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

Multiple solutions of Kazdan-Warner equation on graphs in the negative case

Shuang Liu [email protected] Yunyan Yang111Corresponding author [email protected] Department of Mathematics, Renmin University of China, Beijing 100872, P. R. China
Abstract

Let G=(V,E)G=(V,E) be a finite connected graph, and let κ:V\kappa:V\rightarrow\mathbb{R} be a function such that Vκ𝑑μ<0\int_{V}\kappa d\mu<0. We consider the following Kazdan-Warner equation on GG:

Δu+κKλe2u=0,\Delta u+\kappa-K_{\lambda}e^{2u}=0,

where Kλ=K+λK_{\lambda}=K+\lambda and K:VK:V\rightarrow\mathbb{R} is a non-constant function satisfying maxxVK(x)=0\max_{x\in V}K(x)=0 and λ\lambda\in\mathbb{R}. By a variational method, we prove that there exists a λ>0\lambda^{*}>0 such that when λ(,λ]\lambda\in(-\infty,\lambda^{*}] the above equation has solutions, and has no solution when λλ\lambda\geq\lambda^{\ast}. In particular, it has only one solution if λ0\lambda\leq 0; at least two distinct solutions if 0<λ<λ0<\lambda<\lambda^{*}; at least one solution if λ=λ\lambda=\lambda^{\ast}. This result complements earlier work of Grigor’yan-Lin-Yang [7], and is viewed as a discrete analog of that of Ding-Liu [4] and Yang-Zhu [17] on manifolds.

keywords:
Kazdan-Warnar problem on graph , variation problem on graph
MSC:
[2010] 35R02, 34B45
journal: ***

1 Introduction

Variational method is always a powerful tool in partial differential equations and geometric analysis. Recently, using this tool, Grigor’yan-Lin-Yang [7, 8, 9] obtained existence results for solutions to various partial differential equations on graphs. In particular, Kazdan-warnar equation was proposed on graphs in [7]. The Kazdan-Warner equation arises from the basic geometric problem on prescribing Gaussian curvature of Riemann surface, which systematically studied by Kazdan-Warner [12, 13]. On a closed Riemann surface (Σ,g)(\Sigma,g) with the Gaussian curvature κ\kappa, let g~=e2ug\widetilde{g}=e^{2u}g be a smooth metric conformal to gg and KK be the Gaussian curvature with respect to g~\widetilde{g}. Then uu satisfies the equation

Δgu+κKe2u=0,\Delta_{g}u+\kappa-Ke^{2u}=0, (1)

where Δg\Delta_{g} denotes the Laplace-Beltrami operator with respect to the metric gg. Let vv be a solution to Δgv=κ¯κ\Delta_{g}v=\overline{\kappa}-\kappa and f=2(uv)f=2(u-v), where κ¯\overline{\kappa} is the averaged integral of κ\kappa. Then the above equation is transformed to

Δgf+2κ¯(2Ke2v)ef=0.\Delta_{g}f+2\overline{\kappa}-(2Ke^{2v})e^{f}=0.

Hence, one can free (1) from the geometric situation, and just studies the equation

Δgf+chef=0,\Delta_{g}f+c-he^{f}=0, (2)

where cc is a constant and hh is a function. On graphs, it seems to be out of reach to resemble this topic in terms of Gaussian curvature. Therefore, in [7], the authors focused on the equation similar to the form of (2), namely the Kazdan-Warner equation on graph, and obtained the following: when c=0c=0, it has a solution if and only if hh changes sign and the integral of hh is negative; when c>0c>0, it has a solution if and only if hh is positive somewhere; when c<0c<0, there is a threshold ch<0c_{h}<0 such that it has a solution if c(ch,0)c\in(c_{h},0), but it has no solution for any c<chc<c_{h}. Later, Ge [5] found a solution in the critical case c=chc=c_{h}. More recently Ge-Jiang [6] studied the Kazdan-Warner equation on infinite graphs and Keller-Schwarz [11] on canonically compactifiable graphs; Camilli-Marchi [3] extended the Kazdan-Warner equation on network; for other related works, we refer the readers to [10, 14].

Let us come back to a closed Riemann surface (Σ,g)(\Sigma,g), whose Euler characteristic is negative, or equivalently Σκ𝑑vg<0\int_{\Sigma}\kappa dv_{g}<0. Replacing KK by K+λK+\lambda in (1) with K0K\leq 0, K0K\not\equiv 0, and λ\lambda\in\mathbb{R}, Ding-Liu [4] obtained the following conclusion by using a method of upper and lower solutions and a variational method: there exists a λ>0\lambda^{\ast}>0 such that if λ0\lambda\leq 0, then (1) has a unique solution; if 0<λ<λ0<\lambda<\lambda^{\ast}, then (1) has at least two distinct solutions; if λ=λ\lambda=\lambda^{\ast}, then (1) has at least one solution; if λ>λ\lambda>\lambda^{\ast}, then (1) has no solution. Recently, this result was partly reproved by Borer-Galimberti-Struwe [2] via a monotonicity technique due to Struwe [15, 16], and was extended to the case of conical metrics by Yang-Zhu [17].

Our aim is to extend results of Ding-Liu [4] to graphs. Let us recall some notations from graph theory. Throughout this paper, G=(V,E)G=(V,E) is assumed to be a finite connected graph. The edges on the graph are allowed to be weighted. Weights are given by a function ω:V×V[0,)\omega:V\times V\rightarrow[0,\infty), the edge xyxy from xx to yy has weight ωxy>0\omega_{xy}>0. We assume this weight function is symmetric, ωxy=ωyx\omega_{xy}=\omega_{yx}. Let μ:V+\mu:V\rightarrow\mathbb{R^{+}} be a positive measure on the vertices of the GG. Denote by VV^{\mathbb{R}} the space of real functions on VV. and by μp={fV:xVμ(x)|f(x)|p<}\ell^{p}_{\mu}=\{f\in V^{\mathbb{R}}:\sum_{x\in V}\mu(x)|f(x)|^{p}<\infty\}, for any 1p<1\leq p<\infty, the space of p\ell^{p} integrable functions on VV with respect to the measure μ\mu. For p=p=\infty, let ={fV:supxV|f(x)|<}\ell^{\infty}=\{f\in V^{\mathbb{R}}:\sup_{x\in V}|f(x)|<\infty\} be the set of all bounded functions. As usual, we define the μp\ell^{p}_{\mu} norm of fμp,1pf\in\ell^{p}_{\mu},1\leq p\leq\infty, by

fp=(xVμ(x)|f(x)|p)1/p,1p<,f=supxV|f(x)|.\|f\|_{p}=\left(\sum_{x\in V}\mu(x)|f(x)|^{p}\right)^{{1}/{p}},1\leq p<\infty,\,\|f\|_{\infty}=\sup_{x\in V}|f(x)|.

We define the Laplacian Δ:VV\Delta:V^{\mathbb{R}}\rightarrow V^{\mathbb{R}} on GG by

Δf(x)=1μ(x)yxωxy(f(x)f(y)).\Delta f(x)=\frac{1}{\mu(x)}\sum_{y\sim x}\omega_{xy}(f(x)-f(y)). (3)

Given the weight ω\omega on EE, there are two typical choices of Laplacian as follows:

  • 1.

    μ(x)=deg(x):=yxωxy\mu(x)=\deg(x):=\sum_{y\sim x}\omega_{xy} for all xVx\in V, which is called the normalized graph Laplacian;

  • 2.

    μ(x)1\mu(x)\equiv 1 for all xVx\in V, which is the combinatorial graph Laplacian.

In this paper, we do not restrict μ(x)\mu(x) to the above two forms, but only require μ(x)>0\mu(x)>0 for all xVx\in V. Note that the Laplace operator defined in (3) is the negative usual Laplace operator. The gradient form is defined by

2Γ(f,g)(x)\displaystyle 2\Gamma(f,g)(x) =\displaystyle= (fΔg+gΔfΔ(fg))(x)\displaystyle(f\cdot\Delta g+g\cdot\Delta f-\Delta(f\cdot g))(x)
=\displaystyle= 1μ(x)yxωxy(f(x)f(y))(g(x)g(y)).\displaystyle\frac{1}{\mu(x)}\sum_{y\sim x}\omega_{xy}(f(x)-f(y))(g(x)-g(y)).

For the sake of simplicity, we write Γ(f,f)=Γ(f)\Gamma(f,f)=\Gamma(f). Sometimes we use the notation fg=Γ(f,g)\nabla f\nabla g=\Gamma(f,g). The length of the gradient is denoted by

|f|(x)=Γ(f)(x).|\nabla f|(x)=\sqrt{\Gamma(f)}(x).

