Home Heterogeneous Elliptic BVPs with a Bifurcation-Continuation Parameter in the Nonlinear Mixed Boundary Conditions
Article Open Access

Heterogeneous Elliptic BVPs with a Bifurcation-Continuation Parameter in the Nonlinear Mixed Boundary Conditions

  • Santiago Cano-Casanova ORCID logo EMAIL logo
Published/Copyright: July 18, 2019

Abstract

This article ascertains the global structure of the diagram of positive solutions of a very general class of elliptic boundary value problems with spatial heterogeneities and nonlinear mixed boundary conditions, considering as bifurcation-continuation parameter a certain parameter γ that appears in the boundary conditions. In particular, in this work are obtained, in terms of such a parameter γ, the exact decay rate to zero and blow-up rate to infinity of the continuum of positive solutions of the problem, at the bifurcations from the trivial branch and from infinity. The new findings of this work complement, in some sense, those previously obtained for Robin linear boundary conditions by J. García-Melián, J. D. Rossi and J. C. Sabina de Lis in 2007. The main technical tools used to develop the mathematical analysis carried out in this paper are local and global bifurcation, continuation, comparison and monotonicity techniques and blow-up arguments.

1 Introduction, Previous Results and Main Theorems

In this paper, we will consider the boundary value problem with nonlinear mixed boundary conditions and spatial heterogeneities given by

(1.1) { - Δ u = λ u - a ( x ) u p in Ω , p > 1 , u = 0 on Γ 0 , u + V ( x ) u = γ b ( x ) u q on Γ 1 , q > 1 ,

where

  1. Ω is a bounded domain of N, N1, of class 𝒞2, with boundary Ω=Γ0Γ1, where Γ0 and Γ1 are disjoint open and closed subsets of Ω,

  2. - Δ stands for the minus Laplacian operator in N and λ,

  3. a 𝒞 ( Ω ¯ ) with a0 in Ω measures the spatial heterogeneities in Ω,

  4. the potentials V,b𝒞(Γ1), where b0 on Γ1 and V possesses arbitrary sign in each point xΓ1, represent the spatial heterogeneities on the boundary,

  5. γ will be the bifurcation-continuation parameter, and u(x) stands for the outer normal derivative of u at xΓ1.

This work is devoted to ascertain the global structure of the set of positive solutions of (1.1) considering γ as the bifurcation-continuation parameter, for some fixed λ in a suitable interval. The previous works [5, 6, 7, 9, 11] play a crucial role in obtaining the results of this paper. We will distinguish the following cases:

  1. Case (1): a=0 in Ω and

    (1.2) b ( x ) b ¯ > 0 for all x Γ 1 .

    In this case, we set Ω0:=Ω.

  2. Case (2): a0 in Ω, b0 on Γ1,

    (1.3) Ω 0 : = int { x Ω : a ( x ) = 0 } , Ω 0 𝒞 2 , Ω ¯ 0 Ω and a is bounded away from zero in any compact subset of ( Ω Ω ¯ 0 ) Γ 1 .

    By a0 in Ω, it is meant that a(x)0 for all xΩ and a0. Similarly, b0 on Γ1 means that b(x)0 for all xΓ1 and b0.

The case when b(x) changes sign on Γ1 and/or a(x) changes sign in Ω remains outside the general scope of this work. It will be analyzed elsewhere.

The global structure of the diagram of positive solutions of (1.1) considering λ as bifurcation parameter, for some fixed γ, was already analyzed in [11, 5] (for γ=0) and in [6] (for γ0), in the particular case when a0 in Ω vanishes in Ω0 with Ω¯0Ω, and in [7], in the special case when a=0 in Ω, that is, when Ω0=Ω. In this work, this perspective will be completed by analyzing the structure of the diagram of positive solutions of (1.1) regarding to γ as the main bifurcation-continuation parameter for a fixed value of λ in a suitable interval.

In both cases, (1) and (2), we analyze the existence, uniqueness, asymptotic behavior and stability of the positive solutions of (1.1) and ascertain the exact decay rate to zero as γ- of the positive solutions of (1.1), in terms of the parameter γ in the nonlinear mixed boundary conditions. In both cases, it will be shown that the decay rate in L(Ω) as γ- of the positive solutions of (1.1) is, up multiplicative constants, of order (-γ)-1q-1. Also, in Case (1), we will ascertain the exact blow-up rate to infinity, as γ0, of the continuum of positive solutions of (1.1), in terms of the parameter γ. Up to a multiplicative constant, it will be established that such blow-up rate also is of order (-γ)-1q-1 in L(Ω).

By a positive solution of (1.1), it is meant any couple (λ,u)×Wr2(Ω) for some r>N, with u>0 in Ω satisfying (1.1). It should be noted that Wr2(Ω)𝒞2-Nr(Ω¯) and that any function uWr2(Ω), r>N, is a.e. twice differentiable (cf. [18, Theorem VIII.1]). We will say that a positive solution (λ,u) of (1.1) is strongly positive in Ω, and we will denote it by u0, if u(x)>0 for all xΩΓ1 and u(x)<0 for all xΓ0.

Hereafter, for each γ, we will denote

Λ ( γ ) : = { λ : (1.1) possesses a positive solution } ,

by 𝒞λ+(λ=λ~,γI) the continuum of positive solutions of (1.1) which, considering λ as bifurcation parameter, emanates from the trivial branch (λ,u)=(λ,0) at λ=λ~ for some fixed γ in the interval I, and by 𝒞γ+(γ=γ~,λJ) the continuum of positive solutions of (1.1) which, considering γ as bifurcation parameter, emanates from the trivial branch (γ,u)=(γ,0) at γ=γ~ for some fixed λ in the interval J.

Also, for each V𝒞(Γ1), 𝔅(V(x)) stands for the boundary operator

𝔅 ( V ( x ) ) u : = { u on Γ 0 , u + V ( x ) u on Γ 1 ,

and 𝔇 stands for the Dirichlet operator on Ω. It is known (cf. [3]) that

𝔅 ( V ( x ) ) ( W r 2 ( Ω ) , W r 2 - 1 r ( Γ 0 ) × W r 1 - 1 r ( Γ 1 ) ) .

Owing to the results in [3] and [16, Sections 7.2, 7.4], we have that, for any K𝒞(Ω¯) and V𝒞(Γ1), the boundary eigenvalue problem

(1.4) { ( - Δ + K ( x ) ) φ = σ φ in Ω , 𝔅 ( V ( x ) ) φ = 0 ¯ on Ω

possesses a unique eigenvalue that possesses a positive eigenfunction, unique up to a multiplicative constant, named principal eigenvalue of (1.4). In the sequel, we will denote it by σ1Ω[-Δ+K(x),𝔅(V(x))]. Also, the principal eigenvalue of (1.4) is simple and dominant in the sense that any other eigenvalue λ of (1.4) satisfies

Re ( λ ) > σ 1 Ω [ - Δ + K ( x ) , 𝔅 ( V ( x ) ) ] ,

where Re(λ) stands for the real part of λ. In [16, Theorem 7.8], the dominance of the principal eigenvalue σ1Ω[-Δ+K(x),𝔅(V(x))] is proved, and using the generalized Krein–Rutman theorem given in [16, Theorem 6.3], the dominance of it among the real numbers is definitely proved.

In addition, if φ* stands for the positive eigenfunction of (1.4) associated to σ1Ω[-Δ+K(x),𝔅(V(x))], unique up to a multiplicative constant, then

(1.5) φ * 0 in Ω ,

and

(1.6) φ * W 2 ( Ω ) : = r > N W r 2 ( Ω ) 𝒞 1 + α ( Ω ¯ ) for all α ( 0 , 1 ) .

Hereafter, we will denote σ1:=σ1Ω[-Δ,𝔅(V(x))], and by φ1>0 the principal eigenfunction associated to the principal eigenvalue σ1, normalized so that φ1L(Ω)=1. By (1.5) and (1.6), we get

φ 1 0 in Ω and φ 1 𝒞 1 + α ( Ω ¯ ) for all α ( 0 , 1 ) .

Also, we will denote

(1.7) σ 0 * : = σ 1 Ω [ - Δ , 𝔇 ] , σ 0 : = σ 1 Ω 0 [ - Δ , 𝔇 ] ,

that is, the principal eigenvalue of the -Δ operator in the domains Ω0 and Ω, respectively, under Dirichlet boundary conditions. Owing to [8, Propositions 3.1, 3.2], we know that, in the case when Ω¯0Ω, then

(1.8) σ 1 < σ 0 * < σ 0 ,

and in the case when Ω0=Ω, then σ1<σ0*=σ0. Therefore, in any of the previous cases, σ1<σ0.

