This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

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.

Longfei Fanga,b, Huiqiu Lina, Yongtang Shic
a School of Mathematics, East China University of Science and Technology,
Shanghai 200237, China
b School of Mathematics and Finance, Chuzhou University,
Chuzhou, Anhui 239012, China
cCenter for Combinatorics and LPMC, Nankai University, Tianjin 300071, China
Corresponding author: [email protected] (H. Lin).

Abstract

Given a planar graph family β„±\mathcal{F}, let ex𝒫​(n,β„±){\rm ex}_{\mathcal{P}}(n,\mathcal{F}) and spex𝒫​(n,β„±){\rm spex}_{\mathcal{P}}(n,\mathcal{F}) be the maximum size and maximum spectral radius over all nn-vertex β„±\mathcal{F}-free planar graphs, respectively. Let t​Cβ„“tC_{\ell} be the disjoint union of tt copies of β„“\ell-cycles, and tβ€‹π’žt\mathcal{C} be the family of tt vertex-disjoint cycles without length restriction. Tait and Tobin [Three conjectures in extremal spectral graph theory, J. Combin. Theory Ser. B 126 (2017) 137–161] determined that K2+Pnβˆ’2K_{2}+P_{n-2} is the extremal spectral graph among all planar graphs with sufficiently large order nn, which implies the extremal graphs of both spex𝒫​(n,t​Cβ„“){\rm spex}_{\mathcal{P}}(n,tC_{\ell}) and spex𝒫​(n,tβ€‹π’ž){\rm spex}_{\mathcal{P}}(n,t\mathcal{C}) for tβ‰₯3t\geq 3 are K2+Pnβˆ’2K_{2}+P_{n-2}. In this paper, we first determine spex𝒫​(n,t​Cβ„“){\rm spex}_{\mathcal{P}}(n,tC_{\ell}) and spex𝒫​(n,tβ€‹π’ž){\rm spex}_{\mathcal{P}}(n,t\mathcal{C}) and characterize the unique extremal graph for 1≀t≀21\leq t\leq 2, β„“β‰₯3\ell\geq 3 and sufficiently large nn. Secondly, we obtain the exact values of ex𝒫​(n,2​C4){\rm ex}_{\mathcal{P}}(n,2C_{4}) and ex𝒫​(n,2β€‹π’ž){\rm ex}_{\mathcal{P}}(n,2\mathcal{C}), which solve a conjecture of Li [Planar TurΓ‘n number of the disjoint union of cycles, Discrete Appl. Math. 342 (2024) 260–274] for nβ‰₯2661n\geq 2661.

Keywords: Spectral radius; TurΓ‘n number; Planar graph; Vertex-disjoint cycles; Quadrilateral

AMS Classification: 05C35; 05C50

1 Introduction

Given a graph family β„±\mathcal{F}, a graph is said to be β„±\mathcal{F}-free if it does not contain any Fβˆˆβ„±F\in\mathcal{F} as a subgraph. When β„±={F}\mathcal{F}=\{F\}, we write FF-free instead of β„±\mathcal{F}-free. One of the earliest results in extremal graph theory is the TurΓ‘n’s theorem, which gives the maximum number of edges in an nn-vertex KkK_{k}-free graph. The TurΓ‘n number ex​(n,β„±){\rm ex}(n,\mathcal{F}) is the maximum number of edges in an β„±\mathcal{F}-free graph on nn vertices. Much attention has been given to TurΓ‘n numbers on cycles. FΓΌredi and Gunderson [12] determined ex​(n,C2​k+1){\rm ex}(n,C_{2k+1}) for all nn and kk. However, the exact value of ex​(n,C2​k){\rm ex}(n,C_{2k}) is still open. ErdΕ‘s [8] determined ex​(n,t​C3){\rm ex}(n,tC_{3}) for nβ‰₯400​(tβˆ’1)2n\geq 400(t-1)^{2}, and the unique extremal graph is characterized. Subsequently, Moon [21] showed that ErdΕ‘s’s result is still valid whenever n>9​tβˆ’112n>\frac{9t-11}{2}. ErdΕ‘s and PΓ³sa [9] showed that ex​(n,tβ€‹π’ž)=(2​tβˆ’1)​(nβˆ’t){\rm ex}(n,t\mathcal{C})=(2t-1)(n-t) for tβ‰₯2t\geq 2 and nβ‰₯24​tn\geq 24t. 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 β„±\mathcal{F}, denoted by spex𝒫​(n,β„±){\rm spex}_{\mathcal{P}}(n,\mathcal{F}), is the maximum spectral radius over all nn-vertex β„±\mathcal{F}-free planar graphs. An β„±\mathcal{F}-free planar graph on nn vertices with maximum spectral radius is called an extremal graph to spex𝒫​(n,β„±){\rm spex}_{\mathcal{P}}(n,\mathcal{F}). Boots and Royle [2] and independently Cao and Vince [3] conjectured that K2+Pnβˆ’2K_{2}+P_{n-2} is the unique planar graph with the maximum spectral radius where β€˜+’ means the join product. The conjecture was finally proved by Tait and Tobin [27] for sufficiently large nn.

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 FF be a planar graph and nβ‰₯max⁑{2.16Γ—1017,2​|V​(F)|}n\geq\max\{2.16\times 10^{17},2|V(F)|\}. If FF is a subgraph of 2​K1+Pn/22K_{1}+P_{n/2} but not of K2,nβˆ’2K_{2,n-2}, then the extremal graph to spex𝒫​(n,F){\rm spex}_{\mathcal{P}}(n,F) contains a copy of K2,nβˆ’2K_{2,n-2}.

Let t​Cβ„“tC_{\ell} be the disjoint union of tt copies of β„“\ell-cycles, and tβ€‹π’žt\mathcal{C} be the family of tt vertex-disjoint cycles without length restriction. For tβ‰₯3t\geq 3, it is easy to check that K2+Pnβˆ’2K_{2}+P_{n-2} is t​Cβ„“tC_{\ell}-free and tβ€‹π’žt\mathcal{C}-free. Theorem 1.1 implies that K2+Pnβˆ’2K_{2}+P_{n-2} is the extremal graph to spex𝒫​(n,t​Cβ„“){\rm spex}_{\mathcal{P}}(n,tC_{\ell}) and spex𝒫​(n,tβ€‹π’ž){\rm spex}_{\mathcal{P}}(n,t\mathcal{C}) for tβ‰₯3t\geq 3 and sufficiently large nn. For three positive integers n,n1,n2n,n_{1},n_{2} with n1β‰₯n2n_{1}\geq n_{2} and nβ‰₯n1+2​n2+2n\geq n_{1}+2n_{2}+2, we can find integers Ξ±\alpha and Ξ²\beta such that 1≀β≀n21\leq\beta\leq n_{2} and nβˆ’2=n1+α​n2+Ξ²n-2=n_{1}+\alpha n_{2}+\beta. Let

H​(n1,n2):=Pn1βˆͺα​Pn2βˆͺPΞ².H(n_{1},n_{2}):=P_{n_{1}}\cup\alpha P_{n_{2}}\cup P_{\beta}.

In the paper, we give answers to spex𝒫​(n,t​Cβ„“){\rm spex}_{\mathcal{P}}(n,tC_{\ell}) for t∈{1,2}t\in\{1,2\} as follows.

Theorem 1.2.

For integers β„“β‰₯3\ell\geq 3 and nβ‰₯max⁑{2.16Γ—1017,9Γ—2β„“βˆ’1+3}n\geq\max\{2.16\times 10^{17},9\times 2^{\ell-1}+3\}, the graph K2+H​(2β€‹β„“βˆ’3,β„“βˆ’2)K_{2}+H(2\ell-3,\ell-2) is the extremal graph to spex𝒫​(n,2​Cβ„“){\rm spex}_{\mathcal{P}}(n,2C_{\ell}).

Theorem 1.4 implies that K2+H​(3,1)K_{2}+H(3,1) is 2β€‹π’ž2\mathcal{C}-free for nβ‰₯5n\geq 5. By Theorem 1.2, K2+H​(3,1)K_{2}+H(3,1) is the extremal graph to spex𝒫​(n,2​C3){\rm spex}_{\mathcal{P}}(n,2C_{3}) for nβ‰₯2.16Γ—1017n\geq 2.16\times 10^{17}. Moreover, a planar graph is 2​C32C_{3}-free when it is 2β€‹π’ž2\mathcal{C}-free. Hence, one can easily get answer to spex𝒫​(n,2β€‹π’ž){\rm spex}_{\mathcal{P}}(n,2\mathcal{C}).

Corollary 1.1.

For nβ‰₯2.16Γ—1017n\geq 2.16\times 10^{17}, K2+H​(3,1)K_{2}+H(3,1) is the extremal graph to spex𝒫​(n,2β€‹π’ž){\rm spex}_{\mathcal{P}}(n,2\mathcal{C}).

We use JnJ_{n} to denote the graph obtained from K1+(nβˆ’1)​K1K_{1}+(n-1)K_{1} by embedding a maximum matching within its independent set. Nikiforov [23] and Zhai and Wang [28] showed that JnJ_{n} is the extremal graph to spex​(n,C4){\rm spex}(n,C_{4}) for odd and even nn, respectively. Clearly, JnJ_{n} is planar. This implies that JnJ_{n} is the extremal graph to spex𝒫​(n,C4){\rm spex}_{\mathcal{P}}(n,C_{4}). Nikiforov [22] and CioabΔƒ, Desai, and Tait [5] determined the spectral extremal graph among Cβ„“C_{\ell}-free graphs for odd β„“\ell and even β„“β‰₯6\ell\geq 6, respectively. Next we give the characterization of the spectral extremal graphs among Cβ„“C_{\ell}-free planar graphs.

Theorem 1.3.

For integers β„“β‰₯3\ell\geq 3 and nβ‰₯max⁑{2.16Γ—1017,9Γ—2βŒŠβ„“βˆ’12βŒ‹+3,62532β€‹βŒŠβ„“βˆ’32βŒ‹2+2}n\geq\max\{2.16\times 10^{17},9\times 2^{\lfloor\frac{\ell-1}{2}\rfloor}+3,\frac{625}{32}\lfloor\frac{\ell-3}{2}\rfloor^{2}+2\},
(i) K2,nβˆ’2K_{2,n-2} is the unique extremal graph to spex𝒫​(n,C3){\rm spex}_{\mathcal{P}}(n,C_{3}) for β„“=3\ell=3;
(ii) K2+H​(βŒˆβ„“βˆ’32βŒ‰,βŒŠβ„“βˆ’32βŒ‹)K_{2}+H(\lceil\frac{\ell-3}{2}\rceil,\lfloor\frac{\ell-3}{2}\rfloor) is the unique extremal graph to spex𝒫​(n,Cβ„“){\rm spex}_{\mathcal{P}}(n,C_{\ell}) for β„“β‰₯5\ell\geq 5.

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 nn-vertex β„±\mathcal{F}-free planar graph? This extremal number is called planar TurΓ‘n number of β„±\mathcal{F} and denoted by ex𝒫​(n,β„±){\rm ex}_{\mathcal{P}}(n,\mathcal{F}). The planar TurΓ‘n number for short cycles are studied in [4, 6, 7, 14, 15, 16, 25, 26], but ex𝒫​(n,Cβ„“){\rm ex}_{\mathcal{P}}(n,C_{\ell}) is still open for general β„“\ell. 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 ex𝒫​(n,tβ€‹π’ž)=nβˆ’1{\rm ex}_{\mathcal{P}}(n,t\mathcal{C})=n-1 for t=1t=1. Lan, Shi and Song [17] showed that ex𝒫​(n,tβ€‹π’ž)=3​nβˆ’6{\rm ex}_{\mathcal{P}}(n,t\mathcal{C})=3n-6 for tβ‰₯3t\geq 3, and the double wheel 2​K1+Cnβˆ’22K_{1}+C_{n-2} is an extremal graph. We prove the case of t=2t=2, which will be used to prove our main theorems.

Theorem 1.4.

For nβ‰₯5n\geq 5, we have ex𝒫​(n,2β€‹π’ž)=2​nβˆ’1{\rm ex}_{\mathcal{P}}(n,2\mathcal{C})=2n-1. The extremal graphs are obtained from 2​K1+C32K_{1}+C_{3} and an independent set of size nβˆ’5n-5 by joining each vertex of the independent set to any two vertices of the triangle.

Moreover, Lan, Shi and Song [17] also proved that ex𝒫​(n,t​Cβ„“)=3​nβˆ’6{\rm ex}_{\mathcal{P}}(n,tC_{\ell})=3n-6 for all t,β„“β‰₯3t,\ell\geq 3. They [19] further showed that ex𝒫​(n,2​C3)=⌈5​n2βŒ‰βˆ’5{\rm ex}_{\mathcal{P}}(n,2C_{3})=\lceil\frac{5n}{2}\rceil-5 and obtained lower bounds of ex𝒫​(n,2​Cβ„“){\rm ex}_{\mathcal{P}}(n,2C_{\ell}) for β„“β‰₯4\ell\geq 4, which was improved by Li [20] for sufficiently large nn recently. Li [20] also raised the following conjecture.

Conjecture 1.1.

If nβ‰₯23n\geq 23, then ex𝒫​(n,2​C4)≀197​(nβˆ’2){\rm ex}_{\mathcal{P}}(n,2C_{4})\leq\frac{19}{7}(n-2) and the bound is tight for 14∣(nβˆ’2)14\mid(n-2).

In this paper, we determine the exact value of ex𝒫​(n,2​C4){\rm ex}_{\mathcal{P}}(n,2C_{4}) for large nn, which solve the conjecture for nβ‰₯2661n\geq 2661.

Theorem 1.5.

For nβ‰₯2661n\geq 2661,

ex𝒫​(n,2​C4)={19​n7βˆ’6ifΒ 7∣n,⌊19​nβˆ’347βŒ‹otherwise.{\rm ex}_{\mathcal{P}}(n,2C_{4})=\left\{\begin{array}[]{ll}\frac{19n}{7}-6&\hbox{if $7\mid n$,}\\ \big{\lfloor}\frac{19n-34}{7}\big{\rfloor}{}{}{}{}{}{}{}&\hbox{otherwise.}\end{array}\right.

2 Proof of Theorem 1.1

Let A​(G)A(G) be the adjacency matrix of a planar graph GG, and ρ​(G)\rho(G) be its spectral radius, i.e., the maximum modulus of eigenvalues of A​(G)A(G). Throughout Sections 2 and 3, let GG be an extremal graph to spex𝒫​(n,F){\rm spex}_{\mathcal{P}}(n,F), and ρ\rho denote this spectral radius. By Perron-Frobenius theorem, there exists a positive eigenvector X=(x1,…,xn)TX=(x_{1},\ldots,x_{n})^{\mathrm{T}} corresponding to ρ\rho. Choose uβ€²βˆˆV​(G)u^{\prime}\in V(G) with xuβ€²=max⁑{xi|i=1,2,…,n}=1x_{u^{\prime}}=\max\{x_{i}~{}|~{}i=1,2,\dots,n\}=1. For a vertex uu and a positive integer ii, let Ni​(u)N_{i}(u) denote the set of vertices at distance ii from uu in GG. For two disjoint subset S,TβŠ‚V​(G)S,T\subset V(G), denote by G​[S,T]G[S,T] the bipartite subgraph of GG with vertex set SβˆͺTS\cup T that consist of all edges with one endpoint in SS and the other endpoint in TT. Set e​(S)=|E​(G​[S])|e(S)=|E(G[S])| and e​(S,T)=|E​(G​[S,T])|e(S,T)=|E(G[S,T])|. Since GG is a planar graph, we have

e​(S)≀3​|S|βˆ’6​and​e​(S,T)≀2​(|S|+|T|)βˆ’4.\displaystyle e(S)\leq 3|S|-6~{}~{}~{}\text{and}~{}~{}~{}e(S,T)\leq 2(|S|+|T|)-4. (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 ρ\rho.

Claim 2.1.

For nβ‰₯2.16Γ—1017n\geq 2.16\times 10^{17}, we have ρβ‰₯2​nβˆ’4.\rho\geq\sqrt{2n-4}.

Proof.

Note that K2,nβˆ’2K_{2,n-2} is planar and FF-free. Then, ρβ‰₯ρ​(K2,nβˆ’2)=2​nβˆ’4\rho\geq\rho(K_{2,n-2})=\sqrt{2n-4}, as GG is an extremal graph to spex𝒫​(n,F){\rm spex}_{\mathcal{P}}(n,F). ∎

Claim 2.2.

Set LΞ»={u∈V​(G)|xuβ‰₯1103​λ}L^{\lambda}=\{u\in V(G)~{}|~{}x_{u}\geq\frac{1}{10^{3}\lambda}\} for some constant Ξ»β‰₯1103\lambda\geq\frac{1}{10^{3}}. Then |LΞ»|≀λ​n105|L^{\lambda}|\leq\frac{\lambda n}{10^{5}}.

Proof.

By Claim 2.1, ρβ‰₯2​nβˆ’4\rho\geq\sqrt{2n-4}. Hence,

2​nβˆ’4103​λ≀ρ​xu=βˆ‘v∈NG​(u)xv≀dG​(u)\frac{\sqrt{2n-4}}{10^{3}\lambda}\leq\rho x_{u}=\sum_{v\in N_{G}(u)}x_{v}\leq d_{G}(u)

for each u∈Lλu\in L^{\lambda}. Summing this inequality over all vertices u∈Lλu\in L^{\lambda}, we obtain

|LΞ»|​2​nβˆ’4103β€‹Ξ»β‰€βˆ‘u∈LΞ»dG​(u)β‰€βˆ‘u∈V​(G)dG​(u)≀2​(3​nβˆ’6).\displaystyle|L^{\lambda}|\frac{\sqrt{2n-4}}{10^{3}\lambda}\leq\sum_{u\in L^{\lambda}}d_{G}(u)\leq\sum_{u\in V(G)}d_{G}(u)\leq 2(3n-6).

It follows that |LΞ»|≀3Γ—103​λ​2​nβˆ’4≀λ​n105|L^{\lambda}|\leq 3\times 10^{3}\lambda\sqrt{2n-4}\leq\frac{\lambda n}{10^{5}} as nβ‰₯2.16Γ—1017n\geq 2.16\times 10^{17}. ∎

Claim 2.3.

We have |L1|≀6Γ—104|L^{1}|\leq 6\times 10^{4}.

Proof.

Let u∈V​(G)u\in V(G) be an arbitrary vertex. For convenience, we use NiN_{i}, LiΞ»L_{i}^{\lambda} and Liλ¯\overline{L_{i}^{\lambda}} instead of Ni​(u)N_{i}(u), Ni​(u)∩LΞ»N_{i}(u)\cap L^{\lambda} and Ni​(u)βˆ–LΞ»N_{i}(u)\setminus L^{\lambda}, respectively. By Claim 2.1, ρβ‰₯2​nβˆ’4\rho\geq\sqrt{2n-4}. Then

(2​nβˆ’4)​xu≀ρ2​xu=dG​(u)​xu+βˆ‘v∈N1βˆ‘w∈N1​(v)βˆ–{u}xw.\displaystyle(2n-4)x_{u}\leq\rho^{2}x_{u}=d_{G}(u)x_{u}+\sum_{v\in N_{1}}\sum_{w\in N_{1}(v)\setminus\{u\}}x_{w}. (2)

Note that N1​(v)βˆ–{u}βŠ†N1βˆͺN2=L1Ξ»βˆͺL2Ξ»βˆͺL1λ¯βˆͺL2λ¯N_{1}(v)\setminus\{u\}\subseteq N_{1}\cup N_{2}=L_{1}^{\lambda}\cup L_{2}^{\lambda}\cup\overline{L_{1}^{\lambda}}\cup\overline{L_{2}^{\lambda}}. We can calculate βˆ‘v∈N1βˆ‘w∈N1​(v)βˆ–{u}xw\sum_{v\in N_{1}}\sum_{w\in N_{1}(v)\setminus\{u\}}x_{w} according to two cases w∈L1Ξ»βˆͺL2Ξ»w\in L_{1}^{\lambda}\cup L_{2}^{\lambda} or w∈L1λ¯βˆͺL2λ¯w\in\overline{L_{1}^{\lambda}}\cup\overline{L_{2}^{\lambda}}. We first consider the case w∈L1Ξ»βˆͺL2Ξ»w\in L_{1}^{\lambda}\cup L_{2}^{\lambda}. Clearly, N1=L1Ξ»βˆͺL1λ¯N_{1}=L_{1}^{\lambda}\cup\overline{L_{1}^{\lambda}} and xw≀1x_{w}\leq 1 for w∈L1Ξ»βˆͺL2Ξ»w\in L_{1}^{\lambda}\cup L_{2}^{\lambda}. We can see that

βˆ‘v∈N1βˆ‘w∈(L1Ξ»βˆͺL2Ξ»)xw≀(2​e​(L1Ξ»)+e​(L1Ξ»,L2Ξ»))+βˆ‘v∈L1Ξ»Β―βˆ‘w∈(L1Ξ»βˆͺL2Ξ»)xw.\displaystyle\sum_{v\in N_{1}}\sum_{w\in(L_{1}^{\lambda}\cup L_{2}^{\lambda})}\!\!x_{w}\leq\big{(}2e(L_{1}^{\lambda})+e(L_{1}^{\lambda},L_{2}^{\lambda})\big{)}+\sum_{v\in\overline{L_{1}^{\lambda}}}\sum_{w\in(L_{1}^{\lambda}\cup L_{2}^{\lambda})}\!\!\!x_{w}. (3)

By Claim 2.2, we have |LΞ»|≀λ​n105|L^{\lambda}|\leq\frac{\lambda n}{10^{5}}. Moreover, L1Ξ»βˆͺL2Ξ»βŠ†LΞ»L_{1}^{\lambda}\cup L_{2}^{\lambda}\subseteq L^{\lambda}. Then, by (1), we have

2​e​(L1Ξ»)+e​(L1Ξ»,L2Ξ»)≀2​(3​|L1Ξ»|βˆ’6)+(2​(|L1Ξ»|+|L2Ξ»|)βˆ’4)<8​|LΞ»|≀8​λ​n105.\displaystyle 2e(L_{1}^{\lambda})+e(L_{1}^{\lambda},L_{2}^{\lambda})\leq 2(3|L_{1}^{\lambda}|-6)+(2(|L_{1}^{\lambda}|+|L_{2}^{\lambda}|)-4)<8|L^{\lambda}|\leq\frac{8\lambda n}{10^{5}}. (4)

Next, we consider the remain case w∈L1λ¯βˆͺL2λ¯w\in\overline{L_{1}^{\lambda}}\cup\overline{L_{2}^{\lambda}}. Clearly, xw≀1103​λx_{w}\leq\frac{1}{10^{3}\lambda} for w∈L1λ¯βˆͺL2λ¯w\in\overline{L_{1}^{\lambda}}\cup\overline{L_{2}^{\lambda}}. Then

βˆ‘v∈N1βˆ‘w∈L1λ¯βˆͺL2λ¯xw≀(e​(L1Ξ»,L1λ¯βˆͺL2λ¯)+2​e​(L1λ¯)+e​(L1λ¯,L2λ¯))​1103​λ<6​n103​λ,\displaystyle\sum_{v\in N_{1}}\sum_{w\in\overline{L_{1}^{\lambda}}\cup\overline{L_{2}^{\lambda}}}\!\!\!x_{w}\leq\Big{(}e(L_{1}^{\lambda},\overline{L_{1}^{\lambda}}\cup\overline{L_{2}^{\lambda}})+2e(\overline{L_{1}^{\lambda}})+e(\overline{L_{1}^{\lambda}},\overline{L_{2}^{\lambda}})\Big{)}\frac{1}{10^{3}\lambda}<\frac{6n}{10^{3}\lambda}, (5)

where e​(L1Ξ»,L1λ¯βˆͺL2λ¯)+2​e​(L1λ¯)+e​(L1λ¯,L2λ¯)≀2​e​(G)<6​ne(L_{1}^{\lambda},\overline{L_{1}^{\lambda}}\cup\overline{L_{2}^{\lambda}})+2e(\overline{L_{1}^{\lambda}})+e(\overline{L_{1}^{\lambda}},\overline{L_{2}^{\lambda}})\leq 2e(G)<6n.

