Skip to content
BY 4.0 license Open Access Published by De Gruyter October 15, 2021

Sharp Blow-Up Profiles of Positive Solutions for a Class of Semilinear Elliptic Problems

  • Wan-Tong Li , Julián López-Gómez and Jian-Wen Sun EMAIL logo

Abstract

This paper analyzes the behavior of the positive solution θε of the perturbed problem

{ - Δ u = λ m ( x ) u - [ a ε ( x ) + b ε ( x ) ] u p = 0 in Ω , B u = 0 on Ω ,

as ε0, where aε(x)εαa(x) and bε(x)εβb(x) for some α0 and β0, and some Hölder continuous functions a(x) and b(x) such that a0 (i.e., a0 and a0) and minΩ¯b>0. The most intriguing and interesting case arises when a(x) degenerates, in the sense that Ω0inta-1(0) is a non-empty smooth open subdomain of Ω, as in this case a “blow-up” phenomenon appears due to the spatial degeneracy of a(x) for sufficiently large λ. In all these cases, the asymptotic behavior of θε will be characterized according to the several admissible values of the parameters α and β. Our study reveals that there may exist two different blow-up speeds for θε in the degenerate case.

MSC 2010: 35B40; 35K57; 92D25

1 Introduction and Main Results

In this paper, we characterize the limiting behavior as ε0 of the classical positive solutions of the semilinear boundary value problem

(1.1) { - Δ u = λ m ( x ) u - [ a ε ( x ) + b ε ( x ) ] u p = 0 in Ω , B u = 0 on Ω ,

where λ>0 is a real parameter, p>1 is constant, ε(0,1] is a perturbation parameter, Ω is a bounded domain of class 𝒞2+μ of N (N1) for some μ(0,1), whose boundary consists of two disjoint open and closed subsets Γ0 and Γ1 of class 𝒞2+μ, Ω=Γ0Γ1, and B stands for the boundary operator defined for every u𝒞(Γ0)𝒞1(ΩΓ1) by

(1.2) B u : = { u on Γ 0 , u ν + γ ( x ) u on Γ 1 .

In (1.2), ν stands for the outward unit normal to Ω, and γ𝒞μ(Γ1) is an arbitrary non-negative function, γ0. Necessarily, Γ0 and Γ1 must possess finitely many components. As Γ0, or Γ1, can be empty, our mixed boundary operator B includes Dirichlet, Neumann and Robin boundary conditions. By simplicity, we set BD if Γ0=Ω, and BN if Γ1=Ω and γ=0.

A function u:Ω¯ is said to be of class 𝒞κ+μ for some integer κ1 if it is of class 𝒞κ(Ω¯) and Dαu is Hölder continuous in Ω¯, with exponent μ, for all multi-index αN such that |α|=α1++αN=κ. A subdomain DΩ is said to be of class 𝒞κ+μ if the components of its local charts in D are functions of class 𝒞κ+μ.

Throughout this paper, for every W𝒞μ(Ω¯), we denote by σ1[-Δ+W;B,Ω] the principal eigenvalue of the linear eigenvalue problem

{ ( - Δ + W ) φ = σ φ in Ω , B φ = 0 on Ω .

By applying [26, Theorem 7.10] with h1, it becomes apparent that σ1[-Δ;B,Ω]>0 if either Γ0, or Γ0= but γ0. On the contrary, σ1[-Δ;N,Ω]=0. Thus, under the previous general assumptions, σ1[-Δ;B,Ω]0.

As far as m(x), aε(x) and mε(x) are concerned, for every ε(0,1], they are assumed to be functions of class 𝒞μ(Ω¯) such that

  1. m ( x + ) > 0 for some x+Ω,

  2. a ε 0 in Ω for all ε(0,1], and there exists a constant α0 such that

    lim ε 0 a ε ε α = a uniformly in Ω ¯ ,

    for some non-negative function a𝒞μ(Ω¯) such that a-1(0)=Ω¯0Ω, where Ω0 is an open subdomain of class 𝒞2+μ, possibly empty,

  3. min Ω ¯ b ε > 0 for all ε(0,1], and there exists a constant β0 such that

    lim ε 0 b ε ε β = b uniformly in Ω ¯ ,

    for some b𝒞μ(Ω¯) such that b(x)>0 for all xΩ¯.

Thus, aεεαa and bεεβb for small ε>0. This entails that, as soon as α>0 and β>0, the limiting problem of (1.1) as ε0 becomes linear. Thus, it is of a different nature than the classical problems in “homogenization theory”, where, e.g., aε(x)=a(εx). In this case, limε0a(εx)=a(0), and hence, the limiting problem is regular if a(0)>0. Our problem here is of another nature. As it will become apparent later, once having read the proofs of the main results of this paper, assumptions (Hm), (Ha) and (Hb) are optimal for their validity as they stand.

Under these general assumptions, aε(x)+bε(x)>0 for all xΩ¯ and ε(0,1], and hence, for every ε(0,1], problem (1.1) admits a unique positive solution, denoted by θε in this paper, if and only if λ>λ1B(Ω) (see, e.g., Fraile et al. [15] and Aleja, Antón and López-Gómez [1]), where λ1B(Ω)0 stands for the maximal non-negative principal eigenvalue of the weighted eigenvalue problem

(1.3) { - Δ u = λ m ( x ) u in Ω , B u = 0 on Ω .

It is well known that λ1N(Ω)=0, whereas λ1B(Ω)>0 if either Γ0, or Γ0= but γ0 (see [21, Section 1]). According to this notation, for any given subdomain Ω~ of class 𝒞μ of Ω, λ1D(Ω~) stands for the principal eigenvalue of

{ - Δ u = λ m ( x ) u in Ω ~ , u = 0 on Ω ~ .

Their existence goes back to Hess and Kato [19] and Brown and Lin [7] and can be derived, very easily, from the general theory of [26, Chapter 9].

The main goal of this paper is to analyze how the sharp behavior of θε as ε0 depends on the several possible relations between the exponents α0 and β0, which will provide us with each of the shadow limiting problems of (1.1) as ε0. In all cases, the sharp asymptotic profiles of θε as ε0 are regulated by either the positive solutions, or the minimal metasolutions in ΩΩ¯0, of a semilinear elliptic boundary value problem of the form

(1.4) { - Δ u = λ m ( x ) u - c ( x ) u p ( x ) = 0 in Ω , B u = 0 on Ω ,

where the concrete value of c{a,b,a+b} is determined by the values of α and β; the precise concept of metasolution has been discussed in the monograph [27]. Note that (1.4) is non-degenerate, unless c=a and Ω0, because we are assuming that b(x)>0 for all xΩ¯ and a+bb. Throughout this paper, for every c𝒞μ(Ω¯), c0, we will denote by Uc the unique positive solution of (1.4) if it exists. According to Fraile et al. [15], Uc exists if and only if λ>λ1B(Ω) unless c=a and Ω0. In the degenerate case when c=a and Ω0, problem (1.4) possesses a positive solution if and only if

(1.5) λ 1 B ( Ω ) < λ < λ 1 D ( Ω 0 ) .

