Skip to content
BY 4.0 license Open Access Published by De Gruyter November 29, 2021

Bifurcation analysis for a modified quasilinear equation with negative exponent

  • Siyu Chen , Carlos Alberto Santos , Minbo Yang EMAIL logo and Jiazheng Zhou

Abstract

In this paper, we consider the following modified quasilinear problem:

ΔuκuΔu2=λa(x)uα+b(x)uβinΩ,u>0inΩ,u=0onΩ,

where Ω ⊂ ℝN is a smooth bounded domain, N ≥ 3, a, b are two bounded continuous functions, α > 0, 1 < β ≤ 22* − 1 and λ > 0 is a bifurcation parameter. We use the framework of analytic bifurcation theory to obtain an analytic global unbounded path of solutions to the problem. Moreover, we get the direction of solution curve at the asmptotic point.

MSC 2010: 35B32; 35J10; 35J62; 35J75

1 Introduction

In this paper we consider the following modified quasilinear problem

ΔuκuΔu2=λa(x)uα+b(x)uβinΩ,u>0inΩ,u=0onΩ, (1.1)

where Ω is a smooth bounded domain in ℝN, N ≥ 3, 0 < a(x) ∈ 𝓒(Ω), b(x) ∈ 𝓒(Ω) ∩ L(Ω) and may change sign, κ > 0 is a real constant, α>0,1<β221,2=2NN2,λ>0 is a bifurcation parameter. Problem (1.1) is related to the standing wave solutions for the quasilinear Schrödinger equations

itψ=Δψ+ψ+η(|ψ|2)ψκΔρ(|ψ|2)ρ(|ψ|2)ψ, (1.2)

where ψ = ψ(t, x), ψ:ℝ × Ω → ℂ, κ > 0 is a real constant. Equation (1.2) has been applied extensively in many areas of physical phenomena, for the progress of this topic and the other modified Schrödinger equations one may refer to [3, 12, 13, 24, 25, 26, 29, 30, 31, 32, 42, 44, 45, 46].

For κ = 0, problem (1.1) can be transformed into a semilinear one. In recent years, this type of equations has been studied extensively in both bounded and unbounded domains due to its wide applications in non-Newtonian fluids. For instance, Lazer and McKenna [27] studied the following semilinear problem

Δu=p(x)uγinΩ,u=0onΩ. (1.3)

For p(x) > 0 with some smoothness conditions, they showed that there exists a solution which is smooth on Ω and continuous on Ω, the Lazer-McKenna obstruction then was firstly presented: the equation has a H01 -solution if and only if γ < 3. For the power of 3, Sun and Zhang [40] provided an extension of the classical Lazer-McKenna obstruction and revealed the role of 3, they also gave a local description of the solution set. Lair and Shaker [28] proved that (1.3) has a unique weak H01 -solution on a bounded domain provided 0ε f(s)ds < ∞ and p(x) ∈ L2(Ω). For the regularity of (1.3), Gui and Lin [17] obtained the positive solutions that are Hölder-continuous up to the boundary and has even better regularity in some special cases. The problem was also studied by Ma and Wei [34] when p(x) = −1, they showed that the gradient estimates, L1-estimates, global upper bounds, Liouville properties, classification of stable and finite Morse index solutions, and symmetry properties.

For the perturbed singular problem, Yang [43] considered the following problem with singular nonlinearity,

Δu=λuγ+upinΩ,u=0onΩ. (1.4)

For 0 < γ < 1 < p ≤ (N + 2)/(N − 2), Yang carried out a direct analysis in an H1-neighborhood and proved that (1.4) has a solution which is a local minimiser with respect to the H1-topology. Then the existence of the second solution was given by making use of Ekeland’s variational principle. Arcoya and Moreno-Mérida [2] extended the results of [43] to all γ > 0 in subcritical case by establishing suitable approximated problems, they showed that there exists Λ > 0 such that (1.4) has two positive solutions for every λ ∈ (0, Λ).

Apart from the existence and regularity of solutions for this type of equations, there are many researchers have obtained global bifurcation and local multiplicity results. For instance, the authors in [11] considered the following singular elliptic problem with exponential type growth in a bounded smooth domain Ω ⊂ ℝ2,

Δu=λ(uδ+h(u)eup)inΩ,u=0onΩ,

where 1 ≤ p ≤ 2, 0 < Δ < 1 and h(t) is a smooth “perturbation” of etp as t → ∞. For the radial case, they made a detailed study of the blow-up/convergence of the solution branch as it approaches to the asymptotic bifurcation point at infinity. For the critical case p = 2, they also interpreted all previous works on multiplicity in terms of the corresponding bifurcation diagrams and the asymptotic profile of large solutions along the branch at infinity. Later, Bougherara et al. [5] considered the following semilinear elliptic problem with a strong singular term in a bounded smooth domain Ω ⊂ ℝN (N ≥ 2),

Δu=λ(uδ+f(u))inΩ,u=0onΩ. (1.5)

They improved the results of [11] to all Δ > 0, and obtained an analytic global unbounded path of solutions of (1.5) by using the framework of analytic bifurcation theory as developed in the work [4]. In two dimensions, for 0 < Δ < 1 and certain classes of nonlinearities f with critical growth, it was shown that the existence of an analytic unbounded path of solutions of (1.5) whose Morse index is unbounded along the path and admits infinitely many turning points. Specially, for p-Laplacian differential operator, such bifurcation type results were obtained by Bai et al. [6], Papageorgiou, Rădulescu and Repovš [35, 36]. For more results about this type of equations, one may see [1, 8, 14, 18, 19, 20, 22, 38] and the references therein.

Motivated by the above papers and by the increasing interest on problems with singular nonlinearities, our main purpose in this paper is to investigate the analytic global bifurcation in the case of κ = 1 for (1.1). The quasilinear term u2 makes the problem much more complicated. By using a change of variables, the authors in [30] transformed the quasilinear Schrödinger equations into a new semilinear one and showed that the existence of ground states of soliton-type solutions by variational method. Involving the quasilinear operator and singular nonlinearities in bounded domain, there are some results as well. For instance, the authors in [16] considered the following singular quasilinear problem for Δ ∈ (0, 1) in a ball Ω ⊂ ℝN,

ΔuuΔu2=λu3uuδinΩ,u>0inΩ,u=0onΩ. (1.6)

They obtained the existence of solutions of (1.6) when λ belongs to a certain neighborhood of the first eigenvalue λ1. Moameni and Offin [33] obtained the same results as in [16] by considering a more general class of equations. The authors in [37] considered the following class of singular quasilinear problem,

ΔuuΔu2=a(x)uδ+h(x,u)inΩ,u>0inΩ,u=0onΩ. (1.7)

The function a(x) is nonnegative, Δ > 0 is a constant and the nonlinearity h(x, u) is continuous. They showed that the existence of a solution for the problem via sub-supersolution method when h has an arbitrary polynomial growth. For the second result, they showed that the existence of the second solution by applying the mountain pass theorem when h has subcritical growth. Recently, the authors in [39] studied problem (1.1) in the case of 0 < α < 1, they showed that the existence of a minimal solution as a minimum critical point of the energy functional, and then the second solution was also given by constrained critical point theory.

As far as we know in the literature there is little research on the bifurcation analysis of solutions to this type of quasilinear equation. The present paper is mainly consider the case of κ = 1, and regard (1.1) as a bifurcation problem with λ being the bifurcation parameter. There seem to be some difficulties to transform the quasilinear equation to a semilinear one by making a change of variables, such that the existence of solutions of (1.1) be equivalent to the existence of solutions to new transformed equation, and there holds similar properties. Inspired by [4], we use the framework of analytic bifurcation theory to obtain an analytic global unbounded path of solutions to the transformed equation and so to problem (1.1). Then the direction of the solution curve at the asmptotic point under some conditions is given by making use of local bifurcation theory. %There are some difficulties to overcome. First of all, we need deal with the quasilinear term. Besides, the existence of solution branch established for any α > 0. For that, we first transform the quasilinear equation to a semilinear one by changing of variable such that the existence of solutions to (1.1) be equivalent to the existence of solutions to new transformed equations, and there holds similar properties. Inspired by [4], we use the framework of analytic bifurcation theory to obtain an analytic global unbounded path of solutions to the transformed equation and so to problem (1.1). In addition, we will be also able to get the direction of the solution curve at the asmptotic point under some conditions.

