Extremal spectral results of planar graphs without vertex-disjoint cycles111Lin was supported by NSFC grant 12271162 and Natural Science Foundation of Shanghai (No. 22ZR1416300). Shi was supported by NSFC grant 12161141006.
Abstract
Given a planar graph family , let and be the maximum size and maximum spectral radius over all -vertex -free planar graphs, respectively. Let be the disjoint union of copies of -cycles, and be the family of vertex-disjoint cycles without length restriction. Tait and Tobin [Three conjectures in extremal spectral graph theory, J. Combin. Theory Ser. B 126 (2017) 137β161] determined that is the extremal spectral graph among all planar graphs with sufficiently large order , which implies the extremal graphs of both and for are . In this paper, we first determine and and characterize the unique extremal graph for , and sufficiently large . Secondly, we obtain the exact values of and , which solve a conjecture of Li [Planar TurΓ‘n number of the disjoint union of cycles, Discrete Appl. Math. 342 (2024) 260β274] for .
Keywords: Spectral radius; TurΓ‘n number; Planar graph; Vertex-disjoint cycles; Quadrilateral
AMS Classification: 05C35; 05C50
1 Introduction
Given a graph family , a graph is said to be -free if it does not contain any as a subgraph. When , we write -free instead of -free. One of the earliest results in extremal graph theory is the TurΓ‘nβs theorem, which gives the maximum number of edges in an -vertex -free graph. The TurΓ‘n number is the maximum number of edges in an -free graph on vertices. Much attention has been given to TurΓ‘n numbers on cycles. FΓΌredi and Gunderson [12] determined for all and . However, the exact value of is still open. ErdΕs [8] determined for , and the unique extremal graph is characterized. Subsequently, Moon [21] showed that ErdΕsβs result is still valid whenever . ErdΕs and PΓ³sa [9] showed that for and . For more results on TurΓ‘n-type problem, we refer the readers to the survey paper [13]. In this paper, we initiate the study of spectral extremal values and TurΓ‘n numbers on cycles in planar graphs.
1.1 Planar spectral extremal values on cycles
One extension of the classical TurΓ‘n number is to study extremal spectral radius in a planar graph with a forbidden structure. The planar spectral extremal value of a given graph family , denoted by , is the maximum spectral radius over all -vertex -free planar graphs. An -free planar graph on vertices with maximum spectral radius is called an extremal graph to . Boots and Royle [2] and independently Cao and Vince [3] conjectured that is the unique planar graph with the maximum spectral radius where β+β means the join product. The conjecture was finally proved by Tait and Tobin [27] for sufficiently large .
In order to study the spectral extremal problems on planar graphs, we first give a useful theorem which will be frequently used in the following.
Theorem 1.1.
Let be a planar graph and . If is a subgraph of but not of , then the extremal graph to contains a copy of .
Let be the disjoint union of copies of -cycles, and be the family of vertex-disjoint cycles without length restriction. For , it is easy to check that is -free and -free. Theorem 1.1 implies that is the extremal graph to and for and sufficiently large . For three positive integers with and , we can find integers and such that and . Let
In the paper, we give answers to for as follows.
Theorem 1.2.
For integers and , the graph is the extremal graph to .
Theorem 1.4 implies that is -free for . By Theorem 1.2, is the extremal graph to for . Moreover, a planar graph is -free when it is -free. Hence, one can easily get answer to .
Corollary 1.1.
For , is the extremal graph to .
We use to denote the graph obtained from by embedding a maximum matching within its independent set. Nikiforov [23] and Zhai and Wang [28] showed that is the extremal graph to for odd and even , respectively. Clearly, is planar. This implies that is the extremal graph to . Nikiforov [22] and CioabΔ, Desai, and Tait [5] determined the spectral extremal graph among -free graphs for odd and even , respectively. Next we give the characterization of the spectral extremal graphs among -free planar graphs.
Theorem 1.3.
For integers and ,
(i) is the unique extremal graph to for ;
(ii) is the unique extremal graph to for .
1.2 Planar TurΓ‘n numbers on cycles
We also study another extension of the classical TurΓ‘n number, i.e., the planar TurΓ‘n number. Dowden [6] initiated the following problem: what is the maximum number of edges in an -vertex -free planar graph? This extremal number is called planar TurΓ‘n number of and denoted by . The planar TurΓ‘n number for short cycles are studied in [4, 6, 7, 14, 15, 16, 25, 26], but is still open for general . For more results on planar TurΓ‘n-type problem, we refer the readers to a survey of Lan, Shi and Song [18]. It is easy to see that for . Lan, Shi and Song [17] showed that for , and the double wheel is an extremal graph. We prove the case of , which will be used to prove our main theorems.
Theorem 1.4.
