210
Views
0
CrossRef citations to date
0
Altmetric
Research Article

A characterization of star-perfect graphs

ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon
Received 12 Sep 2022, Accepted 07 Mar 2024, Published online: 02 Apr 2024

Abstract

Motivated by Berge perfect graphs, we define star-perfect graphs and characterize them. For a finite simple graph G(V, E), let θs(G) denote the minimum number of induced stars contained in G such that the union of their vertex sets is V(G), and let αs(G) denote the maximum number of vertices in G such that no two of them are contained in the same induced star of G. We call a graph G star-perfect if αs(H)=θs(H), for every induced subgraph H of G. A graph G is star-perfect if and only if G is (C3,C3k+1,C3k+2)-free, for every k1. A bipartite graph G is star-perfect if and only if every induced cycle in G is of length 6k,k1. The minimum parameter θs(G) and the maximum parameter αs(G) have been extensively studied in various contexts.

Mathematics Subject Classification:

1 Introduction

All our graphs are finite and simple. We follow West [Citation11] and Bondy and Murty [Citation3] for basic terminologies and notations. If G is a graph, then V(G) and E(G) are its vertex set and edge set, respectively. As usual, Kn(n1), Cn(n3), Pn(n1) denote the complete graph, cycle and path on n vertices, respectively. If DV(G), the subgraph of G induced by D is obtained by deleting the vertices of V(G)D and is denoted by G[D]. Given a graph H, we say that a graph G is H-free, if G does not contain an induced subgraph isomorphic to H. Given a family {H1,H2,} of graphs, we say that G is (H1,H2,)-free if G is Hi - free, for every i1.

The neighborhood of a vertex v is the set of vertices adjacent to v and is denoted by NG(v) (or N(v)). The closed neighborhood of v is the set NG(v){v} and is denoted by NG[v] (or N[v]). Suppose H1 and H2 are two induced subgraphs in G. We say that H1 and H2 are adjacent in G, and is denoted by H1H2, if either (i) there exists at least a common vertex between H1 and H2 (V(H1)V(H2)), or (ii) there is an edge in G such that one of the end vertices of the edge is in H1 and the other is in H2. In a graph G, the distance between two vertices u and v, denoted by d(u, v), is the length of a shortest path connecting them. Let A,BV(G). The distance between A and B, denoted by d(A, B), is the length of a shortest (a, b)-path where aA and bB. The symbol [A,B] denotes the set {eE(G):e has one end vertex in A and the other end vertex in B}.

A star is a graph on t + 1 vertices, t0, say v0,v1,v2,,vt such that v0 is adjacent to every vertex in W={v1,v2,,vt} and no two vertices in W are adjacent. The vertex v0 is called the central vertex of the star and we say that the star is centered at v0 and all other vertices of the star are non-central vertices. Note that K1 and K2 are stars. Berge’s perfect graphs [Citation2] have inspired various researchers to study many variations of perfect graphs [Citation5]. Ravindra introduced the concept of F-perfect graphs [Citation10]. Litjens et al. defined and characterized sum-perfect graphs [Citation8].

Let G be a graph. A star-cover of G, denoted by S, is a family of induced stars contained in G such that every vertex of G is in one of the induced stars. The star-covering number of G is the minimum cardinality of a star-cover and is denoted by θs(G). A θs-cover of G is a star-cover of G containing θs(G) stars.

A set TV(G) is a star-independent set of G if no two vertices of T are contained in the same induced star in G (equivalently, any two vertices in T are at a distance at least 3). The maximum cardinality of a star-independent set is called the star-independence number of G and is denoted by αs(G). An αs- independent set of G is a star-independent set with αs(G) vertices.

The minimum parameter θs(G) and the maximum parameter αs(G) have been extensively studied in various contexts with different terminology. For example, star-independent sets are studied in [Citation4], where a star-independent set in G corresponds to a color class in a (k,3)-coloring of G. It is also interesting to note that for a triangle free graph G, the central vertices of stars in star cover of G is a dominating set of G (Recall that a set DV(G) is called a dominating set of G if every vertex in V(G)D is adjacent to at least one vertex in D). So for a triangle free graph G, θs(G) is the minimum dominating set γ(G) of G. Domination is a much studied topic, for example see [Citation1] and references therein. It is well known that DOMINATION PROBLEM is NP-complete for triangle free graphs and several subfamilies of triangle free graphs; see [Citation6]. By our earlier remark that θs(G) is γ(G) for triangle free graphs, the decision problem of finding θs(G) is also NP-complete for triangle free graphs.

The union of vertex disjoint graphs G1,G2,,Gk, written i=1kGi is the graph with vertex set i=1kV(Gi) and edge set i=1kE(Gi).

It is easy to see that the graph parameters αs(G) and θs(G) both satisfy the additive property, that is

  • αs(i=1kGi)=i=1kαs(Gi), and

  • θs(i=1kGi)=i=1kθs(Gi).

Clearly, for any graph G, θs(G)αs(G). We call a graph G a star-perfect graph if θs(H)=αs(H), for every induced subgraph H of G. We call a graph star-imperfect if it is not star-perfect. For example, the cycle C6 is a star-perfect graph, since θs(C6)=2=αs(C6) and θs(H)=αs(H) for every proper induced subgraph H of G, while C4 is a star-imperfect graph, since θs(C4)=2 and αs(C4)=1. Similarly, C3 is star-imperfect graph, since θs(C3)=2 and αs(C3)=1.