From now on, we write Vu𝑑μ=xVμ(x)u(x)\int_{V}ud\mu=\sum_{x\in V}\mu(x)u(x). Define a Sobolev space with a norm on the graph GG by

W1,2(V)={uV:V(|u|2+u2)𝑑μ<+},W^{1,2}(V)=\left\{u\in V^{\mathbb{R}}:\int_{V}(|\nabla u|^{2}+u^{2})d\mu<+\infty\right\},

and

uW1,2(V)=(V(|u|2+u2)𝑑μ)1/2\|u\|_{W^{1,2}(V)}=\left(\int_{V}(|\nabla u|^{2}+u^{2})d\mu\right)^{1/2}

respectively. Since GG is a finite graph, we have that W1,2(V)W^{1,2}(V) is exactly VV^{\mathbb{R}}, a finite dimensional linear space. This implies the following Sobolev embedding:

Lemma 1 ([7], Lemma 5).

If GG is a finite graph, then the Sobolev space W1,2(V)W^{1,2}(V) is precompact. Namely, if {uj}\{u_{j}\} is bounded in W1,2(V)W^{1,2}(V), then there exists some uW1,2(V)u\in W^{1,2}(V) such that up to a subsequence, ujuu_{j}\rightarrow u in W1,2(V)W^{1,2}(V).

The Kazdan-Warner equation we are interested in this paper reads as

Δu+κKλe2u=0onV,\Delta u+\kappa-K_{\lambda}e^{2u}=0\quad\mbox{on}\quad V, (4)

where κV\kappa\in V^{\mathbb{R}} is a function, and Kλ=K+λK_{\lambda}=K+\lambda, λ\lambda\in\mathbb{R}, KVK\in V^{\mathbb{R}} is a function. Now we are ready to state our main results.

Theorem 2.

Let G=(V,E)G=(V,E) be a finite graph, κ\kappa and KλK_{\lambda} be given as in (4) such that Vκ𝑑μ<0\int_{V}\kappa d\mu<0, KmaxVK=0K\leq\max_{V}K=0, and K0K\not\equiv 0. Then there exists a λ(0,minVK)\lambda^{*}\in(0,-\min_{V}K) satisfying

  1. 1.

    if λ0\lambda\leq 0, then (4) has a unique solution;

  2. 2.

    if 0<λ<λ0<\lambda<\lambda^{*}, then (4) has at least two distinct solutions;

  3. 3.

    if λ=λ\lambda=\lambda^{*}, then (4) has at least one solution;

  4. 4.

    if λ>λ\lambda>\lambda^{*}, then (4) has no solution;

Remark 1.

The assertion of λ<minVK\lambda^{*}<-\min_{V}K comes from the conclusion of Step 2 in Subsection 3.3.

Remark 2.

Compared to the existence of solutions in the literature (see for example [7, 5]), the above results firstly reveal the multiple solution problem of Kazdan-Warner equation on graphs in the negative case.

The proof of Theorem 2 is based on the method of variation. It can be viewed as a discrete analog of the result of Ding-Liu [4]. The remaining part of this paper will be organized as follows: In Section 2, we give several preliminary lemmas for our use later; In Section 3, we finish the proof of Theorem 2.

2 Preliminaries

In this section, we provide discrete versions of the maximum principle, the Palais-Smale condition and the upper and lower solution principle. Note that G=(V,E)G=(V,E) is a finite connected graph.

2.1 maximum principle

To proceed, we need the following maximum principles, which are known for experts (see for examples [7, 8]). For readers’ convenience, we include the detailed proofs here.

Lemma 3 (Weak maximum principle).

For any constant c>0c>0, if uu satisfies Δu+cu0\Delta u+cu\geq 0, then u0u\geq 0 on VV.

Proof.

Let u=min{u,0}u^{-}=\min\{u,0\}. For any xVx\in V, we claim that

Δu(x)+cu(x)0,\Delta u^{-}(x)+cu^{-}(x)\geq 0, (5)

from which, one has

VΓ(u)𝑑μ+cuμ22=u,Δu+cu0.\int_{V}\Gamma(u^{-})d\mu+c\|u^{-}\|^{2}_{\ell^{2}_{\mu}}=\langle u^{-},\Delta u^{-}+cu^{-}\rangle\leq 0.

This leads to u0u^{-}\equiv 0 on VV.

To prove this claim, we first consider the case u(x)0u(x)\geq 0. Therefore, cu(x)=0cu^{-}(x)=0 and

Δu(x)=1μ(x)yxωxy(u(x)u(y))=1μ(x)yxωxyu(y)0,\Delta u^{-}(x)=\frac{1}{\mu(x)}\sum_{y\sim x}\omega_{xy}(u^{-}(x)-u^{-}(y))=-\frac{1}{\mu(x)}\sum_{y\sim x}\omega_{xy}u^{-}(y)\geq 0,

due to u(z)0u^{-}(z)\leq 0 for any zVz\in V. In the case u(x)<0u(x)<0, one has cu(x)=cu(x)cu^{-}(x)=cu(x) and thus

Δu(x)=1μ(x)yxωxy(u(x)u(y))=1μ(x)yxωxy(u(x)u(y))1μ(x)yxωxy(u(x)u(y))=Δu(x).\begin{split}\Delta u^{-}(x)=\frac{1}{\mu(x)}\sum_{y\sim x}\omega_{xy}(u^{-}(x)-u^{-}(y))&=\frac{1}{\mu(x)}\sum_{y\sim x}\omega_{xy}(u(x)-u^{-}(y))\\ &\geq\frac{1}{\mu(x)}\sum_{y\sim x}\omega_{xy}(u(x)-u(y))=\Delta u(x).\end{split}

It follows that Δu(x)+cu(x)Δu(x)+cu(x)0\Delta u^{-}(x)+cu^{-}(x)\geq\Delta u(x)+cu(x)\geq 0, which confirms (5) and ends the proof of the lemma. ∎

Lemma 4 (Strong maximum principle).

Suppose that u0u\geq 0, and that Δu+cu0\Delta u+cu\geq 0 for some constant c>0c>0. If there exists x0Vx_{0}\in V such that u(x0)=0u(x_{0})=0, then u0u\equiv 0 on V.V.

Proof.

Let x=x0x=x_{0}, we have

1μ(x0)yx0ωyx0(u(x0)u(y))+cu(x0)0,\frac{1}{\mu(x_{0})}\sum_{y\sim x_{0}}\omega_{yx_{0}}(u(x_{0})-u(y))+cu(x_{0})\geq 0,

which implies

1μ(x0)yx0ωyx0u(y)0.\frac{1}{\mu(x_{0})}\sum_{y\sim x_{0}}\omega_{yx_{0}}u(y)\leq 0.

Since u0u\geq 0 and ωyx0>0\omega_{yx_{0}}>0 for all yx0y\sim x_{0}, we obtain

u(y)=0,for allyx0.u(y)=0,~{}~{}~{}~{}\mbox{for all}~{}~{}y\sim x_{0}.

Therefore, u0u\equiv 0 on VV by the connectedness of G.G.

2.2 Palais-Smale condition

We define a functional Eλ:W1,2(V)E_{\lambda}:W^{1,2}(V)\rightarrow\mathbb{R} by

Eλ(u)=V(|u|2+2κuKλe2u)𝑑μ,E_{\lambda}(u)=\int_{V}(|\nabla u|^{2}+2\kappa u-K_{\lambda}e^{2u})d\mu,

where κ\kappa and KλK_{\lambda} are given as in the assumptions of Theorem 2, in particular Vκ𝑑μ<0\int_{V}\kappa d\mu<0. For any ϕW1,2(V)\phi\in W^{1,2}(V), denote by dEλ(u)(ϕ)dE_{\lambda}(u)(\phi) the Frechet derivative of the functional, by dkEλ(u)(ϕ,,ϕ)d^{k}E_{\lambda}(u)(\phi,\cdots,\phi) the Frechet derivative of order k2k\geq 2.

Lemma 5 (Palais-Smale condition).

Suppose that Vλ={xV:Kλ(x)<0}V_{\lambda}^{-}=\{x\in V:K_{\lambda}(x)<0\} is nonempty for some λ\lambda\in\mathbb{R}. Then EλE_{\lambda} satisfies the (PS)c(PS)_{c} condition for all cc\in\mathbb{R}, i.e. if (uj)(u_{j}) is a sequence of functions in W1,2(V)W^{1,2}(V) such that Eλ(uj)cE_{\lambda}(u_{j})\rightarrow c and dEλ(uj)(ϕ)0dE_{\lambda}(u_{j})(\phi)\rightarrow 0 for all ϕW1,2(V)\phi\in W^{1,2}(V) as jj\rightarrow\infty, then there exists some u0W1,2(V)u_{0}\in W^{1,2}(V) satisfying uju0u_{j}\rightarrow u_{0} in W1,2(V).W^{1,2}(V).