Hereafter, we will denote

𝒞 Γ 0 1 ( Ω ¯ ) : = { u 𝒞 1 ( Ω ¯ ) : u | Γ 0 = 0 } , W r , Γ 0 2 ( Ω ) : = { u W r 2 ( Ω ) : u | Γ 0 = 0 } .

For each fixed λ in a suitable interval, the solutions of (1.1) are the roots of the operator 𝔉λ:×𝒰𝒰 defined by

(1.9) 𝔉 λ ( γ , u ) : = u - 𝔎 ~ 1 ( ( λ + ω ) u - a ( x ) u p ) - 𝔎 ~ 2 ( γ b u q ) ,

for 𝒰:=𝒞Γ01(Ω¯), where 𝔎~1,𝔎~2:𝒰𝒰 are the compact operators defined by

𝔎 ~ 1 : = i K 1 j : 𝒰 𝒰 , 𝔎 ~ 2 : = i K 2 i * t 1 : 𝒰 𝒰 ,

where, for each r>N, fLr(Ω) and gWr1-1r(Γ1), u1:=K1(f)Wr,Γ02(Ω) and u2:=K2(g)Wr,Γ02(Ω) stand for the unique solutions of the problems

{ ( - Δ + ω ) u 1 = f in Ω , 𝔅 ( V ( x ) ) u 1 = 0 on Ω , { ( - Δ + ω ) u 2 = 0 in Ω , u 2 = 0 on Γ 0 , u 2 + V ( x ) u 2 = g on Γ 1 ,

respectively, for some fixed

(1.10) ω > - σ 1 ,

i , j and i* being the inclusion operators

i : W r , Γ 0 2 ( Ω ) 𝒰 , i * : 𝒞 1 ( Γ 1 ) W r 1 - 1 r ( Γ 1 ) , j : 𝒰 L r ( Ω )

and t1:𝒰𝒞1(Γ1) the trace operator on Γ1.

Hereafter, given a solution u0 of (1.1) for the values (λ,γ)=(λ0,γ0) of the parameters, we will denote by ((λ0,u0),(γ0,u0)) the linearization of (1.1) around u0, where (λ0,u0) and (γ0,u0) stand for the linear operators

(1.11) ( λ 0 , u 0 ) : = - Δ - λ 0 + p a ( x ) u 0 p - 1 , ( γ 0 , u 0 ) : = 𝔅 ( V ( x ) - q γ 0 b ( x ) u 0 q - 1 ) .

Owing to (1.10), we get

(1.12) Ker ( D u 𝔉 λ 0 ( γ 0 , u 0 ) ) = Ker ( ( λ 0 , u 0 ) , ( γ 0 , u 0 ) ) .

Also, we will say that a non-negative solution u0 of (1.1) for the values (λ,γ)=(λ0,γ0) is

  1. linearly asymptotically stable if σ1Ω[(λ0,u0),(γ0,u0)]>0,

  2. linearly unstable if σ1Ω[(λ0,u0),(γ0,u0)]<0,

  3. neutrally stable if σ1Ω[(λ0,u0),(γ0,u0)]=0,

  4. linearly stable if σ1Ω[(λ0,u0),(γ0,u0)]0.

Owing to [6, Proposition 2.1], it is known that if uλ,γ is a positive solution of (1.1) for the values (λ,γ) of the parameters, then

(1.13) λ = σ 1 Ω [ - Δ + a ( x ) u λ , γ p - 1 , 𝔅 ( V ( x ) - γ b ( x ) u λ , γ q - 1 ) ] ,

u λ , γ 𝒞 1 + α ( Ω ¯ ) for all α(0,1) and uλ,γ is strongly positive in Ω.

1.1 Case (1): Previous Results and Main Theorem

In Case (1), (1.1) becomes

(1.14) { - Δ u = λ u in Ω , u = 0 on Γ 0 , u + V ( x ) u = γ b ( x ) u q on Γ 1 , q > 1 ,

Ω 0 : = Ω , σ0=σ0*=σ1Ω[-Δ,𝔇], and owing to the results in [3, 7], the following is known:

  1. If γ=0, then (1.14) becomes

    (1.15) { - Δ u = λ u in Ω , 𝔅 ( V ( x ) ) u = 0 on Ω ,

    which possesses a positive solution if and only if λ=σ1, and in this case, all the positive solutions of (1.15) are a positive multiple of φ1, where φ1 is the principal eigenfunction associated to the principal eigenvalue σ1 of (1.15), normalized so that φ1L(Ω)=1. That is, Λ(0)={σ1}. In this case, for γ=0, considering λ as bifurcation parameter, we get a genuine vertical bifurcation to positive solutions from the trivial branch (λ,u)=(λ,0) at λ=σ1, and the global bifurcation diagram of positive solutions of (1.14) is like shown in Figure 1, where +:=𝒞λ+(λ=σ1,γ=0).

  2. For any γ<0, independently of its size, Λ(γ)=(σ1,σ0), and hence (cf. [7, Theorem 3.1])

    (1.16) ( σ 1 , σ 0 ) = γ < 0 Λ ( γ ) .

    In this case, owing to the previous results in [7], we know that the global bifurcation diagram of positive solutions is like shown in Figure 2, where +:=𝒞λ+(λ=σ1,γ<0).

  3. If γ>0, then Λ(γ)(-,σ1) and Λ(γ) might depend on γ in the sense that Λ(γ1)Λ(γ2) if γ1>0, γ2>0 with γ1γ2, and it might occur that γ>0Λ(γ)= (cf. [7, Theorem 4.5]).

Then, in order to fix λ in a suitable interval for considering γ as the main bifurcation-continuation parameter, owing to the previous results, we get in Case (1):

  1. If λσ0, then λΛ(γ) for any γ,

  2. If λ=σ1, then (1.14) possesses a positive solution if and only if γ=0, and in this case, all the positive solutions are positive multiples of φ1. Then, considering γ as bifurcation parameter at λ=σ1, we get a vertical bifurcation to positive solutions from the trivial branch (γ,u)=(γ,0) at γ=0, and the global bifurcation diagram of positive solutions in the γ parameter of (1.14) looks like shown in Figure 3, where +:=𝒞γ+(γ=0,λ=σ1)

  3. If λ(σ1,σ0), then the structure of the set of positive solutions of (1.14), as well as their point-wise behavior regarding to γ as the main bifurcation parameter, are given by Theorem 1.1, which is the main result of this paper for Case (1).

Hereafter, we will set

γ ~ : = - γ and b ¯ : = b L ( Γ 1 ) > 0 .

Figure 1 
            Vertical bifurcation to positive solutions in the λ-parameter at λ=σ1{\lambda=\sigma_{1}} (γ=0{\gamma=0}).
Figure 1

Vertical bifurcation to positive solutions in the λ-parameter at λ=σ1 (γ=0).

Figure 2 
            Bifurcation to positive solutions in the λ-parameter (γ<0{\gamma<0}).
Figure 2

Bifurcation to positive solutions in the λ-parameter (γ<0).

Figure 3 
            Vertical bifurcation to positive solutions in the γ-parameter at γ=0{\gamma=0} (λ=σ1{\lambda=\sigma_{1}}).
Figure 3

Vertical bifurcation to positive solutions in the γ-parameter at γ=0 (λ=σ1).

Theorem 1.1.

Assume that

(1.17) λ ( σ 1 , σ 0 )

and (1.2) holds.

  1. Problem ( 1.14 ) possesses a positive solution if and only if γ < 0 , and it is unique if it exists. In the sequel, we will denote it by u γ .

  2. Any positive solution of ( 1.14 ) is linearly asymptotically stable.

  3. The map ( - , 0 ) 𝒞 Γ 0 1 ( Ω ¯ ) , γuγ is differentiable, and

    (1.18) u ˙ γ : = d u γ d γ 0 𝑖𝑛 Ω .

    In particular, we get u ˙ γ ( x ) > 0 for all x Ω Γ 1 , and therefore the map ( - , 0 ) 𝒞 Γ 0 ( Ω ¯ ) , γuγ is strictly increasing.

  4. All the positive solutions of ( 1.14 ) belong to the branch

    (1.19) Γ + ( λ ) : = { ( γ , u γ ) : γ < 0 } ,

    which is increasing with respect to γ in L(Ω). Moreover, there exist C(λ)>0 and D(λ)>0 such that, for all γ<0, the following holds:

    (1.20) C ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1 u γ L ( Ω ) D ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1 .