A graph G is said to be minimal star-imperfect if (i) θs(G)αs(G) and (ii) Gv is star-perfect for every vertex v of G. Every minimal star-imperfect graph is a connected graph, since a graph G is star-perfect if and only if every component of G is star-perfect. Moreover, every star-imperfect graph contains a minimal star-imperfect graph.

Since star-perfectness is a hereditary property, one expects a forbidden graph characterization for star perfect graphs. We realize this in this paper as given below.

Main Theorem: A graph G is star-perfect if and only if G is (C3,C3k+1,C3k+2)-free, k1.

The theorem is a min-max theorem for star-perfect graphs. Our characterization goes parallel to a characterization of β-perfect graphs due to Markossyan, Gasparyan and Reed [Citation9]. They have proved that the graphs G and G¯ are β-perfect if and only if G and G¯ are even-hole-free (C2k-free, k2). Yet another concept that comes close to our concept of star-perfectness is the concept of neighborhood perfect graphs due to Lehel and Tuza [Citation7].

For ready reference, we collect in one place the observations and remarks made above. Throughout the paper we often use these properties without explicitly referring to them.

  • (1.1)The graph parameters αs(G) and θs(G) both satisfy the additive property.

  • (1.2)For any graph G, θs(G)αs(G). So, if for some graph H, θs(G)αs(H), then αs(H)=θs(H).

  • (1.3)Every minimal star-imperfect graph is a connected graph, since a graph G is star-perfect if and only if every component of G is star-perfect.

  • (1.4) Every star-imperfect graph contains a minimal star-imperfect graph.

  • (1.5)If G is a star-perfect graph and S is a θs -cover and I is an αs -independent set of G, then every star in S contains exactly one vertex of I.

Before proving the main theorem we establish a few properties of minimal star-imperfect graphs.

2 Minimal star-imperfect graphs

Throughout the paper, we fix a clockwise orientation to a cycle C in G.

Lemma 2.1.

Let k be any positive integer. Then,

  1. θs(Pk)=k3 and αs(Pk)=k3.

  2. Pk is star-perfect.

  3. The disjoint union of paths is a star-perfect graph.

Proof.

Let Pk be the path v1v2vk,k1.

(i) If k0(mod3), let k=3p,p1. Then the family of stars {v1v2v3,v4v5v6,v7v8v9,v3p2v3p1v3p} is a star cover of Pk , so θs(Pk)k. On the other hand, every star contained in Pk has at most three vertices, therefore, θs(Pk)p. Combining the two inequalities, we have θs(Pk)=p=k3. Let I be any star-independent set in P. Since any two vertices in I are at a distance at least 3, |I|p=k3. So αs(Pk)p. The set T={v2,v5,,v3p1} is a star-independent set of Pk , since any two vertices in T are at a distance at least 3 in Pk . Thus αs(Pk)|T|=p=k3. Combining the two inequalities, we have αs(Pk)=k3.

If k1(mod3), let k=3p+1,p0. Then the family of stars {v1v2v3,v4v5v6,v7v8v9,v3p2v3p1v3p,[{v3p+1}]} is a star cover of Pk , so θs(Pk)p+1=k3. On the other hand, every star contained in Pk has at most three vertices, therefore, θs(Pk)k3=3p+13=p+1. Combining the two inequalities, we have θs(Pk)=p+1=k3. Let I be any star-independent set in P. Since any two vertices in I are at a distance at least 3, |I|p+1=k3. So αs(Pk)p+1. The set T={v2,v5,,v3p+1} is a star-independent set of Pk , since any two vertices in T are at a distance at least 3 in Pk . Thus αs(Pk)|T|=p+1=k3. Combining the two inequalities, we have αs(Pk)=k+1=k3 and θs(Pk)=k+1=k3.

If k2(mod3) one can show that θs(Pk)=k3 by using similar arguments as above.

Combining the three cases we have θs(Pk)=k3 and αs(Pk)=k3.

(ii) Every proper induced subgraph of P is a disjoint union of paths. Therefore, by (i), and additive property it follows that Pk is star-perfect.

(iii) The assertion follows since a graph G is star-perfect if and only if every component of G is star-perfect. □

Lemma 2.2.

For any integer k2 the following statements are true.

  1. θs(C3k)=k and αs(C3k)=k.

  2. C3k is star-perfect.

Proof.

(i) Let C be a 3k-cycle v1v2v3kv1. We first prove that θs(C)=k. Every star contained in C has at most 3 vertices, so θs(C)3k3=k. Next the set of stars S={v3kv1v2,v3v4v5,,v3k3v3k2v3k1} is a star-cover of C with k stars. Hence θs(C)k. Combining the two inequalities, we have θs(G)=k.

Next we prove that αs(C3k)=k. Let I be any star-independent set in C. Since any two vertices in I are at a distance at least 3, |I|3k3=k. So αs(C)k. The set T={v1,v4,,v3k2} is a star-independent set of C, since any two vertices in T are at a distance at least 3 in C. Thus k=|T|αs(C). Combining the two inequalities, we have αs(C)=k.

Therefore, αs(C)=k, and θs(C)=k.

(ii) Every proper induced subgraph of C is a disjoint union of paths. Therefore, by Lemma 2.1 C3k is star-perfect. □

In our next lemma, we identify minimal star-imperfect cycles.

