Loading [MathJax]/jax/output/SVG/jax.js
Research article

Bifurcation results of positive solutions for an elliptic equation with nonlocal terms

  • Received: 10 April 2021 Accepted: 17 June 2021 Published: 24 June 2021
  • MSC : 35J25, 35J61, 35B32, 35B45

  • In this paper, we investigate the local and global nature for the connected components of positive solutions set of an elliptic equation with nonlocal terms. The local bifurcation results of positive solutions are obtained by using the local bifurcation theory, Lyapunov-Schmidt reduction technique, etc. Under suitable conditions, we show two proofs of priori estimates by using blow-up technique, upper and lower solution method, etc. Finally, the global bifurcation results of positive solutions are obtained by using priori bounds, global bifurcation theory.

    Citation: Jiaqing Hu, Xian Xu, Qiangqiang Yang. Bifurcation results of positive solutions for an elliptic equation with nonlocal terms[J]. AIMS Mathematics, 2021, 6(9): 9547-9567. doi: 10.3934/math.2021555

    Related Papers:

    [1] Lin Zhao . Monotonicity and symmetry of positive solution for 1-Laplace equation. AIMS Mathematics, 2021, 6(6): 6255-6277. doi: 10.3934/math.2021367
    [2] Liye Wang, Wenlong Wang, Ruizhi Yang . Stability switch and Hopf bifurcations for a diffusive plankton system with nonlocal competition and toxic effect. AIMS Mathematics, 2023, 8(4): 9716-9739. doi: 10.3934/math.2023490
    [3] CaiDan LaMao, Shuibo Huang, Qiaoyu Tian, Canyun Huang . Regularity results of solutions to elliptic equations involving mixed local and nonlocal operators. AIMS Mathematics, 2022, 7(3): 4199-4210. doi: 10.3934/math.2022233
    [4] Changmu Chu, Yuxia Xiao, Yanling Xie . Infinitely many sign-changing solutions for a semilinear elliptic equation with variable exponent. AIMS Mathematics, 2021, 6(6): 5720-5736. doi: 10.3934/math.2021337
    [5] Mengyu Wang, Xinmin Qu, Huiqin Lu . Ground state sign-changing solutions for fractional Laplacian equations with critical nonlinearity. AIMS Mathematics, 2021, 6(5): 5028-5039. doi: 10.3934/math.2021297
    [6] Hui Wang, Xue Wang . Bifurcations of traveling wave solutions for the mixed Korteweg-de Vries equation. AIMS Mathematics, 2024, 9(1): 1652-1663. doi: 10.3934/math.2024081
    [7] Choukri Derbazi, Hadda Hammouche . Caputo-Hadamard fractional differential equations with nonlocal fractional integro-differential boundary conditions via topological degree theory. AIMS Mathematics, 2020, 5(3): 2694-2709. doi: 10.3934/math.2020174
    [8] Labudan Suonan, Yonglin Xu . Existence of solutions to mixed local and nonlocal anisotropic quasilinear singular elliptic equations. AIMS Mathematics, 2023, 8(10): 24862-24887. doi: 10.3934/math.20231268
    [9] Xiangrui Li, Shuibo Huang, Meirong Wu, Canyun Huang . Existence of solutions to elliptic equation with mixed local and nonlocal operators. AIMS Mathematics, 2022, 7(7): 13313-13324. doi: 10.3934/math.2022735
    [10] Tomas Godoy . Existence of positive weak solutions for a nonlocal singular elliptic system. AIMS Mathematics, 2019, 4(3): 792-804. doi: 10.3934/math.2019.3.792
  • In this paper, we investigate the local and global nature for the connected components of positive solutions set of an elliptic equation with nonlocal terms. The local bifurcation results of positive solutions are obtained by using the local bifurcation theory, Lyapunov-Schmidt reduction technique, etc. Under suitable conditions, we show two proofs of priori estimates by using blow-up technique, upper and lower solution method, etc. Finally, the global bifurcation results of positive solutions are obtained by using priori bounds, global bifurcation theory.



    In this paper, we consider the elliptic problem with nonlocal terms

    {Δu=λm(x)u+h(x)up+Ωuβ,xΩ,un=0,xΩ, (1.1)

    where Ω is a bounded domain of RN with a smooth boundary, N2; n is the outward unit normal to Ω; m(x), h(x)Cα(ˉΩ) for some α(0,1) and m(x), h(x) may change sign in Ω; p,β>1 and p<N+2N2 for N3, λR is a parameter.

    Many physical phenomena were formulated into nonlocal mathematical models [1,2,3,11,12] and studied by many authors. For example, J. Bebernes and A. Bressan [11] studied an ignition model similar to (1.1) for a compressible reactive gas which is a nonlocal reaction-diffusion equation. In [11], u is the temperature perturbation of the gas and nonlocal term is due to the compressibility of the gas. Subsequently, some researchers [2,3,12] discussed the parabolic problems related to the equation

    {Δu=f(x,u)+g(Ωuβ),xΩ,u=0,xΩ.

    This type of problem is frequently encountered in nuclear reaction process, where it is known that the reaction is very strong, say like f(x,u)=λm(x)u+h(x)up with p>1 and constant functions m(x) and h(x), but the rate with respect to this power is unknown, say like g(Ωuβ)=Ωuβ. The above mathematical problem can also be used to population dynamics and biological science where the total mass is often conserved or known, but the growth of a certain cell is known to be of some form (see [12]). Thus, the problem (1.1) is worthy to be considered.

    Mathematically, the problem (1.1) combines local and nonlocal terms. It is well known [2,3] that the authors discussed that the case m(x)=0, h(x)h0<0, β>1 and p1. In two articles, the parabolic problem related to the equation was studied and the authors showed that the value p=β represents a critical blow-up exponent. They proved that if β>p or β=p and h0>|Ω|, the blow-up phenomenon can occur in finite time. If β<p or β=p and h0|Ω|, all the solutions are global and bounded. The authors also proved the existence of positive solution for h0 small in the particular case h0<0, p>β>1. In [1], F. Corrˆea and A. Suˊarez made a further study for the problem and proved the existence, uniqueness, stability and asymptotic properties of positive solutions for some values of p1 and β>0.

