Home Existence of Three Positive Solutions for a Nonlocal Singular Dirichlet Boundary Problem
Article Open Access

Existence of Three Positive Solutions for a Nonlocal Singular Dirichlet Boundary Problem

  • Jacques Giacomoni EMAIL logo , Tuhina Mukherjee and Konijeti Sreenadh
Published/Copyright: March 21, 2018

Abstract

In this article, we prove the existence of at least three positive solutions for the following nonlocal singular problem:

{ ( - Δ ) s u = λ f ( u ) u q , u > 0  in  Ω , u = 0 in  n Ω ,

where (-Δ)s denotes the fractional Laplace operator for s(0,1), n>2s, q(0,1), λ>0 and Ω is a smooth bounded domain in n. Here f:[0,)[0,) is a continuous nondecreasing map satisfying

lim u f ( u ) u q + 1 = 0 .

We show that under certain additional assumptions on f, the above problem possesses at least three distinct solutions for a certain range of λ. We use the method of sub-supersolutions and a critical point theorem by Amann [H. Amann, Fixed point equations and nonlinear eigenvalue problems in ordered Banach spaces, SIAM Rev. 18 1976, 4, 620–709] to prove our results. Moreover, we prove a new existence result for a suitable infinite semipositone nonlocal problem which played a crucial role to obtain our main result and is of independent interest.

MSC 2010: 35R11; 35R09; 35A15

1 Introduction

In the present paper, we consider the following nonlocal singular problem:

(Pλ) { ( - Δ ) s u = λ f ( u ) u q , u > 0  in  Ω , u = 0 in  n Ω ,

where s(0,1), q(0,1), λ>0 and Ω is a smooth bounded domain in n. We have the following assumptions on fC1([0,)):

  1. f ( 0 ) > 0 .

  2. lim u f ( u ) u q + 1 = 0 .

  3. u f ( u ) is nondecreasing in +.

  4. There exists a 0<σ1<σ2 such that f(u)uq is nondecreasing on (σ1,σ2).

Note that from (f1) we have limu0f(u)uq=.

Remark 1.1.

For instance, the function f defined by f(t)=eαt/(α+t) for any t0 with α>4q satisfies assumptions (f1)–(f4).

The fractional Laplace operator (-Δ)s is defined by

( - Δ ) s u ( x ) = 2 C s n P . V . n u ( x ) - u ( y ) | x - y | n + 2 s d y ,

where P.V. denotes the Cauchy principal value and

C s n = π - n 2 2 2 s - 1 s Γ ( n + 2 s 2 ) Γ ( 1 - s ) ,

with Γ being the Gamma function. The fractional Laplacian is the infinitesimal generator of Lévy stable diffusion processes and arises in anomalous diffusion in plasma, population dynamics, geophysical fluid dynamics, flames propagation, chemical reactions in liquids and American options in finance; see [5, 16, 36] for instance. Fractional Sobolev spaces were introduced mainly in the framework of harmonic analysis in the middle part of the last century and are the natural setting to study weak solutions to problems involving the fractional Laplacian. In this regard, the paper of Caffarelli and Silvestre [8] on the harmonic extension problem brought to light the subject of nonlocal equations, and has subsequently motivated many works on equations and systems involving the fractional Laplacian (-Δ)s, s(0,1). We refer the reader respectively to [16, 34, 31] for an introduction and surveys about fractional Sobolev spaces and fractional elliptic problems.

In the local case, i.e. s=1, the study of elliptic singular problems starts mainly with the pioneering work of Crandal, Rabinowitz and Tartar [11]. This seminal work inspired a huge list of articles where authors have investigated many different issues (existence/nonexistence, uniqueness/multiplicity, regularity of solutions, etc.) about singular problems in the local and more recently in the nonlocal set up. We cite here some related works with no intent to furnish an exhaustive list. The multiplicity of solutions for singular problems with critical nonlinearity has been studied in [25, 28, 27], while the exponential critical nonlinearity case has been dealt with in [13]. Semilinear elliptic and singular problems with convection term were first studied in [19], whereas [15] brought existence results to elliptic equations involving a singular absorption term. Hölder regularity of weak solutions is discussed in [24]. We refer to the surveys [20, 26] for further details on singular elliptic equations in the local setting. In the nonlocal case, singular problems with critical nonlinearity have been studied in [7, 22, 32]. Recently, Adimurthi, Giacomoni and Santra [2] studied the following nonlocal singular problem:

( - Δ ) s u = λ ( K ( x ) u - δ + f ( u ) ) , u > 0  in  Ω , u = 0  in  n Ω ,

where δ,λ>0, K:Ω+ is a Hölder continuous function in Ω that behaves like dist(x,Ω)-β, β[0,2s), and f is a positive real-valued C2 function. They established existence, regularity and bifurcation results using the framework of weighted spaces.

Recently, Düzgün and Iannizzotto [17] have established the existence of three non-zero solutions for a Dirichlet-type boundary value problem involving the fractional Laplacian. But the study of three solutions for singular nonlocal problems was completely open till now. Our work brings new results in this regard. We use the method of sub- and supersolutions combined with a fixed point theorem due to Amann to achieve the objective. For the construction of the barrier functions, we have taken some ideas from [29].

The salient feature of this work is the presence of the singular term u-q which is a primary hindrance in making the operator monotone. We slightly transform the problem to a new one and show that the operator associated with it becomes monotone increasing and compact. This idea had been formerly used by Dhanya, Ko and Shivaji [14] in the local case. But here itself we remark that their approach can not be directly applied to problem () due to the presence of the nonlocal operator (-Δ)s instead of Δ. Most substantially, [14, Theorem 3.6] can not be adapted here due to the lack of an explicit form of (-Δ)sδs(x), where δ(x)=dist(x,Ω) denotes the distance function up to the boundary. To overcome this difficulty, we construct a subsolution v satisfying

( - Δ ) s v + c v δ q ( x ) 0 in  Ω r ,

where c>0 is constant and ΩrΩ. To obtain this, we separately study, by bifurcation arguments, a nonlocal infinite semipositone problem () in Section 6. This leads naturally to a solution of the required problem. In the local setting, we refer the readers to [33, 30, 23] concerning infinite semipositone problems. But we indicate that a “nonlocal” infinite semipositone problem has not been studied in the past. So our results are completely new in this regard. The main result of our paper is accomplished by using a well-known critical point theorem by Amann [3]. Last but not the least, we additionally prove the uniqueness of solutions to () when λ becomes sufficiently large under appropriate conditions of f. This result is motivated by the paper [9] of Castro, Eunkyung and Shivaji. Nevertheless, we point out that their approach can not be exactly applied here due to the presence of the nonlocal operator (-Δ)s. We still succeed to obtain the result by an appropriate application of Hardy’s inequality for the fractional Laplacian (see Section 5). Now we state the main results of our paper as follows.

Theorem 1.2.

There exist constants λ1,λ2>0 such that if λ[λ1,λ2], then problem () has at least three solutions in Cϕ1,s+(Ω).

Remark 1.3.

Theorem 1.2 still holds if (f3) is replaced by the weakened assumption (f3’): there exists k>0 such that uf(u)+ku is nondecreasing in +.

Theorem 1.4.

There exists a λ*>0 such that () has a unique solution when λ>λ*.

Remark 1.5.

Since f(0)>0, it is not difficult to show that for λ>0 small enough there exists a unique solution with small norm. Then Theorems 1.2 and 1.4 entail that the bifurcation curve of solutions to () emanating from (0,0) is S-shaped.

The outline of this paper is as follows: In Section 2, we give some useful preliminaries about the main equation in (). In Section 3, we construct the sub- and supersolutions used to apply the fixed point theorem of Amann. In Section 4, we prove the main result of our paper: Theorem 1.2. In Section 5, we prove our main uniqueness result: Theorem 1.4. Finally, in Section 6 we investigate the fractional and singular semipositone problem () used crucially for proving the strong increasingness of the operator T in Section 4.

2 Preliminaries

We start with defining the function spaces. Given any ϕC0(Ω¯) such that ϕ>0 in Ω, we define

C ϕ ( Ω ) := { u C 0 ( Ω ) : there is  c 0  such that  | u ( x ) | c ϕ ( x )  for all  x Ω } ,

with the usual norm uϕL(Ω) and the associated positive cone. We define the following open convex subset of Cϕ(Ω):

C ϕ + ( Ω ) := { u C ϕ ( Ω ) : inf x Ω u ( x ) ϕ ( x ) > 0 } .

In particular, Cϕ+ contains all functions uC0(Ω) with k1ϕuk2ϕ in Ω for some k1,k2>0. We consider the following fractional Sobolev space:

H ~ s ( Ω ) := { u H s ( n ) : u = 0  in  n Ω }

equipped with the norm

u = ( Q | u ( x ) - u ( y ) | 2 | x - y | n + 2 s d x d y ) 1 2 , where  Q = 2 n ( 𝒞 Ω × 𝒞 Ω ) .

Definition 2.1.

We say that uH~s(Ω) is a weak solution to () if infKu>0 for every compact subset KΩ, and for any φH~s(Ω),

Ω ( - Δ ) s u φ = C s n Q ( u ( x ) - u ( y ) ) ( φ ( x ) - φ ( y ) ) | x - y | n + 2 s d x d y = λ Ω f ( u ) u q φ d x .

Definition 2.2.

By a subsolution of problem (), we mean a function vH~s(Ω) which satisfies (weakly)

(2.1) ( - Δ ) s v λ f ( v ) v q , v > 0  in  Ω , v = 0  in  n Ω .