Combining (2)-(5), we obtain

(2​nβˆ’4)​xu<dG​(u)​xu+βˆ‘v∈L1Ξ»Β―βˆ‘w∈(L1Ξ»βˆͺL2Ξ»)xw+(8​λ10+60Ξ»)​n104.\displaystyle(2n-4)x_{u}<d_{G}(u)x_{u}+\sum_{v\in\overline{L_{1}^{\lambda}}}\sum_{w\in(L_{1}^{\lambda}\cup L_{2}^{\lambda})}\!\!\!x_{w}+\Big{(}\frac{8\lambda}{10}+\frac{60}{\lambda}\Big{)}\frac{n}{10^{4}}. (6)

Now we prove that dG​(u)β‰₯n104d_{G}(u)\geq\frac{n}{10^{4}} for each u∈L1u\in L^{1}. Suppose to the contrary that there exists a vertex u~∈L1\widetilde{u}\in L^{1} with dG​(u~)<n104d_{G}(\widetilde{u})<\frac{n}{10^{4}}. Note that xu~β‰₯1103x_{\widetilde{u}}\geq\frac{1}{10^{3}} as u~∈L1\widetilde{u}\in L^{1}. Setting u=u~u=\widetilde{u}, Ξ»=10\lambda=10 and combining (6), we have

2​nβˆ’4103<dG​(u~)​xu~+βˆ‘v∈L110Β―βˆ‘w∈(L110βˆͺL210)xw+14​n104.\displaystyle\frac{2n-4}{10^{3}}<d_{G}(\widetilde{u})x_{\widetilde{u}}+\sum_{v\in\overline{L_{1}^{10}}}\sum_{w\in(L_{1}^{10}\cup L_{2}^{10})}x_{w}+\frac{14n}{10^{4}}. (7)

By (1), we have

e​(L110Β―,L110βˆͺL210)<2​(|L110Β―|+|L110βˆͺL210|)≀2​(|N1|+|L10|)≀4​n104,\displaystyle e\big{(}\overline{L_{1}^{10}},L_{1}^{10}\cup L_{2}^{10}\big{)}<2\big{(}|\overline{L_{1}^{10}}|+|L_{1}^{10}\cup L_{2}^{10}|\big{)}\leq 2\big{(}|N_{1}|+|L^{10}|\big{)}\leq\frac{4n}{10^{4}},

where |N1|=dG​(u~)<n104|N_{1}|=d_{G}(\widetilde{u})<\frac{n}{10^{4}} and |L10|≀n104|L^{10}|\leq\frac{n}{10^{4}} by Claim 2.2. Combining this with dG​(u~)<n104d_{G}(\widetilde{u})<\frac{n}{10^{4}} gives

dG​(u~)​xu~+βˆ‘v∈L110Β―βˆ‘w∈(L110βˆͺL210)xw+14​n104≀dG​(u~)+e​(L110Β―,L110βˆͺL210)+14​n104≀19​n104,d_{G}(\widetilde{u})x_{\widetilde{u}}+\sum_{v\in\overline{L_{1}^{10}}}\sum_{w\in(L_{1}^{10}\cup L_{2}^{10})}x_{w}+\frac{14n}{10^{4}}\leq d_{G}(\widetilde{u})+e\big{(}\overline{L_{1}^{10}},L_{1}^{10}\cup L_{2}^{10}\big{)}+\frac{14n}{10^{4}}\leq\frac{19n}{10^{4}},

which contradicts (7). Therefore, dG​(u)β‰₯n104d_{G}(u)\geq\frac{n}{10^{4}} for each u∈L1u\in L^{1}. Summing this inequality over all vertices u∈L1u\in L^{1}, we obtain

|L1|​n104β‰€βˆ‘u∈L1dG​(u)≀2​e​(G)≀6​n,|L^{1}|\frac{n}{10^{4}}\leq\sum_{u\in L^{1}}d_{G}(u)\leq 2e(G)\leq 6n,

which yields that |L1|≀6Γ—104|L^{1}|\leq 6\times 10^{4}. ∎

For convenience, we use LL, LiL_{i} and LiΒ―\overline{L_{i}} instead of L1L^{1}, Ni​(u)∩L1N_{i}(u)\cap L^{1} and Ni​(u)βˆ–L1N_{i}(u)\setminus L^{1}, respectively. Now we give a lower bound for degrees of vertices in L1L^{1}.

Claim 2.4.

For every u∈Lu\in L, we have dG​(u)β‰₯(xuβˆ’41000)​nd_{G}(u)\geq(x_{u}-\frac{4}{1000})n.

Proof.

Let L1Β―β€²\overline{L_{1}}^{\prime} be the subset of L1Β―\overline{L_{1}} in which each vertex has at least 22 neighbors in L1βˆͺL2L_{1}\cup L_{2}. We first prove that |L1Β―β€²|≀|L1βˆͺL2|2|\overline{L_{1}}^{\prime}|\leq|L_{1}\cup L_{2}|^{2}. If |L1βˆͺL2|=1|L_{1}\cup L_{2}|=1, then L1Β―β€²\overline{L_{1}}^{\prime} is empty, as desired. It remains the case |L1βˆͺL2|β‰₯2|L_{1}\cup L_{2}|\geq 2. Suppose to the contrary that |L1Β―β€²|>|L1βˆͺL2|2|\overline{L_{1}}^{\prime}|>|L_{1}\cup L_{2}|^{2}. Since there are only (|L1βˆͺL2|2)\binom{|L_{1}\cup L_{2}|}{2} options for vertices in L1Β―β€²\overline{L_{1}}^{\prime} to choose a set of 22 neighbors from L1βˆͺL2L_{1}\cup L_{2}, we can find a set of 22 vertices in L1βˆͺL2L_{1}\cup L_{2} with at least ⌈|L1Β―β€²|/(|L1βˆͺL2|2)βŒ‰β‰₯3\Big{\lceil}|\overline{L_{1}}^{\prime}|/\binom{|L_{1}\cup L_{2}|}{2}\Big{\rceil}\geq 3 common neighbors in L1Β―β€²\overline{L_{1}}^{\prime}. Moreover, one can observe that uβˆ‰L1βˆͺL2u\notin L_{1}\cup L_{2} and L1Β―β€²βŠ†L1Β―βŠ†N1​(u)\overline{L_{1}}^{\prime}\subseteq\overline{L_{1}}\subseteq N_{1}(u). Hence, GG contains a copy of K3,3K_{3,3}, contradicting that GG is planar. Hence, |L1Β―β€²|≀|L1βˆͺL2|2|\overline{L_{1}}^{\prime}|\leq|L_{1}\cup L_{2}|^{2}. It follows that

e​(L1Β―,L1βˆͺL2)≀|L1Β―βˆ–L1Β―β€²|+|L1βˆͺL2|​|L1Β―β€²|≀dG​(u)+(6Γ—104)3≀dG​(u)+n1000,\displaystyle e(\overline{L_{1}},L_{1}\cup L_{2})\leq|\overline{L_{1}}\setminus\overline{L_{1}}^{\prime}|+|L_{1}\cup L_{2}||\overline{L_{1}}^{\prime}|\leq d_{G}(u)+(6\times 10^{4})^{3}\leq d_{G}(u)+\frac{n}{1000},

where the second last inequality holds as |L1Β―β€²|≀|L1βˆͺL2|2|\overline{L_{1}}^{\prime}|\leq|L_{1}\cup L_{2}|^{2} and |L1βˆͺL2|≀|L|≀6Γ—104|L_{1}\cup L_{2}|\leq|L|\leq 6\times 10^{4}, and the last inequality holds as nβ‰₯2.16Γ—1017n\geq 2.16\times 10^{17}. Setting Ξ»=1\lambda=1 and combining the above inequality with (6), we have

(2​nβˆ’4)​xu≀dG​(u)+(dG​(u)+n103)+61​n104,\displaystyle(2n-4)x_{u}\leq d_{G}(u)+\Big{(}d_{G}(u)+\frac{n}{10^{3}}\Big{)}+\frac{61n}{10^{4}},

which yields that dG​(u)β‰₯(nβˆ’2)​xuβˆ’71​n2Γ—104β‰₯(xuβˆ’41000)​nd_{G}(u)\geq(n-2)x_{u}-\frac{71n}{2\times 10^{4}}\geq(x_{u}-\frac{4}{1000})n. ∎

Claim 2.5.

There exists a vertex uβ€²β€²βˆˆL1βˆͺL2u^{\prime\prime}\in L_{1}\cup L_{2} such that xuβ€²β€²β‰₯9881000x_{u^{\prime\prime}}\geq\frac{988}{1000}.

Proof.

Setting u=uβ€²u=u^{\prime}, Ξ»=1\lambda=1 and combining (6), we have

2​nβˆ’4<dG​(uβ€²)+βˆ‘v∈L1Β―βˆ‘w∈L1βˆͺL2xw+61​n104,\displaystyle 2n-4<d_{G}(u^{\prime})+\sum_{v\in\overline{L_{1}}}\sum_{w\in L_{1}\cup L_{2}}\!\!\!x_{w}+\frac{61n}{10^{4}},

which yields that

βˆ‘v∈L1Β―βˆ‘w∈L1βˆͺL2xwβ‰₯2​nβˆ’4βˆ’61​n104βˆ’dG​(uβ€²)β‰₯993​n1000.\sum_{v\in\overline{L_{1}}}\sum_{w\in L_{1}\cup L_{2}}x_{w}\geq 2n-4-\frac{61n}{10^{4}}-d_{G}(u^{\prime})\geq\frac{993n}{1000}.

From Claim 2.4 we have dG​(uβ€²)β‰₯996​n1000d_{G}(u^{\prime})\geq\frac{996n}{1000} as uβ€²βˆˆLu^{\prime}\in L. For simplicity, set NL1​(uβ€²)=NG​(uβ€²)∩L1N_{L_{1}}(u^{\prime})=N_{G}(u^{\prime})\cap L_{1} and dL1​(uβ€²)=|NL1​(uβ€²)|d_{L_{1}}(u^{\prime})=|N_{L_{1}}(u^{\prime})|. By Claim 2.3, |L|≀6Γ—104|L|\leq 6\times 10^{4}. Then, dL1​(uβ€²)≀|L1|≀|L|≀n1000d_{L_{1}}(u^{\prime})\leq|L_{1}|\leq|L|\leq\frac{n}{1000} as nβ‰₯2.16Γ—1017n\geq 2.16\times 10^{17}. It infers that

dL1¯​(uβ€²)=dG​(uβ€²)βˆ’dL1​(uβ€²)β‰₯dG​(uβ€²)βˆ’|L|β‰₯995​n1000.d_{\overline{L_{1}}}(u^{\prime})=d_{G}(u^{\prime})-d_{L_{1}}(u^{\prime})\geq d_{G}(u^{\prime})-|L|\geq\frac{995n}{1000}.

Combining this with (1) gives

e​(L1Β―,L1βˆͺL2)≀e​(L1Β―,L)βˆ’dL1¯​(uβ€²)≀(2​nβˆ’4)βˆ’995​n1000≀1005​n1000.e(\overline{L_{1}},L_{1}\cup L_{2})\leq e(\overline{L_{1}},L)-d_{\overline{L_{1}}}(u^{\prime})\leq(2n-4)-\frac{995n}{1000}\leq\frac{1005n}{1000}.

By averaging, there is a vertex uβ€²β€²βˆˆL1βˆͺL2u^{\prime\prime}\in L_{1}\cup L_{2} such that

xuβ€²β€²β‰₯βˆ‘v∈L1Β―βˆ‘w∈(L1βˆͺL2)xwe​(L1Β―,L1βˆͺL2)β‰₯993​n10001005​n1000β‰₯9881000,x_{u^{\prime\prime}}\geq\frac{\sum_{v\in\overline{L_{1}}}\sum_{w\in(L_{1}\cup L_{2})}x_{w}}{e(\overline{L_{1}},L_{1}\cup L_{2})}\geq\frac{\frac{993n}{1000}}{\frac{1005n}{1000}}\geq\frac{988}{1000},

as desired. ∎

Notice that xuβ€²=1x_{u^{\prime}}=1 and xuβ€²β€²β‰₯9881000x_{u^{\prime\prime}}\geq\frac{988}{1000}. By Claim 2.4, we have

dG​(uβ€²)β‰₯9961000​n​and​dG​(uβ€²β€²)β‰₯9841000​n.\displaystyle d_{G}(u^{\prime})\geq\frac{996}{1000}n~{}~{}~{}\text{and}~{}~{}~{}d_{G}(u^{\prime\prime})\geq\frac{984}{1000}n. (8)

Now, let D={uβ€²,uβ€²β€²}D=\{u^{\prime},u^{\prime\prime}\}, R=NG​(uβ€²)∩NG​(uβ€²β€²)R=N_{G}(u^{\prime})\cap N_{G}(u^{\prime\prime}), and R1=V​(G)βˆ–(DβˆͺR)R_{1}=V(G)\setminus(D\cup R). Thus, |R1|≀(nβˆ’dG​(uβ€²))+(nβˆ’dG​(uβ€²β€²))≀2​n100|R_{1}|\leq(n-d_{G}(u^{\prime}))+(n-d_{G}(u^{\prime\prime}))\leq\frac{2n}{100}. Now, we prove that the eigenvector entries of vertices in RβˆͺR1R\cup R_{1} are small.

Claim 2.6.

Let u∈RβˆͺR1u\in R\cup R_{1}. Then xu≀110x_{u}\leq\frac{1}{10}.

Proof.

For any vertex u∈R1u\in R_{1}, we can see that

dD​(u)≀1​and​dR​(u)≀2.\displaystyle d_{D}(u)\leq 1~{}~{}~{}\text{and}~{}~{}~{}d_{R}(u)\leq 2. (9)

In fact, if dR​(u)β‰₯3d_{R}(u)\geq 3, then G​[NG​(u)βˆͺD]G[N_{G}(u)\cup D] contains a copy of K3,3K_{3,3}, contradicting that GG is planar. By (9), dG​(u)=dD​(u)+dR​(u)+dR1​(u)≀3+dR1​(u)d_{G}(u)=d_{D}(u)+d_{R}(u)+d_{R_{1}}(u)\leq 3+d_{R_{1}}(u). Note that |R1|≀2​n100|R_{1}|\leq\frac{2n}{100} and e​(R1)≀3​|R1|e(R_{1})\leq 3|R_{1}| by (1). Thus,

Οβ€‹βˆ‘u∈R1xuβ‰€βˆ‘u∈R1dG​(u)β‰€βˆ‘u∈R1(3+dR1​(u))≀3​|R1|+2​e​(R1)≀9​|R1|≀18​n100,\rho\sum_{u\in R_{1}}x_{u}\leq\sum_{u\in R_{1}}d_{G}(u)\leq\sum_{u\in R_{1}}(3+d_{R_{1}}(u))\leq 3|R_{1}|+2e(R_{1})\leq 9|R_{1}|\leq\frac{18n}{100},

which yields βˆ‘u∈R1xu≀18​n100​ρ\sum_{u\in R_{1}}x_{u}\leq\frac{18n}{100\rho}. For any u∈RβˆͺR1u\in R\cup R_{1}, dR​(u)≀2d_{R}(u)\leq 2 as GG is K3,3K_{3,3}-free. It follows that

ρ​xu=βˆ‘v∈NG​(u)xvβ‰€βˆ‘v∈ND​(u)xv+βˆ‘v∈NR​(u)xv+βˆ‘v∈NR1​(u)xv≀4+18​n100​ρ.\rho x_{u}=\sum_{v\in N_{G}(u)}x_{v}\leq\sum_{v\in N_{D}(u)}x_{v}+\sum_{v\in N_{R}(u)}x_{v}+\sum_{v\in N_{R_{1}}(u)}x_{v}\leq 4+\frac{18n}{100\rho}.

Dividing both sides by ρ\rho, we get xu≀110x_{u}\leq\frac{1}{10} as ρβ‰₯2​nβˆ’4\rho\geq\sqrt{2n-4}. ∎

Refer to caption
Figure 1: A local structure of Gβ€²G^{\prime}.
Claim 2.7.

R1R_{1} is empty.

Proof.

Suppose to the contrary that a=|R1|>0a=|R_{1}|>0. Assume that Gβˆ—G^{*} is a planar embedding of G​[DβˆͺR]G[D\cup R], and u1,…,u|R|u_{1},\dots,u_{|R|} are around uβ€²u^{\prime} in clockwise order in Gβˆ—G^{*}, with subscripts interpreted modulo |R||R|. Recall that |R1|≀2​n100|R_{1}|\leq\frac{2n}{100}. Then |R|=nβˆ’2βˆ’|R1|>97​n100>|R1||R|=n-2-|R_{1}|>\frac{97n}{100}>|R_{1}|. By the pigeonhole principle, there exists an integer i≀|R|i\leq|R| such that u′​ui​u′′​ui+1​uβ€²u^{\prime}u_{i}u^{\prime\prime}u_{i+1}u^{\prime} is a face of the plane graph Gβˆ—G^{*}. We modify the graph Gβˆ—G^{*} by joining each vertex in R1R_{1} to each vertex in DD and making these edges cross the face u′​ui​u′′​ui+1​uβ€²u^{\prime}u_{i}u^{\prime\prime}u_{i+1}u^{\prime}, to obtain the graph Gβ€²G^{\prime} (see Gβ€²G^{\prime} in Figure 1). Then, Gβ€²G^{\prime} is a plane graph.

We first show that Gβ€²G^{\prime} is FF-free. Suppose to the contrary, then Gβ€²G^{\prime} contains a subgraph Fβ€²F^{\prime} isomorphic to FF. From the modification, we can see that V​(Fβ€²)∩R1V(F^{\prime})\cap R_{1} is not empty. Moreover, since |R|>97​n100>|V​(F)|=|V​(Fβ€²)||R|>\frac{97n}{100}>|V(F)|=|V(F^{\prime})|, we have

|Rβˆ–V​(Fβ€²)|=|R|βˆ’|R∩V​(Fβ€²)|>|V​(Fβ€²)|βˆ’|V​(Fβ€²)∩R|β‰₯|V​(Fβ€²)∩R1|.|R\setminus V(F^{\prime})|=|R|-|R\cap V(F^{\prime})|>|V(F^{\prime})|-|V(F^{\prime})\cap R|\geq|V(F^{\prime})\cap R_{1}|.

Then, we may assume that V​(Fβ€²)∩R1={v1,…,vb}V(F^{\prime})\cap R_{1}=\{v_{1},\dots,v_{b}\} and {w1,…,wb}∈Rβˆ–V​(Fβ€²)\{w_{1},\dots,w_{b}\}\in R\setminus V(F^{\prime}). Clearly, NG′​(vi)=DβŠ†NG′​(wi)N_{G^{\prime}}(v_{i})=D\subseteq N_{G^{\prime}}(w_{i}) for each i∈{1,…,b}i\in\{1,\dots,b\}. This indicates that a copy of FF is already present in GG, a contradiction. Therefore, Gβ€²G^{\prime} is FF-free.