    We want to further consider the global bifurcation structure of the positive solutions set of the problem (1.1) when the function m(x) and h(x) are nonconstant functions because for general function h(x), especially sign-changing function, comparing with the local elliptic equation, we see that many methods that prove the boundedness of positive solutions cannot be used in nonlocal elliptic equation, such as the extremum principle, parameter control. Finally, we can obtain the existence, multiplicity and nonexistence of the positive solution for the problem (1.1) when a bounded connected branch of the positive solutions set is established by the global bifurcation theory.

    In an early paper, K. J. Brown [4] studied the local and global bifurcation of the semilinear elliptic boundary value problem

    {Δu=λm(x)u+b(x)uγ,xΩ,un=0,xΩ, (1.2)

    where 1<γ<N+2N2, m(x), b(x) may change sign in Ω. The cases where Ωm(x)dx0 and Ωm(x)dx=0 were discussed respectively and the author concluded that there are continua of positive solutions of (1.2) connecting λ=0 to the other principal eigenvalue for Ωm(x)dx0 when m(x) and b(x) are under suitable conditions. It was also showed that the closed loops of positive solutions occur naturally and properties of these loops are investigated.

    In this paper, we are interested in the problem (1.1), namely, the problem (1.2) added a nonlocal term. We want to investigate whether the local and global structures of positive solutions set for the problem (1.1) have similar properties to the problem (1.2). We also investigate sufficient conditions for a bounded continuum of positive solutions. In Theorem 2.2, we see that the direction of bifurcation curve is related to β and p. In Theorem 3.1, we get a priori bound of positive solution when m(x) is under suitable conditions and β>max{p,N(p1)/N(p1)22} by using upper and lower solution method, blow-up technique and boot-strapping method. Moreover, another way of proving boundedness shows that the priori bound still exists when β>max{p,N(p1)/N(p1)22} vanishes.

    Before proceeding to the study of local and global nature of positive solutions, we need to introduce some notations. If u>0 in Ω, we say u is a positive solution of the problem (1.1). (λ,u) is called a nonnegative solution of the problem (1.1) if u is a nonnegative solution of the problem (1.1) with λ. Obviously, (λ,0) is a nonnegative solution of the problem (1.1), we say it is a trivial solution.

    To investigate the bifurcation of problem (1.1) at the trivial solution (λ,0), we discuss the linear eigenvalue problem

    {Δw=λm(x)w,xΩ,wn=0,xΩ, (1.3)

    where m(x) changes sign in Ω. According to [5], we have the following results.

    (i) If Ωm<0, the problem (1.3) has the principal eigenvalues λ+>0 and λ0=0.

    (ii) If Ωm>0, the problem (1.3) has the principal eigenvalues λ<0 and λ0=0.

    (iii) If Ωm=0, the problem (1.3) has the unique principal eigenvalue λ0=0.

    The usual norms of the space Lp(Ω) for p[1,) and C(ˉΩ) are, respectively,

    up=(Ω|u(x)|pdx)1/1pp,uC(ˉΩ)=maxˉΩ|u(x)|.

    Let

    Ωh+={xΩ:h(x)>0},Ωh={xΩ:h(x)<0}.

    We use the following hypothesis.

    (H1)Λ={uH10(Ωh+):Ωh+mu2>0}.

    If m(x) changes sign in Ωh+ and (H1) holds, then the equation

    {Δu=λm(x)u,xΩh+,u=0,xΩh+

    has unique positive principal eigenvalue [5,6].

    We have the following main global results of positive solutions in two cases where Ωm<0 and Ωm>0 by using priori bounds, global bifurcation theory.

    Theorem 1.1. Assume Ωm<0 and

    β>p,Ωhφp+11<0,

    where φ1 is the positive eigenfunction of λ+. Then there exists a continuous curve (λ(s),u(s)) of s(0,ε) such that (λ(0),u(0))=(λ+,0),λ(s)>λ+, (λ(s),u(s)) are positive solutions of the problem (1.1), and for any positive solution (λ,u) of the problem (1.1) in a neighborhood of (λ+,0), we have (λ,u)=(λ(s),u(s)).

    Moreover, if m(x) changes sign in Ωh+ and (H1) holds, then the connected branch C+ of positive solutions set containing (λ(s),u(s)) satisfies the following conclusions.

    (i) The projection of C+ on λ-axis is bounded, that is,

    λ=sup{|λ|>0:(λ,u)C+}<+.

    More generally, there is no positive solution of the problem (1.1) for any sufficiently large |λ|.

    (ii) If (λ,u)C+, then u is bounded in C(ˉΩ).

    (iii) The closure ¯C+ of C+ in R×C(ˉΩ) satisfies

    ¯C+{(λ,0):λR,λλ+,0}=.

    The assertions of Theorem 1.1 may be illustrated by the bifurcation diagram shown in Figure 1.

    Figure 1.  Global bifurcation in case Ωm<0.

    Remark 1.1. (i) Theorem 1.1 shows the conditions that the connected branch is supercritical at (λ+,0). In fact, according to Theorem 2.2, we obtain the conditions that the connected branch is supercritical or subcritical at (0,0). In Figure 1, it is clear that if β>p, Ωh<0 (resp. β>p, Ωh>0), the connected branch at (0,0) is subcritical (resp. supercritical). Furthermore, by using Rabinowitz global bifurcation theory, we have that C+ bifurcates from (λ+,0) and backs to (0,0).

    (ii) Comparing with the results of [4], we see that directions of bifurcation curve depend on not only the sign of Ωm and hφp+11 but also the relationship of β and p. In Theorem 1.1, we only show the case of β>p, while the cases of β<p and β=p can also be listed.

    (iii) According to Theorem 1.1, we can obtain the existence, multiplicity and nonexistence of positive solutions for the problem (1.1). For example, the conditions of Theorem 1.1 hold and Ωh<0, then there exist constants σ1<0<λ+<σ2, such that

    the problem (1.1) has at least one positive solution for λ[0,λ+];

    the problem (1.1) has at least two positive solutions for λ(σ1,0)(λ+,σ2).

    the problem (1.1) has no positive solution for λ(,σ1)(σ2,+).

    These results are clearly shown in Figure 1.

    Similarly, we have the global bifurcation of the positive solutions set for Ωm>0.