Whereas if the reverse inequality holds in (2.1), we call v a supersolution of (Pλ). Also we call respectively v a strict sub- and supersolution if the inequality in (2.1) is strict.

We define the distance function as δ(x):=dist(x,Ω), xΩ. Let ϕ1,s denote the first positive eigenfunction of (-Δ)s in H~s(Ω) corresponding to its principal eigenvalue λ1,s such that ϕ1,sL(Ω)=1. We recall that ϕ1,,sCs(n) and also ϕ1,sCδs+(Ω) (see for instance [35, Proposition 1.1 and Theorem 1.2]).

3 Sub- and Supersolutions of ()

In this section, we show the existence of two pairs of sub-supersolutions (ζ1,ϑ1) and (ζ2,ϑ2) such that ζ1ζ2ϑ1, ζ1ϑ2ϑ1 and ζ2⩽̸ϑ2. Moreover, it holds that ζ2,ϑ2 are strict sub- and supersolutions of (). Let w denote the unique solution of the problem

(3.1) ( - Δ ) s w = 1 w q , w > 0  in  Ω , w = 0  in  n Ω .

Then from the proof of [2, Theorem 1.1] we know that wH~s(Ω)Cϕ1,s+(Ω) and wCs(n). We construct our supersolution ϑ1 first. Since (f2) holds, we get

lim u f ( u ) u q + 1 = 0 .

This implies that if we choose a constant Mλ1 sufficiently large such that

f ( M λ w L ( Ω ) ) ( M λ w L ( Ω ) ) q + 1 1 λ w L ( Ω ) q + 1 , i.e. M λ q + 1 λ f ( M λ w L ( Ω ) ) ,

then ϑ1=MλwH~s(Ω)Cϕ1,s+(Ω) forms a supersolution of (). Indeed, using the nondecreasing nature of f, we get

( - Δ ) s ϑ 1 = M λ q + 1 ( M λ w ) q λ f ( M λ w L ( Ω ) ) ( M λ w ) q λ f ( M λ w ) ( M λ w ) q = λ f ( ϑ 1 ) ( ϑ 1 ) q .

Now since limu0f(u)uq=, we can choose mλ>0 sufficiently small so that

λ 1 , s m λ ϕ 1 , s λ f ( m λ ϕ 1 , s ) ( m λ ϕ 1 , s ) q for each  λ > 0 .

Now we define ζ1=mλϕ1,sH~s(Ω)Cϕ1,s+(Ω), and it is easy to see that

( - Δ ) s ζ 1 = m λ λ 1 , s ϕ 1 , s λ f ( m λ ϕ 1 , s ) ( m λ ϕ 1 , s ) q = λ f ( ζ 1 ) ζ 1 q .

Therefore, ζ1 is a subsolution of (Pλ). It is not hard to see that we always choose mλ small enough so that ζ1ϑ1. This completes our construction of the first pair of sub-supersolution.

Our next step is to construct the second pair of sub-supersolution of (). We first construct our positive supersolution ϑ2 such that ϑ2L(Ω)=σ1 (see (f4)). Let us define

ϑ 2 = σ 1 w w L ( Ω ) H ~ s ( Ω ) C ϕ 1 , s + ( Ω )

and assume that

0 < λ σ 1 q + 1 f ( σ 1 ) w L ( Ω ) q + 1 .

Then using the nondecreasing nature of f, we find that it satisfies

( - Δ ) s ϑ 2 = σ 1 w L ( Ω ) w q λ f ( σ 1 ) w L ( Ω ) q ( σ 1 w ) q λ f ( σ 1 w w L ( Ω ) ) ( σ 1 w w L ( Ω ) ) q = λ f ( ϑ 2 ) ϑ 2 q .

Now we construct our second positive supersolution of (Pλ), which is one of the crucial part of our paper. For this, we let σ(0,σ1] be such that f*(σ)=min0<xσf(x)xq, and also define hC([0,)) such that

h ( u ) = { f * ( σ ) if  u σ , f ( u ) u q if  u σ 1 ,

so that h is a nondecreasing function on (0,σ1] and h(u)f(u)uq for all u0. With this definition of h, we consider the nonsingular problem

(Qλ) ( - Δ ) s u = λ h ( u )  in  Ω , u = 0  in  n Ω .

Let Gs(x,y) denote the Green function associated to (-Δ)s with homogeneous Dirichlet boundary conditions in Ω. Then we have

u ( x ) = { λ Ω G s ( x , y ) h ( y ) d y if  x Ω , 0 if  x n Ω .

Let BR^(0) denote the ball (centered at 0, where without loss of generality we assume that 0Ω) of largest radius R^ that is inscribed in Ω, and also let R<R^. Suppose K:L2(Ω)L2(Ω) to be the linear map defined by

K ( g ) ( x ) = Ω G s ( x , y ) g ( y ) d y .

Let χR:Ω be the characteristic function defined by

χ R ( x ) = { 1 if  x B R ( 0 ) , 0 if  x Ω B R ( 0 ) .

Then from [10, Theorem 1.1], for each (x,y)Ω×Ω we have that

(3.2) 0 G s ( x , y ) C min { δ s ( x ) δ s ( y ) | x - y | n , δ s ( x ) | x - y | n - s } .

Therefore, there exists a constant C1>0 depending on R such that

K ( χ R ) ( x ) = Ω G s ( x , y ) d y C δ s ( x ) B R ( 0 ) d y | x - y | n - s C 1

uniformly in xΩ. Let M2=(minxΩK(χR)(x))-1>0 and a(σ1,σ2]. Also we define v=aχR in Ω. Then if v1 denotes a solution to the problem

(3.3) ( - Δ ) s v 1 = h ( v )  in  Ω , v 1 = 0  in  n Ω ,

then for each xΩ we get

v 1 ( x ) = λ Ω G s ( x , y ) h ( a χ R ) ( y ) d y λ h ( a ) Ω G s ( x , y ) d y ,

where we used the fact that h is nondecreasing and χR1 in Ω. Therefore, v1σ2 in BR(0) if

λ M 3 σ 2 h ( a ) , where  M 3 = ( max x Ω Ω G s ( x , y ) d y ) - 1 < + .

Claim: v1v in Ω for a certain range of λ. Let xΩBR(0). Since Gs(x,y)0 for all x,yΩ and h(u)>0 for all u0, we get v(x)=0v1(x). Now let xBR(0). Then

v 1 ( x ) = λ Ω G s ( x , y ) h ( a χ R ) ( y ) d y
= λ ( B R ( 0 ) G s ( x , y ) h ( a ) d y + Ω B R ( 0 ) G s ( x , y ) h ( 0 ) d y )
λ h ( a ) B R ( 0 ) G s ( x , y ) d y (since  h ( 0 ) = f ( σ ) σ q > 0  and  G s ( x , y ) 0 )
λ h ( a ) M 2 .

So if we assume that λM2ah(a), then by the definition of v we get

v 1 ( x ) a v ( x ) for all  x Ω .

Hence we finally get that 0vv1σ2 in Ω if

M 2 a h ( a ) λ M 3 σ 2 h ( a ) .

Since we assumed h to be nondecreasing in (0,σ2) (because of (f4)), we get h(v(x))h(v1(x)) for xΩ. So v1 weakly satisfies the problem

( - Δ ) s v 1 λ h ( v 1 )  in  Ω , v 1 = 0  in  n Ω .

Since h(v)L(Ω), using [35, Theorem 1.2], we get that v1Cs(n). This gives us that v1 forms a subsolution of (). Moreover, (3.3) and the strong maximum principle imply that v1>0 in Ω. Therefore, using the fact that h(u)f(u)uq for all u0 implies that ζ2=v1 forms a positive subsolution of (Pλ).

Therefore, we got that if

λ 1 := M 2 a h ( a ) λ min { σ 1 q + 1 f ( σ 1 ) w L ( Ω ) q + 1 , M 3 σ 2 h ( a ) } = : λ 2 ,

then we obtain a positive subsolution ζ2 and a positive supersolution ϑ2 of () such that ζ2ϑ2. Indeed, ϑ2L(Ω)=σ1 and ζ2L(Ω)a>σ1.

4 Proof of the Main Result

In this section, we prove our main result after establishing some necessary results. We begin by noticing that our problem () can be rewritten as

(P’λ) ( - Δ ) s u - λ f ( 0 ) u q = f ~ ( u ) , u > 0  in  Ω , u = 0  in  n Ω ,

where f~(u)=λ(f(u)-f(0)uq). We fix that λ[λ1,λ2]. Since fC1([0,)), by the mean value theorem we get f~(u)=λf(v)u1-q for some v(0,u). Also this implies f~(0)=0 because limt0|f(t)|< and q(0,1). Therefore, f~ can be considered as a continuous function on [0,) such that f~(0)=0. We assume also the following:

  1. There exists a constant k~>0 such that f~(t)+k~t is increasing in [0,).

Definition 4.1.

We say that zH~s(Ω) is a weak solution of (P’λ) if infKz>0 for every compact subset KΩ, and for any φCc(Ω),

(4.1) C s n Q ( z ( x ) - z ( y ) ) ( φ ( x ) - φ ( y ) ) | x - y | n + 2 s d x d y - λ f ( 0 ) Ω φ z q d x = Ω f ~ ( z ) φ d x .

