Home Loop Type Subcontinua of Positive Solutions for Indefinite Concave-Convex Problems
Article Open Access

Loop Type Subcontinua of Positive Solutions for Indefinite Concave-Convex Problems

  • Uriel Kaufmann , Humberto Ramos Quoirin and Kenichiro Umezu EMAIL logo
Published/Copyright: September 10, 2018

Abstract

We establish the existence of loop type subcontinua of nonnegative solutions for a class of concave-convex type elliptic equations with indefinite weights, under Dirichlet and Neumann boundary conditions. Our approach depends on local and global bifurcation analysis from the zero solution in a nonregular setting, since the nonlinearities considered are not differentiable at zero, so that the standard bifurcation theory does not apply. To overcome this difficulty, we combine a regularization scheme with a priori bounds, and Whyburn’s topological method. Furthermore, via a continuity argument we prove a positivity property for subcontinua of nonnegative solutions. These results are based on a positivity theorem for the associated concave problem proved by us, and extend previous results established in the powerlike case.

MSC 2010: 35J25; 35J61; 35B32

1 Introduction and Main Results

Let ΩN (N1) be a bounded domain with smooth boundary Ω. In this paper, we consider nonnegative solutions of the problem

(PB) { - Δ u = λ a ( x ) f ( u ) + b ( x ) g ( u ) in Ω , u = 0 on Ω ,

where

  1. Δ is the usual Laplacian in N;

  2. u : = u (Dirichlet) or u:=u𝐧 (Neumann), where 𝐧 is the outward unit normal to Ω;

  3. λ is a bifurcation parameter;

  4. a , b C ( Ω ¯ ) are such that a changes sign in Ω and b(x0)>0 for some x0Ω;

  5. f , g : [ 0 , ) are continuous functions with f(0)=g(0)=0.

It follows that (PB) possesses the trivial line (λ,0) of zero solutions. The prototype of f,g to be considered in this paper is

(1.1) f ( s ) = s q , g ( s ) = s p , with  0 < q < 1 < p ,

so that the nonlinearity λf(s)+g(s) has a concave-convex nature.

More precisely, we assume that fC1((0,)) with f(s)>0 for s>0 and gC1([0,)) satisfy

(1.2) lim s 0 + f ( s ) s = ,
(1.3) lim s 0 + g ( s ) s = 0 ,
(1.4) lim s f ( s ) s = 0 ,
(1.5) lim s g ( s ) s = .

Let r>N. A function uW2,r(Ω) (and consequently, uC1(Ω¯)) is said to be a nonnegative solution of (PB) if u0 in Ω, u satisfies the equation pointwisely a.e. in Ω, and u=0 on Ω. If, in addition, u satisfies

{ u > 0 in Ω and u 𝐧 < 0 on Ω if u = u , u > 0 on Ω ¯ if u = u 𝐧 ,

then we write u0. In this case u lies in the interior of the positive cone {uV:u0}, where

V : = { C 0 1 ( Ω ¯ ) : = { u C 1 ( Ω ¯ ) : u = 0 on Ω } if u = u , C 1 ( Ω ¯ ) if u = u 𝐧 .

A nonnegative solution u of (PB) is called positive if u0.

Our first goal is to establish, under certain conditions on a and b, the existence of loop type subcontinua{(λ,u)} (i.e., nonempty, closed and connected subsets in ×V) composed by (0,0) and nontrivial nonnegative solutions (λ,u) of (PB). We shall prove the existence of a loop type subcontinuum 𝒞0 such that

(1.6) 𝒞 0 { ( λ , 0 ) : λ } = { ( 0 , 0 ) } .

It should be emphasized that, in general, one can not deduce that nontrivial nonnegative solutions of (PB) satisfy u0, since the strong maximum principle does not apply. This is due to the fact that a(x)f() does not satisfy the slope condition [1, p. 623], see Remark 2.1 (ii) below. As a matter of fact, (PB) may have solutions u satisfying u>0 in Ω but not u0 (for concrete examples one may argue as in the proof of [15, Proposition 2.9] after a slight modification). In view of this difficulty, our second purpose is to show that nontrivial solutions lying on 𝒞0 satisfy u0.

Bounded subcontinua of positive solutions for indefinite superlinear equations of the form

- Δ u = λ a ( x ) u + b ( x ) u p in Ω ,

with Ω bounded (under different boundary conditions) or Ω=N, have been studied by several authors, see e.g. [4, 5, 8, 6, 21, 7, 19, 18]. According to [6, 7, 19], a bounded subcontinuum linking two different points on (λ,0) is called a mushroom, one that meets a single point on (λ,0) is called a loop, and one that does not touch (λ,0) is called an isola. Cingolani and Gámez studied both the Dirichlet condition case and the case Ω=N, proving the existence of mushrooms [8, Theorems 4.4 and 5.5]. Cano-Casanova considered a mixed boundary condition (with a second order uniformly strongly elliptic operator), and proved the existence of a mushroom [5, Theorem 1.4]. López-Gómez and Molina-Meyer dealt with the Dirichlet condition, and established existence results for a mushroom, a loop and an isola in three cases, respectively [19, Theorems 3.1, 5.1 and 5.2]. In the case of Neumann boundary conditions, Brown proved the existence of a mushroom and a loop in two situations, respectively [4, Sections 2 and 5]. Finally, we refer to [26, Section 3] for the existence of a mushroom of positive solutions for a semilinear elliptic problem with a logistic nonlinearity and an indefinite weight, coupled with a nonlinear boundary condition. Let us emphasize that all the previous works hold in the regular case, i.e., when the nonlinearity considered is C1 at u=0, so that the general theory on local and global bifurcation from simple eigenvalues can be directly applied.

Regarding existence results for positive solutions of concave-convex problems, a large number of works have been devoted to (PB) in the “definite case” (i.e. with a0, a0) since the classical work of Ambrosetti, Brezis and Cerami [3], which treats the model case (1.1) with a=b1 and pN+2N-2 under the Dirichlet boundary condition. In [3] it is proved that (PB) has two positive solutions for λ>0 sufficiently small. This result was extended by De Figueiredo, Gossez and Ubilla [11] to the non-powerlike case, with a0. In addition, in [10], the authors allowed a to change sign and proved the existence of two nontrivial nonnegative solutions of (PB) for λ>0 small. We refer to [23] for a discussion on concave-convex problems under the Neumann boundary condition.

To the best of our knowledge, besides [15, 23] there are no works providing the existence of solutions that are positive in Ω for indefinite concave-convex problems (i.e., with a changing sign). In [15, 16, 14], we first established a positivity property for (PB) in the powerlike and concave case, i.e. with f(s)=sq and b0. Thanks to these results, we obtained a positivity result for (PB) with f(s)=sq and b1 (see [15, Section 4]). Finally, let us mention that in the model case (1.1), the existence of a loop type subcontinuum of nonnegative solutions for the Neumann case was obtained by means of a bifurcation approach in [23]. Furthermore, the asymptotic profile of nonnegative solutions as λ0+ enables one to deduce in some cases their positivity for λ>0 small, cf. [23, Corollary 1.3].

For our first purpose, we assume that there exist two balls B,BΩ and constants a0,a0,b0,b0>0 such that

(1.7) { a ( x ) a 0 and b ( x ) b 0 in B , - a ( x ) a 0 and b ( x ) b 0 in B .

Let ψC(Ω¯) be such that

Ω + ψ : = { x Ω : ψ ( x ) > 0 } .

Then, we introduce the condition

(Hψ) Ω + ψ consists of a finite number of connected components of Ω .

We shall assume this condition for ψ=a and ψ=-a.

Motivated by the model case (1.1), we assume that

(1.8) lim s 0 + s 1 - q f ( s ) = : f 0 ( 0 , ) for some q ( 0 , 1 ) .

We will see that under this condition f behaves like f0qsq when s0+, and satisfies the slope condition, see Remark 2.1. In addition, the following strong concavity (resp. convexity) condition on f (resp. g) shall be used:

(1.9) ( f ( s ) s q ) 0 for s > 0 ,
(1.10) ( g ( s ) s ) > 0 for s > 0 ,

where q(0,1) is given by (1.8). We introduce now the Gidas–Spruck condition [12, Theorem 1.1], which is stronger than (1.5),

(1.11) 0 < lim s g ( s ) s p < for some p > 1 , where p < N + 2 N - 2 if N > 2 .

Finally, we shall use condition (Hb), which will be precisely stated in Remark 2.3 and goes back to Amann and López-Gómez [2].

For our second purpose, we focus on the case f(s)=sq, q(0,1). In association with the sublinear problem

(1.12) { - Δ u = a ( x ) u q in Ω , u = 0 on Ω ,

we introduce the set

(1.13) 𝒜 a : = { q ( 0 , 1 ) : u 0 for any nontrivial nonnegative solution u of (1.12) } .

We know [15, Corollary 1.5 and Theorem 1.9] that under () with ψ=a, there exists qa[0,1) such that 𝒜a=(qa,1), assuming additionally Ωa<0 if u=u𝐧. Let us point out that the condition Ωa<0 is necessary and sufficient for the existence of a positive solution of (1.12) with u=u𝐧, for some q(0,1), see [16, Corollary 1.3].

