Positive radial solutions for Dirichlet problems in the ball

Abstract


We are concerned with existence, localization and multiplicity of positive radial solutions to Dirichlet problems with φ-Laplacians in a ball, in both scalar and system cases. Our approach essentially relies on fixed point index computations and a main feature is that it avoids any Harnack type inequality. Applications to some problems involving operators with Uhlenbeck structure are discussed.

Authors

Petru Jebelean
Institute for Advanced Environmental Research, West University of Timişoara, Timişoara, Romania

Department of Mathematics Babes-Bolyai University, Cluj-Napoca, Romania
Tiberiu Popoviciu Institute of Numerical Analysis, Romanian Academy

Jorge Rodríguez-López
CITMAga & Departamento de Estatística, Análise Matemática e Optimización, Universidade de Santiago de Compostela, 15782, Facultade de Matemáticas, Campus Vida, Santiago, Spain

Keywords

Dirichlet problem; Operator with Uhlenbeck structure; Positive radial solution; Fixed point index; Mean curvature operator; p-Laplacian

Paper coordinates

P. Jebelean, R. Precup, J. Rodríguez-López, Positive radial solutions for Dirichlet problems in the ball, Nonlinear Analysis, 240 (2024), art. id. 113470, https://doi.org/10.1016/j.na.2023.113470

PDF

??

About this paper

Journal

Nonlinear Analysis

Publisher Name

Elsevier

Print ISSN
Online ISSN

google scholar link

[1] Amann H., Fixed point equations and nonlinear eigenvalue problems in ordered Banach spaces, SIAM Rev., 18 (1976), pp. 620-709,  
[3] Candito P., Guarnotta U., Livrea R., Existence of two solutions for singular Φ-Laplacian problems, Adv. Nonlinear Stud., 22 (2022), pp. 659-683,   CrossRefView in ScopusGoogle Scholar
[4] Cheng X., Lü H., Multiplicity of positive solutions for a (p1,p2)– Laplacian system and its applications, Nonlinear Anal. RWA, 13 (2012), pp. 2375-2390
[5] Corrêa F.J.S.A., Carvalho M.L., Gonçalves J.V.A., Silva K.O., Positive solutions of strongly nonlinear elliptic problems, Asymptot. Anal., 93 (2015), pp. 1-20
[7] Drábek P., García-Huidobro M., Manásevich R., Positive solutions for a class of equations with a p-Laplace like operator and weights Nonlinear Anal., 71 (2009), pp. 1281-1300View PDFView articleView in ScopusGoogle Scholar
[8] García-Huidobro M., Manásevich R., Schmitt K., Positive radial solutions of quasilinear elliptic partial differential equations on a ball Nonlinear Anal., 35 (1999), pp. 175-190, View PDFView articleView in ScopusGoogle Scholar
[9] García-Huidobro M., Manásevich R., Ubilla P., Existence of positive solutions for some Dirichlet problems with an asymptotically homogeneous operator, Electron. J. Differential Equations, 1995 (1995), pp. 1-22Google Scholar
[10] García-Huidobro M., Manásevich R., Zanolin F., Infinitely many solutions for a Dirichlet problem with a nonhomogeneous p-Laplacian-like operator in a ball, Adv. Differential Equations, 2 (1997), pp. 203-230View in ScopusGoogle Scholar

[11] Granas A., Dugundji J., Fixed Point Theory, Springer-Verlag, New York (2003), Google Scholar
[12] Guo D., Lakshmikantham V., Nonlinear Problems in Abstract Cones, Academic Press, San Diego (1988), Google Scholar
[13] Gurban D., Jebelean P., Positive radial solutions for systems with mean curvature operator in Minkowski space, Rend. Istit. Mat. Univ. Trieste, 49 (2017), pp. 245-264, View in ScopusGoogle Scholar
[14] Gurban D., Jebelean P., Positive radial solutions for multiparameter Dirichlet systems with mean curvature operator in Minkowski space and Lane–Emden type nonlinearities, J. Differential Equations, 266 (2019), pp. 5377-5396, View PDFView articleView in ScopusGoogle Scholar
[15] Lee Y.-H., Existence of multiple positive radial solutions for a semilinear elliptic system on an unbounded domain, Nonlinear Anal., 47 (2001), pp. 3649-3660, View PDFView articleView in ScopusGoogle Scholar
[16] Precup R., A vector version of Krasnosel’skiĭ’s fixed point theorem in cones and positive periodic solutions of nonlinear systems, J. Fixed Point Theory Appl., 2 (2007), pp. 141-151, View article, CrossRefView in ScopusGoogle Scholar
[17] Precup R., Rodríguez-López J., Multiplicity results for operator systems via fixed point index, Results Math., 74 (2019), pp. 1-14, Google Scholar

[18] Precup R., Rodríguez-López J., Positive radial solutions for Dirichlet problems via a Harnack-type inequality, Math. Methods Appl. Sci. (2022), pp. 1-14, 10.1002/mma.8682, View PDF , This article is free to access, Google Scholar
[19] Rodríguez-López J., A fixed point index approach to Krasnosel’skiĭ–Precup fixed point theorem in cones and applications, Nonlinear Anal., 226 (113138) (2023), p. 19, Google Scholar
[20] Uhlenbeck K, Regularity for a class of non-linear elliptic systems, Acta Math., 138 (1977), pp. 219-240, View in ScopusGoogle Scholar
[21] Walter W., Differential and Integral Inequalities, Springer-Verlag, New York (1970), Google Scholar
[22] Zou H., A priori estimates for a semilinear elliptic system without variational structure and their applications, Math. Ann., 323 (2002), pp. 713-735, View in ScopusGoogle Scholar

Paper (preprint) in HTML form

Positive radial solutions for Dirichlet problems in the ball

Positive radial solutions for Dirichlet problems in the ball

Petru Jebelean Institute for Advanced Environmental Research, West University of Timişoara, Blvd. V. Pârvan, no. 4, 300223 Timişoara, Romania, E-mail: petru.jebelean@e-uvt.ro Radu Precup Faculty of Mathematics and Computer Science and Institute of Advanced Studies in Science and Technology, Babeş-Bolyai University, 400084 Cluj-Napoca, Romania & Tiberiu Popoviciu Institute of Numerical Analysis, Romanian Academy, P.O. Box 68-1, 400110 Cluj-Napoca, Romania, E-mail: r.precup@math.ubbcluj.ro  and  Jorge Rodríguez-López CITMAga & Departamento de Estatística, Análise Matemática e Optimización, Universidade de Santiago de Compostela, 15782, Facultade de Matemáticas, Campus Vida, Santiago, Spain, E-mail: jorgerodriguez.lopez@usc.es – corresponding author
Abstract.

We are concerned with existence, localization and multiplicity of positive radial solutions to Dirichlet problems with ϕ-Laplacians in a ball, in both scalar and system cases. Our approach essentially relies on fixed point index computations and a main feature is that it avoids any Harnack type inequality. Applications to some problems involving operators with Uhlenbeck structure are discussed.

Mathematics Subject Classification: 35J25, 35J60, 34B18, 35J92, 35J93.