    In particular,

    (1.21) lim γ - u γ L ( Ω ) = 0 𝑎𝑛𝑑 lim γ 0 u γ L ( Ω ) = ,

    that is, the branch Γ + ( λ ) bifurcates to positive solutions from the trivial solution as γ - , while it bifurcates to positive solutions from infinity as γ 0 .

  5. The following holds:

    (1.22) lim γ 0 u γ ( x ) = ,

    uniformly in compact subsets K Ω Γ 1 .

  6. For each γ < 0 , the unique positive solution u γ of ( 1.14 ) is globally asymptotically stable as steady-state solution of the parabolic problem associated to ( 1.14 ).

Owing to Theorem 1.1, we get that, for each fixed λ(σ1,σ0), the global bifurcation diagram of positive solutions of (1.14) in the γ-parameter is like shown in Figure 4, where Γ+(λ) stands for the branch of positive solutions defined by (1.19). As illustrated by Figure 4, Γ+(λ) bifurcates to positive solutions from the trivial branch (γ,u)=(γ,0) as γ-, and from infinity as γ0, with exactly the same rate γ~-1q-1, up to some multiplicative constant.

Figure 4 
            Global bifurcation diagram of positive solutions in the γ-parameter for any fixed λ∈(σ1,σ0){\lambda\in(\sigma_{1},\sigma_{0})}.
Figure 4

Global bifurcation diagram of positive solutions in the γ-parameter for any fixed λ(σ1,σ0).

1.2 Case (2): Previous Results and Main Theorem

In Case (2), owing to the previous results of [11, 5, 9, 6], the following is known:

  1. If γ=0, then (1.1) becomes the boundary value problem with linear mixed boundary conditions

    { - Δ u = λ u - a ( x ) u p in Ω , 𝔅 ( V ( x ) ) u = 0 on Ω ,

    which has been widely analyzed in [11, 5, 9], and recently in [10] in a much more general framework with no special restriction on the structure of a-1(0)=Ω0. In this case, it is well known that the problem possesses a positive solution if and only if λ(σ1,σ0), that is, Λ(0)=(σ1,σ0) (cf. [11, Theorem 3.5], [5, Theorem 4.2]). The global bifurcation diagram of positive solutions in this case, considering λ as bifurcation parameter, is like shown in Figure 5, where γ=0+:=𝒞λ+(λ=σ1,γ=0).

  2. For any γ<0, independently of its size, Λ(γ)=(σ1,σ0), that is, γ<0Λ(γ)=(σ1,σ0), (cf. [6, Theorem 3.1]), and its global bifurcation diagram of positive solutions, considering λ as bifurcation parameter, is like shown in Figure 5, where γ<0+:=𝒞λ+(λ=σ1,γ<0).

  3. If γ>0 and p>2q-1, then (cf. [6, Theorem 4.7])

    (1.23) ( σ 1 , σ 0 ) Λ ( γ ) ,

    and its global bifurcation diagram of positive solutions, considering λ as bifurcation parameter, should be like shown in Figure 6, where +:=𝒞λ+(λ=σ1,γ>0).

Figure 5 
            Global bifurcation diagram of positive solutions in the λ-parameter (γ≤0{\gamma\leq 0}).
Figure 5

Global bifurcation diagram of positive solutions in the λ-parameter (γ0).

Figure 6 
            Global bifurcation diagram of positive solutions in the λ-parameter (γ>0{\gamma>0}).
Figure 6

Global bifurcation diagram of positive solutions in the λ-parameter (γ>0).

Therefore, owing to the previous results, in Case (2), we find that γ0Λ(γ)=(σ1,σ0), and if p>2q-1, then (1.23) holds, and hence γΛ(γ)=(σ1,σ0). Then, to get the structure of the set of positive solutions of (1.1), considering γ as bifurcation-continuation parameter, we will fix λ(σ1,σ0). Next follows the main theorem in Case (2).

Theorem 1.2.

Assume that (1.17) holds, and let us denote by u0 the unique positive solution of (1.1) for γ=0.

  1. For each γ 0 , ( 1.1 ) possesses a unique positive solution u γ , which is linearly asymptotically stable. Moreover, the map ( - , 0 ] 𝒞 Γ 0 1 ( Ω ¯ ) , γuγ is differentiable and

    u ˙ γ : = d u γ d γ 0 𝑖𝑛 Ω ,

    and in particular, the map

    (1.24) ( - , 0 ] 𝒞 Γ 0 ( Ω ¯ ) , γ u γ

    is strictly increasing.

  2. If ( 1.2 ) holds and λ ( σ 1 , σ 0 * ) (cf. ( 1.7 )), then there exists D ( λ ) > 0 such that

    (1.25) u γ L ( Ω ) D ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1 for all γ < 0 .

    In particular,

    (1.26) lim γ - u γ L ( Ω ) = 0 ,

    that is, the problem exhibits bifurcation to positive solutions from the trivial branch ( γ , u ) = ( γ , 0 ) when γ - .

  3. There exists ε 0 > 0 and a differentiable map

    u : ( - ε 0 , ε 0 ) 𝒞 Γ 0 1 ( Ω ¯ ) , γ u γ *

    such that u γ * = u γ for all γ ( - ε 0 , 0 ] , there exists a neighborhood 𝔘 of ( γ , u ) = ( 0 , u 0 ) in ( - ε 0 , ε 0 ) × 𝒞 Γ 0 1 ( Ω ¯ ) such that if ( γ , u ~ ) 𝔘 is a positive solution of ( 1.1 ), then u ~ = u γ * , and in addition, ( γ , u γ * ) is a positive linearly asymptotically stable solution of ( 1.1 ) for all γ ( - ε 0 , ε 0 ) . Moreover,

    u ˙ γ * : = d u γ * d γ 0 𝑖𝑛 Ω ,

    and in particular, the map ( - ε 0 , ε 0 ) 𝒞 Γ 0 ( Ω ¯ ) , γuγ* is strictly increasing.

  4. Any positive solution u ^ γ of ( 1.1 ) for γ > 0 satisfies

    (1.27) u ^ γ u 0

  5. If p > 2 q - 1 , then the following hold:

    1. For each γ > 0 , ( 1.1 ) possesses at least a positive solution.

    2. For each γ > 0 , ( 1.1 ) possesses a minimal positive solution u γ min satisfying

      (1.28) u γ min u 0

      and

      (1.29) u γ min = u γ * 𝑓𝑜𝑟 γ ( 0 , ε 0 ) ,

      where ε 0 and u γ * are defined by (iii).

    3. There exist uniform L ( Ω ) bounds for the positive solutions of ( 1.1 ) in compact intervals of values of γ.

Theorem 1.2 establishes in Case (2) that, for each fixed λ(σ1,σ0), the global bifurcation diagram in the γ-parameter of the positive solutions of (1.1) should be like shown by Figure 7, where the continuous line stands for the exact structure of the set of positive solutions for γ<ε0 (cf. Theorem 1.2 (iii)) and the dashed line stands for a possible configuration of the set of positive solutions of (1.1) for γ>ε0.

Figure 7 
            Global bifurcation diagram of positive solutions in the γ parameter (λ∈(σ1,σ2){\lambda\in(\sigma_{1},\sigma_{2})}).
Figure 7

Global bifurcation diagram of positive solutions in the γ parameter (λ(σ1,σ2)).

The distribution of the rest of this paper is the following. Section 2 contains some results about the local structure of the bifurcation diagram of positive solutions of (1.1) in a neighborhood of the positive linearly asymptotically stable or neutrally stable solutions of (1.1). Section 3 contains the proof of Theorem 1.1. To discuss in a particular example the results established by Theorem 1.1, Section 4 analyzes a one-dimensional example of boundary value problem of Case (1), which is explicitly solved for every γ<0 obtaining in addition its branch of positive solutions. The explicit branch of positive solutions obtained, blows up to infinity in L(Ω) as γ0, and decays to zero in L(Ω) as γ-, with the rate provided by Theorem 1.1, that is, with rate γ~-1q-1 in both cases. Finally, Section 5 contains the proof of Theorem 1.2.

2 Local Structure of the Bifurcation Diagram of Positive Solutions of (1.1)

This section analyzes the local structure of the bifurcation diagram of positive solutions of (1.1) in a neighborhood of the positive linearly asymptotically stable and positive neutrally stable solutions of (1.1), assuming that λ(σ1,σ0) is fixed. The results of this section are based on [14, Sections 2, 3], [15, Section 2] and [17, Chapter 2], though here the underlying technicalities are substantially more sophisticated by the intrinsic nature of our problem.

As it will be shown in the proofs of Theorems 1.1 and 1.2, when λ(σ1,σ0) and a=0 in Ω (Case (1)), then the positive solutions of (1.14) exist only for γ<0 and all of them are linearly asymptotically stable, while if a0 in Ω (Case (2)), all the positive solutions for γ0 are linearly asymptotically stable and only may exist neutrally stable solutions if γ>0.