In what follows, we shall obtain a contradiction by showing that ρ​(Gβ€²)>ρ\rho(G^{\prime})>\rho. Clearly, G​[R1]G[R_{1}] is planar. Then, there exists a vertex v1∈R1v_{1}\in R_{1} with dR1​(v1)≀5d_{R_{1}}(v_{1})\leq 5. Define R2=R1βˆ–{v1}R_{2}=R_{1}\setminus\{v_{1}\}. Repeat this step, we obtain a sequence of sets R1,R2,β‹―,RaR_{1},R_{2},\cdots,R_{a} with dRi​(vi)≀5d_{R_{i}}(v_{i})\leq 5 for each i∈{1,…,a}i\in\{1,\ldots,a\}. Combining this with (9), we have

βˆ‘w∈NDβˆͺRβˆͺRi​(vi)xw≀1+βˆ‘w∈NR​(vi)xw+βˆ‘w∈NRi​(vi)xw≀1710,\displaystyle\sum_{w\in N_{D\cup R\cup R_{i}}(v_{i})}x_{w}\leq 1+\sum_{w\in N_{R}(v_{i})}x_{w}+\sum_{w\in N_{R_{i}}(v_{i})}x_{w}\leq\frac{17}{10}, (10)

where the last inequality holds as xw<110x_{w}<\frac{1}{10} for any w∈RβˆͺR1w\in R\cup R_{1} by Claim 2.6. It is not hard to verify that in graph GG the set of edges incident to vertices in R1R_{1} is ⋃i=1a{w​vi|w∈NDβˆͺRβˆͺRi​(vi)}\bigcup_{i=1}^{a}\{wv_{i}~{}|~{}w\in N_{D\cup R\cup R_{i}}(v_{i})\}. Note that xuβ€²+xuβ€²β€²β‰₯19881000x_{u^{\prime}}+x_{u^{\prime\prime}}\geq\frac{1988}{1000}. Combining these with (10), we have

ρ​(Gβ€²)βˆ’Οβ‰₯2XT​Xβ€‹βˆ‘i=1axvi​((xuβ€²+xuβ€²β€²)βˆ’βˆ‘w∈NDβˆͺRβˆͺRi​(vi)xw)>0,\rho(G^{\prime})-\rho\geq\frac{2}{X^{\mathrm{T}}X}\sum_{i=1}^{a}x_{v_{i}}\left((x_{u^{\prime}}+x_{u^{\prime\prime}})-\sum_{w\in N_{D\cup R\cup R_{i}}(v_{i})}x_{w}\right)>0,

which contradicts that GG is extremal to spex𝒫​(n,F){\rm spex}_{\mathcal{P}}(n,F). Therefore, R1R_{1} is empty. ∎

By Claim 2.7, GG contains a copy of K2,nβˆ’2K_{2,n-2}. The proof of Theorem 1.1 is complete. ∎

3 Proofs of Theorems 1.2 and 1.3

Recall that GG is an extremal graph to spex𝒫​(n,F){\rm spex}_{\mathcal{P}}(n,F). In the case F=C3F=C_{3}, by Theorem 1.1, GG contains a copy of K2,nβˆ’2K_{2,n-2}. We further obtain that Gβ‰…K2,nβˆ’2G\cong K_{2,n-2} as GG is triangle-free (otherwise, adding any edge increases triangles, a contradiction). Now we consider the case F∈{Cβ„“|β„“β‰₯5}βˆͺ{2​Cβ„“|β„“β‰₯3}F\in\{C_{\ell}~{}|~{}\ell\geq 5\}\cup\{2C_{\ell}~{}|~{}\ell\geq 3\}. Let PnP_{n} denote a path on nn 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 GG, which help us to present an approach to prove Theorems 1.2 and 1.3.

Lemma 3.1.

Let nβ‰₯max⁑{2.16Γ—1017,2​|V​(F)|}n\geq\max\{2.16\times 10^{17},2|V(F)|\} and F∈{Cβ„“|β„“β‰₯5}βˆͺ{2​Cβ„“|β„“β‰₯3}F\in\{C_{\ell}~{}|~{}\ell\geq 5\}\cup\{2C_{\ell}~{}|~{}\ell\geq 3\}. Then Gβ‰…K2+G​[R]G\cong K_{2}+G[R] and G​[R]G[R] is a disjoint union of paths.

Proof.

Since nβ‰₯2​|V​(F)|n\geq 2|V(F)| and F∈{Cβ„“|β„“β‰₯5}βˆͺ{2​Cβ„“|β„“β‰₯3}F\in\{C_{\ell}~{}|~{}\ell\geq 5\}\cup\{2C_{\ell}~{}|~{}\ell\geq 3\}, we can see that FF is a subgraph of 2​K1+Pn/22K_{1}+P_{n/2} but not of K2,nβˆ’2K_{2,n-2}. Hence, by Theorem 1.1, GG contains a copy of K2,nβˆ’2K_{2,n-2}, and uβ€²,uβ€²β€²u^{\prime},u^{\prime\prime} are the vertices of degree nβˆ’2n-2 and RR is the set of vertices of degree two in K2,nβˆ’2K_{2,n-2}.

Claim 3.1.

u′​uβ€²β€²βˆˆE​(G)u^{\prime}u^{\prime\prime}\in E(G).

Proof.

Suppose to the contrary that u′​uβ€²β€²βˆ‰E​(G)u^{\prime}u^{\prime\prime}\notin E(G). Assume that Gβˆ—G^{*} is a planar embedding of GG, and u1,…,unβˆ’2u_{1},\dots,u_{n-2} are around uβ€²u^{\prime} in clockwise order in Gβˆ—G^{*}, with subscripts interpreted modulo nβˆ’2n-2. If RR induces an cycle u1​u2​…​unβˆ’2​u1u_{1}u_{2}\dots u_{n-2}u_{1}, then we can easily check that Gβˆ—G^{*} contains a copy of FF, a contradiction. Thus, there exists an integer i≀nβˆ’2i\leq n-2 such that ui​ui+1βˆ‰E​(Gβˆ—β€‹[R])u_{i}u_{i+1}\notin E(G^{*}[R]). Furthermore, u′​ui​u′′​ui+1​uβ€²u^{\prime}u_{i}u^{\prime\prime}u_{i+1}u^{\prime} is a face in Gβˆ—G^{*}.

We modify the graph Gβˆ—G^{*} by adding the edge u′​uβ€²β€²u^{\prime}u^{\prime\prime} and making u′​uβ€²β€²u^{\prime}u^{\prime\prime} cross the face u′​ui​u′′​ui+1​uβ€²u^{\prime}u_{i}u^{\prime\prime}u_{i+1}u^{\prime}, to obtain the graph Gβ€²G^{\prime}. Clearly, Gβ€²G^{\prime} is a plane graph. We shall show that Gβ€²G^{\prime} is FF-free. Suppose to the contrary that Gβ€²G^{\prime} contains a subgraph Fβ€²F^{\prime} isomorphic to FF. If Fβ€²=Cβ„“F^{\prime}=C_{\ell} for some β„“β‰₯5\ell\geq 5, then Gβ€²G^{\prime} contains an β„“\ell-cycle containing u′​uβ€²β€²u^{\prime}u^{\prime\prime}, say u′​u′′​u1′​u2′​…​uβ„“βˆ’2′​uβ€²u^{\prime}u^{\prime\prime}u_{1}^{\prime}u_{2}^{\prime}\dots u_{\ell-2}^{\prime}u^{\prime}. However, an β„“{\ell}-cycle u′​u1′​u′′​u2′​…​uβ„“βˆ’2′​uβ€²u^{\prime}u_{1}^{\prime}u^{\prime\prime}u_{2}^{\prime}\dots u_{\ell-2}^{\prime}u^{\prime} is already present in GG, a contradiction. If Fβ€²=2​Cβ„“F^{\prime}=2C_{\ell} for some β„“β‰₯3\ell\geq 3, then Fβ€²F^{\prime} contains two vertex-disjoint β„“\ell-cycles C1C^{1} and C2C^{2}. From the modification, we can see that one of C1C^{1} and C2C^{2} (say C1C^{1}) contains the edge u′​uβ€²β€²u^{\prime}u^{\prime\prime}. This implies that C2C^{2} is a subgraph of G​[R]G[R]. However, G​[V​(C2)βˆͺ{uβ€²,uβ€²β€²}]G[V(C^{2})\cup\{u^{\prime},u^{\prime\prime}\}] contains a K5K_{5}-minor, contradicting the fact that GG is planar. Hence, Gβ€²G^{\prime} is FF-free. However, ρ​(Gβ€²)>ρ\rho(G^{\prime})>\rho, contradicting that GG is extremal to spex𝒫​(n,F){\rm spex}_{\mathcal{P}}(n,F). Therefore, u′​uβ€²β€²βˆˆE​(G)u^{\prime}u^{\prime\prime}\in E(G). ∎

From Claim 3.1 we know that G=K2+G​[R]G=K_{2}+G[R]. It remains to show that G​[R]G[R] is a disjoint union of paths. Theorem 1.1 and Claim 3.1 imply that uβ€²u^{\prime} and uβ€²β€²u^{\prime\prime} are dominating vertices. Furthermore, since GG is K5K_{5}-minor-free and K3,3K_{3,3}-minor-free, we can see that G​[R]G[R] is K3K_{3}-minor-free and K1,3K_{1,3}-minor-free. This implies that G​[R]G[R] is an acyclic graph with maximum degree at most 2. Thus, G​[R]G[R] is a disjoint union of paths. The result follows. ∎

Now we give a transformation that we will use in subsequent proof.

Definition 3.1.

Let s1s_{1} and s2s_{2} be two integers with s1β‰₯s2β‰₯1s_{1}\geq s_{2}\geq 1, and let H=Ps1βˆͺPs2βˆͺH0H=P_{s_{1}}\cup P_{s_{2}}\cup H_{0}, where H0H_{0} is a disjoint union of paths. We say that Hβˆ—H^{*} is an (s1,s2)(s_{1},s_{2})-transformation of HH if

Hβˆ—:={Ps1+1βˆͺPs2βˆ’1βˆͺH0ifΒ s2β‰₯2,Ps1+s2βˆͺH0ifΒ s2=1.H^{*}:=\left\{\begin{array}[]{ll}P_{s_{1}+1}\cup P_{s_{2}-1}\cup H_{0}&\hbox{if $s_{2}\geq 2$,}\\ P_{s_{1}+s_{2}}\cup H_{0}{}{}{}{}{}{}{}&\hbox{if $s_{2}=1$.}\end{array}\right.

Clearly, Hβˆ—H^{*} is a disjoint union of paths, which implies that K2+Hβˆ—K_{2}+H^{*} is planar. If G​[R]β‰…HG[R]\cong H, then we shall show that ρ​(K2+Hβˆ—)>ρ\rho(K_{2}+H^{*})>\rho for sufficiently large nn.

Lemma 3.2.

Let HH and Hβˆ—H^{*} be defined as in Definition 3.1, nβ‰₯max⁑{2.16Γ—1017,2​|V​(F)|,9Γ—2s2+1+3}n\geq\max\{2.16\times 10^{17},2|V(F)|,9\times 2^{s_{2}+1}+3\} and F∈{Cβ„“|β„“β‰₯5}βˆͺ{2​Cβ„“|β„“β‰₯3}F\in\{C_{\ell}~{}|~{}\ell\geq 5\}\cup\{2C_{\ell}~{}|~{}\ell\geq 3\}. If G​[R]β‰…HG[R]\cong H, then ρ​(K2+Hβˆ—)>ρ\rho(K_{2}+H^{*})>\rho.

Proof.

By Lemma 3.1, Gβ‰…K2+G​[R]G\cong K_{2}+G[R] and G​[R]G[R] is a disjoint union of paths. If G​[R]β‰…HG[R]\cong H, then G=K2+HG=K_{2}+H. Assume that P1:=v1​v2​…​vs1P^{1}:=v_{1}v_{2}\dots v_{s_{1}} and P2:=w1​w2​…​ws2P^{2}:=w_{1}w_{2}\dots w_{s_{2}} are two components of HH. If s2=1s_{2}=1, then HβŠ‚Hβˆ—H\subset H^{*}, and so GβŠ‚K2+Hβˆ—G\subset K_{2}+H^{*}. It follows that ρ​(P2+Hβˆ—)>ρ\rho(P_{2}+H^{*})>\rho, the result holds. Next, we deal with the case s2=2s_{2}=2. If xv1≀xw1x_{v_{1}}\leq x_{w_{1}}, then let Hβ€²H^{\prime} be obtained from HH by deleting the edge v1​v2v_{1}v_{2} and adding the edge v2​w1v_{2}w_{1}. Clearly, Hβ€²β‰…Hβˆ—H^{\prime}\cong H^{*}. Moreover,

ρ​(K2+Hβˆ—)βˆ’Οβ‰₯XT​(A​(K2+Hβˆ—)βˆ’A​(G))​XXT​Xβ‰₯2XT​X​(xw1βˆ’xv1)​xv2β‰₯0.\displaystyle\rho(K_{2}+H^{*})-\rho\geq\frac{X^{\mathrm{T}}\big{(}A(K_{2}+H^{*})-A(G)\big{)}X}{X^{\mathrm{T}}X}\geq\frac{2}{X^{\mathrm{T}}X}(x_{w_{1}}-x_{v_{1}})x_{v_{2}}\geq 0.

Since XX is a positive eigenvector of GG, we have ρ​xv1=2+xv2\rho x_{v_{1}}=2+x_{v_{2}}. If ρ​(K2+Hβˆ—)=ρ\rho(K_{2}+H^{*})=\rho, then XX is also a positive eigenvector of K2+Hβˆ—K_{2}+H^{*}, and so ρ​(K2+Hβˆ—)​xv1=2\rho(K_{2}+H^{*})x_{v_{1}}=2, contradicting that ρ​xv1=2+xv2\rho x_{v_{1}}=2+x_{v_{2}}. Thus, ρ​(K2+Hβˆ—)>ρ\rho(K_{2}+H^{*})>\rho, the result holds. The case xv1>xw1x_{v_{1}}>x_{w_{1}} is similar and hence omitted here.

It remains the case s2β‰₯3s_{2}\geq 3. We shall give characterizations of eigenvector entries of vertices in RR in the following two claims.

Claim 3.2.

For any vertex u∈Ru\in R, we have xu∈[2ρ,2ρ+6ρ2]x_{u}\in[\frac{2}{\rho},\frac{2}{\rho}+\frac{6}{\rho^{2}}].

Proof.

By Lemma 3.1, uβ€²u^{\prime} and uβ€²β€²u^{\prime\prime} are dominating vertices of GG. So, xuβ€²=xuβ€²β€²=1x_{u^{\prime}}=x_{u^{\prime\prime}}=1. Hence

ρ​xu=xuβ€²+xuβ€²β€²+βˆ‘v∈NR​(u)xv=2+βˆ‘v∈NR​(u)xv.\displaystyle\rho x_{u}=x_{u^{\prime}}+x_{u^{\prime\prime}}+\sum_{v\in N_{R}(u)}x_{v}=2+\sum_{v\in N_{R}(u)}x_{v}. (11)

Moreover, by Lemma 3.1, dR​(v)≀2d_{R}(v)\leq 2 for any v∈Rv\in R. Combining this with Claim 2.6 and (11) gives xu∈[2ρ,3ρ]x_{u}\in\big{[}\frac{2}{\rho},\frac{3}{\rho}\big{]}. Furthermore, again by (11), ρ​xu∈[2,2+6ρ]\rho x_{u}\in\big{[}2,2+\frac{6}{\rho}\big{]}, which yields that xu∈[2ρ,2ρ+6ρ2]x_{u}\in\big{[}\frac{2}{\rho},\frac{2}{\rho}+\frac{6}{\rho^{2}}\big{]}. ∎

Claim 3.3.

Let ii be a positive integer. Set Ai=[2Οβˆ’6Γ—2iρ2,2ρ+6Γ—2iρ2]A_{i}=\big{[}\frac{2}{\rho}-\frac{6\times 2^{i}}{\rho^{2}},\frac{2}{\rho}+\frac{6\times 2^{i}}{\rho^{2}}\big{]} and Bi=[βˆ’6Γ—2iρ2,6Γ—2iρ2]B_{i}=\big{[}-\frac{6\times 2^{i}}{\rho^{2}},\frac{6\times 2^{i}}{\rho^{2}}\big{]}. Then,
(i) for any i∈{1,…,⌊s1βˆ’12βŒ‹}i\in\{1,\dots,\lfloor\frac{s_{1}-1}{2}\rfloor\}, ρi​(xvi+1βˆ’xvi)∈Ai\rho^{i}(x_{v_{i+1}}-x_{v_{i}})\in A_{i};
(ii) for any i∈{1,…,⌊s2βˆ’12βŒ‹}i\in\{1,\dots,\lfloor\frac{s_{2}-1}{2}\rfloor\}, ρi​(xwi+1βˆ’xwi)∈Bi\rho^{i}(x_{w_{i+1}}-x_{w_{i}})\in B_{i};
(iii) for any i∈{1,…,⌊s22βŒ‹}i\in\{1,\dots,\lfloor\frac{s_{2}}{2}\rfloor\}, ρi​(xviβˆ’xwi)∈Bi\rho^{i}(x_{v_{i}}-x_{w_{i}})\in B_{i}.

Proof.

(i) It suffices to prove that for any i∈{1,…,⌊s1βˆ’12βŒ‹}i\in\{1,\dots,\lfloor\frac{s_{1}-1}{2}\rfloor\},

ρi​(xvj+1βˆ’xvj)∈{AiifΒ j=i,BiifΒ i+1≀j≀s1βˆ’iβˆ’1.\displaystyle\rho^{i}(x_{v_{j+1}}-x_{v_{j}})\in\left\{\begin{array}[]{ll}A_{i}&\hbox{if $j=i$,}\\ B_{i}&\hbox{if $i+1\leq j\leq s_{1}-i-1$.}\end{array}\right.

We shall proceed the proof by induction on ii. Clearly,

ρ​xvj=βˆ‘v∈NG​(vj)xv={2+xv2ifΒ j=1,2+xvjβˆ’1+xvj+1ifΒ 2≀j≀s1βˆ’1.\displaystyle\rho x_{v_{j}}=\sum_{v\in N_{G}(v_{j})}x_{v}=\left\{\begin{array}[]{ll}2+x_{v_{2}}&\hbox{if $j=1$,}\\ 2+x_{v_{j-1}}+x_{v_{j+1}}&\hbox{if $2\leq j\leq s_{1}-1$.}\end{array}\right. (15)

By Claim 3.2, we have

ρ​(xvj+1βˆ’xvj)={xv1+xv3βˆ’xv2∈A1ifΒ j=1,(xvjβˆ’xvjβˆ’1)+(xvj+2βˆ’xj+1)∈B1ifΒ 2≀j≀s1βˆ’2.\displaystyle\rho(x_{v_{j+1}}-x_{v_{j}})=\left\{\begin{array}[]{ll}x_{v_{1}}+x_{v_{3}}-x_{v_{2}}\in A_{1}&\hbox{if $j=1$,}\\ (x_{v_{j}}-x_{v_{j-1}})+(x_{v_{j+2}}-x_{j+1})\in B_{1}&\hbox{if $2\leq j\leq s_{1}-2$.}\end{array}\right. (18)

So the result is true when i=1i=1. Assume then that 2≀iβ‰€βŒŠs1βˆ’12βŒ‹2\leq i\leq\lfloor\frac{s_{1}-1}{2}\rfloor, which implies that s1β‰₯2​i+1s_{1}\geq 2i+1. For i≀j≀s1βˆ’iβˆ’1i\leq j\leq s_{1}-i-1, ρ​(xvj+1βˆ’xvj)=(xvjβˆ’xvjβˆ’1)+(xvj+2βˆ’xvj+1)\rho(x_{v_{j+1}}-x_{v_{j}})=(x_{v_{j}}-x_{v_{j-1}})+(x_{v_{j+2}}-x_{v_{j+1}}), and so

ρi​(xvj+1βˆ’xvj)=ρiβˆ’1​(xvjβˆ’xvjβˆ’1)+ρiβˆ’1​(xvj+2βˆ’xvj+1).\displaystyle\rho^{i}(x_{v_{j+1}}-x_{v_{j}})=\rho^{i-1}(x_{v_{j}}-x_{v_{j-1}})+\rho^{i-1}(x_{v_{j+2}}-x_{v_{j+1}}). (19)

By the induction hypothesis, ρiβˆ’1​(xviβˆ’xviβˆ’1)∈Aiβˆ’1\rho^{i-1}(x_{v_{i}}-x_{v_{i-1}})\in A_{i-1} and ρiβˆ’1​(xvi+2βˆ’xvi+1)∈Biβˆ’1\rho^{i-1}(x_{v_{i+2}}-x_{v_{i+1}})\in B_{i-1}. Setting j=ij=i and combining (19), we have ρi​(xvi+1βˆ’xvi)∈Ai\rho^{i}(x_{v_{i+1}}-x_{v_{i}})\in A_{i}, as desired. If i+1≀j≀s1βˆ’iβˆ’1i+1\leq j\leq s_{1}-i-1, then by the induction hypothesis, ρiβˆ’1​(xvjβˆ’xvjβˆ’1)∈Biβˆ’1\rho^{i-1}(x_{v_{j}}-x_{v_{j-1}})\in B_{i-1} and ρiβˆ’1​(xvj+2βˆ’xvj+1)∈Biβˆ’1\rho^{i-1}(x_{v_{j+2}}-x_{v_{j+1}})\in B_{i-1}. Thus, by (19), we have ρi​(xvj+1βˆ’xvj)∈Bi\rho^{i}(x_{v_{j+1}}-x_{v_{j}})\in B_{i}, 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 i∈{1,…,⌊s22βŒ‹}i\in\{1,\dots,\lfloor\frac{s_{2}}{2}\rfloor\} and j∈{i,…,s2βˆ’i}j\in\{i,\dots,s_{2}-i\}, ρi​(xvjβˆ’xwj)∈Bi.\rho^{i}(x_{v_{j}}-x_{w_{j}})\in B_{i}. We shall proceed the proof by induction on ii. Clearly,

ρ​xwj=βˆ‘w∈NG​(wj)xw={2+xw2ifΒ j=1,2+xwjβˆ’1+xwj+1ifΒ 2≀j≀s2βˆ’1.\rho x_{w_{j}}=\sum_{w\in N_{G}(w_{j})}x_{w}=\left\{\begin{array}[]{ll}2+x_{w_{2}}&\hbox{if $j=1$,}\\ 2+x_{w_{j-1}}+x_{w_{j+1}}&\hbox{if $2\leq j\leq s_{2}-1$.}\end{array}\right.

Combining this with (15) and Claim 3.2 gives

ρ​(xvjβˆ’xwj)={xv2βˆ’xw2∈[βˆ’6ρ2,6ρ2]βŠ‚B1ifΒ j=1,(xvj+1βˆ’xwj+1)+(xvjβˆ’1βˆ’xwjβˆ’1)∈B1ifΒ 2≀j≀s2βˆ’1.\rho(x_{v_{j}}-x_{w_{j}})=\left\{\begin{array}[]{ll}x_{v_{2}}-x_{w_{2}}\in\big{[}-\frac{6}{\rho^{2}},\frac{6}{\rho^{2}}\big{]}\subset B_{1}&\hbox{if $j=1$,}\\ (x_{v_{j+1}}-x_{w_{j+1}})+(x_{v_{j-1}}-x_{w_{j-1}})\in B_{1}&\hbox{if $2\leq j\leq s_{2}-1$.}\end{array}\right.