Moreover, it is unique if it exists (see Fraile et al. [15] and Aleja, Antón and López-Gómez [1]). By adopting the notation λ1D(Ω0)=+ if Ω0=, (1.5) characterizes the existence of a positive solution in all these cases. This convention is consistent with the property that

lim | Ω ~ | 0 λ 1 D ( Ω ~ ) = + ,

where |Ω~| stands for the Lebesgue measure of Ω~, which can be easily derived from the iso-perimetric inequality of Faber [14] and Krahn [20] (see [22, Theorem 5.1]).

The dynamics of the positive solutions of the parabolic counterpart of (1.4) are governed by the maximal non-negative solution of (1.4) if c{b,a+b}, or c=a with Ω0 empty, but it is regulated by its minimal positive metasolution when c=a, Ω0 is non-empty and λλ1D(Ω0). Thus, if the coefficient a(x) has a spatial degeneracy in Ω0, i.e., Ω0, then the asymptotic behavior of the positive solutions to problem (1.1) can be quite different from the behavior of the classical perturbation problem when a(x) admits no spatial degeneracy.

The analysis of these degenerate problems goes back to Brézis and Oswald [6], by means of variational techniques, Ouyang [32], through global continuation, and Fraile et al. [15], where the method of sub and supersolutions was incorporated to the theory of singular problems, and the problem of the dynamics of the parabolic counterparts of these degenerate elliptic models was addressed by the first time. Then, based on [15], the theory advanced very quickly with García-Melián et al. [16], Gómez-Reñasco and López-Gómez [17], and Du and Huang [13] (see [27] for further information).

The spatial degeneracy makes a fundamental change on the behavior of the positive solutions of (1.4) when c=a and Ω0 (see, e.g., López-Gómez [23] and Du [12] for the simplest case when m1). Assume that m is sufficiently smooth in Ω0 with m(x)>0 for all xΩ0 and m0 in Ω¯0. Then, by a recent result of the authors [21, Theorem 1.1], Ua satisfies

lim λ λ 1 D ( Ω 0 ) U a = { + uniformly in Ω ¯ 0 , L min in Ω ¯ Ω ¯ 0 ,

where Lmin stands for the minimal positive solution of the singular problem

(1.6) { - Δ u = λ 1 D ( Ω 0 ) m ( x ) u - a ( x ) u p ( x ) in Ω Ω ¯ 0 , B u = 0 on Ω , u = + on Ω 0 .

The existence of Lmin for (1.6) was established by [21, Theorem 2.1]. The “blow-up” of Ua in Ω¯0 as λλ1D(Ω0) appears due to the spatial degeneracy of a(x) in Ω0.

On the other hand, if c=a0 in Ω¯, then (1.4) reduces to the linear eigenvalue problem (1.3). In this case, the sharp changes of the positive solutions between the nonlinear problem (1.4) with c=a and the linear problem (1.3) have been investigated by Sun in [33], where it was established that, whenever α0,

lim ε 0 U a ε = + locally uniformly in Ω .

These two basic results tell us that the “blow-up” phenomenon occurs when either a(x) exhibits some degeneration or the whole nonlinear problem degenerates to the linear one. At the end of the day, the second situation is a limiting case of the former one when Ω0Ω. Our interest in analyzing those sharp changes explains why we have chosen the perturbed model (1.1), where α0 is the quenching speed of the nonlinear function and β0 is the speed of the spatial degeneracy.

Problem (1.4) is a basic semilinear elliptic boundary value problem used in the study of a variety of phenomena in the applied sciences (see, e.g., [18, 32, 5, 6, 15, 17, 10, 21, 28, 3]). It is also the most paradigmatic model in population dynamics, the diffusive logistic model [4, 9, 31, 2, 11, 17, 13, 23, 29]. In this context, Ω is the region inhabited by the population with species density u, λm(x) measures the birth or death rates in a heterogenous environment, according to whether it is positive or negative, and c(x) measures the carrying capacity of Ω, i.e., the capacity of Ω to support the species u(x). From a biological point of view, the region {xΩ:m(x)>0} acts like a source for u, whereas {xΩ:m(x)<0} can be thought as a sink. The case when the growth function m(x) changes sign seems to be far more interesting from the point of view of the applications than the case when m(x) is everywhere positive, besides much harder to deal with from a mathematical point of view. Under the above assumptions, the semilinear problems (1.4) was well studied (see [15, 18, 31, 29, 1, 21] and the references therein).

We now deliver the main results of this paper. The first one provides us with the sharp profile of θε as ε0 when α=β0. It can be stated as follows.

Theorem 1.1.

Suppose (Hm), (Ha) and (Hb), with α=β0, and λ>λ1B(Ω). Then

(1.7) lim ε 0 + ( ε α p - 1 θ ε ) = U a + b uniformly in Ω ¯ .

In particular, limε0θε=Ua+b uniformly in Ω¯ if α=0, while limε0θε=+ locally uniformly in Ω if α>0, which covers the result of Sun [33], because limε0(aε+bε)=0 in Ω if α=β>0. Moreover, by (1.7), when α=β>0, the positive solutions of (1.1) have a blow-up rate ε-αp-1 because θεε-αp-1Ua+b as ε0.

Our second result provides us with the sharp profile of θε as ε0 when α>β0.

Theorem 1.2.

Suppose (Hm), (Ha) and (Hb), with α>β0, and λ>λ1B(Ω). Then

(1.8) lim ε 0 + ( ε β p - 1 θ ε ) = U b uniformly in Ω ¯ .

In particular, limε0θε=Ub uniformly in Ω¯ if β=0, while limε0θε=+ locally uniformly in Ω if β>0. Moreover, when α>β>0, the positive solutions of (1.1) have a blow-up rate ε-βp-1 because, due to (1.8),

θ ε ε - β p - 1 U b as ε 0 .

The most intriguing case arises when β>α0, where the asymptotic profiles of θε as ε0 make a sharp change with respect to the previous ones. Our main result in this case can be stated as follows. We are denoting by r:{0} the function defined by r(1)=0, r(2)=r(3)=r(4)=r(5)=1, and, for every n2,

r ( 4 n - 2 ) = r ( 4 n - 1 ) = r ( 4 n ) = r ( 4 n + 1 ) = r ( 4 ( n - 1 ) ) + 2 .

It was introduced by the authors in [21].

Theorem 1.3.

Suppose (Hm), (Ha) and (Hb), with β>α0, m0 in Ω0, m(x)>0 for all xΩ0, and mCr(N)(Ω0), where N is the spatial dimension. Then

  1. when λ 1 B ( Ω ) < λ < λ 1 D ( Ω 0 ) ,

    (1.9) lim ε 0 ( ε α p - 1 θ ε ) = U a uniformly in Ω ¯ .