Keywords and phrases: Dirichlet problem, operator with Uhlenbeck structure, positive radial solution, fixed point index, mean curvature operator, p-Laplacian

1. Introduction

In this paper, we first deal with existence and localization of positive solutions to the mixed boundary value problem

(1.1) (rn1ϕ(u))=rn1f(r,u),u(0)=u(1)=0,

where ϕ:(a,a)(b,b) (0<a,b+) is an increasing odd homeomorphism and f:[0,1]×+[0,b) is continuous. According to an already usual terminology, ϕ is called classical, singular or bounded if a=b=+, a<+ and b=+ or a=+ and b<+. To cover all possible cases, in the remaining situation a<+ and b<+, we will say that ϕ is singular-bounded. By a solution of (1.1) we mean a nonnegative function uC1[0,1] with u(0)=u(1)=0, |u(r)|<a for all r[0,1], such that rn1ϕ(u)C1[0,1] and (1.1) is satisfied. A solution is said to be positive if it is distinct from the identically zero function.

Our results here generalize those obtained in [18], where some monotonicity assumptions are required on f and also some additional restrictions are imposed on ϕ. Our approach is based on the fixed point index computation and it is worth to notice that no Harnack type inequality is needed.

A main motivation of this study concerns the class of Dirichlet problems in the unit open ball in n, involving operators with Uhlenbeck structure [20]. Such an operator has the form

𝒰θv:=div(θ(|v|)v),

where θ:(0,a)(0,+) is a C1 function. When v(z)=u(|z|), setting r=|z|, it is straightforward to check that

𝒰θv(z)=1rn1(rn1θ(|v(r)|)v(r)).

Assuming that

(1.2) limx0+xθ(x)=0

together with the ellipticity condition

(1.3) θ(x)+xθ(x)>0(x(0,a))

and putting

(1.4) ϕθ(x):=xθ(|x|)(x(a,a))

we have that ϕθ:(a,a)(ϕθ(a),ϕθ(a)) is an increasing odd homeomorphism. Thus, under assumptions (1.2) and (1.3), finding radially symmetric solutions (i.e., solutions of the form v(x)=u(r) with r=|x|) to the Dirichlet problem

𝒰θv=f(|z|,v) in ,v=0 on 

reduces to solving the mixed boundary value problem of type (1.1):

(1.5) (rn1ϕθ(u))=rn1f(r,u),u(0)=u(1)=0,

with ϕθ in (1.4). Note that allowing a to be finite, unlike [20] and most of the subsequent works (see e.g. [3, 5, 7, 10, 9, 8]), also includes singular and singular-bounded ϕθ homeomorphisms.

The following three particular cases are standard models in this context:

θ:(0,+)(0,+), θ(x)=xp2, where p>1, when 𝒰θ becomes the p-Laplacian operator

Δpv=div(|v|p2v)

and the corresponding ϕϕθ:, ϕ(x)=|x|p2x is classical;

θ:(0,1)(0,+), θ(x)=(1x2)12, when 𝒰θ will be the mean extrinsic curvature operator in Minkowski space

v=div(v1|v|2)

and the corresponding ϕϕθ:(1,1), ϕ(x)=x(1x2)12 is singular;

θ:(0,+)(0,+), θ(x)=(1+x2)12, when 𝒰θ engenders the mean curvature operator in Euclidean space

v=div(v1+|v|2)

and the corresponding ϕϕθ:(1,1), ϕ(x)=x(1+x2)12 is bounded.

To complete the table with the remaining case, let θ:(0,1)(0,+) be given by θ(x)=arcsinx. Then the corresponding 𝒰θ becomes

𝒜v=div(varcsin|v|)

with ϕϕθ:(1,1)(1,1), ϕ(x)=xarcsin|x| which is singular-bounded.

Furthermore, we deal with existence and localization of positive solutions for the system

(1.6) (rn1ϕ1(u1))=rn1f1(r,u1,u2),(rn1ϕ2(u2))=rn1f2(r,u1,u2),u1(0)=u1(1)=0=u2(0)=u2(1),

where for each j=1,2, ϕj:(aj,aj)(bj,bj) (0<aj,bj+) is an increasing odd homeomorphism and fj:[0,1]×+×+[0,bj) is continuous. Here, a notable feature is the fact that in the system the homeomorphisms ϕ1 and ϕ2 engendering the differential operators can be different. Among others, this allows the study of radial solutions in for systems invoking, for example, any two of the operators Δp, , or 𝒜 or some combinations of them (see Example 3.5). We extend here existence and multiplicity results obtained in [17], weakening the global monotonicity conditions on the nonlinearities f1 and f2 therein, and making use - as in the scalar case, of an approach that avoids any Harnack type inequality and relying exclusively on the fixed point index estimation.

The rest of the paper is organized as follows. In Section 2 we reformulate (1.1) as a fixed point problem and we obtain localization of the positive solutions under a homotopic boundary condition, as well as under conditions concerning the behavior of the nonlinearity f on some appropriate subintervals in connection with the range of ϕ1. Applications in terms of asymptotic conditions when the nonlinearity is with separated variables are provided. Section 3 is devoted to extensions of the results in Section 2 to system (1.6). Multiplicity of positive solutions of (1.6) is obtained in Section 4 by refining the estimations of the fixed point index under some additional suitable conditions on the nonlinearities f1 and f2.

2. Scalar case

The space X:=C[0,1] will be endowed with the usual sup-norm ||. We denote B(d):={uX:|u|<d} and let the cone P:={uX:u0 on [0,1]}. It is not difficult to see that a nonnegative function u is a solution of problem (1.1) if and only if u is a fixed point of the operator T:PP given by

(2.1) T(u)(r)=r1ϕ1(τ1n0τsn1f(s,u(s))𝑑s)𝑑τ(r[0,1]).

It is a standard matter that the operator T is completely continuous.

Let us now consider the closed, convex cone

(2.2) K={uP:u is nonincreasing on [0,1]}.
Proposition 2.1.

The operator T maps K into itself.

Proof.

Indeed, take uK and let us show that v:=Tu belongs to K. Since f is nonnegative, v0 and, moreover, from the sign of the homeomorphism ϕ, one deduces that v is nonincreasing on [0,1]. Therefore, vK, as claimed. ∎

For any real numbers α>β>0 and η(0,1), consider the bounded sets

Uα:=KB(α) and Vβ:={uU¯α:u(η)<β}.

Observe that if uK then u(η)=minr[0,η]u and u(0)=|u|. Since K is a closed set, the closure of Uα in both X and K is the same and it is easily seen that the sets Uα and Vβ are open in U¯α. Also, as U¯α is closed, the closure of any of its subsets is the same in both X and U¯α. In the sequel, the boundary of the subsets of U¯α will be understood with respect to the topology induced on U¯α, unless otherwise specified. It is a standard matter to check that

U¯α={uK:|u|α},V¯β={uK:|u|α,u(η)β},
U¯αV¯β¯={uK:|u|α,u(η)β}=U¯αVβ,
Vβ={uK:|u|α,u(η)=β}=(U¯αV¯β).

Next, it is clear that U¯α being closed and convex, it is a retract. Thus, if 𝒪 is an open subset in U¯α, T(𝒪¯)U¯α and T is fixed points free on the boundary of 𝒪, then i(T,𝒪)i(T,𝒪,U¯α) – the fixed point index of the completely continuous operator T:𝒪¯U¯α on 𝒪 with respect to U¯α is well defined [1, 12, 11].

The following result guarantees the existence of a non-trivial fixed point in K of the operator T.