For , we have . The extremal graphs are obtained from and an independent set of size by joining each vertex of the independent set to any two vertices of the triangle.
Moreover, Lan, Shi and Song [17] also proved that for all . They [19] further showed that and obtained lower bounds of for , which was improved by Li [20] for sufficiently large recently. Li [20] also raised the following conjecture.
Conjecture 1.1.
If , then and the bound is tight for .
In this paper, we determine the exact value of for large , which solve the conjecture for .
Theorem 1.5.
For ,
2 Proof of Theorem 1.1
Let be the adjacency matrix of a planar graph , and be its spectral radius, i.e., the maximum modulus of eigenvalues of . Throughout Sections 2 and 3, let be an extremal graph to , and denote this spectral radius. By Perron-Frobenius theorem, there exists a positive eigenvector corresponding to . Choose with . For a vertex and a positive integer , let denote the set of vertices at distance from in . For two disjoint subset , denote by the bipartite subgraph of with vertex set that consist of all edges with one endpoint in and the other endpoint in . Set and . Since is a planar graph, we have
(1) |
Now, we are ready to give the proof of Theorem 1.1.
Proof.
We give the proof in a sequence of claims. Firstly, we give the lower bound of .
Claim 2.1.
For , we have
Proof.
Note that is planar and -free. Then, , as is an extremal graph to . β
Claim 2.2.
Set for some constant . Then .
Proof.
By Claim 2.1, . Hence,
for each . Summing this inequality over all vertices , we obtain
It follows that as . β
Claim 2.3.
We have .
Proof.
Let be an arbitrary vertex. For convenience, we use , and instead of , and , respectively. By Claim 2.1, . Then
(2) |
Note that . We can calculate according to two cases or . We first consider the case . Clearly, and for . We can see that
(3) |
By Claim 2.2, we have . Moreover, . Then, by (1), we have
(4) |
Next, we consider the remain case . Clearly, for . Then
(5) |
where .
Now we prove that for each . Suppose to the contrary that there exists a vertex with . Note that as . Setting , and combining (6), we have
(7) |
By (1), we have
where and by Claim 2.2. Combining this with gives
which contradicts (7). Therefore, for each . Summing this inequality over all vertices , we obtain
which yields that . β
For convenience, we use , and instead of , and , respectively. Now we give a lower bound for degrees of vertices in .
Claim 2.4.
For every , we have .
Proof.
Let be the subset of in which each vertex has at least neighbors in . We first prove that . If , then is empty, as desired. It remains the case . Suppose to the contrary that . Since there are only options for vertices in to choose a set of neighbors from , we can find a set of vertices in with at least common neighbors in . Moreover, one can observe that and . Hence, contains a copy of , contradicting that is planar. Hence, . It follows that
where the second last inequality holds as and , and the last inequality holds as . Setting and combining the above inequality with (6), we have
which yields that . β
Claim 2.5.
There exists a vertex such that .
Proof.
Notice that and . By Claim 2.4, we have
(8) |
Now, let , , and . Thus, . Now, we prove that the eigenvector entries of vertices in are small.
Claim 2.6.
Let . Then .
Proof.

Claim 2.7.
is empty.
Proof.
Suppose to the contrary that . Assume that is a planar embedding of , and are around in clockwise order in , with subscripts interpreted modulo . Recall that . Then . By the pigeonhole principle, there exists an integer such that is a face of the plane graph . We modify the graph by joining each vertex in to each vertex in and making these edges cross the face , to obtain the graph (see in Figure 1). Then, is a plane graph.
We first show that is -free. Suppose to the contrary, then contains a subgraph isomorphic to . From the modification, we can see that is not empty. Moreover, since , we have
Then, we may assume that and . Clearly, for each . This indicates that a copy of is already present in , a contradiction. Therefore, is -free.
In what follows, we shall obtain a contradiction by showing that . Clearly, is planar. Then, there exists a vertex with . Define . Repeat this step, we obtain a sequence of sets with for each . Combining this with (9), we have
(10) |
where the last inequality holds as for any by Claim 2.6. It is not hard to verify that in graph the set of edges incident to vertices in is . Note that . Combining these with (10), we have
which contradicts that is extremal to . Therefore, is empty. β
3 Proofs of Theorems 1.2 and 1.3
Recall that is an extremal graph to . In the case , by Theorem 1.1, contains a copy of . We further obtain that as is triangle-free (otherwise, adding any edge increases triangles, a contradiction). Now we consider the case . Let denote a path on vertices. For convenience, an isolated vertex is referred to as a path of order 1. We first give two lemmas to characterize the structural features of the extremal graph , which help us to present an approach to prove Theorems 1.2 and 1.3.
Lemma 3.1.
Let and . Then and is a disjoint union of paths.
Proof.