  2. When λ λ 1 D ( Ω 0 ) ,

    (1.10) lim ε 0 ( ε α p - 1 θ ε ) = + uniformly in Ω ¯ 0 .

    Moreover, for every sequence { ε n } n 1 , with lim n ε n = 0 , there exists a subsequence of { θ ε n } n 1 , relabeled in the same way, such that

    (1.11) lim n ( ε n α p - 1 θ ε n ) = L 𝑖𝑛 Ω ¯ Ω ¯ 0 ,

    where L is a positive solution of the singular problem

    (1.12) { - Δ u = λ m u - a u p 𝑖𝑛 Ω Ω ¯ 0 , B u = 0 𝑜𝑛 Ω , u = + 𝑜𝑛 Ω 0 .

    Furthermore, if, in addition,

    (1.13) lim ε 0 a ε ε β = 0 uniformly in Ω ¯ 0 ,

    a ε + b ε ε α a for sufficiently small ε > 0 , and the minimal positive solution L min of problem ( 1.12 ) satisfies L min L 2 p ( Ω Ω ¯ 0 ) , then L = L min and

    (1.14) lim ε 0 ( ε β p - 1 θ ε ) = U b , Ω 0 uniformly in Ω ¯ 0 ,

    where U b , Ω 0 stands for the unique positive solution of

    (1.15) { - Δ u = λ m u - b u p 𝑖𝑛 Ω 0 , u = 0 𝑜𝑛 Ω 0 .

Note that (1.14) provides us with the exact blow-up rate of θε in Ω0 as ε0. More precisely, according to (1.11),

θ ε ε - α p - 1 L min in Ω ¯ Ω ¯ 0 as ε 0 ,

whereas, thanks to (1.14),

θ ε ε - β p - 1 U b , Ω 0 in Ω ¯ 0 as ε 0 .

Thus, θε has two different blow-up rates in Ω as ε0 if β>α>0.

Unless the singular problem (1.12) admits a unique positive solution, or aε+bεεαa for sufficiently small ε>0, one cannot guarantee that, along two different subsequences of θε, the same positive solution of (1.12) will be approximated. In general, the limit might depend on the decay of aε+bε to 0 as ε0. Some very general conditions on a(x) near Ω0 so that (1.12) can admit a unique large positive solution were given in [24, 25, 30] as well as in the references therein. According to [24, Theorem 1.1], this occurs, e.g., when there are two continuous functions E,C:Ω0(0,) such that, for every x0Ω0,

(1.16) lim x Ω Ω ¯ 0 x x 0 a ( x ) [ dist ( x , Ω 0 ) ] E ( x 0 ) = C ( x 0 ) ,

though a(x) can decay to zero on Ω0 with rather general rates (see [25] for further information). Moreover, under condition (1.16), the unique positive solution L of (1.12) satisfies, for every x0Ω0,

L ( x ) [ 1 C ( x 0 ) E ( x 0 ) + 2 p - 1 ( E ( x 0 ) + 2 p - 1 + 1 ) ] 1 p - 1 [ dist ( x , Ω 0 ) ] - E ( x 0 ) + 2 p - 1

when xΩΩ¯0 approximates x0. Thus, in the special but important case when the decay rate of a(x) to zero on Ω0, measured by E, is constant on Ω0, i.e., there exists a constant q>0 such that E(x0)=q for all x0Ω0, then LL2p(ΩΩ¯0) if and only if

(1.17) Ω Ω ¯ 0 [ dist ( x , Ω 0 ) ] - 2 p q + 2 p - 1 𝑑 x < + .

It is folklore that (1.17) holds true if and only if N>2pq+2p-1. Therefore, in this case, all the assumptions of the last part of Theorem 1.3 are satisfied for sufficiently large N1. The challenge of analyzing whether or not the assumption LL2p(ΩΩ¯0) can be removed from the statement of Theorem 1.3 remains an open problem in this paper.

Finally, note that, as soon as β>α>0, one has that

a ε ε β = ε α - β a ε ε α

with α-β<0. Thus, (Ha) and (1.13) entail that

lim ε 0 a ε ε α = a 0 in Ω ¯ 0

at a rate higher than εβ-α.

The rest of this paper is organized as follows. Section 2 delivers the proof of a general comparison result, Lemma 2.1, which is used throughout this paper, Section 3 gives the proofs of Theorems 1.1 and 1.2, and Section 4 consists of the proof of Theorem 1.3.

2 A Useful Comparison Result

This section studies an extremely useful comparison result, whose proof will be delivered by the sake of completeness.

Lemma 2.1.

Let c0 be a function of class Cμ(Ω¯) such that problem (1.4) possesses a positive solution Uc, necessarily unique by, e.g., [1, Theorem 3.1]. Then, for every positive strict subsolution u¯ of (1.4) such that u¯(x)>0 for all xΩ, one has that Uc-u¯0 in the sense that Uc(x)-u¯(x)>0 for all xΩ, and ν(Uc-u¯)(x)<0 for all xΩ such that u¯(x)=Ua(x). Similarly, for every positive strict supersolution u¯ of (1.4) such that u¯(x)>0 for all xΩ, one has that u¯-Uc0.

Proof.

Since Uc is a positive solution of (1.4), we have that

{ ( - Δ + c U c p - 1 - λ m ) U c = 0 in Ω , B u = 0 on Ω .

Thus, according to [26, Theorem 7.7],

(2.1) σ 1 [ - Δ + c U c p - 1 - λ m ; B , Ω ] = 0 .

Let u¯ be a positive strict subsolution of (1.4). By definition,

{ - Δ u ¯ λ m ( x ) u ¯ - c ( x ) u ¯ p in Ω , B u ¯ 0 on Ω ,

with some of these inequalities strict. Thus,

(2.2) B ( U c - u ¯ ) 0 on Ω ,

and

- Δ ( U c - u ¯ ) λ m ( U c - u ¯ ) - c ( U c p - u ¯ p ) = λ m ( U c - u ¯ ) - p c 0 1 ( t U c + ( 1 - t ) u ¯ ) p - 1 𝑑 t ( U c - u ¯ ) ,

with some of these inequalities strict. Equivalently,

(2.3) ( - Δ + V - λ m ) ( U c - u ¯ ) 0 in Ω ,

where

V p c 0 1 ( t U c + ( 1 - t ) u ¯ ) p - 1 𝑑 t .

Since u¯(x)>0 for all xΩ and c0, it becomes apparent that

p 0 1 ( t U c ( x ) + ( 1 - t ) u ¯ ( x ) ) p - 1 𝑑 t > U c p - 1 ( x ) for all x Ω ,

and hence, VcUcp-1 in Ω. Consequently, by [8, Proposition 3.3], we find that

σ 1 [ - Δ + V - λ m ; B , Ω ] > σ 1 [ - Δ + c U c p - 1 - λ m ; B , Ω ] = 0 .