We remark that for any φH~s(Ω) and z(x)k1δs(x) in Ω, Hardy’s inequality gives that

| Ω φ z q d x | k 1 Ω | φ ( x ) | δ s ( x ) δ s ( 1 - q ) ( x ) d x k 1 ( Ω | φ ( x ) | 2 δ 2 s ( x ) ) 1 2 ( Ω δ 2 s ( 1 - q ) ( x ) d x ) 1 2 C φ ,

where k1,C>0 are constants. So now following the arguments of [22, Lemma 3.2], we can prove that a weak solution zH~s(Ω) of (P’λ) satisfies (4.1) for every φH~s(Ω) if zk1δs(x) in Ω.

We extend the functions f and f~ naturally as f(t)=f(0) and f~(t)=f~(0) for t0. Because of the assumption in (h1), without loss of generality we can assume that f~ is increasing in +. Now we define the map T:C0(Ω¯)Cϕ1,s(Ω) by T(u)=z if and only if z is a weak solution of

(Sλ) ( - Δ ) s z - λ f ( 0 ) z q = f ~ ( u ) , z > 0  in  Ω , z = 0  in  n Ω .

By saying that zH~s(Ω) is a weak solution of () we mean that it satisfies

(4.2) C s n Q ( z ( x ) - z ( y ) ) ( φ ( x ) - φ ( y ) ) | x - y | n + 2 s d x d y - λ f ( 0 ) Ω φ z q d x = Ω f ~ ( u ) φ d x

for all φCc(Ω). But repeating the arguments from above for (P’λ), we can show that zH~s(Ω) of () satisfies (4.2) for every φH~s(Ω) if zk1δs(x) in Ω.

Proposition 4.2.

A function zH~s(Ω)Cϕ1,s(Ω) is a weak solution of () if and only if z is a fixed point of T.

Proof.

Suppose zH~s(Ω)Cϕ1,s(Ω) is a weak solution of (). Then it is clear that z forms a fixed point of the map T. Conversely assume T(z)=z for some zC0(Ω¯). Then it satisfies (4.2), but it remains to show that zCϕ1,s+(Ω). Since z>0 in Ω and f~(z) is locally Ho¨lder continuous in Ω, we can follow the arguments of [2, Theorem 1.2] to obtain that zCϕ1,s+(Ω). ∎

Proposition 4.3.

The map T is well defined from C0(Ω¯) to Cϕ1,s+(Ω).

Proof.

Let uC0(Ω¯) and v=f~u. Then vC0(Ω¯) and v0 in Ω. To show that T is a well-defined map, we need to show that () has a unique solution corresponding to the above u. We introduce the following approximated problem for ϵ>0:

(Sλϵ) ( - Δ ) s z - λ f ( 0 ) ( z + ϵ ) q = f ~ ( u ) , z > 0  in  Ω , z = 0  in  n Ω .

Then (Sλϵ) has a unique solution in H~s(Ω). Indeed, let H~s(Ω)+ denote the positive cone of H~s(Ω) and define the energy functional Eϵ:H~s(Ω)+ by

E ϵ ( z ) = z 2 2 - λ f ( 0 ) 1 - q Ω ( z + ϵ ) 1 - q d x - Ω f ~ ( u ) z d x ,

where zH~s(Ω)+. Then Eϵ is weakly lower semi-continuous, strictly convex and coercive in H~s(Ω)+. Therefore, Eϵ admits a unique minimizer, say zϵ0, in H~s(Ω)+. Since for small t>0 the term t1-qΩ(z+ϵ)1-qdx dominates, Eϵ(tz) can be made small enough and we get infH~s(Ω)+Eϵ<0. We choose m>0 (independent of ϵ) sufficiently small such that

m λ 1 , s ϕ 1 , s v + λ f ( 0 ) ( m ϕ 1 , s + 1 ) q in  Ω .

Then we get that

(4.3) ( - Δ ) s ( m ϕ 1 , s ) = m λ 1 , s ϕ 1 , s v + λ f ( 0 ) ( m ϕ 1 , s + 1 ) q v + λ f ( 0 ) ( m ϕ 1 , s + ϵ ) q in  Ω .

Claim (1): mϕ1,szϵ for each ϵ>0. We define z~ϵ:=(mϕ1,s-zϵ)+ and assume that meas(supp(z~ϵ)) is non-zero. Then η:[0,1] defined by η(t)=Eϵ(zϵ+tz~ϵ) is a convex function since Eϵ|H~s(Ω)+ is convex. Also,

η ( 1 ) = C s n Q ( ( z ϵ + z ~ ϵ ) ( x ) - ( z ϵ + z ~ ϵ ) ( y ) ) ( z ~ ϵ ( x ) - z ~ ϵ ( y ) ) | x - y | n + 2 s d x d y - λ Ω f ( 0 ) z ~ ϵ ( z ϵ + z ~ ϵ + ϵ ) q - Ω f ~ ( u ) z ~ ϵ

in (0,1]. The fact that zϵ is a minimizer of Eϵ gives that limt0+η(t)0 and 0η(0)η(1). Let us recall the following inequality for any ψ being a convex Lipschitz function:

( - Δ ) s ψ ( u ) ψ ( u ) ( - Δ ) s u .

Therefore, using this with ψ(x)=max{x,0} and (4.3), we get η(1)Eϵ(mϕ1,s),z~ϵ<0, which is a contradiction. Hence supp(z~ϵ) must have measure zero, which establishes Claim (1). Thus Eϵ is Gâteaux differentiable at zϵ and zϵ satisfies (Sλϵ) weakly. Since

f ~ ( u ) + λ f ( 0 ) ( z ϵ + ϵ ) q L ( Ω ) for each  ϵ > 0 ,

from [35, Proposition 1.1 and Theorem 1.2] and Claim (1) we get that zϵCs(n)Cϕ1,s+(Ω) for each ϵ>0. Thus following the arguments in the proof of [2, Theorem 1.1, p. 7], we can show that {zϵ}ϵ>0 is a monotone increasing sequence as ϵ0+, that is, for 0<ϵ<ϵ there holds zϵ<zϵ in Ω. Thus we infer that z=limϵ0+zϵmϕ1,s. From zϵ satisfying (Sλϵ) we obtain

(4.4) z ϵ 2 = λ Ω f ( 0 ) z ϵ ( z ϵ + ϵ ) q d x + Ω f ~ ( u ) z ϵ d x .

We recall the function wH~s(Ω)Cϕ1,s+(Ω) satisfying (3.1). Let z¯=Mw for M1 (independent of ϵ) sufficiently large so that

M ( 1 w q - λ f ( 0 ) ( M w ) q ) > f ~ ( u ) in  Ω .

Then z¯ satisfies

( - Δ ) s z ¯ - λ f ( 0 ) ( z ¯ + ϵ ) q = M w q - λ f ( 0 ) ( M w + ϵ ) q > M ( 1 w q - λ f ( 0 ) ( M w ) q ) > f ~ ( u ) in  Ω .

Now we prove that zϵz¯ by using a comparison argument, which we will refer as comparison principle in the future. We know that h=(zϵ-z¯)H~s(Ω) satisfies the equation

(4.5) ( Δ ) s ( z ϵ - z ¯ ) λ f ( 0 ) ( 1 ( z ϵ + ϵ ) q - 1 ( z ¯ + ϵ ) q ) in  Ω .

If we denote h+=max{h,0} and h-=-min{h,0}, then h=h+-h-. Let Ωh+={xΩ:ze>z¯} and Ωh-=ΩΩh+. Then testing (4.5) with h+ gives

(4.6) C s n Q ( h ( x ) - h ( y ) ) ( h + ( x ) - h + ( y ) ) | x - y | n + 2 s d x d y λ f ( 0 ) Ω h + ( 1 ( z ϵ + ϵ ) q - 1 ( z ¯ + ϵ ) q ) h + d x .

It is easy to see that

( h ( x ) - h ( y ) ) ( h + ( x ) - h + ( y ) ) = h ( x ) h + ( x ) 0 on  Ω × 𝒞 Ω ,
( h ( x ) - h ( y ) ) ( h + ( x ) - h + ( y ) ) 0 on  Ω h + × Ω h - .

This gives

0 < Ω h + Ω h + ( h ( x ) - h ( y ) ) ( h + ( x ) - h + ( y ) ) | x - y | n + 2 s d x d y Q ( h ( x ) - h ( y ) ) ( h + ( x ) - h + ( y ) ) | x - y | n + 2 s d x d y .

Therefore, from (4.6) we obtain

0 < C s n Ω h + Ω h + ( h ( x ) - h ( y ) ) ( h + ( x ) - h + ( y ) ) | x - y | n + 2 s d x d y λ f ( 0 ) Ω h + ( 1 ( z ϵ + ϵ ) q - 1 ( z ¯ + ϵ ) q ) h + d x 0 .

Hence it must be that zϵz¯ in Ω for each ϵ>0. Now we use this in (4.4) and Hölder’s inequality to get

z ϵ 2 λ f ( 0 ) Ω z ¯ 1 - q d x + f ~ ( u ) L 2 ( Ω ) z ¯ L 2 ( Ω ) = : m 0 < + ,

which implies lim supϵ>0zϵ<+. Thus {zϵ}ϵ>0 is a bounded sequence in H~s(Ω), and so there must exist a zH~s(Ω) such that, up to a subsequence, zϵz weakly in H~s(Ω) as ϵ0. We already know that zϵz pointwise a.e. in Ω. Moreover, by Hardy’s inequality, for any φH~s(Ω) we get

0 < | φ ( z ϵ + ϵ ) q | | φ ( m ϕ 1 , s ) q | L 1 ( Ω ) .