The next result provides us the local structure of the bifurcation diagram of positive solutions of (1.1) around a positive linearly asymptotically stable solution.

Proposition 2.1.

Let (γ0,u0) be a positive linearly asymptotically stable solution of (1.1). Then there exists ε>0 and a differentiable function

u : ( γ 0 - ε , γ 0 + ε ) 𝒞 Γ 0 1 ( Ω ¯ ) , γ u γ

such that uγ0=u0, there exists a neighborhood U of (γ0,u0) in (γ0-ε,γ0+ε)×CΓ01(Ω¯) such that if (γ,u~)U is a positive solution of (1.1), then u~=uγ, and in addition, (γ,uγ) is a positive linearly asymptotically stable solution of (1.1) for all γ(γ0-ε,γ0+ε). Moreover,

(2.1) u ˙ γ : = d u γ d γ 0 𝑖𝑛 Ω ,

and in particular, the map

(2.2) ( γ 0 - ε , γ 0 + ε ) 𝒞 Γ 0 ( Ω ¯ ) , γ u γ

is strictly increasing.

Proof.

Since the solutions of (1.1) are the roots of operator (1.9), we get 𝔉λ(γ0,u0)=0. Also, since (γ0,u0) is a positive linearly asymptotically stable solution of (1.1), this yields

(2.3) σ 1 Ω [ ( λ , u 0 ) , ( γ 0 , u 0 ) ] > 0 ,

which, owing to (1.12), implies that Du𝔉λ(γ0,u0) is a topological isomorphism. Then the existence, local uniqueness and regularity of the branch of positive solutions

(2.4) { ( γ , u γ ) : γ ( γ 0 - ε , γ 0 + ε ) }

follow from the implicit function theorem applied to the operator 𝔉λ(γ,u) at (γ0,u0). On the other hand, owing to (1.11), to the continuous dependence of the principal eigenvalue with respect to the potential in the differential operator and on the boundary conditions (cf.[8, Corollary 3.4, Theorem 8.2]), to the regularity of the branch (2.4) given by the implicit function theorem and to (2.3), we get

lim γ γ 0 σ 1 Ω [ ( λ , u γ ) , ( γ , u γ ) ] = σ 1 Ω [ ( λ , u 0 ) , ( γ 0 , u 0 ) ] > 0 ,

and hence, shortening ε>0 if it is necessary, we have

(2.5) σ 1 Ω [ ( λ , u γ ) , ( γ , u γ ) ] > 0 for all γ ( γ 0 - ε , γ 0 + ε ) ,

which proves that (γ,uγ) is linearly asymptotically stable for all γ(γ0-ε,γ0+ε).

We now prove (2.1). Indeed, differentiating (1.1) with respect to the γ-parameter along the branch (2.4) yields

(2.6) { ( λ , u γ ) u ˙ γ = 0 in Ω , ( γ , u γ ) u ˙ γ = ( 0 , b ( x ) u γ q ) > 0 on Ω .

Thus, by (2.5), it follows from the characterization of the strong maximum principle that (2.6) satisfies the strong maximum principle, and therefore (2.1) holds. Now (2.2) is straightforward from (2.1). ∎

The following result will be used to prove Proposition 2.3.

Lemma 2.2.

Assume a0 in Ω; let (γ0,u0) be a positive solution of (1.1) with γ0>0 such that

σ 1 Ω [ ( λ , u 0 ) , ( γ 0 , u 0 ) ] = 0 ,

and let ϕ denote a principal eigenfunction associated to σ1Ω[L(λ,u0),B(γ0,u0)]. Then the following holds:

(2.7) ( p - 1 ) Ω a ( x ) u 0 p - 2 ϕ 3 = γ 0 ( q - 1 ) Γ 1 b ( x ) ϕ 3 u 0 q - 2 - 2 Ω ϕ u 0 | ϕ - ϕ u 0 u 0 | 2 .

Proof.

By construction, we have

(2.8) Ω ( ϕ u 0 ) 2 ( - ϕ Δ u 0 + u 0 Δ ϕ ) = ( p - 1 ) Ω a ( x ) u 0 p - 2 ϕ 3 .

Also, since ϕ0, u00 in Ω, we get ϕ(x)<0, u0(x)<0 for all xΓ0 and ϕu0𝒞1(Ω¯). Then, since

ϕ 3 u 0 2 u 0 | Γ 0 = ( ϕ u 0 ) 2 ϕ u 0 | Γ 0 = 0 and ϕ 2 u 0 ϕ | Γ 0 = ( ϕ u 0 ) ϕ ϕ | Γ 0 = 0 ,

integrating by parts on the left-hand side of (2.8) and taking into account

( ϕ 3 u 0 2 ) , u 0 = 3 ϕ 2 u 0 2 ϕ , u 0 - 2 ϕ 3 u 0 3 | u 0 | 2 ,
( ϕ 2 u 0 ) , ϕ = 2 ϕ u 0 | ϕ | 2 - ( ϕ u 0 ) 2 ϕ , u 0 ,

we get

(2.9) Ω ( ϕ u 0 ) 2 ( - ϕ Δ u 0 + u 0 Δ ϕ ) = γ 0 ( q - 1 ) Γ 1 b ( x ) ϕ 3 u 0 q - 2 - 2 Ω ϕ u 0 | ϕ - ϕ u 0 u 0 | 2 .

Now (2.8) and (2.9) imply (2.7). ∎

The next result provides us the local structure of the bifurcation diagram of positive solutions of (1.1) around any positive neutrally stable solution.

Proposition 2.3.

Assume a0 in Ω; let (γ0,u0) be a positive neutrally solution of (1.1) for γ0>0, and let ϕ denote a principal eigenfunction associated to σ1Ω[L(λ,u0),B(γ0,u0)]. Then there exist ε>0 and a differentiable function

( γ , u ) : ( - ε , ε ) × 𝒞 Γ 0 1 ( Ω ¯ ) , s ( γ ( s ) , u ( s ) )

such that (γ(0),u(0))=(γ0,u0), (γ(s),u(s)) is a positive solution of (1.1) for each s(-ε,ε), there exists a neighborhood U of (γ0,u0) in R×CΓ01(Ω¯) such that if (γ,u)U is a positive solution of (1.1), then

( γ , u ) = ( γ ( s ) , u ( s ) ) for some s ( - ε , ε ) ,

and in addition,

(2.10) γ ( s ) = γ 0 + β s 2 + O ( s 3 ) , u ( s ) = u 0 + s ϕ + v ( s ) ,

where v(s)=O(s2) as s0, Ωv(s)ϕ=0 for each s(-ε,ε) and

(2.11) β = p ( p - 1 ) Ω a ( x ) u 0 p - 2 ϕ 3 - γ 0 q ( q - 1 ) Γ 1 b ( x ) u 0 q - 2 ϕ 3 2 Γ 1 b ( x ) u 0 q ϕ

and also,

(2.12) β = γ 0 ( q - 1 ) ( p - q ) Γ 1 b ( x ) u 0 q - 2 ϕ 3 - 2 p Ω ϕ u 0 | ϕ - ϕ u 0 u 0 | 2 2 Γ 1 b ( x ) u 0 q ϕ .

Moreover,

(2.13) sign ( γ ( s ) ) = sign ( σ 1 Ω [ ( λ , u ( s ) ) , ( γ ( s ) , u ( s ) ) ] ) .

In particular, if pq, then β<0,

(2.14) Ω a ( x ) u 0 p - 2 ϕ 3 Γ 1 b ( x ) u 0 q - 2 ϕ 3 < γ 0 q ( q - 1 ) p ( p - 1 ) ,

and the set of positive solutions of (1.1) around (γ0,u0) possesses the structure of a quadratic subcritical turning point, and there exists δ>0 small enough such that (1.1) possesses at least two positive solutions for each γ(γ0-δ,γ0), one of them linearly asymptotically stable in the lower branch, and the other linearly unstable in the upper branch.

Proof.

The existence of the branch of positive solutions of (1.1)

{ ( γ ( s ) , u ( s ) ) : s ( - ε , ε ) } ,

its local uniqueness in a neighborhood 𝔘 of (γ0,u0) and the structure of the expansions (2.10) and (2.11) follow from [1, Theorem 2.1], applied to the operator 𝔉λ(γ,u) at (γ,u)=(γ0,u0) (cf. (1.9)).

Now (2.12) follows from (2.11) and Lemma 2.2.