Therefore, since either (2.2) or (2.3) is strict, we can conclude from [26, Theorem 7.10] that Ua-u¯0.

The second assertion can be easily accomplished by adapting the previous argument. So we will omit any further technical details herein. ∎

3 Proof of Theorems 1.1 and 1.2

The proof of these theorems relies on the next (regular) perturbation result.

Theorem 3.1.

Suppose (Hm) and cεCμ(Ω¯), ε[0,1], satisfies c0(x)>0 for all xΩ¯mega and

(3.1) lim ε 0 c ε = c 0 uniformly in Ω ¯ .

Then

(3.2) lim ε 0 U c ε = U c 0 uniformly in Ω ¯ .

Proof.

By (3.1), for every η>0, there exists ε0=ε0(η) such that

c 0 - η c ε c 0 + η for all ε ( 0 , ε 0 ) .

Hence, for every ε(0,ε0), we have that

λ m U c ε - ( c 0 + η ) U c ε p λ m U c ε - c ε U c ε p = - Δ U c ε λ m U c ε - ( c 0 - η ) U c ε p .

Consequently, for every ε(0,ε0), Ucε is a subsolution of

{ - Δ u = λ m u - ( c 0 - η ) u p in Ω , B u = 0 on Ω ,

and, simultaneously, it is a supersolution of

(3.3) { - Δ u = λ m u - ( c 0 + η ) u p in Ω , B u = 0 on Ω .

Since c0(x)>0 for all xΩ¯, η>0 can be shortened, if necessary, so that c0(x)-η>0 for all xΩ¯. So the positive solutions Uc0±η are well defined because we are assuming that λ>λ1B(Ω). Moreover, thanks to Lemma 2.1, they satisfy

U c 0 + η U c ε U c 0 - η in Ω for all ε ( 0 , ε 0 ) .

Therefore, (3.2) holds true as soon as

(3.4) lim η 0 U c 0 ± η = U c 0 uniformly in Ω ¯ ,

whose validity can be easily shown, e.g., by applying the implicit function theorem to the pair (η,u)=(0,Uc0) viewed as a solution pair of, e.g., (3.3). Indeed, the solutions of (3.3) are given, e.g., by the zeroes of the operator 𝔉:×𝒞B1+μ(Ω¯)𝒞B1+μ(Ω¯) defined by

𝔉 ( ξ , u ) = u - ( - Δ + 1 ) - 1 [ ( λ m + 1 ) u - ( c 0 + ξ ) u p ] , ( ξ , u ) × 𝒞 B 1 + μ ( Ω ¯ ) ,

where (-Δ+1)-1 stands for the resolvent operator of -Δ+1 in Ω subject to the boundary operator B, and

𝒞 B 1 + μ ( Ω ¯ ) { u 𝒞 1 + μ ( Ω ¯ ) : B u = 0 on Ω } .

By definition, 𝔉(0,Uc0)=0. Moreover, since p>1, 𝔉 is of class 𝒞1 and

D u 𝔉 ( 0 , U c 0 ) u = u - ( - Δ + 1 ) - 1 [ ( λ m + 1 ) u - p c 0 U c 0 p - 1 u ]

for all u𝒞B1+μ(Ω¯). Since it is a compact perturbation of the identity map, Du𝔉(0,Uc0) is a Fredholm operator of index zero. Thus, it is an isomorphism if kerDu𝔉(0,Uc0)=[0]. On the contrary, assume that Du𝔉(0,Uc0)u=0 for some u𝒞B1+μ(Ω¯), u0. Then, by elliptic regularity, u𝒞B2+μ(Ω¯) and

( - Δ + p c 0 U c 0 p - 1 - λ m ) u = 0 in Ω .

Thus, σ1[-Δ+pc0Uc0p-1-λm;B,Ω]0. Hence, by [8, Proposition 3.3],

0 σ 1 [ - Δ + p c 0 U c 0 p - 1 - λ m ; B , Ω ] > σ 1 [ - Δ + c 0 U c 0 p - 1 - λ m ; B , Ω ] ,

which contradicts (2.1) and shows that Du𝔉(0,Uc0) indeed establishes an isomorphism. Therefore, by the implicit function theorem, there exist η0>0 and a map of class 𝒞1, u:(-η0,η0)𝒞B1+μ(Ω¯), such that u(0)=Uc0 and 𝔉(η,u(η))=0 for all η(-η0,η0). Since Uc00, it is apparent that u(η)0 for sufficiently small η. Consequently, by the uniqueness of the positive solution, we find that u(η)=Uc0+η. By the continuity of u(η), this shows (3.4) and ends the proof. ∎

3.1 Proof of Theorem 1.1

Suppose that α=β=0. Then limε0(aε+bε)=a+b uniformly in Ω¯, and hence, Theorem 1.1 is a direct consequence of Theorem 3.1.

Suppose that α=β>0. Then the positive function ωεεαp-1θε solves the problem

{ - Δ v = λ m ( x ) v - ( a ε ε α + b ε ε α ) v p in Ω , B v = 0 on Ω .

Thus, since

lim ε 0 ( a ε ε α + b ε ε α ) = a + b uniformly in Ω ¯ ,

we can infer from Theorem 3.1 that limε0ωε=Ua+b. The proof is complete.

3.2 Proof of Theorem 1.2

Suppose that α>β=0. Then limε0(aε+bε)=b uniformly in Ω¯, and, also in this case, Theorem 1.2 follows as a direct consequence of Theorem 3.1.

Suppose that α>β>0. Then the positive function ωεεβp-1θε solves the problem

{ - Δ v = λ m ( x ) v - ( a ε ε β + b ε ε β ) v p in Ω , B v = 0 on Ω .

Since α>β, it follows from (Ha) and (Hb) that

lim ε 0 ( a ε ε β + b ε ε β ) = lim ε 0 ( ε α - β a ε ε α + b ε ε β ) = b uniformly in Ω ¯ .

Therefore, also by Theorem 3.1, limε0ωε=Ub, which ends the proof of Theorem 1.2.

4 Proof of Theorem 1.3

As Theorem 1.3 is a very deep result, its proof is substantially more sophisticated than the proofs of Theorems 1.1 and 1.2. We will divide it into four steps.

Step 1: Proof of (1.9) for β>α=0 and λ1B(Ω)<λ<λ1D(Ω0). Suppose β>α=0. Then, by (Ha) and (Hb),

(4.1) lim ε 0 ( a ε + b ε ) = a uniformly in Ω ¯ .

Suppose, in addition, that λ1B(Ω)<λ<λ1D(Ω0). By [21, Theorem 1.2], the positive solution Ua is well defined and it is unique. Note that, in case α=0, (1.9) simplifies to

(4.2) lim ε 0 θ ε = U a uniformly in Ω ¯ .