Theorem 2.2.

Assume that there exist numbers α>β>0, η(0,1) such that T(U¯α)U¯α and there exists a function hK such that |h|α, h(η)>β and

(2.3) (1λ)T(u)+λhufor uK with |u|α,u(η)=β and λ[0,1].

Then i(T,U¯αV¯β)=1 and thus problem (1.1) has a solution uK, satisfying

β<u(η) and |u|α.
Proof.

This is a modified computation of the fixed point index in [18]. Since T(U¯α)U¯α and U¯α=, we have that the homotopy H:[0,1]×U¯αU¯α, given by

H(λ,u)=λT(u)(λ[0,1],uU¯α)

is admissible (i.e., H(λ,) is fixed points free on U¯α for all λ[0,1]). Then, by the invariance under homotopy of the fixed point index, we get

i(T,U¯α)=i(H(1,),U¯α)=i(H(0,),U¯α)=i(0,U¯α)=1,

where the last equality is due to the fact that 0U¯α.

Next, as h lies in the convex set U¯α, we can define the homotopy H~:[0,1]×U¯αU¯α by

H~(λ,u):=(1λ)T(u)+λh(λ[0,1],uU¯α).

Condition (2.3) means

H~(λ,u)ufor all uVβ and λ[0,1],

showing that H~:[0,1]×V¯βU¯α is admissible. Thus, we have

i(T,Vβ)=i(H~(0,),Vβ)=i(H~(1,),Vβ)=i(h,Vβ)=0,

since hU¯αV¯β.

Then, as {U¯αV¯β,Vβ,Vβ} is a partition of U¯α, using the excision-additivity property of the fixed point index, we infer

i(T,U¯αV¯β)=i(T,U¯α)i(T,Vβ)=1,

which ensures the existence of a fixed point of T in the set U¯αV¯β, that is, a solution of (1.1) with β<u(η) and |u|α. ∎

Remark 2.1.

From the first part of the above proof, we observe that the simple hypothesis that there exists a number α>0 such that T(U¯α)U¯α yields i(T,U¯α)=1. This ensures that T has a fixed point in U¯α, hence (1.1) is solvable; notice, the same conclusion also follows from Schauder’s fixed point theorem.

Now we give sufficient conditions in order to ensure that the assumptions of the previous result hold.

For any α>β>0 and η(0,1) denote

mα,β:=min{f(r,x):r[0,η],x[β,α]},
Mα:=max{f(r,x):r[0,1],x[0,α]}

and note that 0mα,βMα<b.

Theorem 2.3.

If there exist α>β>0 and η(0,1) such that

(2.4) ϕ1(Mα) α,
(2.5) (1η)ϕ1(ηnmα,β/n) >β,

then problem (1.1) has at least one solution uK with β<u(η) and |u|α.

Proof.

We shall apply Theorem 2.2. First, we show that T maps U¯α into itself. Indeed, for uK with |u|α, we have that

0f(s,u(s))Mα(s[0,1])

and thus from (2.4),

|T(u)(r)| 01ϕ1(τ1n0τsn1f(s,u(s))𝑑s)𝑑τ
01ϕ1(0τf(s,u(s))𝑑s)𝑑τϕ1(Mα)α(r[0,1]).

So, |T(u)|α for all uK with |u|α, hence T maps U¯α into itself.

Next, we prove that (2.3) is fulfilled with hα, which clearly satisfies |h|=α and h(η)>β. Assume that (2.3) does not hold. Then there exist uK with |u|α, u(η)=β and λ[0,1] such that

(1λ)T(u)+λh=u.

In particular, one has

β=u(η) =(1λ)T(u)(η)+λα
=(1λ)η1ϕ1(τ1n0τsn1f(s,u(s))𝑑s)𝑑τ+λα
(1λ)η1ϕ1(τ1n0ηsn1f(s,u(s))𝑑s)𝑑τ+λβ.

Since βu(s)α for all s[0,η], it follows

β (1λ)η1ϕ1(τ1n0ηsn1mα,β𝑑s)𝑑τ+λβ
(1λ)(1η)ϕ1(ηnnmα,β)+λβ,

that is,

(1λ)β(1λ)(1η)ϕ1(ηnmα,β/n),

which contradicts (2.5) for any λ[0,1). In the case λ=1 one has the contradiction

β=u(η)=α>β.

Finally, the conclusion follows from Theorem 2.2. ∎

Remark 2.2.

Conditions (2.4)-(2.5) in Theorem 2.3 are of compression type. If a Harnack inequality holds for all nonnegative supersolutions u of (1.1), that is

minr[0,η]uc|u|,

for some c(0,1), then one may consider the cone

K^={uP:minr[0,η]uc|u|}.

In that case, the set V^β={uK^:minr[0,η]u<β} is bounded (indeed, V^βU^β/c={uK^:|u|<β/c}) and so expansion conditions are possible by means of standard fixed point index arguments. In addition, in that case, condition (2.5) can be weakened by replacing mα,β with

m~β:=min{f(r,s):r[0,η],s[β,β/c]},

and the proof follows in a similar way.

Note that if 0<a<+, then condition (2.4) in Theorem 2.3 is trivially satisfied for α large enough (for instance, α=a). Thus, in this case, it suffices to ensure the existence of a positive number β satisfying condition (2.5). Observe that, for β small enough, (2.5) can be rewritten as

(2.6) mα,βϕ(β/(1η))>nηn.

Hence, it is natural to look for asymptotic conditions on the quotient f/ϕ at 0 which guarantee that the above inequality holds. Then a difficulty arises: the number mα,β depends not only of the behavior of f(r,) at β, but on the whole interval [β,α]. That is why it is convenient to assume the following monotonicity assumption on f:

(Hf) the function f(r,) is nondecreasing in (0,a) for every r[0,1].

Under assumption (Hf), we have that

mα,β=min{f(r,β):r[0,η]}:=mβ,

for any α>β, so (2.6) reads as

(2.7) mβϕ(β/(1η))>nηn.

Below, we need to invoke the condition

(2.8) lim supx0+ϕ(τx)ϕ(x)<+for all τ>1,

that is employed in [7] in connection with a classical homeomorphism ϕ and in [2] relative to a singular ϕ.

Theorem 2.4.

Assume that 0<a<+ and (Hf), (2.8) are fulfilled. If there is some η(0,1) such that

(2.9) limx0+f(r,x)ϕ(x)=+ uniformly with r[0,η],

then problem (1.1) has at least one positive nonincreasing solution u with u(η)>0.

Proof.

This resembles the proof of [18, Theorem 3.6]. From (2.8) with τ=1/(1η), there exist numbers L>0 and ρ1(0,a(1η)) such that

(2.10) ϕ(x/(1η))Lϕ(x)for all x(0,ρ1).

Then, by (2.9), we can find ρ2(0,ρ1) so that

f(r,x)ϕ(x)>Lnηnfor all (r,x)[0,η]×(0,ρ2).

Choosing β(0,ρ2), we deduce

mβϕ(β)>Lnηn,

which, together with (2.10), gives (2.7) and the conclusion follows by Theorem 2.3. ∎

In the case a=+, the existence of a positive number α satisfying (2.4) is not trivial and it can be derived from a suitable asymptotic behavior of the ratio f/ϕ at infinity.

Theorem 2.5.

Assume that a=+ and (Hf), (2.8) are fulfilled. If, in addition,

(2.11) limx+f(r,x)ϕ(x)<1 uniformly with r[0,1]

and there exists η(0,1) such that (2.9) is satisfied, then problem (1.1) has at least one positive nonincreasing solution u with u(η)>0.

Proof.

Using (Hf), from (2.11) we easily get that there exists α>0 such that (2.4) holds true. Then, the arguments in the proof of Theorem 2.4 yield the conclusion. ∎

Next, we consider the case in which the function f is with separated variables i.e., f(r,x)=μ(r)g(x), and where μ and g satisfy the following hypotheses:

(Hμ) μ:[0,1](0,+) is continuous;

(Hg) g:+[0,b/μ¯) is a continuous function such that g(0)=0 and g(x)>0 for all x>0, where μ¯:=max{μ(r):r[0,1]}.

Notice that no monotonicity assumptions on f are required.

Theorem 2.6.

Assume that (2.8), (Hμ) and (Hg) are fulfilled. If, in addition, one of the following conditions holds

(i) a<+ and

(2.12) limx0+g(x)ϕ(x)=+;

(ii) a=+ and

(2.13) limx0+g(x)ϕ(x)=+,limx+g(x)ϕ(x)<1μ¯;

then problem

(2.14) (rn1ϕ(u))=rn1μ(r)g(u),u(0)=u(1)=0,

has at least one strictly decreasing solution u>0 on [0,1).

Proof.

Let us show that hypotheses in Theorem 2.3 hold, that is, there exist α>β>0 such that inequalities (2.4) and (2.5) are satisfied.

First, suppose that condition (i) holds. Since a<+, condition (2.4) is satisfied for any αa. Fix such a positive α and let us prove the existence of a number β(0,α) as in (2.5). Taking any η(0,1), by (2.8) with τ=1/(1η), there exist L>0 and ρ1(0,a(1η)) such that (2.10) holds true. The asymptotic behavior of the quotient g/ϕ at zero in (2.12) implies that there exists ρ2>0 (we may assume ρ2<ρ1<α) such that

g(x)>Lμ¯nηnϕ(x)for all x(0,ρ2),

where μ¯ stands for min{μ(r):r[0,η]}>0. This and (2.10) give

(2.15) μ¯g(x)>nηnϕ(x1η)for all x(0,ρ2).

Since g is continuous, vanishes at zero and it is positive on (0,+), there exists x¯(0,ρ2) such that

g(x¯)<min{g(x):x[ρ2,α]}.

We choose a number β[x¯,α] such that

g(β)=min{g(x):x[x¯,α]}.

Clearly, g(β)g(x¯)<min{g(x):x[ρ2,α]} implies that β<ρ2. Moreover, it is obvious that

g(β)=min{g(x):x[β,α]}.

Therefore, we obtain from (2.15) that

mα,β=μ¯g(β)>nηnϕ(β1η),

as wished.

Next, let us suppose that condition (ii) holds. If b=+, the asymptotic behavior of g/ϕ at infinity in (2.13) guarantees that there exists k>0 and l(0,1) such that

μ¯g(x)k+lϕ(x)for all x0.

Since ϕ is unbounded, there exists α>0 large enough, such that k(1l)ϕ(α). Hence, being ϕ an increasing function, we have

μ¯g(x)ϕ(α)for all x[0,α],

which implies that Mαϕ(α). On the other hand, if b<+, then condition (ii) implies that limx+g(x)<b/μ¯ and so there exists l>0 and x0>0 such that μ¯g(x)l<b for all xx0. It suffices to choose αx0 such that ϕ(α)max{l,μ¯maxx[0,x0]g(x)} in order to ensure that Mαϕ(α). Finally, the existence of a positive number β<α satisfying inequality (2.5) can be deduced exactly as in the previous case.

Now, Theorem 2.3 ensures that problem (2.14) has a positive solution u which, by virtue of [2, Lemma 1] is strictly decreasing. ∎

Let us emphasize the previous result in the particular cases of radial solutions to the Dirichlet problem in the unit ball of n involving the operators Δp, , and 𝒜 defined in Section 1.

Corollary 2.7.

Assume that conditions (Hμ) and (Hg) with b=+ hold. If

limx0+g(x)xp1=+andlimx+g(x)xp1<1μ¯,

then problem

Δpv=μ(|z|)g(v) in ,v=0 on 

has at least one radial solution v(z)=u(|z|), with u>0 on [0,1) and strictly decreasing.

Corollary 2.8.

Assume that conditions (Hμ) and (Hg) with b=+ hold. If

limx0+g(x)x=+,

then problem

v=μ(|z|)g(v) in ,v=0 on 

has at least one radial solution v(z)=u(|z|), with u>0 on [0,1) and strictly decreasing.

Corollary 2.9.

Assume that conditions (Hμ) and (Hg) with b=1 hold. If

limx0+g(x)x=+andlimx+g(x)<1μ¯,

then problem

v=μ(|z|)g(v) in ,v=0 on 

has at least one radial solution v(z)=u(|z|), with u>0 on [0,1) and strictly decreasing.

Corollary 2.10.

Assume that conditions (Hμ) and (Hg) with b=1 hold. If

limx0+g(x)xarcsinx=+,

then problem

𝒜v=μ(|z|)g(v) in ,v=0 on 

has at least one radial solution v(z)=u(|z|), with u>0 on [0,1) and strictly decreasing.

3. Systems – existence of positive solutions

Let ϕj and fj (j=1,2) be as in Section 1. We consider the cone P×P in the product Banach space X×X and the completely continuous operator T=(T1,T2):P×PP×P, where, for (u1,u2)P×P,

Tj(u1,u2)(r)=r1ϕj1(1τn10τsn1fj(s,u1(s),u2(s))𝑑s)𝑑τ(r[0,1],j=1,2).

Clearly, u=(u1,u2) is a solution of (1.6) (that is, u solves (1.6) and both u1 and u2 are nonnegative functions on [0,1]) iff u is a fixed point of T. In the case of (1.6), by a positive solution, we mean a solution u=(u1,u2) such that both u1 and u2 are distinct from the identically zero function. Also, arguing as in the proof of Proposition 2.1 we have that T maps the cone K×K into itself.

In accordance with the previous section, for any real numbers η(0,1), αj>βj>0, j=1,2, we define the sets

Uαj:=KB(αj) and Vβj:={uU¯αj:u(η)<βj}.

This time, the retract will be the bounded closed convex set U¯α1×U¯α2. If 𝒪 is an open subset of U¯α1×U¯α2 and A:𝒪¯U¯α1×U¯α2 is a completely continuous operator which is fixed points free on the boundary of 𝒪, we use the notation i(A,𝒪) for the fixed point index of A on 𝒪 with respect to U¯α1×U¯α2.

We can state the following vector version of Theorem 2.2, cf. [16, 17, 19].

Theorem 3.1.

Assume that there are numbers αj>0, j=1,2, such that T(U¯α1×U¯α2)U¯α1×U¯α2. In addition, suppose that there exist η(0,1), βj(0,αj) and functions hjU¯αjV¯βj, j=1,2, such that

(3.1) (1λ)Tj(u)+λhjuj

for all u=(u1,u2)(U¯α1Vβ1)×(U¯α2Vβ2) with uj(η)=βj and λ[0,1] (j=1,2).