To prove (2.13), we will adapt the arguments given in [2, Proposition 20.8] and [17, Proposition 2.8] to the current framework. Indeed, taking into account that (γ(s),u(s)) is a positive solution of (1.1) for s(-ε,ε), differentiating (1.1) with respect to s yields

(2.15) { ( λ , u ( s ) ) u ( s ) = 0 in Ω , ( γ ( s ) , u ( s ) ) u ( s ) = ( 0 , γ ( s ) b ( x ) u q ( s ) ) on Ω .

Now consider ϕ(s) a principal eigenfunction associated to the principal eigenvalue σ1Ω[(λ,u(s)),(γ(s),u(s))]. By definition, the following holds:

(2.16) { ( λ , u ( s ) ) ϕ ( s ) = σ 1 Ω [ ( λ , u ( s ) ) , ( γ ( s ) , u ( s ) ) ] ϕ ( s ) in Ω , ( γ ( s ) , u ( s ) ) ϕ ( s ) = 0 on Ω .

Multiplying the differential equation of (2.15) by ϕ(s) and the differential equation of (2.16) by u(s), and integrating by parts in Ω, the following holds:

σ 1 Ω [ ( λ , u ( s ) ) , ( γ ( s ) , u ( s ) ) ] Ω ϕ ( s ) u ( s ) = γ ( s ) Γ 1 b ( x ) u q ( s ) ϕ ( s ) .

Now, since u(0)=ϕ0 in Ω, by continuity and owing to (2.10), we get u(s)0 in Ω for s0, and hence Ωϕ(s)u(s)>0 for s0. Then, since Γ1b(x)uq(s)ϕ(s)>0, we obtain that (2.13) holds.

Finally, the fact that β<0 if pq follows from (2.12), taking into account that γ0>0, and (2.14) follows from (2.11), taking into account that β<0. The remaining assertions follow from (2.10) and (2.13). ∎

3 Proof of Theorem 1.1

Along this section, it will be assumed that a=0 in Ω. This section delivers the proof of Theorem 1.1. Some preliminaries are imperative for it.

Lemma 3.1.

Assume (1.17). Then there exist C(λ)>0 and, for each γ<0, a positive strict subsolution φ~λ,γ of (1.14) such that

(3.1) φ ~ λ , γ L ( Ω ) = C ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1

and

(3.2) lim γ 0 φ ~ λ , γ ( x ) = for all x Ω Γ 1 .

Moreover, (3.2) holds uniformly in compact subsets contained in ΩΓ1.

Proof.

For each ε>0, let us denote σ1ε:=σ1Ω[-Δ,𝔅(V(x)+ε)]. Owing to [8, Theorem 8.2, Remark 8.3] and [8, Proposition 3.5], we have

(3.3) σ 1 ε > σ 1 for all ε > 0 and lim ε 0 σ 1 ε = σ 1 .

Since λ(σ1,σ0) and (3.3) holds, there exists a sufficiently small ε1=ε1(λ)>0 such that

(3.4) σ 1 < σ 1 ε < λ for all ε ( 0 , ε 1 ) .

Fix ε:=ε(λ)>0 satisfying (3.4), and let us denote by φε the principal eigenfunction associated to σ1ε normalized so that

(3.5) φ ε L ( Ω ) = 1 .

Now let us consider the function φ~λ,γ:=μ(λ,γ)φε, where

(3.6) μ ( λ , γ ) : = C ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1 > 0 , C ( λ ) = C ( ε ( λ ) ) : = ( ε 2 ) 1 q - 1 .

We are going to prove that, for each γ<0, φ~λ,γ is a positive strict subsolution of (1.14). Indeed, by construction and owing to (3.4), the following holds in Ω:

(3.7) ( - Δ - λ ) φ ~ λ , γ = μ ( λ , γ ) ( σ 1 ε - λ ) φ ε < 0 .

As for the boundary conditions,

(3.8) φ ~ λ , γ | Γ 0 = μ ( λ , γ ) φ ε | Γ 0 = 0 ,

and since γ~>0, b0, φε>0 and (3.5) holds, by construction and taking into account (3.6), the following holds on Γ1:

(3.9) φ ~ λ , γ + V ( x ) φ ~ λ , γ + γ ~ b ( x ) φ ~ λ , γ q = μ ( λ , γ ) φ ε ( - ε + γ ~ ( μ ( λ , γ ) ) q - 1 b ( x ) φ ε q - 1 ) μ ( λ , γ ) φ ε ( - ε + γ ~ μ ( λ , γ ) q - 1 b ¯ ) = - ε 2 μ ( λ , γ ) φ ε < 0 .

Moreover, owing to (3.5), we find that

(3.10) φ ~ λ , γ L ( Ω ) = μ ( λ , γ ) = C ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1 .

Then (3.7), (3.8), (3.9) and (3.10) imply that φ~λ,γ is a positive strict subsolution of (1.14) satisfying (3.1).

On the other hand, since φε is strongly positive in Ω, we get φε(x)>0 for all xΩΓ1, and hence

lim γ 0 φ ~ λ , γ ( x ) = lim γ ~ 0 φ ~ λ , γ ( x ) = lim γ ~ 0 ( ε 2 ) 1 q - 1 ( 1 b ¯ γ ~ ) 1 q - 1 φ ε ( x ) = ,

which proves (3.2). To prove that (3.2) holds uniformly in compact subsets contained in ΩΓ1, let K be a compact subset contained in ΩΓ1. Since φε is bounded away from zero in any compact subset KΩΓ1, we get mK:=minxKφε(x)>0. Then, for each xK,

(3.11) φ ~ λ , γ ( x ) = μ ( λ , γ ) φ ε ( x ) μ ( λ , γ ) m K = ( ε 2 ) 1 q - 1 ( 1 b ¯ γ ~ ) 1 q - 1 m K > 0 ,

and since q>1, taking limits in (3.11) when γ0, we get that (3.2) holds uniformly in K. ∎

Lemma 3.2.

Assume (1.17) and (1.2). Then there exist D(λ)>0 and, for each γ<0, a positive strict supersolution φ¯λ,γ of (1.14) satisfying

(3.12) φ ¯ λ , γ L ( Ω ) = D ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1

and

(3.13) φ ¯ λ , γ φ ~ λ , γ ,

where φ~λ,γ is the positive strict subsolution of (1.14) built in Lemma 3.1.

Proof.

For each n𝐍, let us denote σ1n:=σ1Ω[-Δ,𝔅(V(x)+n)], and by φn the principal eigenfunction associated to σ1n, normalized so that

(3.14) φ n L ( Ω ) = 1 .

Since φn is strongly positive in Ω, φn(x)>0 for all xΓ1. Let us denote

(3.15) m n : = min x Γ 1 φ n ( x ) > 0 , g ( n ) : = ( n + 1 ) 1 q - 1 m n > 0

Owing to [8, Proposition 3.1] and [8, Theorem 9.1], we have

(3.16) σ 1 n < σ 0 for all n and lim n σ 1 n = σ 0 .

Since λ(σ1,σ0) and (3.16) holds, there exists n=n(λ)𝐍 large enough such that

(3.17) σ 1 < λ < σ 1 n < σ 0 .

Fix n:=n(λ)𝐍 satisfying (3.17), and let us denote D~(λ):=g(n(λ))>0. Let φε and C(λ) be the principal eigenfunction associated to σ1ε and the expression C(λ) considered in Lemma 3.1, respectively. Since φε and φn are strongly positive in Ω, there exists K(λ)1 such that

(3.18) C ( λ ) φ ε K ( λ ) D ~ ( λ ) φ n .

Now let us denote

(3.19) D ( λ ) : = K ( λ ) D ~ ( λ ) , α ( λ , γ ) : = D ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1 > 0 ,

and let us consider the function φ¯λ,γ=α(λ,γ)φn. We are going to prove that, for each γ<0, φ¯λ,γ is a positive strict supersolution of (1.14) satisfying (3.12) and (3.13). Indeed, by construction, owing to (3.17) and (3.19), the following holds in Ω:

(3.20) ( - Δ - λ ) φ ¯ λ , γ = α ( λ , γ ) ( σ 1 n - λ ) φ n > 0 .

As for the boundary conditions,

(3.21) φ ¯ λ , γ | Γ 0 = α ( λ , γ ) φ n | Γ 0 = 0 ,

and since γ~>0, b(x)b¯>0, φn>0, k(λ)1, q>1 and taking into account the definition of α(λ,γ) (cf. (3.15), (3.19)), by construction, the following holds on Γ1:

(3.22) φ ¯ λ , γ + V ( x ) φ ¯ λ , γ + γ ~ b ( x ) φ ¯ λ , γ q = α ( λ , γ ) φ n ( γ ~ b ( x ) ( α ( λ , γ ) ) q - 1 φ n q - 1 - n ) α ( λ , γ ) φ n ( γ ~ b ¯ ( α ( λ , γ ) ) q - 1 m n q - 1 - n ) = α ( λ , γ ) φ n ( ( n + 1 ) ( k ( λ ) ) q - 1 - n ) α ( λ , γ ) m n > 0 .