Since and , we can see that is a subgraph of but not of . Hence, by Theorem 1.1, contains a copy of , and are the vertices of degree and is the set of vertices of degree two in .
Claim 3.1.
.
Proof.
Suppose to the contrary that . Assume that is a planar embedding of , and are around in clockwise order in , with subscripts interpreted modulo . If induces an cycle , then we can easily check that contains a copy of , a contradiction. Thus, there exists an integer such that . Furthermore, is a face in .
We modify the graph by adding the edge and making cross the face , to obtain the graph . Clearly, is a plane graph. We shall show that is -free. Suppose to the contrary that contains a subgraph isomorphic to . If for some , then contains an -cycle containing , say . However, an -cycle is already present in , a contradiction. If for some , then contains two vertex-disjoint -cycles and . From the modification, we can see that one of and (say ) contains the edge . This implies that is a subgraph of . However, contains a -minor, contradicting the fact that is planar. Hence, is -free. However, , contradicting that is extremal to . Therefore, . β
From Claim 3.1 we know that . It remains to show that is a disjoint union of paths. Theorem 1.1 and Claim 3.1 imply that and are dominating vertices. Furthermore, since is -minor-free and -minor-free, we can see that is -minor-free and -minor-free. This implies that is an acyclic graph with maximum degree at most 2. Thus, is a disjoint union of paths. The result follows. β
Now we give a transformation that we will use in subsequent proof.
Definition 3.1.
Let and be two integers with , and let , where is a disjoint union of paths. We say that is an -transformation of if
Clearly, is a disjoint union of paths, which implies that is planar. If , then we shall show that for sufficiently large .
Lemma 3.2.
Let and be defined as in Definition 3.1, and . If , then .
Proof.
By Lemma 3.1, and is a disjoint union of paths. If , then . Assume that and are two components of . If , then , and so . It follows that , the result holds. Next, we deal with the case . If , then let be obtained from by deleting the edge and adding the edge . Clearly, . Moreover,
Since is a positive eigenvector of , we have . If , then is also a positive eigenvector of , and so , contradicting that . Thus, , the result holds. The case is similar and hence omitted here.
It remains the case . We shall give characterizations of eigenvector entries of vertices in in the following two claims.
Claim 3.2.
For any vertex , we have .
Proof.
Claim 3.3.
Let be a positive integer. Set and . Then,
(i) for any , ;
(ii) for any , ;
(iii) for any ,
.
Proof.
(i) It suffices to prove that for any ,
We shall proceed the proof by induction on . Clearly,
(15) |
By Claim 3.2, we have
(18) |
So the result is true when . Assume then that , which implies that . For , , and so
(19) |
By the induction hypothesis, and . Setting and combining (19), we have , as desired. If , then by the induction hypothesis, and . Thus, by (19), we have , as desired. Hence the result holds.
The proof of (ii) is similar to that of (i) and hence omitted here.
(iii) It suffices to prove that for any and , We shall proceed the proof by induction on . Clearly,
Combining this with (15) and Claim 3.2 gives
So the claim is true when . Assume then that , which implies that . If , then , and so
(20) |
By the induction hypothesis, and . Combining these with (20), we have . Hence the result holds. β
Since , we have , and so
Combining this with Claim 3.3, we obtain
(21) |
and
(22) |
for any . Similarly,
(23) |
Recall that . Let be integers with and , and be obtained from by deleting edges and adding edges . Since , we can see that . Moreover,
(24) |
Now, we consider the following two cases to complete the proof.
Case 1. is odd.
Case 2. is even.
We only consider the subcase . The proof of the subcase is similar and hence omitted here. Set . Then, as . Moreover, , as . If , then is even. This implies that by symmetry, that is, . If , then by (21), . In both situations, we have . From (22) we have , which implies that . Furthermore, we have by (24). If , then is a positive eigenvector of , and so . On the other hand, , since is a positive eigenvector of . It follows that , which is a contradiction. Thus, , as desired.
This completes the proof of Lemma 3.2. β
From Lemma 3.1, we may assume that , where and .
Let be a disjoint union of paths.
We use to denote the order of the -th longest path of for any .
Having Lemmas 3.1 and 3.2,
we are ready to complete the proofs of Theorems 1.2 and 1.3.
Proof of Theorem 1.2. We first give the following claim.
Claim 3.4.
If is a disjoint union of paths, then is -free if and only if and .
Proof.
We first claim that is -free if and only if is -free. Equivalently, contains a copy of if and only if contains a copy of . Assume that contains two vertex-disjoint -cycles and , and . Since is acyclic, we can see that must contain at least one vertex of and for any . Without loss of generality, assume that and . Then, , and so contains a . We can further find that contains a . Conversely, assume that contains two vertex-disjoint paths and such that . Thus, the subgraph induced by contains a copy of . Similarly, the subgraph induced by contains a copy of . This indicates that contains a copy of . So, the claim holds.