Therefore, we can use the Lebesgue dominated convergence theorem to pass through the limit as ϵ0+ in (Sλϵ) to obtain

C s n Q ( z ( x ) - z ( y ) ) ( φ ( x ) - φ ( y ) ) | x - y | n + 2 s d x d y - λ f ( 0 ) Ω φ z q d x = Ω f ~ ( u ) φ d x ,

that is, z is a weak solution of (). Finally, it remains to show that zCϕ1,s+(Ω). But this easily follows from z¯zzϵmϕ1,s in Ω. Thus, T is well defined, and this completes the proof. ∎

Before proving our next result, we recall [12, Theorem 1.2] as follows.

Theorem 4.4.

Let cLloc1(Ω) be a nonpositive function and let uHs(Ω) be a weak supersolution of

( - Δ ) s u = c ( x ) u in  Ω .

  1. If Ω is bounded and u0 a.e. in 𝒞Ω, then either u>0 a.e. in Ω or u=0 a.e. in n.

  2. If u 0 a.e. in n , then either u > 0 a.e. in Ω or u=0 a.e. in n.

Lemma 4.5.

The map T is strictly monotone increasing from C0(Ω¯) to Cϕ1,s+(Ω).

Proof.

First we show that T is monotone increasing. For this, we let u1,u2C0(Ω¯) be such that u1u2. Then f~(u1)f~(u2) since f~ is increasing. Now let zi=T(ui) for i=1,2. Then each zi satisfies

( - Δ ) s z i - λ f ( 0 ) z i q = f ~ ( u i ) , z i > 0  in  Ω , z i = 0  in  n Ω ,

and z^:=(z2-z1)H~s(Ω) satisfies

(4.7) ( - Δ ) s ( z 2 - z 1 ) - λ f ( 0 ) ( 1 z 2 q - 1 z 1 q ) = f ~ ( u 2 ) - f ~ ( u 1 ) 0 in  Ω .

Then testing (4.7) with z^- gives

(4.8) Q ( z ^ ( x ) - z ^ ( y ) ) ( z ^ - ( x ) - z ^ - ( y ) ) | x - y | n + 2 s d x d y λ f ( 0 ) { z 2 < z 1 } ( 1 z 2 q - 1 z 1 q ) z ^ - d x .

It is easy to see that the right-hand side of (4.8) is nonpositive, and the left-hand side can be estimated as

Q ( z ^ ( x ) - z ^ ( y ) ) ( z ^ - ( x ) - z ^ - ( y ) ) | x - y | n + 2 s d x d y - { z 2 < z 1 } { z 2 < z 1 } | z ^ - ( x ) - z ^ - ( y ) | 2 | x - y | n + 2 s d x d y 0 .

Therefore, it must be that z-=0 in Ω, that is, z2z1 in Ω. Now we assume that u2u1 and u1u2. Then we show that z2>z1 in Ω. We already know that z2z1, and by the mean value theorem we get that there exists a ξ(z1,z2) such that (4.7) can be written as

(4.9) ( - Δ ) s ( z 2 - z 1 ) + λ f ( 0 ) ( q ξ q + 1 ) ( z 2 - z 1 ) = f ~ ( u 2 ) - f ~ ( u 1 ) 0 in  Ω .

Let c(x)=1/ξq+1(x). Since ξ(z1,z2) and ziCϕ1,s+(Ω) for i=1,2, we easily get that cLloc1(Ω). Therefore, from Theorem 4.4 we obtain that z2-z1>0 in Ω, that is, T is a strictly monotone increasing map. ∎

The proof of our next result is motivated by the proof of [2, Lemma 4.3].

Proposition 4.6.

The map T:Cϕ1,s(Ω)Cϕ1,s(Ω) is compact.

Proof.

Let uCϕ1,s(Ω) and T(u)=zCϕ1,s(Ω). Then z solves (). We can write z as

z = ( Δ ) - s ( λ f ( 0 ) z q ) + ( - Δ ) - s ( f ~ ( u ) ) in  Ω .

Let {uk}kCϕ1,s(Ω) be a bounded sequence, that is,

sup k u k δ s L ( Ω ) < + and T ( u k ) = z k C ϕ 1 , s ( Ω )

for each k. Then we have

z k = ( Δ ) - s ( λ f ( 0 ) z k q ) + ( - Δ ) - s ( f ~ ( u k ) ) in  Ω .

From the proof of Proposition 4.3 we infer that mϕ1,s and Mw form a sub- and supersolution, respectively, of () for an appropriate choice of positive constants m and M (independent of k). Then by the weak comparison principle we get that

(4.10) m ϕ 1 , s z k M w , that is, k 1 δ s ( x ) z k ( x ) k 2 δ s ( x )  in  Ω

for some constants k1,k2>0. In order to prove compactness of the map T, we need to show that the sequence {zk} is relatively compact in Cϕ1,s+(Ω). Since {uk} is bounded in Cϕ1,s(Ω), we get f~(uk)L(Ω) and supkf~(uk)L(Ω)C1 for some constant C1>0. Therefore, from [35, Theorem 1.2] we obtain

(4.11) ( - Δ ) - s f ~ ( u k ) δ s C 0 , α ( Ω ) C f ~ ( u k ) L ( Ω ) C 2

for some constants C,C2>0 (independent of k) and 0<α<min{s,1-s}. Now for fixed ϵ>0, we define the set

D ϵ := { x Ω : δ ( x ) ϵ }

and let χDϵ denote the corresponding characteristic function on Dϵ. We also define the following functions:

z k 1 , ϵ := ( Δ ) - s ( λ f ( 0 ) χ D ϵ z k q ) + ( - Δ ) - s ( f ~ ( u k ) ) ,
z k 2 , ϵ := ( ( Δ ) - s ( λ f ( 0 ) ( 1 - χ D ϵ ) z k q ) ) χ D 3 ϵ ,
z k 3 , ϵ := ( ( Δ ) - s ( λ f ( 0 ) ( 1 - χ D ϵ ) z k q ) ) ( 1 - χ D 3 ϵ ) .

Clearly, zk=zk1,ϵ+zk2,ϵ+zk3,ϵ. Therefore, it is enough to prove that each {zki,ϵ} for i=1,2,3 is relatively compact in Cϕ1,s(Ω). Because of (4.10) we have

λ f ( 0 ) χ D ϵ z k q λ f ( 0 ) χ D ϵ k 1 δ s q ( x ) λ f ( 0 ) k 1 ϵ s q ,

which implies

sup k λ f ( 0 ) χ D ϵ z k q L ( Ω ) C 3 = C 3 ( ϵ )

for some constant C3>0. So from (4.11) and [35, Theorem 1.2] we infer that

z k 1 , ϵ δ s C 0 , α ( Ω ) C λ f ( 0 ) χ D ϵ z k q L ( Ω ) + ( - Δ ) - s f ~ ( u k ) δ s C 0 , α ( Ω ) C 4 = C 4 ( ϵ )

for some constant C4>0. Thus {zk1,ϵ} is relatively compact in Cϕ1,s(Ω) for each fixed ϵ>0. Considering the sequence {zk2,ϵ}, for any x,xD3ϵ we get

| z k 2 , ϵ ( x ) δ s ( x ) - z k 2 , ϵ ( x ) δ s ( x ) | = λ | Ω ( G s ( x , y ) δ s ( x ) - G s ( x , y ) δ s ( x ) ) f ( 0 ) ( 1 - χ D ϵ ) ( y ) z k q ( y ) d y |
C | Ω ( G s ( x , y ) δ s ( x ) - G s ( x , y ) δ s ( x ) ) ( 1 - χ D ϵ ) ( y ) k 1 δ s q ( y ) d y |

for some constant C>0. It has been proved in [2, Lemma 4.3] that the map xGs(x,y)δs(x) is Hölder continuous in D3ϵ uniformly with respect to yΩDϵ (but still depending on ϵ). This implies that there exists Cϵ>0 constant such that Gs(x,y)Cs(D3ϵ)Cϵ uniformly with respect to yΩDϵ. Therefore, we finally get that

| z k 2 , ϵ ( x ) δ s ( x ) - z k 2 , ϵ ( x ) δ s ( x ) | C ~ ϵ | x - x | s Ω D ϵ 1 δ s q ( y ) d y C ^ ϵ | x - x | s

for some constant C~ϵ,C^ϵ>0. This clearly gives that {zk2,ϵ} is relatively compact in Cϕ1,s(Ω). Lastly, we consider the sequence {zk3,ϵ} and fix β(sq,s). Recalling estimate (3.2) for Gs(x,y), for xΩD3ϵ we get

| z k 3 , ϵ ( x ) δ s ( x ) | = | λ f ( 0 ) δ s ( x ) n G s ( x , y ) ( 1 - χ D ϵ ) ( y ) δ s ( x ) z k q ( y ) d y |
| λ f ( 0 ) δ s ( x ) n D ϵ min ( δ s ( x ) δ s ( y ) | x - y | n , δ s ( x ) | x - y | n - s ) 1 k 1 δ s q ( y ) d y |
λ f ( 0 ) ϵ β - s q n D ϵ min ( δ s ( y ) | x - y | n , 1 | x - y | n - s ) 1 k 1 δ β ( y ) d y
(4.12) O ( ϵ β - s q ) .

Now we show that {zkδs} is relatively compact in L(Ω). Let τ>0 be small enough. Then because of (4.12) we can always choose ϵ small enough such that

z k 3 , ϵ δ s L ( Ω ) τ .