The paper is organized as follows. In section 2, we state some preliminaries and main results including transforming quasilinear problem (1.1) into a semilinear elliptic one and give some definitions and lemmas. In section 3, we give the proof to the main results by using bifurcation theory to the transformed problem to obtain an global unbounded path of solutions, and the properties at turning point.

2 Preliminaries and main results

Taking into account the ideas of [30], we can use the change of variables ω = h−1(u) to transform the quasilinear equation into a semilinear one with singularity at zero and superlinear at infinity, which h is defined by

h(t)=(1+2|h(t)|2)1/2fort>0,h(t)=h(t)fort0.

Now, we list some properties of the function h(t) as given in [41, 42].

Lemma 2.1

Assume h : ℝ → ℝ is given as above, then there hold:

  1. h(t) = −2h(t)(h(t))4, t > 0,

  2. h is unique, invertible, and 𝓒(ℝ)-function,

  3. 0 < h(t) ≤ 1 for all t ∈ ℝ,

  4. h(t)∣ ≤ ∣tfor all t ∈ ℝ,

  5. limt0h(t)t=1,limth(t)t=0andlimth(t)t=21/4,

  6. h(t)h(t)∣ ≤ 1/2 for all t ∈ ℝ,

  7. h(t)/2 ≤ th(t) ≤ h(t) for all t ≥ 0,

  8. h(t)∣ ≥ h(1)∣tfort∣ ≤ 1, andh(t)∣ ≥ h(1)∣t1/2 fort∣ ≥ 1,

  9. the function hα(t)h(t) is decreasing for t > 0 and α > 0,

  10. the function hβ(t)h(t) is increasing for t > 0 and β ≥ 1.

Using Lemma 2.1, we can perform the changing of variables ω = h−1(u) to infer that u Hloc1 (Ω) is a solution of (1.1) if and only if ω Hloc1 (Ω) is a weak solution of

Δω=[λa(x)h(ω)α+b(x)h(ω)β]h(ω)inΩ,ω>0inΩ,ω=0onΩ (2.1)

in the sense of the following definition.

Definition 2.2

We say ω Hloc1 (Ω) ∩ 𝓒0(Ω) is a weak solution of (2.1) if essinfKω>0 for any compact set KΩ, (uε)+ H01 (Ω) for any ε > 0 given, and

Ωωϕ=Ωλa(x)h(ω)αh(ω)ϕ+b(x)h(ω)βh(ω)ϕ,foranyϕC0(Ω).

To state our main result, we denote the following set of all classical solutions to (2.1)

S:={ωC2(Ω)C0(Ω¯),ω>0solves(2.1)}.

In the sequel, let φ1 be the first positive eigenfunction for −Δ in H01 (Ω), ∥φ1L(Ω) = 1, we define

ϕα=φ1,0<α<1,φ1(logφ1)12,α=1,φ12α+1,α>1,

then

Cϕ(Ω):={ωC(Ω)|forsomeC>0,|ω|Cϕ(x),xΩ}

defines a Banach space endowed with the norm

ωCϕ(Ω):=supxΩ|ωϕ(x)|,

consequently,

Cϕ+(Ω):={ωCϕ(Ω)|infxΩωϕ(x)>0}

is an open convex subset of 𝓒ϕ(Ω).

Now, we define the following solution operator associated to (2.1):

F(λ,ω)=ω(Δ)1[λa(x)h(ω)α+b(x)h(ω)β]h(ω),

where (λ,ω) ∈ ℝ+ × Cϕ+ (Ω), λ > 0.

Remark 2.3

Note that, for any ω ∈ 𝓒0(Ω) solves equation (2.1) is indeed twice continuously differentiable in Ω by standard elliptic regularity, see for example [5][10].

We recall some results about global analytic bifurcation theory that introduced in [7]. Let 𝓧,𝓨 be real Banach spaces, 𝓤 ⊂ ℝ × 𝓧 be an open set containing (0, 0) in its closure and F : 𝓤 → 𝓨 be an ℝ-analytic function. Define the solution set

S={(λ,x)U:F(λ,x)=0}

and the non-singular solution set

N={(λ,x)S:ker(xF(λ,x))=0)}.

Definition 2.4

A distinguished arc is a maximal connected subset of 𝓝.

Let us introduce the following assumptions:

  1. Bounded closed subsets of 𝓢 are compact in ℝ × 𝓧.

  2. xF(λ, x) is a Fredholm operator of index zero for all (λ, x) ∈ 𝓢.

  3. There exists an analytic function (λ, u) : (0,ε) ↦ 𝓢 such that xF(λ(s), u(s)) is invertible for all s ∈ (0,ε) and lims0+ (λ(s), u(s)) = (0, 0).

Denote

A:={(λ(s),u(s)):s>0}

and

A+={(λ(s),u(s)):s(0,ε)}.

Evidently, 𝓐+ ⊂ 𝓢. The function (λ, u) from (0, ε) to (0, ∞) in the ℝ-analytic case as follows.

Lemma 2.5

Assume (H1) − (H3) hold. Then (λ, u) can be extended as a continuous map (still called) (λ, u) : (0, ∞) ↦ 𝓢 with the following properties:

  1. 𝓐 ∩ 𝓝 is an at most countable union of distinct distinguished arcs i=0nAi,n.

  2. 𝓐+ ⊂ 𝓐0.

  3. {s > 0 : ker(xF(λ(s), u(s))) ≠ {0}} is a discrete set.

  4. At each of its points, 𝓐 has a local analytic re-parameterization in the following sense: for each s* ∈ (0, ∞), there exists a continuous and injective map ρ* : (−1, 1) ↦ ℝ such that ρ*(0) = s* and the re-parametrisation

    (1,1)t(λ(ρ(t)),u(ρ(t)))Aisanalytic.

    Furthermore, the map sλ(s) is injective in a right neighborhood of s = 0 and for each s* > 0 there exists ε* > 0 such that λ is injective on [s*, s*+ε*] and [s*ε*, s*].

  5. One of the following holds:

    1. ∥(λ(s), u(s))∥ℝ×𝓧 → ∞ as s → ∞,

    2. the sequence {(λ(s), u(s))} approaches the boundary of 𝓤 as s → ∞,

    3. 𝓐 is the closed loop:

      A={(λ(s),u(s)):0sT,(λ(T),u(T))=(0,0)forsomeT>0}.

      In this case, chosing the smallest T > 0 such that

      (λ(s+T),u(s+T))=(λ(s),u(s))foralls0.

  6. Suppose that xF(λ(s1), u(s1)) is invertible for some s1 > 0. If (λ(s1), u(s1)) = (λ(s2), u(s2)) for some s2s1, then (e)(iii) occurs ands1s2is an integer multiple of T. In particular, the map s ↦ (λ(s), u(s)) is injective on [0, T).

From the definition of F, we immediately have the following Lemma.

Lemma 2.6

Let F be given as above. Then

  1. ωF(λ,ω)v = vλ(−Δ)−1 a(x)[h(ω)αh(ω)] v − (−Δ)−1b(x)[h(ω)βh(ω)] v,

  2. λF(λ,ω)v = −(−Δ)−1a(x)h(ω)αh(ω)v,

  3. ωωF(λ,ω)(v, z) = −λ(−Δ)−1 a(x)[h(ω)αh(ω)] vz − (−Δ)−1b(x)[h(ω)βh(ω)] vz.

Now, we are ready to state our main results.

Theorem 2.7

Assume that α > 0, 1 < β ≤ 22*, and b+ ≠ 0. Then there exists Λ ∈ (0, ∞) and an unbounded set 𝓐 ⊂ (0, Λ] × Cϕα+ (Ω) of solutions to problem (2.1) which is globally parametrised by a continuous map s ↦ (λ(s),ω(s)) where s ∈ (0, ∞) and (λ(s),ω(s)) ∈ 𝓐 ⊂ 𝓢. Furthermore, the path 𝓐 has following properties:

  1. (λ(s), ω(s)) → (0, 0) in ℝ × 𝓒ϕα(Ω) as s → 0+,

  2. ω(s)∥𝓒ϕα(Ω) → ∞ as s → ∞,

  3. {s ≥ 0: ωF(λ(s), ω(s)) is not invertible} is a discrete set,

  4. the branch of minimal solutions {(λ, ωλ) : 0 < λ < Λ} of (2.1) coincides with a path-connected portion of 𝓐 which closure containing (0, 0), furthermore, the minimal solution branch is parametrised by an analytic map,

  5. 𝓐 has at least one asymptotic bifurcation point Λa ∈ [0, Λ],

  6. each point of 𝓐 has a local analytic re-parametrization as follows: for each s* ∈ (0, ∞), there exists a continuous and injective map ρ* : (−1, 1) ↦ ℝ such that ρ*(0) = s* and the re-parametrisation

    (1,1)t(λ(ρ(t),ω(ρ(t)))Aisanalytic.