Next, we claim that is -free if and only if and . If is -free, then (otherwise, contains a copy of , a contradiction); (otherwise, contains a copy of as , a contradiction). Conversely, if and , then it is not hard to verify that is -free.
This completes the proof of Claim 3.4. β
Recall that (resp. ) is the order of the -th longest path of (resp. ) for any . By Claim 3.4, and . By a direct computation, we have , and hence
(25) |
We first claim that . Suppose to the contrary that . Let be an -transformation of . Clearly, and . By Claim 3.4, is -free. However, by (25) and Lemma 3.2, we have , contradicting that is extremal to . Thus, , the claim holds.
Note that for each . We then claim that for . If , then , as , for each and the claim holds trivially. Now let . Suppose to the contrary, then set Let be an -transformation of . Clearly, and . By Claim 3.4, is -free. However, by (25) and Lemma 3.2, we have , contradicting that is extremal to . So, the claim holds.
Since , for and , we can see that .
This completes the proof of Theorem 1.2.
Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β
Proof of Theorem 1.3. It remains the case . We first give the following claim.
Claim 3.5.
If is a disjoint union of paths, then is -free if and only if .
Proof.
One can observe that the longest cycle in is of order . Furthermore, contains a cycle for each . Consequently, if and only if is -free. Hence, the claim holds. β
By Claim 3.5, , which implies that as . By a direct computation, we have , and hence
(26) |
Since , we have , which implies that and . Consequently, since , we have
(27) |
as and . On the other hand, , which implies that by (27).
We first claim that . Suppose to the contrary that . Let be an -transformation of . Clearly, and . Then, . By Claim 3.5, is -free. However, by (26) and Lemma 3.2, we have , contradicting that is extremal to . Hence, the claim holds.
Secondly, we claim that for . Suppose to the contrary, then set . Let be an -transformation of . Clearly, and , and so . By Claim 3.5, is -free. However, by (26) and Lemma 3.2, we have , contradicting that is extremal to . Hence, the claim holds. It follows that .
Finally, we claim that . Note that . Then, for . It remains the case . Suppose to the contrary, then as . We use to denote the -th longest path of for . Recall that . Then are paths of order . We may assume that . Since , there exists an endpoint of with . Let be obtained from by (i) deleting and joining to an endpoint of ; (ii) deleting all edges of and joining to an endpoint of for each . Then, is obtained from by deleting edges and adding edges. By Claim 3.2, we obtain
for any vertices . Then
where as . So, . On the other hand, . By Claim 3.5, is -free, contradicting that is extremal to . So, the claim holds.
Since and , we have .
Moreover, since for and ,
we can further obtain that ,
which implies that .
This completes the proof of Theorem 1.3.
Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β Β
4 Proof of Theorem 1.4
Above all, we shall introduce the Jordan Curve Theorem: any simple closed curve in the plane partitions the rest of the plane into two disjoint arcwise-connected open sets (see [1], P. 244). The corresponding two open sets are called the interior and the exterior of . We denote them by and , and their closures by and , respectively. A plane graph is a planar embedding of a planar graph. The Jordan Curve Theorem gives the following lemma.
Lemma 4.1.
Let be a cycle of a plane graph , and let be two vertices of with and , then .
Let be a plane graph. A face in of size is called an -face. Let denote the number of -faces in , and let denote .
Lemma 4.2.
(Proposition 2.5 of [1], P. 250) Let be a planar graph, and let be an arbitrary face in some planar embedding of . Then admits a planar embedding whose outer face has the same boundary as .
Let be the minimum degree of a graph . It is well known that every graph with contains a cycle. In the following, we give a more delicate characterization on planar graphs, which contains an important structural information of the extremal graphs in Theorem 1.4.
Lemma 4.3.
Let be a plane graph on vertices with . Then contains two vertex-disjoint cycles unless .
Proof.
We first deal with some trivial cases. Since , we have . If , then . If , then , and so . On the other hand, , since is planar. Thus, . It is not hard to verify that when and when =8, as desired. If is not connected, then contains at least two components and with for , which implies that each contains a cycle. Thus, contains two vertex-disjoint cycles, as desired. If has a cut vertex , then has at least two components and . Since , we have for , which implies that both and contain a cycle. Thus, also contains two vertex-disjoint cycles.
Next, we only need to consider the case that is a 2-connected graph of order . Since is 2-connected, each face of is a cycle. Let be a face of with minimum size . By Lemma 4.2, we may assume without loss of generality that is the outer face of . Let . If contains a cycle, then contains two vertex-disjoint cycles, as desired. Now assume that is acyclic. Since , we have . This, together with Eulerβs formula , gives . On the other hand,
Hence, , yielding . Subsequently, we shall give several claims.
Claim 4.1.