On the other hand, by construction and owing to (3.14), we obtain

(3.23) φ ¯ λ , γ L ( Ω ) = α ( λ , γ ) = D ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1 .

Then (3.20)–(3.23) prove that φ¯λ,γ is a positive strict supersolution of (1.14) satisfying (3.12).

Finally, (3.13) follows by construction, taking into account (3.18). ∎

Now we are going to prove Theorem 1.1.

Proof of Theorem 1.1.

(i) To prove the necessary condition, we will argue by contradiction. Assume there exists uγ, a positive solution of (1.14) for some γ0. Then (1.13) holds (for a=0), and owing to the monotonicity of the principal eigenvalue with respect to the potential on the boundary (cf. [8, Proposition 3.5]) and since b0, we find that

λ = σ 1 Ω [ - Δ , 𝔅 ( V ( x ) - γ b ( x ) u γ q - 1 ) ] σ 1 Ω [ - Δ , 𝔅 ( V ( x ) ) ] = σ 1 ,

which contradicts (1.17). The sufficient condition follows from the fact that Λ(γ)=(σ1,σ0) for each γ<0 (cf. (1.16), [7, Theorem 3.1]). The proof of the uniqueness of the positive solution of (1.14) for each γ<0 follows arguing in the same way as in the proof of [7, Theorem 3.1 (ii)].

(ii) Let γ<0, and let uγ be the unique positive solution of (1.14) for such a value γ<0 (guaranteed by (i)). The linearization of (1.14) in uγ is given by

( ( λ , u γ ) , 𝔅 ( γ , u γ ) ) : = ( - Δ - λ , 𝔅 ( V ( x ) + q γ ~ b ( x ) u γ q - 1 ) ) .

Now, since, by (1.13),

(3.24) λ = σ 1 Ω [ - Δ , 𝔅 ( V ( x ) + γ ~ b ( x ) u γ q - 1 ) ] ,

owing to the facts that b0, q>1 and γ~>0, it follows from the monotonicity of the principal eigenvalue with respect to the potential on the boundary ([8, Proposition 3.5]) and (3.24) that

σ 1 Ω [ ( λ , u γ ) , 𝔅 ( γ , u γ ) ] > σ 1 Ω [ - Δ - λ , 𝔅 ( V ( x ) + γ ~ b ( x ) u γ q - 1 ) ] = 0 ,

which proves that uγ is linearly asymptotically stable.

(iii) It follows from the existence and uniqueness of positive solution uγ of (1.14) for any γ<0, and the fact that it is linearly asymptotically stable for any γ<0 (guaranteed by (i), (ii)) and Proposition 2.1.

(iv) The existence of the branch of positive solutions Γ+(λ) and the fact that all the positive solutions of (1.14) for λ(σ1,σ0) belong to Γ+(λ) follow from (i). The increasing character of the branch Γ+(λ) in L(Ω)-norm with respect to the parameter γ<0 follows from (1.18). On the other hand, to prove (1.20) and (1.21), let us consider, for each γ<0, the unique positive solution uγ of (1.14), the positive strict supersolution φ¯λ,γ of (1.14) given by Lemma 3.2 and the positive strict subsolution φ~λ,γ of (1.14) given by Lemma 3.1 satisfying (3.13). Owing to the sub-supersolution method (cf. [2]) and the uniqueness of positive solution of (1.14), for each γ<0, we get

(3.25) φ ~ λ , γ < u γ < φ ¯ λ , γ ,

and hence

(3.26) φ ~ λ , γ L ( Ω ) u γ L ( Ω ) φ ¯ λ , γ L ( Ω ) .

Now (3.1), (3.12) and (3.26) imply (1.20). Finally, taking limits in (1.20) as γ0 and as γ-, we obtain, respectively,

lim γ 0 u γ L ( Ω ) = , lim γ - u γ L ( Ω ) = 0 ,

which prove (1.21). This completes the proof of (iv).

(v) It follows from (3.25), taking into account that, owing to Lemma 3.1, (3.2) holds uniformly in compact subsets contained in ΩΓ1.

Finally, (vi) is [7, Theorem 3.1 (ii), Theorem 3.8 (ii)]. ∎

4 A One-Dimensional Example of Case (1)

In this section, we will consider the one-dimensional mixed boundary value problem

(4.1) { - u ′′ ( x ) = λ u ( x ) , 0 < x < 1 , u ( 0 ) = 0 , u ( 1 ) - u ( 1 ) = γ b ( u ( 1 ) ) q , q > 1 ,

where γ<0 and b>0. This problem fits into the framework of (1.14) with

N = 1 , Ω = ( 0 , 1 ) , Γ 0 = { x 0 = 0 } , Γ 1 = { x 1 = 1 } , b ¯ = b ¯ = b , V = - 1

It can be explicitly solved, and the main goal of this section is to show, through the explicit solution of (4.1), the consequences of Theorem 1.1 in the general case. In this framework, σ1=0 is the principal eigenvalue associated to the eigenvalue problem

{ - u ( x ) ′′ = λ u ( x ) , 0 < x < 1 , u ( 0 ) = 0 , u ( 1 ) - u ( 1 ) = 0 ,

and σ0=π2 is the principal eigenvalue associated to the eigenvalue problem

{ - u ( x ) ′′ = λ u ( x ) , 0 < x < 1 , u ( 0 ) = 0 , u ( 1 ) = 0 .

Then, taking any fixed λ(σ1,σ0)=(0,π2), it is straightforward to see that, for each γ<0, (4.1) possesses a unique positive solution (as Theorem 1.1 (i) establishes). Namely,

(4.2) u γ ( x ) = H ( λ ) ( 1 b γ ~ ) 1 q - 1 sin ( λ x ) ,

where

H ( λ ) : = ( sin ( λ ) - λ cos ( λ ) ( sin ( λ ) ) q ) 1 q - 1 > 0

It should be noted that H(λ)>0 for each λ(0,π2). Indeed, if λ(0,π24), then

sin ( λ ) > 0 , cos ( λ ) > 0 , tan ( λ ) > λ ,

which imply H(λ)>0, and if λ[π24,π2), then H(λ)>0 because, in this case, sin(λ)>0 and cos(λ)0. It is easily seen that, for any fixed λ(0,π2) and any γ<0,

(4.3) u γ L ( 0 , 1 ) = u γ ( x λ ) = H ~ ( λ ) ( 1 b γ ~ ) 1 q - 1 ,

where

x λ = { 1 if λ ( 0 , π 2 4 ] , π 2 λ if λ ( π 2 4 , π 2 ) ,    H ~ ( λ ) = { H ( λ ) sin ( λ ) if λ ( 0 , π 2 4 ] , H ( λ ) if λ ( π 2 4 , π 2 ) .

Also, since λ(0,π), for any x(0,1], we get λx(0,π); hence sin(λx)>0, and therefore it follows from (4.2) that

(4.4) lim γ 0 u γ ( x ) = lim γ ~ 0 u γ ( x ) = , 0 < x 1 .

According to Theorem 1.1, in this particular example, (4.3) shows that the branch Γ+(λ):={(γ,uγ):γ<0} is increasing with γ in L(Ω), that (1.20) holds for C(λ)=D(λ)=H~(λ), and that Γ+(λ) bifurcates to positive solutions from the trivial branch (γ,u)=(γ,0) as γ- and from infinity as γ0, with rate γ~-1q-1 in both cases, up to multiplicative constants. Similarly, (4.4) and (4.2) show that (1.22) holds uniformly in compact subsets of ΩΓ1=(0,1].

5 Proof of Theorem 1.2

In the sequel, for each ρ>0, we set

𝒜 ρ : = ( Γ 1 + B ρ ( 0 ) ) Ω , Γ 0 ρ : = 𝒜 ρ Γ 1 Ω , Ω ρ : = Ω 𝒜 ¯ ρ ,

where Bρ(0) denotes the ball of N centered at the origin with radius ρ>0. The following result will be used in the proof of Theorem 1.2.

Lemma 5.1.

Let I=[α,β]R with 0α<β, and let us assume that (1.17) holds and there exist constants ρ~>0 and C:=C(ρ~)>0 such that

(5.1) u γ L ( 𝒜 ρ ~ ) C for all γ [ α , β ] ,

where uγ denotes any positive solution of (1.1) for the value γ>0 of the parameter. Then there exists C1>0 such that

(5.2) u γ L ( Ω ) C 1 for all γ [ α , β ] .