Then i(T,(U¯α1V¯β1)×(U¯α2V¯β2))=1 and thus problem (1.6) has a solution u=(u1,u2)K×K, with

βj<uj(η) and |uj|αj(j=1,2).
Proof.

For the sake of simplicity of the writing, set Q:=U¯α1×U¯α2. Since U¯αiVβi are closed and convex, there exist retractions ρi:XU¯αiVβi (i=1,2). Then, the map ρ:Q(U¯α1Vβ1)×(U¯α2Vβ2), defined by

ρ(u)=(ρ1(u1),ρ2(u2))(u=(u1,u2)Q)

is a retraction, too.

We introduce the operator N=(N1,N2):QQ defined by N:=Tρ. Explicitly, this means

Nj(u)=Tj(ρ1(u1),ρ2(u2))(u=(u1,u2)Q,j=1,2)

and it is easily seen that N is completely continuous. From (3.1) we deduce that, for each j{1,2}, it holds

(3.2) (1λ)Nj(u)+λhjuj

for all u=(u1,u2)Q with uj(η)=βj and λ[0,1].

We first consider the homotopy H:[0,1]×QQ, given by

H(λ,u)=λN(u)(λ[0,1],uQ),

which is admissible and so, we infer

(3.3) i(N,Q)=i(H(1,),Q)=i(H(0,),Q)=i(0,Q)=1.

Next, denote

Ω11 :=Vβ1×Vβ2={(u1,u2)Q:u1(η)<β1,u2(η)<β2},
Ω12 :=Vβ1×(U¯α2V¯β2)={(u1,u2)Q:u1(η)<β1,u2(η)>β2},
Ω21 :=(U¯α1V¯β1)×Vβ2={(u1,u2)Q:u1(η)>β1,u2(η)<β2},
Ω22 :=(U¯α1V¯β1)×(U¯α2V¯β2)={(u1,u2)Q:u1(η)>β1,u2(η)>β2},
Γ :={(u1,u2)Q:u1(η)=β1 or u2(η)=β2}

and notice that {Ω11,Ω12,Ω21,Ω22,Γ} is a partition of Q. Clearly, Γ is closed and Ωjk are open (j,k=1,2). Also, since, on account of (3.2) with λ=0, we have that N(u)u for uΓ, the excision-additivity property of the fixed point index and (3.3) yields

(3.4) i(N,Ω22)=1i(N,Ω11)i(N,Ω12)i(N,Ω21).

To estimate i(N,Ω) with Ω{Ω11,Ω12,Ω21}, first observe that hQΩ¯. Then, let H~:[0,1]×Ω¯Q be the homotopy

H~(λ,u)=(1λ)N(u)+λh(u=(u1,u2)Ω¯,λ[0,1]).

From (3.2) we have that H~(λ,u)u for uΩ(Γ) and λ[0,1], meaning that H~ is admissible. Thus, we get

i(N,Ω)=i(H~(0,))=i(H~(1,))=i(h,Ω)=0.

Therefore, as N=T on Ω¯22=(U¯α1Vβ1)×(U¯α2Vβ2), the conclusion follows from (3.4). ∎

Remark 3.1.

Similarly to Remark 2.1, solely assuming that there exist numbers αj>0, j=1,2, such that T(U¯α1×U¯α2)U¯α1×U¯α2, we have that i(T,U¯α1×U¯α2)=1. This immediately follows by the invariance under homotopy of the fixed point index, as in the beginning of the proof of Theorem 2.2

For any numbers η(0,1), 0<βj<αj, j=1,2, we use the following notations

mα,βj:=min{fj(r,x,y):r[0,η],x[β1,α1],y[β2,α2]},
Mα1,α2j:=max{fj(r,x,y):r[0,1],x[0,α1],y[0,α2]}.

Note that one has 0mα,βjMα1,α2j<bj (j=1,2). Also, in the writing of mα,βj, actually, we mean α=(α1,α2) and β=(β1,β2).

We have the following existence result for the system (1.6).

Theorem 3.2.

If there exist η(0,1), αj>βj>0, such that

ϕj1(Mα1,α2j)αj and (1η)ϕj1(ηnmα,βj/n)>βj(j=1,2),

then problem (1.6) has at least one solution u=(u1,u2)K×K such that

βj<uj(η)and|uj|αj,j=1,2.
Proof.

This can be easily adapted from that of Theorem 2.3 using Theorem 3.1 instead of Theorem 2.2. ∎

Now we consider the following system

(3.5) (rn1ϕ1(u1))=rn1μ1(r)g1(u1,u2),(rn1ϕ2(u2))=rn1μ2(r)g2(u1,u2),u1(0)=u1(1)=0=u2(0)=u2(1),

under the hypotheses:

(Hsμ) μ1,μ2:[0,1](0,+) are continuous;

(Hsg) gj:+×+[0,bj/μ¯j) is a continuous function, where μ¯j:=max{μj(r):r[0,1]} (j=1,2). Also, g1(0,0)=0=g2(0,0) and g1(ξ,0)>0<g2(0,ξ) for all ξ>0.

We say that a function g=g(x,y) is nondecreasing with respect to y (resp. x) if for fixed x (resp. y) one has

g(x,y1)g(x,y2) as y1y2(resp. g(x1,y)g(x2,y) as x1x2).

This property is also known as quasi-monotonicity, see for instance [4, 15, 13, 14, 21].

Proposition 3.3.

Assume (Hsμ) and (Hsg) and that g1(x1,x2), g2(x1,x2) are nondecreasing with respect to x2 and x1, respectively. If u=(u1,u2) is a positive solution of problem (3.5) then both of the components of u are >0 on [0,1) and strictly decreasing.

Proof.

From

(3.6) rn1ϕ1(u1)=0rtn1μ1(t)g1(u1,u2)𝑑t0rtn1μ1(t)g1(u1,0)𝑑t

we have u10, that is u1 is decreasing. Since u1 is positive (not identically zero), it follows that u1(0)>0. Then, using (Hsg) and (3.6), we get u1<0, meaning that u1 is strictly decreasing and >0 on [0,1). Similar reasoning for u2.∎

Now, as a consequence of Theorem 3.2, we prove the existence of positive solutions for problem (3.5) under a sublinear growth condition on g1 and g2. In the particular case of systems involving the mean curvature operator in Minkowski space, we obtain the following result in the line of [13, Theorem 3.1]. To state it, for δ>0, we introduce the notation

Πj(δ):={[0,δ]×+ if j=1,+×[0,δ] if j=2.
Theorem 3.4.

Let conditions (Hsμ) and (Hsg) hold and let condition (2.8) be satisfied with ϕ replaced by ϕj (j=1,2). Assume that g1(x1,x2), g2(x1,x2) are nondecreasing with respect to x2 and x1, respectively, and

(3.7) limx10+g1(x1,0)ϕ1(x1)=+,
(3.8) limx20+g2(0,x2)ϕ2(x2)=+.

In case that for some j{1,2} one has aj=+, assume in addition that one of the following conditions is satisfied:

(A):

(i) if bj=+, then for any δ>0, gj is bounded on Πj(δ) and

(3.9) limxj+gj(x1,x2)ϕj(xj)<1μ¯j, uniformly with xk+(kj),

(ii) if bj<+ then

(3.10) supx1,x2+gj(x1,x2)<bjμ¯j;
(B):

gj(x1,x2) is also nondecreasing in xj and

(3.11) limτ+gj(τ,τ)ϕj(τ)<1μ¯j.