We are now in position to state our main results. First, we deal with the Dirichlet problem.

Theorem 1.1.

Under Bu=u, we assume (1.3), (1.4), (1.7), (1.8), (1.11) and () with ψ=±a. In addition, suppose either

  1. b > 0 on Ω ¯ , and p < N + 2 N - 2 if N > 2 , or

  2. ( 1.9 ), ( 1.10 ) and (Hb).

Then, the following two assertions hold:

  1. ( (PB) ) admits a loop type subcontinuum 𝒞 0 (i.e., a nonempty, closed and connected subset in × C 0 1 ( Ω ¯ ) ) of nonnegative solutions which satisfies ( 1.6 ). Moreover, we have the following properties, see Figure 1 (i):

    1. ( 0 , u 0 ) 𝒞 0 for some positive solution u 0 of ( (PB) ) with λ = 0 .

    2. There exists δ > 0 such that 𝒞 0 does not contain any positive solution u of ( (PB) ) with λ = 0 satisfying u C ( Ω ¯ ) δ .

    3. 𝒞 0 contains closed connected sets 𝒞 0 ± such that { ( 0 , 0 ) } 𝒞 0 ± , and if ( λ , u ) 𝒞 0 ± { ( 0 , 0 ) } , then λ 0 , i.e., 𝒞 0 bifurcates both subcritically and supercritically at ( 0 , 0 ) .

  2. Let f ( s ) = s q , q(0,1). Assume that b0, and additionally that

    (1.14) g ( s ) 0 𝑓𝑜𝑟 s > 0

    when condition (a) holds. If q𝒜a𝒜-a, then u0 for any (λ,u)𝒞0{(0,0)}. In particular, the component of nontrivial nonnegative solutions of ((PB)) including 𝒞0{(0,0)} is bounded.

Figure 1

The loop type subcontinua 𝒞0 and 𝒞.

(i)
(i)
(ii)
(ii)

Next, we consider the Neumann problem under the condition

(1.15) Ω b < 0 .

In this case, we shall obtain a loop type subcontinuum of nonnegative solutions with the same nature the Dirichlet case admits, as in Figure 1 (i) (for the case Ωb0 we refer to Remark 1.3 below). To this end, we need the following decay and positivity condition for g, which is stronger than (1.3):

(1.16) lim s 0 + g ( s ) s σ = : g 0 ( 0 , ) for some σ > 1 , where σ < 2 N N - 2 if N > 2 .

Moreover, we are able to discuss the positivity of (nontrivial) nonnegative solutions for (PB) with (1.1), assuming

(1.17) { p < N + 1 N - 1 if N > 2 , a C α ( Ω ¯ ) for some α ( 0 , 1 ) , Ω a < 0 , b 1 .

It is known [25, Theorem 1] that if Ω+a is connected, then (PB) possesses a loop type subcontinuum 𝒞 in ×C(Ω¯) of nonnegative solutions which satisfies (1.6). Furthermore, we have the following properties, see Figure 1 (ii):

  1. { λ : ( λ , u ) 𝒞 { ( 0 , 0 ) } } = ( 0 , Λ ] for some Λ>0.

  2. 𝒞 possesses at least two nontrivial nonnegative solutions for λ>0 small enough.

Now, we state our main results for the Neumann problem, which are given in a similar way as in Theorem 1.1, and where condition (1.15) provides us with a loop type subcontinuum bifurcating both subcritically and supercritically at (0,0).

Theorem 1.2.

Under Bu=un, assume (1.7), (1.8), (1.9), (1.10), (1.11), (1.16), () with ψ=±a, and (Hb). Then, the following two assertions hold:

  1. If ( 1.15 ) holds, then ( (PB) ) admits a loop type subcontinuum 𝒞 0 in × C 1 ( Ω ¯ ) of nonnegative solutions for which the same assertions in Theorem 1.1 (i) hold true, see Figure 1(i).

  2. Assume ( 1.1 ), ( 1.17 ) and the condition that Ω + a is connected. Let 𝒞 be the loop type subcontinuum stated above. If q 𝒜 a , then the same conclusion in Theorem 1.1 (ii) holds with 𝒞0 replaced by 𝒞.

Remark 1.3.

When u=u𝐧 and Ωb(x)0, the existence of a loop type subcontinuum of nonnegative solutions of (PB) has been established in the particular case f(s)=sq and g(s)=sp with 0<q<1<p (see [25], and as a particular case, see also (1.17)). In this case, although the loop type subcontinuum 𝒞 satisfies (1.6), 𝒞{(0,0)} appears in λ>0. This means that 𝒞 never meets the vertical line {(0,u):0u0}, see [23, Lemma 6.8 (1)]. Thus, the approach used in the proof of Theorem 1.2 (i) does not work for excluding the possibility that 𝒞={(0,0)}, see the argument in Subsection 5.1. Let us mention that in [25] the authors used a suitable rescaling technique (which strongly relies on the homogeneity of f(s)=sq and g(s)=sp) to exclude this possibility.

Remark 1.4.

Theorems 1.1 and 1.2 can be extended to the case a,bL(Ω) except assertion (ii) in Theorem 1.2. This can be done if we formulate () for ψL(Ω) such that ψ0, letting Ω+ψ be the largest open subset of Ω in which ψ>0 a.e., and assuming additionally

{ ψ is bounded away from zero on compact subsets of Ω i for i = 1 , , l , and | ( supp ψ + ) Ω + ψ | = 0 .

2 Preliminaries and Examples

We start this section with some remarks concerning some of our assumptions.

Remark 2.1.

Condition (1.8) implies:

  1. By the L’Hospital rule,

    (2.1) lim s 0 + f ( s ) s q = f 0 q > 0 .

    In particular, since fC1((0,)) and f(0)=0, we can show that fCα([0,s0]) for α(0,q] and s0>0.

  2. f satisfies the slope condition, that is, for any s0>0, there exists M0>0 such that

    (2.2) f ( s ) - f ( t ) s - t > - M 0 for  0 t < s s 0 .

However, even under (1.8), a(x)f() does not satisfy the slope condition for xΩ where a(x)<0, since lims0+f(s)=.

Remark 2.2.

  1. Since f(s)>0 for s>0, we note that if (1.9) holds, then f is concave for s>0, i.e.,

    ( f ( s ) s ) < 0 for s > 0 .

  2. It is easy to check that (1.9) is stronger than (1.4). This is a consequence of the fact that (1.9) yields

    0 lim s f ( s ) s q < .

  3. Let us also note that (1.3) and (1.10) imply that g(s)>0 for s>0. Indeed, assume first g(s0)<0 for some s0>0, and set ε0:=-g(s0)/s0>0. From (1.3), we infer that for some s1(0,s0),

    g ( s 0 ) s 0 = - ε 0 < g ( s 1 ) s 1 ,

    which contradicts (1.10). Hence g(s)0 for all s>0. Next, assume that g(s0)=0 for some s0>0. By (1.10), it follows that g(s)0 for s(0,s0). This implies that g(s1)>0 for some s1(0,s0). It follows that

    g ( s 1 ) s 1 > 0 = g ( s 0 ) s 0 ,

    which contradicts (1.10) again, as desired.

Remark 2.3.

We describe here the explicit formula for the growth condition (Hb) of b+ in a neighborhood of Ω+b used in Theorems 1.1 and 1.2, which originates from Amann and López-Gómez [2, Theorem 4.3]:

  1. Ω + b is a subdomain of Ω with smooth boundary Ω+b and either

    1. Ω + b ¯ Ω and b<0 in Db:=ΩΩ+b¯, or

    2. Ω + b { x Ω : d ( x , Ω ) < σ } for some σ>0 and b<0 in Db.

    In addition, Db is a subdomain of Ω with smooth boundary, and there exist γ>0 and a function β defined in a tubular neighborhood U:={xΩ+b:d(x,Ω+b)<σ} of Ω+b in Ω+b, which is continuous, positive and bounded away from zero, and satisfies b+()=β()d(,Ω+b)γ in U and

    1 < p < min ( N + 2 N - 2 , N + 1 + γ N - 1 ) if N > 2 .

We conclude this section showing some examples of functions satisfying the previous conditions. We start with the following lemma, which characterizes the functions satisfying (1.8) and (1.9).

Lemma 2.4.

Let fC([0,))C1((0,)) with f(0)=0 and f(s)>0 for s>0. Then, the following two conditions are equivalent:

  1. ( 1.8 ) and ( 1.9 ) hold.

  2. f ( s ) = s q h ( s ) for some q ( 0 , 1 ) and h C ( [ 0 , ) ) C 1 ( ( 0 , ) ) such that

    h is nonincreasing , h ( s ) > 0 𝑓𝑜𝑟 s 0 𝑎𝑛𝑑 lim s 0 + s h ( s ) = 0 .

Proof.

It is easy to see that f as in condition (ii) fulfills (1.8) and (1.9). Conversely, if f satisfies the aforementioned conditions, defining h(s):=s-qf(s) for s>0 and h(0):=lims0+h(s), it is also easy to check that h has the desired properties. ∎

As particular cases, we mention h(s)=11+sr (r0) and h(s)=e-s.