For each such ϵ>0, we can get a convergent subsequences {zkm1,ϵ} and {zkm2,ϵ} of {zk1,ϵ} and {zk2,ϵ}, respectively, in L(Ω) since they are relatively compact in Cϕ1,s(Ω). Hence we have

z k m δ s - z k m δ s L ( Ω ) z k m 1 , ϵ δ s - z k m 1 , ϵ δ s L ( Ω ) + z k m 2 , ϵ δ s - z k m 2 , ϵ δ s L ( Ω ) + 2 τ 4 τ

when m,mK for some K. This implies that {zkm} is a Cauchy sequence in Cϕ1,s(Ω), and hence convergent too. This proves that the sequence {zk} is relatively compact in Cϕ1,s(Ω). ∎

We seek the help of a solution to a nonlocal infinite semipositone problem (discussed in Section 6) for proving our next result, that is, the map T is strongly increasing. By strongly increasing we mean that if u1u2 and u1u2, then T(u2)-T(u1)Cϕ1,s+(Ω).

Theorem 4.7.

The map T:Cϕ1,s(Ω)Cϕ1,s(Ω) is strongly increasing.

Proof.

Let u1u2 such that u1u2 and T(ui)=zi for i=1,2. Then from Lemma 4.5 we already know that z1>z2 in Ω and z2-z1Cϕ1,s(Ω). Therefore, it remains to prove that there exist a k1>0 such that k1δs(x)(z2-z1)(x) in Ω. We know that (z2-z1) satisfies (4.9) and since z1(x)ξ(x)z2(x) in Ω and each ziCϕ1,s(Ω), we can get a constant k>0 such that

λ f ( 0 ) q ξ q + 1 k δ s ( q + 1 ) ( x ) in  Ω .

Therefore, if we set z~=(z2-z1), then we obtain

(4.13) ( - Δ ) s z ~ + k δ s ( q + 1 ) ( x ) z ~ 0 , z ~ > 0  in  Ω , z ~ = 0  in  n Ω .

From Theorem 6.2 we know that for sufficiently small θ>0 there exists a vCϕ1,s+(Ω) which satisfies weakly

( - Δ ) s v = v p - θ v γ , v > 0  in  Ω , v = 0  in  n Ω ,

where γ(q,1) and p(0,1). So there exist constants m1,m2>0 such that m1δs(x)v(x)m2δs(x) in Ω. From this we get

(4.14) ( - Δ ) s v + k v δ s ( q + 1 ) ( x ) = v p - θ v γ ( x ) + k v δ s ( q + 1 ) ( x ) v p - m 2 - γ θ δ s γ ( x ) + m 2 k δ s q ( x ) in  Ω .

Since γ(q,1), the term m2θδsγ(x) dominates near the boundary of Ω. We define Ωη={xΩ:δ(x)<η} and choose η>0 small enough so that (4.14) gives

(4.15) ( - Δ ) s v + k v δ s ( q + 1 ) ( x ) 0 , v > 0  in  Ω η , v = 0  in  n Ω .

From (4.13) we have

(4.16) ( - Δ ) s z ~ + k z ~ δ s ( q + 1 ) ( x ) 0 , z ~ > 0  in  Ω η , z ~ = 0  in  n Ω .

We choose m3>0 small enough such that m3vz~ in ΩΩη. Thus (4.15) and (4.16) give

( - Δ ) s ( m 3 v - z ~ ) + k ( m 3 v - z ~ ) δ s ( q + 1 ) ( x ) 0 in  Ω η , ( m 3 v - z ~ ) 0  in  n Ω η .

By the comparison principle, we get m1m3δs(x)m3vz~ in Ωη. Since 0<z~Cϕ1,s(Ω), we get infxΩΩηz~>0. Hence there must exist a constant k1>0 such that k1ϕ1,s(x)z~ in Ω. This proves that (z2-z1)Cϕ1,s+(Ω), and the map T is strongly increasing on Cϕ1,s(Ω). ∎

We recall a fixed point theorem by Amann [3] which will help us to get the desired result.

Theorem 4.8.

Let X be a retract of some Banach space and let f:XX be a compact map. Suppose that X1 and X2 are disjoint subsets of X and let Uk, k=1,2, be open subsets of X such that UkXk, k=1,2. Moreover, suppose that f(Xk)Xk and that f has no fixed points on XkUk, k=1,2. Then f has at least three distinct fixed points x1,x2,x3 with xkXk, k=1,2, and xX(X1X2).

We also recall [3, Corollary 6.2].

Lemma 4.9.

Let X be an ordered Banach space and let [y1,y2] be an ordered interval in X. Let f:[y1,y2]X be an increasing compact map such that f(y1)y1 and f(y2)y2. Then f has a minimal fixed point x¯ and a maximal fixed point x¯.

Now the proof of the main result goes as follows.

Proof of Theorem 1.2.

To obtain solutions of (), or equivalently (P’λ), it is enough to find fixed points of the map T, thanks to Proposition 4.2. We define the sets X=[ζ1,ϑ1], X1=[ζ1,ϑ2] and X2=[ζ2,ϑ1]. Since X and the Xi, for each i=1,2, are nonempty closed and convex subsets of Cϕ1,s(Ω), they form retracts of Cϕ1,s(Ω). By construction (done in Section 2), we know that X1X2= in X. Since ζ1 and ϑ1 are ordered sub- and supersolutions of (), respectively, and T is strictly increasing (Lemma 4.5), by the comparison principle we obtain

ζ 1 T ( ζ 1 ) T ( ϑ 1 ) ϑ 1 .

This implies that T(X)X, and similarly it also holds that T(Xk)Xk for k=1,2. Because of Proposition 4.6 and Theorem 4.7, we get that T:XX is compact and a strongly increasing map. It has been proved that ϑ2 is a strict supersolution of (Pλ) and T(ϑ2)ϑ2. So using Theorem 4.4, we infer that T(ϑ2)<ϑ2, T(ϑ1)<ϑ1, T(ζ1)>ζ1 and T(ζ2)<ζ2. Therefore, Lemma 4.9 implies that T has a maximal fixed point u1X1 such that u1(ζ1,ϑ2), and a minimal fixed point u2X2 such that u2(ζ2,ϑ1). Now repeating the arguments in Theorem 4.7, we can prove that there exist constants a1,a2>0 such that

u 1 a 1 ϕ 1 , s + ζ 1 , ϑ 2 - u 1 a 1 ϕ 1 , s , u 2 + a 2 ϕ 1 , s ϑ 1 , u 2 - ζ 2 a 2 ϕ 1 , s    in  Ω .

We define the open ball B in X by

B := X { φ C ϕ 1 , s ( Ω ) : φ ϕ 1 , s L ( Ω ) < a } with  a = min ( a 1 , a 2 ) .

Then for each i=1,2, we have ui+BXi, and thus the Xi have nonempty interior. We construct open balls around each fixed point of T in Xi for each i=1,2 and take Ui as the largest open set in Xi containing all these open balls and such that XiUi contains no fixed point of T. Now applying Theorem 4.8, we get the existence of a third fixed point u3 of T lying in X(X1X2). This completes the proof. ∎

5 Uniqueness for Large λ

In this section, we prove that () admits a unique solution when λ is sufficiently large. We assume only that fC1([0,)) satisfies (f1)–(f3) and

  1. there exists an α>0 such that f(u)uq is decreasing for u>α.

Theorem 5.1.

Problem () admits a solution for each λ>0.

Proof.

We recall the first pair of sub-supersolution (ζ1,ϑ1) of () constructed in Section 2. Without loss of generality, we can assume that f is nondecreasing in [ζ1,v1]. As in the proof of Theorem 1.2, we can show that if X=[ζ1,ϑ1], then T:XX is strictly increasing and a compact map and that T(X)X. Therefore, we can apply Lemma 4.9 to conclude that T has a fixed point uλ in X. Then Proposition 4.2 gives us that uλCϕ1,s+(Ω) is a solution of (), which completes the proof. ∎

Lemma 5.2.

Any positive solution uλ in Cϕ1,s+(Ω) of () satisfies uλΘλw in Ω, where Θλ=(λf(0))1/(1+q) and w is the solution to (3.1).

Proof.

Let uλ solve (). Since f is nondecreasing, we get

( - Δ ) s u λ ( x ) = λ f ( u λ ( x ) ) u λ q ( x ) λ f ( 0 ) u λ q ( x ) in  Ω ,

and

( - Δ ) s ( Θ λ w ) ( x ) = λ f ( 0 ) ( Θ λ w ) q ( x ) in  Ω .

Therefore, by the weak comparison principle (see Lemma 4.5), we conclude that uλ-Θλw0 in n. ∎

Corollary 5.3.

There exists a minimal solution of () in Cϕ1,s+(Ω) for each λ>0.

Proof.

From Theorem 5.1 we know that () has a solution uλCϕ1,s+(Ω) such that

ζ 1 u λ ϑ 1 in  Ω ,

where both ζ1 and ϑ1 are in Cϕ1,s+(Ω). Now the result follows from Lemma 5.2 and Lemma 4.9. ∎

Theorem 5.4.

There exists a λ*>0 such that () has a unique solution when λ>λ*.

Proof.