Proof.

Let (uj)(u_{j}) be a function sequence such that Eλ(uj)cE_{\lambda}(u_{j})\rightarrow c and dEλ(uj)(ϕ)0,dE_{\lambda}(u_{j})(\phi)\rightarrow 0, or equivalently

V(|uj|2+2κujKλe2uj)𝑑μ=c+oj(1),\int_{V}(|\nabla u_{j}|^{2}+2\kappa u_{j}-K_{\lambda}e^{2u_{j}})d\mu=c+o_{j}(1), (6)
V(ujϕ+κϕKλe2ujϕ)𝑑μ=oj(1)ϕW1,2(V),ϕW1,2(V),\int_{V}(\nabla u_{j}\nabla\phi+\kappa\phi-K_{\lambda}e^{2u_{j}}\phi)d\mu=o_{j}(1)\|\phi\|_{W^{1,2}(V)},~{}~{}~{}~{}\forall\phi\in W^{1,2}(V), (7)

where oj(1)0o_{j}(1)\rightarrow 0 as j.j\rightarrow\infty.

Let ϕ1\phi\equiv 1 in (7), one has

V(κKλe2uj)𝑑μ=oj(1)μ(V)1/2,\int_{V}(\kappa-K_{\lambda}e^{2u_{j}})d\mu=o_{j}(1)\mu(V)^{1/2},

which implies

VKλe2uj𝑑μ=Vκ𝑑μ+oj(1).\int_{V}K_{\lambda}e^{2u_{j}}d\mu=\int_{V}\kappa d\mu+o_{j}(1). (8)

Inserting (8) into (6), we obtain

V(|uj|2+2κuj)𝑑μ=Vκ𝑑μ+c+oj(1).\int_{V}(|\nabla u_{j}|^{2}+2\kappa u_{j})d\mu=\int_{V}\kappa d\mu+c+o_{j}(1). (9)

We now claim that uju_{j} is bounded in μ2\ell^{2}_{\mu}. Suppose not, there holds ujμ2\|u_{j}\|_{\ell^{2}_{\mu}}\rightarrow\infty. We set vj=ujujμ2.v_{j}=\frac{u_{j}}{\|u_{j}\|_{\ell^{2}_{\mu}}}. By the Cauchy-Schwarz inequality, one has

Vκujujμ22𝑑μ=oj(1).\int_{V}\kappa\frac{u_{j}}{\|u_{j}\|^{2}_{\ell^{2}_{\mu}}}d\mu=o_{j}(1).

This together with (9) leads to

V|vj|2𝑑μ=oj(1).\int_{V}|\nabla v_{j}|^{2}d\mu=o_{j}(1). (10)

Hence, vjv_{j} is bounded in W1,2(V)W^{1,2}(V). In view of Lemma 1 and (10), vjγv_{j}\rightarrow\gamma in W1,2(V)W^{1,2}(V) for some constant γ.\gamma. Here and in the sequel, we do not distinguish sequence and subsequence. Since vjμ2=1\|v_{j}\|_{\ell^{2}_{\mu}}=1, we have γ0.\gamma\neq 0. It follows from (9) that

Vκvj𝑑μoj(1).\int_{V}\kappa v_{j}d\mu\leq o_{j}(1).

Passing to the limit jj\rightarrow\infty in the above inequality, we conclude that γ0\gamma\geq 0 since Vκ𝑑μ<0\int_{V}\kappa d\mu<0. Therefore γ>0\gamma>0.

On the other hand, for any xVλx\in V_{\lambda}^{-}, if there exists NN\in\mathbb{N}, if j>Nj>N such that uj(x)0u_{j}(x)\leq 0, then limjvj(x)0\lim_{j\rightarrow\infty}v_{j}(x)\leq 0, which contradicts γ>0\gamma>0 and confirms our claim. If not, let xVλx_{*}\in V_{\lambda}^{-}, due to the finiteness of VV, we can choose a subsequence {jk}k=0\{j_{k}\}_{k=0}^{\infty} such that ujk(x)>0u_{j_{k}}(x_{*})>0. Set

ϕ(x)={ujk(x),x=x0,xx.\phi(x)=\left\{\begin{array}[]{ll}u_{j_{k}}(x_{*}),&\hbox{$x=x_{*}$}\\[5.16663pt] 0,&x\not=x_{\ast}.\end{array}\right.

Then

ϕW1,2(V)2=2yxωxyujk2(x)+μ(x)ujk2(x)=(2deg(x)+μ(x))ujk2(x).\|\phi\|_{W^{1,2}(V)}^{2}=2\sum_{y\sim x_{*}}\omega_{x_{*}y}u_{j_{k}}^{2}(x_{*})+\mu(x_{*})u_{j_{k}}^{2}(x_{*})=(2\deg(x_{*})+\mu(x_{*}))u_{j_{k}}^{2}(x_{*}).

Substituting it into (7), we have

Δujk(x)+κ(x)Kλ(x)e2ujk(x)C.\Delta u_{j_{k}}(x_{*})+\kappa(x_{*})-K_{\lambda}(x_{*})e^{2u_{j_{k}}(x_{*})}\leq C^{\prime}. (11)

Since vjγv_{j}\rightarrow\gamma, ujμ2+\|u_{j}\|_{\ell^{2}_{\mu}}\rightarrow+\infty and VV has finite points, we conclude

uj=(γ+oj(1))ujμ2uniformlyonV.u_{j}=(\gamma+o_{j}(1))\|u_{j}\|_{\ell^{2}_{\mu}}\quad{\rm uniformly\,\,on}\quad V.

This together with (11) leads to

ujkμ22ojk(1)+κ(x)Kλ(x)e2(γ+ojk(1))ujkμ22C,\|u_{j_{k}}\|_{\ell^{2}_{\mu}}^{2}o_{j_{k}}(1)+\kappa(x_{*})-K_{\lambda}(x_{*})e^{2(\gamma+o_{j_{k}}(1))\|u_{j_{k}}\|_{\ell^{2}_{\mu}}^{2}}\leq C^{\prime},

which is impossible since Kλ(x)<0K_{\lambda}(x_{\ast})<0. Then our claim follows immediately.

Since uju_{j} is bounded in μ2\ell_{\mu}^{2}, we have uju_{j} is bounded in W1,2(V)W^{1,2}(V) due to the finiteness of VV. Therefore, by Lemma 1, there exists some u0W1,2(V)u_{0}\in W^{1,2}(V) such that up to subsequence, uju0u_{j}\rightarrow u_{0} in W1,2(V)W^{1,2}(V). ∎

2.3 Upper and lower solutions principle

Let f:V×f:V\times\mathbb{R}\rightarrow\mathbb{R} be a function, and ff is smooth with respect to the second variable. We say that uVu\in V^{\mathbb{R}} is an upper (lower) solution to the following equation

Δu(x)+f(x,u(x))=0,xV,\Delta u(x)+f(x,u(x))=0,~{}~{}x\in V, (12)

if uu satisfies Δu(x)+f(x,u(x))()0\Delta u(x)+f(x,u(x))\geq(\leq)~{}0 for any xVx\in V. We generalize ([7], Lemma 8) to the following:

Lemma 6.

Suppose that φ,ψ\varphi,\psi are lower and upper solution to (12) respectively with φψ\varphi\leq\psi on VV. Then (12) has a solution uu with φuψ\varphi\leq u\leq\psi on VV.

Proof.

This is a discrete version of the argument of Kazdan-Warner ([12], Lemma 9.3), and the method of proof carries over to the setting of graphs.

Since the graph is finite, there exists a constant AA such that AφψA-A\leq\varphi\leq\psi\leq A. One can find a sufficient large constant cc such that F(x,t)=ctf(x,t)F(x,t)=ct-f(x,t) is increasing with respect to t[A,A]t\in[-A,A] for any fixed xVx\in V. We define an operator Lu=Δu+cuLu=\Delta u+cu, and LL is a compact operator and Ker(L)=span{1}\mbox{Ker}(L)=\mbox{span}\{1\} due to the finiteness of the graph. Hence, we can define φj+1,ψj+1\varphi_{j+1},\psi_{j+1} inductively as the unique solution to

φ0=φ,Lφj+1(x)=cφj(x)f(x,φj(x)),j0,xV,\varphi_{0}=\varphi,~{}~{}L\varphi_{j+1}(x)=c\varphi_{j}(x)-f(x,\varphi_{j}(x)),~{}~{}\forall j\geq 0,x\in V,
ψ0=ψ,Lψj+1(x)=cψj(x)f(x,ψj(x)),j0,xV\psi_{0}=\psi,~{}~{}L\psi_{j+1}(x)=c\psi_{j}(x)-f(x,\psi_{j}(x)),~{}~{}\forall j\geq 0,x\in V

respectively. Combining with the definition of upper (lower) solution and the monotonicity of F(x,t)F(x,t) with respect to tt, we obtain

Lφ0(x)Lφ1(x)=F(x,φ(x))F(x,ψ(x))=Lψ1(x)Lψ(x),xV.L\varphi_{0}(x)\leq L\varphi_{1}(x)=F(x,\varphi(x))\leq F(x,\psi(x))=L\psi_{1}(x)\leq L\psi(x),~{}~{}x\in V.