So the claim is true when i=1i=1. Assume then that 2≀iβ‰€βŒŠs22βŒ‹2\leq i\leq\lfloor\frac{s_{2}}{2}\rfloor, which implies that s2β‰₯2​is_{2}\geq 2i. If i≀j≀s2βˆ’ii\leq j\leq s_{2}-i, then ρ​(xvjβˆ’xwj)=(xvjβˆ’1βˆ’xwjβˆ’1)+(xvj+1βˆ’xwj+1)\rho(x_{v_{j}}-x_{w_{j}})=(x_{v_{j-1}}-x_{w_{j-1}})+(x_{v_{j+1}}-x_{w_{j+1}}), and so

ρi​(xvjβˆ’xwj)=ρiβˆ’1​(xvjβˆ’1βˆ’xwjβˆ’1)+ρiβˆ’1​(xvj+1βˆ’xwj+1).\displaystyle\rho^{i}(x_{v_{j}}-x_{w_{j}})=\rho^{i-1}(x_{v_{j-1}}-x_{w_{j-1}})+\rho^{i-1}(x_{v_{j+1}}-x_{w_{j+1}}). (20)

By the induction hypothesis, ρiβˆ’1​(xvjβˆ’1βˆ’xwjβˆ’1)∈Biβˆ’1\rho^{i-1}(x_{v_{j-1}}-x_{w_{j-1}})\in B_{i-1} and ρiβˆ’1​(xvj+1βˆ’xwj+1)∈Biβˆ’1\rho^{i-1}(x_{v_{j+1}}-x_{w_{j+1}})\in B_{i-1}. Combining these with (20), we have ρi​(xvjβˆ’xwj)=Bi\rho^{i}(x_{v_{j}}-x_{w_{j}})=B_{i}. Hence the result holds. ∎

Since nβ‰₯9Γ—2s2+1+3n\geq 9\times 2^{s_{2}+1}+3, we have ρβ‰₯2​nβˆ’4>6Γ—2s2/2\rho\geq\sqrt{2n-4}>6\times 2^{s_{2}/{2}}, and so

2ρi+1βˆ’6Γ—2iρi+2>(2ρi+1βˆ’6Γ—2iρi+2)βˆ’6Γ—2iρi+2>0​for any​i≀s22.\displaystyle\frac{2}{\rho^{i+1}}-\frac{6\times 2^{i}}{\rho^{i+2}}>\left(\frac{2}{\rho^{i+1}}-\frac{6\times 2^{i}}{\rho^{i+2}}\right)-\frac{6\times 2^{i}}{\rho^{i+2}}>0~{}~{}\text{for any}~{}~{}i\leq\frac{s_{2}}{2}.

Combining this with Claim 3.3, we obtain

xvi+1βˆ’xviβ‰₯2ρi+1βˆ’6Γ—2iρi+2>0​for any​i≀min⁑{s22,⌊s1βˆ’12βŒ‹},\displaystyle x_{v_{i+1}}-x_{v_{i}}\geq\frac{2}{\rho^{i+1}}-\frac{6\times 2^{i}}{\rho^{i+2}}>0~{}~{}\text{for any}~{}~{}i\leq\min\Big{\{}\frac{s_{2}}{2},\Big{\lfloor}\frac{s_{1}-1}{2}\Big{\rfloor}\Big{\}}, (21)

and

xvi+1βˆ’xwi=(xvi+1βˆ’xvi)+(xviβˆ’xwi)β‰₯(2ρi+1βˆ’6Γ—2iρi+2)βˆ’6Γ—2iρi+2>0\displaystyle x_{v_{i+1}}-x_{w_{i}}=(x_{v_{i+1}}-x_{v_{i}})+(x_{v_{i}}-x_{w_{i}})\geq\left(\frac{2}{\rho^{i+1}}-\frac{6\times 2^{i}}{\rho^{i+2}}\right)-\frac{6\times 2^{i}}{\rho^{i+2}}>0 (22)

for any i≀min⁑{⌊s22βŒ‹,⌊s1βˆ’12βŒ‹}i\leq\min\big{\{}\big{\lfloor}\frac{s_{2}}{2}\big{\rfloor},\big{\lfloor}\frac{s_{1}-1}{2}\big{\rfloor}\big{\}}. Similarly,

xwi+1>xwi​and​xwi+1>xvi​for any​iβ‰€βŒŠs2βˆ’12βŒ‹.\displaystyle x_{w_{i+1}}>x_{w_{i}}~{}~{}\text{and}~{}~{}x_{w_{i+1}}>x_{v_{i}}~{}~{}\text{for any}~{}~{}i\leq\Big{\lfloor}\frac{s_{2}-1}{2}\Big{\rfloor}. (23)

Recall that s2β‰₯3s_{2}\geq 3. Let t1,t2t_{1},t_{2} be integers with t1,t2β‰₯1t_{1},t_{2}\geq 1 and t1+t2=s2βˆ’1t_{1}+t_{2}=s_{2}-1, and Hβ€²H^{\prime} be obtained from HH by deleting edges vt1​vt1+1,wt2​wt2+1v_{t_{1}}v_{t_{1}+1},w_{t_{2}}w_{t_{2}+1} and adding edges vt1​wt2,vt1+1​wt2+1v_{t_{1}}w_{t_{2}},v_{t_{1}+1}w_{t_{2}+1}. Since t1+t2=s2βˆ’1t_{1}+t_{2}=s_{2}-1, we can see that Hβ€²β‰…Hβˆ—H^{\prime}\cong H^{*}. Moreover,

ρ​(K2+Hβˆ—)βˆ’Οβ‰₯XT​(A​(K2+Hβˆ—)βˆ’A​(G))​XXT​Xβ‰₯2XT​X​(xvt1+1βˆ’xwt2)​(xwt2+1βˆ’xvt1).\displaystyle\rho(K_{2}+H^{*})-\rho\geq\frac{X^{\mathrm{T}}\big{(}A(K_{2}+H^{*})-A(G)\big{)}X}{X^{\mathrm{T}}X}\geq\frac{2}{X^{\mathrm{T}}X}(x_{v_{t_{1}+1}}-x_{w_{t_{2}}})(x_{w_{t_{2}+1}}-x_{v_{t_{1}}}). (24)

Now, we consider the following two cases to complete the proof.

Case 1. s2s_{2} is odd.

Set t1=s2βˆ’12t_{1}=\frac{s_{2}-1}{2}. Then, t2=s2βˆ’12t_{2}=\frac{s_{2}-1}{2} as t1+t2=s2βˆ’1t_{1}+t_{2}=s_{2}-1. By (22) and (23), we have xvt1+1>xwt2x_{v_{t_{1}+1}}>x_{w_{t_{2}}} and xwt2+1>xvt1x_{w_{t_{2}+1}}>x_{v_{t_{1}}}, and so ρ​(K2+Hβˆ—)>ρ\rho(K_{2}+H^{*})>\rho by (24), as desired.

Case 2. s2s_{2} is even.

We only consider the subcase xws2/2β‰₯xvs2/2x_{w_{s_{2}/2}}\geq x_{v_{s_{2}/2}}. The proof of the subcase xws2/2<xvs2/2x_{w_{s_{2}/2}}<x_{v_{s_{2}/2}} is similar and hence omitted here. Set t1=s22t_{1}=\frac{s_{2}}{2}. Then, t2=s2βˆ’22t_{2}=\frac{s_{2}-2}{2} as t1+t2=s2βˆ’1t_{1}+t_{2}=s_{2}-1. Moreover, xwt2+1β‰₯xvt1x_{w_{t_{2}+1}}\geq x_{v_{t_{1}}}, as xws2/2β‰₯xvs2/2x_{w_{s_{2}/2}}\geq x_{v_{s_{2}/2}}. If s1=s2s_{1}=s_{2}, then s1s_{1} is even. This implies that xv(s1+2)/2=xvs1/2x_{v_{(s_{1}+2)/2}}=x_{v_{s_{1}/2}} by symmetry, that is, xvt1+1=xvt1x_{v_{t_{1}+1}}=x_{v_{t_{1}}}. If s1β‰₯s2+1s_{1}\geq s_{2}+1, then by (21), xvt1+1>xvt1x_{v_{t_{1}+1}}>x_{v_{t_{1}}}. In both situations, we have xvt1+1β‰₯xvt1x_{v_{t_{1}+1}}\geq x_{v_{t_{1}}}. From (22) we have xvt1>xwt2x_{v_{t_{1}}}>x_{w_{t_{2}}}, which implies that xvt1+1>xwt2x_{v_{t_{1}+1}}>x_{w_{t_{2}}}. Furthermore, we have ρ​(K2+Hβˆ—)β‰₯ρ\rho(K_{2}+H^{*})\geq\rho by (24). If ρ​(K2+Hβˆ—)=ρ\rho(K_{2}+H^{*})=\rho, then XX is a positive eigenvector of K2+Hβˆ—K_{2}+H^{*}, and so ρ​(K2+Hβˆ—)​xvt1=2+xvt1βˆ’1+xwt2\rho(K_{2}+H^{*})x_{v_{t_{1}}}=2+x_{v_{t_{1}-1}}+x_{w_{t_{2}}}. On the other hand, ρ​xvt1=2+xvt1βˆ’1+xvt1+1\rho x_{v_{t_{1}}}=2+x_{v_{t_{1}-1}}+x_{v_{t_{1}+1}}, since XX is a positive eigenvector of GG. It follows that xvt1+1=xwt2x_{v_{t_{1}+1}}=x_{w_{t_{2}}}, which is a contradiction. Thus, ρ​(K2+Hβˆ—)>ρ\rho(K_{2}+H^{*})>\rho, as desired.

This completes the proof of Lemma 3.2. ∎

From Lemma 3.1, we may assume that G​[R]=βˆͺi=1tPniG[R]=\cup_{i=1}^{t}P_{n_{i}}, where tβ‰₯2t\geq 2 and n1β‰₯n2β‰₯β‹―β‰₯ntn_{1}\geq n_{2}\geq\cdots\geq n_{t}. Let HH be a disjoint union of tt paths. We use ni​(H)n_{i}(H) to denote the order of the ii-th longest path of HH for any i∈{1,…,t}i\in\{1,\dots,t\}. 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 HH is a disjoint union of tβ‰₯2t\geq 2 paths, then K2+HK_{2}+H is 2​Cβ„“2C_{\ell}-free if and only if n1​(H)≀2β€‹β„“βˆ’3n_{1}(H)\leq 2\ell-3 and n2​(H)β‰€β„“βˆ’2n_{2}(H)\leq\ell-2.

Proof.

We first claim that K2+HK_{2}+H is 2​Cβ„“2C_{\ell}-free if and only if Pn1​(H)βˆͺPn2​(H)P_{n_{1}(H)}\cup P_{n_{2}(H)} is 2​Pβ„“βˆ’12P_{\ell-1}-free. Equivalently, K2+HK_{2}+H contains a copy of 2​Cβ„“2C_{\ell} if and only if Pn1​(H)βˆͺPn2​(H)P_{n_{1}(H)}\cup P_{n_{2}(H)} contains a copy of 2​Pβ„“βˆ’12P_{\ell-1}. Assume that K2+HK_{2}+H contains two vertex-disjoint β„“\ell-cycles C1C^{1} and C2C^{2}, and V​(K2)={uβ€²,uβ€²β€²}V(K_{2})=\{u^{\prime},u^{\prime\prime}\}. Since HH is acyclic, we can see that CiC^{i} must contain at least one vertex of uβ€²u^{\prime} and uβ€²β€²u^{\prime\prime} for any i∈{1,2}i\in\{1,2\}. Without loss of generality, assume that uβ€²βˆˆV​(C1)u^{\prime}\in V(C^{1}) and uβ€²β€²βˆˆV​(C2)u^{\prime\prime}\in V(C^{2}). Then, C1βˆ’{uβ€²}β‰…C2βˆ’{uβ€²β€²}β‰…Pβ„“βˆ’1C^{1}-\{u^{\prime}\}\cong C^{2}-\{u^{\prime\prime}\}\cong P_{\ell-1}, and so HH contains a 2​Pβ„“βˆ’12P_{\ell-1}. We can further find that Pn1​(H)βˆͺPn2​(H)P_{n_{1}(H)}\cup P_{n_{2}(H)} contains a 2​Pβ„“βˆ’12P_{\ell-1}. Conversely, assume that Pn1​(H)βˆͺPn2​(H)P_{n_{1}(H)}\cup P_{n_{2}(H)} contains two vertex-disjoint paths P1P^{1} and P2P^{2} such that P1β‰…P2β‰…Pβ„“βˆ’1P^{1}\cong P^{2}\cong P_{\ell-1}. Thus, the subgraph induced by V​(P1)βˆͺ{uβ€²}V(P^{1})\cup\{u^{\prime}\} contains a copy of Cβ„“C_{\ell}. Similarly, the subgraph induced by V​(P2)βˆͺ{uβ€²β€²}V(P^{2})\cup\{u^{\prime\prime}\} contains a copy of Cβ„“C_{\ell}. This indicates that K2+HK_{2}+H contains a copy of 2​Cβ„“2C_{\ell}. So, the claim holds.

Next, we claim that Pn1​(H)βˆͺPn2​(H)P_{n_{1}(H)}\cup P_{n_{2}(H)} is 2​Pβ„“βˆ’12P_{\ell-1}-free if and only if n1​(H)≀2β€‹β„“βˆ’3n_{1}(H)\leq 2\ell-3 and n2​(H)β‰€β„“βˆ’2n_{2}(H)\leq\ell-2. If Pn1​(H)βˆͺPn2​(H)P_{n_{1}(H)}\cup P_{n_{2}(H)} is 2​Pβ„“βˆ’12P_{\ell-1}-free, then n1​(H)≀2β€‹β„“βˆ’3n_{1}(H)\leq 2\ell-3 (otherwise, Pn1​(H)P_{n_{1}(H)} contains a copy of 2​Pβ„“βˆ’12P_{\ell-1}, a contradiction); n2​(H)β‰€β„“βˆ’2n_{2}(H)\leq\ell-2 (otherwise, Pn1​(H)βˆͺPn2​(H)P_{n_{1}(H)}\cup P_{n_{2}(H)} contains a copy of 2​Pβ„“βˆ’12P_{\ell-1} as n1​(H)β‰₯n2​(H)β‰₯β„“βˆ’1n_{1}(H)\geq n_{2}(H)\geq\ell-1, a contradiction). Conversely, if n1​(H)≀2β€‹β„“βˆ’3n_{1}(H)\leq 2\ell-3 and n2​(H)β‰€β„“βˆ’2n_{2}(H)\leq\ell-2, then it is not hard to verify that Pn1​(H)βˆͺPn2​(H)P_{n_{1}(H)}\cup P_{n_{2}(H)} is 2​Pβ„“βˆ’12P_{\ell-1}-free.

This completes the proof of Claim 3.4. ∎

Recall that nin_{i} (resp. ni​(H)n_{i}(H)) is the order of the ii-th longest path of G​[R]G[R] (resp. HH) for any i∈{1,…,t}i\in\{1,\dots,t\}. By Claim 3.4, n1≀2β€‹β„“βˆ’3n_{1}\leq 2\ell-3 and n2β‰€β„“βˆ’2n_{2}\leq\ell-2. By a direct computation, we have 9Γ—2β„“βˆ’1+3β‰₯max⁑{2​|V​(F)|,9Γ—2n2+1+3}9\times 2^{\ell-1}+3\geq\max\{2|V(F)|,9\times 2^{n_{2}+1}+3\}, and hence

nβ‰₯max⁑{2.16Γ—1017,2​|V​(F)|,9Γ—2n2+1+3}.\displaystyle n\geq\max\{2.16\times 10^{17},2|V(F)|,9\times 2^{n_{2}+1}+3\}. (25)

We first claim that n1=2β€‹β„“βˆ’3n_{1}=2\ell-3. Suppose to the contrary that n1≀2β€‹β„“βˆ’4n_{1}\leq 2\ell-4. Let Hβ€²H^{\prime} be an (n1,nt)(n_{1},n_{t})-transformation of G​[R]G[R]. Clearly, n1​(Hβ€²)=n1​(H)+1≀2β€‹β„“βˆ’3n_{1}(H^{\prime})=n_{1}(H)+1\leq 2\ell-3 and n2​(Hβ€²)=n2β‰€β„“βˆ’2n_{2}(H^{\prime})=n_{2}\leq\ell-2. By Claim 3.4, K2+Hβ€²K_{2}+H^{\prime} is 2​Cβ„“2C_{\ell}-free. However, by (25) and Lemma 3.2, we have ρ​(K2+Hβ€²)>ρ\rho(K_{2}+H^{\prime})>\rho, contradicting that GG is extremal to spex𝒫​(n,2​Cβ„“){\rm spex}_{\mathcal{P}}(n,2C_{\ell}). Thus, n1=2β€‹β„“βˆ’3n_{1}=2\ell-3, the claim holds.

Note that ni≀n2β‰€β„“βˆ’2n_{i}\leq n_{2}\leq\ell-2 for each i∈{2,…,tβˆ’1}i\in\{2,\dots,t-1\}. We then claim that ni=β„“βˆ’2n_{i}=\ell-2 for i∈{2,…,tβˆ’1}i\in\{2,\dots,t-1\}. If β„“=3\ell=3, then ni=1n_{i}=1, as niβ‰€β„“βˆ’2n_{i}\leq\ell-2, for each i∈{2,…,t}i\in\{2,\dots,t\} and the claim holds trivially. Now let β„“β‰₯4\ell\geq 4. Suppose to the contrary, then set i0=min⁑{i|2≀i≀tβˆ’1,niβ‰€β„“βˆ’3}.i_{0}=\min\{i~{}|~{}2\leq i\leq t-1,n_{i}\leq\ell-3\}. Let Hβ€²H^{\prime} be an (ni0,nt)(n_{i_{0}},n_{t})-transformation of G​[R]G[R]. Clearly, n1​(Hβ€²)=n1=2β€‹β„“βˆ’3n_{1}(H^{\prime})=n_{1}=2\ell-3 and n2​(Hβ€²)=max⁑{n2,ni0+1}β‰€β„“βˆ’2n_{2}(H^{\prime})=\max\{n_{2},n_{i_{0}}+1\}\leq\ell-2. By Claim 3.4, K2+Hβ€²K_{2}+H^{\prime} is 2​Cβ„“2C_{\ell}-free. However, by (25) and Lemma 3.2, we have ρ​(K2+Hβ€²)>ρ\rho(K_{2}+H^{\prime})>\rho, contradicting that GG is extremal to spex𝒫​(n,2​Cβ„“){\rm spex}_{\mathcal{P}}(n,2C_{\ell}). So, the claim holds.

Since n1=2β€‹β„“βˆ’3n_{1}=2\ell-3, ni=β„“βˆ’2n_{i}=\ell-2 for i∈{2,…,tβˆ’1}i\in\{2,\dots,t-1\} and ntβ‰€β„“βˆ’2n_{t}\leq\ell-2, we can see that G​[R]β‰…H​(2β€‹β„“βˆ’3,β„“βˆ’2)G[R]\cong H(2\ell-3,\ell-2). 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 β–‘\Box

Proof of Theorem 1.3. It remains the case β„“β‰₯5\ell\geq 5. We first give the following claim.

Claim 3.5.

If HH is a disjoint union of tβ‰₯2t\geq 2 paths, then K2+HK_{2}+H is Cβ„“C_{\ell}-free if and only if n1​(H)+n2​(H)β‰€β„“βˆ’3n_{1}(H)+n_{2}(H)\leq\ell\!-\!3.

Proof.

One can observe that the longest cycle in K2+HK_{2}+H is of order n1​(H)+n2​(H)+2n_{1}(H)+n_{2}(H)+2. Furthermore, K2+(Pn1​(H)βˆͺPn2​(H))K_{2}+(P_{n_{1}(H)}\cup P_{n_{2}(H)}) contains a cycle CiC_{i} for each i∈{3,…,n1​(H)+n2​(H)+2}i\in\{3,\dots,n_{1}(H)+n_{2}(H)+2\}. Consequently, n1​(H)+n2​(H)+2β‰€β„“βˆ’1n_{1}(H)+n_{2}(H)+2\leq\ell-1 if and only if K2+HK_{2}+H is Cβ„“C_{\ell}-free. Hence, the claim holds. ∎

By Claim 3.5, n1+n2β‰€β„“βˆ’3n_{1}+n_{2}\leq\ell-3, which implies that n2β‰€βŒŠβ„“βˆ’32βŒ‹n_{2}\leq\lfloor\frac{\ell-3}{2}\rfloor as n1β‰₯n2n_{1}\geq n_{2}. By a direct computation, we have 9Γ—2βŒŠβ„“βˆ’12βŒ‹+3β‰₯max⁑{2​|V​(F)|,9Γ—2n2+1+3}9\times 2^{\lfloor\frac{\ell-1}{2}\rfloor}+3\geq\max\{2|V(F)|,9\times 2^{n_{2}+1}+3\}, and hence

nβ‰₯max⁑{2.16Γ—1017,2​|V​(F)|,9Γ—2n2+1+3}.\displaystyle n\geq\max\{2.16\times 10^{17},2|V(F)|,9\times 2^{n_{2}+1}+3\}. (26)

Since β„“β‰₯5\ell\geq 5, we have βŒŠβ„“βˆ’32βŒ‹β‰₯1\lfloor\frac{\ell-3}{2}\rfloor\geq 1, which implies that βŒŠβ„“βˆ’32βŒ‹2β‰₯βŒŠβ„“βˆ’32βŒ‹\lfloor\frac{\ell-3}{2}\rfloor^{2}\geq\lfloor\frac{\ell-3}{2}\rfloor and 3β€‹βŒŠβ„“βˆ’32βŒ‹2β‰₯2β€‹βŒŠβ„“βˆ’32βŒ‹+1β‰₯β„“βˆ’33\lfloor\frac{\ell-3}{2}\rfloor^{2}\geq 2\lfloor\frac{\ell-3}{2}\rfloor+1\geq\ell-3. Consequently, since nβ‰₯62532β€‹βŒŠβ„“βˆ’32βŒ‹2+2>6β€‹βŒŠβ„“βˆ’32βŒ‹2+2n\geq\frac{625}{32}\lfloor\frac{\ell-3}{2}\rfloor^{2}+2>6\lfloor\frac{\ell-3}{2}\rfloor^{2}+2, we have

n>βŒŠβ„“βˆ’32βŒ‹2+2β€‹βŒŠβ„“βˆ’32βŒ‹+(β„“βˆ’3)+2β‰₯n22+3​n2+n1+2\displaystyle n>\Big{\lfloor}\frac{\ell-3}{2}\Big{\rfloor}^{2}+2\Big{\lfloor}\frac{\ell-3}{2}\Big{\rfloor}+(\ell-3)+2\geq n_{2}^{2}+3n_{2}+n_{1}+2 (27)

as n2β‰€βŒŠβ„“βˆ’32βŒ‹n_{2}\leq\lfloor\frac{\ell-3}{2}\rfloor and n1+n2β‰€β„“βˆ’3n_{1}+n_{2}\leq\ell-3. On the other hand, nβˆ’2=βˆ‘i=1tni≀n1+(tβˆ’1)​n2n-2=\sum_{i=1}^{t}n_{i}\leq n_{1}+(t-1)n_{2}, which implies that tβ‰₯nβˆ’2βˆ’n1n2+1β‰₯n2+4t\geq\frac{n-2-n_{1}}{n_{2}}+1\geq n_{2}+4 by (27).

We first claim that n1+n2=β„“βˆ’3n_{1}+n_{2}=\ell-3. Suppose to the contrary that n1+n2β‰€β„“βˆ’4n_{1}+n_{2}\leq\ell-4. Let Hβ€²H^{\prime} be an (n1,nt)(n_{1},n_{t})-transformation of G​[R]G[R]. Clearly, n1​(Hβ€²)=n1+1β‰€β„“βˆ’3βˆ’n2n_{1}(H^{\prime})=n_{1}+1\leq\ell-3-n_{2} and n2​(Hβ€²)=n2n_{2}(H^{\prime})=n_{2}. Then, n1​(Hβ€²)+n2​(Hβ€²)β‰€β„“βˆ’3n_{1}(H^{\prime})+n_{2}(H^{\prime})\leq\ell-3. By Claim 3.5, K2+Hβ€²K_{2}+H^{\prime} is Cβ„“C_{\ell}-free. However, by (26) and Lemma 3.2, we have ρ​(K2+Hβ€²)>ρ\rho(K_{2}+H^{\prime})>\rho, contradicting that GG is extremal to spex𝒫​(n,Cβ„“){\rm spex}_{\mathcal{P}}(n,C_{\ell}). Hence, the claim holds.

Secondly, we claim that ni=n2n_{i}=n_{2} for i∈{3,…,tβˆ’1}i\in\{3,\dots,t-1\}. Suppose to the contrary, then set i0=min⁑{i|3≀i≀tβˆ’1,ni≀n2βˆ’1}i_{0}=\min\{i~{}|~{}3\leq i\leq t-1,n_{i}\leq n_{2}-1\}. Let Hβ€²H^{\prime} be an (ni0,nt)(n_{i_{0}},n_{t})-transformation of G​[R]G[R]. Clearly, n1​(Hβ€²)=n1n_{1}(H^{\prime})=n_{1} and n2​(Hβ€²)=max⁑{n2,ni0+1}=n2n_{2}(H^{\prime})=\max\{n_{2},n_{i_{0}}+1\}=n_{2}, and so n1​(Hβ€²)+n2​(Hβ€²)=β„“βˆ’3n_{1}(H^{\prime})+n_{2}(H^{\prime})=\ell-3. By Claim 3.5, K2+Hβ€²K_{2}+H^{\prime} is Cβ„“C_{\ell}-free. However, by (26) and Lemma 3.2, we have ρ​(K2+Hβ€²)>ρ\rho(K_{2}+H^{\prime})>\rho, contradicting that GG is extremal to spex𝒫​(n,Cβ„“){\rm spex}_{\mathcal{P}}(n,C_{\ell}). Hence, the claim holds. It follows that G=K2+(Pn1βˆͺ(tβˆ’2)​Pn2βˆͺPnt)G=K_{2}+(P_{n_{1}}\cup(t-2)P_{n_{2}}\cup P_{n_{t}}).

Finally, we claim that n2=βŒŠβ„“βˆ’32βŒ‹n_{2}=\lfloor\frac{\ell-3}{2}\rfloor. Note that 1≀n2β‰€βŒŠβ„“βˆ’32βŒ‹1\leq n_{2}\leq\lfloor\frac{\ell-3}{2}\rfloor. Then, n2=βŒŠβ„“βˆ’32βŒ‹n_{2}=\lfloor\frac{\ell-3}{2}\rfloor for β„“βˆˆ{5,6}\ell\in\{5,6\}. It remains the case β„“β‰₯7\ell\geq 7. Suppose to the contrary, then n2β‰€βŒŠβ„“βˆ’52βŒ‹n_{2}\leq\lfloor\frac{\ell-5}{2}\rfloor as n2β‰€βŒŠβ„“βˆ’32βŒ‹n_{2}\leq\lfloor\frac{\ell-3}{2}\rfloor. We use PiP^{i} to denote the ii-th longest path of G​[R]G[R] for i∈{1,…,t}i\in\{1,\dots,t\}. Recall that tβ‰₯n2+4t\geq n_{2}+4. Then P2,P3,…,Pn2+3P^{2},P^{3},\dots,P^{n_{2}+3} are paths of order n2n_{2}. We may assume that Pn2+3=w1​w2​…​wn2P^{n_{2}+3}=w_{1}w_{2}\dots w_{n_{2}}. Since n1=β„“βˆ’3βˆ’n2β‰₯βŒˆβ„“βˆ’32βŒ‰β‰₯2n_{1}=\ell-3-n_{2}\geq\lceil\frac{\ell-3}{2}\rceil\geq 2, there exists an endpoint wβ€²w^{\prime} of P1P^{1} with w′​wβ€²β€²βˆˆE​(P1)w^{\prime}w^{\prime\prime}\in E(P^{1}). Let Gβ€²G^{\prime} be obtained from GG by (i) deleting w′​wβ€²β€²w^{\prime}w^{\prime\prime} and joining wβ€²w^{\prime} to an endpoint of Pn2+2P^{n_{2}+2}; (ii) deleting all edges of Pn2+3P^{n_{2}+3} and joining wiw_{i} to an endpoint of Pi+1P^{i+1} for each i∈{1,…,n2}i\in\{1,\dots,n_{2}\}. Then, Gβ€²G^{\prime} is obtained from GG by deleting n2n_{2} edges and adding n2+1n_{2}+1 edges. By Claim 3.2, we obtain

4ρ2<xui​xuj<4ρ2+24ρ3+36ρ4<4ρ2+25ρ3\frac{4}{\rho^{2}}<x_{u_{i}}x_{u_{j}}<\frac{4}{\rho^{2}}+\frac{24}{\rho^{3}}+\frac{36}{\rho^{4}}<\frac{4}{\rho^{2}}+\frac{25}{\rho^{3}}

for any vertices ui,uj∈Ru_{i},u_{j}\in R. Then

ρ​(Gβ€²)βˆ’Οβ‰₯XT​(A​(Gβ€²)βˆ’A​(G))​XXT​X>2XT​X​(4​(n2+1)ρ2βˆ’4​n2ρ2βˆ’25​n2ρ3)>0,\rho(G^{\prime})-\rho\geq\frac{X^{\mathrm{T}}(A(G^{\prime})-A(G))X}{X^{\mathrm{T}}X}>\frac{2}{X^{\mathrm{T}}X}\left(\frac{4(n_{2}+1)}{\rho^{2}}-\frac{4n_{2}}{\rho^{2}}-\frac{25n_{2}}{\rho^{3}}\right)>0,

where n2<βŒŠβ„“βˆ’32βŒ‹β‰€425​2​nβˆ’4≀425​ρn_{2}<\lfloor\frac{\ell-3}{2}\rfloor\leq\frac{4}{25}\sqrt{2n-4}\leq\frac{4}{25}\rho as nβ‰₯62532β€‹βŒŠβ„“βˆ’32βŒ‹2+2n\geq\frac{625}{32}\lfloor\frac{\ell-3}{2}\rfloor^{2}+2. So, ρ​(Gβ€²)>ρ\rho(G^{\prime})>\rho. On the other hand, Gβ€²β‰…K2+(Pn1βˆ’1βˆͺ(n2+1)​Pn2+1βˆͺ(tβˆ’n2βˆ’4)​Pn2βˆͺPnt)G^{\prime}\cong K_{2}+(P_{n_{1}-1}\cup(n_{2}+1)P_{n_{2}+1}\cup(t-n_{2}-4)P_{n_{2}}\cup P_{n_{t}}). By Claim 3.5, Gβ€²G^{\prime} is Cβ„“C_{\ell}-free, contradicting that GG is extremal to spex𝒫​(n,Cβ„“){\rm spex}_{\mathcal{P}}(n,C_{\ell}). So, the claim holds.

Since n1+n2=β„“βˆ’3n_{1}+n_{2}=\ell-3 and n2=βŒŠβ„“βˆ’32βŒ‹n_{2}=\lfloor\frac{\ell-3}{2}\rfloor, we have n1=βŒˆβ„“βˆ’32βŒ‰n_{1}=\lceil\frac{\ell-3}{2}\rceil. Moreover, since ni=βŒŠβ„“βˆ’32βŒ‹n_{i}=\lfloor\frac{\ell-3}{2}\rfloor for i∈{2,…,tβˆ’1}i\in\{2,\dots,t-1\} and nt≀n2n_{t}\leq n_{2}, we can further obtain that G​[R]β‰…H​(βŒˆβ„“βˆ’32βŒ‰,βŒŠβ„“βˆ’32βŒ‹)G[R]\cong H(\lceil\frac{\ell-3}{2}\rceil,\lfloor\frac{\ell-3}{2}\rfloor), which implies that Gβ‰…K2+H​(βŒˆβ„“βˆ’32βŒ‰,βŒŠβ„“βˆ’32βŒ‹)G\cong K_{2}+H(\lceil\frac{\ell-3}{2}\rceil,\lfloor\frac{\ell-3}{2}\rfloor). 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 β–‘\Box

4 Proof of Theorem 1.4

Above all, we shall introduce the Jordan Curve Theorem: any simple closed curve CC 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 CC. We denote them by i​n​t​(C)int(C) and e​x​t​(C)ext(C), and their closures by I​n​t​(C)Int(C) and E​x​t​(C)Ext(C), respectively. A plane graph is a planar embedding of a planar graph. The Jordan Curve Theorem gives the following lemma.

Lemma 4.1.

Let CC be a cycle of a plane graph GG, and let x,yx,y be two vertices of GG with x∈i​n​t​(C)x\in int(C) and y∈e​x​t​(C)y\in ext(C), then x​yβˆ‰E​(G)xy\notin E(G).

Let GG be a plane graph. A face in GG of size ii is called an ii-face. Let fi​(G)f_{i}(G) denote the number of ii-faces in GG, and let f​(G)f(G) denote βˆ‘ifi​(G)\sum_{i}f_{i}(G).

Lemma 4.2.

(Proposition 2.5 of [1], P. 250) Let GG be a planar graph, and let ff be an arbitrary face in some planar embedding of GG. Then GG admits a planar embedding whose outer face has the same boundary as ff.

Let δ​(G)\delta(G) be the minimum degree of a graph GG. It is well known that every graph GG with δ​(G)β‰₯2\delta(G)\geq 2 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 GG be a plane graph on nn vertices with δ​(G)β‰₯3\delta(G)\geq 3. Then GG contains two vertex-disjoint cycles unless G∈{2​K1+C3,K1+Cnβˆ’1}G\in\{2K_{1}+C_{3},K_{1}+C_{n-1}\}.

Proof.

We first deal with some trivial cases. Since δ​(G)β‰₯3\delta(G)\geq 3, we have nβ‰₯1+δ​(G)β‰₯4n\geq 1+\delta(G)\geq 4. If n=4n=4, then Gβ‰…K1+C3G\cong K_{1}+C_{3}. If n=5n=5, then 2​e​(G)=βˆ‘v∈V​(G)dG​(v)β‰₯3Γ—5=152e(G)=\sum_{v\in V(G)}d_{G}(v)\geq 3\times 5=15, and so e​(G)β‰₯8e(G)\geq 8. On the other hand, e​(G)≀3​nβˆ’6=9e(G)\leq 3n-6=9, since GG is planar. Thus, e​(G)∈{8,9}e(G)\in\{8,9\}. It is not hard to verify that Gβ‰…2​K1+C3G\cong 2K_{1}+C_{3} when e​(G)=9e(G)=9 and Gβ‰…K1+C4G\cong K_{1}+C_{4} when e​(G)e(G)=8, as desired. If GG is not connected, then GG contains at least two components G1G_{1} and G2G_{2} with δ​(Gi)β‰₯3\delta(G_{i})\geq 3 for i∈{1,2}i\in\{1,2\}, which implies that each GiG_{i} contains a cycle. Thus, GG contains two vertex-disjoint cycles, as desired. If GG has a cut vertex vv, then Gβˆ’{v}G-\{v\} has at least two components G3G_{3} and G4G_{4}. Since δ​(G)β‰₯3\delta(G)\geq 3, we have δ​(Gi)β‰₯2\delta(G_{i})\geq 2 for i∈{3,4}i\in\{3,4\}, which implies that both G3G_{3} and G4G_{4} contain a cycle. Thus, GG also contains two vertex-disjoint cycles.

Next, we only need to consider the case that GG is a 2-connected graph of order nβ‰₯6n\geq 6. Since GG is 2-connected, each face of GG is a cycle. Let CC be a face of GG with minimum size gg. By Lemma 4.2, we may assume without loss of generality that CC is the outer face of GG. Let G1=Gβˆ’V​(C)G_{1}=G-V(C). If G1G_{1} contains a cycle, then GG contains two vertex-disjoint cycles, as desired. Now assume that G1G_{1} is acyclic. Since δ​(G)β‰₯3\delta(G)\geq 3, we have 2​e​(G)=βˆ‘v∈V​(G)dG​(v)β‰₯3​n2e(G)=\sum_{v\in V(G)}d_{G}(v)\geq 3n. This, together with Euler’s formula nβˆ’2=e​(G)βˆ’f​(G)n-2=e(G)-f(G), gives e​(G)≀3​f​(G)βˆ’6e(G)\leq 3f(G)-6. On the other hand,

2​e​(G)=βˆ‘iβ‰₯gi​fi​(G)β‰₯gβ€‹βˆ‘iβ‰₯gfi​(G)=g​f​(G).\displaystyle 2e(G)=\sum_{i\geq g}if_{i}(G)\geq g\sum_{i\geq g}f_{i}(G)=gf(G).

Hence, g​f​(G)≀2​e​(G)≀6​f​(G)βˆ’12gf(G)\leq 2e(G)\leq 6f(G)-12, yielding g≀6​f​(G)βˆ’12f​(G)<6g\leq\frac{6f(G)-12}{f(G)}<6. Subsequently, we shall give several claims.

Claim 4.1.

We have g=3g=3.

Refer to caption
Figure 2: Two possible local structures of GG.
Proof.

Suppose to the contrary that g∈{4,5}g\in\{4,5\}, and let C=v1​v2​…​vg​v1C=v_{1}v_{2}\ldots v_{g}v_{1}. We first consider the case that there exists a vertex of G1G_{1} adjacent to two consecutive vertices of CC. Without loss of generality, let w1∈V​(G1)w_{1}\in V(G_{1}) and {w1,v1,v2}\{w_{1},v_{1},v_{2}\} induces a triangle Cβ€²C^{\prime}. More generally, we define A={w∈V​(G1)|v1,v2∈NC​(w)}A=\{w\in V(G_{1})~{}|~{}v_{1},v_{2}\in N_{C}(w)\}. Clearly, w1∈Aw_{1}\in A. We can select a vertex, say w1w_{1}, in AA such that AβŠ†E​x​t​(Cβ€²)A\subseteq Ext(C^{\prime}) (see Figure 2). Notice that Cβ€²C^{\prime} is not a face of GG, as g∈{4,5}g\in\{4,5\}. Then, i​n​t​(Cβ€²)β‰ βˆ…int(C^{\prime})\neq\varnothing. By Lemma 4.1, every vertex in i​n​t​(Cβ€²)int(C^{\prime}) has no neighbors in e​x​t​(Cβ€²)ext(C^{\prime}). Moreover, by the definitions of AA and w1w_{1}, every vertex in i​n​t​(Cβ€²)int(C^{\prime}) has at most one neighbor in {v1,v2}\{v_{1},v_{2}\}. It follows that every vertex in i​n​t​(Cβ€²)int(C^{\prime}) has at least one neighbor in i​n​t​(Cβ€²)int(C^{\prime}), as δ​(G)β‰₯3\delta(G)\geq 3. Thus, G​[i​n​t​(Cβ€²)]G[int(C^{\prime})] is nonempty, that is, G1​[i​n​t​(Cβ€²)]G_{1}[int(C^{\prime})] is nonempty. Recall that G1G_{1} is acyclic. Then G1​[i​n​t​(Cβ€²)]G_{1}[int(C^{\prime})] contains at least two pendant vertices, one of which (say w2w_{2}) is not adjacent to w1w_{1}. Hence, w2w_{2} is also a pendant vertex of G1G_{1}, as w2w_{2} has no neighbors in e​x​t​(Cβ€²)ext(C^{\prime}). On the other hand, w2w_{2} has at most one neighbor in {v1,v2}\{v_{1},v_{2}\}, and so dC​(w2)≀1d_{C}(w_{2})\leq 1. Therefore, dG​(w2)=dG1​(w2)+dC​(w2)≀2d_{G}(w_{2})=d_{G_{1}}(w_{2})+d_{C}(w_{2})\leq 2, contradicting δ​(G)β‰₯3\delta(G)\geq 3.

Now it remains the case that each vertex of G1G_{1} is not adjacent to two consecutive vertices of CC. Note that δ​(G)β‰₯3\delta(G)\geq 3 and G1G_{1} is acyclic. Then G1G_{1} contains a vertex w0w_{0} with dG1​(w0)≀1d_{G_{1}}(w_{0})\leq 1, and thus dC​(w0)=dG​(w0)βˆ’dG1​(w0)β‰₯2d_{C}(w_{0})=d_{G}(w_{0})-d_{G_{1}}(w_{0})\geq 2. Now, since g∈{4,5}g\in\{4,5\}, we may assume without loss of generality that v1,v3∈NC​(w0)v_{1},v_{3}\in N_{C}(w_{0}). Let Aβ€²={w∈V​(G1)|v1,v3∈NC​(w)}A^{\prime}=\{w\in V(G_{1})~{}|~{}v_{1},v_{3}\in N_{C}(w)\}. Clearly, w0∈Aβ€²w_{0}\in A^{\prime} and v2βˆ‰NC​(w)v_{2}\notin N_{C}(w) for each w∈Aβ€²w\in A^{\prime}. Now, we can select a vertex, say w1w_{1}, in Aβ€²A^{\prime} such that Aβ€²βŠ†E​x​t​(Cβ€²β€²)A^{\prime}\subseteq Ext(C^{\prime\prime}), where Cβ€²β€²=w1​v1​v2​v3​w1C^{\prime\prime}=w_{1}v_{1}v_{2}v_{3}w_{1} (see Figure 2). We can see that i​n​t​(Cβ€²β€²)β‰ βˆ…int(C^{\prime\prime})\neq\varnothing (otherwise, dG​(v2)=|{v1,v3}|=2d_{G}(v_{2})=|\{v_{1},v_{3}\}|=2, a contradiction). By the definition of w1w_{1}, we have i​n​t​(Cβ€²β€²)∩Aβ€²=βˆ…int(C^{\prime\prime})\cap A^{\prime}=\varnothing. Furthermore, every vertex in i​n​t​(Cβ€²β€²)int(C^{\prime\prime}) has no neighbors in e​x​t​(Cβ€²β€²)ext(C^{\prime\prime}) and has at most one neighbor in {v1,v2,v3}.\{v_{1},v_{2},v_{3}\}. Thus, every vertex in i​n​t​(Cβ€²β€²)int(C^{\prime\prime}) has at least one neighbor in i​n​t​(Cβ€²β€²)int(C^{\prime\prime}). By a similar argument as above, we can find a vertex w2∈i​n​t​(Cβ€²β€²)w_{2}\in int(C^{\prime\prime}) with dG​(w2)=dG1​(w2)+dC​(w2)≀2d_{G}(w_{2})=d_{G_{1}}(w_{2})+d_{C}(w_{2})\leq 2, which contradicts δ​(G)β‰₯3\delta(G)\geq 3. ∎

By Claim 4.1, the outer face of GG is a triangle C=v1​v2​v3​v1C=v_{1}v_{2}v_{3}v_{1}. In the following, we denote Bi={w∈V​(G1)|dC​(w)=i}B_{i}=\{w\in V(G_{1})~{}|~{}d_{C}(w)=i\} for i≀3i\leq 3. Since δ​(G)β‰₯3\delta(G)\geq 3, we have w∈B3w\in B_{3} for each isolated vertex ww of G1G_{1}, and w∈B2βˆͺB3w\in B_{2}\cup B_{3} for each pendant vertex ww of G1G_{1}.

Claim 4.2.

|B3|≀1|B_{3}|\leq 1 and |B2|+|B3|β‰₯2|B_{2}|+|B_{3}|\geq 2.

Proof.

Since CC is the outer face of GG, every vertex of G1G_{1} lies in i​n​t​(C)int(C). Furthermore, since GG is planar, it is easy to see that |B3|≀1|B_{3}|\leq 1. This implies that G1G_{1} contains at most one isolated vertex. Recall that |G1|=nβˆ’3β‰₯3|G_{1}|=n-3\geq 3 and G1G_{1} is acyclic. Then G1G_{1} contains at least two pendant vertices w1w_{1} and w2w_{2}. Therefore, |B2|+|B3|β‰₯|{w1,w2}|=2|B_{2}|+|B_{3}|\geq|\{w_{1},w_{2}\}|=2. ∎

Claim 4.3.