Then problem (3.5) has at least one solution u=(u1,u2) with both of the components >0 on [0,1) and strictly decreasing.

Proof.

We first show that Theorem 3.2 applies.

Step 1: Finding α1,α2. First, we claim that there exists τ1>0 large enough, such that

(3.12) ϕ11(Mτ,τ1)<τfor all ττ1.

Indeed, if a1<+, we pick τ1a1 and obviously (3.12) holds true. Otherwise, meaning a1=+, the existence of such a number τ1 can be obtained under each of the conditions (A) and (B) (with j=1), as follows.

Case A: Assume that condition (A) holds. So, if b1=+, (3.9) implies the existence of c10 and l1(0,1) satisfying

μ¯1g1(x1,x2)c1+l1ϕ1(x1)for all x1,x20.

Since ϕ1 is unbounded, we can find τ1>0 such that c1(1l1)ϕ1(τ1). For ττ1 and x1,x2[0,τ], we can estimate

μ¯1g1(x1,x2)(1l1)ϕ1(τ)+l1ϕ1(x1)ϕ1(τ)

which gives (3.12). In case b1<+, (3.10) implies that there is some l1>0 such that

μ¯1g1(x1,x2)l1<b1for all x1,x20.

We choose τ1>0 with ϕ1(τ1)>l1 (such an τ1 exists by virtue of limx+ϕ1(x)=b1). Then for ττ1, it follows

ϕ1(τ)ϕ1(τ1)>Mτ,τ1,

which clearly implies (3.12).

Case B: Assume now that condition (B) holds. The asymptotic condition (3.11) guarantees the existence of l1 and τ10 such that

g1(τ,τ)ϕ1(τ)l1<1μ¯1for all ττ1.

From this, using that μ¯1l1<1, one has

μ¯1g1(τ,τ)μ¯1l1ϕ1(τ)<ϕ1(τ)for all ττ1

and, by the monotonicity of g1 in both variables, we get (3.12), as claimed.

As above, also there exists τ2>0 such that

ϕ21(Mτ,τ2)<τfor all ττ2.

Thus, taking α1=α2=α0:=max{τ1,τ2}, it follows

(3.13) ϕj1(Mα1,α2j)<αj(j=1,2).

Step 2: We prove the existence of numbers β1,β2(0,α0) as required in Theorem 3.2.

Fixed η(0,1), by (2.8) and (3.7), there exists ρ1=ρ1(η)(0,α0) such that

(3.14) μ¯1g1(x,0)>nηnϕ1(x1η)for all x(0,ρ1),

where μ¯1:=min{μ1(r):r[0,η]}. Since g1 is continuous and

g1(0,0)=0<min{g1(x,0):x[ρ1,α0]},

there exists x1(0,ρ1) such that

0<g1(x1,0)<min{g1(x,0):x[ρ1,α0]}.

Choose β1[x1,ρ1) such that g1(β1,0)=min{g1(x,0):x[x1,α0]}. Since g1(x,y) is nondecreasing with respect to y, we have

g1(β1,0)=min{g1(x,y):x[β1,α0],y[0,α0]},

and thus (3.14) implies that

μ¯1min{g1(x,y):x[β1,α0],y[0,α0]}>nηnϕ1(β11η).

Clearly, the inequality (1η)ϕ11(ηnmα,β1/n)>β1 holds for any β=(β1,σ) with σ(0,α0). The number β2 can be fixed similarly using condition (3.8) instead of (3.7) and obtaining that (1η)ϕ21(ηnmα,β2/n)>β2 for all β=(σ,β2) with σ(0,α0). We take the pair β=(β1,β2).

Now, Theorem 3.2 ensures that system (3.5) has a positive solution u=(u1,u2) and the conclusion follows by Proposition 3.3.

Example 3.5.

As emphasized in Section 1, the form of the system (3.5) allows to treat Dirichlet problems in the unit ball with any two of the operators Δp, , or 𝒜, or combinations of them. To exemplify, we consider the autonomous system

(3.15) Δpv1=av1α+bv2r,v2σv2=cv1s+dv2q,v1=v2=0on 

where a,d>0, σ,b,c,s0, α[0,p1), q[0,1) and r[0,p1]. Then (3.15) has at least one radial solution v=(v1,v2) with v1,v2>0 in and radially strictly decreasing, provided that either r<p1 or r=p1 and b<1. Indeed, finding radial solutions reduces to a problem of type (3.5) with

ϕ1(x)=|x|p2x,ϕ2(x)=x(1x2)12+σx(1+x2)12,μ1=μ21,g1(x1,x2)=ax1α+bx2r,g2(x1,x2)=cx1s+dx2q

and the conclusion follows from Theorem 3.4 (condition (B)). Here, the nonlinearities g1 and g2 are of the type considered in [6], [22], where they are superlinear and associated with the classical Laplacian operator. It is worth to observe that – as therein, the presence of the p-Laplacian operator Δp in the first equation induces restrictive conditions on both of the exponents α and r, while the singular operator in the second equation allows only the sublinearity restriction on the exponent q – a situation that, having in view [13, Example 3.4 (ii)], is somehow to be expected. A situation when condition (A) (i) in Theorem 3.4 is fulfilled is, for instance, if in system (3.15) the first equation is replaced by

Δpv1=a(1+arctanv2)v1α+|sinv1|

and a,α,σ,c,s,d,q are kept as above.

4. Systems – multiplicity of positive solutions

We emphasize the fact that the results in this section have a simpler counterpart in the scalar case, and this is quite easy to formulate following the systems model. For the sake of brevity of the exposition, we will restrict ourselves only to the case of systems.

Clearly, Theorem 3.2 provides much more information about solutions of (1.6) than Theorem 3.4 since it gives a concrete localization of them. This is the key in order to derive multiplicity results from Theorem 3.2. Obviously, if there exist several pairs of numbers α and β satisfying the conditions of such existence result, we get multiplicity. In addition, the fixed point index computation obtained in Theorem 3.1 can be useful to this aim. In this line, we can prove the following three-solution theorem.

Theorem 4.1.

Under the hypotheses of Theorem 3.1, assume that there exist α~1(0,α1] and α~2(0,α2] such that either α~1<β1 or α~2<β2 and, moreover,

(4.1) T(u)λufor all u(Uα~1×Uα~2) and all λ1.

Then problem (1.6) has at least three solutions: (u1,u2),(v1,v2),(w1,w2)K×K such that

βj<uj(η) and |uj|αj(j=1,2);
|vj|<α~j(j=1,2);
|wj|αj(j=1,2),α~1<|w1| or α~2<|w2|, and w1(η)<β1 or w2(η)<β2.
Proof.

To simplify the writing, as in the proof of Theorem 3.1, denote Ω22:=(U¯α1V¯β1)×(U¯α2V¯β2), Q:=U¯α1×U¯α2 and let Q:=Uα~1×Uα~2. Since the open sets Ω22 and Q are disjoint, we can write

Q =Ω22Q¯[QΩ22Q¯]
=(Ω22Ω22)(QQ)[QΩ22Q¯]
=Ω22Q[QΩ22Q¯](Ω22Q)

and hence, {Ω22Q,QΩ22Q¯,(Ω22Q)} is a partition of Q. By Theorem 3.1 and Remark 3.1, we have that

(4.2) i(T,Ω22)=1=i(T,Q).