Then the weak maximum principle (see Lemma 3) yields that

φφ1ψ1ψ.\varphi\leq\varphi_{1}\leq\psi_{1}\leq\psi.

Moreover, it turns out that φ1\varphi_{1} and ψ1\psi_{1} are lower and upper solution to (12) respectively. By induction, we have

φφjφj+1ψj+1ψjψ,j=1,2,.\varphi\leq\varphi_{j}\leq\varphi_{j+1}\leq\psi_{j+1}\leq\psi_{j}\leq\psi,~{}~{}j=1,2,\cdots.

Since VV is finite, it is easy to see that up to a subsequence, φju1,ψju2\varphi_{j}\rightarrow u_{1},\psi_{j}\rightarrow u_{2} uniformly on VV, and u=u1u=u_{1} or u2u_{2} is a solution to (12) with φuψ\varphi\leq u\leq\psi on VV. ∎

3 Proof of Theorem 2

3.1 Unique solution in the case λ0\lambda\leq 0.

Claim 1.

EλE_{\lambda} is strictly convex on W1,2(V)W^{1,2}(V).

Proof.

We only need to show that there exists some constant C>0C>0 such that

d2Eλ(u)(h,h)ChW1,2(V)2,u,hW1,2(V).d^{2}E_{\lambda}(u)(h,h)\geq C\|h\|^{2}_{W^{1,2}(V)},~{}~{}\forall u,h\in W^{1,2}(V). (13)

Suppose not, there would be a function uW1,2(V)u\in W^{1,2}(V) and a function sequence hjW1,2(V)h_{j}\in W^{1,2}(V) such that hjW1,2(V)=1\|h_{j}\|_{W^{1,2}(V)}=1 for all jj and d2Eλ(u)(hj,hj)0d^{2}E_{\lambda}(u)(h_{j},h_{j})\rightarrow 0 as jj\rightarrow\infty. From Lemma 1, there exists hW1,2(V)h_{\infty}\in W^{1,2}(V), such that up to a subsequence, hjhh_{j}\rightarrow h_{\infty} as jj\rightarrow\infty in W1,2(V)W^{1,2}(V). Since

d2Eλ(u)(hj,hj)=2V(|hj|22Kλe2uhj2)𝑑μ,d^{2}E_{\lambda}(u)(h_{j},h_{j})=2\int_{V}(|\nabla h_{j}|^{2}-2K_{\lambda}e^{2u}h_{j}^{2})d\mu,

and Kλ0K_{\lambda}\leq 0, it follows that V|hj|2𝑑μ0\int_{V}|\nabla h_{j}|^{2}d\mu\rightarrow 0 and VKλe2uhj2𝑑μ0\int_{V}K_{\lambda}e^{2u}h_{j}^{2}d\mu\rightarrow 0, which lead to hch_{\infty}\equiv c for some constant cc, and moreover

c2VKλe2u𝑑μ=limjVKλe2uhj2𝑑μ=0.c^{2}\int_{V}K_{\lambda}e^{2u}d\mu=\lim_{j\rightarrow\infty}\int_{V}K_{\lambda}e^{2u}h_{j}^{2}d\mu=0.

It is easily seen that VKλe2u𝑑μ<0\int_{V}K_{\lambda}e^{2u}d\mu<0 by K0K\not\equiv 0, thus c=0c=0. This contradicts

hW1,2(V)=limjhjW1,2(V)=1.\|h_{\infty}\|_{W^{1,2}(V)}=\lim_{j\rightarrow\infty}\|h_{j}\|_{W^{1,2}(V)}=1.

Hence (13) holds. ∎

Claim 2.

For any ε>0\varepsilon>0, there exist constants C,C(ε)>0C,C(\varepsilon)>0 such that

Eλ(u)(C2ε)uW1,2(V)22C(ε).E_{\lambda}(u)\geq(C-2\varepsilon)\|u\|_{W^{1,2}(V)}^{2}-2C(\varepsilon).
Proof.

By Young’s inequality, for any ε>0\varepsilon>0, there exists a constant C(ε)>0C(\varepsilon)>0 such that

|Vκu𝑑μ|εuW1,2(V)2+C(ε).\left|\int_{V}\kappa ud\mu\right|\leq\varepsilon\|u\|_{W^{1,2}(V)}^{2}+C(\varepsilon).

Thus, it is sufficient to find some constant C>0C>0 such that for all uW1,2(V)u\in W^{1,2}(V)

V(|u|2Kλe2u)𝑑μCuW1,2(V)2.\int_{V}(|\nabla u|^{2}-K_{\lambda}e^{2u})d\mu\geq C\|u\|^{2}_{W^{1,2}(V)}.

Suppose not, there would exist a sequence of functions uju_{j} satisfying

V(|uj|2+uj2)𝑑μ=1,V(|uj|2Kλe2uj)𝑑μ=oj(1).\int_{V}(|\nabla u_{j}|^{2}+u_{j}^{2})d\mu=1,~{}~{}\int_{V}(|\nabla u_{j}|^{2}-K_{\lambda}e^{2u_{j}})d\mu=o_{j}(1).

Clearly, uju_{j} is bounded in W1,2(V)W^{1,2}(V), it follows from Lemma 1 that there exists some function uW1,2(V)u\in W^{1,2}(V) such that up to a subsequence, ujuu_{j}\rightarrow u in W1,2(V)W^{1,2}(V) as j.j\rightarrow\infty. Due to Kλ0K_{\lambda}\not\equiv 0 and Kλ0,K_{\lambda}\leq 0, we have

0<V(|u|2Kλe2u)𝑑μ=limjV(|uj|2Kλe2uj)𝑑μ=0,0<\int_{V}(|\nabla u|^{2}-K_{\lambda}e^{2u})d\mu=\lim_{j\rightarrow\infty}\int_{V}(|\nabla u_{j}|^{2}-K_{\lambda}e^{2u_{j}})d\mu=0,

which gets a contradiction. ∎

Proof of (1) in Theorem 2.   It is a consequence of Claim 1 and Claim 2. Precisely we denote Λ=infuW1,2(V)Eλ(u)\Lambda=\inf_{u\in W^{1,2}(V)}E_{\lambda}(u). By Claim 2, we see that Λ\Lambda is a definite real number. Take a function sequence ujW1,2(V)u_{j}\in W^{1,2}(V) such that Eλ(uj)ΛE_{\lambda}(u_{j})\rightarrow\Lambda as jj\rightarrow\infty. Applying Claim 2, we have that uju_{j} is bounded in W1,2(V)W^{1,2}(V). Then, in view of Lemma 1, there exists a subsequence of uju_{j} (still denoted by uju_{j}) and a function u0W1,2(V)u_{0}\in W^{1,2}(V) such that uju0u_{j}\rightarrow u_{0} in W1,2(V)W^{1,2}(V). Obviously Eλ(u0)=ΛE_{\lambda}(u_{0})=\Lambda, and thus u0u_{0} is a critical point of EλE_{\lambda}. We also need to explain why EλE_{\lambda} has only one critical point. For otherwise, we assume uu^{\ast} is another critical point of EλE_{\lambda}. Note that dEλ(u0)=dEλ(u)=0dE_{\lambda}(u_{0})=dE_{\lambda}(u^{\ast})=0. It follows from Claim 1, particularly from (13), that

Eλ(u0)\displaystyle E_{\lambda}(u_{0}) =\displaystyle= Eλ(u)+dEλ(u)(u0u)+12d2Eλ(ξ)(u0u,u0u)\displaystyle E_{\lambda}(u^{\ast})+dE_{\lambda}(u^{\ast})(u_{0}-u^{\ast})+\frac{1}{2}d^{2}E_{\lambda}(\xi)(u_{0}-u^{\ast},u_{0}-u^{\ast})
\displaystyle\geq Eλ(u)+Cu0uW1,2(V)2\displaystyle E_{\lambda}(u^{\ast})+C\|u_{0}-u^{\ast}\|_{W^{1,2}(V)}^{2}

for some positive constant CC, where ξ\xi is a function lies between uu^{\ast} and u0u_{0}. Hence we have Eλ(u0)>Eλ(u)E_{\lambda}(u_{0})>E_{\lambda}(u^{\ast}), contradicting the fact that Eλ(u0)=ΛE_{\lambda}(u_{0})=\Lambda. This implies the uniqueness of the critical point of EλE_{\lambda}. \hfill\Box