Lemma 2.3.

Let k be any positive integer. Then,

  1. θs(C3)=2, αs(C3)=1; θs(C3k+1)=k+1, αs(C3k+1)=k; θs(C3k+2)=k+1 and αs(C3k+2)=k.

  2. The cycles C3, C3k+1 and C3k+2 are star-imperfect graphs.

  3. The cycles C3, C3k+1 and C3k+2 are minimal star-imperfect graphs.

Proof.

We prove (i) in three steps depending on the length of the cycles.

Step 1: For C3, αs(C3)=1 and θs(C3)=2. And every proper induced subgraph of C3 is K1 or K2, each of which is star-perfect. Therefore, C3 is minimal star-imperfect.

Step 2: Let C be a (3k+1)-cycle v1v2v3k+1v1. We claim that θs(C)=k+1 and αs(C)=k. Every star contained in C has at most 3 vertices. Hence, θs(C)3k+13=k+1. Next the family of stars {v3k+1v1v2,v3v4v5,,v3k3v3k2v3k1,[{v3k}]} is a star-cover of C with k + 1 stars. Hence, θs(C)k+1. Combining the two inequalities, we have θs(G)=k+1.

Next, let I be any star-independent set in C. Since any two vertices in I are at distance at least 3, we deduce |I|3k+13=k. So αs(C)k. The set T={v1,v4,,v3k2} is a star-independent set of C, since any two vertices in T are at distance at least 3 in C. Thus, k=|T|αs(C). Combining the two inequalities, we have αs(C)=k and hence our claim is proved.

Step 3: We can show that C3k+2 is a minimal star-imperfect graph, by using similar arguments as given in Step 2.

(ii) is a consequence of (i).

(iii) Every proper induced subgraph of C is a disjoint union of paths and so by Lemma 2.1 (iii) it follows that every induced subgraph is star-perfect. We conclude that C is a minimal star-imperfect graph. □

Our next goal is to show that a minimal star-imperfect graph is a block.

Lemma 2.4.

Let G be a connected star-perfect graph and v be a vertex of G. Then one of the following is true.

  1. The vertex v is in an αs -independent set of G.

  2. A neighbor of v is in an αs -independent set of G.

Proof.

If (i) is not true, then there exists an αs -independent set S={u1,u2,,uk} that does not contain v, where k=αs(G). We claim that [N(v),S]ϕ. Assume the contrary that [N(v),S]=ϕ. Therefore, d(N(v),S)2. Let P be a shortest path between N(v) and S. By our assumption, length (P)2. Since P is a shortest path, d(v,S)3, for all uS. Therefore, {v}S is a star-independent set in G of size αs(G)+1, a contradiction. Hence our claim [N(v),S]ϕ holds.

Since [N(v),S]ϕ, let without loss of generality w1u1[N(v),S]. We assert that d(w1,S{u1})3. If not, then d(w1,ui)2, for some uiS{u1}. If d(w1,ui)=1, then u1 and ui belong to a star and therefore, they are not star-independent, a contradiction. Next, if d(w1,ui)=2, then w1 and ur , for some r such that 2rk are adjacent to a vertex in G. Let Q(w1,ur) be a shortest path of length 2.

The vertex v is at a distance at most 2 from some vertex ur,2rk. If not, d(v,ui)3, for all ui,2ik. Therefore, d(v,ui)3 and so {v}{u2,u3,,uk} is an αs -independent set in G, a contradiction to our first assumption. So let R(ur,v) be a shortest path of length 2. Then w1QurRvw1 is an induced C5, a contradiction to the assumption that G is star-perfect. Hence, d(w1,S{u1})3 and our assertion is true. Thus, {w1,u2,,uk} is an αs -independent set of G, and so (ii) holds. This completes the proof. □

Note. The condition that G is star-perfect is necessary. For example, see .

Fig. 1 Bull graph.

Fig. 1 Bull graph.

Bull graph is star-imperfect and {u1,u2} is the only αs -independent set and neither v nor any of its neighbors {w1,w2} is in an αs -independent set of bull graph.

Lemma 2.5.

If G is a minimal star-imperfect graph and v is a cut vertex of G, then every component of Gv has at least two vertices.

Proof.

Let G1,G2,,Gk be the components of Gv where k2. Contrary to the lemma, assume that at least one of the components, say G1, has exactly one vertex u. The star uv together with θs -cover of i=2kGi forms a θs -cover of G. So θs(G)i=2kθs(Gi)+1. The star K1=[{u}] together with θs -cover of i=2kGi forms a star-cover of Gv, since u is isolated in Gv. Thus, i=2kθs(Gi)+1=θs(Gv) and we have θs(G)θs(Gv). The vertex u together with αs -independent set of i=2kGi is a star-independent set of G. Then αs(G)i=2kαs(Gi)+1=αs(Gv). These two inequalities imply that θs(G)θs(Gv)=αs(Gv),since Gv is star‐perfectαs(G)

Next since αs(G)θs(G) is always true, we have αs(G)=θs(G). On the other hand since G is a minimal star-imperfect graph, we have θs(G)αs(G). We have arrived at a contradiction. □

Lemma 2.6.

If G is a minimal star-imperfect graph, then G is a block.

Proof.