Proof.

Let ρ~>0, and let C:=C(ρ~)>0 satisfy (5.1). Let ρ1(0,ρ~) be sufficiently small so that

𝒜 ¯ ρ 1 Ω ¯ 0 = , 𝒜 ¯ ρ 1 Γ 0 = ,

whose existence is guaranteed by the facts that Γ1Γ0= and Ω¯0Ω. Owing to (5.1), any positive solution uγ of (1.1) for γI satisfies

(5.3) { - Δ u γ = λ u γ - a ( x ) u γ p in Ω ρ 1 , u γ = 0 on Γ 0 , u γ C on Γ 0 ρ 1 .

Also, let u0 be the unique positive solution of (1.1) for γ=0, whose existence and uniqueness are guaranteed by [11, Theorem 3.5], [5, Theorem 4.2]. Since u0 is strongly positive in Ω, we get

u ¯ 0 : = min x Γ 0 ρ 1 u 0 ( x ) > 0 ,

and taking

k > max { 1 , C u ¯ 0 } ,

it is easy to see that the function u¯:=ku0 satisfies

(5.4) { ( - Δ - λ + a ( x ) u ¯ p - 1 ) u ¯ > 0 in Ω ρ 1 , u ¯ = 0 on Γ 0 , u ¯ > C > 0 on Γ 0 ρ 1 .

Then, owing to (5.3) and (5.4), the function ψγ:=u¯-uγ satisfies

(5.5) { ( - Δ - λ + a ( x ) u ¯ p - u γ p u ¯ - u γ ) ψ γ > 0 in Ω ρ 1 , ψ γ = 0 on Γ 0 , ψ γ > 0 on Γ 0 ρ 1 .

On the other hand, since u¯>0 fits (5.4), it follows from the characterization of the strong maximum principle (cf. [4, Theorem 2.4]) that

(5.6) σ 1 Ω ρ 1 [ - Δ - λ + a ( x ) u ¯ p - 1 , 𝒟 ] > 0 .

Also, since a0 and

u ¯ p - u γ p u ¯ - u γ > u ¯ p - 1 ,

it follows from the monotonicity of the principal eigenvalue with respect to the potential and (5.6) that

(5.7) σ 1 Ω ρ 1 [ - Δ - λ + a ( x ) u ¯ p - u γ p u ¯ - u γ , 𝒟 ] > σ 1 Ω ρ 1 [ - Δ - λ + a ( x ) u ¯ p - 1 , 𝒟 ] > 0 .

Now, owing again to the characterization of the strong maximum principle, it follows from (5.7) that (5.5) satisfies the strong maximum principle, and hence ψγ0 in Ω¯ρ1, that is,

(5.8) u γ u ¯ in Ω ¯ ρ 1 for all γ [ α , β ] .

Now (5.1) and (5.8) imply

u γ L ( Ω ) C 1 : = max { C , u ¯ L ( Ω ρ 1 ) } for all γ [ α , β ] ,

which proves (5.2). ∎

Proof of Theorem 1.2.

(i) The proofs of the existence and uniqueness of positive solution for each γ0 follow from [11, Theorem 3.5], [5, Theorem 4.2] in the particular case when γ=0, and from [6, Theorem 3.1] when γ<0. The fact that uγ is linearly asymptotically stable for each γ0 follows arguing as in the proof of Theorem 1.1 (ii), taking into account that a0 in Ω and using the monotonicity of the principal eigenvalue with respect to the potential in the differential operator (cf. [8, Proposition 3.3]). The remaining assertions of (i) follow from the fact that uγ is linearly asymptotically stable for each γ0, by Proposition 2.1. This completes the proof of (i).

(ii) Since λ(σ1,σ0*)(σ1,σ0) (cf. (1.7), (1.8)), for each γ<0, let vγ be the unique positive solution of (1.14), whose existence and uniqueness are guaranteed in Theorem 1.1, and let uγ be the unique positive solution of (1.1), whose existence and uniqueness have been proved in (i). Owing to the fact that (1.2) holds, it follows from (1.20) the existence of D(λ)>0 such that

(5.9) v γ L ( Ω ) D ( λ ) ( 1 b ¯ γ ~ ) 1 q - 1 , γ < 0 .

On the other hand, it is easy to prove that uγ is a positive strict subsolution of (1.14), and hence, owing to the characterization of the strong maximum principle (cf. [4, Theorem 2.4]), we obtain (cf.[7, Lemma 3.5])

(5.10) u γ < v γ .

Now (1.25) follows from (5.9) and (5.10), and (1.26) follows letting γ- in (1.25). This completes the proof of (ii).

(iii) It follows from the fact that u0 is linearly asymptotically stable, as it has been proved in (i), and Proposition 2.1.

(iv) Let u^γ be any positive solution of (1.1) for some γ>0. Then, by construction, the function φγ:=u^γ-u0 satisfies

(5.11) { ( - Δ - λ + a ( x ) u ^ γ p - u 0 p u ^ γ - u 0 ) φ γ = 0 in Ω , φ γ = 0 on Γ 0 , φ γ + V ( x ) φ γ = γ b ( x ) u ^ γ q > 0 on Γ 1 .

Also, due to (1.13), we obtain

(5.12) σ 1 Ω [ - Δ - λ + a ( x ) u 0 p - 1 , 𝔅 ( V ( x ) ) ] = 0 ,

and since a0 in Ω and

u ^ γ p - u 0 p u ^ γ - u 0 > u 0 p - 1 0 ,

it follows from the monotonicity of the principal eigenvalue with respect to the potential (cf. [8, Proposition 3.3]) and (5.12) that

(5.13) σ 1 Ω [ - Δ - λ + a ( x ) u ^ γ p - u 0 p u ^ γ - u 0 , 𝔅 ( V ) ] > σ 1 Ω [ - Δ - λ + a ( x ) u 0 p - 1 , 𝔅 ( V ) ] = 0 .

Now, owing to the characterization of the strong maximum principle, it follows from (5.13) that

( - Δ - λ + a ( x ) u ^ γ p - u 0 p u ^ γ - u 0 , 𝔅 ( V ) , Ω )

satisfies the strong maximum principle, and therefore (5.11) implies φγ0 in Ω, that is, (1.27) holds. This completes the proof of (iv).

(v) The proof of (v) (a) follows from [6, Theorem 4.7 (i)] since, in particular, it establishes that

( σ 1 , σ 0 ) γ > 0 Λ ( γ ) .

To prove (v) (b), let u^γ be any positive solution of (1.1) for some γ>0, whose existence is guaranteed by (v) (a). Owing to (iv), (1.27) holds. On the other hand, it is easy to see that u0 is a positive strict subsolution of (1.1) for each γ>0. Then we get that [u0,u^γ] is an ordered sub-supersolution pair of (1.1) for γ>0, and hence [2, Theorem 6.1] implies the existence of a minimal solution uγmin of (1.1) in the interval (u0,u^γ] for each γ>0, which, together with (iv), proves (1.28).

We now prove (1.29). Indeed, by the definition of uγmin and owing to (1.28), we get, for each γ(0,ε0),

(5.14) u 0 u γ min u γ * .

On the other hand, by the continuity of the branch Γ+(λ,ε0)={(γ,uγ*):γ[0,ε0)}, we find that

(5.15) lim γ 0 u γ * - u 0 L ( Ω ) = 0 ,

and hence (5.14) and (5.15) imply that

(5.16) lim γ 0 u γ min - u 0 L ( Ω ) = 0 .

Now (1.29) follows from (5.16) and the fact that the unique positive solution of (1.1) for each γ(-ε0,ε0) in a small neighborhood 𝔘 of (γ,u)=(0,u0) in (-ε0,ε0)×𝒞Γ01(Ω¯) is given by uγ* (cf. (iii)). This ends the proof of (v) (b).

We now prove (v) (c). The existence of uniform L(Ω)-bounds when γ varies in compact intervals I=[α,β](-,0] follows from the existence and uniqueness of positive solutions of (1.1) guaranteed by (i) and the fact that the map (1.24) is increasing in L(Ω). In particular, uγL(Ω)u0L(Ω) for all γ0. To prove the existence of uniform L(Ω)-bounds when γ varies in compact intervals I=[α,β][0,), we will argue by contradiction, using rescaling and blow-up arguments, much like in [6, Theorem 4.4], but now, varying γ instead of λ. Suppose there exists a compact interval I=[α,β] with α0 of values of γ such that the positive solutions of (1.1) are not uniformly bounded in L(Ω). Then Lemma 5.1 implies the existence of a sequence (γk,uk), k1, of positive solutions of (1.1) with γkI, k1, a sequence of points {xk}k=1Ω¯ and a point xΓ1 such that

(5.17) M k : = u k ( x k ) = sup x Ω u k ( x ) , x k x Γ 1 .