By (4.1), we obtain that, for sufficiently small ε>0, aεaε+bεa(x)+1 for all xΩ¯. Subsequently, we will assume that ε has been chosen in this way. Then

λ m θ ε - ( a + 1 ) θ ε p λ m θ ε - ( a ε + b ε ) θ ε p = - Δ θ ε λ m θ ε - a ε θ ε p

and hence, by Lemma 2.1,

(4.3) U a + 1 θ ε U a ε in Ω .

Subsequently, for sufficiently small δ>0, we consider the open subset of Ω defined by

Ω δ { x Ω : dist ( x , Ω 0 ) < δ } .

By construction, Ω¯0Ωδ and, since m0 in Ωδ for sufficiently small δ>0, using the general theory of [26, Chapter 9], it is apparent that δ can be shortened so that

(4.4) λ < λ 1 D ( Ω δ ) < λ 1 D ( Ω 0 ) .

Pick any function a~δ𝒞μ(Ω¯) such that a~δ-1(0)=Ω¯δ and a~δ(x)<a(x) for all xΩ¯Ω¯0. Then, since aεa as ε0 uniformly in Ω¯ and aε(x)>0 for all xΩ¯, one has that a~δaε for sufficiently small ε>0. Thus,

- Δ U a ε = λ m U a ε - a ε U a ε p λ m U a ε - a ~ δ U a ε p

and hence, by Lemma 2.1, we find that UaεUa~δ for sufficiently small ε>0. Therefore, by (4.3), we can infer that

(4.5) U a + 1 θ ε U a ~ δ in Ω .

For sufficiently small ε>0, εε satisfies the integral equation

(4.6) θ ε = ( - Δ + 1 ) - 1 [ ( λ m + 1 ) θ ε - ( a ε + b ε ) θ ε p ] = ( - Δ + 1 ) - 1 [ ( λ m + 1 ) θ ε - a θ ε p ] + ( - Δ + 1 ) - 1 [ ( a - a ε - b ε ) θ ε p ] .

Moreover, by (4.1), we have that

lim ε 0 ( - Δ + 1 ) - 1 [ ( a - a ε - b ε ) θ ε p ] = 0 in 𝒞 2 + μ ( Ω ¯ ) .

Thus, by the compactness of the resolvent operator, it follows from (4.5) that there exists a sequence {εn}n1, with limnεn=0, such that limnθεn=θ0𝒞2+μ(Ω¯). Particularizing (4.6) at ε=εn and letting n yields to θ0=(-Δ+1)-1[(λm+1)θ0-aθ0p], i.e., θ0 solves the problem

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

Moreover, by (4.5), θ00. Therefore, by uniqueness, θ0=Ua. As this argument can be repeated along any sequence of ε’s converging to zero, (4.2) holds.

Step 2: Proof of (1.10) and (1.11) for β>α=0 and λλ1D(Ω0). Assume that β>α=0 and λλ1D(Ω0). As in Step 1, (4.1) holds. Pick a sufficiently small δ>0 such that m0 in Ωδ. Since θε(x)>0 for all xΩδΩ, θε provides us with a positive strict supersolution of the boundary value problem

(4.7) { - Δ u = λ m ( x ) u - ( a ε + b ε ) u p in Ω δ , u = 0 on Ω δ .

According to (4.4), we have that λλ1D(Ω0)>λ1D(Ωδ). Moreover, aε(x)+bε(x)>0 for all xΩ¯δ. Thus, (4.7) admits a unique positive solution, denoted by θ[λ,ε,δ] (see, e.g., [1] if necessary). By Lemma 2.1,

(4.8) θ [ λ , ε , δ ] θ ε in Ω δ .

Similarly, since m0 in Ωδ, for sufficiently small η>0, the function θ[λ,ε,δ] provides us with a positive supersolution of

(4.9) { - Δ u = ( λ 1 D ( Ω 0 ) - η ) m ( x ) u - ( a ε + b ε ) u p in Ω δ , u = 0 on Ω δ .

Hence, again by Lemma 2.1, it follows from (4.8) that

(4.10) θ [ λ 1 D ( Ω 0 ) - η , ε , δ ] θ [ λ , ε , δ ] θ ε in Ω δ .

As for sufficiently small η>0 we have that λ1B(Ω)<λ1D(Ω0)-η<λ1D(Ω0), applying the result already proven in Step 1 to problem (4.9), it is apparent that

lim ε 0 θ [ λ 1 D ( Ω 0 ) - η , ε , δ ] = θ [ λ 1 D ( Ω 0 ) - η , a , δ ] ,

where θ[λ1D(Ω0)-η,a,δ] denotes the unique positive solution of

{ - Δ u = ( λ 1 D ( Ω 0 ) - η ) m ( x ) u - a u p in Ω δ , u = 0 on Ω δ .

Consequently, letting ε0 in (4.10), we obtain that, for sufficiently small η,

(4.11) θ [ λ 1 D ( Ω 0 ) - η , a , δ ] lim inf ε 0 θ ε in Ω δ .

Since under the general assumptions of Theorem 3.1, [21, Theorem 1.1] implies that

(4.12) lim η 0 θ [ λ 1 D ( Ω 0 ) - η , a , δ ] = + uniformly in Ω ¯ 0 = a - 1 ( 0 ) ,

letting η0 in (4.11), we can conclude from (4.12) that lim infε0θε=+ uniformly in Ω¯0. This ends the proof of (1.10) for α=0.

It remains to prove the second assertion of part (b) for α=0. By (Ha) and (Hb), there exists δ0>0 such that, for every δ(0,δ0), there are ε0=ε0(δ)>0 and η0=η0(δ)>0 such that

a ε + b ε a - η min Ω ¯ Ω δ a - η > 0 in D δ Ω Ω ¯ δ

for all ε(0,ε0) and η(0,η0). Thus, thanks to Lemma 2.1,

(4.13) θ ε L a - η , D δ min in D δ Ω Ω ¯ δ for every ε ( 0 , ε 0 ) ,

where La-η,Dδmin stands for the minimal positive solution of the singular problem

{ - Δ u = λ m u - ( a - η ) u p ( x ) in D δ , B u = 0 on Ω , u = + on Ω δ .

Observe that Dδ=ΩΩδ. Now, we consider a decreasing sequence of δ’s, {δn}n1, with δ1<δ0, such that limnδn=0. Then, for every n1,

D δ n D δ n + 1 Ω Ω ¯ 0 , Ω Ω ¯ 0 = m n D δ m .

Moreover, for every n1, ε(0,ε0(δn)) and η(0,η0(δn)), it follows from (4.13) that

(4.14) θ ε L a - η , D δ n min in D δ n Ω Ω ¯ δ n .