G is connected by (1.3). Contrary to the lemma, let v be a cut vertex in G and let C1,C2,,Ck be the components of Gv, where k2. By Lemma 2.5 none of Ci is K1. Let G1 = C1 and G2=i=2kCi. Let L1 be the subgraph induced by [V(G1){v}] and let L2 be the subgraph induced by [V(G2){v}]. Let H1=L1NG[v] and H2=L2NG[v]. Since G is minimal star-imperfect, we have (i) αs(Gi)=θs(Gi), (ii) αs(Li)=θs(Li), and (iii) αs(Hi)=θs(Hi), for all i, i = 1, 2. Therefore, there exists two non empty subgraphs G1 and G2 of Gv such that G1 and G2 have just the vertex v in common. We make two cases and in each case we arrive at a contradiction.

Case 1: There exists some αs -independent set of G, say I containing v.

Let Is1=V(G1)(I{v}) and Is2=V(G2)(I{v}). Note that Is1 and Is2 are αs -independent sets of G1 and G2, respectively. Moreover they are star-independent sets of G. Thus, we deduce αs(G) as follows. αs(G)=|I|=|Is1|+|Is2|+1=αs(G1)+αs(G2)+1=θs(G1)+θs(G2)+1,since G1 and G2 arestar‐perfect graphs as remarked above=θs(Gv)+1θs(G),since θs(Gv)‐set together with the star centered at v covers all the vertices of G

Therefore, αs(G)θs(G). Hence, αs(G)=θs(G), a contradiction to the property that G is minimal star-imperfect graph.

Case 2: No αs -independent set of G contains v.

At the outset we observe that dG(x,y)=dLi(x,y), for every x,yV(Li),i=1,2. Then, every star-independent set in Li(i=1,2) is a star-independent set in G.

Since L1 is star-perfect and v does not belong to any αs -independent set of G, by Lemma 2.4, there exists uNL1(v) such that u is in an αs -independent set of L1, say I. Then I{u} is an αs -independent set of H1. Let T1=I{u}. Let T2 be an αs -independent set of H2. Let T3={u}. Then one can easily verify that T1T2T3 is a star-independent set in G. Hence αs(G)|T1|+|T2|+1. (1)

Next let Si be θs -cover of Hi,i=1,2 and let S3 be the star [N[v]] in G. It follows that S1S2S3 is a star-cover of G. Hence θs(G)|S1|+|S2|+1. (2)

Hence we derive, θs(G)|S1|+|S2|+1,by (2)=|T1|+|T2|+1,since |S1|=|T1| and |S2|=|T2| because H1 and H2 are star‐perfect graphs αs(G), by (1)

Therefore, αs(G)θs(G). Hence, αs(G)=θs(G), a contradiction to the property that G is minimal star-imperfect graph.

Note. The parameters α(G) and θ(G) of Berge perfect graphs satisfy monotone property, that is if H is an induced subgraph of G, then α(H)α(G) and θ(H)θ(G). However, the parameters αs(G) and θs(G) do not satisfy monotone property. That is, if H is an induced subgraph of G, αs(H) (or θs(H)) may exceed αs(G) (or θs(G)). For example if G=K1,n,n2 with central vertex v0, then αs(K1,n)=1 and θs(K1,n)=1. However αs(K1,nv0)=n and θs(K1,nv0)=n,n2.

So the property that the parameters αs and θs do not satisfy monotone property causes difficulty in some of the results related to their equality.

Corollary 2.1.

Every tree is a star-perfect graph.

Proof.

On the contrary, suppose there exists a star-imperfect tree T. By (1.4), T contains a minimal star-imperfect tree T. By Lemma 2.6, T is a block and so it is either K1 or K2. However, K1 and K2 are star-perfect graphs, a contradiction to our supposition. □

The following two lemmas are easily verifiable.

Lemma 2.7.

If G is triangle free and if G has an odd cycle of length at least 5, then G has an induced cycle of length at least 5. □

Lemma 2.8.

If G is triangle free and if G has a cycle of length 8 with at least one chord, then G has an induced cycle of length 4 or 5. □

Before moving on to the next lemma it is convenient to introduce and fix a notation.

In what follows, let G be a minimal star-imperfect graph. So for every vertex v, Gv is a star-perfect graph. Let v0 be a fixed vertex in G and let θs(Gv0)=αs(Gv0)=l, for some positive integer l. Let S1,S2,,Sl be a θs -cover of Gv0, where Si is a star centered at vi , 1il. Thus, L={S0,S1,,Sl} is a star-cover of G where Si is a star centered at vi , 0il. Therefore, θs(G)l+1. Let vi1,vi2, denote the vertices adjacent to vi , that is {vi}{vi1,vi2,}=V(Si). Let M be the set of stars in L that are adjacent to v0 and N be the set of stars in L not adjacent to v0. Let v0 be adjacent to vi1, for all stars SiM with central vertex vi .

Lemma 2.9.

The cardinality of M is at least 2.

Proof.

Contrary to the lemma, let |M|<2. Since G is minimal star imperfect, it is a block and so |M|0. Next possibility is |M|=1. Let v0 be any vertex in G and v0 is adjacent to v11. If v12 exists then v0v12, otherwise we have an induced C4, a contradiction. Hence, v11 is a cut vertex in G and Gv11 has a component which is [{v0}]=K1, a contradiction to Lemma 2.5. Hence |M|2. □

Lemma 2.10.

If the stars in M are mutually vertex disjoint, then each star in M has at least two vertices.