    Theorem 1.2. Assume Ωm>0 and

    β>p,Ωhφp+11<0,

    where φ1 is the positive eigenfunction of λ. Then there exists a continuous curve (λ(s),u(s)) of s(0,ε) such that (λ(0),u(0))=(λ,0),λ(s)<λ, (λ(s),u(s)) are positive solutions of the problem (1.1), and for any positive solution (λ,u) of the problem (1.1) in a neighborhood of (λ,0), we have (λ,u)=(λ(s),u(s)).

    Moreover, if m(x) changes sign in Ωh+ and (H1) holds, then the connected components C+ of positive solutions set containing (λ(s),u(s)) satisfies the claims (i), (ii) of Theorem 1.1 and the closure ¯C+ of C+ in R×C(ˉΩ) satisfies

    ¯C+{(λ,0):λR,λλ,0}=.

    The assertions of Theorem 1.2 may be illustrated by the bifurcation diagram shown in Figure 2.

    Figure 2.  Global bifurcation in case Ωm>0.

    Next, we consider the case Ωm=0. We shall then study the global bifurcation of positive solutions for the problem (1.1) by the approximation method. Let mε(x)=m(x)ε for ε>0, we have the following conclusions.

    Theorem 1.3. Assume Ωm=0, p=2 or 3, m(x) changes sign in Ωh+ and (H1) holds. If

    2NN2>β>p,Ωh<0,

    then there exists a connected components C+ of positive solutions set for the problem (1.1), which bifurcates from the origin and backs to the origin in λ-norm plane, namely, the closure ¯C+ of C+ in R×C(ˉΩ) is a closed loop.

    Remark 1.2. For the case Ωm=0, the hypotheses for the Crandall and Rabinowitz theorem are no longer satisfied, and we use the Lyapunov-Schmidt technique to investigate how bifurcation occurs at (λ,u)=(0,0). Moreover, comparing with the global bifurcation results of [4], we see that the relationship of p and nonlocal term power β also has influence on the continuum.

    The assertions of Theorem 1.3 may be illustrated by the bifurcation diagram shown in Figure 3.

    Figure 3.  ¯C+ε approaches a closed loop as ε0.

    The rest of this article is organized as follows. In Section 2, we discuss the local properties of positive solutions set in the cases where Ωm0 and Ωm=0 by using the local bifurcation theory, Liapunov-Schmidt reduction technique. In Section 3, we show that a priori estimate of positive solutions by using the blow-up technique, global bifurcation theory and upper and lower solution. In Section 4, we complete the proof of main results in two cases.

    Let's investigate local bifurcation in the cases where Ωm0 and Ωm=0, respectively.

    If Ωm<0, let λ1=λ+ or λ0; if Ωm>0, let λ1=λ or λ0. Let φ1 be the positive eigenfunction of λ1. If λ1=0, φ1 is a constant and we take φ1=1. Let 0<ε1 be a constant. We have the following result.

    Theorem 2.1. Assume Ωm0, then there exists a bifurcation curve (λ(s),u(s)) of positive solutions of the problem (1.1) at (λ1,0) parameterized by s(0,ε), which satisfies

    λ(s)=λ1+γ(s),u(s)=s(φ1+z(s)),z()Z,span{φ1}Z=C2+θ(ˉΩ).

    Here, γ(0)=0, z(0)=0. γ(s) and z(s) is analytic at s=0.

    Proof. We define the mapping

    F:R×XCθ(Ω),
    F(λ,u)=Δu+λm(x)u+h(x)up+Ωuβ, (2.1)

    where X={uC2+θ(Ω):un=0,xΩ}.

    For any wX, we have

    Fu(λ,u)[w]=Δw+λm(x)w+ph(x)up1w+βΩuβ1w,

    then

    Fu(λ,0)[w]=Δw+λm(x)w.

    Thus, we have

    ker(Fu(λ1,0))=span{φ1},dim(ker(Fu(λ1,0)))=1.

    The range of Fu(λ1,0) is R(Fu(λ1,0))={uCθ(Ω):Ωuφ1=0}, so

    codimR(Fu(λ1,0))=1.

    Next, we prove

    Fλu(λ1,0)[φ1]R(Fu(λ1,0)). (2.2)

    Since Fλu(λ1,0)[φ1]=m(x)φ1, we have

    (i) if λ1=0, then φ1=1 and m(x)φ1R(Fu(λ1,0));

    (ii) if λ10, then Ω|φ1|2=λ1Ωmφ21>0, namely Ωmφ210. Thus, we have m(x)φ1R(Fu(λ1,0)).

    Hence, we get (2.2). By virtue of the Crandall-Rabinowitz local bifurcation theory, we obtain Theorem 2.1.

    Next, we discuss the direction of bifurcation.

    Theorem 2.2. Assume Ωm0. (λ(s),u(s)) for s(0,ε) is a bifurcation curve of positive solutions obtained by Theorem 2.1, then we have the following conclusions.

    (1) λ1=0.

    (i) β=p, then

    lims0γ(s)sp1=|Ω|2+ΩhΩm.

    If |Ω|2+Ωh and Ωm have same (resp. opposite) sign, then the bifurcation curve at (λ1,0) is subcritical (resp. supercritical).

    (ii) β<p, then

    lims0γ(s)sβ1=|Ω|2Ωm.

    If Ωm>0(resp.<0), then the bifurcation curve at (λ1,0) is subcritical (resp. supercritical).

    (iii) β>p, then

    lims0γ(s)sp1=ΩhΩm.

    If Ωh and Ωm have same (resp. opposite) sign, then the bifurcation curve at (λ1,0) is subcritical (resp. supercritical).

    (2) λ10.

    (i) β=p, if Ωhφp+11+Ωφ1Ωφp1 and Ωm have opposite (resp. same) sign, then the bifurcation curve at (λ1,0) is subcritical (resp. supercritical).

    (ii) β<p, if Ωφ1Ωφβ1 and Ωm have opposite (resp. same) sign, then the bifurcation curve at (λ1,0) is subcritical (resp. supercritical).

    (iii) β>p, if Ωhφp+11 and Ωm have opposite (resp. same) sign, then the bifurcation curve at (λ1,0) is subcritical (resp. supercritical).