3.2 Multiplicity of solutions for 0<λ<λ0<\lambda<\lambda^{*}.

Fixing λ(0,λ)\lambda\in(0,\lambda^{*}), we will seek two different solutions of (4). One is a strict local minimum of the functional EλE_{\lambda}, and the other is from mountain-pass theorem. We firstly prove the existence of λ\lambda^{*}. Consider the case λ=0\lambda=0 in the equation (4) as follows

Δu+κKe2u=0.\Delta u+\kappa-Ke^{2u}=0. (14)

For the solution u0u_{0} of (14), the linearized equation of (14) at u0u_{0}

Δv2Ke2u0v=0\Delta v-2Ke^{2u_{0}}v=0

has only a trivial solution v0,v\equiv 0, since K0K\leq 0 and K0.K\not\equiv 0. Indeed,

0V|v|2𝑑μ=VvΔv𝑑μ=2VKe2u0v2𝑑μ0,0\leq\int_{V}|\nabla v|^{2}d\mu=\int_{V}v\Delta vd\mu=2\int_{V}Ke^{2u_{0}}v^{2}d\mu\leq 0,

which implies v(x)=0v(x)=0 when K(x)<0.K(x)<0. In the case K(x)=0K(x)=0, we have Δv(x)=0\Delta v(x)=0. Thus

V|v|2𝑑μ=VvΔv𝑑μ=0.\int_{V}|\nabla v|^{2}d\mu=\int_{V}v\Delta vd\mu=0.

It follows that vv is a constant function and hence v0.v\equiv 0. By the implicit function theorem, there exists a small enough s>0s>0 such that the equation (4) has a solution for any λ(0,s)\lambda\in(0,s). Indeed, let u=tv+u0,v0u=tv+u_{0},v\not\equiv 0 on VV, we consider G(λ,t)=Δu+κKλeuG(\lambda,t)=\Delta u+\kappa-K_{\lambda}e^{u}. It is easy to see that G(0,0)=0G(0,0)=0, G(λ,t)G(\lambda,t) and tG(λ,t)=Δv2v(K+λ)e2(u0+tv)\partial_{t}G(\lambda,t)=\Delta v-2v(K+\lambda)e^{2(u_{0}+tv)} are continuous on any domain D2D\subset\mathbb{R}^{2}, Furthermore, tG(0,0)=Δv2Ke2u0v0\partial_{t}G(0,0)=\Delta v-2Ke^{2u_{0}}v\not\equiv 0 unless v0v\equiv 0 on VV. Therefore, by the implicit function theorem, there exists s>0s>0 such that t=g(λ)t=g(\lambda) and G(λ,g(λ))=0G(\lambda,g(\lambda))=0 for any λ(0,s)\lambda\in(0,s). In other words, uλ=g(λ)v+u0,λ(0,s)u_{\lambda}=g(\lambda)v+u_{0},\forall\lambda\in(0,s) is the solution of (4). Define

λ=sup{s:the equation (4) has a solution for any λ(0,s)}.\lambda^{*}=\sup\{s:\mbox{the equation \eqref{eq:0} has a solution for any $\lambda\in(0,s)$}\}.

One can see that λminVK.\lambda^{*}\leq-\min_{V}K. For otherwise, Kλ=K+λ>0K_{\lambda}=K+\lambda>0 for some λ<λ\lambda<\lambda^{*}. Adding up the equation (4) for all xVx\in V, we have

0>Vκ𝑑μ=VKλe2u𝑑μ>0,0>\int_{V}\kappa d\mu=\int_{V}K_{\lambda}e^{2u}d\mu>0,

which is impossible. In conclusion, we have 0<λminVK0<\lambda^{*}\leq-\min_{V}K.

Proof of (2) in Theorem 2. We separate the proof into the following three steps.

Step 1. The existence of the upper and lower solution of (4).

Take λ1\lambda_{1} with λ<λ1<λ\lambda<\lambda_{1}<\lambda^{*}, let uλ1u_{\lambda_{1}} be a solution of (4) at λ1\lambda_{1}. It is easily seen that ψ=uλ1\psi=u_{\lambda_{1}} is a strict upper solution of (4) at λ\lambda, namely

Δψ+κKλe2ψ>0.\Delta\psi+\kappa-K_{\lambda}e^{2\psi}>0.

Let vv be the solution to the following equation

Δv=κ+1μ(V)Vκ𝑑μ.\Delta v=-\kappa+\frac{1}{\mu(V)}\int_{V}\kappa d\mu. (15)

The existence of solution to (15) was proved in [7]. Set φ=vs\varphi=v-s, where ss is a sufficiently large constant such that φ<ψ\varphi<\psi on VV and

Δφ+κKλe2φ=1μ(V)Vκ𝑑μKλe2v2s<0\Delta\varphi+\kappa-K_{\lambda}e^{2\varphi}=\frac{1}{\mu(V)}\int_{V}\kappa d\mu-K_{\lambda}e^{2v-2s}<0

since Vκ𝑑μ<0\int_{V}\kappa d\mu<0. Therefore, φ\varphi is a strict lower solution of (4). Let [φ,ψ][\varphi,\psi] be the order interval defined by

[φ,ψ]={uV:φuψonV}.[\varphi,\psi]=\{u\in V^{\mathbb{R}}:\varphi\leq u\leq\psi~{}\mbox{on}~{}V\}.

The upper and lower-solution method (Lemma 6) asserts that (4) has a solution uλ[φ,ψ]u_{\lambda}\in[\varphi,\psi] on VV.

Step 2. uλu_{\lambda} can be chosen as a strict local minimum of EλE_{\lambda}.

Let fλ(x,t)=ctκ(x)+Kλ(x)e2tf_{\lambda}(x,t)=ct-\kappa(x)+K_{\lambda}(x)e^{2t}, where cc is sufficiently large such that fλ(x,t)f_{\lambda}(x,t) is increasing in t[A,A]t\in[-A,A], AA is a constant such that Aφ<ψA-A\leq\varphi<\psi\leq A on VV. Let Fλ(x,u(x))=0u(x)fλ(x,t)𝑑tF_{\lambda}(x,u(x))=\int_{0}^{u(x)}f_{\lambda}(x,t)dt. It is easy to rewrite Eλ(u)E_{\lambda}(u) as

Eλ(u)=V(|u|2+cu2)𝑑μ2xVμ(x)Fλ(x,u(x))VKλ𝑑μ.E_{\lambda}(u)=\int_{V}(|\nabla u|^{2}+cu^{2})d\mu-2\sum_{x\in V}\mu(x)F_{\lambda}(x,u(x))-\int_{V}K_{\lambda}d\mu.

It is obvious that EλE_{\lambda} is bounded from below on [φ,ψ][\varphi,\psi]. Therefore, we denote

a:=infu[φ,ψ]Eλ(u).a:=\inf_{u\in[\varphi,\psi]}E_{\lambda}(u).

Taking a function sequence uj[φ,ψ]u_{j}\subset[\varphi,\psi] such that Eλ(uj)aE_{\lambda}(u_{j})\rightarrow a as j.j\rightarrow\infty. From it, we can get that uju_{j} is bounded in W1,2(V)W^{1,2}(V), and thus up to subsequence, uju_{j} converges to some uλu_{\lambda} in W1,2(V)W^{1,2}(V) and μq\ell^{q}_{\mu} for any q1q\geq 1, and e2uje^{2u_{j}} converges to e2uλe^{2u_{\lambda}} in μ1\ell^{1}_{\mu}. Hence

Eλ(uλ)=infu[φ,ψ]Eλ(u).E_{\lambda}(u_{\lambda})=\inf_{u\in[\varphi,\psi]}E_{\lambda}(u).

As a consequence, uλu_{\lambda} satisfies the Euler-Lagrange equation

Δuλ(x)+cuλ(x)=fλ(x,uλ(x)).\Delta u_{\lambda}(x)+cu_{\lambda}(x)=f_{\lambda}(x,u_{\lambda}(x)).

From it, one can conclude that

φ(x)<uλ(x)<ψ(x),xV.\varphi(x)<u_{\lambda}(x)<\psi(x),\forall x\in V. (16)

Indeed, noting that fλ(x,t)f_{\lambda}(x,t) is increasing with respect to t[A,A]t\in[-A,A], we have

Δφ(x)+cφ(x)fλ(x,φ(x))fλ(x,ψ(x))Δψ(x)+cψ(x).\Delta\varphi(x)+c\varphi(x)\leq f_{\lambda}(x,\varphi(x))\leq f_{\lambda}(x,\psi(x))\leq\Delta\psi(x)+c\psi(x).