We note that oscillatory cases are out of our scope. For instance, consider h(s)=sin(1s)+2. If we put f(s):=sqh(s) for s>0 and f(0):=0, then fC([0,))C1((0,)) with f(0)=0 and f(s)>0 for s>0. Moreover, f fulfills (1.2) and (1.4), but (2.1), (2.2) and (1.9) do not hold.

We now exhibit examples of g satisfying (1.10), (1.11), and (1.16).

(a) We set

g ( s ) : = s p h ( s ) , with 1 < p < N + 2 N - 2 if N > 2 ,

where hC1([0,)) is nondecreasing, bounded, and satisfies one of the following conditions:

  1. 0 = h ( 0 ) < h ( 0 ) .

  2. 0 < h ( 0 ) .

Note that (1.16) holds if we choose σ:=p+1 in (i) and σ:=p in (ii). An example for (i) is h(s)=1-e-s, while h(s)=arctan(s+1) is included in (ii). Another example is given by h(s)=sr1+sr with r=0 or r=1. The case r=1 satisfies (i), whereas r=0 satisfies (ii). More generally, the function g(s)=sp+r1+sr (0r<2NN-2-p) fulfills (1.10), (1.11) and (1.16). Indeed, we can take σ:=p+r for (1.16).

(b) Let k1 and 1<p<N+2N-2 if N>2. We put

g ( s ) : = s p k + s 1 + s .

Then, g satisfies (1.11) and (1.16). It is also clear that (1.10) holds when k=1. Meanwhile, when k>1, it satisfies (1.10) if additionally

p > p 1 ( k ) , where p 1 ( k ) : = 2 k k + 1 .

We note that p1(k) is increasing for k>1, and p1(k)1 as k1+, whereas p1(k)2 as k. Let us finally observe that case (b) is not included in any of the possibilities considered in case (a). Indeed, h(s)=k+s1+s is decreasing for s0.

3 Regularization Schemes and Transversality Conditions

Let us now explain our approach to study bifurcation of nontrivial nonnegative solutions for (PB) from (λ,0). From (1.2), we see that f is not differentiable at s=0, so that we can not directly apply the usual bifurcation theory from simple eigenvalues to (PB). To overcome this difficulty, we proceed as in [23, 25], “regularizing” (PB) at u=0, using ε>0. We refer to [19, Section 5] for a similar approach introducing a new parameter for a different regular problem.

We extend g to as a C1 function and set F: by

F ( s ) : = { s 1 - q f ( s ) , s 0 , f 0 q s , s < 0 .

For ε>0, we shall study the auxiliary problem

(PcalBε) { - Δ u = λ a ( x ) ( u + ε ) q - 1 F ( u ) + b ( x ) g ( u ) in Ω , u = 0 on Ω .

Note that (1.8) and (2.1) imply that FC1(), F(0)=0 and F(0)=f0q, so that

(3.1) s ( s + ε ) q - 1 F ( s ) C 1 ( ( - ε , ) ) .

Observe also that (P,0) corresponds to (PB), as far as nonnegative solutions are concerned.

Let us set

h λ , ε ( x , s ) : = λ a ( x ) ( s + ε ) q - 1 F ( s ) + b ( x ) g ( s ) .

From (3.1), we see that h(x,) satisfies the slope condition. Consequently, given a nontrivial nonnegative solution u of (PcalBε), we can choose M>0 such that (-Δ+M)u0 and 0 in Ω. Thus, by the strong maximum principle and Hopf’s lemma, we deduce u0, see [13], [18, Theorem 7.10].

We shall then consider the linearized eigenvalue problem at u=0 for the regular problem (PcalBε):

(3.2) { - Δ ϕ = λ a ( x ) f 0 q ε q - 1 ϕ in Ω , ϕ = 0 on Ω .

Since a changes sign, (3.2) has exactly two principal eigenvalues λ1,ε-<0<λ1,ε+ (resp. λ1,ε-=0<λ1,ε+) if

u = u ( respectively u = u 𝐧 and Ω a < 0 ) ,

which are both simple, and furthermore, (λ1,ε±,0) satisfy the Crandall–Rabinowitz transversality condition, see [18, Theorem 9.4].

Thanks to the simplicity and transversality condition, the local bifurcation theory [9, Theorem 1.7] ensures the existence and uniqueness of positive solutions of (PcalBε) bifurcating at (λ1,ε±,0). Moreover, the unilateral global bifurcation theory [17, Theorem 6.4.3] (see also [22, Theorem 1.27]) ensures that (PcalBε) possesses two components 𝒞ε±={(λ,u)} (i.e., maximal, nonempty, closed and connected subsets in ×V) of nonnegative solutions emanating from (λ1,ε±,0), respectively (see Remark 3.2). In addition, 𝒞ε+{(λ1,ε±,0)} and 𝒞ε-{(λ1,ε±,0)} consist of positive solutions. This is due to elliptic regularity and the fact (see [1, Proposition 18.1]) that (PcalBε) has no bifurcating positive solutions from (λ,0) at any λλ1,ε±.

Under some additional growth condition on g, we shall verify that 𝒞ε± are bounded in ×V uniformly in ε(0,1], so that 𝒞ε-=𝒞ε+ (:=𝒞ε) (i.e. 𝒞ε is a mushroom). By simple computations, it can be shown easily that

(3.3) λ 1 , ε ± 0 as ε 0 + ,

so that, passing to the limit as ε0+, we shall observe by Whyburn’s topological argument [27, (9.12) Theorem] that

𝒞 0 : = lim sup ε 0 + 𝒞 ε

is a loop type subcontinuum which consists of nonnegative solutions of (PB) and satisfies (1.6).

Remark 3.1.

The critical case

(3.4) u = u 𝐧 and Ω a = 0

can be handled in a similar way. In this case, we replace a by a-ε for ε>0 small in (PcalBε). Then, the above argument remains valid, since we can determine the asymptotic behavior (3.3) (see [23, Lemma 6.6] for the proof). In addition, we can reduce the case Ωa>0 to the case Ωa<0 under u=u𝐧. Indeed, we only have to notice the symmetry property λa(x)=(-λ)(-a(x)). The situation may be illustrated by Figure 2 (i) and (ii).

Figure 2

The bounded component 𝒞ε for (PcalBε) with u=u𝐧.

(i) 
            Case ∫Ωa<0{\int_{\Omega}a<0}.
(i)

Case Ωa<0.

(ii) 
            Case ∫Ωa>0{\int_{\Omega}a>0}.
(ii)

Case Ωa>0.

Remark 3.2.

We shall show that the transversality condition allows us to apply the unilateral global bifurcation result [17, Theorem 6.4.3] to (PcalBε) at (λ1,ε±,0). To this end, we reduce (PcalBε) to an operator equation in C(Ω¯). Given ξLr(Ω), r>N, let uW2,r(Ω) be the unique solution of

(3.5) { ( - Δ + 1 ) u = ξ ( x ) in Ω , u = 0 on Ω ,

where

W 2 , r ( Ω ) : = { { u W 2 , r ( Ω ) : u = 0 on Ω } if u = u , W 2 , r ( Ω ) if u = u 𝐧 .

We introduce the solution operator 𝒮:Lr(Ω)W2,r(Ω) associated with (3.5), implying that 𝒮(ξ)=u, which is bijective and homeomorphic. It follows that 𝒮:L(Ω)W2,r(Ω) is continuous, and moreover, 𝒮:L(Ω)V is compact, since so is the embedding W2,r(Ω)V. Thus, (PcalBε) is reduced to

(3.6) ( λ , u ) : = u - 𝒮 [ ( λ a f 0 q ε q - 1 + 1 ) u + h λ ( x , u ) ] = 0 in C ( Ω ¯ ) ,

where

h λ ( x , s ) : = λ a ( x ) { ( s + ε ) q - 1 F ( s ) - f 0 q ε q - 1 s } + b ( x ) g ( s ) .

Given uC(Ω¯), let

𝒜 λ u : = 𝒮 [ ( λ a f 0 q ε q - 1 + 1 ) u ] ,

and let

Ind ( 0 , 𝒜 λ ) : = deg ( 1 - 𝒜 λ , B R )

be the fixed point index of 𝒜λ at the origin for λλ1,ε± but close to λ1,ε±, where BR is the ball with radius R>0 and centered at the origin. According to [17, Section 4.2, Theorem 5.6.2], the transversality condition at (λ1,ε±,0) implies that Ind(0,𝒜λ) changes sign as λ crosses λ1,ε±. Thus, [17, Theorem 6.4.3] applies, so that equation (3.6) (and so, (PcalBε)) possesses two components 𝒞ε±={(λ,u)} in ×C(Ω¯) of nonnegative solutions u emanating from (λ1,ε±,0), respectively. Finally, we can verify that 𝒞ε± are also components in ×V, by elliptic regularity.

In the forthcoming sections, we will characterize the limiting behavior of 𝒞ε as ε0+ under the conditions on a,b,f and g stated in Theorems 1.1 and 1.2.