Proof.

If the lemma is not true, then for some star, say S1=K1M. Let S1=[{v1}] and by Lemma 2.9, there is one more star, say S2 in M.

S1S2. If otherwise, we have induced C3 or C4 or C5, contradiction to G being minimal star-imperfect. If S1SkN, then we observe that SkSj,j1 and j < k.

If v1vk, then vkv2 or v21, otherwise we have induced C4 or C5, a contradiction. Then the only adjacency possible is vkv22. Also no other vertex of Sk is adjacent to S2, otherwise we have induced C5 or C7, a contradiction. This implies Gvk has two components one containing S2 and the other containing Sk , contradicting that G is a block.

If v1vk1 or v1vk2, we similarly arrive at a contradiction to G being a block. Thus in all the cases we arrive at a contradiction, and hence the lemma is true. □

Note. In view of this lemma, v0 is not adjacent to any of the central vertices of the stars in M.

Lemma 2.11.

There is at most one edge between a pair of stars in M.

Proof.

Let S1={v1,v11,v12} and S2={v2,v21,v22}. Since G is a minimal star-imperfect graph, G is (C3,C3k+1,C3k+2)-free.

Using Lemma 2.10 we show that v0vi,1il. If not, let v0vk, for some i = k. Then by Lemma 2.10 vk1 exists and vk1 is adjacent to v0. Then v0vkvk1v0 is a cycle of length 3, a contradiction. Thus v0 is not adjacent to any central vertices.

Let H be the subgraph of G induced by the vertices {v0,v1,v11,v12,v2,v21,v22}. The only possible edge between S1 and S2 is v1v22 or v2v11 for otherwise, there exists induced C3,C4,C5,C7, a contradiction to G being minimal star-imperfect graph. If both v1v22 and v2v11 exist, then we have an induced C4, a contradiction.

Hence, there exists at most one edge between two stars in M and hence the lemma is proved. □

Note. From now, we assume every star in M has at least 2 vertices and the stars in M are pairwise disjoint. If S1 has two vertices, and S2 has three vertices and if V(S1)V(S2)ϕ, the vertices in V(S1)V(S2) could be retained in S2. By similar arguments V(S1)V(S3), V(S2)V(S3) is empty. Continuing the arguments we see that any two stars in M are disjoint and the stars could be arranged in increasing order with respect to their number of vertices.

Lemma 2.12.

The set N is non-empty.

Proof.

Contrary to the lemma, suppose N is empty. Since G is minimal star-imperfect, G is a block. If there is no edge between any two of the stars in M, then v0 is a cut vertex of G, contradicting the fact G is a block. There exist two stars Si,SjM such that there is at most one edge between them by Lemma 2.11. If both Si and Sj have exactly two vertices, there is no edge between Si and Sj , otherwise we have induced C3, C4 or C5, a contradiction. So one of them has at least two vertices and the other has at least three vertices. Assuming Si has only two vertices, as observed in the proof of Lemma 2.11, vivj2 is the only edge between Si and Sj (See ). We observe that if Sk is a star in M, i,jk, then there is no edge between Sk and Si or Sk and Sj . If there is an edge between Sj and Sk , say vkvj2, then we observe that there is no edge between Si and Sj except vivj2, otherwise we have induced C3, C4 or C5, a contradiction. Let H=[V(Si)V(Sj)V(Sk){v0}]. If the vertices vir,vjs,vkt, r2, s3, t2 exist, they are pendent in G, a contradiction. Then H is as in .

Fig. 2 H=[V(Si)V(Sj)V(Sk){v0}].

Fig. 2 H=[V(Si)∪V(Sj)∪V(Sk)∪{v0}].

In H, the set {v0,vj2} is αs -independent set of H of order 2 and the stars centered at v0,vj2 is θs -cover of H of order 2. Therefore, αs(H)=θs(H). If Sk+1 is a star in M other than these three, we construct H1 with vertex set V(H) and V(Sk+1) and by similar arguments, we see that αs(H1)=θs(H1). Recursively we reach a stage where αs(G)=θs(G), contradiction to G being minimal star-imperfect graph. Hence our assumption is false and N is non-empty. □

Note. Although C5 is a minimal star-imperfect graph, Lemma 2.12 does not hold good for C5.

Lemma 2.13.

Let G be a minimal star-imperfect graph. Let H1 be the induced subgraph of G on {v0,v1,v11,v12,v2,v21,v22,v3,v31,v32} where {v1,v11,v12}S1M, {v2,v21,v22}S2M and {v3,v31,v32}S3N. Then H1 does not contain induced C6 or C9.

Proof.

If the lemma is not true, H1 possibly contains an induced C6 or C9.

Case 1: v0 is adjacent to a vertex of the stars S1 and S2 and H1 contains a C6. (See for possible C6)

Fig. 3 v0v11v31v3v2v21v0 is a cycle of length 6.

Fig. 3 v0v11v31v3v2v21v0 is a cycle of length 6.

The set L1 of stars of H1 centered at v1,v2,v3,v0 is a star-cover of H1; so θs(H1)|L1|=4. The set T1={v12,v22,v32,v0} is a star-independent set in H1, since d(x,y)3 for all x,yT1 else H1 contains a forbidden cycle (some edges of the forbidden cycle are edges of the cycle shown in ). It follows that θs(H1)4αs(H1), that is θs(H1)αs(H1) and θs(H1)=αs(H1).