Let w0,w1w_{0},w_{1} be two vertices in V​(G1)V(G_{1}) such that NC​(w0)βŠ‡{v3}N_{C}(w_{0})\supseteq\{v_{3}\} and NC​(w1)βŠ‡{v1,v2}N_{C}(w_{1})\supseteq\{v_{1},v_{2}\} (see Figure 3). Then
(i) v3,w0∈e​x​t​(Cβ€²β€²β€²)v_{3},w_{0}\in ext(C^{\prime\prime\prime}), where Cβ€²β€²β€²=w1​v1​v2​w1C^{\prime\prime\prime}=w_{1}v_{1}v_{2}w_{1};
(ii) if w0βˆ‰B3w_{0}\notin B_{3}, then G1G_{1} contains a pendant vertex in e​x​t​(Cβ€²β€²β€²)ext(C^{\prime\prime\prime}).

Refer to caption
Figure 3: A local structure in Claim 4.3.
Proof.

(i) Since CC is the outer face and v3∈V​(C)βˆ–V​(Cβ€²β€²β€²)v_{3}\in V(C)\setminus V(C^{\prime\prime\prime}), we have v3∈e​x​t​(Cβ€²β€²β€²)v_{3}\in ext(C^{\prime\prime\prime}). Furthermore, using w0​v3∈E​(G)w_{0}v_{3}\in E(G) and Lemma 4.1 gives w0∈e​x​t​(Cβ€²β€²β€²)w_{0}\in ext(C^{\prime\prime\prime}).

(ii) Since w0βˆ‰B3w_{0}\notin B_{3}, we have dC​(w0)≀2d_{C}(w_{0})\leq 2, and so dG1​(w0)=dG​(w0)βˆ’dC​(w0)β‰₯1d_{G_{1}}(w_{0})=d_{G}(w_{0})-d_{C}(w_{0})\geq 1. By (i), we know that w0∈e​x​t​(Cβ€²β€²β€²)w_{0}\in ext(C^{\prime\prime\prime}). If dG1​(w0)=1d_{G_{1}}(w_{0})=1, then w0w_{0} is a desired pendant vertex. It remains the case that dG1​(w0)β‰₯2d_{G_{1}}(w_{0})\geq 2. Now, whether w1w_{1} is a neighbor of w0w_{0} or not, w0w_{0} has at least one neighbor in V​(G1)∩e​x​t​(Cβ€²β€²β€²)V(G_{1})\cap ext(C^{\prime\prime\prime}). Thus, G1​[e​x​t​(Cβ€²β€²β€²)]G_{1}[ext(C^{\prime\prime\prime})] is nonempty. Recall that G1G_{1} is acyclic. Then G1​[e​x​t​(Cβ€²β€²β€²)]G_{1}[ext(C^{\prime\prime\prime})] contains at least two pendant vertices, one of which (say w2w_{2}) is not adjacent to w1w_{1}. Hence, w2w_{2} is also a pendant vertex of G1G_{1}, as w2w_{2} has no neighbors in i​n​t​(Cβ€²β€²β€²)int(C^{\prime\prime\prime}). ∎

Refer to caption
Figure 4: Two possible local structures in Claim 4.4.
Claim 4.4.

Let w1,w2∈V​(G1)w_{1},w_{2}\in V(G_{1}) with NC​(w1)∩NC​(w2)βŠ‡{v1,v2}N_{C}(w_{1})\cap N_{C}(w_{2})\supseteq\{v_{1},v_{2}\}. Assume that Cβ€²β€²β€²=w1​v1​v2​w1C^{\prime\prime\prime}=w_{1}v_{1}v_{2}w_{1} and w2∈i​n​t​(Cβ€²β€²β€²)w_{2}\in int(C^{\prime\prime\prime}) (see Figure 4). Then GG contains a cycle CviC_{v_{i}} such that V​(Cvi)βŠ†I​n​t​(Cβ€²β€²β€²)V(C_{v_{i}})\subseteq Int(C^{\prime\prime\prime}) and V​(Cvi)∩V​(C)={vi}V(C_{v_{i}})\cap V(C)=\{v_{i}\} for each i∈{1,2}i\in\{1,2\}.

Proof.

We first claim that NC​(w2)={v1,v2}N_{C}(w_{2})=\{v_{1},v_{2}\}. By Claim 4.3, we know that v3∈e​x​t​(Cβ€²β€²β€²)v_{3}\in ext(C^{\prime\prime\prime}). Now, since w2∈i​n​t​(Cβ€²β€²β€²)w_{2}\in int(C^{\prime\prime\prime}), we have w2​v3βˆ‰E​(G)w_{2}v_{3}\notin E(G) by Lemma 4.1, and so NC​(w2)={v1,v2}N_{C}(w_{2})=\{v_{1},v_{2}\}. Furthermore, we have dG1​(w2)β‰₯1d_{G_{1}}(w_{2})\geq 1. Then G1G_{1} contains a path PP with endpoints w2w_{2} and w3w_{3}, where w3w_{3} is a pendant vertex of G1G_{1}. If V​(P)⊈i​n​t​(Cβ€²β€²β€²)V(P)\not\subseteq int(C^{\prime\prime\prime}), then by w2∈i​n​t​(Cβ€²β€²β€²)w_{2}\in int(C^{\prime\prime\prime}) and Lemma 4.1, we have V​(P)∩V​(Cβ€²β€²β€²)={w1}V(P)\cap V(C^{\prime\prime\prime})=\{w_{1}\} as v1,v2βˆ‰V​(G1)v_{1},v_{2}\notin V(G_{1}). Now let Pβ€²P^{\prime} be the subpath of PP with endpoints w2w_{2} and w1w_{1}. Then V​(Pβ€²)βˆ–{w1}βŠ†i​n​t​(Cβ€²β€²β€²)V(P^{\prime})\setminus\{w_{1}\}\subseteq int(C^{\prime\prime\prime}), and GG contains a cycle C​(vi)=vi​w1​P′​w2​viC(v_{i})=v_{i}w_{1}P^{\prime}w_{2}v_{i} for each i∈{1,2}i\in\{1,2\}, as desired. Next, assume that V​(P)βŠ†i​n​t​(Cβ€²β€²β€²)V(P)\subseteq int(C^{\prime\prime\prime}). Then, w3∈i​n​t​(Cβ€²β€²β€²)w_{3}\in int(C^{\prime\prime\prime}). By v3∈e​x​t​(Cβ€²β€²β€²)v_{3}\in ext(C^{\prime\prime\prime}) and Lemma 4.1, we get that w3​v3βˆ‰E​(G)w_{3}v_{3}\notin E(G), and so w3βˆ‰B3w_{3}\notin B_{3}. Moreover, dG1​(w3)=1d_{G_{1}}(w_{3})=1 and δ​(G)β‰₯3\delta(G)\geq 3 give w3∈B2w_{3}\in B_{2}. Thus, NC​(w3)={v1,v2}N_{C}(w_{3})=\{v_{1},v_{2}\}. Therefore, GG contains a cycle Cvi=vi​w2​P​w3​viC_{v_{i}}=v_{i}w_{2}Pw_{3}v_{i} for each i∈{1,2}i\in\{1,2\}, as desired. ∎

Having above four claims, we are ready to give the final proof of Lemma 4.3. By Claim 4.2, we have |B3|≀1|B_{3}|\leq 1 and |B2|β‰₯1|B_{2}|\geq 1. We may without loss of generality that w1∈B2w_{1}\in B_{2} and NC​(w1)={v1,v2}N_{C}(w_{1})=\{v_{1},v_{2}\}. For each i∈{1,2}i\in\{1,2\}, let i¯∈{1,2}βˆ–{i}\overline{i}\in\{1,2\}\setminus\{i\}. Since dC​(w1)=2d_{C}(w_{1})=2, we have dG1​(w1)β‰₯1d_{G_{1}}(w_{1})\geq 1. Hence, G1G_{1} is nonempty, and so G1G_{1} contains at least two pendant vertices. According to the size of B3B_{3}, we now distinguish two cases to complete the proof.

Case 1. |B3|=1|B_{3}|=1.

Assume that B3={w0}B_{3}=\{w_{0}\}. Then NC​(w0)={v1,v2,v3}N_{C}(w_{0})=\{v_{1},v_{2},v_{3}\} (see Figure 5). We then consider two subcases according to the size of B2B_{2}.

Refer to caption
Figure 5: Three possible structures in Case 1.

Subcase 1.1. |B2|=1|B_{2}|=1, that is, B2={w1}B_{2}=\{w_{1}\}.

For each pendant vertex ww of G1G_{1}, we have dC​(w)=dG​(w)βˆ’dG1​(w)β‰₯2d_{C}(w)=d_{G}(w)-d_{G_{1}}(w)\geq 2, consequently, w∈B2βˆͺB3={w1,w0}w\in B_{2}\cup B_{3}=\{w_{1},w_{0}\}. This indicates that G1G_{1} contains exactly two pendant vertices w1w_{1} and w0w_{0}. Furthermore, we can see that G1G_{1} contains no isolated vertices (otherwise, every isolated vertex of G1G_{1} has at least three neighbors in V​(C)V(C) and so belongs to B3B_{3}, while the unique vertex w0∈B3w_{0}\in B_{3} is a pendant vertex of G1G_{1}). Therefore, G1G_{1} is a path of order nβˆ’|C|n-|C| with endpoints w1w_{1} and w0w_{0}.

Now we know that G1G_{1} is a path with |G1|=nβˆ’3β‰₯3|G_{1}|=n-3\geq 3. Let NG1​(w0)={w2}N_{G_{1}}(w_{0})=\{w_{2}\} and Pβ€²β€²=G1βˆ’{w0}P^{\prime\prime}=G_{1}-\{w_{0}\}. Then Pβ€²β€²P^{\prime\prime} is a path with endpoints w1w_{1} and w2w_{2}. Since dG1​(w2)=2d_{G_{1}}(w_{2})=2, we have dC​(w2)β‰₯1d_{C}(w_{2})\geq 1. If w2​v3∈E​(G)w_{2}v_{3}\in E(G), then GG contains two vertex-disjoint cycles v3​w0​w2​v3v_{3}w_{0}w_{2}v_{3} and w1​v1​v2​w1w_{1}v_{1}v_{2}w_{1}, as desired. If w2​vi∈E​(G)w_{2}v_{i}\in E(G) for some i∈{1,2}i\in\{1,2\}, then GG contains two vertex-disjoint cycles vi​w1​P′′​w2​viv_{i}w_{1}P^{\prime\prime}w_{2}v_{i} and w0​vi¯​v3​w0w_{0}v_{\overline{i}}v_{3}w_{0}, as desired.

Subcase 1.2. |B2|β‰₯2|B_{2}|\geq 2.

Let w2∈B2βˆ–{w1}w_{2}\in B_{2}\setminus\{w_{1}\}. If NC​(w1)=NC​(w2)N_{C}(w_{1})=N_{C}(w_{2}), then we may assume that w2∈i​n​t​(Cβ€²β€²β€²)w_{2}\in int(C^{\prime\prime\prime}) by the symmetry of w1w_{1} and w2w_{2}, where Cβ€²β€²β€²=w1​v1​v2​w1C^{\prime\prime\prime}=w_{1}v_{1}v_{2}w_{1}. By Claim 4.4, GG contains a cycle Cv1C_{v_{1}} such that V​(Cv1)βŠ†I​n​t​(Cβ€²β€²β€²)V(C_{v_{1}})\subseteq Int(C^{\prime\prime\prime}) and V​(Cv1)∩V​(C)={v1}V(C_{v_{1}})\cap V(C)=\{v_{1}\}. On the other hand, Claim 4.3 implies that w0∈e​x​t​(Cβ€²β€²β€²)w_{0}\in ext(C^{\prime\prime\prime}). Hence, w0βˆ‰V​(Cv1)w_{0}\notin V(C_{v_{1}}). Therefore, GG contains two vertex-disjoint cycles Cv1C_{v_{1}} and w0​v2​v3​w0w_{0}v_{2}v_{3}w_{0}, as desired.

It remains the case that NC​(w1)β‰ NC​(w2)N_{C}(w_{1})\neq N_{C}(w_{2}). Now NC​(w2)={vi,v3}N_{C}(w_{2})=\{v_{i},v_{3}\} for some i∈{1,2}i\in\{1,2\}. We define Cβ€²β€²β€²=w0​v1​v2​w0C^{\prime\prime\prime}=w_{0}v_{1}v_{2}w_{0} instead of the original one in Claim 4.4. Then w1∈i​n​t​(Cβ€²β€²β€²)w_{1}\in int(C^{\prime\prime\prime}). Moreover, w2∈e​x​t​(Cβ€²β€²β€²)w_{2}\in ext(C^{\prime\prime\prime}) as w2​v3∈E​(G)w_{2}v_{3}\in E(G). By Claim 4.4, there exists a cycle CviΒ―C_{v_{\overline{i}}} such that V​(CviΒ―)βŠ†I​n​t​(Cβ€²β€²β€²)V(C_{v_{\overline{i}}})\subseteq Int(C^{\prime\prime\prime}) and V​(CviΒ―)∩V​(C)={viΒ―}V(C_{v_{\overline{i}}})\cap V(C)=\{v_{\overline{i}}\}. Therefore, GG contains two vertex-disjoint cycles CviΒ―C_{v_{\overline{i}}} and w2​vi​v3​w2w_{2}v_{i}v_{3}w_{2}, as desired.

Case 2. |B3|=0|B_{3}|=0.

Recall that A={w∈V​(G1)|v1,v2∈NC​(w)}A=\{w\in V(G_{1})~{}|~{}v_{1},v_{2}\in N_{C}(w)\}. Since |B3|=0|B_{3}|=0, we can see that NC​(w)=NC​(w1)={v1,v2}N_{C}(w)=N_{C}(w_{1})=\{v_{1},v_{2}\} for each w∈Aw\in A. We may assume without loss of generality that AβŠ†I​n​t​(Cβ€²β€²β€²)A\subseteq Int(C^{\prime\prime\prime}) by the symmetry of vertices in AA. By Claim 4.3, there exists a pendant vertex w3w_{3} of G1G_{1} in e​x​t​(Cβ€²β€²β€²)ext(C^{\prime\prime\prime}), which implies that dC​(w3)β‰₯2d_{C}(w_{3})\geq 2. Since |B3|=0|B_{3}|=0, we have w3∈B2w_{3}\in B_{2}, and thus B2βŠ‡{w1,w3}B_{2}\supseteq\{w_{1},w_{3}\}. Moreover, w3βˆ‰Aw_{3}\notin A as AβŠ†I​n​t​(Cβ€²β€²β€²)A\subseteq Int(C^{\prime\prime\prime}). Assume without loss of generality that NC​(w3)={v1,v3}N_{C}(w_{3})=\{v_{1},v_{3}\} (see Figure 6). We also consider two subcases according to |B2||B_{2}|.

Refer to caption
Figure 6: Three possible structures in Case 2.

Subcase 2.1. |B2|=2|B_{2}|=2, that is, B2={w1,w3}B_{2}=\{w_{1},w_{3}\}.

Since dC​(w3)=2d_{C}(w_{3})=2, we have dG1​(w3)β‰₯1d_{G_{1}}(w_{3})\geq 1, which implies that G1G_{1} is non-empty and has at least two pendant vertices. On the other hand, since δ​(G)β‰₯3\delta(G)\geq 3 while B3=βˆ…B_{3}=\varnothing, we can see that G1G_{1} contains no isolated vertices, and w∈B2={w1,w3}w\in B_{2}=\{w_{1},w_{3}\} for each pendant vertex ww of G1G_{1}. Therefore, G1G_{1} contains exactly two pendant vertices w1w_{1} and w3w_{3}, more precisely, G1G_{1} is a path with endpoints w1w_{1} and w3w_{3}. Let ww be an arbitrary vertex in V​(G1)βˆ–{w1,w3}V(G_{1})\setminus\{w_{1},w_{3}\}. Then, dC​(w)=dG​(w)βˆ’dG1​(w)=dG​(w)βˆ’2β‰₯1.d_{C}(w)=d_{G}(w)-d_{G_{1}}(w)=d_{G}(w)-2\geq 1.

If w​v2∈E​(G)wv_{2}\in E(G), then GG contains two vertex-disjoint cycles v2​w1​P′′′​w​v2v_{2}w_{1}P^{\prime\prime\prime}wv_{2} and v1​w3​v3​v1v_{1}w_{3}v_{3}v_{1}, where Pβ€²β€²β€²P^{\prime\prime\prime} is the subpath of G1G_{1} from w1w_{1} to ww (see Figure 6(aa)). If w​v3∈E​(G)wv_{3}\in E(G), then GG contains two vertex-disjoint cycles v3​w3​P′′′​w​v3v_{3}w_{3}P^{\prime\prime\prime}wv_{3} and v1​w1​v2​v1v_{1}w_{1}v_{2}v_{1}, where Pβ€²β€²β€²P^{\prime\prime\prime} is the subpath of G1G_{1} from w3w_{3} to ww (see Figure 6(aa)). If NC​(w)={v1}N_{C}(w)=\{v_{1}\} for each w∈V​(G1)βˆ–{w1,w3},w\in V(G_{1})\setminus\{w_{1},w_{3}\}, then Gβ‰…K1+Cnβˆ’1G\cong K_{1}+C_{n-1}, as desired.

Subcase 2.2. |B2|β‰₯3|B_{2}|\geq 3.

For each vertex w∈B2w\in B_{2}, it is clear that NC​(w)N_{C}(w) is one of {v1,v2}\{v_{1},v_{2}\}, {v1,v3}\{v_{1},v_{3}\} and {v2,v3}\{v_{2},v_{3}\}. We first consider the case that there exist two vertices in B2B_{2} which have the same neighbors in CC. Without loss of generality, assume that we can find a vertex w2∈B2w_{2}\in B_{2} with NC​(w2)=NC​(w1)={v1,v2}N_{C}(w_{2})=N_{C}(w_{1})=\{v_{1},v_{2}\}. Then w2∈Aw_{2}\in A. Recall that AβŠ†I​n​t​(Cβ€²β€²β€²)A\subseteq Int(C^{\prime\prime\prime}) and Cβ€²β€²β€²=w1​v1​v2​w1C^{\prime\prime\prime}=w_{1}v_{1}v_{2}w_{1} (see Figure 6(bb)). Then, we can further get that w2∈i​n​t​(Cβ€²β€²β€²)w_{2}\in int(C^{\prime\prime\prime}). By Claim 4.4, there exists a cycle Cv2C_{v_{2}} such that V​(Cv2)βŠ‚I​n​t​(Cβ€²β€²β€²)V(C_{v_{2}})\subset Int(C^{\prime\prime\prime}) and V​(Cv2)∩V​(C)={v2}V(C_{v_{2}})\cap V(C)=\{v_{2}\}. On the other hand, Claim 4.3 implies that w3∈e​x​t​(Cβ€²β€²β€²)w_{3}\in ext(C^{\prime\prime\prime}). Hence, w3βˆ‰V​(Cv2)w_{3}\notin V(C_{v_{2}}). Therefore, GG contains two vertex-disjoint cycles Cv2C_{v_{2}} and w3​v1​v3​w3w_{3}v_{1}v_{3}w_{3}, as desired.

Now it remains the case that any two vertices in B2B_{2} have different neighborhoods in CC. This implies that |B2|=3|B_{2}|=3 and we can find a vertex w2∈B2w_{2}\in B_{2} with NC​(w2)={v2,v3}N_{C}(w_{2})=\{v_{2},v_{3}\}. Now we have B2={w1,w2,w3}B_{2}=\{w_{1},w_{2},w_{3}\}. Furthermore, since δ​(G)β‰₯3\delta(G)\geq 3 and B3=βˆ…B_{3}=\varnothing, we have dG1​(w)β‰₯1d_{G_{1}}(w)\geq 1 for each w∈V​(G1)w\in V(G_{1}), and if dG1​(w)=1d_{G_{1}}(w)=1, then w∈B2w\in B_{2}. Since |B2|=3|B_{2}|=3, we can see that G1G_{1} has only one connected component, that is, G1G_{1} is a tree and some wiw_{i}, say w2w_{2}, is a pendant vertex of G1G_{1}. Now, G1βˆ’{w2}G_{1}-\{w_{2}\} contains a subpath Pβ€²β€²β€²β€²P^{\prime\prime\prime\prime} with endpoints w1w_{1} and w3w_{3} (see Figure 6(cc)). Then GG contains two vertex-disjoint cycles v1​w1​P′′′′​w3​v1v_{1}w_{1}P^{\prime\prime\prime\prime}w_{3}v_{1} and w2​v2​v3​w2w_{2}v_{2}v_{3}w_{2}, as desired.

This completes the proof of Lemma 4.3. ∎

Let 𝒒nβˆ—\mathcal{G}^{*}_{n} be the family of graphs obtained from 2​K1+C32K_{1}+C_{3} and an independent set of size nβˆ’5n-5 by joining each vertex of the independent set to arbitrary two vertices of the triangle (see Figure 7). Clearly, every graph in 𝒒nβˆ—\mathcal{G}^{*}_{n} is planar. Now, let 𝒒n\mathcal{G}_{n} be the family of planar graphs obtained from 2​K1+C32K_{1}+C_{3} by iteratively adding vertices of degree 2 until the resulting graph has nn vertices. Then 𝒒nβˆ—βŠ†π’’n\mathcal{G}^{*}_{n}\subseteq\mathcal{G}_{n}.

Refer to caption
Figure 7: An extremal graph in 𝒒nβˆ—\mathcal{G}^{*}_{n}.
Lemma 4.4.

For any graph Gβˆˆπ’’nG\in\mathcal{G}_{n}, GG is 2β€‹π’ž2\mathcal{C}-free if and only if Gβˆˆπ’’nβˆ—G\in\mathcal{G}^{*}_{n}.

Proof.