One can conclude (16) by the strong maximum principle (Lemma 4), and the fact φ<ψ\varphi<\psi. For any hW1,2(V)h\in W^{1,2}(V), we define a function η(t)=Eλ(uλ+th),t\eta(t)=E_{\lambda}(u_{\lambda}+th),t\in\mathbb{R}. There holds φuλ+thψ\varphi\leq u_{\lambda}+th\leq\psi for sufficiently small |t||t|. Since uλu_{\lambda} is a minimum of EλE_{\lambda} on (φ,ψ)(\varphi,\psi), we have η(0)=dEλ(uλ)(h)=0\eta^{\prime}(0)=dE_{\lambda}(u_{\lambda})(h)=0 and η′′(0)=d2Eλ(uλ)(h,h)0.\eta^{\prime\prime}(0)=d^{2}E_{\lambda}(u_{\lambda})(h,h)\geq 0. Furthermore, there exists a constant C>0C>0 such that

d2Eλ(uλ)(h,h)ChW1,2(V),hW1,2(V),d^{2}E_{\lambda}(u_{\lambda})(h,h)\geq C\|h\|_{W^{1,2}(V)},~{}~{}~{}~{}\forall h\in W^{1,2}(V), (17)

which implies uλu_{\lambda} is strict local minimum of EλE_{\lambda} on W1,2(V)W^{1,2}(V). It remains to prove (17). We first denote

θ:=infhW1,2(V)=1d2Eλ(uλ)(h,h),\theta:=\inf_{\|h\|_{W^{1,2}(V)}=1}d^{2}E_{\lambda}(u_{\lambda})(h,h),

which is nonnegative. It is sufficient to prove θ>0\theta>0, (17) follows. Suppose θ=0\theta=0, we claim that there exists some h~\tilde{h} with h~W1,2(V)=1\|\tilde{h}\|_{W^{1,2}(V)}=1 such that d2Eλ(uλ)(h~,h~)=0d^{2}E_{\lambda}(u_{\lambda})(\tilde{h},\tilde{h})=0. To see this, let hjh_{j} be a function sequence satisfying hjW1,2(V)=1\|h_{j}\|_{W^{1,2}(V)}=1 for all jj and d2Eλ(uλ)(hj,hj)0d^{2}E_{\lambda}(u_{\lambda})(h_{j},h_{j})\rightarrow 0 as j.j\rightarrow\infty. Up to subsequence, hjh~h_{j}\rightarrow\tilde{h} in W1,2(V)W^{1,2}(V) from Lemma 1, and confirms our claim. To put it another way, the functional vd2Eλ(uλ)(v,v)v\mapsto d^{2}E_{\lambda}(u_{\lambda})(v,v) attains its minimum at v=h~v=\tilde{h}, it follows that d2Eλ(uλ)(h~,v)=0d^{2}E_{\lambda}(u_{\lambda})(\tilde{h},v)=0 for all vW1,2(V).v\in W^{1,2}(V). Hence, h~\tilde{h} is a solution of the following equation

Δh=2Kλe2uλh,λ(0,λ).\Delta h=2K_{\lambda}e^{2u_{\lambda}}h,~{}~{}~{}~{}\lambda\in(0,\lambda^{*}). (18)

It is easy to see that h~\tilde{h} is not a constant. For otherwise (18) yields

0>Vκ𝑑μ=VKλe2uλ𝑑μ=0,0>\int_{V}\kappa d\mu=\int_{V}K_{\lambda}e^{2u_{\lambda}}d\mu=0,

which is impossible. Multiplying (18) by h~3\tilde{h}^{3}, we obtain

d4Eλ(uλ)(h~,h~,h~,h~)=16VKλe2uλh~4𝑑μ=8Vh~3Δh~𝑑μ=8x,yVωxy(h~3(x)h~3(y))(h~(x)h~(y))=8x,yVωxy(h~2(x)+h~(x)h~(y)+h~2(y))(h~(x)h~(y))2<0.\begin{split}d^{4}E_{\lambda}(u_{\lambda})(\tilde{h},\tilde{h},\tilde{h},\tilde{h})&=-16\int_{V}K_{\lambda}e^{2u_{\lambda}}\tilde{h}^{4}d\mu\\ &=-8\int_{V}\tilde{h}^{3}\Delta\tilde{h}d\mu\\ &=-8\sum_{x,y\in V}\omega_{xy}(\tilde{h}^{3}(x)-\tilde{h}^{3}(y))(\tilde{h}(x)-\tilde{h}(y))\\ &=-8\sum_{x,y\in V}\omega_{xy}(\tilde{h}^{2}(x)+\tilde{h}(x)\tilde{h}(y)+\tilde{h}^{2}(y))(\tilde{h}(x)-\tilde{h}(y))^{2}<0.\end{split}

The last inequality is due to the fact a2+ab+b2>0a^{2}+ab+b^{2}>0 for any a,ba,b\in\mathbb{R} satisfying ab0ab\neq 0. Since d2Eλ(uλ+th~)(h~,h~)d^{2}E_{\lambda}(u_{\lambda}+t\tilde{h})(\tilde{h},\tilde{h}) attains its minimum at t=0t=0, we have d3Eλ(uλ)(h~,h~,h~)=0d^{3}E_{\lambda}(u_{\lambda})(\tilde{h},\tilde{h},\tilde{h})=0, which together with dEλ(uλ)(h~)=0dE_{\lambda}(u_{\lambda})(\tilde{h})=0 and d2Eλ(uλ)(h~,h~)=0d^{2}E_{\lambda}(u_{\lambda})(\tilde{h},\tilde{h})=0 leads to

Eλ(uλ+ϵh~)=Eλ(uλ)+ϵ424d4Eλ(uλ)(h~,h~,h~,h~)+0(ϵ5)<Eλ(uλ)E_{\lambda}(u_{\lambda}+\epsilon\tilde{h})=E_{\lambda}(u_{\lambda})+\frac{\epsilon^{4}}{24}d^{4}E_{\lambda}(u_{\lambda})(\tilde{h},\tilde{h},\tilde{h},\tilde{h})+0(\epsilon^{5})<E_{\lambda}(u_{\lambda}) (19)

for small ϵ>0.\epsilon>0. Let ϵ\epsilon small enough such that φuλ+ϵh~ψ\varphi\leq u_{\lambda}+\epsilon\tilde{h}\leq\psi, thus by (19),

Eλ(uλ+ϵh~)<Eλ(uλ),E_{\lambda}(u_{\lambda}+\epsilon\tilde{h})<E_{\lambda}(u_{\lambda}),

which contradicts the fact that uλu_{\lambda} is the minimum of EλE_{\lambda} on [φ,ψ][\varphi,\psi]. Therefore θ>0\theta>0, which concludes (17).

Step 3. The second solution of (4) is given by the mountain-pass theorem.

We shall use the mountain-pass theorem due to Ambrosetti and Rabinowitz [1], which reads as follows: Let (X,)(X,\|\cdot\|) be a Banach space, JC1(X,)J\in C^{1}(X,\mathbb{R}), e0,eXe_{0},e\in X and r>0r>0 be such that ee0>r\|e-e_{0}\|>r and

b:=infue0=rJ(u)>J(e0)J(e).b:=\inf_{\|u-e_{0}\|=r}J(u)>J(e_{0})\geq J(e).

If JJ satisfies the (PS)c(PS)_{c} condition with c:=infγΓmaxt[0,1]J(γ(t))c:=\inf_{\gamma\in\Gamma}\max_{t\in[0,1]}J(\gamma(t)), where

Γ:={γC([0,1],X):γ(0)=e0,γ(1)=e},\Gamma:=\{\gamma\in C([0,1],X):\gamma(0)=e_{0},\gamma(1)=e\},

then cc is a critical value of JJ. In our case, W1,2(V)W^{1,2}(V) is a Banach space, and Eλ:W1,2(V)E_{\lambda}:W^{1,2}(V)\rightarrow\mathbb{R} is a smooth functional.