    Moreover, the map sλ(s) is injective in a right neighborhood of s = 0 and for each s* > 0 there exists ε* > 0 such that λ is injective on [s*, s*+ε*] and [s*ε*, s*].

Corollary 2.8

In addition to assertions in Theorem 2.7, if (Λ, ωΛ) ∈ 𝓐 for some ωΛ Cϕα+ (Ω) and α52. Then 𝓐 turns to the left of {λ = Λ} at the point (Λ, ωΛ) ∈ 𝓐.

Fig. 2.1 
Possible bifurcation branch
Fig. 2.1

Possible bifurcation branch

3 Local and global bifurcation analysis

In this section, we establish local and global bifurcation to problem (2.1). Firstly, we consider the properties of the linearised operator for the corresponding function F to the problem.

3.1 Analysis of the solution operator and linearised operator

Proposition 3.1

Assume that the changing of variables h is defined section 2. Then the map F : ℝ × Cϕα+ → 𝓒ϕα is well defined and analytic.

Proof

We split the proof in three steps.

  1. h(ω) ∈ Cϕα+ for any ω Cϕα+ . We just consider the case 0 < α < 1, because the case α ≥ 1 is similar. Since ϕα = φ1, it follows from the properties of h(t) and ∥φ1L(Ω) = 1, that there exist positive constants C1, C2 depend on ω such that

    0<C1h(ω)φ1(x)ωφ1(x)C2<,xΩ.

    Then h(ω)α Cϕαα+ (Ω) for ω Cϕα+ . By the fact that λ > 0, 0 < a(x) ∈ 𝓒(Ω) and h(ω) ∈ 𝓒(Ω), we conclude that λ a(x)h(ω)αh(ω) ∈ Cϕαα+ (Ω).

  2. For any ω Cϕαα (Ω), by similar ideas made in the proofs of Proposition 2.3 in [5], we are able to obtain that ω ↦ (−Δ)−1ω ∈ 𝓒ϕα(Ω) is a linear continuous map and hence analytic.

  3. Due to α > 0, 1 < β ≤ 22* − 1 and the properties of h(t), it is easy to see that (−Δ)−1b(x)h(ω)βh(ω) ∈ 𝓒ϕα(Ω) for any ω Cϕα+ .

Hence, by the above three steps, we can obtain the result.□

Now, to show the existence of the analytic global path of solutions to F in ℝ+ × Cϕα+ (Ω), we consider the following problem:

Δw+kw=λa(x)h(w)αh(w)+g(x)inΩ,w=0onΩ, (3.1)

where k ≥ 0 and g(x) is a local Hölder continuous function in Ω.

To begin with, we have the following comparison principle.

Lemma 3.2

Assume that there exist u and v satisfying the following inequalities in the weak sense,

Δuλa(x)h(u)αh(u)+g(x)kuinΩ,Δvλa(x)h(v)αh(v)+g(x)kvinΩ,uvonΩ, (3.2)

then there holds uv in Ω.

Proof

Arguing by contradiction, assume that Ω0 := {xΩ: u(x) > v(x)} ≠ ∅. For fixed ϵ > 0, let us define

uϵ(x)=u(x)+ϵ,vϵ(x)=v(x)+ϵ

and

φϵ=(uϵ2vϵ2)+uϵ,ψϵ=(uϵ2vϵ2)+vϵ.

By pointwise limit, we get

φϵφ:=u2v2uχΩ0,ψϵψ:=u2v2vχΩ0,

where χΩ0 represent the characteristic function of Ω0.

By taking derivatives, we get

φϵ=u2v+ϵu+ϵv+(v+ϵ)2(u+ϵ)2u,

and

ψϵ=2u+ϵv+ϵu(u+ϵ)2(v+ϵ)2vv

in Ω0. From the above argument, we have that φϵ, ψϵ Hloc1 (Ω) ∩ 𝓒0(Ω). So, by density arguments, we are able to test the first and second inequalities in (3.2) against φϵ and ψϵ, respectively, to obtain

Ω0[uφϵvψϵ]dx=Ω0[|u|22v+ϵu+ϵuv+(v+ϵ)2(u+ϵ)2|u|22u+ϵv+ϵuv+(u+ϵ)2(v+ϵ)2|v|2+|v|2]dx=Ω[|u|22vϵuϵuv+vϵ2uϵ2|u|2dx2uϵvϵuv+uϵ2vϵ2|v|2+|v|2]dx. (3.3)

Now, set W1:=lnuϵ=uuϵ and W2:=lnvϵ=vvϵ in Ω0, it follows from (3.3) and Lemma 4.2 of [23], that

Ω[uφϵvψϵ]dx=Ω0[|uϵ2|W1|22vϵ2W1W2+vϵ2+2uϵ2W1W2+uϵ2|W2|2+vϵ2|W2|2]dx=Ω0uϵ2[|W1|2|W2|22W2(W1W2)]+vϵ2[|W2|2|W1|22V1(V2V1)]dxΩ0(uϵ2+vϵ2)|W1W2|2dx0. (3.4)

On the other hand, since a(x) > 0 and h(t)αh(t) is decreasing for t > 0 and α > 0, we have

Ωλa(x)[h(u)αh(u)(uϵ2vϵ2)+uϵh(v)αh(v)(uϵ2vϵ2)+vϵ]dx=Ω0λa(x)(uϵ2vϵ2)+[h(u)αh(u)1uϵh(v)αh(v)1vϵ]dxΩ0λa(x)(uϵ2vϵ2)+h(v)αh(v)(1uϵ1vϵ)<0. (3.5)

Besides, we are able to obtain

Ω[kv(uϵ2vϵ2)+vϵku(uϵ2vϵ2)+uϵ]dx=Ω0k(uϵ2vϵ2)(vvϵuuϵ)<0.

This, together with (3.4) and (3.5), implies

0Ω[uφϵvψϵ]dx<0,

which is a contradiction. Thus, Ω0 = ∅, and hence the proof is completed.□

Lemma 3.3

There exists a unique weak solution w W01,q (Ω) ∩ 𝓒0(Ω) for some q > 1 to problem (3.1). Furthermore, w ∈ [α, ϕ] for c > 0 small enough if there exists ϕ Cϕα+ (Ω) which is a super-solution of (3.1). In particular, w Cϕα+ (Ω).

Proof

Given k > 0, 0 ≤ g(x) ∈ L(Ω), we consider the following approximated problem:

Δw+kw=λa(x)h(w+ε)αh(w+ε)+g(x)inΩ,w=0onΩ. (3.6)

It is easy to check that for c > 0 small enough, ψ_ϵ=(c1+α2φ1+ε1+α2)21+αε is a sub-solution of (3.6) for ε > 0. The unique positive solution ψε H01 (Ω) of

Δψ¯ε+kψ¯ε=λa(x)h(ε)αh(ε)+g(x)L(Ω)

is a super-solution of (3.6). Indeed, it follows from the monotonicity of h(t)αh(t) that

Δψ¯ε+kψ¯ε=λa(x)h(ε)αh(ε)+g(x)L(Ω)λa(x)(h(ψ¯ε)+ε)αh(ψ¯ε)+g(x).

By comparison principle, it is obvious that ψϵ < ψε. Then we obtain a solution wε ∈ [ψϵ, ψε] to (3.6) by standard arguments, and uniquely by the non-increasing nature of the right hand side in (3.6). Thus, we can infer that wε is Hölder continuous on Ω through elliptic regularity. In addition, wε > 0 in Ω by maximum principle.

Now we prove that wε is monotone as ε → 0+ by a comparison argument: let 0 < ε < ε, then we have

Δ(wεwε)+k(wεwε)=λa(x)[h(wε+ε)αh(wε+ε)h(wε+ε)αh(wε+ε)].