Let V1:={v1,v2,v3}V_{1}:=\{v_{1},v_{2},v_{3}\} be the set of vertices of degree 4 and V2:={w1,w2}V_{2}:=\{w_{1},w_{2}\} be the set of vertices of degree 3 in 2​K1+C32K_{1}+C_{3}, respectively. Then V1V_{1} induces a triangle. We first show that every graph GG in 𝒒nβˆ—\mathcal{G}^{*}_{n} is 2β€‹π’ž2\mathcal{C}-free. It suffices to prove that every cycle of GG contains at least two vertices in V1V_{1}. Let CC be an arbitrary cycle of GG. If V​(C)βŠ†V1V(C)\subseteq V_{1}, then there is nothing to prove. It remains the case that there exists a vertex w∈V​(C)βˆ–V1w\in V(C)\setminus V_{1}. By the definition of 𝒒nβˆ—\mathcal{G}^{*}_{n}, we can see that NC​(w)βŠ†NG​(w)βŠ†V1N_{C}(w)\subseteq N_{G}(w)\subseteq V_{1}. Note that |NC​(w)|β‰₯2|N_{C}(w)|\geq 2. Hence, CC contains at least two vertices in V1V_{1}.

In the following, we will show that every graph Gβˆˆπ’’nβˆ–π’’nβˆ—G\in\mathcal{G}_{n}\setminus\mathcal{G}^{*}_{n} contains two vertex-disjoint cycles. By the definition of 𝒒n\mathcal{G}_{n}, GG is obtained from 2​K1+C32K_{1}+C_{3} by iteratively adding nβˆ’5n-5 vertices u1,u2,…,unβˆ’5u_{1},u_{2},\dots,u_{n-5} of degree 2. Now, let Gnβˆ’5=GG_{n-5}=G, and Giβˆ’1=Giβˆ’{ui}G_{i-1}=G_{i}-\{u_{i}\} for i∈{1,2,…,nβˆ’5}i\in\{1,2,\ldots,n-5\}. Then G0β‰…2​K1+C3G_{0}\cong 2K_{1}+C_{3}. Moreover, |Gi|=i+5|G_{i}|=i+5 and dGi​(ui)=2d_{G_{i}}(u_{i})=2 for each i∈{1,2,…,nβˆ’5}i\in\{1,2,\dots,n-5\}. Now let

iβˆ—=max⁑{i|0≀i≀nβˆ’5,Giβˆˆπ’’i+5βˆ—}.i^{*}=\max\{i~{}|~{}0\leq i\leq n-5,~{}G_{i}\in\mathcal{G}^{*}_{i+5}\}.

Since G0=2​K1+C3βˆˆπ’’5βˆ—G_{0}=2K_{1}+C_{3}\in\mathcal{G}^{*}_{5} and Gnβˆ’5βˆ‰π’’nβˆ—G_{n-5}\notin\mathcal{G}^{*}_{n}, we have 0≀iβˆ—β‰€nβˆ’60\leq i^{*}\leq n-6. By the choice of iβˆ—i^{*}, we know that Giβˆ—βˆˆπ’’iβˆ—+5βˆ—G_{i^{*}}\in\mathcal{G}^{*}_{i^{*}+5} and Giβˆ—+1βˆ‰π’’iβˆ—+6βˆ—G_{i^{*}+1}\notin\mathcal{G}^{*}_{i^{*}+6}, which implies that NGiβˆ—β€‹(ui)βŠ†V1N_{G_{i^{*}}}(u_{i})\subseteq V_{1} and NGiβˆ—+1​(uiβˆ—+1)⊈V1N_{G_{i^{*}+1}}(u_{i^{*}+1})\not\subseteq V_{1}.

Refer to caption
Figure 8: An extremal graph in 𝒒nβˆ—\mathcal{G}^{*}_{n}.

Now we may assume that Gnβˆ’5G_{n-5} is a planar embedding of GG, and G0G_{0} is a plane subgraph of Gnβˆ’5G_{n-5}. Observe that 2​K1+C32K_{1}+C_{3} has six planar embeddings (see Figure 8). Without loss of generality, assume that G0G_{0} is the leftmost graph in Figure 8. Then, uiβˆ—+1u_{i^{*}+1} lies in one of the following regions (see Figure 8):

e​x​t​(w2​v1​v2​w2),i​n​t​(w2​v1​v3​w2),i​n​t​(w2​v2​v3​w2),ext(w_{2}v_{1}v_{2}w_{2}),int(w_{2}v_{1}v_{3}w_{2}),int(w_{2}v_{2}v_{3}w_{2}),
i​n​t​(w1​v1​v2​w1),i​n​t​(w1​v1​v3​w1),i​n​t​(w1​v2​v3​w1).int(w_{1}v_{1}v_{2}w_{1}),int(w_{1}v_{1}v_{3}w_{1}),int(w_{1}v_{2}v_{3}w_{1}).

By Lemma 4.2, we can assume that uiβˆ—+1u_{i^{*}+1} lies in the outer face, that is, uiβˆ—+1∈e​x​t​(w2​v1​v2​w2)u_{i^{*}+1}\in ext(w_{2}v_{1}v_{2}w_{2}). For simplify, we denote Cβ€²=w2​v1​v2​w2C^{\prime}=w_{2}v_{1}v_{2}w_{2}. Let uu be an arbitrary vertex with u​v3∈E​(Giβˆ—+1)uv_{3}\in E(G_{i^{*}+1}). Then by Lemma 4.1 and v3∈i​n​t​(Cβ€²)v_{3}\in int(C^{\prime}), we have u∈i​n​t​(Cβ€²)u\in int(C^{\prime}), and thus u​uiβˆ—+1βˆ‰E​(Giβˆ—+1)uu_{i^{*}+1}\notin E(G_{i^{*}+1}). This implies that NGiβˆ—+1​(uiβˆ—+1)βŠ†V​(Cβ€²)βˆͺW12N_{G_{i^{*}+1}}(u_{i^{*}+1})\subseteq V(C^{\prime})\cup W_{12}, where W12={u|u∈V​(Giβˆ—+1),NGiβˆ—+1​(u)={v1,v2}}W_{12}=\{u~{}|~{}u\in V(G_{i^{*}+1}),~{}N_{G_{i^{*}+1}}(u)=\{v_{1},v_{2}\}\}. Recall that dGiβˆ—+1​(uiβˆ—+1)=2d_{G_{i^{*}+1}}(u_{i^{*}+1})=2 and NGiβˆ—+1​(uiβˆ—+1)⊈V1N_{G_{i^{*}+1}}(u_{i^{*}+1})\not\subseteq V_{1}. Then, |NGiβˆ—+1​(uiβˆ—+1)∩{v1,v2}|≀1|N_{G_{i^{*}+1}}(u_{i^{*}+1})\cap\{v_{1},v_{2}\}|\leq 1. If |NGiβˆ—+1​(uiβˆ—+1)∩{v1,v2}|=1|N_{G_{i^{*}+1}}(u_{i^{*}+1})\cap\{v_{1},v_{2}\}|=1, then we may assume without loss of generality that v1∈NGiβˆ—+1​(uiβˆ—+1)v_{1}\in N_{G_{i^{*}+1}}(u_{i^{*}+1}), and uβ€²βˆˆNGiβˆ—+1​(uiβˆ—+1)βˆ–{v1}u^{\prime}\in N_{G_{i^{*}+1}}(u_{i^{*}+1})\setminus\{v_{1}\}. Since NGiβˆ—+1​(uiβˆ—+1)βŠ†V​(Cβ€²)βˆͺW12N_{G_{i^{*}+1}}(u_{i^{*}+1})\subseteq V(C^{\prime})\cup W_{12}, we have uβ€²βˆˆ{w2}βˆͺW12u^{\prime}\in\{w_{2}\}\cup W_{12}, and so u′​v1∈E​(Giβˆ—+1)u^{\prime}v_{1}\in E(G_{i^{*}+1}). Thus, Gnβˆ’5G_{n-5} contains two vertex-disjoint cycles uiβˆ—+1​v1​u′​uiβˆ—+1u_{i^{*}+1}v_{1}u^{\prime}u_{i^{*}+1} and w1​v2​v3​w1w_{1}v_{2}v_{3}w_{1}, as desired. Now consider the case that |NGiβˆ—+1​(uiβˆ—+1)∩{v1,v2}|=0|N_{G_{i^{*}+1}}(u_{i^{*}+1})\cap\{v_{1},v_{2}\}|=0. This implies that NGiβˆ—+1​(uiβˆ—+1)βŠ†{w2}βˆͺW12N_{G_{i^{*}+1}}(u_{i^{*}+1})\subseteq\{w_{2}\}\cup W_{12}. Let NGiβˆ—+1​(uiβˆ—+1)={uβ€²,uβ€²β€²}N_{G_{i^{*}+1}}(u_{i^{*}+1})=\{u^{\prime},u^{\prime\prime}\}. Then u′​v1,u′′​v1∈E​(Giβˆ—+1)u^{\prime}v_{1},u^{\prime\prime}v_{1}\in E(G_{i^{*}+1}). Therefore, Gnβˆ’5G_{n-5} contains two vertex-disjoint cycles uiβˆ—+1​u′​v1​u′′​uiβˆ—+1u_{i^{*}+1}u^{\prime}v_{1}u^{\prime\prime}u_{i^{*}+1} and w1​v2​v3​w1w_{1}v_{2}v_{3}w_{1}. ∎

Given a graph GG, let G~\widetilde{G} be the largest induced subgraph of GG with minimal degree at least 3. It is easy to see that G~\widetilde{G} can be obtained from GG 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 G~\widetilde{G} 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 nβ‰₯5n\geq 5 and GG be an extremal graph corresponding to ex𝒫​(n,2β€‹π’ž){\rm ex}_{\mathcal{P}}(n,2\mathcal{C}). Observe that K2+(P3βˆͺ(nβˆ’5)​K1)K_{2}+(P_{3}\cup(n-5)K_{1}) is a planar graph which contains no two vertex-disjoint cycles (see Figure 7). Thus, e​(G)β‰₯e​(K2+(P3βˆͺ(nβˆ’5)​K1))=2​nβˆ’1e(G)\geq e(K_{2}+(P_{3}\cup(n-5)K_{1}))=2n-1.

If G~\widetilde{G} is empty, then we define Gβ€²G^{\prime} as the graph obtained from GG by iteratively removing the vertices of degree at most 2 until the resulting graph has 4 vertices. By the planarity of Gβ€²G^{\prime}, we have e​(Gβ€²)≀3​|Gβ€²|βˆ’6=6e(G^{\prime})\leq 3|G^{\prime}|-6=6, and thus

e​(G)≀e​(Gβ€²)+2​(nβˆ’4)≀6+2​nβˆ’8=2​nβˆ’2,e(G)\leq e(G^{\prime})+2(n-4)\leq 6+2n-8=2n-2,

a contradiction.

Now we know that G~\widetilde{G} is nonempty. Then, G~\widetilde{G} contains no two vertex-disjoint cycles as G~βŠ†G\widetilde{G}\subseteq G. By the definition of G~\widetilde{G}, we have δ​(G~)β‰₯3\delta(\widetilde{G})\geq 3. By Lemma 4.3, we get that G~∈{2​K1+C3,K1+C|G~|βˆ’1}\widetilde{G}\in\{2K_{1}+C_{3},K_{1}+C_{|\widetilde{G}|-1}\}. If G~β‰…K1+C|G~|βˆ’1\widetilde{G}\cong K_{1}+C_{|\widetilde{G}|-1}, then

e​(G)≀e​(G~)+2​(nβˆ’|G~|)=2​(|G~|βˆ’1)+2​(nβˆ’|G~|)=2​nβˆ’2,e(G)\leq e(\widetilde{G})+2(n-|\widetilde{G}|)=2(|\widetilde{G}|-1)+2(n-|\widetilde{G}|)=2n-2,

a contradiction. Thus, G~β‰…2​K1+C3\widetilde{G}\cong 2K_{1}+C_{3}. Now, e​(G)≀e​(G~)+2​(nβˆ’5)=2​nβˆ’1e(G)\leq e(\widetilde{G})+2(n-5)=2n-1. Therefore, e​(G)=2​nβˆ’1e(G)=2n-1, which implies that ex𝒫​(n,2β€‹π’ž)=2​nβˆ’1{\rm ex}_{\mathcal{P}}(n,2\mathcal{C})=2n-1 and Gβˆˆπ’’nG\in\mathcal{G}_{n}. By Lemma 4.4, we have Gβˆˆπ’’nβˆ—G\in\mathcal{G}^{*}_{n}. This completes the proof of Theorem 1.4. ∎

5 Proof of Theorem 1.5

We shall further introduce some notations on a plane graph GG. A vertex or an edge of GG is said to be incident with a face FF, if it lies on the boundary of FF. Clearly, every edge of GG is incident with at most two faces. A face of size ii is called an ii-face. The numbers of ii-faces and total faces are denoted by fi​(G)f_{i}(G) and f​(G)f(G), respectively. Let E3​(G)E_{3}(G) be the set of edges incident with at least one 33-face, and particularly, let E3,3​(G)E_{3,3}(G) be the set of edges incident with two 33-faces. Moreover, let e3​(G)e_{3}(G) and e3,3​(G)e_{3,3}(G) denote the cardinalities of E3​(G)E_{3}(G) and E3,3​(G)E_{3,3}(G), respectively. We can easily see that 3​f3​(G)=e3​(G)+e3,3​(G).3f_{3}(G)=e_{3}(G)+e_{3,3}(G).

Lan, Shi and Song proved that ex𝒫​(n,K1+P3)≀12​(nβˆ’2)5{\rm ex}_{\mathcal{P}}(n,K_{1}+P_{3})\leq\frac{12(n-2)}{5}, with equality when n≑12(mod20)n\equiv 12\pmod{20} (see [16]), and ex𝒫​(n,K1+Pk+1)≀13​k​n4​k+2βˆ’12​k2​k+1{\rm ex}_{\mathcal{P}}(n,K_{1}+P_{k+1})\leq\frac{13kn}{4k+2}-\frac{12k}{2k+1} for k∈{3,4,5}k\in\{3,4,5\} (see [17]). For kβ‰₯6k\geq 6, one can easily see that ex𝒫​(n,K1+Pk+1)=3​nβˆ’6{\rm ex}_{\mathcal{P}}(n,K_{1}+P_{k+1})=3n-6. In [10], the authors obtained the following sharp result.

Lemma 5.1.

([10]) Let n,kn,k be two integers with k∈{2,3,4,5}k\in\{2,3,4,5\} and nβ‰₯126βˆ’k+1n\geq\frac{12}{6-k}+1. Then ex𝒫​(n,K1+Pk+1)≀24​k7​k+6​(nβˆ’2){\rm ex}_{\mathcal{P}}(n,K_{1}+P_{k+1})\leq\frac{24k}{7k+6}(n-2), with equality if n≑12​(k+2)6βˆ’k(mod28​k+246βˆ’k)n\equiv{\frac{12(k+2)}{6-k}}\pmod{{\frac{28k+24}{6-k}}}.

To prove Theorem 1.5, we also need an edge-extremal result on outerplanar graphs. Let exπ’ͺ​𝒫​(n,Ck){\rm ex}_{\mathcal{OP}}(n,C_{k}) denote the maximum number of edges in an nn-vertex CkC_{k}-free outerplanar graph.

Lemma 5.2.

([11]) Let n,k,Ξ»n,k,\lambda be three integers with nβ‰₯kβ‰₯3n\geq k\geq 3 and Ξ»=⌊k​nβˆ’2​kβˆ’1k2βˆ’2​kβˆ’1βŒ‹+1\lambda=\lfloor\frac{kn-2k-1}{k^{2}-2k-1}\rfloor+1. Then

exπ’ͺ​𝒫​(n,Ck)={2​nβˆ’Ξ»+2β€‹βŒŠΞ»kβŒ‹βˆ’3ifΒ k∣λ,2​nβˆ’Ξ»+2β€‹βŒŠΞ»kβŒ‹βˆ’2otherwise.{\rm ex}_{\mathcal{OP}}(n,C_{k})=\left\{\begin{array}[]{ll}2n-\lambda+2\big{\lfloor}\frac{\lambda}{k}\big{\rfloor}-3&\hbox{if $k\mid\lambda$,}\\ 2n-\lambda+2\big{\lfloor}\frac{\lambda}{k}\big{\rfloor}-2&\hbox{otherwise.}\end{array}\right.

In particular, we can obtain the following corollary.

Corollary 5.1.

Let nβ‰₯4n\geq 4. Then

exπ’ͺ​𝒫​(nβˆ’1,C4)={127​nβˆ’5ifΒ 7∣n,⌊12​nβˆ’277βŒ‹otherwise.{\rm ex}_{\mathcal{OP}}(n-1,C_{4})=\left\{\begin{array}[]{ll}\frac{12}{7}n-5&\hbox{if $7\mid n$,}\\ \big{\lfloor}\frac{12n-27}{7}\big{\rfloor}{}{}{}&\hbox{otherwise.}\end{array}\right.
Refer to caption
Figure 9: The constructions of G,G1,G2,…,GaG,G_{1},G_{2},\ldots,G_{a}.

For arbitrary integer nβ‰₯4n\geq 4, we can find a unique (a,b)(a,b) such that aβ‰₯0a\geq 0, 1≀b≀71\leq b\leq 7 and nβˆ’1=7​a+b+2n-1=7a+b+2. Let GG be a 99-vertex outerplanar graph and G1,…,GaG_{1},\ldots,G_{a} be aa copies of GG (see Figure 9). Then, we define G0G_{0} as the subgraph of GG induced by {u1,u2}βˆͺ{v1,v2,…,vb}\{u_{1},u_{2}\}\cup\{v_{1},v_{2},\dots,v_{b}\}. One can check that |G0|=b+2|G_{0}|=b+2 and

e​(G0)={12​(b+2)βˆ’237ifΒ 7∣(b+2βˆ’6),⌊12​(b+2)βˆ’157βŒ‹otherwise.e(G_{0})=\left\{\begin{array}[]{ll}\frac{12(b+2)-23}{7}&\hbox{if $7\mid(b+2-6)$,}\\ \big{\lfloor}\frac{12(b+2)-15}{7}\big{\rfloor}{}{}{}&\hbox{otherwise.}\end{array}\right.

We now construct a new graph Gβˆ—G^{*} from G0,G1,…,GaG_{0},G_{1},\ldots,G_{a} by identifying the edges e2​ie_{2i} and e2​i+1e_{2i+1} for each i∈{0,…,aβˆ’1}i\in\{0,\ldots,a-1\}. Clearly, Gβˆ—G^{*} is a connected C4C_{4}-free outerplanar graph with

|Gβˆ—|=βˆ‘i=0a|Gi|βˆ’2​a=(2+b)+9​aβˆ’2​a=nβˆ’1.|G^{*}|=\sum_{i=0}^{a}|G_{i}|-2a=(2+b)+9a-2a=n-1.

Moreover, since n≑b+2βˆ’6(mod7),n\equiv b+2-6\,(\!\!\!\mod 7), we have

e​(Gβˆ—)=βˆ‘i=0ae​(Gi)βˆ’a=e​(G0)+12​a={127​nβˆ’5ifΒ 7∣n,⌊12​nβˆ’277βŒ‹otherwise.e(G^{*})=\sum_{i=0}^{a}e(G_{i})-a=e(G_{0})+12a=\left\{\begin{array}[]{ll}\frac{12}{7}n-5&\hbox{if $7\mid n$,}\\ \big{\lfloor}\frac{12n-27}{7}\big{\rfloor}{}{}{}&\hbox{otherwise.}\end{array}\right.

Combining Corollary 5.1, Gβˆ—G^{*} is an extremal graph corresponding to exπ’ͺ​𝒫​(nβˆ’1,C4){\rm ex}_{\mathcal{OP}}(n-1,C_{4}).

Lemma 5.3.

Let nβ‰₯2661n\geq 2661 and Gβˆ—βˆ—G^{**} be an extremal plane graph corresponding to ex𝒫​(n,2​C4){\rm ex}_{\mathcal{P}}(n,2C_{4}). Then Gβˆ—βˆ—G^{**} contains at least fourteen quadrilaterals and all of them share exactly one vertex.

Proof.

Note that e​(Gβˆ—)=exπ’ͺ​𝒫​(nβˆ’1,C4)β‰₯127​nβˆ’5e(G^{*})={\rm ex}_{\mathcal{OP}}(n-1,C_{4})\geq\frac{12}{7}n-5 and Gβˆ—G^{*} is a C4C_{4}-free outerplanar graph of order nβˆ’1n-1. Then K1+Gβˆ—K_{1}+G^{*} is an nn-vertex 2​C42C_{4}-free planar graph, and thus

e​(Gβˆ—βˆ—)β‰₯e​(K1+Gβˆ—)=e​(Gβˆ—)+nβˆ’1β‰₯197​nβˆ’6.\displaystyle e(G^{**})\geq e(K_{1}+G^{*})=e(G^{*})+n-1\geq\frac{19}{7}n-6.

On the other hand, by Lemma 5.1, we have

ex𝒫​(n,K1+P4)≀83​(nβˆ’2).{\rm ex}_{\mathcal{P}}(n,K_{1}+P_{4})\leq\frac{8}{3}(n-2).

Note that 197​nβˆ’6>83​(nβˆ’2)\frac{19}{7}n-6>\frac{8}{3}(n-2) for nβ‰₯2661n\geq 2661. Then Gβˆ—βˆ—G^{**} contains a copy, say H1H_{1}, of K1+P4K_{1}+P_{4}. Let G1G_{1} be the graph obtained from Gβˆ—βˆ—G^{**} by deleting all edges within V​(H1)V(H_{1}). Since |H1|=5|H_{1}|=5, we have e​(G1)β‰₯e​(Gβˆ—βˆ—)βˆ’(3​|H1|βˆ’6)=e​(Gβˆ—βˆ—)βˆ’9>83​(nβˆ’2)e(G_{1})\geq e(G^{**})-(3|H_{1}|-6)=e(G^{**})-9>\frac{8}{3}(n-2). Thus, G1G_{1} contains a copy, say H2H_{2}, of K1+P4K_{1}+P_{4}. Now we can obtain a new graph G2G_{2} from G1G_{1} by deleting all edges within V​(H2)V(H_{2}). Note that e​(Gβˆ—βˆ—)βˆ’14Γ—9>83​(nβˆ’2)e(G^{**})-14\times 9>\frac{8}{3}(n-2). Repeating above steps, we can obtain a graph sequence G1,G2,…,G14G_{1},G_{2},\ldots,G_{14} and fourteen copies H1,H2,β‹―,H14H_{1},H_{2},\cdots,H_{14} of K1+P4K_{1}+P_{4} such that HiβŠ†Giβˆ’1H_{i}\subseteq G_{i-1} and GiG_{i} is obtained from Giβˆ’1G_{i-1} by deleting all edges within V​(Hi)V(H_{i}). This also implies that Gβˆ—βˆ—G^{**} contains at least fourteen quadrilaterals. We next give four claims on those copies of K1+P4K_{1}+P_{4}.