    Proof. Since (λ(s),u(s)) is a positive solution of the problem (1.1), we have

    {Δ[s(φ1+z(s))]=λ(s)ms(φ1+z(s))+hsp(φ1+z(s))p+Ω[s(φ1+z(s))]β,xΩ(s(φ1+z(s)))n=0,xΩ,

    then

    {Δz(s)=γ(s)mφ1+λ(s)mz(s)+hsp1(φ1+z(s))p+sβ1Ω(φ1+z(s))β,xΩ,(z(s))n=0,xΩ. (2.3)

    Multiplying the Eq (2.3) by φ1, integrating in Ω, and using the Green formula, it follows that

    0=Ωγ(s)mφ21+Ωγ(s)mzφ1+sp1Ωh(φ1+z(s))pφ1+sβ1Ωφ1Ω(φ1+z(s))β.

    (i) If β=p, we get

    γ(s)sp1=Ωh(x)(φ1+z(s))pφ1+Ωφ1Ω(φ1+z(s))pΩmφ21+Ωmzφ1.

    So

    lims0γ(s)sp1=Ωhφp+11+Ωφ1Ωφp1Ωmφ21.

    (ii) If β<p, we get

    γ(s)sβ1=spβΩh(φ1+z(s))pφ1+Ωφ1Ω(φ1+z(s))βΩmφ21+Ωmzφ1.

    So

    lims0γ(s)sβ1=Ωφ1Ωφβ1Ωmφ21.

    (iii) If β>p, we get

    γ(s)sp1=Ωh(φ1+z(s))pφ1+sβpΩφ1Ω(φ1+z(s))βΩmφ21+Ωmzφ1.

    So

    lims0γ(s)sp1=Ωhφp+11Ωmφ21.

    Since Ωmφ21 and Ωm have opposite sign, then we get Theorem 2.2.

    If Ωm=0, (2.2) doesn't work, then the hypotheses for the Crandall and Rabinowitz theorem are no longer satisfied. However we can use the Lyapunov-Schmidt technique to investigate how bifurcation occurs.

    Assume uX is the solution of the problem (1.1), let u=s+w, where s is a constant and Ωw=0. Let Q be the projection of X onto W, where W={wX,Ωw=0}. Then w=Q[u]=u1|Ω|Ωu, so u=s+w is the solution of the problem (1.1) if and only if

    Δw=λm(x)(s+w)+h(x)(s+w)p+Ω(s+w)β.

    The condition

    Q[Δw]=Q[λm(x)(s+w)+h(x)(s+w)p+Ω(s+w)β],

    implies that

    Δw=Q[λm(x)(s+w)+h(x)(s+w)p]. (2.4)

    The condition

    (IQ)[Δw]=(IQ)[λm(x)(s+w)+h(x)(s+w)p+Ω(s+w)β],

    implies that

    Ω(λmw+h(s+w)p+|Ω|(s+w)β)=0. (2.5)

    We consider F(λ,s,w)=0. Here F:R×R×WW,

    F(λ,s,w)=ΔwQ[λm(x)(s+w)+h(x)(s+w)p].

    Note that F(0,0,0)=0, Fw(0,0,0)w=Δw: WW is homeomorphism, by using implicit function theorem, there exists a unique solution w=w(λ,s) of the equation F(λ,s,w)=0 around (λ,s,w)=(0,0,0), being analytic at (0,0) and having the condition w(0,0)=0.

    Since W is complete, so w satisfies Ωkwsklλk(0,0)=0, namely kwsklλk(0,0)W. We substitute w=w(λ,s) for (2.5), then

    Φ(λ,s):=Ω(λmw(λ,s)+h(s+w(λ,s))p+|Ω|(s+w(λ,s))β)=0,

    where (λ,s) is in a neighborhood of (0,0) and Φ(λ,s) is analytic at (0,0).

    Let wm be the solution of the problem

    {Δw=m,xΩ,wn=0,xΩ,Ωw=0.

    We have the following conclusions.

    Theorem 2.3. Assume Ωm=0, p=2 or 3. If β=p, let |Ω|+Ωh<0; if β>p, let Ωh<0. Then there exists a continuous curve (λ(s),u(s)) of s(0,ε) such that (λ(0),u(0))=(0,0), (λ(s),u(s)) are positive solutions of the problem (1.1), and for any positive solution (λ,u) of the problem (1.1) in a neighborhood of (0,0), we have (λ,u)=(λ(s),u(s)). Moreover, if β=p, we have

    lims0+u(s)|λ(s)|2p1=(Ω|wm|2(|Ω|+Ωh))1p1>0. (2.6)

    If β>p, we have

    lims0+u(s)|λ(s)|2p1=(Ω|wm|2Ωh)1p1>0. (2.7)

    Proof. We shall solve that Φ(λ,s)=0 by considering the Taylor expansion of Φ at (λ,s)=(0,0). Since w(0,0)=0, we have Φ(0,0)=0. From F(λ,s,w)=0, we can calculate the partial derivative of w=w(λ,s) with respect to λ and s at (0,0) respectively.

    Calculating derivative of F(λ,s,w)=0 with respect to λ, when λ=0 and s=0, we have

    {Δwλ(0,0)=0,xΩ,wλn(0,0)=0,xΩ.

    So wλ(0,0) is a constant. By virtue of kwsklλk(0,0)W, we get wλ(0,0)=0. Similarly, we have

    kwλk(0,0)=0,k1.

    Calculating derivative of F(λ,s,w)=0 with respect to s, when λ=0 and s=0, we get

    kwsk(0,0)={0,1kp1,wp,k=p,

    where wp is the solution of the equation

    {Δw=p!Q[h(x)],xΩ,wn=0,xΩ,Ωw=0.

    Moreover, we have

    2wsλ(0,0)=wm.

    Next, we calculate partial derivative of Φ(λ,s) with respect to λ and s at (0,0) respectively. By direct calculations, we have

    kΦλk(0,0)=0,k1,

    (i) β=p,

    kΦsk(0,0)={0,1kp1,p!(|Ω|+Ωh),k=p,

    (ii) β>p,

    kΦsk(0,0)={0,1kp1,p!Ωh,k=p,

    (iii) β<p,

    kΦsk(0,0)={0,1kβ1,p!|Ω|,k=β.

    Moreover, we have

    2Φsλ(0,0)=Ωm=0,
    3Φs2λ(0,0)=0,p=3,β3,
    3Φsλ2(0,0)=2Ωmwm=2Ω|wm|2>0.