We have .

Proof.
Suppose to the contrary that , and let . We first consider the case that there exists a vertex of adjacent to two consecutive vertices of . Without loss of generality, let and induces a triangle . More generally, we define . Clearly, . We can select a vertex, say , in such that (see Figure 2). Notice that is not a face of , as . Then, . By Lemma 4.1, every vertex in has no neighbors in . Moreover, by the definitions of and , every vertex in has at most one neighbor in . It follows that every vertex in has at least one neighbor in , as . Thus, is nonempty, that is, is nonempty. Recall that is acyclic. Then contains at least two pendant vertices, one of which (say ) is not adjacent to . Hence, is also a pendant vertex of , as has no neighbors in . On the other hand, has at most one neighbor in , and so . Therefore, , contradicting .
Now it remains the case that each vertex of is not adjacent to two consecutive vertices of . Note that and is acyclic. Then contains a vertex with , and thus . Now, since , we may assume without loss of generality that . Let . Clearly, and for each . Now, we can select a vertex, say , in such that , where (see Figure 2). We can see that (otherwise, , a contradiction). By the definition of , we have . Furthermore, every vertex in has no neighbors in and has at most one neighbor in Thus, every vertex in has at least one neighbor in . By a similar argument as above, we can find a vertex with , which contradicts . β
By Claim 4.1, the outer face of is a triangle . In the following, we denote for . Since , we have for each isolated vertex of , and for each pendant vertex of .
Claim 4.2.
and .
Proof.
Since is the outer face of , every vertex of lies in . Furthermore, since is planar, it is easy to see that . This implies that contains at most one isolated vertex. Recall that and is acyclic. Then contains at least two pendant vertices and . Therefore, . β
Claim 4.3.
Let be two vertices in such that
and (see Figure 3).
Then
(i) , where ;
(ii) if ,
then contains a pendant vertex in .

Proof.
(i) Since is the outer face and , we have . Furthermore, using and Lemma 4.1 gives .
(ii) Since , we have , and so . By (i), we know that . If , then is a desired pendant vertex. It remains the case that . Now, whether is a neighbor of or not, has at least one neighbor in . Thus, is nonempty. Recall that is acyclic. Then contains at least two pendant vertices, one of which (say ) is not adjacent to . Hence, is also a pendant vertex of , as has no neighbors in . β

Claim 4.4.
Let with . Assume that and (see Figure 4). Then contains a cycle such that and for each .
Proof.
We first claim that . By Claim 4.3, we know that . Now, since , we have by Lemma 4.1, and so . Furthermore, we have . Then contains a path with endpoints and , where is a pendant vertex of . If , then by and Lemma 4.1, we have as . Now let be the subpath of with endpoints and . Then , and contains a cycle for each , as desired. Next, assume that . Then, . By and Lemma 4.1, we get that , and so . Moreover, and give . Thus, . Therefore, contains a cycle for each , as desired. β
Having above four claims, we are ready to give the final proof of Lemma 4.3. By Claim 4.2, we have and . We may without loss of generality that and . For each , let . Since , we have . Hence, is nonempty, and so contains at least two pendant vertices. According to the size of , we now distinguish two cases to complete the proof.
Case 1. .
Assume that . Then (see Figure 5). We then consider two subcases according to the size of .

Subcase 1.1. , that is, .
For each pendant vertex of , we have , consequently, . This indicates that contains exactly two pendant vertices and . Furthermore, we can see that contains no isolated vertices (otherwise, every isolated vertex of has at least three neighbors in and so belongs to , while the unique vertex is a pendant vertex of ). Therefore, is a path of order with endpoints and .
Now we know that is a path with . Let and . Then is a path with endpoints and . Since , we have . If , then contains two vertex-disjoint cycles and , as desired. If for some , then contains two vertex-disjoint cycles and , as desired.
Subcase 1.2. .
Let . If , then we may assume that by the symmetry of and , where . By Claim 4.4, contains a cycle such that and . On the other hand, Claim 4.3 implies that . Hence, . Therefore, contains two vertex-disjoint cycles and , as desired.
It remains the case that . Now for some . We define instead of the original one in Claim 4.4. Then . Moreover, as . By Claim 4.4, there exists a cycle such that and . Therefore, contains two vertex-disjoint cycles and , as desired.
Case 2. .
Recall that . Since , we can see that for each . We may assume without loss of generality that by the symmetry of vertices in . By Claim 4.3, there exists a pendant vertex of in , which implies that . Since , we have , and thus . Moreover, as . Assume without loss of generality that (see Figure 6). We also consider two subcases according to .

Subcase 2.1. , that is, .