Let uλ and u¯λ be two distinct positive solutions of () in Cϕ1,s+(Ω) such that uλ is the minimal solution as obtained from Corollary 5.3. So it holds that uλu¯λ in Ω. We have

Q ( - Δ ) s ( u λ - u ¯ λ ) ( u λ - u ¯ λ ) d x = λ Ω ( f ( u λ ) u λ q - f ( u ¯ λ ) u ¯ λ q ) ( u λ - u ¯ λ ) d x ,

which gives

C s n ( u λ - u ¯ λ ) 2 = λ Ω ( f ( u λ ) u λ q - f ( u ¯ λ ) u ¯ λ q ) ( u λ - u ¯ λ ) d x
(5.1) = λ Ω ( 0 1 f 0 ( u λ + t ( u ¯ λ - u λ ) ) d t ) ( u λ - u ¯ λ ) 2 d x ,

where f0(u)=f(u)uq. From Lemma 5.2 we know that uλΘλwΘλk1δs(x) in Ω for some k1>0. So if we define

Ω 0 := { x Ω : f 1 ( x ) 0 } { x Ω : δ s ( x ) α Θ λ k 1 } ,

where f1(x)=01f0(uλ+t(u¯λ-uλ))dt, then uλ(x)α in Ω0. From (5.1) we obtain

C s n ( u λ - u ¯ λ ) 2 = λ ( Ω 0 f 1 ( x ) ( u λ - u ¯ λ ) 2 d x + Ω Ω 0 f 1 ( x ) ( u λ - u ¯ λ ) 2 d x ) .

Since f10 in ΩΩ0,

C s n ( u λ - u ¯ λ ) 2 λ Ω 0 f 1 ( x ) ( u λ - u ¯ λ ) 2 d x .

We also notice that limλ+δ(x)=0 for xΩ0. Since by (f5) we have uλ+t(u¯λ-uλ)(x)α for xΩ0, there exists an M2>0 such that |f((uλ+t(u¯λ-uλ))(x))|M2 for all xΩ0. Therefore, we also have the following estimate by using Hardy’s inequality:

λ Ω 0 f 1 ( x ) ( u λ - u ¯ λ ) 2 ( x ) d x λ Ω 0 ( 0 1 f ( ( u λ + t ( u ¯ λ - u λ ) ) ( x ) ) ( u λ + t ( u ¯ λ - u λ ) ) q ( x ) d t ) ( u λ - u ¯ λ ) 2 ( x ) d x
λ M 2 Ω 0 ( u λ - u ¯ λ ) 2 ( x ) u λ q ( x ) d x
λ M 2 ( Θ λ k 1 ) - q Ω 0 ( u λ - u ¯ λ ) 2 ( x ) δ ( 2 - q ) s ( x ) δ 2 s ( x ) d x (since  u λ Θ λ w )
λ M 2 ( Θ λ k 1 ) - q ( σ Θ λ k 1 ) 2 - q Ω 0 ( u λ - u ¯ λ ) 2 ( x ) δ 2 s ( x ) d x
C ( λ ) ( u λ - u ¯ λ ) 2 ,

where C(λ)=O(λ-(1-q)/(1+q)). This gives a contradiction for λ large enough since q(0,1). Therefore, we state that uλu¯λ when λ is sufficiently large, and this completes the proof. ∎

6 A Fractional and Singular Semipositone Problem

We devote this section to prove the existence of a weak solution for the following nonlocal infinite semipositone problem:

(Iθ) ( - Δ ) s v = v p - θ v γ , v > 0  in  Ω , v = 0  in  n Ω ,

where p(0,1), θ is a positive parameter and γ(q,1). Before this, we consider the problem

(I0) ( - Δ ) s v = v p , v > 0  in  Ω , v = 0  in  n Ω

for p(0,1). The energy functional E0:H~s(Ω) associated to (I0) is given by

E 0 ( v ) := C s n v 2 2 - 1 p + 1 Ω | v | p + 1 d x

for vH~s(Ω). Then E0 is weakly lower semicontinuous and coercive, which implies that E0 possesses a global minimizer, say v0H~s(Ω). Since infH~s(Ω)E0<0 and E0(|v0|)E0(v0), we get v00 and we can assume that v00 in Ω. Now it is easy to see that v0 solves problem (I0) weakly. By [6, Proposition 2.2], we obtain v0L(Ω), and thus, using [35, Theorem 1.2], we conclude that v0Cs(n)Cϕ1,s(Ω). Now by the strong maximum principle it follows that v0>0 in Ω. For η>0 small enough, ηϕ1,s forms a subsolution of (I0), and then it is easy to show by the weak comparison principle that v0Cϕ1,s+(Ω). The uniqueness of v0 as a solution of (I0) follows by using the Picone identity [4, Lemma 6.2] and following the arguments in [21, Theorem 5.2].

For fixed μ>0, let us consider the solution operator G(θ,v):{|θ|<μ}×Bϵ(v0)Cϕ1,s(Ω) defined by

G ( θ , v ) := v - ( - Δ ) - s ( v p - θ v γ )

for (θ,v){|θ|<μ}×Bϵ(v0), where Bϵ(v0) denotes the open ball in Cϕ1,s+(Ω) with center v0 and radius ϵ>0. We point out that for ϵ>0 small enough, Bϵ(v0)Cϕ1,s+(Ω). Furthermore, G(θ,u)=0 if and only if u solves (). Let (θ,v){|θ|<μ}×Bϵ(v0). Then (-Δ)-svpCϕ1,s(Ω) by [35, Theorem 1.2]. Since vL(Ω), we have that (-Δ)-svpCϕ1,s(Ω). Moreover, (-Δ)-s(θvγ)Cϕ1,s(Ω) follows from [2, Theorem 1.2] and the fact that vCϕ1,s+(Ω). Thus the map G(,) is well defined.

Lemma 6.1.

The map G is continuously Fréchet differentiable.

Proof.

We begin with showing that G is continuous. Let v,vkBϵ(v0), |θ|<μ and τn be such that (vk-vCϕ1,s(Ω)+|τ|)0 as k. Then

| G ( θ + τ , v k ) - G ( θ , v ) | = | ( v k - v ) - ( - Δ ) - s ( v k q - v q ) + ( - Δ ) - s ( θ + τ v k γ - θ v γ ) |
( v k - v ) C ϕ 1 , s ( Ω ) ϕ 1 , s + C 1 ( v k q - v q ) C ϕ 1 , s ( Ω ) ϕ 1 , s
+ θ | ( - Δ ) - s ( γ ( v k - v ) ( v + ξ ( v k - v ) ) γ + 1 + τ v k γ ) |
( v k - v ) C ϕ 1 , s ( Ω ) ϕ 1 , s ( x ) + C 1 ( v k - v ) C ϕ 1 , s ( Ω ) ϕ 1 , s
+ C 3 θ | ( - Δ ) - s ( γ ( v k - v ) C ϕ 1 , s ( Ω ) δ s γ ( x ) + τ δ s γ ( x ) ) |

for appropriate constants Ci>0, i=1,2,3. Now [1, Proposition 1.2.9] gives that

| G ( θ + τ , v k ) - G ( θ , v ) | O ( v k - v C ϕ 1 , s ( Ω ) + | τ | ) ϕ 1 , s ,

which implies that G is continuous on {|θ|<μ}×Bϵ(v0). Following similar arguments, we can show that

lim t 0 + G ( θ , v + t ϕ ) - G ( θ , v ) t = ϕ - p ( - Δ ) - s ( v p - 1 ϕ ) - θ ( - Δ ) - s ( γ v - γ - 1 ϕ )

for v,ϕBϵ(v0). This implies that G(θ,) is Gâteaux differentiable and

D v G ( θ , v ) ( ϕ ) = ϕ - p ( - Δ ) - s ( v p - 1 ϕ ) - θ ( - Δ ) - s ( γ v - γ - 1 ϕ ) .

Next, to prove that G is Fréchet differentiable we first consider

| G ( θ , v + ϕ ) - G ( θ , v ) - D v G ( θ , v ) ( ϕ ) |
= | - ( - Δ ) - s ( ( v + ϕ ) p - v p - p v p - 1 ϕ ) + θ ( - Δ ) - s ( 1 ( v + ϕ ) γ - 1 v γ + γ ϕ v γ + 1 ) | .

Since vCϕ1,s+(Ω) and ϕCϕ1,s(Ω), we get |ϕv|K for some constant K0. This along with Taylor series expansion gives for some θ0(0,1),

| ( v + ϕ ) p - v p - p v p - 1 ϕ = p ( p - 1 ) ϕ 2 2 ( v + θ 0 ϕ ) 2 - p | C ϕ C ϕ 1 , s ( Ω ) 2 .

So applying [35, Theorem 1.2], we get that

| ( - Δ ) - s ( ( v + ϕ ) p - v p - p v p - 1 ϕ ) | O ( ϕ C ϕ 1 , s ( Ω ) 2 ) .

Also a similar idea gives, for some ξ1(0,1),

1 ( v + ϕ ) γ - 1 v γ + γ ϕ v γ + 1 = - γ ϕ ( v + ξ 1 ϕ ) γ + 1 + γ ϕ v γ + 1 = γ ϕ ( ( v + ξ 1 ϕ ) γ + 1 - v γ + 1 ( v ( v + ξ 1 ϕ ) ) γ + 1 ) C 4 ϕ C ϕ 1 , s ( Ω ) 2 δ s γ

for an appropriate constant C4>0. So again using [1, Proposition 1.2.9], we get

| ( - Δ ) - s ( 1 ( v + ϕ ) γ - 1 v γ + γ ϕ v γ + 1 ) | O ( ϕ C ϕ 1 , s ( Ω ) 2 ) .