From (4.1) we deduce that the homotopy H:[0,1]×Q¯Q defined by

H(λ,u)=λT(u)(u=(u1,u2)Q¯,λ[0,1])

is admissible and, by the invariance under homotopy of the fixed point index, it follows

(4.3) i(T,Q)=i(H(1,),Q)=i(H(0,),Q)=i(0,Q)=1.

Thus, using (4.2) and (4.3) and the excision-additivity property of the fixed point index, we infer

i(T,U¯α1×U¯α2[(U¯α1Vβ1)×(U¯α2Vβ2)U¯α~1×U¯α~2])=i(T,QΩ22Q¯)=1.

Finally, the conclusion follows in a straightforward way from the existence property of the index. ∎

Remark 4.1.

Obviously, under the assumptions of Theorem 4.1, both components of the solution (u1,u2) are positive, (v1,v2) may be the trivial solution while (w1,w2) may be semi-trivial, meaning that either w1 or w2 may be the identically zero function (see Figure 1).

KK|u1|=α1|u2|=α2|u1|=α~1|u2|=α~2(u1,u2)(w1,w2)(v1,v2)min[0,η]u1=β1min[0,η]u2=β2Uα~1×Uα~2(U¯α1V¯β1)×(U¯α2V¯β2)
Figure 1. Localization of the solutions provided by Theorem 4.1.

As a direct consequence of Theorem 4.1, we deduce the following multiplicity result.

Theorem 4.2.

Under the hypotheses of Theorem 3.2, assume that there exist α~1(0,α1] and α~2(0,α2] such that either α~1<β1 or α~2<β2 and, moreover,

(4.4) ϕj1(Mα~1,α~2j)<α~j(j=1,2).

Then problem (1.6) has at least three solutions: (u1,u2),(v1,v2),(w1,w2)K×K such that

βj<uj(η) and |uj|αj(j=1,2);
|vj|<α~j(j=1,2);
|wj|αj(j=1,2),α~1<|w1| or α~2<|w2|, and w1(η)<β1 or w2(η)<β2.
Proof.

Hypotheses of Theorem 3.2 guarantee that conditions of Theorem 3.1 are satisfied. So, to apply Theorem 4.1, we only have to verify (4.1). But, (4.4) ensures that

T(U¯α~1×U¯α~2)Uα~1×Uα~2

(see the proof of Theorem 2.3), which implies (4.1). ∎

In the particular case when α~1=α1, we obtain the following corollary and a similar result also holds true when α~2=α2.

Corollary 4.3.

Assume that there are numbers η(0,1), α1>β1>0, α2>β2>α~2>0, such that

ϕ11(Mα1,α21)<α1,ϕ21(Mα1,α22)α2,ϕ21(Mα1,α~22)<α~2

and

(1η)ϕj1(ηnmα,βj/n)>βj(j=1,2).

Then problem (1.6) has at least three solutions: (u1,u2),(v1,v2),(w1,w2)K×K such that

βj<uj(η) and |uj|αj(j=1,2);
|v1|<α1,|v2|<α~2;
|wj|αj(j=1,2),α~2<|w2|, and w1(η)<β1 or w2(η)<β2.

Existence of at least two positive solutions can be obtained if we strengthen the previous assumptions. See their localization in Figure 2.

Theorem 4.4.

Under the hypotheses of Theorem 3.1, assume that there exist α~1(0,β1) and α~2(0,β2) such that

(4.5) T1(U¯α~1×U¯α2)Uα~1,
(4.6) T2(U¯α1×U¯α~2)Uα~2.

Then problem (1.6) has at least two positive solutions: (u1,u2),(v1,v2)K×K such that

βj<uj(η) and |uj|αj(j=1,2);
α~j<|vj|αj(j=1,2), and v1(η)<β1 or v2(η)<β2.
KK|u1|=α1|u2|=α2|u1|=α~1|u2|=α~2(u1,u2)(v1,v2)min[0,η]u1=β1min[0,η]u2=β2Ω0(U¯α1V¯β1)×(U¯α2V¯β2)
Figure 2. Localization of the solutions provided by Theorem 4.4.
Proof.

Let Q, Q and Ω22 be as in the proof of Theorem 4.1. First, observe that conditions (4.5) and (4.6) imply T(Q¯)Q. Then, using

(Q) =(Uα~1×U¯α~2)(U¯α~1×Uα~2)
={(u1,u2)Q:|u1|=α~1,|u2|α~2 or |u1|α~1,|u2|=α~2},
(Uα~1×U¯α2) =Uα~1×U¯α2
={(u1,u2)Q:|u1|=α~1,|u2|α2},
(U¯α1×Uα~2) =U¯α1×Uα~2
={(u1,u2)Q:|u1|α1,|u2|=α~2}

and the invariance under homotopy of the fixed point index (as in the proof of Theorem 4.1), we deduce that

i(T,Q)=i(T,Uα~1×U¯α2)=i(T,U¯α1×Uα~2)=1.

Let

γ1 :={(u1,u2)Q:|u1|α~1,|u2|=α~2 or |u1|=α~1,|u2|α2},
γ2 :={(u1,u2)Q:|u1|=α~1,|u2|α~2 or |u1|α1,|u2|=α~2}

and notice that γ1, γ2 are closed, γ1Q(Uα~1×U¯α2), γ2Q(U¯α1×Uα~2) and T is fixed points free on both γ1 and γ2. Also, we have that {Q,Uα~1×(U¯α2U¯α~2),γ1} and {Q,(U¯α1U¯α~1)×Uα~2,γ2} are partitions of U¯α~1×U¯α2 and U¯α1×U¯α~2, respectively. Then, excision-additivity property of the index yields

i(T,Uα~1×(U¯α2U¯α~2)) =i(T,Uα~1×U¯α2)i(T,Q)=0,
i(T,(U¯α1U¯α~1)×Uα~2) =i(T,Uα~1×U¯α2)i(T,Q)=0.

Consider the relatively open set

Ω0:=(Uα~1×U¯α2)(U¯α1×Uα~2)

and observe that setting

γ:=γ1γ2={(u1,u2)Q:|u1|=α~1,|u2|α2 or |u1|α1,|u2|=α~2}.

one has that {Q,Uα~1×(U¯α2U¯α~2),(U¯α1U¯α~1)×Uα~2,γ} is a partition of Ω¯0. Thus, we obtain

(4.7) i(T,Ω0)=i(T,Q)+i(T,Uα~1×(U¯α2U¯α~2))+i(T,(U¯α1U¯α~1)×Uα~2)=1.

Finally, consider the relative complement of Ω¯0Ω¯22 with respect to the set Q, that is,

Ω:=U¯α1×U¯α2[Ω¯0(U¯α1Vβ1)×(U¯α2Vβ2)].

Using (4.7), Theorem 3.1 and Remark 3.1, we deduce

(4.8) i(T,Ω)=i(T,Q)i(T,Ω0)i(T,Ω22)=1.

In conclusion, the existence property of the index together with (4.8) and Theorem 3.1 ensure that T has at least two non-trivial positive fixed points located in the disjoint open sets Ω and (U¯α1V¯β1)×(U¯α2V¯β2). ∎

Remark 4.2.