    Therefore, the Taylor expansion of Φ at (0,0) is Φ(λ,s)=sψ(λ,s).

    (1) p=2.

    (i) If β=2, then

    ψ(λ,s)=s(|Ω|+Ωh)+λ2Ω|wm|2+higherorderterms.

    (ii) If β>2, then

    ψ(λ,s)=sΩh+λ2Ω|wm|2+higherorderterms.

    (2) p=3.

    (i) If β=2, then

    ψ(λ,s)=s|Ω|+λ2Ω|wm|2+higherorderterms.

    (ii) If β=3, then

    ψ(λ,s)=λ2Ω|wm|2+s2(|Ω|+Ωh)+higherorderterms.

    (iii) If β>3, then

    ψ(λ,s)=λ2Ω|wm|2+s2Ωh+higherorderterms.

    For (1) (i), we note that ψs(0,0)=|Ω|+Ωh<0, and that by using implicit function theorem, there exists a unique solution s=s(λ) of the equation ψ(λ,s)=0 around (0,0), which satisfies

    s(0)=ψλ(0,0)ψs(0,0)=0,s(0)=2ψλ2(0,0)ψs(0,0)=2Ω|wm|2|Ω|+Ωh.

    So

    s(λ)=s(0)+s(0)λ+s2λ2+o(λ2)=λ2(Ω|wm|2|Ω|+Ωh+o(1)).

    Moreover, we have

    u=s+w(λ,s)=s+w(0,0)+(wλ(0,0)λ+ws(0,0)s)+12(wλλ(0,0)λ2+2wλs(0,0)λs+wss(0,0)s2)+...=s+wmλs+12wps2+...=s(1+o(1)).

    Therefore, combining the above two equations, we obtain (2.6).

    For (1) (ii), ψs(0,0)=Ωh<0, using a similar argument as that of (1) (i), we have

    s(λ)=λ2(Ω|wm|2Ωh+o(1)),u=s(1+o(1)).

    For (2) (i), since |Ω|>0,Ω|wm|2>0, so the problem (1.1) is no positive solution in the neighborhood of (0,0).

    For (2) (ii), since

    detD2ψ(0,0)=det(2Ω|wm|2002(|Ω|+Ωh))=4(|Ω|+Ωh)Ω|wm|2<0,

    by using the Morse lemma, we see that for any s>0, ψ(λ,s)=0 has a unique solution s=s(λ) in a neighborhood of (0,0) and we have

    s(λ)=((Ω|wm|2(|Ω|+Ωh)λ2)12+o(1)),u=s(1+o(1)).

    So we get the conclusion.

    For (2) (iii), since

    detD2ψ(0,0)=det(2Ω|wm|2002Ωh)=4ΩhΩ|wm|2<0,

    using a similar argument as that of (2) (ii), we get

    s(λ)=((Ω|wm|2Ωhλ2)12+o(1)),u=s(1+o(1)).

    We first prove that under the suitable conditions, the problem (1.1) has no positive solution for any sufficiently large |λ|. More precisely, we have the following results.

    Proposition 3.1. Assume that m(x) changes sign in Ωh+, (H1) holds and (λ,u) is a positive solution of the problem (1.1). Then λ(λ1,λ+1), where λ1<0 and λ+1>0 are the principal eigenvalue of the equation

    {Δu=λm(x)u,xΩh+,u=0,xΩh+.

    Proof. If (˜λ,˜u) is a positive solution of the problem (1.1), then

    {Δ˜u=˜λm(x)˜u+h(x)˜up+Ω˜uβ,xΩ,˜un=0,xΩ,

    thus, we have

    {Δ˜u˜λm(x)˜u=h(x)˜up+Ω˜uβ>0,xΩh+,˜u0,xΩh+.

    It follows that the principal eigenvalue μ1(˜λ) of the eigenvalue problem

    {Δu˜λm(x)u=μ(˜λ)u,xΩh+,u=0,xΩh+

    is positive. Then λ1<˜λ<λ+1.

    Assume m(x) changes sign in Ωh+, (H1) holds. For any λ(λ1,λ+1), eλ is the unique positive solution of the equation

    {Δuλm(x)u=1,xΩh+,u=0,xΩh+.

    We have the following lemma.

    Lemma 3.1. Assume m(x) changes sign in Ωh+, (H1) holds and (λ,u) is a positive solution of the problem (1.1). Then

    ueλΩuβ,xΩh+.

    Proof. If (λ,u) is a positive solution of the problem (1.1), then λ(λ1,λ+1), and we have

    {Δuλm(x)u=h(x)up+ΩuβΩuβ,xΩh+,u0,xΩh+,

    so

    {Δ(uΩuβ)λm(x)(uΩuβ)1,xΩh+,uΩuβ0,xΩh+.

    Thus, we have uΩuβeλ, namely, ueλΩuβ for xΩh+.

    We will use the method of Gidas-Spruck [7] to discuss the priori estimate of positive solutions.

    Lemma 3.2. Assume m(x) changes sign in Ωh+, (H1) holds and {(λk,uk)} is a sequence of positive solutions of the problem (1.1) with ukC(ˉΩh+) as k. Then there exists a constant C>0 such that ukC(ˉΩh+)Ct1pk for sufficiently large k, where tk=Ωuβk.

    Proof. If (λk,uk) is a positive solution of the problem (1.1), then λk(λ1,λ+1). Choose xkˉΩh+ such that

    uk(xk)=maxˉΩh+uk,k=1,2...

    and let Mk=uk(xk). Assume Mkt1pk as k+. Take a change of variables

    uk(x)=Mkvk(y),y=ρk(xxk),xΩh+, (3.1)

    where ρk=Mp12k, yΩk:=ρk(Ωh+{xk}). It is clear that 0<vk1, vk(0)=1. Substituting (3.1) into the problem (1.1), by direct calculations, we have

    Δvk(y)=λkˉmk(y)M1pkvk(y)+ˉhk(y)vpk(y)+(Mkt1pk)p,yΩk, (3.2)

    where ˉmk(y)=m(ρ1ky+xk), ˉhk(y)=h(ρ1ky+xk). Since ˉΩh+ is compact, then there exists a subsequence of {xk}, still denoted by {xk}, such that xkx0ˉΩh+. Now, we distinguish two cases.