Let H2 be the induced subgraph of G on V(G)V(H1). If V(H2)=ϕ, then G = H1. So αs(G)=αs(H1)=θs(H1)=θs(G). Thus, αs(G)=θs(G), a contradiction to G being minimal star-imperfect. So let V(H2)ϕ and T2 be a star-independent set in G containing T1 and has maximum cardinality. Since V(H2)ϕ, it follows that H2 is star-perfect.

Claim 1: If KH1, then T1={v12,v22,v32,v0} is also a star independent set in K.

For any two vertices x,yT1, we know dH1(x,y)3. We set to prove that dK(x,y)3. Contrary to this let dK(x,y)2. We know that dK(x,y)1, since H1 is an induced subgraph of K. If dK(x,y)=2, then say a pair of vertices in T1 is adjacent to a common vertex uV(K)V(H1). Then it is easy to verify that we obtain an induced forbidden cycle such as C5, C7 or C8, a contradiction. Hence, dK(x,y)1 or 2. This proves that T1 is star-independent set in K as well.

Claim 2: (T2T1)ϕ

If not, then T2T1=ϕ and T2 = T1. This implies that there exists no vertex vV(G)V(H1) such that dG(v,t)3, for any tT1. Therefore, KH1, we have αs(K)=4, or else we have contradiction to the maximality of T1. (1)

Since V(H2)ϕ, let H be a maximum induced subgraph of G containing H1 such that θs(H)=θs(H1). (2)

If H=G, then θs(G)=θs(H)=θs(H1),by (2)=αs(H1),since H1 isstar‐perfect=αs(H),by (1)=αs(G), by Claim 1

Therefore, θs(G)=αs(G), a contradiction to G being minimal star-imperfect graph.

So let HG, then there exists a vertex u in GH such that θs(H{u})θs(H1). Using θs(H{u})θs(H1) and (1) we have, αs(H{u})=4=αs(H1)=θs(H1)θs(H{u}). Hence for the proper induced subgraph H{u} of G, θs(H{u})αs(H{u}), a contradiction to G being minimal star-imperfect graph. Therefore, T2T1=ϕ is not possible and hence our claim.

The sets T1, T2 are star-independent sets in G. Let L2 be a set of stars in H2 such that each of the star contains a vertex of T2T1 (such an L2 exists because stars centered at each vertex of T2T1 is such an L2). Note that L2 need not be a star-cover of H2. Let H3 be an induced subgraph on V(H1)SL2V(S). Let t be a vertex in H=[V(G)V(H3)]. If t is star-independent with T2 then we have a contradiction to maximality of T2. By similar arguments as in claim 2, it implies that θs(H3)αs(H3) where H3 is the induced subgraph on V(H3){t}, a contradiction to G being minimal star-imperfect graph.

Hence our assumption that H1 contains an induced C6 is not true and hence the lemma is proved.

Case 2: v0 is adjacent to a vertex of the stars S1 and S2 and H1 contains a C9. (See for possible C9)

Fig. 4 v0v11v1v12v32v3v22v2v21v0 is a cycle of length 9.

Fig. 4 v0v11v1v12v32v3v22v2v21v0 is a cycle of length 9.

The set L1 of stars of H1 centered at v11,v32,v3 is a star-cover of H1; so θs(H1)|L1|=3. The set T1={v12,v22,v0} is a star-independent set in H1, since d(x,y)3 for all x,yT1 else H1 contains a forbidden cycle (some edges of the forbidden cycle are edges of the cycle shown in ). It follows that θs(H1)3αs(H1), that is θs(H1)αs(H1) and θs(H1)=αs(H1).

By repeating the same arguments as in Case 1, we see that H1 contains an induced C9 is not true and hence the lemma.

Lemma 2.14.

If G is minimal star-imperfect, then θs(G)=αs(G)+1.

Proof.

Using Lemma 2.3, we have that G is (C3,C3k+1,C3k+2)-free, k1.

Let H1 be the induced subgraph on SiMV(Si){v0} and H2 be the induced subgraph on SjNV(Sj). By Lemma 2.9 we already know that |M|2 and H2 is not a null graph. Let T1 and T2 be αs -independent sets of H1 and H2, respectively. Let T2 be a maximum subset (with respect to cardinality) of T2 such that T1T2 is a star-independent set of G. We consider the following two cases.

Case 1: T2=T2

Since G is minimal star-imperfect graph, θs(Gv0)=αs(Gv0). Therefore, there exists a θs -cover, say F of Gv0 such that every star in F contains exactly one vertex of αs -independent set, say I of Gv0.

Moreover, I is a star-independent set of G, since H1=SiMV(Si){v0}. αs(G)|I|, by above remark=αs(Gv0),since I is an αs‐independentset of Gv0=l, by notation=θs(G)1,by notation

So we have, αs(G)θs(G)1, implying either αs(G)=θs(G)1 or αs(G)θs(G). In the former case, αs(G)=θs(G)1, that is, θs(G)=αs(G)+1, which we set out to prove and in the later case, αs(G)θs(G), a contradiction since G is a minimal star-imperfect graph ().

Fig. 5 H1=[SiMV(Si){v0}] and H2=[SjNV(Sj)].

Fig. 5 H1=[∪Si∈MV(Si)∪{v0}] and H2=[∪Sj∈NV(Sj)].