Since uλu_{\lambda} is a strict local minimum of EλE_{\lambda} on W1,2(V)W^{1,2}(V), there exists a small enough number r>0r>0 such that

infuuλW1,2(V)=rEλ(u)>Eλ(uλ).\inf_{\|u-u_{\lambda}\|_{W^{1,2}(V)}=r}E_{\lambda}(u)>E_{\lambda}(u_{\lambda}). (20)

Moreover, for any λ>0,\lambda>0, EλE_{\lambda} has no lower bound on W1,2(V)W^{1,2}(V), namely, there exists vW1,2(V)v\in W^{1,2}(V) such that

Eλ(v)<Eλ(uλ),vuλW1,2(V)>r.E_{\lambda}(v)<E_{\lambda}(u_{\lambda}),~{}~{}~{}~{}\|v-u_{\lambda}\|_{W^{1,2}(V)}>r. (21)

To see this, we set Vε={xV:Kλ(x)>ε}V_{\varepsilon}=\{x\in V:K_{\lambda}(x)>\varepsilon\} for small ε>0\varepsilon>0. Note that VεV_{\varepsilon} is nonempty since maxVK=0\max_{V}K=0 and λ>0\lambda>0. Let fW1,2(V)f\in W^{1,2}(V) be a function which equals to 11 in VεV_{\varepsilon} and vanishes on V/VεV/V_{\varepsilon}, then

Eλ(tf)=t2V|f|2𝑑μ+tVεκ𝑑μVεKλe2t𝑑μV/VεKλ𝑑μAt2+Bt+Cεμ(Vε)e2t,t+.\begin{split}E_{\lambda}(tf)&=t^{2}\int_{V}|\nabla f|^{2}d\mu+t\int_{V_{\varepsilon}}\kappa d\mu-\int_{V_{\varepsilon}}K_{\lambda}e^{2t}d\mu-\int_{V/V_{\varepsilon}}K_{\lambda}d\mu\\ &\leq At^{2}+Bt+C-\varepsilon\mu(V_{\varepsilon})e^{2t}\rightarrow-\infty,~{}~{}~{}~{}t\rightarrow+\infty.\end{split}

In view of (20), (21) and Lemma 5, the mountain pass theorem of Ambrosetti and Rabinowitz gives another critical point uλu^{\lambda} of EλE_{\lambda} other than uλu_{\lambda}. In particular,

Eλ(uλ)=minγΓmaxuγEλ(u),E_{\lambda}(u^{\lambda})=\min_{\gamma\in\Gamma}\max_{u\in\gamma}E_{\lambda}(u),

where Γ={γC([0,1],W1,2(V)):γ(0)=uλ,γ(1)=v}\Gamma=\{\gamma\in C([0,1],W^{1,2}(V)):\gamma(0)=u_{\lambda},\gamma(1)=v\}, and dEλ(uλ)(h)=0dE_{\lambda}(u^{\lambda})(h)=0 for any hW1,2(V)h\in W^{1,2}(V). \hfill\Box

3.3 Solvability at λ\lambda^{*}.

For any λ,0<λ<λ\lambda,0<\lambda<\lambda^{*}, let uλu_{\lambda} be the local minimum of EλE_{\lambda} obtained in the previous subsection. That is, uλu_{\lambda} is the solution of (4), and

d2Eλ(uλ)(h,h)=2V(|h|22Kλe2uλh2)𝑑μ0,hW1,2(V).d^{2}E_{\lambda}(u_{\lambda})(h,h)=2\int_{V}(|\nabla h|^{2}-2K_{\lambda}e^{2u_{\lambda}}h^{2})d\mu\geq 0,~{}~{}~{}~{}\forall h\in W^{1,2}(V). (22)

Proof of (3) in Theorem 2. The crucial point in this proof is to show that uλu_{\lambda} is uniformly bounded in W1,2(V)W^{1,2}(V) as λλ.\lambda\rightarrow\lambda^{*}. If it is true, then up to subsequence, uλu_{\lambda} converges to some uu in W1,2(V)W^{1,2}(V), and uu is the solution of

Δu+κKλe2u=0.\Delta u+\kappa-K_{\lambda^{*}}e^{2u}=0.

Hence, we aim to prove the W1,2(V)W^{1,2}(V) boundedness of uλ.u_{\lambda}. To this end, we divide the proof into the following three steps.

Step 1. There exists a constant C>0C>0 such that

uλCon Vu_{\lambda}\geq-C~{}~{}~{}~{}\mbox{on $V$} (23)

uniformly for any 0<λ<λ0<\lambda<\lambda^{*}.

Let vv satisfy (15), and φs=vs\varphi_{s}=v-s for s>0s>0. Then for sufficient large ss, say ss0s\geq s_{0}, φs\varphi_{s} is a continuous family with respect to ss of strict lower solution of (4) at λ=0\lambda=0, i.e.

Δφs+κKe2φs<0.\Delta\varphi_{s}+\kappa-Ke^{2\varphi_{s}}<0.

It is clear that φs\varphi_{s} is also a strict lower solution of (4) at λ(0,λ)\lambda\in(0,\lambda^{*}) for any s[s0,)s\in[s_{0},\infty). Now, we prove that uλφs0u_{\lambda}\geq\varphi_{s_{0}}, and thus (23) holds. For otherwise, we can find for some s(s0,)s\in(s_{0},\infty)

uλφson V,anduλ(x~)=φs(x~)for some x~V.u_{\lambda}\geq\varphi_{s}~{}~{}~{}\mbox{on $V$},~{}~{}~{}~{}\mbox{and}~{}~{}u_{\lambda}(\tilde{x})=\varphi_{s}(\tilde{x})~{}~{}\mbox{for some $\tilde{x}\in V$}.

It follows from the strong maximum principle (Lemma 4) that uλφsu_{\lambda}\equiv\varphi_{s}, which contradicts that φs\varphi_{s} is strict low solution of (4) at λ[0,λ).\lambda\in[0,\lambda^{*}).

Step 2. The set Vλ={xV:Kλ(x)<0}V_{\lambda^{*}}^{-}=\{x\in V:K_{\lambda^{*}}(x)<0\} is not empty.

From the case (1) in Theorem 2, there exists unique solution w0w_{0} of the equation

Δw+κ+e2w=0.\Delta w+\kappa+e^{2w}=0.

Together with the solution uλu_{\lambda} to the equation (4) at λ\lambda, and let vλ=uλw0v_{\lambda}=u_{\lambda}-w_{0}, we have

ΔvλKλe2uλe2w0=0.\Delta v_{\lambda}-K_{\lambda}e^{2u_{\lambda}}-e^{2w_{0}}=0.

Multiplying the above equation by e2vλe^{-2v_{\lambda}} and intergrading by parts, one has

VKλew0𝑑μ=Ve2vλΔvλ𝑑μVe2(uλ2w0)𝑑μ0.\int_{V}K_{\lambda}e^{w_{0}}d\mu=\int_{V}e^{-2v_{\lambda}}\Delta v_{\lambda}d\mu-\int_{V}e^{-2(u_{\lambda}-2w_{0})}d\mu\leq 0.

Indeed, by Green formula

Ve2vλΔvλ𝑑μ=x,yVωxy(vλ(x)vλ(y))(e2vλ(x)e2vλ(y))0\int_{V}e^{-2v_{\lambda}}\Delta v_{\lambda}d\mu=\sum_{x,y\in V}\omega_{xy}(v_{\lambda}(x)-v_{\lambda}(y))(e^{-2v_{\lambda}(x)}-e^{-2v_{\lambda}(y)})\leq 0

since (vλ(x)vλ(y))(e2vλ(x)e2vλ(y))0(v_{\lambda}(x)-v_{\lambda}(y))(e^{-2v_{\lambda}(x)}-e^{-2v_{\lambda}(y)})\leq 0 for any x,yVx,y\in V. Therefore

VKλew0𝑑μ=limλλVKλew0𝑑μ0.\int_{V}K_{\lambda^{*}}e^{w_{0}}d\mu=\lim_{\lambda\rightarrow\lambda^{*}}\int_{V}K_{\lambda}e^{w_{0}}d\mu\leq 0.

Suppose that Kλ0,K_{\lambda^{*}}\geq 0, thus Kλ0K_{\lambda^{*}}\equiv 0. This contradicts the assumption that KλK_{\lambda^{*}} is not a constant.

Step 3. uλu_{\lambda} is uniformly bounded in W1,2(V)W^{1,2}(V) as λλ\lambda\rightarrow\lambda^{*}.

Fixing x0Vλx_{0}\in V_{\lambda^{*}}^{-}, we set ρV\rho\in V^{\mathbb{R}} to be a function which vanishes besides x0x_{0} and ρ(x0)<0\rho(x_{0})<0. Consider the equation

Δw+κρe2w=0,\Delta w+\kappa-\rho e^{2w}=0,

which always has the unique solution from the case (1) in Theorem 2, set w1.w_{1}. As above, the function vλ=uλw1v_{\lambda}=u_{\lambda}-w_{1} satisfies the equation

Δvλ+ρe2w1Kλe2(vλ+w1)=0.\Delta v_{\lambda}+\rho e^{2w_{1}}-K_{\lambda}e^{2(v_{\lambda}+w_{1})}=0. (24)

Multiplying (24) by e2vλe^{2v_{\lambda}} and integrating by parts over VV gives

Ve2vλΔvλ𝑑μ+Vρe2(vλ+w1)𝑑μVKλe2(2vλ+w1)𝑑μ=0.\int_{V}e^{2v_{\lambda}}\Delta v_{\lambda}d\mu+\int_{V}\rho e^{2(v_{\lambda}+w_{1})}d\mu-\int_{V}K_{\lambda}e^{2(2v_{\lambda}+w_{1})}d\mu=0. (25)

Utilizing the fact that

(e2ae2b)(ab)(eaeb)2,a,b,(e^{2a}-e^{2b})(a-b)\geq(e^{a}-e^{b})^{2},~{}~{}~{}~{}a,b\in\mathbb{R},

one can estimate the first term in (25) as follows.

