran, S., Patel, H. D. (2014). A graph associated with the set of all nonzero annihilating ideals of a commutative ring. Discrete Math. Algorithms Appl. 6(4): 1450047–22 pages).
  • West, D. B. (2001). Introduction to Graph Theory, 2nd ed. Hoboken, NJ: Prentice Hall.
  • Ye, M., Wu, T. S. (2012). Comaximal ideal graphs of commutative rings. J. Algebra Appl. 11(6): 1250114–14 pages).