Claim 5.1.

Let i,ji,j be two integers with 1≀i<j≀141\leq i<j\leq 14 and v∈V​(Hi)∩V​(Hj)v\in V(H_{i})\cap V(H_{j}). Then, V​(Hi)∩NHj​(v)=βˆ…V(H_{i})\cap N_{H_{j}}(v)=\varnothing.

Proof.

Suppose to the contrary that there exists a vertex w∈V​(Hi)∩NHj​(v)w\in V(H_{i})\cap N_{H_{j}}(v). Note that v,w∈V​(Hi)v,w\in V(H_{i}). By the definition of GiG_{i}, whether v​w∈E​(Hi)vw\in E(H_{i}) or not, we can see that v​wβˆ‰E​(Gi)vw\notin E(G_{i}). On the other hand, note that HjβŠ†Gjβˆ’1βŠ†GiH_{j}\subseteq G_{j-1}\subseteq G_{i}, then v​w∈E​(Hj)βŠ†E​(Gi)vw\in E(H_{j})\subseteq E(G_{i}), contradicting v​wβˆ‰E​(Gi).vw\notin E(G_{i}). Hence, the claim holds. ∎

Claim 5.2.

|V​(Hi)∩V​(Hj)|∈{1,2}|V(H_{i})\cap V(H_{j})|\in\{1,2\} for any two integers i,ji,j with 1≀i<j≀141\leq i<j\leq 14.

Proof.

If HiH_{i} and HjH_{j} are vertex-disjoint, then Gβˆ—βˆ—G^{**} contains 2​C42C_{4}, a contradiction. Now suppose that there exist three vertices v1,v2,v3∈V​(Hi)∩V​(Hj)v_{1},v_{2},v_{3}\in V(H_{i})\cap V(H_{j}). Observe that K1+P4K_{1}+P_{4} contains no an independent set of size 33. Then Hj​[{v1,v2,v3}]H_{j}[\{v_{1},v_{2},v_{3}\}] is nonempty. Assume without loss of generality that v1​v2∈E​(Hj)v_{1}v_{2}\in E(H_{j}). Then v2∈V​(Hi)∩NHj​(v1)v_{2}\in V(H_{i})\cap N_{H_{j}}(v_{1}), which contradicts Claim 5.1. Therefore, 1≀|V​(Hi)∩V​(Hj)|≀21\leq|V(H_{i})\cap V(H_{j})|\leq 2. ∎

Now for convenience, a vertex vv in a graph GG is called a 22-vertex if dG​(v)=2d_{G}(v)=2, and a 2+2^{+}-vertex if dG​(v)>2.d_{G}(v)>2. Clearly, every copy of K1+P4K_{1}+P_{4} contains two 2-vertices and three 2+2^{+}-vertices.

Claim 5.3.

Let β„‹\mathcal{H} be the family of graphs HiH_{i} (1≀i≀141\leq i\leq 14) such that every 2-vertex in HiH_{i} is a 2+2^{+}-vertex in H1H_{1}. Then |β„‹|≀3|\mathcal{H}|\leq 3.

Proof.

Note that H1H_{1} contains only three 2+2^{+}-vertices, say v1,v2v_{1},v_{2} and v3v_{3}. Then every graph Hiβˆˆβ„‹H_{i}\in\mathcal{H} must contain two of v1,v2v_{1},v_{2} and v3v_{3} as 22-vertices. Suppose to the contrary that |β„‹|β‰₯4|\mathcal{H}|\geq 4. By pigeonhole principle, there exist two graphs Hi1,Hi2βˆˆβ„‹H_{i_{1}},H_{i_{2}}\in\mathcal{H} such that they contain the same two 2-vertices, say v1,v2v_{1},v_{2}. It follows that Hijβˆ’{vj}H_{i_{j}}-\{v_{j}\} contains a 44-cycle for j∈{1,2}j\in\{1,2\}. By Claim 5.2, we have V​(Hi1)∩V​(Hi2)={v1,v2}V(H_{i_{1}})\cap V(H_{i_{2}})=\{v_{1},v_{2}\}, which implies that Hi1βˆ’{v1}H_{i_{1}}-\{v_{1}\} and Hi2βˆ’{v2}H_{i_{2}}-\{v_{2}\} are vertex-disjoint. Hence, Gβˆ—βˆ—G^{**} contains two vertex-disjoint 4-cycles, a contradiction. ∎

Claim 5.4.

Let jj be an integer with 2≀j≀142\leq j\leq 14 and Hjβˆ‰β„‹H_{j}\notin\mathcal{H}. Then, there exists a vertex v∈V​(H1)∩V​(Hj)v\in V(H_{1})\cap V(H_{j}) such that dH1​(v)β‰₯3d_{H_{1}}(v)\geq 3 and dHj​(v)β‰₯3d_{H_{j}}(v)\geq 3.

Proof.

By Claim 5.2, we have 1≀|V​(H1)∩V​(Hj)|≀21\leq|V(H_{1})\cap V(H_{j})|\leq 2. We first assume that V​(H1)∩V​(Hj)={u}V(H_{1})\cap V(H_{j})=\{u\}. If dH1​(u)β‰₯3d_{H_{1}}(u)\geq 3 and dHj​(u)β‰₯3d_{H_{j}}(u)\geq 3, then there is nothing to prove. If dH1​(u)=2d_{H_{1}}(u)=2, then Gβˆ—βˆ—G^{**} contains two vertex-disjoint subgraphs H1βˆ’{u}H_{1}-\{u\} and HjH_{j}, and thus 2​C42C_{4}, a contradiction. If dHj​(u)=2d_{H_{j}}(u)=2, then we can similarly get a contradiction. Therefore, |V​(H1)∩V​(Hj)|=2|V(H_{1})\cap V(H_{j})|=2.

Now, assume that V​(H1)∩V​(Hj)={u1,u2}V(H_{1})\cap V(H_{j})=\{u_{1},u_{2}\}. We first deal with the case dHj​(u1)d_{H_{j}}(u_{1}) =dHj​(u2)=2=d_{H_{j}}(u_{2})=2. Since Hjβˆ‰β„‹H_{j}\notin\mathcal{H}, one of {u1,u2}\{u_{1},u_{2}\}, say u1u_{1}, is a 2-vertex in H1H_{1}. Hence, Gβˆ—βˆ—G^{**} contains two vertex-disjoint subgraphs H1βˆ’{u1}H_{1}-\{u_{1}\} and Hjβˆ’{u2}H_{j}-\{u_{2}\}, and so 2​C42C_{4}, a contradiction. Thus, there exists some i∈{1,2}i\in\{1,2\} with dHj​(ui)β‰₯3d_{H_{j}}(u_{i})\geq 3. If dH1​(ui)β‰₯3d_{H_{1}}(u_{i})\geq 3, then we are done. If dH1​(ui)=2d_{H_{1}}(u_{i})=2, then we define Hjβ€²H_{j}^{\prime} as the subgraph of HjH_{j} induced by NHj​(ui)βˆͺ{ui}N_{H_{j}}(u_{i})\cup\{u_{i}\}. Since dHj​(ui)β‰₯3d_{H_{j}}(u_{i})\geq 3, we can check that Hjβ€²H_{j}^{\prime} always contains a C4C_{4}. Moreover, since dH1​(ui)=2d_{H_{1}}(u_{i})=2, we can see that H1βˆ’{ui}H_{1}-\{u_{i}\} also contains a C4C_{4}. On the other hand, by Claim 5.1, we have NHj​(ui)∩V​(H1)=βˆ…N_{H_{j}}(u_{i})\cap V(H_{1})=\varnothing, which implies that Hjβ€²H_{j}^{\prime} and H1βˆ’{ui}H_{1}-\{u_{i}\} are vertex-disjoint. Therefore, Gβˆ—βˆ—G^{**} contains 2​C42C_{4}, a contradiction. ∎

By Claim 5.3, |β„‹|≀3|\mathcal{H}|\leq 3, thus there are at least ten graphs in {Hj|2≀j≀14}βˆ–β„‹\{H_{j}~{}|~{}2\leq j\leq 14\}\setminus\mathcal{H}. However, H1H_{1} has only three 2+2^{+}-vertices. By Claim 5.4 and pigeonhole principle, there exists a 2+2^{+}-vertex ww in H1H_{1} and four graphs, say H2,H3,H4,H5H_{2},H_{3},H_{4},H_{5}, of {Hj|2≀j≀14}βˆ–β„‹\{H_{j}~{}|~{}2\leq j\leq 14\}\setminus\mathcal{H}. By Claim 5.1, we get that NHj​(w)∩V​(Hi)=βˆ…N_{H_{j}}(w)\cap V(H_{i})=\varnothing, and so NHj​(w)∩NHi​(w)=βˆ…N_{H_{j}}(w)\cap N_{H_{i}}(w)=\varnothing, for any i,ji,j with 1≀i<j≀51\leq i<j\leq 5. If Gβˆ—βˆ—βˆ’{w}G^{**}-\{w\} contains a quadrilateral Cβ€²C^{\prime}, then there exists some j′≀5j^{\prime}\leq 5 such that NHj′​(w)∩V​(Cβ€²)=βˆ…N_{H_{j^{\prime}}}(w)\cap V(C^{\prime})=\varnothing as |Cβ€²|=4|C^{\prime}|=4. Since ww is a 2+2^{+}-vertex in Hjβ€²H_{j^{\prime}}, we can observe that the subgraph of Hjβ€²H_{j^{\prime}} induced by NHj′​(w)βˆͺ{w}N_{H_{j^{\prime}}}(w)\cup\{w\} must contain a C4C_{4}. Consequently, Gβˆ—βˆ—G^{**} is not 2​C42C_{4}-free, a contradiction. Thus, Gβˆ—βˆ—βˆ’{w}G^{**}-\{w\} is C4C_{4}-free, which implies that all quadrilaterals of Gβˆ—βˆ—G^{**} share exactly one vertex. This completes the proof of Lemma 5.3. ∎

Now we are ready to give the proof of Theorem 1.5.

Proof.

Recall that Gβˆ—G^{*} is an extremal graph corresponding to exπ’ͺ​𝒫​(nβˆ’1,C4){\rm ex}_{\mathcal{OP}}(n-1,C_{4}). Then K1+Gβˆ—K_{1}+G^{*} is planar and 2​C42C_{4}-free. By Corollary 5.1, we have

e​(K1+Gβˆ—)=exπ’ͺ​𝒫​(nβˆ’1,C4)+nβˆ’1={197​nβˆ’6ifΒ 7∣n,⌊19​nβˆ’347βŒ‹otherwise.\displaystyle e(K_{1}+G^{*})={\rm ex}_{\mathcal{OP}}(n-1,C_{4})+n-1=\left\{\begin{array}[]{ll}\frac{19}{7}n-6&\hbox{if $7\mid n$,}\\ \big{\lfloor}\frac{19n-34}{7}\big{\rfloor}{}&\hbox{otherwise.}\end{array}\right. (30)

To prove Theorem 1.5, it suffices to show ex𝒫​(n,2​C4)=e​(K1+Gβˆ—).{\rm ex}_{\mathcal{P}}(n,2C_{4})=e(K_{1}+G^{*}). Since Gβˆ—βˆ—G^{**} is an extremal plane graph corresponding to ex𝒫​(n,2​C4){\rm ex}_{\mathcal{P}}(n,2C_{4}), we have e​(Gβˆ—βˆ—)β‰₯e​(K1+Gβˆ—)e(G^{**})\geq e(K_{1}+G^{*}). In the following, we show that e​(Gβˆ—βˆ—)≀e​(K1+Gβˆ—)e(G^{**})\leq e(K_{1}+G^{*}).

Refer to caption
Figure 10: Two possible structures of H​(e)H(e).

By Lemma 5.3, all quadrilaterals of Gβˆ—βˆ—G^{**} share a vertex ww. Thus, Gβˆ—βˆ—βˆ’{w}G^{**}-\{w\} is C4C_{4}-free. Assume that dGβˆ—βˆ—β€‹(w)=sd_{G^{**}}(w)=s and w1,…,wsw_{1},\dots,w_{s} are around ww in clockwise order, with subscripts interpreted modulo ss. Let ee be an arbitrary edge in E3,3​(Gβˆ—βˆ—)E_{3,3}(G^{**}), that is, ee is incident with two 3-faces, say FF and Fβ€²F^{\prime}. We define H​(e)H(e) as the plane subgraph induced by all edges incident with FF and Fβ€²F^{\prime}. Clearly, H​(e)β‰…K1+P3H(e)\cong K_{1}+P_{3} and so it contains a C4C_{4}. Recall that all quadrilaterals of Gβˆ—βˆ—G^{**} share exactly one vertex ww. Then, w∈V​(He)w\in V(H_{e}) and ww is incident with at least one face of FF and Fβ€²F^{\prime} (see Figure 10). Note that ee is incident with FF. Then, either e=w​wie=ww_{i} or e=wi​wi+1e=w_{i}w_{i+1} for some i∈{1,2,…,s}i\in\{1,2,\dots,s\}. By the choice of ee, we have

E3,3​(Gβˆ—βˆ—)βŠ†{w​wi,wi​wi+1|1≀i≀s}.\displaystyle E_{3,3}(G^{**})\subseteq\{ww_{i},w_{i}w_{i+1}~{}|~{}1\leq i\leq s\}. (31)

Assume first that f4​(Gβˆ—βˆ—)=tβ‰₯1f_{4}(G^{**})=t\geq 1 and F1,…,FtF_{1},\dots,F_{t} are 4-faces in Gβˆ—βˆ—G^{**}. Since every 4-face is a quadrilateral, ww is incident with each 4-face. Consequently, there exists ji∈{1,…,s}j_{i}\in\{1,\dots,s\} such that w​wji,w​wji+1ww_{j_{i}},ww_{j_{i}+1} are incident with FiF_{i} for each i∈{1,…,t}i\in\{1,\dots,t\}. Thus, w​wjiβˆ‰E3,3​(Gβˆ—βˆ—)ww_{j_{i}}\notin E_{3,3}(G^{**}) for 1≀i≀t1\leq i\leq t. On the other hand, if wji​wji+1∈E3,3​(Gβˆ—βˆ—)w_{j_{i}}w_{j_{i}+1}\in E_{3,3}(G^{**}), then H​(wji​wji+1)H(w_{j_{i}}w_{j_{i}+1}) contains a C4C_{4}, and so w∈V​(H​(wji​wji+1))w\in V(H(w_{j_{i}}w_{j_{i}+1})). This implies that w​wji​wji+1​www_{j_{i}}w_{j_{i}+1}w is a 3-face in Gβˆ—βˆ—G^{**}, contradicting the fact that w​wji,w​wji+1ww_{j_{i}},ww_{j_{i}+1} are incident with the 4-face FiF_{i}. Thus, we also have wji​wji+1βˆ‰E3,3​(Gβˆ—βˆ—)w_{j_{i}}w_{j_{i}+1}\notin E_{3,3}(G^{**}) for 1≀i≀t1\leq i\leq t.

By the argument above, we can see that

E3,3​(Gβˆ—βˆ—)∩{w​wji,wji​wji+1|1≀i≀t}=βˆ….\displaystyle E_{3,3}(G^{**})\cap\{ww_{j_{i}},w_{j_{i}}w_{j_{i}+1}~{}|~{}1\leq i\leq t\}=\varnothing. (32)

Using (31) and (32) gives e3,3​(Gβˆ—βˆ—)≀2​sβˆ’2​t=2​sβˆ’2​f4​(Gβˆ—βˆ—)e_{3,3}(G^{**})\leq 2s-2t=2s-2f_{4}(G^{**}). Hence,

3​f3​(Gβˆ—βˆ—)=e3​(Gβˆ—βˆ—)+e3,3​(Gβˆ—βˆ—)≀e​(Gβˆ—βˆ—)+2​sβˆ’2​f4​(Gβˆ—βˆ—).\displaystyle 3f_{3}(G^{**})=e_{3}(G^{**})+e_{3,3}(G^{**})\leq e(G^{**})+2s-2f_{4}(G^{**}). (33)

On the other hand,

2​e​(Gβˆ—βˆ—)=βˆ‘iβ‰₯3i​fi​(Gβˆ—βˆ—)β‰₯3​f3​(Gβˆ—βˆ—)+4​f4​(Gβˆ—βˆ—)+5​(f​(Gβˆ—βˆ—)βˆ’f3​(Gβˆ—βˆ—)βˆ’f4​(Gβˆ—βˆ—)),\displaystyle 2e(G^{**})=\sum_{i\geq 3}if_{i}(G^{**})\geq 3f_{3}(G^{**})+4f_{4}(G^{**})+5(f(G^{**})-f_{3}(G^{**})-f_{4}(G^{**})),

which yields f​(Gβˆ—βˆ—)≀15​(2​e​(Gβˆ—βˆ—)+2​f3​(Gβˆ—βˆ—)+f4​(Gβˆ—βˆ—)).f(G^{**})\leq\frac{1}{5}\left(2e(G^{**})+2f_{3}(G^{**})+f_{4}(G^{**})\right). Combining this with Euler’s formula f​(Gβˆ—βˆ—)=e​(Gβˆ—βˆ—)βˆ’(nβˆ’2)f(G^{**})=e(G^{**})-(n-2), we obtain

e​(Gβˆ—βˆ—)≀53​(nβˆ’2)+23​f3​(Gβˆ—βˆ—)+13​f4​(Gβˆ—βˆ—).\displaystyle e(G^{**})\leq\frac{5}{3}(n-2)+\frac{2}{3}f_{3}(G^{**})+\frac{1}{3}f_{4}(G^{**}). (34)

If f4​(Gβˆ—βˆ—)=t=0f_{4}(G^{**})=t=0, then (33) and (34) hold directly. Combining (33) and (34), we have e​(Gβˆ—βˆ—)≀157​(nβˆ’2)+47​sβˆ’17​f4​(Gβˆ—βˆ—)e(G^{**})\leq\frac{15}{7}(n-2)+\frac{4}{7}s-\frac{1}{7}f_{4}(G^{**}). Recall that dGβˆ—βˆ—β€‹(w)=s≀nβˆ’1d_{G^{**}}(w)=s\leq n-1. If s≀nβˆ’2s\leq n-2, then e​(Gβˆ—βˆ—)β‰€βŒŠ197​(nβˆ’2)βŒ‹β‰€e​(K1+Gβˆ—)e(G^{**})\leq\lfloor\frac{19}{7}(n-2)\rfloor\leq e(K_{1}+G^{*}) by (30), as desired. If s=nβˆ’1s=n-1, then ww is a dominating vertex of the planar graph Gβˆ—βˆ—G^{**}, which implies that Gβˆ—βˆ—βˆ’{w}G^{**}-\{w\} is outerplanar. Recall that Gβˆ—βˆ—βˆ’{w}G^{**}-\{w\} is C4C_{4}-free. Thus, e​(Gβˆ—βˆ—βˆ’{w})≀exπ’ͺ​𝒫​(nβˆ’1,C4)e(G^{**}-\{w\})\leq{\rm ex}_{\mathcal{OP}}(n-1,C_{4}), and so e​(Gβˆ—βˆ—)≀exπ’ͺ​𝒫​(nβˆ’1,C4)+nβˆ’1e(G^{**})\leq{\rm ex}_{\mathcal{OP}}(n-1,C_{4})+n-1. Combining (30), we get e​(Gβˆ—βˆ—)≀e​(K1+Gβˆ—)e(G^{**})\leq e(K_{1}+G^{*}), as required. This completes the proof of Theorem 1.5. ∎

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), 251–257.
  • [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. CioabΔƒ, D.N. Desai, M. Tait, The spectral even cycle problem, arxiv:2205.00990v1 (2022).
  • [6] C. Dowden, Extremal C4C_{4}-free/C5C_{5}-free planar graphs, J. Graph Theory 83 (2016), 213–230.
  • [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), 2395–2405.
  • [8] P. ErdΕ‘s, Über ein Extremal problem in der Graphentheorie, Arch. Math. 13 (1962), 222–227.
  • [9] P. ErdΕ‘s, L. PΓ³sa, On the maximal number of disjoint circuits of a graph, Publ. Math. Debrecen 9 (1962), 3–12.
  • [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), 641–645.
  • [13] Z. FΓΌredi, M. Simonovits, The history of degenerate (bipartite) extremal graph problems, Bolyai Soc. Studies (The ErdΕ‘s Centennial) 25 (2013), 167–262.
  • [14] D. Ghosh, E. GyΕ‘ri, R.R. Martin, A. Paulos, C.Q. Xiao, Planar TurΓ‘n number of the 6-cycle, SIAM J. Discrete Math. 36 (3) (2022), 2028–2050.
  • [15] E. GyΕ‘ri, 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 HH-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), 200–216.
  • [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), 260–274.
  • [21] J.W. Moon, On independent complete subgraphs in a graph, Canadian J. Math. 20 (1968) 95–102.
  • [22] V. Nikiforov, A spectral condition for odd cycles in graphs, Linear Algebra Appl. 428 (2008), no. 7, 1492–1498.
  • [23] V. Nikiforov, Bounds on graph eigenvalues II, Linear Algebra Appl. 427 (2-3) (2007), 183–189.
  • [24] B. Pittel, J. Spencer, N. Wormald, Sudden emergence of a giant kk-core in a random graph, J. Combin. Theory Ser. B 67 (1996), 111–151.
  • [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), 137–161.
  • [28] M.Q. Zhai, B. Wang, Proof of a conjecture on the spectral radius of C4C_{4}-free graphs, Linear Algebra Appl. 437 (7) (2012), 1641–1647.