Since γkI, k1, taking a subsequence if it is necessary, we can assume without loss of generality that

(5.18) lim k γ k = γ [ α , β ] .

Also, since xΓ1 and the potential a is bounded away from zero in any compact subset of (ΩΩ¯0)Γ1 (cf. (1.3)), we obtain

(5.19) a ( x ) > 0

Set

(5.20) ρ k : = M k - p - 1 2 .

Since p>1, (5.17) implies

(5.21) lim k ρ k = 0 .

We will denote by xj the j-th coordinate of x. Since Γ1 is smooth enough, by straightening Γ1 in a neighborhood of x by a non-singular change of coordinates, we can assume that

B δ ( x ) Ω = { x N : x N > 0 } B δ ( x ) ,
B δ ( x ) Γ 1 = { x N : x N = 0 } B δ ( x ) ,

for some fixed δ small enough, where Bδ(x) stands for the ball of radius δ>0 centered at the point x. Since xkx as k, |xk-x|<δ2, for k large enough, and hence |x-xk|<δ2 implies |x-x|<δ. Thus, for k large enough, Bδ2(xk)Bδ(x), and hence

Ω k : = B δ 2 ( x k ) { x N : x N > 0 } B δ ( x ) Ω .

Now, taking into account (5.20) and making the change of variable

(5.22) y = x - x k ρ k , v k ( y ) : = u k ( x ) M k = u k ( x k + ρ k y ) M k ,

Ω k transforms into

Ω ~ k : = B δ 2 ρ k ( 0 ) { y N : y N > - x k N ρ k } ,

the differential equation

- Δ u k = λ u k ( x ) - a ( x ) u k p ( x ) , x Ω k ,

into

- Δ v k ( y ) = λ ρ k 2 v k ( y ) - a ( x k + ρ k y ) v k p ( y ) , y Ω ~ k

and the boundary condition

u k + V ( x ) u k ( x ) = γ k b ( x ) u k q ( x ) , x Ω ¯ k Γ 1 ,

into

- v k ( y ) y N + V ( x k + ρ k y ) ρ k v k ( y ) = γ k b ( x k + ρ k y ) M k 2 q - 1 - p 2 v k q ( y ) ,

with

y { y N : y N = - x k N ρ k } B ¯ δ 2 ρ k ( 0 )

By construction, we get that, for k1 large enough, vk(y) is well defined in Ω~k, and owing to (5.22), the following holds:

(5.23) 0 < v k ( y ) v k ( 0 ) = u k ( x k ) M k - 1 = 1 , y Ω ~ k .

Now, arguing as in [6, Theorem 4.4], distinguishing the different behaviors of the sequence xkNρk as k, and taking into account that p>2q-1, (5.17)–(5.19), (5.21) and (5.23), we get that if the sequence xkNρk is not bounded away from zero, then there exists a non-negative function wWr2(BR+(0)) for some R>0, attaining its maximum at y=0 and satisfying

- Δ w ( 0 ) = - a ( x ) w p ( 0 ) = - a < 0 ,

which contradicts the fact that w(y) attains its maximum at y=0, and if the sequence xkNρk is either not bounded above, or it is bounded above and bounded away from zero, then there exists of a non-negative function v(y)Wr2(BR(0)) for some R>0, satisfying

- Δ v ( y ) = - a ( x ) v p ( y ) < 0 in B R ( 0 )

and attaining its maximum at y=0, which gives again a contradiction. This completes the proof of (v) (c). ∎


Communicated by Julian Lopez Gomez


Award Identifier / Grant number: MTM2015-65899-P

Funding statement: Supported by the Ministry of Economy and Competitiveness under grant MTM2015-65899-P.

Acknowledgements

The author would like to thank the student Ignacio Sanz Soriano for generating all the graphics of this work, so substantially clarifying the exposition of the results.

References

[1] H. Amann, Multiple positive fixed points of asymptotically linear maps, J. Funct. Anal. 17 (1974), 174–213. 10.1016/0022-1236(74)90011-1Search in Google Scholar

[2] H. Amann, Fixed point equations and nonlinear eigenvalue problems in ordered Banach spaces, SIAM Rev. 18 (1976), 620–709. 10.1137/1018114Search in Google Scholar

[3] H. Amann, Dual semigroups and second order linear elliptic boundary value problems, Israel J. Math. 45 (1983), 225–254. 10.1007/BF02774019Search in Google Scholar

[4] H. Amann and J. López-Gómez, A priori bounds and multiple solutions for superlinear indefinite elliptic problems, J. Differential Equations 146 (1998), 336–374. 10.1006/jdeq.1998.3440Search in Google Scholar

[5] S. Cano-Casanova, Existence and structure of the set of positive solutions of a general class of sublinear elliptic non-classical mixed boundary value problems, Nonlinear Anal. 49 (2002), 361–430. 10.1016/S0362-546X(01)00116-XSearch in Google Scholar

[6] S. Cano-Casanova, Nonlinear mixed boundary conditions in BVPs of Logistic type with spatial heterogeneities and a nonlinear flux on the boundary with arbitrary sign. The case p>2q-1, J. Differential Equations 256 (2014), 82–107. 10.1016/j.jde.2013.08.011Search in Google Scholar

[7] S. Cano-Casanova, Linear elliptic and parabolic PDEs with nonlinear mixed boundary conditions and spatial heterogeneities, Electron. J. Differential Equations 2018 (2018), 1–27. Search in Google Scholar

[8] S. Cano-Casanova and J. López-Gómez, Properties of the principal eigenvalues of a general class of non-classical mixed boundary value problems, J. Differential Equations 178 (2002), 123–211. 10.1006/jdeq.2000.4003Search in Google Scholar

[9] S. Cano-Casanova and J. López-Gómez, Permanence under strong aggressions is possible, Ann. Inst. Henri Poincaré D 20 (2003), no. 6, 99–1041. 10.1016/s0294-1449(03)00022-2Search in Google Scholar

[10] D. Daners and J. López-Gómez, Global dynamics of generalized logistic equations, Adv. Nonlinear Stud. 18 (2018), no. 2, 217–236. 10.1515/ans-2018-0008Search in Google Scholar

[11] J. M. Fraile, P. Koch Medina, J. López-Gómez and S. Merino, Elliptic eigenvalue problems and unbounded continua of positive solutions of a semilinear elliptic equation, J. Differential Equations 127 (1996), no. 1, 295–319. 10.1006/jdeq.1996.0071Search in Google Scholar

[12] J. Garcia-Melián, J. D. Rossi and J. C. Sabina de Lis, A bifurcation problem governed by the boundary condition. I, NoDEA Nonlinear Differential Equations Appl. 14 (2007), 499–525. 10.1007/s00030-007-4064-xSearch in Google Scholar

[13] J. Garcia-Melián, J. D. Rossi and J. C. Sabina de Lis, A bifurcation problem governed by the boundary condition. II, Proc. Lond. Math. Soc. (3) 94 (2007), 1–25. 10.1112/plms/pdl001Search in Google Scholar

[14] R. Gómez-Reñasco and J. López-Gómez, The effect of varying coefficients on the dynamics of a class of superlinear indefinite reaction-diffusion equations, J. Differential Equations 167 (2000), 36–72. 10.1006/jdeq.2000.3772Search in Google Scholar

[15] R. Gómez-Reñasco and J. López-Gómez, The uniqueness of the stable positive solutions for a class of superlinear indefinite reaction diffusion equations, Differential and Integral Equations 14 (2001), no. 6, 751–768. 10.57262/die/1356123245Search in Google Scholar

[16] J. López-Gómez, Linear Second Order Elliptic Operators, World Scientific, Hackensack, 2013. 10.1142/8664Search in Google Scholar

[17] J. López-Gómez, M. Molina-Meyer and A. Tellini, The uniqueness of the linearly stable positive solution for a class of superlinear indefinite problems with nonhomogeneous boundary conditions, J. Differential Equations 255 (2013), 503–523. 10.1016/j.jde.2013.04.019Search in Google Scholar

[18] E. M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University, Princeton, 1970. 10.1515/9781400883882Search in Google Scholar

Received: 2019-03-29
Accepted: 2019-05-23
Published Online: 2019-07-18
Published in Print: 2020-02-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 3.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2019-2051/html
Scroll to top button