Since , we have , which implies that is non-empty and has at least two pendant vertices. On the other hand, since while , we can see that contains no isolated vertices, and for each pendant vertex of . Therefore, contains exactly two pendant vertices and , more precisely, is a path with endpoints and . Let be an arbitrary vertex in . Then,
If , then contains two vertex-disjoint cycles and , where is the subpath of from to (see Figure 6()). If , then contains two vertex-disjoint cycles and , where is the subpath of from to (see Figure 6()). If for each then , as desired.
Subcase 2.2. .
For each vertex , it is clear that is one of , and . We first consider the case that there exist two vertices in which have the same neighbors in . Without loss of generality, assume that we can find a vertex with . Then . Recall that and (see Figure 6()). Then, we can further get that . By Claim 4.4, there exists a cycle such that and . On the other hand, Claim 4.3 implies that . Hence, . Therefore, contains two vertex-disjoint cycles and , as desired.
Now it remains the case that any two vertices in have different neighborhoods in . This implies that and we can find a vertex with . Now we have . Furthermore, since and , we have for each , and if , then . Since , we can see that has only one connected component, that is, is a tree and some , say , is a pendant vertex of . Now, contains a subpath with endpoints and (see Figure 6()). Then contains two vertex-disjoint cycles and , as desired.
This completes the proof of Lemma 4.3. β
Let be the family of graphs obtained from and an independent set of size by joining each vertex of the independent set to arbitrary two vertices of the triangle (see Figure 7). Clearly, every graph in is planar. Now, let be the family of planar graphs obtained from by iteratively adding vertices of degree 2 until the resulting graph has vertices. Then .

Lemma 4.4.
For any graph , is -free if and only if .
Proof.
Let be the set of vertices of degree 4 and be the set of vertices of degree 3 in , respectively. Then induces a triangle. We first show that every graph in is -free. It suffices to prove that every cycle of contains at least two vertices in . Let be an arbitrary cycle of . If , then there is nothing to prove. It remains the case that there exists a vertex . By the definition of , we can see that . Note that . Hence, contains at least two vertices in .
In the following, we will show that every graph contains two vertex-disjoint cycles. By the definition of , is obtained from by iteratively adding vertices of degree 2. Now, let , and for . Then . Moreover, and for each . Now let
Since and , we have . By the choice of , we know that and , which implies that and .

Now we may assume that is a planar embedding of , and is a plane subgraph of . Observe that has six planar embeddings (see Figure 8). Without loss of generality, assume that is the leftmost graph in Figure 8. Then, lies in one of the following regions (see Figure 8):
By Lemma 4.2, we can assume that lies in the outer face, that is, . For simplify, we denote . Let be an arbitrary vertex with . Then by Lemma 4.1 and , we have , and thus . This implies that , where . Recall that and . Then, . If , then we may assume without loss of generality that , and . Since , we have , and so . Thus, contains two vertex-disjoint cycles and , as desired. Now consider the case that . This implies that . Let . Then . Therefore, contains two vertex-disjoint cycles and . β
Given a graph , let be the largest induced subgraph of with minimal degree at least 3. It is easy to see that can be obtained from by iteratively removing the vertices of degree at most 2 until the resulting graph has minimum degree at least 3 or is empty. It is well known that is unique and does not depend on the order of vertex deletion (see [24]).
In the following, we give the proof of Theorem 1.4.
Proof.
Let and be an extremal graph corresponding to . Observe that is a planar graph which contains no two vertex-disjoint cycles (see Figure 7). Thus, .
If is empty, then we define as the graph obtained from by iteratively removing the vertices of degree at most 2 until the resulting graph has 4 vertices. By the planarity of , we have , and thus
a contradiction.
5 Proof of Theorem 1.5
We shall further introduce some notations on a plane graph . A vertex or an edge of is said to be incident with a face , if it lies on the boundary of . Clearly, every edge of is incident with at most two faces. A face of size is called an -face. The numbers of -faces and total faces are denoted by and , respectively. Let be the set of edges incident with at least one -face, and particularly, let be the set of edges incident with two -faces. Moreover, let and denote the cardinalities of and , respectively. We can easily see that
Lan, Shi and Song proved that , with equality when (see [16]), and for (see [17]). For , one can easily see that . In [10], the authors obtained the following sharp result.
Lemma 5.1.
([10]) Let be two integers with and . Then , with equality if .
To prove Theorem 1.5, we also need an edge-extremal result on outerplanar graphs. Let denote the maximum number of edges in an -vertex -free outerplanar graph.
Lemma 5.2.
([11]) Let be three integers with and . Then
In particular, we can obtain the following corollary.
Corollary 5.1.
Let . Then

For arbitrary integer , we can find a unique such that , and . Let be a -vertex outerplanar graph and be copies of (see Figure 9). Then, we define as the subgraph of induced by . One can check that and
We now construct a new graph from by identifying the edges and for each . Clearly, is a connected -free outerplanar graph with
Moreover, since we have
Combining Corollary 5.1, is an extremal graph corresponding to .
Lemma 5.3.
Let and be an extremal plane graph corresponding to . Then contains at least fourteen quadrilaterals and all of them share exactly one vertex.
Proof.
Note that and is a -free outerplanar graph of order . Then is an -vertex -free planar graph, and thus
On the other hand, by Lemma 5.1, we have
Note that for . Then contains a copy, say , of . Let be the graph obtained from by deleting all edges within . Since , we have . Thus, contains a copy, say , of . Now we can obtain a new graph from by deleting all edges within . Note that . Repeating above steps, we can obtain a graph sequence and fourteen copies of such that and is obtained from by deleting all edges within . This also implies that contains at least fourteen quadrilaterals. We next give four claims on those copies of .
Claim 5.1.
Let be two integers with and . Then, .
Proof.
Suppose to the contrary that there exists a vertex . Note that . By the definition of , whether or not, we can see that . On the other hand, note that , then , contradicting Hence, the claim holds. β
Claim 5.2.
for any two integers with .
Proof.
If and are vertex-disjoint, then contains , a contradiction. Now suppose that there exist three vertices . Observe that contains no an independent set of size . Then is nonempty. Assume without loss of generality that . Then , which contradicts Claim 5.1. Therefore, . β
Now for convenience, a vertex in a graph is called a -vertex if , and a -vertex if Clearly, every copy of contains two 2-vertices and three -vertices.
Claim 5.3.
Let be the family of graphs () such that every 2-vertex in is a -vertex in . Then .
Proof.
Note that contains only three -vertices, say and . Then every graph must contain two of and as -vertices. Suppose to the contrary that . By pigeonhole principle, there exist two graphs such that they contain the same two 2-vertices, say . It follows that contains a -cycle for . By Claim 5.2, we have , which implies that and are vertex-disjoint. Hence, contains two vertex-disjoint 4-cycles, a contradiction. β
Claim 5.4.
Let be an integer with and . Then, there exists a vertex such that and .
Proof.
By Claim 5.2, we have . We first assume that . If and , then there is nothing to prove. If , then contains two vertex-disjoint subgraphs and , and thus , a contradiction. If , then we can similarly get a contradiction. Therefore, .
Now, assume that . We first deal with the case . Since , one of , say , is a 2-vertex in . Hence, contains two vertex-disjoint subgraphs and , and so , a contradiction. Thus, there exists some with . If , then we are done. If , then we define as the subgraph of induced by . Since , we can check that always contains a . Moreover, since , we can see that also contains a . On the other hand, by Claim 5.1, we have , which implies that and are vertex-disjoint. Therefore, contains , a contradiction. β
By Claim 5.3, , thus there are at least ten graphs in . However, has only three -vertices. By Claim 5.4 and pigeonhole principle, there exists a -vertex in and four graphs, say , of . By Claim 5.1, we get that , and so , for any with . If contains a quadrilateral , then there exists some such that as . Since is a -vertex in , we can observe that the subgraph of induced by must contain a . Consequently, is not -free, a contradiction. Thus, is -free, which implies that all quadrilaterals of share exactly one vertex. This completes the proof of Lemma 5.3. β
Now we are ready to give the proof of Theorem 1.5.
Proof.
Recall that is an extremal graph corresponding to . Then is planar and -free. By Corollary 5.1, we have
(30) |
To prove Theorem 1.5, it suffices to show Since is an extremal plane graph corresponding to , we have . In the following, we show that .

By Lemma 5.3, all quadrilaterals of share a vertex . Thus, is -free. Assume that and are around in clockwise order, with subscripts interpreted modulo . Let be an arbitrary edge in , that is, is incident with two 3-faces, say and . We define as the plane subgraph induced by all edges incident with and . Clearly, and so it contains a . Recall that all quadrilaterals of share exactly one vertex . Then, and is incident with at least one face of and (see Figure 10). Note that is incident with . Then, either or for some . By the choice of , we have
(31) |
Assume first that and are 4-faces in . Since every 4-face is a quadrilateral, is incident with each 4-face. Consequently, there exists such that are incident with for each . Thus, for . On the other hand, if , then contains a , and so . This implies that is a 3-face in , contradicting the fact that are incident with the 4-face . Thus, we also have for .
By the argument above, we can see that
(32) |
Using (31) and (32) gives . Hence,
(33) |
On the other hand,
which yields Combining this with Eulerβs formula , we obtain
(34) |
If , then (33) and (34) hold directly. Combining (33) and (34), we have . Recall that . If , then by (30), as desired. If , then is a dominating vertex of the planar graph , which implies that is outerplanar. Recall that is -free. Thus, , and so . Combining (30), we get , as required. This completes the proof of Theorem 1.5. β
References
- [1] J.A. Bondy, U.S.R. Murty, Graph Theory, Springer-Verlag, Berlin, 2008.
- [2] B.N. Boots, G.F. Royle, A conjecture on the maximum value of the principal eigenvalue of a planar graph, Geogr. Anal. 23 (3) (1991), 276-282.
- [3] D.S. Cao, A. Vince, The spectral radius of a planar graph, Linear Algebra Appl. 187 (1993), 251β257.
- [4] D.W. Cranston, B. LidickΓ½, X.N. Liu, A. Shantanam, Planar TurΓ‘n numbers of cycles: a counterexample, Electron. J. Combin. 29 (3) (2022), 3.31.
- [5] S. CioabΔ, D.N. Desai, M. Tait, The spectral even cycle problem, arxiv:2205.00990v1 (2022).
- [6] C. Dowden, Extremal -free/-free planar graphs, J. Graph Theory 83 (2016), 213β230.
- [7] L.L. Du, B. Wang, M.Q. Zhai, Planar TurΓ‘n numbers on short cycles of consecutive lengths, Bull. Iranian Math. Soc. 48 (5) (2022), 2395β2405.
- [8] P. ErdΕs, Γber ein Extremal problem in der Graphentheorie, Arch. Math. 13 (1962), 222β227.
- [9] P. ErdΕs, L. PΓ³sa, On the maximal number of disjoint circuits of a graph, Publ. Math. Debrecen 9 (1962), 3β12.
- [10] L.F. Fang, B. Wang, M.Q. Zhai, Planar TurΓ‘n number of intersecting triangles, Discrete Math. 345 (5) (2022), 112794.
- [11] L.F. Fang, M.Q. Zhai, Outerplanar TurΓ‘n numbers of cycles and paths, Discrete Math. 346 (12) (2023), 113655.
- [12] Z. FΓΌredi, D.S. Gunderson, Extremal numbers for odd cycles, Combin. Probab. Comput. 24 (2015), 641β645.
- [13] Z. FΓΌredi, M. Simonovits, The history of degenerate (bipartite) extremal graph problems, Bolyai Soc. Studies (The ErdΕs Centennial) 25 (2013), 167β262.
- [14] D. Ghosh, E. GyΕri, R.R. Martin, A. Paulos, C.Q. Xiao, Planar TurΓ‘n number of the 6-cycle, SIAM J. Discrete Math. 36 (3) (2022), 2028β2050.
- [15] E. GyΕri, A. Li, R.T. Zhou, The planar TurΓ‘n number of the seven-cycle, arxiv:2307.06909v2.
- [16] Y.X. Lan, Y.T. Shi, Z-X. Song, Extremal Theta-free planar graphs, Discrete Math. 342 (12) (2019), 111610.
- [17] Y.X. Lan, Y.T. Shi, Z-X. Song, Extremal -free planar graphs, Electron. J. Combin. 26 (2019) no. 2, Paper No. 2.11, 17 pp.
- [18] Y.X. Lan, Y.T. Shi, Z.X. Song, Planar TurΓ‘n number and planar anti-Ramsey number of graphs, Oper. Res. Trans. 25 (2021), 200β216.
- [19] Y.X. Lan, Y.T. Shi, Z-X. Song, Planar TurΓ‘n number of cubic graphs and disjoint union of cycles, arxiv: 2202.09216v2 (2022).
- [20] P. Li, Planar TurΓ‘n number of the disjoint union of cycles, Discrete Appl. Math. 342 (2024), 260β274.
- [21] J.W. Moon, On independent complete subgraphs in a graph, Canadian J. Math. 20 (1968) 95β102.
- [22] V. Nikiforov, A spectral condition for odd cycles in graphs, Linear Algebra Appl. 428 (2008), no. 7, 1492β1498.
- [23] V. Nikiforov, Bounds on graph eigenvalues II, Linear Algebra Appl. 427 (2-3) (2007), 183β189.
- [24] B. Pittel, J. Spencer, N. Wormald, Sudden emergence of a giant -core in a random graph, J. Combin. Theory Ser. B 67 (1996), 111β151.
- [25] R.L. Shi, Z. Walsh, X.X. Yu, Planar TurΓ‘n number of the 7-cycle, arxiv:2306.13594v1.
- [26] R.L. Shi, Z. Walsh, X.X. Yu, Dense circuit graphs and the planar TurΓ‘n number of a cycle, arxiv:2310.06631v1.
- [27] M. Tait, J. Tobin, Three conjectures in extremal spectral graph theory, J. Combin. Theory Ser. B 126 (2017), 137β161.
- [28] M.Q. Zhai, B. Wang, Proof of a conjecture on the spectral radius of -free graphs, Linear Algebra Appl. 437 (7) (2012), 1641β1647.