4 A Priori Bounds

In this section, we establish an a priori bound for positive solutions of the regularized problem (PcalBε) in ×V (Corollary 4.6).

We start with an a priori bound on λ, uniformly in ε[0,1].

Proposition 4.1.

Assume (1.3), (1.5), (1.8) and (1.7). Then, there exist λ¯,ε¯>0 such that if 0εε¯, then (PcalBε) has no positive solution for |λ|λ¯.

Proof.

Let us suppose first that we are not in the case (3.4). Let 0ε1, and assume that (PcalBε) has a positive solution u for some λ>0. Let B be given by (1.7), λB>0 be the first eigenvalue of the problem

(4.1) { - Δ ϕ = λ ϕ in B , ϕ = 0 on B ,

and ϕBC2(B¯) be a positive eigenfunction associated to λB. We extend ϕB to Ω¯ by setting ϕB=0 in Ω¯B¯, so that ϕBH01(Ω). Since u>0 on B¯ and ϕBν<0 on B, the divergence theorem yields that

λ B B ϕ B u = B - Δ ϕ B u = B ϕ B u - B ϕ B ν u > B ϕ B u ,

where ν is the outward unit normal to B. On the other hand, we see that

B u ϕ B = B b ( x ) g ( u ) ϕ B + λ B a ( x ) ( u + ε ) q - 1 F ( u ) ϕ B ,

where q is given by (1.8). It follows that

(4.2) B u q ϕ B { b ( x ) g ( u ) u q + λ a ( x ) ( u u + ε ) 1 - q F ( u ) u - λ B u 1 - q } < 0 .

Now, for (x,s)B×(0,), we set

h ( x , s ) : = b ( x ) g ( s ) s q + λ a ( x ) ( s s + ε ) 1 - q F ( s ) s - λ B s 1 - q .

By (1.5), there exists s0>0 such that

g ( s ) λ B b 0 s for s > s 0 ,

where b0 is from (1.7). Hence, since f(s)>0 for s>0 and a(x)0 a.e. in B, we deduce that if λ>0, xB and s>s0, then

(4.3) h ( x , s ) b ( x ) g ( s ) s q - λ B s 1 - q s 1 - q ( b 0 g ( s ) s - λ B ) 0 .

Let us now consider the case 0<ss0. From (1.3), we can choose K0>0 such that

| g ( s ) s | K 0 for  0 < s s 0 .

Recalling (1.8) (or (2.1)), we set

M 1 : = inf 0 < s s 0 F ( s ) s > 0 .

By putting b:=b, it follows that

h ( x , s ) = s 1 - q { b ( x ) g ( s ) s + λ a ( x ) ( 1 s + ε ) 1 - q F ( s ) s - λ B } s 1 - q { λ a 0 ( 1 s 0 + 1 ) 1 - q M 1 - ( λ B + b K 0 ) } ,

so that h(x,s)0 for xB and 0<ss0 if

(4.4) λ λ ¯ : = ( λ B + b K 0 ) ( s 0 + 1 ) 1 - q a 0 M 1 .

Consequently, by (4.2), (4.3) and (4.4) we deduce that λ<λ¯.

Next, let us verify the existence of a lower bound on λ<0 for the existence of a positive solution of (PcalBε). In order to check this, we notice that if (PcalBε) has a positive solution u for some λ<0 and ε[0,1], then

- Δ u = ( - λ ) ( - a ( x ) ) ( u + ε ) q - 1 F ( u ) + b ( x ) g ( u ) in Ω .

From (1.7), the desired conclusion follows arguing as above with B now replaced by B.

It remains to consider case (3.4). However, it suffices to note that (1.7) implies that if ε is small enough, then

a ε ( x ) a 0 2 in B and - a ε ( x ) a 0 in B .

The proof now follows in the same way as above. ∎

Next, given a compact interval I, we establish an a priori upper bound for positive solutions of (PcalBε) whenever λI and ε[0,1]. We start with the following preliminary lemma (see also [2, Theorem 4.1]).

Lemma 4.2.

Assume (1.5), (1.8), (1.9), (1.10) and (Hb). Let Λ>0. Suppose there exists a constant C1>0 such that uC(Ω+b¯)C1 for all positive solutions u of (PcalBε) with λ[-Λ,Λ] and ε[0,1]. Then, there exists C2>0 such that uC(Ω¯)C2 for all positive solutions u of (PcalBε) with λ[-Λ,Λ] and ε[0,1].

Proof.

(i) First, we consider the Dirichlet case. We use a comparison principle for concave problems inspired by the one in [3, Lemma 3.3].

(1) Assume first that Ω+b¯Ω, and recall that Db is given by (Hb). Let λ[0,Λ], and consider the problem

(4.5) { - Δ v = λ a + ( x ) ( v + ε ) q - 1 F ( v ) - b - ( x ) g ( v ) in D b , v = C 1 on Ω + b , v = 0 on Ω .

Let u be a positive solution of (PcalBε), with λ[0,Λ] and ε[0,1]. It follows that u>0 in Db. Since λ0, b=-b- in Db, and f(s)>0 for s>0, it is easy to check that u is a subsolution of (4.5).

Next, we construct a supersolution of (4.5). Consider the unique positive solution w0 of the problem

{ - Δ w = 1 in D b , w = 0 on Ω + b Ω .

Set w¯:=C(w0+1) for C>0. If CC1, then w¯C1 on Ω+b and w¯0 on Ω. Moreover, we claim that if C is sufficiently large, then

- Δ w ¯ λ a + ( x ) ( w ¯ + ε ) q - 1 F ( w ¯ ) - b - ( x ) g ( w ¯ ) in D b .

Indeed, let

δ : = 1 Λ a + ( w 0 C ( D b ¯ ) + 1 ) > 0 .

Then, from (1.9) and (1.5), there exists s1>0 large enough such that if ss1, then

0 g ( s ) and f ( s ) δ s ( so that F ( s ) δ s 2 - q ) .

It follows that if Cs1, then

- Δ w ¯ - { λ a + ( x ) ( w ¯ + ε ) q - 1 F ( w ¯ ) - b - ( x ) g ( w ¯ ) } C - Λ a + w ¯ q - 1 δ w ¯ 2 - q C { 1 - δ Λ a + ( w 0 C ( D b ¯ ) + 1 ) } = 0 in D b .

Thus the claim has been verified, and w¯ is a supersolution of (4.5) if Cmax(C1,s1). Note that C can be chosen independently of λ[0,Λ] and ε[0,1].

Now, we see from (1.9) and (1.10) that the nonlinearity in (4.5) is concave, that is, if we set

j ( x , s ) : = λ a + ( x ) ( s + ε ) q - 1 F ( s ) - b - ( x ) g ( s ) , x D b , s > 0 ,

then

s ( j ( x , s ) s ) < 0 for x D b , s > 0 .

Indeed,

d d s ( ( s + ε ) q - 1 F ( s ) s ) = ( q - 1 ) ( s + ε ) q - 2 ( s - q f ( s ) ) + ( s + ε ) q - 1 d d s ( s - q f ( s ) ) < 0 .

Reasoning as in [24, Proposition A.1] (whose argument is based on [3, Lemma 3.3]), we may deduce that uw¯ in Db, so that

u C 1 + w ¯ C ( D b ¯ ) on Ω ¯ .

It remains to verify the case -Λλ<0. Note that any positive solution u of (PcalBε) with λ[-Λ,0) and ε[0,1] satisfies

- Δ u = μ ( - a ( x ) ) ( u + ε ) q - 1 F ( u ) + b ( x ) g ( u ) in D b ,

with μ:=-λ(0,Λ]. Instead of (4.5), we consider the following concave problem:

(4.6) { - Δ v = μ a - ( x ) ( v + ε ) q - 1 F ( v ) - b - ( x ) g ( v ) in D b , v = C 1 on Ω + b , v = 0 on Ω .

Then, we see that u is a subsolution of this problem. The remainder of the argument is identical to the one in the case λ[0,Λ].

(2) Assume now that Ω+b{xΩ:d(x,Ω)<σ} for some σ>0. This case can be verified identically. Indeed, it suffices to replace (4.5) by the problem

{ - Δ v = λ a + ( x ) ( v + ε ) q - 1 F ( v ) - b - ( x ) g ( v ) in D b , v = C 1 on Ω + b Ω .

(ii) Lastly, we verify the Neumann case. Since aεa+ and -aεa-+1, it suffices to replace (4.5) by

{ - Δ v = λ a + ( x ) ( v + ε ) q - 1 F ( v ) - b - ( x ) g ( v ) in D b , v = C 1 on Ω + b , v 𝐧 = 0 on Ω ,

and (4.6) by

{ - Δ v = μ ( a - ( x ) + 1 ) ( v + ε ) q - 1 F ( v ) - b - ( x ) g ( v ) in D b , v = C 1 on Ω + b , v 𝐧 = 0 on Ω .

The proof of Lemma 4.2 is now complete. ∎

The following result is due to Gidas and Spruck [12, Theorem 1.1] and Amann and López-Gómez [2, Section 4] (see also López-Gómez, Molina-Meyer and Tellini [20, Section 6]).

Proposition 4.3.