Case 2: T2T2

This implies T2T2. Then there exists uT2T2.

If u is star-independent with T1{v32, we have a contradiction to choice of T2. So, let u be not star-independent with T1{v32}. Then, d(u,v12)2 where u=v42.

When d(u,v12)=1, u is not adjacent to any vertex in the set {v1,v11,v12,v2,v22}, or else induced forbidden cycle.

When d(u,v12)=2, u is not adjacent to any vertex in the set {v1,v11,v12,v2,v21}, or else induced forbidden cycle.

For cases d(u,v12)=1 and uv21, we obtain C6 and d(u,v12=2 and uv22, we obtain C9. Using Lemma 2.13 we have a contradiction.

If v41 and v4 are not star-independent to T1{v32} then no vertex of S4 is star-independent with T1{v32}. This implies that θs(H3)αs(H3) where H3 is the induced subgraph on V(H1)V(S3)V(S4). This implies that v4 or V41 is star-independent with T1{v32}. This contradicts choice of T2. Hence case 2 does not arise at all.

This completes the proof of the lemma. □

Lemma 2.15.

Let G be a minimal star-imperfect graph. Let H1 be the induced subgraph of G on {v0,vi,vi1,vi2,vj,vj1,vj2,vk,vk1,vk2} where {vi,vi1,vi2}SiM, {vj,vj1,vj2}SjM and {vk,vk1,vk2}SkN. If Sk is adjacent to Sj , then Sk is not adjacent to any other star in M.

Proof.

If no star in M is adjacent to any star in N, then v0 is a cut vertex in G, a contradiction by Lemma 2.6. So let SjM be adjacent to SkN. The following adjacencies are possible between Sj and Sk .

If Sk is not adjacent to any other stars in M, then we are done. If not, let Sk be also adjacent to SiM.

We look into each case mentioned in in view of Sk adjacent to Si as well.

Table 1 Adjacencies between SjM and SkN.

Since G is minimal star-imperfect, G is (C3,C3k+1,C3k+2)-free, k1. In each of the following cases other than those mentioned in each of them, we obtain induced C4,C5,C7,C8 or C10, a contradiction.

Case I:vjvk ()

Fig. 6 vjvk.

Fig. 6 vj↔vk.

If Sk is adjacent to SiM, the following adjacencies are possible.

If vkp(p1)vi1 (), then vkpvkvjvj1v0vi1vkp is an induced cycle of length 6, a contradiction to Lemma 2.13.

Fig. 7 vkpvkvjvj1v0vi1vkp is an induced C6.

Fig. 7 vkpvkvjvj1v0vi1vkp is an induced C6.

Case II:vjvk1 ()

Fig. 8 vjvk1.

Fig. 8 vj↔vk1.

If Sk is adjacent to SiM, the following adjacencies are possible:

If vk1vi, then vk1vjvj1v0vi1vivk1 forms C6, a contradiction to Lemma 2.13.

If vk1viq(q2), then vk1vjvj1v0vi1viqvk1 is an induced cycle of length 6, a contradiction to Lemma 2.13.

If vkp(p2)viq(q2), then vkpvkvk1vjvj1v0vi1viviqvkp is an induced C9, a contradiction to Lemma 2.13.

If vkvi1, then vkvk1vjvj1v0vi1vk is an induced cycle of length 6, a contradiction to Lemma 2.13.

Case III:vj1vk ()

Fig. 9 vj1vk.

Fig. 9 vj1↔vk.

If Sk is adjacent to SiM, the following adjacencies are possible: If vk1vi, then vk1vkvj1v0vi1vivk1 is an induced cycle of length 6, a contradiction to Lemma 2.13.

If vkviq(q2), then vkvj1v0vi1viviqvk is an induced cycle of length 6, a contradiction to Lemma 2.13.

Case IV:vj1vk1 ()

Fig. 10 vj1vk1.

Fig. 10 vj1↔vk1.

If Sk is adjacent to SiM, the following adjacencies are possible:

If vkvi, then vkvk1vj1v0vi1vivk is an induced cycle of length 6, a contradiction to Lemma 2.13.

If vk1viq(q2), then vk1vj1v0vi1viviqvk1 is an induced cycle of length 6, a contradiction to Lemma 2.13.

If vkp(p2)vi1, then vkpvkvk1vj1v0vi1vkp is an induced cycle of length 6, a contradiction to Lemma 2.13.

Case V:vjpvk, for some p2 ()

Fig. 11 vjpvk, for some p2.

Fig. 11 vjp↔vk, for some p≥2.

If Sk is adjacent to SiM, the following adjacencies are possible:

If vkvi1, then vkvi1v0vj1vjvjpvk is an induced cycle of length 6, a contradiction to Lemma 2.13.

If vkp1(p11)viq(q2), then vkp1viqvivi1v0vj1vjvjpvkvkp1 is an induced cycle of length 9, a contradiction to Lemma 2.13.

Case VI:vjpvk1, for some p2 ()

Fig. 12 vjpvk1, for some p2.

Fig. 12 vjp↔vk1, for some p≥2.

If Sk is adjacent to SiM, the following adjacencies are possible:

If vkviq(q2), then vkviqvivi1v0vj1vjvjpvk1vk is an induced cycle of length 9, a contradiction to Lemma 2.13.

If vk1vi1, then vk1vivi1v0vj1vjvj3vkvk1 is an induced cycle of length 6, a contradiction to Lemma 2.13.