In the proof of Theorem 4.4 we obtain that i(T,Ω0)=1, which ensures that T has a fixed point in Ω0. Hence, problem (1.6) has a third solution, but it may be the trivial one.

Remark 4.3.

When we speak about two or three solutions, it is understood that they are supposed to be distinct. In case of systems (of two equations), where by a solution we mean a pair (u1,u2), two solutions are distinct if they differ at least on one of their components, not necessarily on both. Thus regarding Theorem 4.4, there is the possibility to have v2=u2 in case where v1(η)<β1, or v1=u1 in case where v2(η)<β2.

Theorem 4.5.

Under the hypotheses of Theorem 3.2, assume that there exist α~1(0,β1) and α~2(0,β2) such that

(4.9) ϕ11(Mα~1,α21) <α~1,
(4.10) ϕ21(Mα1,α~22) <α~2.

Then problem (1.6) has at least two positive solutions: (u1,u2),(v1,v2)K×K such that

βj<uj(η) and |uj|αj(j=1,2);
α~j<|vj|αj(j=1,2), and v1(η)<β1 or v2(η)<β2.
Proof.

Hypotheses of Theorem 3.2 ensure that conditions of Theorem 3.1 are fulfilled. So, to apply Theorem 4.4, we only have to verify (4.5) and (4.6). But, this is straightforward from (4.9) and (4.10), respectively (see the proof of Theorem 2.3). ∎

Example 4.6.

For any γ>3n2n/3, the system

(4.11) (rn1(u1/1(u1)2))=rn1u12[16γ+(u21/4)2],(rn1(u2/1(u2)2))=rn1u22[16γ+(u11/4)2],u1(0)=u1(1)=0=u2(0)=u2(1),

has at least two positive solutions: (u1,u2),(v1,v2)K×K such that

14<uj(12) and |uj|1(j=1,2);
α~<|vj|1(j=1,2), and v1(12)<14 or v2(12)<14

where α~(0,1/4) is such that

(4.12) α~(16γ+916)[1+α~4(16γ+916)2]12<1.

To see this, first observe that the choice of γ ensures that

(4.13) γn2n[1+(γn2n)2]12>12,

then Theorem 4.5 applies with the following choices:

ϕj(x)=x(1x2)12,fj(r,x1,x2)=xj2[16γ+(xk1/4)2](j,k{1,2} and kj)
η=1/2,α1=α2=1,β1=β2=1/4,α~1=α~2=α~ (see (4.12)).

Indeed, using that

ϕj1(x)=x(1+x2)12,mα,βjm(1, 1),(1/4, 1/4)j=γ(j{1,2}),
Mα~1,α21=Mα1,α~22=α~2(16γ+916),

together with (4.13) and (4.12), hypotheses of Theorem 4.5 are easily checked by direct computations. Notice that none of the functions fj(r,x1,x2) is nondecreasing with respect to xk (j,k{1,2} and kj). As a consequence of the above, we have that, for any γ>3n2n/3, the Dirichlet system in the unit ball n,

v1=v12[16γ+(v21/4)2],v2=v22[16γ+(v11/4)2],v1=v2=0on 

has at least two radial solutions, each of them having nonnegative and nontrivial components.

References

  • [1] H. Amann, Fixed point equations and nonlinear eigenvalue problems in ordered Banach spaces, SIAM Rev., 18 (1976), 620–709.
  • [2] C. Bereanu, P. Jebelean and P. J. Torres, Positive radial solutions for Dirichlet problems with mean curvature operators in Minkowski space, J. Funct. Anal., 264 (2013), 270–287.
  • [3] P. Candito, U. Guarnotta and R. Livrea, Existence of two solutions for singular Φ-Laplacian problems, Adv. Nonlinear Stud., 22 (2022), 659–683.
  • [4] X. Cheng and H. Lü, Multiplicity of positive solutions for a (p1,p2)– Laplacian system and its applications, Nonlinear Anal. Real World Appl., 13 (2012), 2375–2390.
  • [5] F.J.S.A. Corrêa, M.L. Carvalho, J.V.A. Gonçalves, K.O. Silva, Positive solutions of strongly nonlinear elliptic problems, Asymptot. Anal., 93 (2015), 1–20.
  • [6] R. Dalmasso, Existence of positive radial solutions for a semilinear elliptic system without variational structure, Nonlinear Anal., 67 (2007), 1255–1259.
  • [7] P. Drábek, M. García–Huidobro, R. Manásevich, Positive solutions for a class of equations with a p-Laplace like operator and weights, Nonlinear Anal., 71 (2009), 1281–1300.
  • [8] M. García–Huidobro, R. Manásevich and K. Schmitt, Positive radial solutions of quasilinear elliptic partial differential equations on a ball, Nonlinear Anal., 35 (1999), 175–190.
  • [9] M. García-Huidobro, R. Manásevich and P. Ubilla, Existence of positive solutions for some Dirichlet problems with an asymptotically homogeneous operator, Electron. J. Differential Equations, 1995 (1995), 1–22.
  • [10] M. García-Huidobro, R. Manásevich and F. Zanolin, Infinitely many solutions for a Dirichlet problem with a nonhomogeneous p-Laplacian-like operator in a ball, Adv. Differential Equations, 2 (1997), 203–230.
  • [11] A. Granas and J. Dugundji, Fixed Point Theory, Springer-Verlag, New York, 2003.
  • [12] D. Guo and V. Lakshmikantham, Nonlinear Problems in Abstract Cones. Academic Press, San Diego, 1988.
  • [13] D. Gurban and P. Jebelean, Positive radial solutions for systems with mean curvature operator in Minkowski space, Rend. Istit. Mat. Univ. Trieste, 49 (2017), 245–264.
  • [14] D. Gurban and P. Jebelean, Positive radial solutions for multiparameter Dirichlet systems with mean curvature operator in Minkowski space and Lane-Emden type nonlinearities, J. Differential Equations, 266 (2019), 5377–5396.
  • [15] Y.-H. Lee, Existence of multiple positive radial solutions for a semilinear elliptic system on an unbounded domain, Nonlinear Anal., 47 (2001), 3649–3660.
  • [16] R. Precup, A vector version of Krasnosel’skiĭ’s fixed point theorem in cones and positive periodic solutions of nonlinear systems, J. Fixed Point Theory Appl., 2 (2007), 141–151.
  • [17] R. Precup and J. Rodríguez-López, Multiplicity results for operator systems via fixed point index, Results Math. 74 (2019), 1–14.
  • [18] R. Precup and J. Rodríguez-López, Positive radial solutions for Dirichlet problems via a Harnack-type inequality, Math. Meth. Appl. Sci., (2022), 1–14. doi:10.1002/mma.8682
  • [19] J. Rodríguez-López, A fixed point index approach to Krasnosel’skiĭ-Precup fixed point theorem in cones and applications, Nonlinear Anal., 226 (2023), No. 113138, pp. 19.
  • [20] K. Uhlenbeck, Regularity for a class of non-linear elliptic systems, Acta Math., 138 (1977), 219–240.
  • [21] W. Walter, Differential and Integral Inequalities, Springer-Verlag, New York, 1970.
  • [22] H. Zou, A priori estimates for a semilinear elliptic system without variational structure and their applications, Math. Ann., 323 (2002), 713–735.
2024

Related Posts