Therefore, we get

G ( θ , v + ϕ ) - G ( θ , v ) - D v G ( θ , v ) ( ϕ ) C ϕ 1 , s ( Ω ) 0 as  ϕ C ϕ 1 , s ( Ω ) 0 .

Now we prove the continuity of DvG(θ,v). Consider {vk}kBϵ(v0) such that vk-vCϕ1,s(Ω)0 as k. Then

D v G ( θ , v k ) - D v G ( θ , v ) = sup 0 ϕ C ϕ 1 , s ( Ω ) ( D v G ( θ , v k ) - D v G ( θ , v ) ) ϕ C ϕ 1 , s ( Ω ) ϕ C ϕ 1 , s ( Ω ) .

We have that

( D v G ( θ , v k ) - D v G ( θ , v ) ) ϕ = - p ( - Δ ) - s ( ( v k p - 1 - v p - 1 ) ϕ ) - γ θ ( - Δ ) - s ( ϕ v k γ + 1 - ϕ v γ + 1 ) .

But using ϕCϕ1,s(Ω) and again using arguments similar to before, we get

| ( D v G ( θ , v k ) - D v G ( θ , v ) ) ϕ | O ( ϕ C ϕ 1 , s ( Ω ) ( v k p - 1 - v p - 1 C ϕ 1 , s ( Ω ) + v k - v C ϕ 1 , s ( Ω ) ) ) ,

which implies that DvG(θ,v) is continuous. Similarly, we can prove that Dθ(θ,v) exists and is continuous, where Dθ(θ,v)=(-Δ)-s(1vγ). ∎

The linearization of the map G with respect to the second variable at (θ,v){|θ|<μ}×Bϵ(v0) given by vG(θ,v):Cϕ1,s+(Ω)Cϕ1,s(Ω) is defined by

v G ( θ , v ) h = h - ( - Δ ) - s ( p v p - 1 h + θ γ h u γ + 1 ) for  h C ϕ 1 , s ( Ω ) .

Since v0 solves (I0), clearly G(0,v0)=0. Now we will show that the map hvG(0,v0)h is invertible. To do this, we will be studying an eigenvalue problem. Let us define

Λ := inf 0 u H ~ s ( Ω ) ( u 2 - p Ω v 0 p - 1 u 2 d x Ω u 2 d x ) = inf S ( u 2 - p Ω v 0 p - 1 u 2 d x ) ,

where S={uH~s(Ω):Ωu2dx=1}. Then by Hardy’s inequality and v0Cϕ1,s+(Ω) it follows that

Ω v 0 p - 1 u 2 d x k 2 Ω u 2 δ 2 s ( x ) δ s ( p + 1 ) d x < + .

So the functional

I 00 ( u ) = u 2 - p Ω v 0 p - 1 u 2 d x

is well defined on H~s(Ω). Following standard minimization arguments and using the compact embedding of H~s(Ω) in L2(Ω), it is easy to show that infI00(S)=I00(ψ)=Λ for some ψS. Also since I00(|ψ|)I00(ψ), by minimality of ψ we assert that without loss of generality we may assume that ψ0. Then ψ satisfies

(6.1) ( - Δ ) s ψ = Λ ψ + p v 0 p - 1 ψ in  Ω , ψ = 0  in  n Ω ,

which implies that ψ is an eigenfunction corresponding to the eigenvalue Λ for the operator (-Δ)s-pv0p-1 with homogeneous Dirichlet boundary condition in Ω. We obtain ψL(Ω) by following the arguments in [18, Theorem 3.2, p. 379]. So local regularity results assert that ψClocs(Ω). We claim that ψ>0 in Ω because if this would not be true, then there would exist a x0KΩ such that ψ(x0)=0. But this gives

0 > 2 C s n n ( ψ ( x 0 ) - ψ ( y ) ) | x - y | n + 2 s d y = Λ ψ + p v 0 p - 1 ( x 0 ) ψ ( x 0 ) = 0 in  K ,

which is a contradiction. Therefore, ψ>0 in Ω.

Claim (1): Λ is a principal eigenvalue. We have to show that any eigenfunction (say ψ) associated to it does not change sign. Assume by contradiction that ψ+0 and ψ-0. Then

C s n Q ( ψ ( x ) - ψ ( y ) ) ( ψ + ( x ) - ψ + ( y ) ) | x - y | n + 2 s d y d x
= C s n ( Q ( ψ + ( x ) - ψ + ( y ) ) ( ψ + ( x ) - ψ + ( y ) ) | x - y | n + 2 s d y d x + 2 Q ψ - ( x ) ψ + ( y ) | x - y | n + 2 s d y d x )
= Λ Ω ψ ψ + d x + p Ω v 0 p - 1 ψ ψ + d x .

Now if ψ+,ψ-0, then

ψ + 2 - p Ω v 0 p - 1 ( ψ + ) 2 d x = Λ Ω ( ψ + ) 2 d x - 2 Q ψ - ( x ) ψ + ( y ) | x - y | n + 2 s d y d x < Λ Ω ( ψ + ) 2 d x ,

but this contradicts the definition of Λ. This proves the claim.

Claim (2): Λ is the unique principal eigenvalue. Suppose not, that is, let Λ0 be another principal eigenvalue of (-Δ)s-pv0p-1 and let ψ0H~s(Ω) denote the corresponding eigenfunction. Then ψ0>0 in Ω, and it satisfies

(6.2) ( - Δ ) s ψ 0 = Λ 0 ψ 0 + p v 0 p - 1 ψ 0  in  Ω , ψ 0 = 0  in  n Ω .

Testing (6.1) with ψ0 and (6.2) with ψ gives

Λ Ω ψ ψ 0 d x = Λ 0 Ω ψ 0 ψ d x .

Since ψ,ψ0>0 in Ω, we get Λ=Λ0.

Claim (3): Any nonnegative eigenfunction ψ is in Cϕ1,s+(Ω). For sufficiently small η>0, by using Theorem 4.4 and the Hopf Lemma, it is easy to show that ψηϕ1,s. Assume first that Λ0. Then

ψ = Λ ( - Δ ) - s ψ + p ( - Δ ) - s ( ψ v 0 1 - p ) in  Ω .

Since ψL(Ω), we get (-Δ)-sψCϕ1,s(Ω). Also since v0Cϕ1,s+(Ω), we get that

ψ v 0 1 - p C δ s ( 1 - p ) ( x ) in  Ω

for some C>0. So

( - Δ ) - s ( ψ v 0 1 - p ) C ϕ 1 , s + ( Ω )

follows from [1, Proposition 1.2.9] together with s(1-p)<s. Therefore, ψCϕ1,s+(Ω) when Λ0. In the other case, if Λ<0, then ψ satisfies

( - Δ ) s ψ + ( - Λ ) ψ = p ( ψ v 0 1 - p ) C δ s ( 1 - p ) ( x ) 0 in  Ω .

So by using [12, Theorem 1.5 (1)] and [35, Theorem 1.2], we infer that ψCϕ1,s+(Ω). This proves the claim.

Claim (4): Λ>0. First we show that Λ is non-zero. Suppose it is equal to zero. Then (6.1) reduces to

( - Δ ) s ψ = p v 0 p - 1 ψ in  Ω .

Using v0 as a test function in the expression above gives

n ( - Δ ) s ψ v 0 d x = p Ω v 0 p ψ d x .

But we know that v0 is a unique solution of (I0). Therefore, we get

p Ω v 0 p ψ d x = Ω v 0 p ψ d x ,

which gives p=1 since ψ,v0>0 in Ω. This gives a contradiction, and thus Λ0. Now we assume by contradiction that Λ<0. For ϵ>0 small enough, we consider the function v0-ϵψ. Then since Λ<0 and p-1<0, we get that

( - Δ ) s ( v 0 - ϵ ψ ) = v 0 p - ϵ Λ ψ - ϵ p v 0 p - 1 ψ > v 0 p - ϵ p v 0 p - 1 ψ ( v 0 - ϵ ψ ) p in  Ω .

This implies that v0-ϵψ forms a strict supersolution of (I0). It is already known that ηϕ1,s for a sufficiently small choice of η forms a subsolution of (I0). Therefore, there must be a function v^H~s(Ω) such that ηϕ1,sv^(v0-ϵψ), which is a solution of (I0). But this contradicts the uniqueness of v0 due to ϵψ>0 in Ω. Hence Λ>0.

Theorem 6.2.

For a small range of θ, problem () admits a solution.

Proof.

We already proved that G is a continuously Fréchet differentiable map. Now we consider the problem

(6.3) ( - Δ ) s u - p v 0 p - 1 u = ψ , u > 0  in  Ω , u = 0  in  n Ω .

By defining the energy functional corresponding to it and by minimization arguments, it is easy to show that the above problem has a solution uH~s(Ω). Now suppose u1,u2H~s(Ω) are two distinct solutions of (6.3). Then

0 = n ( - Δ ) s ( u 1 - u 2 ) ( u 1 - u 2 ) d x - p Ω v 0 p - 1 ( u 1 - u 2 ) 2 d x Λ Ω ( u 1 - u 2 ) 2 d x > 0

since Λ>0. This implies that the solution must be unique. Also using an argument similar to that in Claim (3) gives uCϕ1,s(Ω). All this, along with the previous claims guarantees that the map

v G ( 0 , v 0 ) : C ϕ 1 , s ( Ω ) C ϕ 1 , s ( Ω )

is invertible. Hence we apply the implicit function theorem to get that there exist a subset

{ | θ | < μ } × B ϵ ( v 0 ) { | θ | < μ } × B ϵ ( v 0 )

for μ<μ and ϵ<ϵ, and a C1-map h:{|θ|<μ}Bϵ(v0) such that G(θ,v)=0 in {|θ|<μ}×Bϵ(v0) coincides with the graph of h. This completes the proof. ∎


Communicated by Laurent Veron


Funding statement: The authors were partially funded by IFCAM (Indo-French Centre for Applied Mathematics) UMI CNRS 3494 under the project “Singular phenomena in reaction diffusion equations and in conservation laws”.