On the other hand, assume x0=argminΩ¯(wεwε)Ω, and (wεwε)(x0) ≤ 0, then it follows that

Δ(wεwε)(x0)+k(wεwε)(x0)λa(x)[h(wε(x0)+ε)αh(wε(x0)+ε)h(wε(x0)+ε)αh(wε(x0)+ε)]<0,

which is a contradiction with the last equation. Thus we have wε > wε in Ω if 0 < ε < ε. Therefore, we obtain that

w=limε0+wεcϕαandwC0(Ω¯) (3.7)

and satisfies in the sense of distributions of (3.1).

Furthermore, w W01,q (Ω) for some q > 1. Indeed, from g(x) ∈ L(Ω) and (3.7), it is easy to get that g(x)+λ a(x)h(w)αh(w) ∈ L1(Ω, d(x, Ω)s) for some s < 1. Then, it follows from Theorems 3 and 4 of [15], that w W01,q (Ω) for some q > 1. Using the comparison principle again, we can derive that w is the unique weak solution of (3.1).

Now, assume that (3.1) has a super-solution ϕ Cϕα+ (Ω). It is clear that ϕ is also a super-solution of (3.6). Thus, for c > 0 small enough, we get αϕ. Hence ψε can be replaced by ϕ and repeat the above argument to get a solution w Cϕα+ (Ω). Thus we complete the proof.□

Since 0 < a(x) ∈ 𝓒(Ω) and h(t)αh(t), t > 0, is decreasing, then (h(t)αh(t)) ≤ 0. From the idea of [5], let m(x) := −a(x)(h(ω)αh(ω)), then we have m(x) ≥ 0. Now, we consider the following problem

Δv+m(x)v=m(x)zinΩ,zCϕα(Ω).

We have the following result.

Lemma 3.4

Assume that m(x) is defined as above, 0 ≤ m(x) ≤ m1d(x, Ω)−2 for some positive constant m1. Then, for given z ∈ 𝓒ϕα(Ω), there exists a unique v ∈ 𝓒ϕα(Ω) solvesΔ v + m(x)v = m(x)z in Ω. Furthermore, vCϕα(Ω)CmzCϕαα(Ω) for some constant C > 0 independent of z.

To this end, we need the next Lemma.

Lemma 3.5

If m(x) is defined as above, then there exists positive constant m1 such that m(x) ≤ m1d(x, Ω)−2.

Proof

Since a(x) ∈ 𝓒(Ω) be bounded, it suffices to prove that there exists a positive constant C such that −(h(ω)αh(ω)) ≤ Cd(x, Ω)−2 for all ω > 0. In the following process, the C represents different positive constant. A direct calculation shows that

(h(ω)αh(ω))=h(ω)α1[α(h(ω))2h(ω)h(ω)]=h(ω)α1[α(h(ω))2+2h2(ω)(h(ω))4]=h(ω)α1(h(ω))2[α+2h2(ω)(h(ω))2].

Note that h(ω)α−1d(x, Ω)−2 near Ω. Besides this, by Lemma 2.1-(3) and −(6), we have (h(ω))2C and [h(ω)h(ω)]2C for all ω > 0. Now, it is obvious that −(h(ω)αh(ω)) ≤ Cd(x, Ω)−2 for all ω > 0. Consequently, one can choose a positive constant m1 such that m(x) ≤ m1d(x, Ω)−2.□

Corollary 3.6

Let g(x) ∈ Cϕαα (Ω), m(x) ∈ 𝓒(Ω) and 0 ≤ m(x) ≤ m1 d(x, Ω)−2 for some positive constant m1. If v ∈ 𝓒ϕα(Ω) ∩ 𝓒2(Ω) is a classical solution of

Δv+m(x)v=g(x)inΩ,

then vCϕα(Ω)cgCϕαα(Ω) for c > 0 independent of g.

Lemma 3.7

The map ωF(λ, ω) is Fredholm with index 0 for all (λ, ω) ∈ ℝ+ × Cϕα+ (Ω). %for b+ ≠ 0.

Proof

Rewrite ωF(λ, ω) = I + Aλ + B, where Aλv = −λ(−Δ)−1 a(x)(h(ω)αh(ω))v, Bv = −(−Δ)−1b(x)(h(ω)βh(ω)) v.

Applying Lemma 3.4 with m(x) = −λ a(x)(h(ω)αh(ω)), which turns out that I + Aλ is invertible on 𝓒ϕα(Ω). On the other hand, B is compact on 𝓒ϕα(Ω). Thus, ωF(λ, ω) is Fredholm with index 0.□

3.2 Local and global bifurcation analysis

In this section, we shall show the existence of minimal solution to problem (2.1) for λ ∈ (0, Λ), and then state that the full set of minimal solution can be parametrised by an analytic curve. Besides this, we shall illustrate some bifurcation results for λ = Λ, where

Λ:=sup{λ>0:(2.1)hasaweaksolution}.

Lemma 3.8

It holds 0 < Λ < ∞ and (2.1) admits a minimal solution ωλ Cϕα+ (Ω) for all 0 < λ < Λ with b+(x) ≠ 0.

Proof

Let ϕ_λ=cλ11+αϕα, then we can check that ϕλ is a sub-solution of (2.1) for all λ > 0 if c > 0 is chosen small enough. Next, we find a super-solution to (2.1). Consider the following problem

Δω=λ¯a(x)h(ω)αh(ω)inΩ,ω>0inΩ,ω=0onΩ. (3.8)

Since the conditions (g1) and (g2) of Theorem 2.2 in [10] are fulfilled by the nonlinear perturbation of (3.8), we conclude that (3.8) admits a unique solution ψλ Cϕα+ (Ω) for any α > 0. Let ν solves

Δν=1inΩ,ν>0inΩ,ν=0onΩ. (3.9)

Define ϕλ := ψλ + . For an appropriate constant M > 0 and some λ > 0, ϕλ is a super-solution of (2.1) for 0 < λ < λ. Indeed, combining the monotonicity of h(t)αh(t) with the properties of h(t), we have

Δϕ¯λ=λ¯a(x)h(ψλ¯)αh(ψλ¯)+Mλa(x)h(ϕ¯λ)αh(ϕ¯λ)+b(x)h(ϕ¯λ)βh(ϕ¯λ),

if λ + M > λ + b(x)h(ϕλ)βh(ϕλ) sufficiently large. This also implies λ > 0. Note that we can obtain ϕλ Cϕα+ (Ω), then there holds liminfλ0+infxΩϕ¯λλ11+αϕα>0. This implies for all 0 < λ < λ, there holds ϕλ < ϕλ in Ω if choose c > 0 small enough.

Consider the following monotone iterative scheme for all λ ∈ (0, λ),

Δωnλa(x)h(ωn)αh(ωn)+kωn=b(x)h(ωn1)βh(ωn1)+kωn1inΩ,ωn=0onΩ, (3.10)

with ω0 = ϕλ and k = k(λ) > 0 large enough such that b(x)h(t)βh(t) + kt is non-decreasing on [0, ∥ϕλL(Ω)]. From the comparison principle and Lemma 3.3, we can get the existence of ωn, and ϕλ < ωn < ϕλ. It is easy to check that ϕλ, ϕλ are respectively sub and super-solution to (3.10). Furthermore, due to the monotonicity of the left side of (3.10), it follows that the monotonicity of the iterates, that is ωnωn−1. By Ascoli-Arzela Theorem, there exists ωλ such that ωnωλ in Cϕα+ (Ω) as n → ∞ and ϕλωλϕλ. That is, ωλ is a minimal solution of (2.1) for 0 < λ < λ.

Now, we set

Λ:=sup{λ>0:(2.1)hasaweaksolution}.

From the above argument, we have Λ > 0. We claim that Λ < ∞. In fact, taking φ1 as the test function in (2.1), we obtain

Ω(λa(x)h(ω)αh(ω)φ1+b(x)h(ω)βh(ω)φ1)dx=ωφ1(Δ)ωdx=ωω(Δ)φ1dx=λ1Ωωφ1dx. (3.11)

If we chose a λ > 0, large enough if necessary, such that λ a(x)h(t)αh(t) + b(x)h(t)βh(t) > 2λ1t for all t > 0, which is a contradiction with (3.11). Thus Λ < ∞ must holds. Now, we can further get the existence of minimal solution ωλ Cϕα+ (Ω) for (2.1) with any 0 < λ < Λ. In fact, taking ϕ_λ=cλ11+αϕα as a sub-solution, and ϕλ, solves problem (2.1)λ, for appropriate λ < λ < Λ as a super-solution of (2.1). Then by the similar proceeding above, we can conclude that there exists a minimal solution ωλ Cϕα+ (Ω) for all 0 < λ < Λ, which completes the proof.□

Remark 3.9

Suppose that there exists M0 > 0 such that

λa(x)[h(t)αh(t)]+b(x)[h(t)βh(t)]0in(0,M0),

if further choose λ0 small enough such that sup0<λ<λ0ωλC(Ω¯)<M0. Then ωλ is the unique solution in (0, λ0) × {ω C0+ (Ω) : ∥ω𝓒(Ω) < M0}.

Indeed, suppose ω̃λ is another solution and satisfiesω̃λ𝓒(Ω) < M0 with λ < λ0. Let ψλ = ωλω̃λ, then ψλ solves

Δψλ[λa(x)(h(ξλ)αh(ξλ))+b(x)(h(ξλ)βh(ξλ))]ψλ=0,

where ξλ lies between ωλ and ω̃λ. It is easy to see λ a(x)(h(ξλ)αh(ξλ)) + b(x)(h(ξλ)βh(ξλ)) ≤ 0 and hence ψλ ≡ 0.

Lemma 3.10

Let ω ∈ 𝓒2(Ω) ∩ Cϕα+ (Ω), λ > 0. If ωF(λ, ω)φ = 0 for some φ ∈ 𝓒2(Ω) ∩ 𝓒ϕα(Ω), then φ H01 (Ω) ∩ 𝓒φ1(Ω) and is a H1-weak solution forΔφ − [λ a(x)(h(ω)αh(ω)) + b(x)(h(ω)βh(ω))]φ = 0. Inversely, if φ H01 (Ω) is a non-negative H1-weak solution forΔφ − [λ a(x)(h(ω)αh(ω)) + b(x)(h(ω)βh(ω))]φ = θφ for some θ ∈ ℝ, then φ ∈ 𝓒2(Ω) ∩ 𝓒φ1(Ω).

Proof

For some φ ∈ 𝓒2(Ω) ∩ 𝓒ϕα(Ω), define the minimization problem infψH01(Ω)F(ψ), where

F(ψ)=Ω|ψ|2dxΩλa(x)(h(ω)αh(ω))ψ2dxΩb(x)(h(ω)βh(ω))φψdx.

By the properties of h(t), a(x) and b(x), it is easy to show that the above functional is coercive and weakly lower semicontinuous on H01 (Ω). Then there exists a minimiser ψ0 H01 (Ω) and is a non-trivial H1-weak solution of

Δψ0λa(x)(h(ω)αh(ω))ψ0=b(x)(h(ω)βh(ω))φ.

By standard elliptic regularity, we have ψ0 ∈ 𝓒2(Ω). Now, in H1 -weak sense, let us take a comparison with the solution ξ H01 (Ω) of −Δξ = M in Ω, where M=supΩ¯|b(x)(h(ω)βh(ω))φ|, which infer that ψ0 ∈ 𝓒φ1(Ω). Then we get that φψ0 ∈ 𝓒2(Ω) ∩ 𝓒0(Ω) is a solution of

Δ(φψ0)λa(x)(h(ω)αh(ω))(φψ0)=0.

It follows from maximum principle that φψ0. Hence φ H01 (Ω) ∩ 𝓒φ1(Ω) and is a H1-weak solution for −Δφ − [λ a(x)(h(ω)αh(ω))+b(x)(h(ω)βh(ω))]φ = 0.

Inversely, by similar ideas made in the proofs of Theorem 8.15 in [21], we can easily see that if φ H01 (Ω) is a non-negative H1-weak solution of

Δφ[λa(x)(h(ω)αh(ω))+b(x)(h(ω)βh(ω))]φ=θφ

for some θ ∈ ℝ, then φ ∈ 𝓒φ1(Ω) ∩ 𝓒2(Ω).□

Definition 3.11

Let Γ := {(λ, ωλ) : 0 < λ < Λ, ωλ is the minimal solution of (2.1)}.

Lemma 3.12

For any (λ, ωλ) ∈ Γ, ωF(λ, ωλ) is invertible, and further the full set of minimal solutions Γ is parametrised by an analytic map.

Proof

We just consider the case of α > 1, the case of 0 < α ≤ 1 is similar. Consider the following problem

Δv=λa(x)h(v+ε)αh(v+ε)+b(x)h(v)βh(ω)inΩ,v>0inΩ,v=0onΩ, (3.12)

where ε > 0. Let ψε=(cα+12φ1+εα+12)2α+1ε, then we can check that ψε is a sub-solution of (3.12) if we chose c = cλ > 0 small enough and λ > 0. On the other hand, we can find that ωλ, obtained above, is a super-solution of (3.12). Then ψεωλ and ψεϕα can be guaranteed by restricting cλ if necessary. Hence there exists a minimal solution vλε to (3.12) by using the method of monotone iteration again, which satisfies ψε vλε ωλ.

Now, let us define

Λ1(λ)=infϕH01(Ω),Ωϕ2=1Ω|ϕ|2λa(x)[h(ωλ)αh(ωλ)]ϕ2dxΩb(x)[h(ωλ)βh(ωλ)]ϕ2dx (3.13)

and

Λ1ε(λ)=infϕH01(Ω),Ωϕ2=1Ω|ϕ|2dxΩλa(x)[h(vλε+ε)αh(vλε+ε)]ϕ2dxΩb(x)[h(vλε)βh(vλε)]ϕ2dx, (3.14)

where ωλ is the minimal solution of (2.1), λ ∈ (0, Λ).

Let φλεH01(Ω) be a nonnegative minimiser of (3.14). We can obtain Λ1ε (λ) ≥ 0 for any λ ∈ (0, Λ). Indeed, assume that Λ1ε (λ) < 0 for some ε > 0 and λ ∈ (0, Λ). It is easy to verify that vλεμφλε is a super-solution of (3.12) for μ > 0 small enough. Using the method of monotone iteration again, we can conclude that there exists a solution to (3.12), λ, such that ψεv~λvλεμφλε, which is a contradiction with vλε is a minimal solution. It follows from vλε ωλ and elliptic regularity, that there exists vλ such that vλε vλ in 𝓒loc(Ω), that is, vλ solves (2.1) and vλωλ by the minimality of ωλ.

Furthermore, we can obtain Λ1(λ) ≥ 0 for any λ ∈ (0, Λ). Indeed, note that vλε + εψε + εcλϕα, −(h(t + ε)αh(t + ε)) = h(t + ε)α−1(h(t + ε))2[α + 2h2(t + ε)(h(t + ε))2]. Applying Lemma 2.1 and Hardy’s inequality, there exists C > 0 such that 2h2(t + ε)(h(t + ε))2C and

Ωh(vλ+ε)α1φλ2cλα1Ωd2φλ2<,

where φλ be the nonnegative minimiser of (3.13). Applying dominated convergence theorem, we have

Λ1(λ)=Ω|φλ|2dx+λa(x)h(ωλ)α1(h(ωλ))2[α+2h2(ωλ)(h(ωλ))2]φλ2dxΩb(x)[h(ωλ)βh(ωλ)]φλ2dx=Ω|φλ|2dx+λa(x)h(vλε+ε)α1(h(vλε+ε))2[α+2h2(vλε+ε)(h(vλε+ε))2]φλ2dxΩb(x)[h(vλε)βh(vλε)]φλ2dx+oε(1)Λ1ε+oε(1),

which yields Λ1(λ) ≥ 0 for any 0 < λ < Λ.

Now we prove Λ1(λ) > 0. Suppose there exists some λ0 ∈ (0, Λ) such that Λ1(λ0) = 0. Without loss of generality, we may assume that Λ1(λ) > 0 for λ0 < λ < Λ. Then we have

0=Ω|ϕλ0|2dxλ0Ωa(x)[h(ωλ0)αh(ωλ0)]ϕλ02dxΩb(x)[h(ωλ0)βh(ωλ0)]ϕλ02dx,

which implies

λ0Ωa(x)[h(ωλ0)αh(ωλ0)]ϕλ02dx+Ωb(x)[h(ωλ0)βh(ωλ0)]ϕλ02dx>0.

For any λ < λ0, combining 0 < a(x) ∈ 𝓒(Ω) with the monotonicity of h(t)αh(t), we obtain

Ω|ϕλ0|2dxλΩa(x)[h(ωλ0)αh(ωλ0)]ϕλ02dxΩb(x)[h(ωλ0)βh(ωλ0)]ϕλ02dx<0,

which implies Λ1(λ) < 0 for λ < λ0, a contradiction with Λ1(λ) ≥ 0 for all λ ∈ (0, Λ). Therefore, Λ1(λ) > 0 for all 0 < λ < Λ.

Suppose ωF(λ, ωλ) is not invertible for some λ ∈ (0, Λ). Then there exists φ ∈ 𝓒2(Ω) ∩ 𝓒ϕα(Ω) with Ω φ2 = 1 satisfying

Δφλa(x)[h(ωλ)αh(ωλ)]φb(x)[h(ωλ)βh(ωλ)]φ=0. (3.15)

From Lemma 3.10, we have φ H01 (Ω) is a H1-weak solution of (3.15), that is, Λ1(λ) = 0, which is a contradiction with the above attained. Then we can apply the implicit function theorem at any (λ, ωλ) for λ ∈ (0, Λ) to obtain that the minimal solution branch is parametrised by an analytic map.□

Lemma 3.13

There exist closed and bounded subsets of 𝓢 = {(λ, ω) ∈ ℝ+ × Cϕα+ (Ω) : F(λ, ω) = 0} are compact.

Proof

Assume that (λ, ω) ∈ 𝓢 and ω solves (2.1). We claim that

infSinfΩ(ωλ1α+1ϕα)c (3.16)

for some positive constant c. Since ωωλ, where ωλ is obtained as above, then we can find a sufficiently small constant c > 0 such that cλ1α+1ϕα is a sub-solution of (2.1) for λ > 0. Thus ωλ cλ1α+1ϕα , which implies the claim is true.

Let 𝓞 be a closed and bounded subset of 𝓢. Then there exists a constant M > 0 such that λ + ∥ω𝓒ϕα(Ω)} ≤ M for all (λ, ω) ∈ 𝓞. It follows from (3.16) and the properties of h(t), that there exists small enough c such that

|Δω|λa(x)h(cλ1α+1ϕα)αh(cλ1α+1ϕα)+|b|sup[0,M]h(ω)βh(ω). (3.17)

Now, according to Lemma 2.1-(8), if ∣ cλ1α+1ϕα ∣ ≤ 1, the another case is similar, then together with Lemma 2.1-(3), we can continue to calculate (3.17) as follows:

|Δω|λa(x)h(1)α(cλ1α+1ϕα)α+|b|sup[0,M]h(ω)βh(ω)λ11+αa(x)h(1)αcαϕαα+|b|sup[0,M]h(ω)βh(ω)M11+αa(x)h(1)αcαϕαα+|b|sup[0,M]h(ω)βh(ω). (3.18)

By the above estimate and applying Proposition 3.4 of [17], it is easy to see

supQωCτ(Ω¯)<forsomeτ(0,1).

Let {(λn,ωn)} ⊂ 𝓞. Then there exists, up to a subsequence, (λn,ωn) → (λ0,ω0) in 𝓒(Ω). We claim that (λn,ωn) → (λ0, ω0) in 𝓒ϕα(Ω). First, we claim λ0 ≠ 0. Otherwise, we have ωλn H01 (Ω) satisfies ∥ωλn∥ → 0. By the above Lemma, ωλn → 0 in 𝓒ϕα(Ω). By Lemma 3.8, we have ωλn = ωn and implies (λn,ωn) → (0, 0) in ℝ+ × 𝓒ϕα(Ω), which is a contradiction with (0, 0) does not belonging to 𝓞. Thus λ0 > 0. Then we have −Δω0 Cϕαα+ (Ω) by combining inequality (3.18) with bounded ωn in 𝓒ϕα(Ω). Furthermore, by Lemma 3.4, it follows that ω0 Cϕα+ (Ω). Let vn = ω0ωn, in virtue of (3.16), we have vn solves

Δvn+[λ0a(x)(h(ξn)αh(ξn))b(x)(h(ζn)βh(ζn))]vn=(λ0λn)a(x)h(ωn)αh(ωn)=o(1)ϕαα,

where ξn and ζn lie between ωn and ω0. Applying Corollary 3.6 with

m(x)=λ0a(x)(h(ξn)αh(ξn))b(x)(h(ζn)βh(ζn)),

and meets the hypothesis what m(x) needs if b ≥ 0 small enough, and then we can obtain that ωnω0 in 𝓒ϕα(Ω). This implies the Lemma holds.□

Remark 3.14

It is clear that the above lemma remains true if b(x) < 0.

Next we consider the bifurcation analysis at Λ = λ.

Lemma 3.15

The solution of F(λ, ω) = 0 near (Λ, ωΛ) is described by a curve (λ(s), ω(s)) = (Λ + τ(s), ωΛ + Λ + x(s), where s → (τ(s), x(s)) ∈ ℝ × X is a continuously differentiable function near s = 0 with τ(0) = τ(0) = 0, x(0) = x(0) = 0. Furthermore, τ is of class 𝓒2 near 0 and τ(0) < 0 if α 5 -2 and b+ ≠ 0.

Proof

From Lemma 2.6, we have the following function at λ = Λ:

ωF(Λ,ωΛ)=IΛ(Δ)1a(x)[h(ωΛ)αh(ωΛ](Δ)1b(x)[h(ωΛ)βh(ωΛ].

Then we have Λ1(Λ) ≥ 0 by the above obtained, and in fact Λ1(Λ) = 0 by the implicit function theorem and (2.1) has no solution for λ > Λ. We now verify the conditions of the local bifurcation result of Cranduall-Rabinowitz [9]. From obtained above, the map ωF(λ, ω) is Fredholm with index 0 for all (λ, ω) ∈ ℝ+ × Cϕα+ (Ω), then we can easily get ker(ωF(Λ, ωΛ)) is one dimensional and spanned by φΛ which is the associated eigenfunction of Λ. We can also get codimRange(ωF(Λ, ωΛ)) = 1. Now we claim that λF(Λ, ωΛ) ∉ Range(ωF(Λ, ωΛ)), where λF(λ, ω)v = −(−Δ)−1a(x)h(ω)αh(ω)v. If it is not true, then there exists v C01 (Ω) such that

v(Δ)1[Λa(x)(h(ωΛ)αh(ωΛ))vb(x)(h(ωΛ)βh(ωΛ))v]=(Δ)1a(x)h(ωΛ)αh(ωΛ).

By Lemma 3.10, we have v ∈ 𝓒2(Ω) ∩ 𝓒φ1(Ω) solves

Δv=Λa(x)(h(ωΛ)αh(ωΛ))vb(x)(h(ωΛ)βh(ωΛ))va(x)h(ωΛ)αh(ωΛ).

Multiplying the above equation by ϕΛ and integrating by parts, we obtain

Ωa(x)h(ωΛ)αh(ωΛ)ϕΛdx=0,

which is a contradiction with the properties of a(x) and h(t). Thus the claim holds.

Now, let X be any complement of the span {ϕΛ} in C01 (Ω). Define θ : ℝ × ℝ × X H01 (Ω) with θ(s, τ, x) = F(Λ + τ, ωΛ + Λ + x). It is easy to see that

τ,xθ(0,0,0)=(λF(Λ,ωΛ),ωF(Λ,ωΛ))

is an isomorphism from ℝ × X onto 𝓒1(Ω) ∩ 𝓒0(Ω). In fact, it follows from

codimRange(ωF(Λ,ωΛ))=1andλF(Λ,ωΛ)Range(ωF(Λ,ωΛ)),

that the map

(τ,x)τλF(Λ,ωΛ)+ωF(Λ,ωΛ)x (3.19)

is injective for (τ, x) ∈ ℝ × X. Then, for ϵ > 0, there exists a neighborhood V of 0 in ℝ and a unique 𝓒2 function g : (−ϵ, ϵ) → V × X so that g(s) := (τ(s), x(s)) with s ∈ (−ϵ, ϵ), τ(0) = 0, x(0) = 0, θ(τ(s), x(s)) = F(Λ+τ(s), ωΛ+sϕΛx(s)) = 0. Differentiating the expression (3.19) with respect to s at s = 0, we have

τ(0)λF(Λ,ωΛ)+ωF(Λ,ωΛ)(ϕΛ+x(0))=0.

Then τ(0) = 0 and x(0) = 0 since ωF(Λ, ωΛ)ϕΛ = 0. Differentiating the above equation again with respect to s at s = 0, we have

τ(0)λF(Λ,ωΛ)+ωF(Λ,ωΛ)x(0)+ωωF(Λ,ωΛ)(ϕΛ,ϕΛ)=0.

Applying the dual product with −ΔϕΛ, we obtain

τ(0)λF(Λ,ωΛ)+ωF(Λ,ωΛ)x(0)+ωωF(Λ,ωΛ)(ϕΛ,ϕΛ),ΔϕΛ=τ(0)(Δ)1a(x)h(ωΛ)αh(ωΛ)+{x(0)Λ(Δ)1[a(x)(h(ωΛ)αh(ωΛ))x(0)](Δ)1[b(x)h(ωΛ)βh(ωΛ)x(0)]}Λ(Δ)1[a(x)(h(ωΛ)αh(ωΛ))ϕΛ2](Δ)1[b(x)(h(ωΛ)βh(ωΛ))ϕΛ2],ΔϕΛ=τ(0)(a(x)h(ωΛ)αh(ωΛ),ϕΛ+x(0),ΔϕΛΛa(x)(h(ωΛ)αh(ωΛ)ϕΛb(x)(h(ωΛ)βh(ωΛ))ϕΛΛa(x)(h(ωΛ)αh(ωΛ))ϕΛ2+b(x)(h(ωΛ)βh(ωΛ))ϕΛ2,ϕΛ=τ(0)(a(x)h(ωΛ)αh(ωΛ),ϕΛΛa(x)(h(ωΛ)αh(ωΛ))ϕΛ2+b(x)(h(ωΛ)βh(ωΛ))ϕΛ2,ϕΛ=0.

By computing, we can easily get a(x)(h(ω)αh(ω))+b(x)(h(ω)βh(ω)) > 0 for α 5 -2 and b+ ≠ 0, and hence τ(0) < 0.□

Remark 3.16

  1. From the above results, we can infer that the direction of the solution curve at the neighborhood λ = Λ. That is, τ(0) < 0 means the curve of solution from right turns to left at {λ = Λ}.

  2. If 0 < α < 1/3 and b < 0, there holds λ a(x)(h(ω)αh(ω))+b(x)(h(ω)βh(ω)) < 0, and hence τ(0) > 0. The case of τ(0) > 0 needs further analysis.

  3. The sign of τ(0) is indefinite when 0 < α < 5 -2 and b+ ≠ 0 or 1/3 ≤ α < 1 and b < 0.

3.3 The proof of main results

Proof

Proof of Theorem 2.7. Let 𝓤 = ℝ+ × 𝓧 = ℝ+ × 𝓒ϕα(Ω) and the positive cone 𝓦 = Cϕα+ (Ω). Clearly 𝓦 is open. Conditions (H1)-(H3) of Lemma 2.5 hold because of Lemma 3.13, Proposition 3.1, Lemma 3.7 and Lemma 3.12. In fact, by Lemma 3.12, we may fix 𝓐+ = {(λ(s), w(s)) : 0 < s < s0} for some s0 > 0 be an analytic parametrisation which is one portion of minimal solution branch which given by {(λ, ωλ) ∈ Γ : 0 < λ < λ0}. Then Lemma 2.5 holds. Next, we apply Lemma 2.5 to prove Theorem 2.7. It is clear that assertions (iii) and (vi) of Theorem 2.7 are true. From the definition of 𝓐 and 𝓐+, assertion (i) easily obtained. Assertion (iv) can be get from Lemma 3.8 and Lemma 3.12.

It remains to prove assertion (ii), clearly, (v) is a consequence of (ii). In order to show assertion (ii), we only need verify the property (e)(i) of Lemma 2.5 occurring. If case (e)(ii) is occur, then there exists (λ(sn), ω(sn)) → (0, ω0) as sn → ∞ in 𝓤, where (0, ω0) is a boundary point. Then Δω(sn) → 0 in 𝓒loc(Ω), i.e. ω0 = 0 and hence ω0 ≡ 0. By Lemma 3.8, it is easy to see that ω(sn) is the minimal solution for all large sn. However, the minimal solution arc 𝓐0 starting from (0, 0) is isolated from other solutions, and hence, the distinguished arc corresponding to all large s coincide with 𝓐, which is a contradiction with (a) of Lemma 2.5. By the same argument, we can rule out case (e)(iii). Hence case (e)(i) holds. Therefore, ∥ω(s)∥𝓒ϕα(Ω) → ∞ as s → ∞ since (2.1) has no solution for λ > Λ. This completes the proof.□

Proof

Proof of Corollary 2.8. It follows from Lemma 3.15 and Lemma 2.5, that one can easily get the analytic path 𝓐 turns to the left at the point (Λ, ωΛ).□

Acknowledgement

Minbo Yang is the corresponding author who was partially supported by NSFC(11971436, 12011530199) and ZJNSF(LD19A010001).

  1. Conflict of interest: Authors state no conflict interest.

References

[1] Adimurthi and J. Giacomoni, Multiplicity of positive solutions for a singular and critical elliptic problem in ℝ2, Commun. Contemp. Math. 8 (2006), 621-656.10.1142/S0219199706002222Search in Google Scholar

[2] D. Arcoya and L. Moreno-Mérida, Multiplicity of solutions for a Dirichlet problem with a strongly singular nonlinearity, Nonlinear Anal. 95 (2014), 281-291.10.1016/j.na.2013.09.002Search in Google Scholar

[3] C.O. Alves, Y. Wang and Y. Shen, Soliton solutions for a class of quasilinear Schrödinger equations with a parameter, J. Differential Equations 259 (2015), 318-343.10.1016/j.jde.2015.02.030Search in Google Scholar

[4] B. Buffoni, E.N. Dancer and J.F. Toland, The regularity and local bifurcation of steady periodic water waves, Arch. Ration. Mech. Anal. 152 (2000), 207-240.10.1007/s002050000086Search in Google Scholar

[5] B. Bougherara, J. Giacomoni and S. Prashanth, Analytic global bifurcation and infinite turning points for very singular problems, Calc. Var. 52 (2015), 829-856.10.1007/s00526-014-0735-8Search in Google Scholar

[6] Yunru Bai, Dumitru Motreanu and Shengda Zeng, Continuity results for parametric nonlinear singular Dirichlet problems, Adv. Nonlinear Anal. 9 (2020), 372-387.10.1515/anona-2020-0005Search in Google Scholar

[7] B. Buffoni and J.F. Toland, Analytic Theory of Global Bifurcation. An introduction, Princeton Series in Applied Mathematics, Princeton University Press, Princeton (2003).10.1515/9781400884339Search in Google Scholar

[8] M.M. Coclite and G. Palmieri, On a singular nonlinear Dirichlet problem, Comm. Partial Differential Equations 14 (1989), 1315-1327.10.1080/03605308908820656Search in Google Scholar

[9] M.G. Crandall and P.H. Rabinowitz, Bifurcation from simple eigenvalues, J. Funct. Anal. 8 (1971), 321-340.10.1016/0022-1236(71)90015-2Search in Google Scholar

[10] M.G. Crandall, P.H. Rabinowitz and L. Tartar, On a Dirichlet problem with a singular nonlinearity, Comm. Partial Differential Equations 2 (1977), 193-222.10.1080/03605307708820029Search in Google Scholar

[11] R. Dhanya, J. Giacomoni, S. Prashanth and K. Saoudi, Global bifurcation and local multiplicity results for elliptic equations with singular nonlinearity of super exponential growth in ℝ2, Adv. Differential Equations 3 (2012).10.57262/ade/1355703090Search in Google Scholar

[12] Y. Ding, F. Gao and M. Yang, Semiclassical states for Choquard type equations with critical growth: critical frequency case, Nonlinearity 33 (2020), 6695-6728.10.1088/1361-6544/aba88dSearch in Google Scholar

[13] L. Du and M. Yang. Uniqueness and nondegeneracy of solutions for a critical nonlocal equation. Discrete Contin. Dyn. Syst. 39 (2019), 5847-5866.10.3934/dcds.2019219Search in Google Scholar