Pick a sequence {εm}m1 of ε’s such that limmεm=0, as well as an integer n1. Then, thanks to (4.14), for sufficiently large m1, say mm0(n), we have that θεmLa-η0(δn+2)/2,Dδn+2min in Dδn+2. In particular, there exists a constant Cn+1>0 such that

θ ε m L a - η 0 ( δ n + 2 ) / 2 , D δ n + 2 min C n + 1 in D δ n + 1 .

Thus, by the global Schauder estimates, there exists a constant C~n such that

θ ε m 𝒞 2 + μ ( D ¯ δ n ) C ~ n for all m m 0 ( n ) .

Consequently, as the injection 𝒞2+μ(D¯δn)𝒞2(D¯δn) is compact, there exists a subsequence of θεm, denoted by {θm,n}m1, and a function Θn𝒞2(D¯δn) such that

lim m θ m , n = Θ n in 𝒞 2 ( D ¯ δ n ) .

As this scheme can be repeated for every integer n1, we can extract a subsequence of θεm, denoted by {θm,1}m1, such that

lim m θ m , 1 = Θ 1 in 𝒞 2 ( D ¯ δ 1 ) ,

and a subsequence of {θm,1}m1, denoted by {θm,2}m1, such that

lim m θ m , 2 = Θ 2 in 𝒞 2 ( D ¯ δ 2 ) ,

and, more generally, for every n2, a subsequence of {θm,n-1}m1, denoted by {θm,n}m1, such that

lim m θ m , n = Θ n in 𝒞 2 ( D ¯ δ n ) ,

where Θn𝒞2(D¯δn) for all n1. By the uniqueness of the limits,

Θ n + 1 | D ¯ δ n = Θ n for all n 1 .

So Θn+1 provides us with an extension of Θn to D¯δn+1 for all n1. Moreover, by (4.1), letting m in the differential equation of θm,n, it becomes apparent that, for every n1, the function Θn solves the problem

{ - Δ u = λ m ( x ) u - a u p ( x ) in D δ n , B u = 0 on Ω .

Furthermore, since limnδn=0, it follows from (4.10) that the diagonal subsequence, {θn,n}n1, converges to a large solution of the singular problem (1.12). This ends the proof of (1.11) for α=0.

Step 3: Proof of (1.9)–(1.11) for β>α>0. Suppose β>α>0. In this case, the function ωεεαp-1θε solves the problem

{ - Δ w = λ m ( x ) w - ( a ε ε α + b ε ε α ) w p in Ω , B w = 0 on Ω .

Since β>α, (Ha) and (Hb) imply that

lim ε 0 ( a ε ε α + b ε ε α ) = a + lim ε 0 ( ε β - α b ε ε β ) = a uniformly in Ω ¯ .

Therefore, (1.9), (1.10) and (1.11) follow easily by applying Steps 1 and 2 to ωε.

Step 4: Proof of (1.14) for β>α0. Set ϑεεβ-αp-1ωε=εβp-1θε. Then ϑε satisfies

(4.15) { - Δ ϑ ε = λ m ϑ ε - ( a ε ε β + b ε ε β ) ϑ ε p in Ω , B ϑ ε = 0 on Ω .

According to (Ha), (Hb) and (1.13), we have that

lim ε 0 ( a ε ε β + b ε ε β ) = lim ε 0 ( ε α - β a ε ε α ) + b = { + in Ω Ω ¯ 0 , b in Ω ¯ 0 .

Set η:=12minΩ¯b(0,minΩ¯b), and let ε0(0,1) be such that

b ε ε β b - η for all ε ( 0 , ε 0 ) .

Then

c ε a ε ε β + b ε ε β b - η for all ε ( 0 , ε 0 ) ,

and it follows from (4.15) that

- Δ ϑ ε = λ m ϑ ε - c ε ϑ ε p λ m ϑ ε - ( b - η ) ϑ ε p .

Thus, by Lemma 2.1, it is apparent that

(4.16) ϑ ε U b - η in Ω for all ε ( 0 , ε 0 ) .

In particular, {ϑε}ε(0,ε0) is bounded in L(Ω). Moreover, multiplying the differential equation of (4.15) by ϑε and integrating by parts in Ω yields to

(4.17) Ω | ϑ ε | 2 𝑑 x + Γ 1 γ ϑ ε 2 𝑑 S + Ω c ε ϑ ε p + 1 𝑑 x = λ Ω m ϑ ε 2 𝑑 x .

Thus, since γ0 on Γ1 and cε0 in Ω, it follows from (4.17) that {ϑε}ε(0,ε0) is bounded in W1,2(Ω). As the injection W1,2(Ω)L2(Ω) is compact, for any given sequence εn(0,ε0), n1, such that limnεn=0, one can extract a subsequence, relabeled by n1, such that, for some ϑ0L2(Ω), limnϑεn=ϑ0 in L2(Ω). Moreover, by (4.16) and (4.17), we also find that

Ω Ω ¯ 0 a ε n ε n β ϑ ε n p + 1 𝑑 x Ω c ε n ϑ ε n p + 1 𝑑 x λ Ω m ϑ ε 2 𝑑 x λ max Ω ¯ m Ω U b - η 2 𝑑 x .

Hence, there is a constant C>0 such that

(4.18) Ω Ω ¯ 0 a ε n ε n β ϑ ε n p + 1 𝑑 x C for all n 1 .

Consequently, since β>0 and

lim n a ε n ε n β = + pointwise in Ω Ω ¯ 0 ,

it follows from (4.18) that

lim n ϑ ε n = 0 almost everywhere in Ω Ω ¯ 0 .

Thus, by the uniqueness of the limit, ϑ0=0 in ΩΩ¯0. To complete the proof, it remains to show that ϑ0=Ub,Ω0 in Ω0, where Ub,Ω0 stands for the unique positive solution of (1.15). This can be proven with the following argument.

Note that, by the definition of ϑεn, we have that

(4.19) c ε n ϑ ε n p = c ε n ε n ( β - α ) p p - 1 ω ε n p = c ε n ε n β - α + β - α p - 1 ω ε n p = a ε n + b ε n ε n α ε n β - α p - 1 ω ε n p .

Thus, since

lim n a ε n + b ε n ε n α = a + lim n ( b ε n ε n β ε n β - α ) = a

uniformly in Ω¯ and, owing to (1.11), we already know that limnωεn=L in Ω¯Ω¯0, it becomes apparent that

c ε n ϑ ε n p a L p ε n β - α p - 1 as n .

In particular,

lim n ( c ε n ϑ ε n p ) = 0 uniformly in compact subsets of ( Ω Ω ) Ω ¯ 0 .

This nice convergence is lost at Ω0 because L= on Ω0. This technical difficulty is overcome with the integrability condition LminL2p(ΩΩ0) and the assumption that

(4.20) a ε + b ε ε α a for sufficiently small ε > 0 .