Assume (1.4), (1.8) and (1.11). In addition, suppose either

  1. b > 0 on Ω ¯ and p < N + 2 N - 2 for the Dirichlet case ( p < N + 1 N - 1 for the Neumann case) if N > 2 , or

  2. ( 1.9 ), ( 1.10 ) and (Hb) hold.

Then, given Λ>0, there exists C>0 such that if 0ε1, then uC in Ω for all positive solutions u of (PcalBε) with |λ|Λ.

Remark 4.4.

  1. Condition (i) for the Neumann case is based on (Hb) with γ=0 from Amann and López-Gómez [2, Section 4], not on Gidas and Spruck [12, Theorem 1.1].

  2. Under condition (ii), we can handle the case where b>0 in Ω and b=0 somewhere on Ω.

Proof.

Case (i) under u=u is verified in the same way as in the proof of [12, Theorem 1.1].

Case (i) under u=u𝐧 and case (ii) can be handled using the arguments in [23, Proposition 6.5] and [24, Proposition 4.2], which are based on [20, Section 6]. Indeed, based on (1.4), (1.11) and (Hb), we employ the argument developed by Amann and López-Gómez [2] to deduce that the hypothesis of Lemma 4.2 is fulfilled, so that the desired conclusion follows. ∎

By H we denote the usual norm of H=H01(Ω) when u=u and of H=H1(Ω) when u=u𝐧. The next lemma follows easily by a bootstrap argument based on elliptic regularity and the Sobolev embedding theorem.

Lemma 4.5.

Assume that there exist C>0 and p(1,N+2N-2) such that

| g ( s ) | C ( 1 + s p ) 𝑓𝑜𝑟 s 0 .

Let Λ>0 and u be a nonnegative solution of (PcalBε) for λ[-Λ,Λ] and ε[0,1]. Then, given c1>0, there exists c2>0 such that if uHc1, then uVc2.

With the aid of Lemma 4.5, Propositions 4.1 and 4.3 provide us with an a priori bound in ×V for positive solutions of (PcalBε) uniformly in ε[0,1].

Corollary 4.6.

Under the assumptions of Propositions 4.1 and 4.3, there exists ε¯,C¯>0 such that if 0εε¯, then |λ|+uVC¯ for all positive solutions u of (PcalBε).

Proof.

Propositions 4.1 and 4.3 imply that there exist ε¯,C>0 such that if 0εε¯, then |λ|+uHC for all positive solutions u of (PcalBε), since u satisfies

Ω | u | 2 = λ Ω { a ( x ) ( u + ε ) q - 1 F ( u ) u + b ( x ) g ( u ) u } C

for some C>0. Lemma 4.5 provides then the desired conclusion. ∎

We discuss now bifurcation of nontrivial nonnegative solutions for (PB) from (λ,0) for λ>0. We prove the following preliminary lemma.

Lemma 4.7.

Assume (1.8) and () with ψ=a. Let Λ>0, Ω be a nonempty connected open subset of Ω+a, and BΩ be a ball. Then there exists CΛ>0 such that if λΛ, then uC(B¯)CΛ for all nontrivial nonnegative solutions u of (PB) such that u0 in Ω.

Proof.

We use an argument based on subsolutions and supersolutions. First of all, we remark that (1.8) implies (1.2) and (2.2), see Remark 2.1.

Let Λ,Ω and B be as in the statement of this lemma, and let λΛ. Assume that u is a nontrivial nonnegative solution of (PB) such that u0 in Ω. Set

(4.7) K 1 : = max 0 s u C ( Ω ¯ ) | g ( s ) | 0 .

Since g(0)=0, the mean value theorem provides some constant θ=θx(0,1) such that g(u)=g(θu)u. Thus, using (4.7) we get that

( - Δ + b K 1 + 1 ) u = λ a ( x ) f ( u ) + ( b K 1 + 1 + b ( x ) g ( θ u ) ) u u 0 and 0 in Ω .

The strong maximum principle yields that u>0 in Ω. Now, let s0>0 be fixed. Then, the following two possibilities may occur: (i) us0 on B¯; (ii) u>s0 somewhere on B¯.

We consider case (i). We have a0:=infB¯a>0, so that u is a supersolution of the problem

(4.8) { - Δ v + b K 2 v = λ a 0 f ( v ) in B , v = 0 on B ,

where

K 2 : = max 0 s s 0 | g ( s ) | 0 .

Indeed, u0 on B. Moreover, since f(s)>0 for s>0, the mean value theorem shows that

- Δ u + b K 2 u - λ a 0 f ( u ) = λ ( a ( x ) - a 0 ) f ( u ) + ( b K 2 + b ( x ) g ( θ u ) ) u 0 in B .

To construct a subsolution of (4.8), we use the positive eigenfunction ϕB associated to the first eigenvalue λB of (4.1) and such that ϕBC(B¯)=1. From (1.2), we find a constant s1>0 small enough such that

(4.9) f ( s ) s λ B + b K 2 Λ a 0 for  0 < s s 1 .

If 0<ss1, then we observe that

- Δ ( s ϕ B ) + b K 2 s ϕ B - λ a 0 f ( s ϕ B ) s ϕ B { λ B + b K 2 - Λ a 0 f ( s ϕ B ) s ϕ B } 0 in B .

This implies that sϕB is a subsolution of (4.8) whenever 0<ss1. Now, since u>0 in Ω, it follows that u>0 on B¯. Furthermore, we assert that

(4.10) u s 1 ϕ B on B ¯ .

By contradiction, we assume that us1ϕB. Then, since u>0=s1ϕB on B, we can choose σ(0,1) such that u-σs1ϕB0 on B¯ and u-σs1ϕB=0 somewhere in B. From (2.2), we fix M0>0 such that

f ( s ) - f ( t ) s - t > - M 0 for  0 t < s s 0 .

Putting M1:=λa0M0>0, we see that the mapping

s M 1 s + λ a 0 f ( s )

is nondecreasing for 0ss0. Indeed, if 0t<ss0, then

M 1 s + λ a 0 f ( s ) - ( M 1 t + λ a 0 f ( t ) ) = ( M 1 + λ a 0 f ( s ) - f ( t ) s - t ) ( s - t ) ( M 1 - λ a 0 M 0 ) ( s - t ) = 0 ,

as desired. Thus, using this monotonicity and having in mind that (recall (4.9))

( - Δ + b K 2 ) σ s 1 ϕ B λ a 0 f ( σ s 1 ϕ B ) in B ,

we deduce that

( - Δ + b K 2 + M 1 ) ( u - σ s 1 ϕ B ) M 1 u + λ a 0 f ( u ) - ( M 1 σ s 1 ϕ B + λ a 0 f ( σ s 1 ϕ B ) ) 0 in B

and u-σs1ϕB=u>0 on B. The strong maximum principle yields that u-σs1ϕB>0 in B, a contradiction. Thus, we have verified (4.10). By taking into account case (ii), CΛ:=min{s0,s1} is as desired. ∎

By virtue of Lemma 4.7, there are no nontrivial nonnegative solutions of (PB) bifurcating from (λ,0) for λ>0, and moreover, there exist no small positive solutions of (PB) for λ=0. In view of this fact, although we shall observe that (PB) possesses a bounded subcontinuum of nontrivial nonnegative solutions bifurcating at (0,0), we infer that the bifurcation subcontinuum is of loop type.

Proposition 4.8.

Assume (1.3). Then, the following assertions hold:

  1. Assume ( (Hψ) ) with ψ = a , and additionally ( 1.8 ) if u = u 𝐧 . Then, given λ 0 > 0 , there exist δ 0 , c 0 > 0 such that u C ( Ω ¯ ) c 0 for all nontrivial nonnegative solutions u of ( (PB) ) for λ ( λ 0 - δ 0 , λ 0 + δ 0 ) .

  2. Assume ( 1.15 ) and ( 1.16 ) if u = u 𝐧 . Then, there exists C > 0 such that u C ( Ω ¯ ) C for all positive solutions u of ( (PB) ) for λ = 0 .

Proof.

(i) We recall that (1.8) also implies (2.1), see Remark 2.1. First, we verify the Dirichlet case. By contradiction, we assume that λnλ0>0, and un are nontrivial nonnegative solutions of (PB) with λ=λn such that unC(Ω¯)0. Then, we claim that, up to a subsequence, Ωa(x)f(un)un0. If not, then we may suppose that Ωa(x)f(un)un>0 for all n. It follows that un0 in Ω+a. Indeed, if un0 in Ω+a, then, using that f(s)>0 for s>0, we find that

0 < Ω a ( x ) f ( u n ) u n Ω + a a ( x ) f ( u n ) u n = 0 ,

which is a contradiction. Employing () with ψ=a, we may deduce that there exists a connected open subset ΩΩ+a such that un0 in Ω for all n1. Let BΩ be a ball. We apply Lemma 4.7 with Λ=λ02, to derive that unC(B¯)c0 for some c0>0 independent of n, which contradicts unC(Ω¯)0. Thus, the claim follows.

Now, we observe from the definition of un that

u n H 2 : = Ω | u n | 2 = λ n Ω a ( x ) f ( u n ) u n + Ω b ( x ) g ( u n ) u n Ω b | g ( u n ) | u n .