References

[1] N. Abatangelo, Large S-harmonic functions and boundary blow-up solutions for the fractional Laplacian, Discrete Contin. Dyn. Syst. 35 (2015), no. 12, 5555–5607. 10.3934/dcds.2015.35.5555Search in Google Scholar

[2] A. Adimurthi, J. Giacomoni and S. Santra, Positive solutions to a fractional equation with singular nonlinearity, preprint (2017), https://arxiv.org/abs/1706.01965. 10.1016/j.jde.2018.03.023Search in Google Scholar

[3] 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

[4] S. Amghibech, On the discrete version of Picone’s identity, Discrete Appl. Math. 156 (2008), no. 1, 1–10. 10.1016/j.dam.2007.05.013Search in Google Scholar

[5] D. Applebaum, Lévy processes—from probability to finance and quantum groups, Notices Amer. Math. Soc. 51 (2004), no. 11, 1336–1347. Search in Google Scholar

[6] B. Barrios, E. Colorado, R. Servadei and F. Soria, A critical fractional equation with concave-convex power nonlinearities, Ann. Inst. H. Poincaré Anal. Non Linéaire 32 (2015), no. 4, 875–900. 10.1016/j.anihpc.2014.04.003Search in Google Scholar

[7] B. N. Barrios, I. De Bonis, M. Medina and I. Peral, Semilinear problems for the fractional Laplacian with a singular nonlinearity, Open Math. 13 (2015), 390–407. 10.1515/math-2015-0038Search in Google Scholar

[8] L. Caffarelli and L. Silvestre, An extension problem related to the fractional Laplacian, Comm. Partial Differential Equations 32 (2007), no. 7–9, 1245–1260. 10.1080/03605300600987306Search in Google Scholar

[9] A. Castro, E. Ko and R. Shivaji, A uniqueness result for a singular nonlinear eigenvalue problem, Proc. Roy. Soc. Edinburgh Sect. A 143 (2013), no. 4, 739–744. 10.1017/S0308210512000418Search in Google Scholar

[10] Z.-Q. Chen and R. Song, Estimates on Green functions and Poisson kernels for symmetric stable processes, Math. Ann. 312 (1998), no. 3, 465–501. 10.1007/s002080050232Search in Google Scholar

[11] M. G. Crandall, P. H. Rabinowitz and L. Tartar, On a Dirichlet problem with a singular nonlinearity, Comm. Partial Differential Equations 2 (1977), no. 2, 193–222. 10.1080/03605307708820029Search in Google Scholar

[12] L. M. Del Pezzo and A. Quaas, A Hopf’s lemma and a strong minimum principle for the fractional p-Laplacian, J. Differential Equations 263 (2017), no. 1, 765–778. 10.1016/j.jde.2017.02.051Search in Google Scholar

[13] R. Dhanya, J. Giacomoni, S. Prashanth and K. Saoudi, Global bifurcation and local multiplicity results for elliptic equations with singular nonlinearity of super exponential growth in 2, Adv. Differential Equations 17 (2012), no. 3–4, 369–400. 10.57262/ade/1355703090Search in Google Scholar

[14] R. Dhanya, E. Ko and R. Shivaji, A three solution theorem for singular nonlinear elliptic boundary value problems, J. Math. Anal. Appl. 424 (2015), no. 1, 598–612. 10.1016/j.jmaa.2014.11.012Search in Google Scholar

[15] J. I. Diaz, J.-M. Morel and L. Oswald, An elliptic equation with singular nonlinearity, Comm. Partial Differential Equations 12 (1987), no. 12, 1333–1344. 10.1080/03605308708820531Search in Google Scholar

[16] E. Di Nezza, G. Palatucci and E. Valdinoci, Hitchhiker’s guide to the fractional Sobolev spaces, Bull. Sci. Math. 136 (2012), no. 5, 521–573. 10.1016/j.bulsci.2011.12.004Search in Google Scholar

[17] F. Düzgün and A. Iannizzotto, Three nontrivial solutions for nonlinear fractional Laplacian equations, Adv. Nonlinear Anal. (2016), 10.1515/anona-2016-0090. 10.1515/anona-2016-0090Search in Google Scholar

[18] G. Franzina and G. Palatucci, Fractional p-eigenvalues, Riv. Math. Univ. Parma (N.S.) 5 (2014), no. 2, 373–386. Search in Google Scholar

[19] M. Ghergu and V. D. Rădulescu, Multi-parameter bifurcation and asymptotics for the singular Lane–Emden–Fowler equation with a convection term, Proc. Roy. Soc. Edinburgh Sect. A 135 (2005), no. 1, 61–83. 10.1017/S0308210500003760Search in Google Scholar

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

[21] J. Giacomoni, T. Mukherjee and K. Sreenadh, Existence and stabilization results for a singular parabolic equation involving the fractional Laplacian, Discrete Contin. Dyn. Syst. Ser. S, to appear; https://arxiv.org/abs/1709.01906. 10.3934/dcdss.2019022Search in Google Scholar

[22] J. Giacomoni, T. Mukherjee and K. Sreenadh, Positive solutions of fractional elliptic equation with critical and singular nonlinearity, Adv. Nonlinear Anal. 6 (2017), no. 3, 327–354. 10.1515/anona-2016-0113Search in Google Scholar

[23] J. Goddard, II, E. K. Lee, L. Sankar and R. Shivaji, Existence results for classes of infinite semipositone problems, Bound. Value Probl. 2013 (2013), Article ID 97. 10.1186/1687-2770-2013-97Search in Google Scholar

[24] C. Gui and F.-H. Lin, Regularity of an elliptic problem with a singular nonlinearity, Proc. Roy. Soc. Edinburgh Sect. A 123 (1993), no. 6, 1021–1029. 10.1017/S030821050002970XSearch in Google Scholar

[25] Y. Haitao, Multiplicity and asymptotic behavior of positive solutions for a singular semilinear elliptic problem, J. Differential Equations 189 (2003), no. 2, 487–512. 10.1016/S0022-0396(02)00098-0Search in Google Scholar

[26] J. Hernández and F. J. Mancebo, Singular elliptic and parabolic equations, Handbook of Differential Equations: Stationary Partial Differential Equations. Vol. III, Elsevier, Amsterdam (2006), 317–400. 10.1016/S1874-5733(06)80008-2Search in Google Scholar

[27] N. Hirano, C. Saccon and N. Shioji, Existence of multiple positive solutions for singular elliptic problems with concave and convex nonlinearities, Adv. Differential Equations 9 (2004), no. 1–2, 197–220. 10.57262/ade/1355867973Search in Google Scholar

[28] N. Hirano, C. Saccon and N. Shioji, Brezis–Nirenberg type theorems and multiplicity of positive solutions for a singular elliptic problem, J. Differential Equations 245 (2008), no. 8, 1997–2037. 10.1016/j.jde.2008.06.020Search in Google Scholar

[29] E. Ko, E. K. Lee and R. Shivaji, Multiplicity results for classes of infinite positone problems, Z. Anal. Anwend. 30 (2011), no. 3, 305–318. 10.4171/ZAA/1436Search in Google Scholar

[30] E. K. Lee, R. Shivaji and J. Ye, Positive solutions for infinite semipositone problems with falling zeros, Nonlinear Anal. 72 (2010), no. 12, 4475–4479. 10.1016/j.na.2010.02.022Search in Google Scholar

[31] G. Molica Bisci, V. D. Radulescu and R. Servadei, Variational Methods for Nonlocal Fractional Problems, Encyclopedia Math. Appl. 162, Cambridge University Press, Cambridge, 2016. 10.1017/CBO9781316282397Search in Google Scholar

[32] T. Mukherjee and K. Sreenadh, Critical growth fractional elliptic problem with singular nonlinearities, Electron. J. Differential Equations 54 (2016), 1–23. Search in Google Scholar

[33] M. Ramaswamy, R. Shivaji and J. Ye, Positive solutions for a class of infinite semipositone problems, Differential Integral Equations 20 (2007), no. 12, 1423–1433. 10.57262/die/1356039073Search in Google Scholar

[34] X. Ros-Oton, Nonlocal elliptic equations in bounded domains: A survey, Publ. Mat. 60 (2016), no. 1, 3–26. 10.5565/PUBLMAT_60116_01Search in Google Scholar

[35] X. Ros-Oton and J. Serra, The Dirichlet problem for the fractional Laplacian: Regularity up to the boundary, J. Math. Pures Appl. (9) 101 (2014), no. 3, 275–302. 10.1016/j.matpur.2013.06.003Search in Google Scholar

[36] L. Silvestre, Regularity of the obstacle problem for a fractional power of the Laplace operator, Comm. Pure Appl. Math. 60 (2007), no. 1, 67–112. 10.1002/cpa.20153Search in Google Scholar

Received: 2018-01-19
Accepted: 2018-03-13
Published Online: 2018-03-21
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 23.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2018-0011/html
Scroll to top button