Next, we will show that {ϑεn}n1 is actually a Cauchy sequence in W1,2(Ω). This entails that ϑ0W1,2(Ω) and that limnϑεn=ϑ0 in W1,2(Ω). Actually, since ϑ0=0 in ΩΩ¯0, it becomes apparent that ϑ0W01,2(ΩΩ¯δ) for sufficiently small δ, say δ<δ0, where we are denoting Ωδ{xΩ:dist(x,Ω0)<δ}. Thus, since Ω0 is stable, as discussed, e.g., in [8], we find that

ϑ 0 δ ( 0 , δ 0 ) W 0 1 , 2 ( Ω Ω ¯ δ ) = W 0 1 , 2 ( Ω Ω ¯ 0 ) .

To prove that {ϑεn}n1 is a Cauchy sequence in W1,2(Ω), we argue as usual. By some direct straightforward manipulations from (4.15), it is easily seen that, for every pair of integers, n1 and 1,

Ω | ϑ ε n - ϑ ε | 2 𝑑 x λ Ω m ( ϑ ε - ϑ ε n ) 2 𝑑 x + Ω c ε ϑ ε p | ϑ ε - ϑ ε n | 𝑑 x + Ω c ε n ϑ ε n p | ϑ ε n - ϑ ε | 𝑑 x - Γ 1 γ ( ϑ ε - ϑ ε n ) 2 𝑑 σ .

Thus, since γ0 on Γ1, we find that

(4.21) Ω | ϑ ε n - ϑ ε | 2 𝑑 x λ max Ω ¯ m Ω ( ϑ ε - ϑ ε n ) 2 𝑑 x + Ω c ε ϑ ε p | ϑ ε - ϑ ε n | 𝑑 x + Ω c ε n ϑ ε n p | ϑ ε n - ϑ ε | 𝑑 x .

To estimate the right-hand side of (4.21), we proceed as follows. By (4.20), there exists ε1(0,ε0) such that, for every ε(0,ε1), (4.20) implies that

- Δ ω ε = λ m ω ε - ( a ε ε α + b ε ε α ) ω ε p λ m ω ε - a ω ε p

for all ε(0,ε1). Thus, by Lemma 2.1, it becomes apparent that

(4.22) ω ε L min in Ω Ω ¯ 0

for all ε(0,ε1). In particular, by (1.11), we find that L=Lmin because LminL by definition of minimal solution. By the Hölder inequality,

(4.23) Ω c ε n ϑ ε n p | ϑ ε n - ϑ ε | 𝑑 x ( Ω c ε n 2 ϑ ε n 2 p 𝑑 x ) 1 2 ϑ ε n - ϑ ε L 2 ( Ω ) .

Moreover,

Ω c ε n 2 ϑ ε n 2 p 𝑑 x = Ω Ω ¯ 0 c ε n 2 ϑ ε n 2 p 𝑑 x + Ω 0 c ε n 2 ϑ ε n 2 p 𝑑 x .

According to (4.19) and (4.22), there exists a constant C>0 such that

Ω Ω ¯ 0 c ε n 2 ϑ ε n 2 p 𝑑 x = ε n 2 ( β - α ) p - 1 Ω Ω ¯ 0 a ε n + b ε n ε n α ω ε n 2 p 𝑑 x ε n 2 ( β - α ) p - 1 C Ω Ω ¯ 0 ( L min ) 2 p 𝑑 x .

Hence, since LminL2p(ΩΩ0), we find that

lim n Ω Ω ¯ 0 c ε n 2 ϑ ε n 2 p 𝑑 x = 0 .

On the other hand, as, due to (1.13),

lim ε 0 c ε = b uniformly in Ω ¯ 0 ,

by (4.16), we find that, for sufficiently large n1,

Ω 0 c ε n 2 ϑ ε n 2 p 𝑑 x Ω 0 ( b + η ) 2 U b - η 2 p 𝑑 x C

for some constant C>0. Therefore, by (4.23), there exists a constant C~>0 such that

Ω c ε n ϑ ε n p | ϑ ε n - ϑ ε | 𝑑 x C ~ ϑ ε n - ϑ ε L 2 ( Ω ) .

By symmetry,

Ω c ε ϑ ε p | ϑ ε n - ϑ ε | 𝑑 x C ~ ϑ ε n - ϑ ε L 2 ( Ω ) .

From these estimates, it readily follows that ϑεn is a Cauchy sequence in W1,2(Ω0). The fact that ϑ0 provides us with a weak non-negative solution of (1.15) is routine from the weak formulation of the ϑεn-problem by letting n. To show that ϑ00, we can use the following argument. Since, for every n1, ϑεn(x)>0 for all xΩ0, it becomes apparent that ϑεn is a positive supersolution of the problem

(4.24) { - Δ v = λ m v - ( a ε n ε n β + b ε n ε n β ) v p in Ω 0 , v = 0 on Ω 0 .

Thus, by Lemma 2.1, we have that, for every n1,

(4.25) ϑ ε n v ε n in Ω 0 ,

where vεn denotes the unique positive solution of (4.24). On the other hand, by our assumptions of the weight functions aε and bε and, in particular, by (1.13) and (Hb), we have that

lim ε 0 ( a ε ε β + b ε ε β ) = b uniformly in Ω ¯ 0 .

Consequently, according to Theorem 1.1, it is apparent that

lim n v ε n = U b , Ω 0 ,

where Ub,Ω0 stands for the (unique) positive solution of (1.15). Therefore, letting n in (4.25) yields to ϑ0Ub,Ω0. Actually, by the uniqueness of the weak positive solution, we have that ϑ0=Ub,Ω0. As this argument can be repeated along any sequence {εn}n1, the proof of Theorem 1.3 is completed.


Communicated by Guozhen Lu


Award Identifier / Grant number: 11731005

Award Identifier / Grant number: 11401277

Award Identifier / Grant number: PGC2018-097104-B-I00

Award Identifier / Grant number: lzujbky-2021-52

Funding statement: W. T. Li was partially supported by NSF of China (11731005), J. López-Gómez was partially supported by the IMI of Complutense University, and the Ministry of Science, Innovation and Universities of Spain under Grant PGC2018-097104-B-I00, and J. W. Sun was partially supported by the NSF of China (11401277) and the Fundamental Research Funds for the Central Universities (lzujbky-2021-52).