Thus we find that none of these six cases arise at all and hence our lemma. □

Theorem 2.1

(Main Theorem). A graph G is star-perfect if and only if G is (C3,C3k+1,C3k+2)-free, k1.

Proof.

If G is star-perfect, then by Lemma 2.3, G is (C3,C3k+1,C3k+2)-free, k1.

Conversely, suppose G is (C3,C3k+1,C3k+2)-free, k1. If G is not star-perfect, assume G is a minimal star-imperfect graph.

Let v0 be a vertex in G such that deg(v0)=Δ(G). We assume Δ(G)3. Since G is minimal star-imperfect, Gv0 is star-perfect and αs(Gv0)=θs(Gv0). Let S1,S2,,Sl be a θs -cover of Gv0. Then S0={v0},S1,S2,,Sl will be a θs -cover of G and θs(G)=l+1 and αs(G)=l, by Lemma 2.14. Then obviously v0 is not adjacent to a central vertex of S1,S2,,Sl.

Let Si,SjM and SkN be three stars such that i1,j2,i<j<k. Without loss of generality, let SkSj. By Lemma 2.15, if SkN is adjacent to SjM, then Sk is not adjacent to any other star Sr in M where rj. (1)

Since deg(v0)=Δ(G)3, there exists S1,S2,SjM. Further SkSj and by Lemma 2.11, there is at most one edge between them. If S1 and S2 are adjacent, then only possible edge is v1v22 or v2v12. Then Sj is not adjacent to S1 and S2, otherwise we obtain an induced induced C5, C7 or C8. This implies that Gv0 has two different components one containing S1 and S2 and the other containing Sk , by (1). This contradicts that G is a block.

Next if S1 and S2 are not adjacent, and if S1 is adjacent to Sj then S2 is not adjacent to Sj , otherwise we obtain an induced C5, C7 or C8. This implies that Gv0 has two different components one containing S2 and the other containing Sk , by (1), a contradiction to G being a block. Therefore, our assumption that deg(v0)=Δ(G)3 is not true.

This implies that deg(v0)=Δ(G)2. If deg(v0)=Δ(G)=1, then G is K2, a contradiction to G being minimal star-imperfect graph. Therefore, only possibility is Δ(G)=2 and G is C3k,k2 which is star-perfect, by Lemma 2.2, a contradiction to G being minimal star-imperfect. This completes the proof of the main theorem. □

Corollary 2.2.

A bipartite graph G is star-perfect if and only if every induced cycle of G is C6k,k1.

Proof.

From the above main theorem, a bipartite graph G is star-perfect if and only if every induced cycle of G is of the form C3k,k2. It is well know that a graph is bipartite if and only if every induced cycle is of the form C2k,k0. Combining these two statements we get that a bipartite graph G is star-perfect if and only if every induced cycle of G is C6k,k1. □

Immediately we get the following corollary.

Corollary 2.3.

A perfect graph G is star-perfect if and only if G is bipartite graph in which all induced cycles of G are of the form C6k,k1.

3 Future directions

All the stars in a star-cover of G studied in this paper are induced. We can extend this study to θS(G) which is the minimum number of stars (not necessarily induced) contained in G such that the union of their vertex sets is V(G) and let αS(G) denote the maximum number of vertices in G such that no two of them are contained in the same star (not necessarily induced) of G. We call a graph G S-perfect if αS(H)=θS(H), for every induced subgraph H of G. It will be an interesting problem to characterize S-perfect graphs.

Acknowledgments

We thank profusely Professor S. A. Choudum, Christ (Deemed to be University), for all fruitful discussions and critical comments on the entire paper.

References

  • Ananchuen, N., Ananchuen, W., Plummer, M. D. (2011). Domination in graphs. In: Dehmer, M., ed. Structural Analysis of Complex Networks. Boston: Birkhäuser, 73–104.
  • Berge, C. (1960). Les problèmes de coloration en théorie des graphes. Publ. Inst. Statist. Univ. Paris 9: 123–160.
  • Bondy, J. A., Murty, U. S. R. (1976). Graph Theory with Applications. London: Macmillan.
  • Chen, G., Gyárfás, A., Schelp, R. H. (1998). Vertex colorings with a distance restriction. Discrete Math. 191(1–3): 65–82.
  • Jensen, T. R., Toft, B. (1995) Graph Coloring Problems. New York, NY: Wiley.
  • Korobitsin, D. V. (1992). On the complexity of domination number determination in monogenic classes of graphs. Discret. Math. Appl. 2(2): 191–200.
  • Lehel, J., Tuza, Z. (1986). Neighborhood perfect graphs. Discret. Math. 61(1): 93–101.
  • Litjens, B., Polak, S., Sivaraman, V. (2019). Sum-perfect graphs. Discret. Appl. Math. 259: 232–239.
  • Markossyan, S. E., Gasparyan, G. S., Reed, B. A. (1996). β-perfect graphs. J. Combin. Theory, Ser. B 67(1): 1–11.
  • Ravindra, G. (2011) F-perfect graphs. In: Second India-Taiwan Conference on Discrete Mathematics. Available at: http://www.adma.co.in/news/openproblems/SecondIndiaTaiwanConferenceOpenProblemsSession.pdf
  • West, D. B. (2017). Introduction to Graph Theory. Upper Saddle River, NJ: Pearson.