Ve2vλΔvλ𝑑μ=12x,yVωxy(e2vλ(x)e2vλ(y))(vλ(x)vλ(y))12x,yVωxy(evλ(x)evλ(y))2=V|evλ|2𝑑μ.\begin{split}\int_{V}e^{2v_{\lambda}}\Delta v_{\lambda}d\mu&=\frac{1}{2}\sum_{x,y\in V}\omega_{xy}(e^{2v_{\lambda}(x)}-e^{2v_{\lambda}(y)})(v_{\lambda}(x)-v_{\lambda}(y))\\ &\geq\frac{1}{2}\sum_{x,y\in V}\omega_{xy}(e^{v_{\lambda}(x)}-e^{v_{\lambda}(y)})^{2}\\ &=\int_{V}|\nabla e^{v_{\lambda}}|^{2}d\mu.\end{split}

Inserting this estimate into (25), we obtain

V|evλ|2𝑑μ+Vρe2(vλ+w1)𝑑μVKλe2(2vλ+w1)𝑑μ0.\int_{V}|\nabla e^{v_{\lambda}}|^{2}d\mu+\int_{V}\rho e^{2(v_{\lambda}+w_{1})}d\mu-\int_{V}K_{\lambda}e^{2(2v_{\lambda}+w_{1})}d\mu\leq 0. (26)

On the other hand, let h=evλh=e^{v_{\lambda}} in (22), yields

V|evλ|2𝑑μ2VKλe2(2vλ+w1)𝑑μ0.\int_{V}|\nabla e^{v_{\lambda}}|^{2}d\mu-2\int_{V}K_{\lambda}e^{2(2v_{\lambda}+w_{1})}d\mu\geq 0.

Together with (26), we have

V|evλ|2𝑑μ2Vρe2(vλ+w1)𝑑μ=2ρ(x0)e2uλ(x0).\int_{V}|\nabla e^{v_{\lambda}}|^{2}d\mu\leq-2\int_{V}\rho e^{2(v_{\lambda}+w_{1})}d\mu=-2\rho(x_{0})e^{2u_{\lambda}(x_{0})}. (27)

One may derive that uλ(x0)u_{\lambda}(x_{0}) is the uniform bound, i.e. there exists a constant C>0C>0 such that

e2uλ(x0)C.e^{2u_{\lambda}(x_{0})}\leq C.

Indeed, if uλ(x0)0,u_{\lambda}(x_{0})\leq 0, the above inequality is obvious; if uλ(x0)>0,u_{\lambda}(x_{0})>0, as in the proof of Lemma 5, see (11), one can get the boundedness of uλ(x0)u_{\lambda}(x_{0}) as well. Together with (27), yields

V|evλ|2𝑑μC.\int_{V}|\nabla e^{v_{\lambda}}|^{2}d\mu\leq C^{\prime}. (28)

Next, we claim that evλe^{v_{\lambda}} and thus euλe^{u_{\lambda}} is uniformly bounded in μ2\ell^{2}_{\mu}. Suppose not, we may assume that evλμ2\|e^{v_{\lambda}}\|_{\ell^{2}_{\mu}}\rightarrow\infty as λλ.\lambda\rightarrow\lambda^{*}. Let

wλ=evλevλμ2,w_{\lambda}=\frac{e^{v_{\lambda}}}{\|e^{v_{\lambda}}\|_{\ell^{2}_{\mu}}},

then wλμ2=1\|w_{\lambda}\|_{\ell^{2}_{\mu}}=1, and wλμ20\|\nabla w_{\lambda}\|_{\ell^{2}_{\mu}}\rightarrow 0 from (28). It follows that wλw_{\lambda} converges to a constant in W1,2(V)W^{1,2}(V). From (3.3), we have wλ(x0)0w_{\lambda}(x_{0})\rightarrow 0, and hence wλ0w_{\lambda}\equiv 0 on VV, which contradicts wλμ2=1\|w_{\lambda}\|_{\ell^{2}_{\mu}}=1 and then confirms our claim. Since uλu_{\lambda} is bounded below by (23) in Step 1, one has the μ2\ell^{2}_{\mu}-boundedness of uλu_{\lambda}, and thus the W1,2(V)W^{1,2}(V)-boundedness of uλu_{\lambda}. \hfill\Box

3.4 No solution when λ>λ\lambda>\lambda^{*}.

Proof of this case is a consequence of the upper and lower solutions principle, as follows.

Proof of (4) in Theorem 2. Let uλ1u_{\lambda_{1}} be the solution of (4) at some λ1>λ\lambda_{1}>\lambda^{*}. For any 0<λ<λ10<\lambda<\lambda_{1}, uλ1u_{\lambda_{1}} is an upper solution of (4) at λ\lambda. Indeed,

Δuλ1+κ(K+λ)e2uλ1=(λ1λ)e2uλ1>0.\Delta u_{\lambda_{1}}+\kappa-(K+\lambda)e^{2u_{\lambda_{1}}}=(\lambda_{1}-\lambda)e^{2u_{\lambda_{1}}}>0.

From (15), it is easy to get a lower bound solution φ\varphi of (4) at λ\lambda such that φuλ1\varphi\leq u_{\lambda_{1}}. By the upper and lower solutions principle, there exists a solution of (4) at λ\lambda, which contradicts the definition of λ.\lambda^{*}. \hfill\Box

References

References

  • [1] A. Ambrosetti, P. Rabinowitz, Dual variational methods in critical point theory and applications, J. Funct. Anal. 14 (1973) 349-381.
  • [2] F. Borer, L. Galimberti, M. Struwe, “Large” conformal metrics of prescribing Gauss curvature on surfaces of high genus, Comment. Math. Helv. 90 (2015) 407-428.
  • [3] F. Camilli, C. Marchi, A note on Kazdan-Warner equation on networks, arXiv:1909.08472.
  • [4] W. Ding, J. Liu, A note on the problem of prescribing Gaussian curvature on surfaces, Trans. Amer. Math. Soc. 347 (1995) 1059-1066.
  • [5] H. Ge, Kazdan-Warner equation on graph in the negative case, J. Math. Anal. Appl. 453 (2017) 1022-1027.
  • [6] H. Ge, W. Jiang, Kazdan-warner equation on infinite graphs, J. Korean Math. Soc. 55 (2018) 1091-1101.
  • [7] A. Grigor’yan, Y. Lin, Y. Yang, Kazdan-Warner equation on graph, Calc. Var. Partial Differential Equations 55 (2016), no. 4, Art. 92, 13 pp.
  • [8] A. Grigor’yan, Y. Lin, Y. Yang, Yamabe type equations on graphs. J. Differential Equations 261 (2016) 4924-4943.
  • [9] A. Grigor’yan, Y. Lin, Y. Yang, Existence of positive solutions to some nonlinear equations on locally finite graphs, Sci. China Math. 60 (2017) 1311-1324.
  • [10] X. Han, M. Sha, L. Zhao, Existence and convergence of solutions for nonlinear biharmonic equations on graphs, to appear in J. Differential Equations, https://doi.org/10.1016/j.jde.2019.10.007
  • [11] M. Keller, M. Schwarz, The Kazdan-Warner equation on canonically compactifiable graphs, Calc. Var. Partial Differential Equations 57 (2018), no. 2, Art. 70, 18 pp.
  • [12] J. Kazdan, F. Warner, Curvature functions for compact 2-manifolds, Ann. of Math. 99 (1974) 14-47.
  • [13] J. Kazdan, F. Warner, Curvature functions for open 2-manifolds, Ann. of Math. 99 (1974) 203-219.
  • [14] Y. Lin, Y. Wu, The existence and nonexistence of global solutions for a semilinear heat equation on graphs,Calc. Var. Partial Differential Equations 56 (2017), no. 4, Art. 102, 22 pp.
  • [15] M. Struwe, Critical points of embeddings of H01,nH_{0}^{1,n} into Orlicz spaces, Ann. Inst. H. Poincaré Anal. Non Linéaire 5 (1988) 425-464.
  • [16] M. Struwe, The existence of surfaces of constant mean curvature with free boundaries, Acta Math. 160 (1988) 19-64.
  • [17] Y. Yang, X. Zhu, Prescribing Gaussian curvature on closed Riemann surface with conical singularity in the negative case, Ann. Acad. Sci. Fenn. Math. 44 (2019) 167-181.