References

[1] D. Aleja, I. Antón and J. López-Gómez, Solution components in a degenerate weighted BVP, Nonlinear Anal. 192 (2020), Article ID 111690. 10.1016/j.na.2019.111690Search in Google Scholar

[2] D. Aleja and J. López-Gómez, Dynamics of a class of advective-diffusive equations in ecology, Adv. Nonlinear Stud. 15 (2015), no. 3, 557–585. 10.1515/ans-2015-0303Search in Google Scholar

[3] H. Amann and J. López-Gómez, A priori bounds and multiple solutions for superlinear indefinite elliptic problems, J. Differential Equations 146 (1998), no. 2, 336–374. 10.1006/jdeq.1998.3440Search in Google Scholar

[4] F. Belgacem and C. Cosner, The effects of dispersal along environmental gradients on the dynamics of populations in heterogeneous environments, Canad. Appl. Math. Quart. 3 (1995), no. 4, 379–397. Search in Google Scholar

[5] H. Berestycki, Le nombre de solutions de certains problèmes semi-linéaires elliptiques, J. Funct. Anal. 40 (1981), no. 1, 1–29. 10.1016/0022-1236(81)90069-0Search in Google Scholar

[6] H. Brezis and L. Oswald, Remarks on sublinear elliptic equations, Nonlinear Anal. 10 (1986), no. 1, 55–64. 10.1016/0362-546X(86)90011-8Search in Google Scholar

[7] K. J. Brown and S. S. Lin, On the existence of positive eigenfunctions for an eigenvalue problem with indefinite weight function, J. Math. Anal. Appl. 75 (1980), no. 1, 112–120. 10.1016/0022-247X(80)90309-1Search 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), no. 1, 123–211. 10.1006/jdeq.2000.4003Search in Google Scholar

[9] R. S. Cantrell and C. Cosner, Spatial Ecology via Reaction-Diffusion Equations, John Wiley & Sons, Chichester, 2003. 10.1002/0470871296Search in Google Scholar

[10] E. N. Dancer and J. López-Gómez, Semiclassical analysis of general second order elliptic operators on bounded domains, Trans. Amer. Math. Soc. 352 (2000), no. 8, 3723–3742. 10.1090/S0002-9947-00-02534-4Search in Google Scholar

[11] 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

[12] Y. Du, Spatial patterns for population models in a heterogeneous environment, Taiwanese J. Math. 8 (2004), no. 2, 155–182. 10.11650/twjm/1500407619Search in Google Scholar

[13] Y. Du and Q. Huang, Blow-up solutions for a class of semilinear elliptic and parabolic equations, SIAM J. Math. Anal. 31 (1999), no. 1, 1–18. 10.1137/S0036141099352844Search in Google Scholar

[14] C. Faber, Beweis das unter allen homogenen Membranen von gleicher Fläche und gleicher Spannung die kreisförmige den tiefsten Grundton gibt, Sitzungsber. Bayer. Akad. der Wiss. Math. Phys. (1923), 169–172. Search in Google Scholar

[15] 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

[16] J. García-Melián, R. Gómez-Reñasco, J. López-Gómez and J. C. Sabina de Lis, Pointwise growth and uniqueness of positive solutions for a class of sublinear elliptic problems where bifurcation from infinity occurs, Arch. Ration. Mech. Anal. 145 (1998), no. 3, 261–289. 10.1007/s002050050130Search in Google Scholar

[17] R. Gómez-Reñasco and J. López-Gómez, On the existence and numerical computation of classical and non-classical solutions for a family of elliptic boundary value problems, Nonlinear Anal. 48 (2002), no. 4, 567–605. 10.1016/S0362-546X(00)00208-XSearch in Google Scholar

[18] D. Henry, Geometric Theory of Semilinear Parabolic Equations, Lecture Notes in Math.840, Springer, Berlin, 1981. 10.1007/BFb0089647Search in Google Scholar

[19] P. Hess and T. Kato, On some linear and nonlinear eigenvalue problems with an indefinite weight function, Comm. Partial Differential Equations 5 (1980), no. 10, 999–1030. 10.1080/03605308008820162Search in Google Scholar

[20] E. Krahn, Über eine von Rayleigh formulierte Minimaleigenschaft des Kreises, Math. Ann. 94 (1925), no. 1, 97–100. 10.1007/BF01208645Search in Google Scholar

[21] W.-T. Li, J. López-Gómez and J.-W. Sun, Sharp patterns of positive solutions for some weighted semilinear elliptic problems, Calc. Var. Partial Differential Equations 60 (2021), no. 3, Paper No. 85. 10.1007/s00526-021-01993-9Search in Google Scholar

[22] J. López-Gómez, The maximum principle and the existence of principal eigenvalues for some linear weighted boundary value problems, J. Differential Equations 127 (1996), no. 1, 263–294. 10.1006/jdeq.1996.0070Search in Google Scholar

[23] J. López-Gómez, Approaching metasolutions by solutions, Differential Integral Equations 14 (2001), no. 6, 739–750. 10.57262/die/1356123244Search in Google Scholar

[24] J. López-Gómez, The boundary blow-up rate of large solutions, J. Differential Equations 195 (2003), no. 1, 25–45. 10.1016/j.jde.2003.06.003Search in Google Scholar

[25] J. López-Gómez, Optimal uniqueness theorems and exact blow-up rates of large solutions, J. Differential Equations 224 (2006), no. 2, 385–439. 10.1016/j.jde.2005.08.008Search in Google Scholar

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

[27] J. López-Gómez, Metasolutions of Parabolic Equations in Population Dynamics, CRC Press, Boca Raton, 2016. 10.1201/b19418Search in Google Scholar

[28] J. López-Gómez and P. H. Rabinowitz, The effects of spatial heterogeneities on some multiplicity results, Discrete Contin. Dyn. Syst. 36 (2016), no. 2, 941–952. 10.3934/dcds.2016.36.941Search in Google Scholar

[29] Y. Lou, On the effects of migration and spatial heterogeneity on single and multiple species, J. Differential Equations 223 (2006), no. 2, 400–426. 10.1016/j.jde.2005.05.010Search in Google Scholar

[30] M. Marcus and L. Véron, Uniqueness and asymptotic behavior of solutions with boundary blow-up for a class of nonlinear elliptic equations, Ann. Inst. H. Poincaré Anal. Non Linéaire 14 (1997), no. 2, 237–274. 10.1016/s0294-1449(97)80146-1Search in Google Scholar

[31] J. D. Murray, Mathematical Biology, 2nd ed., Springer, New York, 1998. Search in Google Scholar

[32] T. Ouyang, On the positive solutions of semilinear equations Δu+λu-hup=0 on the compact manifolds, Trans. Amer. Math. Soc. 331 (1992), no. 2, 503–527. 10.1090/S0002-9947-1992-1055810-7Search in Google Scholar

[33] J.-W. Sun, Asymptotic profiles in diffusive logistic equations, Z. Angew. Math. Phys. 72 (2021), no. 4, Paper No. 152. 10.1007/s00033-021-01582-ySearch in Google Scholar

Received: 2021-09-10
Revised: 2021-10-01
Accepted: 2021-10-02
Published Online: 2021-10-15
Published in Print: 2021-11-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 29.5.2024 from https://www.degruyter.com/document/doi/10.1515/ans-2021-2149/html
Scroll to top button