Set vn:=un/unH, so that

(4.11) v n H 2 b Ω | g ( u n ) | u n u n H 2 .

From (1.3), given ε>0, there exists sε>0 such that

| g ( s ) | ε b s for  0 < s s ε .

Also, for n large enough, we have that unC(Ω¯)sε, so that

(4.12) b Ω | g ( u n ) | u n u n H 2 ε Ω v n 2 .

From (4.11) and (4.12), we derive that vn0 in H01(Ω), a contradiction.

Next, we verify the Neumann case. Assume to the contrary that λnλ0>0, and un are nontrivial nonnegative solutions of (PB) with λ=λn such that unC(Ω¯)0. We remark that unH0, since un are nonnegative solutions of (PB) with λ=λn. Arguing as in the proof for the Dirichlet case, we have that, up to a subsequence, Ωa(x)f(un)un0, and consequently, Ω|un|2bΩ|g(un)|un.

Set vn:=un/unH, so that vnH=1. We may assume that there exists v0H1(Ω) such that vnv0 in H1(Ω), vnv0 a.e. in Ω, and vnv0 in Lt(Ω) for t<2*. By (1.3), for any ε>0, there exists s0>0 such that

(4.13) | g ( s ) | ε b s for  0 s s 0 .

Thus, for n large enough, we have that unC(Ω¯)s0, so that

Ω | v n | 2 b Ω | g ( u n ) | u n H v n ε Ω v n 2 ε .

This implies that Ω|vn|20, and it follows that vnv0, and v0 is a positive constant.

Since un is a nonnegative solution of (PB) with λ=λn, we see that, for every ϕH1(Ω),

(4.14) ( Ω v n ϕ ) u n H 1 - q = λ n Ω a ( x ) f ( u n ) u n H q ϕ + Ω b ( x ) g ( u n ) u n H q ϕ .

Since unC(Ω¯)0, (4.13) implies that |g(un)|un for n large enough, so that

| Ω b ( x ) g ( u n ) u n H q ϕ | b Ω u n u n H q | ϕ | b u n H 1 - q ( Ω v n 2 ) 1 2 ( Ω ϕ 2 ) 1 2 0 .

We use this inequality to deduce from (4.14) that, passing to the limit as n,

Ω a ( x ) f ( u n ) u n H q ϕ 0 .

On the other hand, since f(0)=0, we have that

Ω a ( x ) f ( u n ) u n H q ϕ = v n > 0 a ( x ) f ( u n H v n ) ( u n H v n ) q v n q ϕ .

Thus, using (2.1) and the fact that un0 in H1(Ω), vnv0 in Lt(Ω), vnv0 a.e. in Ω and v0 is a positive constant, the Lebesgue dominated convergence theorem yields that

Ω a ( x ) f ( u n ) u n H q ϕ f 0 q v 0 q Ω a ( x ) ϕ .

Therefore,

Ω a ( x ) ϕ = 0 .

Since ϕH1(Ω) is arbitrary, we find that a0, which is a contradiction.

(ii) In the Dirichlet case, we argue as in the proof of assertion (i) to prove assertion (ii), by taking λn=λ0=0 therein.

Next, we verify the Neumann case. Assume to the contrary that there exist positive solutions un of (PB) for λ=0 such that unC(Ω¯)0. Then, as in the proof of assertion (i), we may deduce from (1.3) that vn:=un/unH1(Ω)v0 in Lt(Ω) for t<2*, and v0 is a positive constant. Since un is a positive solution of (PB) with λ=0, we obtain Ωb(x)g(un)=0. Recalling (1.16), we see that

0 = Ω b ( x ) g ( u n ) u n H 1 ( Ω ) σ g 0 v 0 σ Ω b ( x ) ,

so that Ωb(x)=0, which contradicts (1.15).

The proof is now complete. ∎

Assuming additionally () with ψ=-a, we can extend Proposition 4.8 (i) to λ<0, and in this case, bifurcation of nontrivial nonnegative solutions of (PB) from (λ,0) can only occur at (0,0).

Corollary 4.9.

Under the assumptions of Proposition 4.8, assume in addition () with ψ=-a. Then the conclusion of Proposition 4.8(i) holds for all λ00. In particular, given δ(0,1), the set of nontrivial nonnegative solutions of (PB) is away from the set {(λ,0);δ|λ|δ-1}.

Proof.

In view of Proposition 4.8, it remains to verify the case λ<0. Assume to the contrary that λnλ0<0, and un are nontrivial nonnegative solutions of (PB) with λ=λn such that un0 in C(Ω¯). Then, we have that

- Δ u n = ( - λ n ) ( - a ( x ) ) f ( u n ) + b ( x ) g ( u n ) in Ω .

By the same arguments used in Lemma 4.7 and Proposition 4.8, we get the desired conclusion. ∎

5 Proofs of Theorems 1.1 and 1.2

If X is a metric space and EnX, then we set

lim inf n E n : = { x X : lim n dist ( x , E n ) = 0 } ,
lim sup n E n : = { x X : lim inf n dist ( x , E n ) = 0 } .

We shall use the following result due to Whyburn [27, (9.12) Theorem]:

Theorem 5.1.

Assume {En} is a sequence of connected sets satisfying that

  1. n 1 E n is precompact;

  2. lim inf n E n .

Then, lim supnEn is nonempty, closed and connected.

As stated in Remark 3.1, we only have to prove Theorem 1.2 (i) in the case Ωa<0.

5.1 Proof of Assertion (i) in Theorems 1.1 and 1.2

When u=u𝐧, we employ the following lemma, which concerns the direction of the bifurcation component 𝒞ε at (0,0), see [23, Theorem 5.1] for the proof.

Lemma 5.2.

Let Bu=un. Assume (1.8), (1.16) and Ωa0. Let Z be any complement of 1 in W2,r(Ω). Then, for ε>0 small enough, the set {(λ,u)} of nontrivial solutions of (PcalBε) around (0,0) is parameterized as

( λ , u ) = ( μ ( s ) , s ( 1 + z ( s ) ) ) ,

with s(-s0,s0) for some s0>0. Here, μ:(-s0,s0)R and z:(-s0,s0)Z are continuous, and satisfy μ(0)=z(0)=0. Therefore, Cε is precisely described by {(μ(s),s(1+z(s))):s[0,s0)} around (0,0). Furthermore, the following holds:

(5.1) lim s 0 + μ ( s ) s σ - 1 = - ε 1 - q q g 0 Ω b ( x ) f 0 Ω a ( x ) .

In particular, under (1.15) and Ωa<0, the bifurcation of Cε is subcritical at (0,0).

Now, we consider the metric space X:=×V with the metric given by

d ( ( λ , u ) , ( μ , v ) ) : = | λ - μ | + u - v V for ( λ , u ) , ( μ , v ) × V .

From Corollary 4.6, if ε(0,1], then the components 𝒞ε± of positive solutions of (PcalBε), emanating from (λ1,ε±,0), satisfy

(5.2) 𝒞 ε ± { ( λ , u ) × V : | λ | + u V C ¯ } ,

where C¯ does not depend on ε(0,1]. This implies that 𝒞ε± are both bounded, and consequently, we deduce that 𝒞ε-=𝒞ε+ (cf. [1, Proposition 18.1]). Then, 𝒞ε:=𝒞ε± is nonempty and connected. In addition,

(5.3) ( 0 , 0 ) lim inf ε 0 + 𝒞 ε ,

since λ1,ε±0 as ε0+. Moreover, by elliptic regularity, we obtain that

(5.4) ε > 0 𝒞 ε is precompact .

Indeed, for any {(λn,un)}ε>0𝒞ε, we have that (λn,un)𝒞εn for some εn(0,1]. From (5.2), we may assume that {λn} is a convergent sequence. Using (5.2) again, we deduce that unW2,r(Ω) are solutions of

{ - Δ u n = λ n a ( x ) ( u n + ε n ) q - 1 F ( u n ) + b ( x ) g ( u n ) in Ω , u n = 0 on Ω .

In particular, using a bootstrap argument and the Sobolev embedding theorem, we deduce that unC1+θ(Ω¯) is bounded for some θ(0,1). The compact embedding C1+θ(Ω¯)C1(Ω¯) implies that {un} has a convergent subsequence in V, as desired. Now, by (5.3) and (5.4), we may apply Theorem 5.1 to infer that 𝒞0:=lim supε0+𝒞ε is nonempty, closed and connected in ×V. From (5.2), 𝒞0 is bounded in ×V. In addition, 𝒞0 is contained in the nonnegative solutions set of (PB). Indeed, given (λ,u)𝒞0, there exists (λn,un)𝒞εn such that εn0+ and (λn,un)(λ,u) in ×V. Thus u is a nonnegative weak solution of (PB), and eventually, a nonnegative solution in W2,r(Ω) by elliptic regularity.