    Case 1. x0Ωh+. It is seen in this case that ΩkRN as k. Hence, for any compact subset K1, we have K1Ωk for sufficiently large k. Since 0<vk1, there exist a positive constant C2, such that

    |λkˉmk(y)M1pkvk(y)+ˉhk(y)vpk(y)+(Mkt1pk)p|C2.

    By using the regularity theory of the elliptic equation, we know that, there exists a subsequence of {vk}, still denoted by itself, such that

    vkvinC1(K1),k+,

    where vC1(K1). Since K1⊂⊂Ωk is arbitrarily given, by a diagonal process, we can choose a subsequence, still denoted by {vk}, such that

    vkvinC1loc(RN),k+.

    Thus, we have

    Δv(y)=h(x0)vp(y),yRN. (3.3)

    Note that v(0)=1, by (3.3) and a linear change of coordinates, we find that there exists a nontrivial non-negative function wC2(RN) satisfying Δw=wp, which contradicts [7].

    Case 2. x0Ωh+. By an additional change of coordinates, we can assume that a neighborhood of x0 in Ωh+ is a hyperplane xN=0 and ˉΩh+H={xRN,xN>0}. Hence, given R>0, there exists kR such that for kkR, vk is well defined on

    HR,k:=B(0,R){yN>Mp12kxNk}.

    Now, we have the following three cases.

    (i) {Mp12kxNk} is not bounded from upper. Assume without loss of generality that Mp12kxNk. Then, we have

    HR,kB(0,R),k.

    We may argue exactly as in Case 1.

    (ii) {Mp12kxNk} is not bounded from below. Assume without loss of generality that Mp12kxNk0. Then, we have

    HR,kB(0,R)H,k.

    Arguing as in Case 1, there exists vC2(ˉH) such that v0, v(0)=1, and v satisfies

    Δv=h(x0)vp,inRN+.

    This contradicts Corollary 2.1 of [7].

    (iii) {Mp12kxNk} is bounded from below. Assume without loss of generality that Mp12kxNks,s>0. Then, we have

    HR,kB(0,R){yRN:yN>s}=BRHs.

    We can proceed as in (ii) and there exists vC2loc(Hs) such that v0, v(0)=1, and v satisfies

    Δv=h(x0)vp,inHs.

    Taking a change of variable through yN=s, we have that vC2(ˉH), v0, v(0)=1, and v satisfies

    Δv=h(x0)vp,inRN+,

    a contradiction.

    Lemma 3.3. Assume m(x) changes sign in Ωh+ and (H1) holds, then there exists a constant C>0 such that uC(ˉΩh+)C for any positive solution (λ,u) of the problem (1.1).

    Proof. If (λk,uk) is a positive solution of the problem (1.1), then λk(λ1,λ+1). Assume

    λkλ0(λ1,λ+1),ukC(ˉΩh+),k.

    By virtue of Lemma 3.2, we have

    ukukC(ˉΩh+)Ct1pk,xΩh+,

    where tk=Ωuβk. Moreover, we have tk. But by using Lemma 3.1, we have ukeλktk, a contradiction. Therefore, there exists a constant C>0 such that uC(ˉΩh+)C.

    Theorem 3.1. Assume m(x) changes sign in Ωh+, (H1) holds and β>max{p,N(p1)/N(p1)22}. Then there exists a constant C>0 such that uC(ˉΩ)C for any positive solution (λ,u) of the problem (1.1).

    Proof. Let f(u)=λm(x)u+h(x)up+Ωuβ. By Lemma 3.3, there exists a constant C1>0, such that uuC(ˉΩh+)C1,xΩh+. It follows that Ωuβ is bounded by Lemma 3.1, so f(u) is bounded in Lβ/βpp(Ω). Thus, u is bounded in W2,β/βpp(Ω). By using boot-strapping method [1], it follows that there exists a constant C>0, such that uC(ˉΩ)C.

    Next, we give another way of proving boundedness. We first show two results about eigenvalue problem. We consider the eigenvalue problem

    {Δu+a(x)u=σu,xΩ,BΩu=0,xΩ, (3.4)

    where

    BΩu={v,xΓ1(Ω),un,xΓ2(Ω),

    Ω=Γ1(Ω)Γ2(Ω), Γ1(Ω) is nonempty. σ1(Δ+a(x),BΩ,Ω) is the principal eigenvalue of problem (3.4). Generally, we denote BΩ by B. We have the following results.

    (1) If a(x)˜a(x), then σ1(Δ+a(x),B,Ω)σ1(Δ+˜a(x),B,Ω).

    Proof. Let ˜a(x)=a(x)+b(x), then b(x)0. Let F(u)=Ω|u|2+Ωau2, ˜F(u)=Ω|u|2+Ω˜au2, then

    σ1(Δ+a(x),B,Ω)=inf{F(u):uH1(Ω),uL2(Ω)=1,Bu=0}inf{F(u)+Ωbu2:uH1(Ω),uL2(Ω)=1,Bu=0}=inf{˜F(u):uH1(Ω),uL2(Ω)=1,Bu=0}σ1(Δ+˜a(x),B,Ω).

    (2) If ΩΩ and int(Ω)Γ2(Ω)=, then

    σ1(Δ+a(x),B,Ω)σ1(Δ+a(x),B,Ω).

    Proof. Let FΩ(u)=Ω|u|2+Ωau2, ˜u denote that the function uH1(Ω) extends to Ω and int(Ω)Γ2(Ω)=. Then ˜u satisfies ˜u|ΩΩ=0, BΩ˜u|ΩΩ=0. Therefore,

    σ1(Δ+a(x),B,Ω)=inf{FΩ(u):uH1(Ω),uL2(Ω)=1,BΩu=0}=inf{FΩ(˜u):˜uH1(Ω),˜uL2(Ω)=1,BΩ˜u|ΩΩ=0,˜u|ΩΩ=0}inf{FΩ(u):uH1(Ω),uL2(Ω)=1,BΩu|ΩΩ=0,u|ΩΩ=0}=inf{FΩ(u):uH1(Ω),uL2(Ω)=1,BΩu|ΩΩ=0,u|Γ1(ΩΩ)=0,u|Γ2(ΩΩ)=0}inf{FΩ(u):uH1(Ω),uL2(Ω)=1,BΩu|ΩΩ=0,u|Γ1(ΩΩ)=0,un|Γ2(ΩΩ)=0}=inf{FΩ(u):uH1(Ω),uL2(Ω)=1,BΩu=0}=σ1(Δ+a(x),B,Ω).