[14] L. Dupaigne, M. Ghergu and V. Rădulescu, Lane-Emden-Fowler equations with convection and singular potential, J. Math. Pures Appl. 87 (2007), 563-581.10.1016/j.matpur.2007.03.002Search in Google Scholar

[15] J.I. Díaz and J.M. Rakotoson, On the differentiability of very weak solutions with right-hand side data integrable with respect to the distance to the boundary, J. Funct. Anal. 257 (2009), 807-831.10.1016/j.jfa.2009.03.002Search in Google Scholar

[16] J.M. do Ó and A. Moameni, Solutions for singular quasilinear Schrödinger equations with one parameter, Commun. Pure Appl. Anal. 9(2010), 1011-1023.10.3934/cpaa.2010.9.1011Search in Google Scholar

[17] C.F. Gui and F.H. Lin, Regularity of an elliptic problem with a singular nonlinearity, Proc. Roy. Soc. Edinburgh Sect. A 123 (1993), 1021-1029.10.1017/S030821050002970XSearch in Google Scholar

[18] M. Ghergu and V. Rădulescu, Sublinear singular elliptic problems with two parameters, J. Differential Equations 195 (2003), 520-536.10.1016/S0022-0396(03)00105-0Search in Google Scholar

[19] M. Ghergu and V. Rădulescu, Multi-parameter bifurcation and asymptotics for the singular Lane-Emden-Fowler equation with a convection term, Proc. Roy. Soc. Edinburgh Sect. A 135 (2005), 61-83.10.1017/S0308210500003760Search in Google Scholar

[20] M. Ghergu and V. Rădulescu, Singular Elliptic Problems: Bifurcation and Asymptotic Analysis, Oxford Lecture Ser. Math. Appl. 37 (2008), Oxford University Press.10.1093/oso/9780195334722.003.0002Search in Google Scholar

[21] D. Gilbarg, N.S. Trudinger, Elliptic partial differential equations of second order, Springer-Verlag, Berlin (1977).10.1007/978-3-642-96379-7Search in Google Scholar

[22] N. Hirano, C. Saccon and N. Shioji, Existence of multiple positive solutions for singular elliptic problems with concave and convex nonlinearities, Adv. Differential Equations 9 (2004), 197-220.10.57262/ade/1355867973Search in Google Scholar

[23] P. Lindqvist, On the equation div(∣∇ up−2u) + λup−2u = 0, Proc. Amer. Math. Soc. 109 (1990), 157-164.10.1090/S0002-9939-1990-1007505-7Search in Google Scholar

[24] J.Q. Liu, X.Q. Liu and Z.Q. Wang, Quasilinear elliptic equations via perturbation method, Proc. Amer. Math. Soc. 141 (2013), 253-263.10.1090/S0002-9939-2012-11293-6Search in Google Scholar

[25] J.Q. Liu, X.Q. Liu and Z.Q. Wang, Quasilinear elliptic equations with critical growth via perturbation method, J. Differential Equations 254 (2013), 102-124.10.1016/j.jde.2012.09.006Search in Google Scholar

[26] J.Q. Liu, X.Q. Liu and Z.Q. Wang, Multiple sign-changing solutions for quasilinear elliptic equations via perturbation method, Comm. Partial Differential Equations 39 (2014), 2216-2239.10.1080/03605302.2014.942738Search in Google Scholar

[27] A.C. Lazer and P.J. McKenna, On a singular nonlinear elliptic boundary-value problem, Proc. Amer. Math. Soc. 111 (1991), 721-730.10.1090/S0002-9939-1991-1037213-9Search in Google Scholar

[28] A.V. Lair and A.W. Shaker, Classical and weak solutions of a singular semilinear elliptic problem, J. Math. Anal. Appl. 211 (1997), 371-385.10.1006/jmaa.1997.5470Search in Google Scholar

[29] J.Q. Liu and Z.Q. Wang, Soliton solutions for quasilinear Schrödinger equations I, Proc. Amer. Math. Soc 131 (2002), 441-448.10.1090/S0002-9939-02-06783-7Search in Google Scholar

[30] J.Q. Liu, Y.Q. Wang and Z.Q. Wang, Soliton solutions for quasilinear Schrödinger equations II, J. Differential Equations 187 (2003), 473-493.10.1016/S0022-0396(02)00064-5Search in Google Scholar

[31] J.Q. Liu, Y.Q. Wang and Z.Q. Wang, Solutions for quasilinear Schrödinger equations via the nehari method, Comm. Partial Differential Equations 29 (2004), 879-901.10.1081/PDE-120037335Search in Google Scholar

[32] S. Liu and J. Zhou, Standing waves for quasilinear Schrödinger equations with indefinite potentials, J. Differential Equations 265 (2018), 3970-3987.10.1016/j.jde.2018.05.024Search in Google Scholar

[33] A. Moameni and D.C. Offin, Positive solutions for singular quasilinear Schrödinger equations with one parameter, II, J. Partial Differential Equations 23 (2010), 223-234.10.4208/jpde.v23.n3.2Search in Google Scholar

[34] L. Ma and J.C. Wei, Properties of positive solutions to an elliptic equation with negative exponent, J. Funct. Anal. 254 (2008), 1058-1087.10.1016/j.jfa.2007.09.017Search in Google Scholar

[35] Nikolaos S. Papageorgiou, Vicenţiu D. Rădulescu and Dušan D. Repovš, Positive solutions for nonlinear parametric singular Dirichlet problems, Bull. Math. Sci. 9 (2019), no.3, 1950011, 21pp.10.1142/S1664360719500115Search in Google Scholar

[36] Nikolaos S. Papageorgiou, Vicenţiu D. Rădulescu and Dušan D. Repovš, Nonlinear nonhomogeneous singular problems, Calc. Var. Partial Differential Equations 59 (2020), no.1, Paper No.9, 31pp.10.1007/s00526-019-1667-0Search in Google Scholar

[37] G. dos Santos, G.M. Figueiredo and U.B. Severo, Multiple solutions for a class of singular quasilinear problems, J. Math. Anal. Appl. 480 (2019), 123405.10.1016/j.jmaa.2019.123405Search in Google Scholar

[38] Y.J. Sun, S.P. Wu and Y.M. Long, Combined effects of singular and superlinear nonlinearities in some singular boundary value problems, J. Differential Equations 176 (2001), 511-531.10.1006/jdeq.2000.3973Search in Google Scholar

[39] C.A. Santos, M.B. Yang and J.Z. Zhou, Global multiplicity of solutions for a modified elliptic problem with singular terms, preprint.10.1088/1361-6544/ac2a50Search in Google Scholar

[40] Y.J. Sun and D.Z. Zhang, The role of the power 3 for elliptic equations with negative exponents, Calc. Var. 49 (2014), 909-922.10.1007/s00526-013-0604-xSearch in Google Scholar

[41] L.L. Wang, Existence and uniqueness of solutions to singular quasilinear Schrödinger equations, Elec. J. Differential Equations 38 (2018), 1-9.Search in Google Scholar

[42] X. Wu, Multiple solutions for quasilinear Schrödinger equations with a parameter, J. Differential Equations 256 (2014), 2619-2632.10.1016/j.jde.2014.01.026Search in Google Scholar

[43] H.T. Yang, Multiplicity and asymptotic behavior of positive solutions for a singular semilinear elliptic problem, J. Differential Equations 189 (2003), 487-512.10.1016/S0022-0396(02)00098-0Search in Google Scholar

[44] Minbo Yang, Fukun Zhao and Shunneng Zhao, Classification of solutions to Hartree equation with double Hardy-Littlewood-Sobolev critical parts, Discrete Contin. Dyn. Syst. A 41 (2021), 5209–5241.10.3934/dcds.2021074Search in Google Scholar

[45] Minbo Yang, Xianmei Zhou, On a coupled Schrodinger system with Stein-Weiss type convolution part, The Journal of Geometric Analysis 31 (2021), 10263-10303.10.1007/s12220-021-00645-wSearch in Google Scholar

[46] Y. Zhen, F. Gao, Z. Shen, M. Yang, On a class of coupled critical Hartree system with deepening potential, Math. Meth. Appl. Sci. 44 (2021), 772-798.10.1002/mma.6785Search in Google Scholar

Received: 2021-06-30
Accepted: 2021-10-23
Published Online: 2021-11-29

© 2021 Siyu Chen et al., published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 9.6.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2021-0215/html
Scroll to top button