Now, we show that 𝒞0 is nontrivial. By construction, we see that for ε0+, there exists a positive solution uε of (PcalBε) such that (0,uε)𝒞ε. Indeed, we used Lemma 5.2 if u=u𝐧. In this case, we observe from (5.1) that when (1.15) and Ωa<0 hold, the bounded component 𝒞ε bifurcates subcritically at (0,0), provided that ε is small enough. This implies that 𝒞ε cuts {(0,u):0u0}, and consequently, the desired assertion follows. Since uεVC¯, it follows by combining elliptic regularity and standard compactness arguments as above that there exist εn0+ and un:=uεn such that un converges in V to a nonnegative solution u0 of (PB) for λ=0. By definition, we have that (0,u0)𝒞0. From Proposition 4.8 (ii), we infer that u0 is nontrivial, and so, u00 by the strong maximum principle and Hopf’s lemma. Assertion (i) (1) has been now verified. We use Proposition 4.8 (ii) again to deduce assertion (i) (2).

Since 𝒞0 is nontrivial, we infer from Corollary 4.9 that 𝒞0 does not contain any (λ,0) with λ0. Assertion (1.6) has been verified.

Finally, we verify assertion (i) (3). For ρ>0 and (λ1,u1)×V, we set

B ρ ( ( λ 1 , u 1 ) ) : = { ( λ , u ) × V : | λ - λ 1 | + u - u 1 V < ρ } ,
S ρ ( ( λ 1 , u 1 ) ) : = { ( λ , u ) × V : | λ - λ 1 | + u - u 1 V = ρ } .

We note that Bρ((λ1,u1))¯=Bρ((λ1,u1))Sρ((λ1,u1)).

Let Σε+ and Σε- be closed connected subsets of {(λ,u)𝒞ε:λ0} and {(λ,u)𝒞ε:λ0}, respectively, such that (λ1,ε±,0),(0,uε±)Σε± for some positive solutions uε± of (PB) for λ=0, see Figure 3. This is well defined thanks to Proposition 4.8 (ii). Since Σε±𝒞ε, we observe that

Σ 0 ± : = lim sup ε 0 + Σ ε ± lim sup ε 0 + 𝒞 ε = 𝒞 0 .

Repeating the argument above, Whyburn’s topological approach yields that Σ0± are nonempty, closed and connected sets consisting of nonnegative solutions of (PB) and such that (0,0)lim infε0+Σε±Σ0±. Proposition 4.8 (ii) tells us that (0,u0±)Σ0± for some positive solutions u0± of (PB) with λ=0. It follows that Σ0±{(0,0)}, and by virtue of Corollary 4.9, that Σ0±{(0,0)} consists of nontrivial nonnegative solutions of (PB).

By definition, (λ,u)Σ0+ (resp. Σ0-) implies λ0 (resp. λ0). Lastly, by using Proposition 4.8 (ii) again, there exists ρ>0 small such that Σ0,ρ±:=Σ0±Bρ((0,0))¯ is closed and connected, and if (λ,u)Σ0,ρ±{(0,0)}, then λ0. So, 𝒞0±:=Σ0,ρ± have the desired properties. ∎

Figure 3 
            The situations of Σε±{\Sigma_{\varepsilon}^{\pm}}.
Figure 3

The situations of Σε±.

5.2 Proof of Assertion (ii) in Theorems 1.1 and 1.2

We consider the positivity of nontrivial nonnegative solutions of (PB) with f(s)=sq, q(0,1). Let 𝔖 be the nontrivial nonnegative solutions set of (PB), i.e.,

𝔖 : = { ( λ , u ) × V : 0 u 0 solves ( P ) } .

Let be a nonempty connected subset of 𝔖, and let

: = { ( λ , u ) : u 0 } .

The following lemma ([15, Theorem 1.7]) provides us with a nonexistence result for nontrivial nonnegative solutions of (1.12), which plays an important role in our argument when u=u𝐧 and Ωa0.

Lemma 5.3.

Let Bu=un. Assume () with ψ=a. If Ωa0, then there exists qa*(0,1) such that (1.12) has no nontrivial nonnegative solutions for any q(qa*,1).

We give now sufficient conditions for the positivity of the nontrivial nonnegative solutions on as follows. We recall that the sets 𝒜±a are given by (1.13).

Proposition 5.4.

Let f(s)=sq, q(0,1). Suppose (1.14), () with ψ=±a and the condition b0 in Ω. Assume additionally Ωa<0 if Bu=un. If

  1. q 𝒜 a 𝒜 - a (Dirichlet),

  2. q 𝒜 a ( q - a , 1 ) (Neumann),

where q-a is as in Lemma 5.3, then C is open and closed in C. Consequently, C=C if C.

Proof.

It is straightforward that is open in , since u0 for (λ,u). Next, we verify that is closed in . Assume that (λn,un) and (λn,un)(λ0,u0) in ×V. We shall show that (λ0,u0). We discuss the following three cases, in accordance with the sign of λ0:

Case (i): λ0>0. We use the condition q𝒜a to deduce the desired assertion. In this case, λn>0 for sufficiently large n. By the change of variables vn=λn-11-qun, we find that

{ - Δ v n = a ( x ) v n q + λ n - 1 1 - q b ( x ) g ( λ n 1 1 - q v n ) in Ω , v n = 0 on Ω .

Since vnv0=λ0-11-qu0 in V, we find that v0 is a nonnegative weak solution of the problem

{ - Δ v = a ( x ) v q + λ 0 - 1 1 - q b ( x ) g ( λ 0 1 1 - q v ) in Ω , v = 0 on Ω .

In addition, v00 in Ω+a. Indeed, since (1.14) holds and b0, we see that vn is a supersolution of (1.12) which is positive in Ω. So, condition () with ψ=a allows us to apply [15, Lemma 2.2], and deduce that there exist a ball BΩ+a and a continuous function ψ on B¯ such that vnψ>0 in B. Passing to the limit, we have that v0ψ in B, as desired.

By (1.14) and the condition b0, v0 is a supersolution of (1.12) and v0>0 in B. On the other hand, we can construct a nonnegative subsolution ψ0 of (1.12) such that ψ00 in B, ψ00 in ΩB and ψ0v0. The subsolutions and supersolutions method provides us with a solution v1 of (1.12) such that ψ0v1v0, so that v10, since q𝒜a. Consequently, we conclude that u0=λ01/(1-q)v00, as desired.

Case (ii): λ0<0. (1) Case u=u: We use the condition q𝒜-a to deduce the desired assertion. In this case, λn<0 for sufficiently large n. Setting μ=-λ, (PB) turns into

{ - Δ u = - μ a ( x ) u q + b ( x ) g ( u ) in Ω , u = 0 on Ω .

Set μn=-λn>0, so that vn:=μn-11-qun satisfies

{ - Δ v n = - a ( x ) v n q + μ n - 1 1 - q b ( x ) g ( μ n 1 1 - q v n ) in Ω , v n = 0 on Ω .

Since vnv0=μ0-11-qu0 in C01(Ω¯), we infer that v0 is a nonnegative weak solution of the problem

{ - Δ v = - a ( x ) v q + μ 0 - 1 1 - q b ( x ) g ( μ 0 1 1 - q v ) in Ω , v = 0 on Ω .

By using the condition q𝒜-a, the rest of the argument is carried out similarly to the previous case.

(2) Case u=u𝐧: Under q(q-a,1), we shall see that case λ0<0 does not occur, using Lemma 5.3. We have λn<0 for n sufficiently large. By setting μ=-λ, (PB) becomes

{ - Δ u = - μ a ( x ) u q + b ( x ) g ( u ) in Ω , u 𝐧 = 0 on Ω .

Setting μn=-λn>0 and vn=μn-11-qun, we find that

{ - Δ v n = - a ( x ) v n q + μ n - 1 1 - q b ( x ) g ( μ n 1 1 - q v n ) in Ω , v n 𝐧 = 0 on Ω .

Since vnv0=μ0-11-qu0 in C1(Ω¯), we infer that v0 is a nonnegative weak solution of the problem

{ - Δ v = - a ( x ) v q + μ 0 - 1 1 - q b ( x ) g ( μ 0 1 1 - q v ) in Ω , v 𝐧 = 0 on Ω .

In addition, v00 in Ω+-a. Indeed, since (1.14) holds, and b0, we see that vn is a positive supersolution of

(5.5) { - Δ v = - a ( x ) v q in Ω , v 𝐧 = 0 on Ω ,

so that, by [15, Lemma 2.2], there exists a ball BΩ+-a and a continuous function ψ on B¯ such that vnψ>0 in B. Passing to the limit, we obtain v0ψ in B, as desired.

Now, we see that v0 is also a nonnegative supersolution of (5.5) such that v0>0 in B. Since we can construct a nonnegative subsolution ψ0 of (5.5) such that ψ00 in B, ψ00 in ΩB and ψ0v0, the subsolutions and supersolutions method provides us with a solution v1 of (5.5) such that ψ0v1v0 on Ω¯. So, v1 is nontrivial and nonnegative. However, this contradicts Lemma 5.3, since Ω(-a)>0 and q-a*<q<1.

Case (iii): λ0=0. In this case, u0 solves the problem

{ - Δ u 0 = b ( x ) g ( u 0 ) in Ω , u 0 = 0 on Ω .

Since u0 is nontrivial and nonnegative, the strong maximum principle and Hopf’s lemma yield that u00, as desired.