    Let S is the positive solution set of the problem (1.1) and ΛS:={λR;(λ,u)S} is bounded. we have

    Theorem 3.2. If sup(λ,u)SsupΩh+u<, then sup(λ,u)SsupΩu<.

    Proof. If (λ,u) is a positive solution of the problem (1.1), then

    {Δuλm(x)u=h(x)up+Ωuβ>0,xΩˉΩh,u0,xΓ1(ΩˉΩh),un=0,xΓ2(ΩˉΩh),

    where Γ1(ΩˉΩh)Ω, Γ2(ΩˉΩh)Ω, Γ1(Ω) is nonempty. Then (λ,u) is a strict upper solution of the equation

    {Δvλm(x)v=0,xΩˉΩh,v=0,xΓ1(ΩˉΩh),vn=0,xΓ2(ΩˉΩh),

    so

    σ1(Δλm(x),B,ΩˉΩh)>0.

    By Ωh0ΩˉΩh and result (2), we have

    σ1(Δλm(x),B,Ωh0)σ1(Δλm(x),B,ΩˉΩh).

    Let

    Ωδ=Ωh0{xΩh:d(x,Ωh)<δ}{Ωh(ΩΩh+)},

    then ΩδΩh0 when δ0. So there exists a sufficiently small δ>0, such that σ1(Δλm(x),B,Ωδ)>0, so that σ1(Δλm(x),B,Ωδ)>0 satisfies the strong maximum principle.

    Let M=sup(λ,u)SsupΩh+u and ψ be the unique solution of

    {Δvλm(x)v=1,xΩδ,v=M,xΓ1(Ωδ),vn=0,xΓ2(Ωδ).

    Since σ1(Δλm(x),B,Ωδ)>0, we have ψ>0,xΩδ by the strong maximum principle. Denote by w an extension of ψ|Ωδ/δ22 with minˉΩw>0, wn|Γ2(Ω)=0. Then ˉu=kw is for sufficiently large k>0 a positive strict upper solution of

    {Δv=λm(x)v+h(x)vp+k2,xΩˉΩh+,v=M,xΓ1(ΩˉΩh+),vn=0,xΓ2(ΩˉΩh+),

    where k2>Ωuβ. Indeed, in Ωδ/δ22 we have

    Δkw=kΔw=kλm(x)w+kkλm(x)w+h(x)(kw)p+kλm(x)(kw)+h(x)(kw)p+k2.

    In Σδ={xΩh:d(x,Ωh)δ2}, since w(x) and h(x) are positive and bounded away from zero, then h(x)kp1wp as k, namely

    Δwλm(x)w+h(x)kp1wp+k2/k2kk.

    On Γ1(ΩˉΩh+), since minˉΩw>0, we know kwM; on Γ2(ΩˉΩh+), we have kwn=0. Thus ˉu=kw is a positive strict upper solution. Moreover, by the relationship between the strict upper solution and the principal eigenvalues, we have

    σ1(Δλm(x)ˉup1h(x),B,ΩˉΩh+)>0.

    If (λ,u)S then it follows that v=ˉuu satisfies

    {Δvλm(x)v(ˉup1+ˉup2u+...+up1)h(x)v>0,xΩˉΩh+,v0,xΓ1(ΩˉΩh+),vn=0,xΓ2(ΩˉΩh+).

    Indeed, in ΩˉΩh+, since

    Δˉuλm(x)ˉu+h(x)ˉup+k2,
    Δu=λm(x)u+h(x)up+Ωuβ,

    so

    Δv>λm(x)v+h(x)(ˉupup)+(k2Ωuβ)>λm(x)v+(ˉup1+ˉup2u+...+up1)h(x)v.

    Thus by result (1) we have

    σ1(Δλm(x)(ˉup1+ˉup2u+...+up1)h(x),B,ΩˉΩh+)σ1(Δλm(x)ˉup1h(x),B,ΩˉΩh+)>0.

    By the relationship between the principal eigenvalues and the strong maximum principle, we have uˉu. The proof is completed.

    proof of Theorem 1.1. If Ωm<0, by virtue of Theorem 2.1, we obtain the bifurcation curve of positive solutions of the problem (1.1), and the direction of bifurcation is given by Theorem 2.2. When m(x) changes sign in Ωh+ and (H1) holds, we have (i) by Proposition 3.1. Finally, we have (ii) by the priori estimate of positive solutions of Theorem 3.2.

    Next, we prove (iii). We assume by contradiction that (λk,uk) are positive solutions of the problem (1.1), and (λk,uk)(γ,0) in R×C(ˉΩ), where γλ+ and 0. Let vk=ukukC(ˉΩ), by the problem (1.1), we have

    {Δvk=λkm(x)vk+h(x)vpkukp1C(ˉΩ)+ukβ1C(ˉΩ)Ωvβk,xΩ,vkn=0,xΩ.

    By virtue of the regularity theory of the elliptic equation, we know that there is a subsequence, still denoted by {vk}, such that vkv0 in C2(ˉΩ), v0 is a solution of the equation

    {Δv0=γm(x)v0,xΩ,v0n=0,xΩ.

    So γ=λ+ or 0, a contradiction. Moreover, by using Rabinowitz global bifurcation theory, we see that C+ bifurcates from (λ+,0) and backs to (0,0).

    Similarly, we obtain the global bifurcation results of the problem (1.1) for Ωm>0.

    proof of Theorem 1.3. Since m(x) changes sign in Ωh+ and mε(x)=m(x)ε for ε>0, so mε changes sign and Ωmε<0 for sufficiently small ε. Thus, we know that

    {Δu=λmε(x)u,xΩ,un=0,xΩ

    has the principal eigenvalue 0 and λ+(mε)>0. Substituting mε(x) for m(x), we can see that u is bounded in C(ˉΩ) with mε(x) through the proof of a priori estimate of positive solutions. By Theorem 1.1, we see that the problem (1.1) with mε has a connected branch C+ε of positive solutions set in R×C(ˉΩ) such that its closure ¯C+ε contains (0,0) and (λ+(mε),0). Suppose that φε>0 is the principal eigenfunction corresponding to λ+(mε) normalized so that φεW1,2(Ω)=1. According to [4], we have limε0λ+(mε)=0 and φεC in W1,2(Ω), where C is a positive constant.