Lastly, since is connected, we conclude that = if . ∎

Introducing the following growth condition on g:

(5.6) 0 < lim s g ( s ) s p < for some p > 1 , where p < N + 1 N - 1 if N > 2 ,

we can deduce that , as shown in (i) and (ii) below.

Remark 5.5.

  1. If (0,u0) with u00, then u00, i.e., (0,u0). Indeed, this is a direct application of the strong maximum principle and Hopf’s lemma.

  2. Assume (1.3) and () with ψ=a. Assume also (1.11) if u=u; and (5.6) and Ωa<0 if u=u𝐧. Let b1 and q𝒜a. If there exist (λn,un) with λn0+, then un0 for sufficiently large n, i.e., (λn,un) for such n. This is an immediate consequence of [15, Theorem 4.1, Theorem 4.5].

  3. When u=u𝐧, Proposition 5.4 is valid for Ωa>0, where we now assume q𝒜-a(qa*,1) instead of q𝒜a(q-a*,1). Indeed, when Ωa>0, the case λ0>0 does not occur, based on Lemma 5.3, whereas the case λ0<0 is verified as in the proof of Proposition 5.4, using λa(x)=(-λ)(-a(x)) and relying on [15, Lemma 2.2].

  4. Proposition 5.4 and item (ii) hold more generally in the framework ×C(Ω¯), which can been seen by using elliptic regularity.

Proof of assertion (ii) in Theorem 1.1.

We note from Remark 2.2 (iii) that (1.14) holds in case (b) of Theorem 1.1. Based on the result stated in Remark 5.5 (i), this assertion is verified by a direct application of Proposition 5.4. ∎

Proof of assertion (ii) in Theorem 1.2.

Based on the result stated in Remark 5.5 (ii), this assertion is straightforward from Proposition 5.4 and Remark 5.5 (iv). Indeed, we do not need to assume () with ψ=-a for applying Proposition 5.4 to the loop 𝒞 given in Theorem 1.2 (ii), since it lies in λ0 (see Figure 1 (ii)). Note that the condition () with ψ=-a is used only for case (ii) in the proof of Proposition 5.4. ∎


Communicated by Julian Lopez Gomez


Award Identifier / Grant number: 30720150100019CB

Award Identifier / Grant number: 1161635

Award Identifier / Grant number: 1171532

Award Identifier / Grant number: 1171691

Award Identifier / Grant number: 15K04945

Award Identifier / Grant number: 18K03353

Funding statement: U. Kaufmann was partially supported by Secyt-UNC 30720150100019CB. H. Ramos Quoirin was supported by FONDECYT grants 1161635, 1171532 and 1171691. K. Umezu was supported by JSPS KAKENHI Grant Numbers 15K04945 and 18K03353.

References

[1] H. Amann, Fixed point equations and nonlinear eigenvalue problems in ordered Banach spaces, SIAM Rev. 18 (1976), no. 4, 620–709. 10.1137/1018114Search in Google Scholar

[2] 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

[3] A. Ambrosetti, H. Brezis and G. Cerami, Combined effects of concave and convex nonlinearities in some elliptic problems, J. Funct. Anal. 122 (1994), no. 2, 519–543. 10.1006/jfan.1994.1078Search in Google Scholar

[4] K. J. Brown, Local and global bifurcation results for a semilinear boundary value problem, J. Differential Equations 239 (2007), no. 2, 296–310. 10.1016/j.jde.2007.05.013Search in Google Scholar

[5] S. Cano-Casanova, Compact components of positive solutions for superlinear indefinite elliptic problems of mixed type, Topol. Methods Nonlinear Anal. 23 (2004), no. 1, 45–72. 10.12775/TMNA.2004.003Search in Google Scholar

[6] S. Cano-Casanova, J. López-Gómez and M. Molina-Meyer, Isolas: Compact solution components separated away from a given equilibrium curve, Hiroshima Math. J. 34 (2004), no. 2, 177–199. 10.32917/hmj/1150998161Search in Google Scholar

[7] S. Cano-Casanova, J. López-Gómez and M. Molina-Meyer, Bounded components of positive solutions of nonlinear abstract equations (in Russian), Ukr. Mat. Visn. 2 (2005), no. 1, 38–51; translation in Ukr. Math. Bull. 2 (2005), 39–52. Search in Google Scholar

[8] S. Cingolani and J. L. Gámez, Positive solutions of a semilinear elliptic equation on 𝐑N with indefinite nonlinearity, Adv. Differential Equations 1 (1996), no. 5, 773–791. 10.57262/ade/1366896019Search in Google Scholar

[9] M. G. Crandall and P. H. Rabinowitz, Bifurcation from simple eigenvalues, J. Funct. Anal. 8 (1971), 321–340. 10.1016/0022-1236(71)90015-2Search in Google Scholar

[10] D. G. De Figueiredo, J.-P. Gossez and P. Ubilla, Local superlinearity and sublinearity for indefinite semilinear elliptic problems, J. Funct. Anal. 199 (2003), no. 2, 452–467. 10.1007/978-3-319-02856-9_36Search in Google Scholar

[11] D. G. de Figueiredo, J.-P. Gossez and P. Ubilla, Multiplicity results for a family of semilinear elliptic problems under local superlinearity and sublinearity, J. Eur. Math. Soc. (JEMS) 8 (2006), no. 2, 269–286. 10.4171/JEMS/52Search in Google Scholar

[12] B. Gidas and J. Spruck, A priori bounds for positive solutions of nonlinear elliptic equations, Comm. Partial Differential Equations 6 (1981), no. 8, 883–901. 10.1080/03605308108820196Search in Google Scholar

[13] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics Math., Springer, Berlin, 2001. 10.1007/978-3-642-61798-0Search in Google Scholar

[14] U. Kaufmann, H. Ramos Quoirin and K. Umezu, A curve of positive solutions for an indefinite sublinear Dirichlet problem, preprint (2017), https://arxiv.org/abs/1709.04822. 10.3934/dcds.2020063Search in Google Scholar

[15] U. Kaufmann, H. Ramos Quoirin and K. Umezu, Positivity results for indefinite sublinear elliptic problems via a continuity argument, J. Differential Equations 263 (2017), no. 8, 4481–4502. 10.1016/j.jde.2017.05.021Search in Google Scholar

[16] U. Kaufmann, H. Ramos Quoirin and K. Umezu, Positive solutions of an elliptic Neumann problem with a sublinear indefinite nonlinearity, NoDEA Nonlinear Differential Equations Appl. 25 (2018), no. 2, Article ID 12. 10.1007/s00030-018-0502-1Search in Google Scholar

[17] J. López-Gómez, Spectral Theory and Nonlinear Functional Analysis, Chapman & Hall/CRC Res. Notes Math. 426, Chapman & Hall/CRC, Boca Raton, 2001. 10.1201/9781420035506Search in Google Scholar

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

[19] J. López-Gómez and M. Molina-Meyer, Bounded components of positive solutions of abstract fixed point equations: Mushrooms, loops and isolas, J. Differential Equations 209 (2005), no. 2, 416–441. 10.1016/j.jde.2004.07.018Search in Google Scholar

[20] 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), no. 3, 503–523. 10.1016/j.jde.2013.04.019Search in Google Scholar

[21] J. López-Gómez and A. Suárez, Combining fast, linear and slow diffusion, Topol. Methods Nonlinear Anal. 23 (2004), no. 2, 275–300. 10.12775/TMNA.2004.012Search in Google Scholar

[22] P. H. Rabinowitz, Some global results for nonlinear eigenvalue problems, J. Funct. Anal. 7 (1971), 487–513. 10.1016/0022-1236(71)90030-9Search in Google Scholar

[23] H. Ramos Quoirin and K. Umezu, An indefinite concave-convex equation under a Neumann boundary condition I, Israel J. Math. 220 (2017), no. 1, 103–160. 10.1007/s11856-017-1512-0Search in Google Scholar

[24] H. Ramos Quoirin and K. Umezu, An indefinite concave-convex equation under a Neumann boundary condition II, Topol. Methods Nonlinear Anal. 49 (2017), no. 2, 739–756. 10.12775/TMNA.2017.007Search in Google Scholar

[25] H. Ramos Quoirin and K. Umezu, A loop type component in the non-negative solutions set of an indefinite elliptic problem, Commun. Pure Appl. Anal. 17 (2018), no. 3, 1255–1269. 10.3934/cpaa.2018060Search in Google Scholar

[26] K. Umezu, Bifurcation approach to a logistic elliptic equation with a homogeneous incoming flux boundary condition, J. Differential Equations 252 (2012), no. 2, 1146–1168. 10.1016/j.jde.2011.08.043Search in Google Scholar

[27] G. T. Whyburn, Topological Analysis, 2nd ed., Princeton Math. Ser. 23, Princeton University Press, Princeton, 1964. 10.1515/9781400879335Search in Google Scholar

Received: 2018-02-17
Revised: 2018-07-17
Accepted: 2018-08-01
Published Online: 2018-09-10
Published in Print: 2019-05-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 26.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2018-2027/html?lang=en
Scroll to top button