    If β>p and Ωh<0, then C+ε is subcritical at (0,0) by Theorem 2.2. Since λ+(mε)0 and φεC, so for sufficiently small ε, we have Ωhφp+11 and Ωh have same sign for β>p. So C+ε is supercritical at (λ+(mε),0) for β>p and Ωh<0. Therefore, ¯C+ε is likely to approach a closed loop as ε0, which bifurcates from the origin and backs to the origin, as Figure 3.

    We now investigate ¯C+ε as ε0. Although it seems likely in Figure 3 that ¯C+ε approaches a closed loop joining the origin to itself as ε0, this seems difficult to establish. We can, however, prove that ¯C+ε does not shrink to a point. For sets En, nN, we define

    limninfEn={x: there exists N0N such that any neighborhood of x intersects En for all nN0},

    limnsupEn={x: any neighborhood of x intersects En for infinitely many n}.

    According to [8], if n1En is precompact in M and lim_nEn, then ¯limnEn is non-empty, closed and connected. Here, {En} is a sequence of connected sets in a complete metric space M.

    Obviously, (0,0)¯C+ε, so (0,0)lim_ε0C+ε. It follows from the results of the priori bounds that ε>0C+ε is precompact in C2(ˉΩ). Then ¯limε0C+ε is non-empty, closed and connected. We note that (0,0)¯limε0C+ε, and also, from the definition, that ¯limε0C+ε consists of nonnegative solutions of the problem (1.1).

    Since ¯C+ε joining (0,0) and (λ+(mε),0) is subcritical at (0,0) and supercritical at (λ+(mε),0), then ¯C+ε must join (0,uε), uε is a positive solution of the equation

    {Δu=h(x)up+Ωuβ,xΩ,un=0,xΩ. (4.1)

    By a priori estimate of positive solutions, we see that uε is bounded. By virtue of the regularity theory of the elliptic equation, it follows that {uε} must have a convergent subsequence in C2(ˉΩ) converging to u, where u is a solution of the Eq (4.1). Moreover, we have (0,u)¯limε0C+ε.

    Next, we prove u0. Otherwise, we have uε0. Let vε=uεuε, by the Eq (4.1), we see that

    Ω|uε|2=Ωh(x)up+1ε+ΩuεΩuβε,

    then

    Ω|vε|2=uεp1Ωh(x)vp+1ε+uεβ1ΩvεΩvβε. (4.2)

    Thus, we have Ω|vε|20 as ε0. Since vε is bounded, so we may assume that vεv0 in W1,2(Ω), vεv0 in Lp+1(Ω) and Lβ(Ω), hence, we claim that vεv0 in W1,2(Ω). Otherwise, we have

    Ω|v0|2<lim_ε0Ω|vε|20,

    a contradiction, so v0 is a positive constant c, then vεc in Lp+1(Ω) and Lβ(Ω). Thus, when ε is sufficiently small, we have ΩvεΩvβε<0 for β>p, but this is impossible because of the equality in (4.2).

    Therefore, under the conditions of the Theorem 1.3, there exists a connected components C+ of positive solutions set such that its closure ¯C+ includes ¯limε0C+ε, which bifurcates from the origin and backs to the origin, namely, ¯C+ is a closed loop.

    The authors would like to thank the referees for their valuable comments and suggestions. This paper is partially supported by NSFC, PR China 11871250.

    The authors declare no conflict of interest.



    [1] F. Corrˆea, A. Suˊarez, Combining local and nonlocal terms in a nonlinear elliptic problem, Math. Method. Appl. Sci., 35 (2012), 547–563. doi: 10.1002/mma.1592
    [2] M. Wang, Y. Wang, Properties of positive solutions for non-local reaction-diffusion problems, Math. Method. Appl. Sci., 19 (1996), 1141–1156. doi: 10.1002/(SICI)1099-1476(19960925)19:14<1141::AID-MMA811>3.0.CO;2-9
    [3] P. Quittner, P. Souplet, Superlinear parabolic problems, Blow-up, Global Existence and Steady States, Basel: Birkhäuser Verlag, 2007.
    [4] K. J. Brown, Local and global bifurcation results for a semilinear boundary value problem, J. Differ. Equations, 239 (2007), 296–310. doi: 10.1016/j.jde.2007.05.013
    [5] K. J. Brown, S. S. Lin, On the existence of positive eigenfunctions for an eigenvalue problem with indefinite weight function, J. Math. Anal. Appl., 75 (1980), 112–120. doi: 10.1016/0022-247X(80)90309-1
    [6] M. Wang, Nonlinear elliptic equations (in Chinaese), Beijing: Science Press, 2010.
    [7] B. Gidas, J. Spruck, A priori bounds for positive solutions of nonlinear elliptic equations, Commun. Part. Diff. Eq., 6 (1981), 883–901. doi: 10.1080/03605308108820196
    [8] G. T. Whyburn, Topological analysis, B. Am. Math. Soc., 62 (1956), 119–121.
    [9] K. Umezu, Bifurcation approach to a logistic elliptic equation with a homogeneous incoming flux boundary condition, J. Differ. Equations, 252 (2012), 1146–1168. doi: 10.1016/j.jde.2011.08.043
    [10] K. Umezu, Global structure of supercritical bifurcation with turning points for the logistic elliptic equation with nonlinear boundary conditions, Nonlinear Anal-Theor, 89 (2013), 250–266. doi: 10.1016/j.na.2013.05.011
    [11] J. Bebernes, A. Bressan, Thermal behavior for a confined reactive gas, J. Differ. Equations, 44 (1982), 118–133. doi: 10.1016/0022-0396(82)90028-6
    [12] H. Bei, H. M. Yin, Semilinear parabolic equations with prescribed energy, Rendiconti Del Circolo Matematico Di Palermo, 44 (1995), 479–505. doi: 10.1007/BF02844682
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2461) PDF downloads(118) Cited by(0)

Figures and Tables

Figures(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog