Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

The Gelfand problem for the Infinity Laplacian

  • We study the asymptotic behavior as p of the Gelfand problem

    {Δpu=λeuin ΩRnu=0on Ω.

    Under an appropriate rescaling on u and λ, we prove uniform convergence of solutions of the Gelfand problem to solutions of

    {min{|u|Λeu,Δu}=0in Ω,u=0 on Ω.

    We discuss existence, non-existence, and multiplicity of solutions of the limit problem in terms of Λ.

    Citation: Fernando Charro, Byungjae Son, Peiyong Wang. The Gelfand problem for the Infinity Laplacian[J]. Mathematics in Engineering, 2023, 5(2): 1-28. doi: 10.3934/mine.2023022

    Related Papers:

    [1] Karl K. Brustad, Erik Lindgren, Peter Lindqvist . The infinity-Laplacian in smooth convex domains and in a square. Mathematics in Engineering, 2023, 5(4): 1-16. doi: 10.3934/mine.2023080
    [2] Federico Cluni, Vittorio Gusella, Dimitri Mugnai, Edoardo Proietti Lippi, Patrizia Pucci . A mixed operator approach to peridynamics. Mathematics in Engineering, 2023, 5(5): 1-22. doi: 10.3934/mine.2023082
    [3] Antonio Iannizzotto, Giovanni Porru . Optimization problems in rearrangement classes for fractional p-Laplacian equations. Mathematics in Engineering, 2025, 7(1): 13-34. doi: 10.3934/mine.2025002
    [4] Catharine W. K. Lo, José Francisco Rodrigues . On the obstacle problem in fractional generalised Orlicz spaces. Mathematics in Engineering, 2024, 6(5): 676-704. doi: 10.3934/mine.2024026
    [5] Chao Xia, Jiabin Yin . Two overdetermined problems for anisotropic p-Laplacian. Mathematics in Engineering, 2022, 4(2): 1-18. doi: 10.3934/mine.2022015
    [6] Antonio Greco, Francesco Pisanu . Improvements on overdetermined problems associated to the p-Laplacian. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022017
    [7] María Ángeles García-Ferrero, Angkana Rüland . Strong unique continuation for the higher order fractional Laplacian. Mathematics in Engineering, 2019, 1(4): 715-774. doi: 10.3934/mine.2019.4.715
    [8] David Cruz-Uribe, Michael Penrod, Scott Rodney . Poincaré inequalities and Neumann problems for the variable exponent setting. Mathematics in Engineering, 2022, 4(5): 1-22. doi: 10.3934/mine.2022036
    [9] Stefano Biagi, Serena Dipierro, Enrico Valdinoci, Eugenio Vecchi . A Hong-Krahn-Szegö inequality for mixed local and nonlocal operators. Mathematics in Engineering, 2023, 5(1): 1-25. doi: 10.3934/mine.2023014
    [10] Patrizia Pucci, Letizia Temperini . On the concentration–compactness principle for Folland–Stein spaces and for fractional horizontal Sobolev spaces. Mathematics in Engineering, 2023, 5(1): 1-21. doi: 10.3934/mine.2023007
  • We study the asymptotic behavior as p of the Gelfand problem

    {Δpu=λeuin ΩRnu=0on Ω.

    Under an appropriate rescaling on u and λ, we prove uniform convergence of solutions of the Gelfand problem to solutions of

    {min{|u|Λeu,Δu}=0in Ω,u=0 on Ω.

    We discuss existence, non-existence, and multiplicity of solutions of the limit problem in terms of Λ.



    Dedicated to the memory of Ireneo Peral, with love and admiration.

    We are interested in the asymptotic behavior as p of sequences of solutions of the problem

    {Δpu=λeuin ΩRnu=0on Ω. (1.1)

    In the case p=2, problem (1.1) is known as the Liouville-Bratu-Gelfand problem [5,22,37]; see also [15,26]. It appears in connection with prescribed Gaussian curvature problems [9,37], emission of electricity from hot bodies [40], and the equilibrium of gas spheres and the structure of stars [8,17,42]. Problem (1.1) with p=2 was also studied by Barenblatt in relation to combustion theory in a volume edited by Gelfand [22]. For general p, problem (1.1) is often known in the literature as the "Gelfand problem" or a "Gelfand-type problem". It was studied by García-Azorero, Peral, and Puel in [20,21]; see also [7,24,41] and the references therein.

    The asymptotic study of p-Laplacian problems as p offers a qualitative and quantitative understanding of their solution sets for large p, see [4,10,11,12,13,18,30]. Additionally, they have been used in [23] to obtain optimal bounds for the diameter of manifolds in terms of their curvature.

    In [4,10,11,12,13,18,30], the authors study limits of p-Laplacian equations with power-type right-hand sides and combinations of these. In all these cases, the parameter λ is allowed to vary with p in order to get nontrivial limits of sequences {uλ,p}p of solutions to the corresponding p-Laplacian problem; namely,

    λ1/ppΛanduλp,puas p.

    With an exponential right-hand side, the solution sets change more drastically as p and more severe rescalings become necessary. To take limits in (1.1), we consider

    {Δpuλp,p=λpeuλp,pin Ωuλp,p=0on Ω, (1.2)

    with the rescaling

    λ1/pppΛas p. (1.3)

    Under this normalization, we prove that any uniform limit

    uλp,ppuas p (1.4)

    is a viscosity solution of the limit problem

    {min{|u|Λeu,Δu}=0in Ω,u=0on Ω. (1.5)

    It is worth noting that in [38], the authors consider problem (1.1) without the rescalings (1.3) and (1.4). They obtain that, regardless of λ, the solutions up converge uniformly as p to the unique viscosity solution of

    {min{|u|1,Δu}=0in Ω,u=0on Ω,

    which is the distance function to the boundary of the domain. As the authors of [38] acknowledge in their paper, this result is not unexpected since for each nonnegative function fL(Ω){0}, the sequence of unique solutions of

    {Δpvp=f(x)in Ωvp=0on Ω,

    converges uniformly in Ω to the distance function to the boundary of the domain; see [3,27,31]. This highlights a critical feature of these problems, a precise scaling between u and λ that balances reaction and diffusion and produces a nontrivial limit problem.

    Therefore, in this paper, we prove passage to the limit of the sequence of minimal solutions of problem (1.2) under the rescaling (1.3), (1.4). Furthermore, we show that the resulting limit is a minimal solution of (1.5). Note that the fact that the limit solution is minimal is nontrivial; in principle, limit and minimal solutions could differ. To prove this, we use a comparison principle for "small solutions" of problem (1.5), which we prove in Section 4. As it turns out, minimal solutions to problem (1.5) are "small" in the sense of this comparison principle. To the best of our knowledge, no corresponding comparison and uniqueness results for small solutions were known in the literature for p<.

    In Section 8, we find a second solution to the limit problem (1.5) under certain geometric assumptions on the domain Ω. Furthermore, we show that both solutions lie on an explicit curve of solutions. Some examples of domains satisfying the geometric condition are the ball, the annulus, and the stadium (convex hull of two balls of the same radius); squares or ellipses do not verify the condition. We conjecture that this second solution is a limit of appropriately rescaled mountain-pass solutions of (1.2).

    The paper is organized as follows. In Section 2, we provide some necessary preliminaries, and Section 3 formally introduces the limit problem. We have chosen to introduce the limit problem before proving any convergence results to streamline the presentation. In Section 4, we prove the comparison principle for small solutions of the limit equation (1.5). Section 5 concerns non-existence of solutions to (1.5) for large values of Λ. In Section 6, we find a branch of minimal solutions to (1.5) up to a maximal Λ. Section 7 discusses uniform convergence as p of p-minimal solutions to minimal solutions of (1.5). Finally, in Section 8, we show the multiplicity result and exhibit a curve of explicit solutions under a geometric condition on the domain.

    In this section, we state some necessary preliminaries and notation. First, let us recall that weak solutions of problem (1.1) are also viscosity solutions. The proof, which we omit here, follows [30,Lemma 1.8]; see also [3].

    Lemma 2.1. If u is a continuous weak solution of (1.1), then it is a viscosity solution of the same problem, rewritten as

    {Fp(u,D2u)=λeuin Ωu=0on Ω,

    where

    Fp(ξ,X)=|ξ|p2trace((I+(p2)ξξ|ξ|2)X). (2.1)

    The divergence form of the p-Laplacian, i.e., div(|u|p2u), is better suited for variational techniques, while the expanded form (2.1) is preferable in the viscosity framework. In the sequel, we will always consider the most suitable form without further mention.

    In [31] the problem

    {Δpvp=1in ΩvpW1,p0(Ω)

    is studied in connection with torsional creep problems when Ω is a general bounded domain. Since we are interested in the case p, we can assume p>n without loss of generality. Then every function in vpW1,p0(Ω) can be considered continuous in ¯Ω and 0 on the boundary in the classical sense. The existence result we will need below is the following. We refer the interested reader to [31] and [27,Theorem 3.11 and Remark 4.23] for the proof.

    Proposition 2.2. Let Ω be a bounded domain and n<p<. Then, there exists a unique solution vpW1,p0(Ω)C(¯Ω) of the p-torsion problem

    {Δpvp=1in Ωvp=0on Ω, (2.2)

    and vp converge uniformly as p to the unique viscosity solution to

    {min{|v|1,Δv}=0in Ω,v=0on Ω.

    Moreover, v(x)=dist(x,Ω).

    The uniqueness of the solution in Proposition 2.2 follows from the following comparison principle.

    Lemma 2.3. Let f:ΩR be a continuous, bounded, and positive function. Suppose that u,v:¯ΩR are bounded, u is upper semicontinuous and v is lower semicontinuous in ¯Ω. If u and v are, respectively, a viscosity sub- and supersolution of

    min{|w|f(x),Δw}=0in Ω,

    and uv on Ω, then uv in Ω.

    We refer the interested reader for instance to [27,Theorem 4.18 and Remark 4.23] and also [25,Theorem 2.1] (for the proof of [27,Theorem 4.18], notice that every -superharmonic function is Lipschitz continuous, see [35]).

    We also need some facts about first eigenvalues and eigenfunctions of the p-Laplacian. Let us recall that the first eigenvalue λ1(p;Ω) is characterized by the nonlinear Rayleigh quotient

    λ1(p;Ω)=infϕW1,p0(Ω)Ω|ϕ|pdxΩ|ϕ|pdx.

    In [32] (see also [33]), it is proved that the first eigenvalue of the p-Laplacian is simple (that is, the first eigenfunction is unique up to multiplication by constants) when Ω is a bounded domain; see also [2,19,39] and the references in [32]. Moreover, it is also proved in [32] that in a bounded domain, only the first eigenfunction is positive and that the first eigenvalue is isolated (there exists ϵ>0 such that there are no eigenvalues in (λ1,λ1+ϵ]).

    Proposition 2.4 ([32]). Let Ω be a bounded domain and n<p<. Then, there exists a solution ψpW1,p0(Ω)C(¯Ω) of

    {Δpψp=λ1(p;Ω)|ψp|p2ψpin Ωψp=0on Ω.

    Moreover, λ1(p;Ω) is simple and isolated.

    Lastly, we recall the behavior as p of the first eigenvalue of the p-Laplacian, see [30] for the proof.

    Lemma 2.5. limpλ1(p,Ω)1/p=Λ1(Ω)=dist(,Ω)1.

    We denote the first -eigenvalue by Λ1(Ω), see [30].

    In the present section, we characterize uniform limits of appropriate rescalings of solutions of (1.2) as solutions of a PDE. See [4,10,11,12,13,18,30] for related results.

    Proposition 3.1. Consider a sequence {(λp,uλp,p)}p of solutions of (1.2) and assume

    limpλ1/ppp=Λ.

    Then, any uniform limit

    uΛ=limpuλp,pp

    is a viscosity solution of the problem

    {min{|u|Λeu,Δu}=0in Ω,u=0on Ω. (3.1)

    Proof. Consider a point x0Ω and a function ϕC2(Ω) such that uΛϕ has a strict local minimum at x0. As uΛ is the uniform limit of uλp,p/p, there exists a sequence of points xpx0 such that uλp,ppϕ attains a local minimum at xp for each p. As uλp,p is a continuous weak solution of (1.2), it is also a viscosity solution and a supersolution. Then, we get

    (p2)pp1|ϕ(xp)|p4{|ϕ(xp)|2p2Δϕ(xp)+D2ϕ(xp)ϕ(xp),ϕ(xp)}=pp1Δpϕ(xp)λpeuλ,p(xp).

    Rearranging terms, we obtain

    (p2)[|ϕ(xp)|(λppp1euλ,p(xp))1p4]p4{|ϕ(xp)|2p2Δϕ(xp)+D2ϕ(xp)ϕ(xp),ϕ(xp)}1.

    If we suppose that |ϕ(x0)|<ΛeuΛ(x0) we obtain a contradiction letting p in the previous inequality. Thus, it must be

    |ϕ(x0)|ΛeuΛ(x0)0. (3.2)

    We also have that

    Δϕ(x0)=D2ϕ(x0)ϕ(x0),ϕ(x0)0, (3.3)

    because we would get a contradiction otherwise. Therefore, we can put together (3.2) and (3.3) writing

    min{|ϕ(x0)|ΛeuΛ(x0),Δϕ(x0)}0,

    and conclude that uΛ is a viscosity supersolution of (3.1).

    It remains to show that uΛ is a viscosity subsolution of the limit equation (3.1). More precisely, we have to show that, for each x0Ω and ϕC2(Ω) such that uΛϕ attains a strict local maximum at x0 (note that x0 and ϕ are not the same than before) we have

    min{|ϕ(x0)|ΛeuΛ(x0),Δϕ(x0)}0.

    We can suppose that

    |ϕ(x0)|>ΛeuΛ(x0),

    since we are done otherwise. Again, the uniform convergence of uλ,p/p to uΛ provides a sequence of points xpx0 which are local maxima of uλ,ppϕ. Recalling the definition of viscosity subsolution we have

    (p2)[|ϕ(xp)|(λppp1euλ,p(xp))1p4]p4{|ϕ(xp)|2p2Δϕ(xp)+D2ϕ(xp)ϕ(xp),ϕ(xp)}1,

    for each p. Letting p, we find Δϕ(x0)0, or else we get a contradiction.

    In the previous argument, the fact that euΛ(x0) is strictly positive independently of the value of uΛ(x0) makes a difference with the case with a power-type right-hand side (see [10,11,12,13,18,30]), where one needs to make sure that uΛ>0 in Ω. Furthermore, in the power-type right-hand side case, one can consider sign-changing solutions, see [10,29] and get a more involved limit equation that takes into account sign changes. In the next result, we show that all solutions to the limit problem (3.1) are positive. Moreover, we show that solutions cannot be arbitrarily small for every given Λ and must grow (at least) linearly from the boundary.

    Proposition 3.2. Let ΩRn be a bounded domain and Λ>0. Then, every solution uΛ of (3.1) verifies

    uΛΛdist(,Ω)in Ω.

    In particular, every solution of (3.1) is strictly positive and satisfies the estimate

    uΛL(Ω)ΛΛ1(Ω)1.

    Proof. Let uΛ be a solution of (3.1). Then, uΛ0 in Ω by Lemma 2.3. Let us show that

    min{|uΛ|Λ,ΔuΛ}0in Ω

    in the viscosity sense. To see this, consider x0Ω and ϕC2 such that uΛϕ has a minimum at x0. Since uΛ(x) is a solution of (3.1), we have

    min{|ϕ(x0)|ΛeuΛ(x0),Δϕ(x0)}0in Ω.

    We deduce Δϕ(x0)0 and |ϕ(x0)|ΛeuΛ(x0)Λ and we get

    min{|ϕ(x0)|Λ,Δϕ(x0)}0in Ω

    as desired.

    On the other hand, vΛ(x)=Λdist(x,Ω) is the unique viscosity solution of

    min{|vΛ|Λ,ΔvΛ}=0in Ω.

    Then, one gets uΛvΛ=Λdist(,Ω) by comparison, see Lemma 2.3.

    In this section, we prove a comparison principle for small solutions of the limit equation (1.5). This result is interesting for two main reasons. Firstly, Eq (1.5) is not proper in the terminology of [14], a basic requirement for comparison. Secondly, based on the multiplicity results for the p-Laplacian equation (1.2), see [20,21], one cannot expect comparison to hold in general. The key idea is a change of variables that allows us to obtain a proper equation for solutions with u<1. Remarkably, minimal solutions of (1.5) verify this condition (see Section 6 below), and we can conclude they are the only ones with u<1. The change of variables we use here is the same that was used to prove comparison for the limit problem with concave right-hand side in [12].

    We prove a more general result with a "right-hand" side f(u) that satisfies a hypothesis reminiscent of the celebrated Brezis-Oswald condition, see [6] and Remark 4.2 below.

    Theorem 4.1. Let f:RR be a continuous function for which there exist c(0,] and q(0,1) such that

    f(t)tqispositiveandnonincreasingforall t(0,c). (4.1)

    Let ΩRn be a bounded domain and let u,vC(¯Ω) with max{u,v}<c be, respectively, a positive viscosity sub- and supersolution of

    min{|w|f(w),Δw}=0in Ω. (4.2)

    Then, whenever uv on Ω, we have uv in ¯Ω.

    Remark 4.2. It is possible to prove a comparison principle for Eq (4.2) under the Brezis-Oswald [6] condition

    f(t)tis decreasing for all t>0.

    Under this condition, the power-type change of variables used in [12] and in the proof of Theorem 4.1 no longer applies. Instead, we need a logarithmic change of variables, similarly to the comparison principle for the eigenvalue problem for the infinity Laplacian in [30]. However, a viscosity comparison principle obtained through a logarithmic change of variables requires that either the sub- or the supersolution are strictly positive in ¯Ω and does not allow us to conclude uniqueness of solutions for the Dirichlet problem with homogeneous boundary data, which our result does.

    Before going into the proof of Theorem 4.1, let us discuss an important consequence of Theorem 4.1, the uniqueness of "small" solutions of problem (3.1).

    Corollary 4.3. Let ΩRn be a bounded domain. For every Λ>0, the problem

    {min{|u|Λeu,Δu}=0in Ω,u=0on Ω, (4.3)

    has at most one viscosity solution with u<1.

    Proof of Corollary 4.3. Suppose for the sake of contradiction that there are two viscosity solutions, u,v of (4.3) with max{u,v}<1. Notice that both u and v are strictly positive in Ω by Proposition 3.2. In this case we have f(t)=Λet and (4.1) is satisfied with c=q for every q(0,1). Then, we can choose q(0,1) such that max{u,v}<q<1, and all the hypotheses of Theorem 4.1 are satisfied. Because u=v on Ω, we conclude uv.

    We devote the rest of the section to the proof of Theorem 4.1. In the next lemma we apply a change of variables to Eq (4.2) to obtain a proper equation for small solutions.

    Lemma 4.4. Let q(0,1) and let v be a positive viscosity supersolution (respectively, subsolution) of (4.2) in Ω. Then, ˜v(x)=v1q(x) is a viscosity supersolution (subsolution) of

    min{|˜w(x)|(1q)f(˜w(x)11q)˜w(x)q1q,Δ˜w(x)q1q|˜w(x)|4˜w(x)}=0 (4.4)

    in every subdomain U compactly contained in Ω.

    Proof. Let ˜ϕC2(Ω) touch ˜v from below at x0Ω. If we define ϕ(x)=˜ϕ(x)11q, then ϕ touches v from below at x0. Note that ϕ(x) is C2 in a neighborhood of x0, since v>0 in Ω implies ˜ϕ(x)>0 around x0. Then

    ϕ(x0)=11q˜ϕ(x0)q1q˜ϕ(x0),D2ϕ(x0)=11q˜ϕ(x0)q1qD2˜ϕ(x0)+q(1q)2˜ϕ(x0)2q11q˜ϕ(x0)˜ϕ(x0).

    Because v is a viscosity supersolution of (4.2) and ϕ(x0)=v(x0)>0, we have

    0min{|ϕ(x0)|f(ϕ(x0)),D2ϕ(x0)ϕ(x0),ϕ(x0)}=min{11q˜ϕ(x0)q1q(|˜ϕ(x0)|(1q)f(˜ϕ(x0)11q)˜ϕ(x0)q1q),(11q˜ϕ(x0)q1q)3(Δ˜ϕ(x0)+q1q|˜ϕ(x0)|4˜ϕ(x0))}.

    Therefore,

    min{|˜ϕ(x0)|(1q)f(˜ϕ(x0)11q)˜ϕ(x0)q1q,Δ˜ϕ(x0)q1q|˜ϕ(x0)|4˜ϕ(x0)}0,

    that is, ˜v is a viscosity supersolution of (4.4). The subsolution case is analogous.

    Equation (4.4) is given by the functional

    F:R+×Rn×SnR(t,p,X)min{|p|(1q)f(t11q)tq1q,Xp,pq1q|p|4t},

    which is degenerate elliptic and non-decreasing in t for 0<t<c1q by hypothesis (4.1). Under these conditions, it is well-known (see [14,Section 5.C]) that it is possible to establish a comparison principle when the supersolution or the subsolution are strict. In the next lemma we show that we can find a perturbation of the supersolution that is a strict supersolution, see [12,27,30] for related constructions.

    Lemma 4.5. Consider a subdomain U compactly contained in Ω, and q(0,1),c>0 as in (4.1). Let ˜v>0 with ˜v<c1q be a viscosity supersolution of (4.4) in U. Define

    ˜vϵ(x)=(1+ϵ)(˜v(x)+ϵ). (4.5)

    Then, ˜vϵ˜v uniformly in ¯U as ϵ0, and for every ϵ>0 small enough, there exists a positive constant C=C(ϵ,q,˜v) such that

    min{|˜vϵ(x)|(1q)f(˜vϵ(x)11q)˜vϵ(x)q1q,Δ˜vϵ(x)q1q|˜vϵ(x)|4˜vϵ(x)}C>0in U, (4.6)

    in the viscosity sense, that is, ˜vϵ is a strict viscosity supersolution of (4.4) in U with ˜vϵ<c1q.

    Proof. Let ˜ϕϵC2 touch ˜vϵ(x) from below at x0U. Define

    ˜ϕ(x)=11+ϵ˜ϕϵ(x)ϵ,

    which clearly touches ˜v(x) from below at x0. Then,

    ˜ϕ(x0)=(1+ϵ)1˜ϕϵ(x0)andD2˜ϕ(x0)=(1+ϵ)1D2˜ϕϵ(x0). (4.7)

    Since ˜v(x) is a viscosity supersolution of (4.4) in U, we deduce

    |˜ϕ(x0)|(1q)f(˜v(x0)11q)˜v(x0)q1q, (4.8)

    and

    D2˜ϕ(x0)˜ϕ(x0),˜ϕ(x0)q1q|˜ϕ(x0)|4˜v(x0)0. (4.9)

    In the sequel we assume ϵ small enough so that ˜v<˜vϵ=(1+ϵ)(˜v+ϵ)<c1q. Then, from (4.1), (4.5), (4.7) and (4.8), we obtain

    |˜ϕϵ(x0)|(1q)f(˜vϵ(x0)11q)˜vϵ(x0)q1qϵ(1q)f(˜v(x0)11q)˜v(x0)q1q+(1q)(f(˜v(x0)11q)˜v(x0)q1qf(˜vϵ(x0)11q)˜vϵ(x0)q1q)ϵ(1q)f(˜v11q)˜vq1q. (4.10)

    Similarly, from (4.1), (4.5), (4.7), (4.8), and (4.9) we arrive at

    D2˜ϕϵ(x0)˜ϕϵ(x0),˜ϕϵ(x0)q1q|˜ϕϵ(x0)|4˜vϵ(x0)(1+ϵ)3q1q(1˜v(x0)1˜v(x0)+ϵ)|˜ϕ(x0)|4ϵ(1+ϵ)3q(1q)3˜v(˜v+ϵ)(f(˜v(x0)11q)˜v(x0)q1q)4ϵ(1+ϵ)3q(1q)3˜v(˜v+ϵ)(f(˜v11q)˜vq1q)4. (4.11)

    Finally, we get (4.6) from (4.10) and (4.11) as desired, which concludes the proof.

    Proof of Theorem 4.1. Since uvC(¯Ω) and ¯Ω is compact, uv attains its maximum in ¯Ω. Suppose, for the sake of contradiction, that max¯Ω(uv)>0. Let

    ˜u(x)=u(x)1q,˜v(x)=v(x)1q,

    and define ˜vϵ(x) as in (4.5). Notice that uv0 on Ω gives

    ˜u˜vϵ=˜u(1+ϵ)˜v(1+ϵ)ϵ<0on Ω.

    Moreover, by uniform convergence, we have max¯Ω(˜u˜vϵ)>0 for ϵ small enough. Therefore, we can fix ϵ>0 small as in Lemma 4.5 for the rest of the proof and assume there exists U compactly contained in Ω that contains all maximum points of ˜u˜vϵ. We have proved in Lemmas 4.4 and 4.5 that ˜u and ˜vϵ are, respectively, a viscosity subsolution and strict supersolution of (4.4) in U.

    For every τ>0, let (xτ,yτ) be a maximum point of ˜u(x)˜vϵ(y)τ2|xy|2 in ¯ΩׯΩ. By the compactness of ¯Ω, we can assume that xτˆx as τ for some ˆx¯Ω (notice that also yτˆx). Then, [14,Proposition 3.7] implies that ˆx is a maximum point of ˜u˜vϵ and, therefore, an interior point of U. We also have that

    limτ(˜u(xτ)˜vϵ(yτ)τ2|xτyτ|2)=˜u(ˆx)˜vϵ(ˆx)>0,

    and, consequently, both xτ and yτ are interior points of U for τ large enough and

    ˜u(xτ)˜vϵ(yτ)τ2|xτyτ|2>0. (4.12)

    The definition of viscosity solution and the maximum principle for semicontinuous functions, see [14], imply that there exist symmetric matrices Xτ, Yτ with XτYτ such that

    min{τ|xτyτ|(1q)f(˜u(xτ)11q)˜u(xτ)q1q,τ2Xτ(xτyτ),(xτyτ)q1qτ4|xτyτ|4˜u(xτ)}0,

    and

    min{τ|xτyτ|(1q)f(˜vϵ(yτ)11q)˜vϵ(yτ)q1q,τ2Yτ(xτyτ),(xτyτ)q1qτ4|xτyτ|4˜vϵ(yτ)}C(ϵ,q,˜v)>0.

    Subtracting both equations, we get

    0<C(ϵ,q,˜v)min{τ|xτyτ|(1q)f(˜vϵ(yτ)11q)˜vϵ(yτ)q1q,τ2Yτ(xτyτ),(xτyτ)q1qτ4|xτyτ|4˜vϵ(yτ)} (4.13)
    min{τ|xτyτ|(1q)f(˜u(xτ)11q)˜u(xτ)q1q,τ2Xτ(xτyτ),(xτyτ)q1qτ4|xτyτ|4˜u(xτ)}. (4.14)

    We consider four cases, depending on the values where the minima in (4.13) and (4.14) are attained. In all cases we obtain a contradiction using that XτYτ and ˜vϵ(yτ)˜u(xτ), which follows from (4.12).

    1) Both minima are attained by the first terms and (4.1) implies a contradiction, i.e.,

    0<C(ϵ,q,˜v)(1q)(f(˜u(xτ)11q)˜u(xτ)q1qf(˜vϵ(yτ)11q)˜vϵ(yτ)q1q)0.

    2) Both minima are attained by the second terms. Then,

    0<C(ϵ,q,˜v)τ2(YτXτ)(xτyτ),(xτyτ)+q1qτ4|xτyτ|4(1˜u(xτ)1˜vϵ(yτ))0,

    a contradiction.

    3) The minima in (4.13) and (4.14) are attained by the second and first term, respectively. This case can be reduced to case (1) above and we again obtain a contradiction. Namely,

    0<C(ϵ,q,˜v)τ2Yτ(xτyτ),(xτyτ)q1qτ4|xτyτ|4˜vϵ(yτ)τ|xτyτ|+(1q)f(˜u(xτ)11q)˜u(xτ)q1q(1q)(f(˜u(xτ)11q)˜u(xτ)q1qf(˜vϵ(yτ)11q)˜vϵ(yτ)q1q)0.

    4) Finally, if the minima in (4.13) and (4.14) are respectively attained by the first and second term, we obtain a contradiction as in case (2) above, i.e.,

    0<C(ϵ,q,˜v)τ|xτyτ|(1q)f(˜vϵ(yτ)11q)˜vϵ(yτ)q1q+τ2Xτ(xτyτ),(xτyτ)+q1qτ4|xτyτ|4˜u(xτ)τ2(YτXτ)(xτyτ),(xτyτ)+q1qτ4|xτyτ|4(1˜u(xτ)1˜vϵ(yτ))0.

    {}Since all the alternatives lead to a contradiction, the proof is complete.

    We show here that due to the structure of the limit problem (3.1), there exists a threshold Λmax beyond which the problem has no solutions.

    Proposition 5.1. Let ΩRn be a bounded domain. Problem (3.1) has no solutions for Λ>Λmax, where

    Λmax=e1Λ1(Ω), (5.1)

    and Λ1(Ω)=dist(,Ω)1 is the first -eigenvalue, see [30].

    Proof. Define μ=Λ1(Ω)+ϵ with ϵ>0. Suppose for contradiction that problem (3.1) has a solution uΛ for some Λ>e1μ.

    First we are going to use this uΛ to construct a supersolution to the eigenvalue problem with parameter μ. More precisely, we are going to show that

    min{|uΛ|μuΛ,ΔuΛ}0in Ω (5.2)

    in the viscosity sense. To this aim, let x0Ω and ϕC2 such that uΛϕ has a minimum in x0. Since uΛ(x) is a solution of problem (3.1) we have

    min{|ϕ(x0)|ΛeuΛ(x0),Δϕ(x0)}0in Ω.

    We deduce that Δϕ(x0)0 and |ϕ(x0)|ΛeuΛ(x0). Hence,

    |ϕ(x0)|μuΛ(x0)ΛeuΛ(x0)μuΛ(x0).

    To deduce (5.2) it is enough to show that

    mintRΦΛ(t)0whereΦΛ(t)=Λetμt.

    It is elementary to check that the function ΦΛ is convex and has a unique minimum point at tmin=log(μΛ1). Notice that limt±ΦΛ(t)=+, and hence tmin is a global minimum. Then, it is easy to check that Λ>e1μ implies ΦΛ(tmin)0.

    Next, we notice that any first -eigenfunction is a subsolution of the eigenvalue problem with parameter μ. So, let v be a first eigenfunction, that is, a solution of

    {min{|v|Λ1(Ω)v,Δv}=0in Ω,v>0in Ωv=0on Ω

    normalized in such a way that v<e1. Clearly, by definition of μ,

    min{|v|μv,Δv}0in Ω.

    Now, we have to show that uΛ and v are ordered, namely, that 0<vuΛ in Ω. Indeed, using that v<e1 and Λ1(Ω)<μ<Λe, it is easy to see that

    min{|v|Λ,Δv}0in Ω,

    and using that euΛ(x)1 in Ω one gets

    min{|uΛ|Λ,ΔuΛ}0in Ω.

    As v=uΛ=0 on Ω, we get 0<vuΛ by comparison, see Lemma 2.3.

    So far, we have a subsolution v and a supersolution uΛ of the eigenvalue problem

    min{|w|μw,Δw}=0in Ω (5.3)

    which verify 0<vuΛ. Next we claim that it is possible to construct a solution of (5.3) iterating between v and uΛ. The argument finishes noticing that we have constructed a positive eigenfunction associated to μ=Λ1+ϵ, which is a contradiction with the fact that Λ1 is isolated (see [29,Theorem 8.1] and [30,Theorem 3.1]). Since the argument above works for every ϵ>0, we conclude that there is no solution of (3.1) for Λ>Λmax.

    We conclude by proving the claim. First, define w1(x), viscosity solution of

    {min{|w1|μv,Δw1}=0in Ωw1=0on Ω.

    To prove that such a w1 exists, notice that v is a subsolution of the problem and that uΛ is a supersolution, since, from (5.2) and vuΛ we deduce

    min{|uΛ|μv,ΔuΛ}0.

    Then, we can apply the comparison principle in Lemma 2.3 as above and apply the Perron method ([14,Theorem 4.1]), to get a unique w1 such that

    vw1uΛin Ω.

    Then, we define w2, the solution of

    {min{|w2|μw1,Δw2}=0in Ωw2=0on Ω.

    In this case, w1 is a subsolution and uΛ is a supersolution, since

    min{|w1|μv,Δw1}=0min{|w1|μw1,Δw1}0,

    while

    min{|uΛ|μuΛ,ΔuΛ}0min{|uΛ|μw1,ΔuΛ}0.

    As w1=uΛ=0 on Ω, by comparison and the Perron method, we obtain that there exists a unique w2 satisfying

    vw1w2uΛin Ω.

    Iterating this procedure, we construct a non-decreasing sequence

    vw1w2wk1wkuΛ

    of solutions of

    {min{|wk|μwk1,Δwk}=0in Ωwk=0on Ω. (5.4)

    Notice that wk is uniformly bounded by construction. On the other hand, as Δwk0 in Ω, we have (see [34,35] and also [28] for a related construction) that

    |wk(x)|wk(x)dist(x,Ω)uΛ(x)dist(x,Ω)a.e. xΩ,

    for all k>1. From there, both wk and wk are uniformly bounded in compact subsets of Ω. We observe that v,uΛ are barriers in Ω for each wk. Hence by the Ascoli-Arzela theorem and the monotonicity of the sequence {wk}, the whole sequence converges uniformly in Ω to some wC(Ω) which verifies w=0 on Ω. Then, we can take limits in the viscosity sense in (5.4) and obtain that the limit w is a viscosity solution of (5.3), which proves the claim.

    In this section we show that for every Λ(0,Λmax] there is a minimal solution of the problem

    {min{|u|Λeu,Δu}=0in Ω,u=0on Ω. (6.1)

    The proof is based on the ideas in [20], although our construction is different in order to take advantage of Corollary 4.3, our result of uniqueness for small solutions (the construction in [20] would only allow us to conclude that the minimal solution satisfies uuΛmax=1, and Corollary 4.3 requires a strict inequality).

    Theorem 6.1. Let ΩRn be a bounded domain. Then, problem (6.1) has a minimal solution uΛ for every Λ(0,Λmax], where Λmax is given by (5.1). Moreover,

    1) We have the estimate

    Λdist(x,Ω)uΛ(x)eΛdist(x,Ω).

    In particular, uΛeΛΛ1(Ω)1<1 for Λ(0,Λmax).

    2) For every Λ(0,Λmax), uΛ is the only solution of (6.1) with u<1.

    3) The branch of minimal solutions is a non-decreasing continuum, in the sense that if 0<Λ<Υ<Λmax, then uΛuΥ and whenever ΥΛ(0,Λmax), then uΥuΛ uniformly.

    Proof. 1) Let u_ and ¯u be the unique viscosity solutions of

    {min{|u_|Λ,Δu_}=0in Ωu_=0on Ω (6.2)

    and

    {min{|¯u|eΛ,Δ¯u}=0in Ω¯u=0on Ω, (6.3)

    respectively. By Proposition 2.2, we have the explicit expressions

    u_(x)=Λdist(x,Ω)and¯u(x)=eΛdist(x,Ω) (6.4)

    and u_¯u follows trivially (alternatively, this can be proved by comparison, Lemma 2.3, using that ¯u is a viscosity supersolution of (6.2)).

    2) Define now u1, viscosity solution of

    {min{|u1|Λeu_,Δu1}=0in Ωu1=0on Ω. (6.5)

    Let us show that

    u_u1¯uin Ω. (6.6)

    First, we prove u1¯u. We aim to show that min{|u1|eΛ,Δu1}0 in the viscosity sense and then apply comparison for Eq (6.3), see Lemma 2.3. Therefore, let x0Ω and ϕC2(Ω) such that u1ϕ attains a local maximum at x0. We can assume that Δϕ(x0)>0 because we are done otherwise. Then, from (6.5), (6.4), and (5.1), we have

    |ϕ(x0)|Λeu_(x0)ΛeΛmaxΛ1(Ω)1<eΛ.

    In order to show that u1u_, we prove that min{|u1|Λ,Δu1}0 in the viscosity sense and then proceed by comparison for Eq (6.2). Indeed, since u1 is a supersolution of (6.5), we have Δu10 and |u1|Λeu_Λ in the viscosity sense, as desired.

    3) For each k0, we define uk+1 as the viscosity solution of

    {min{|uk+1|Λeuk,Δuk+1}=0in Ωuk+1=0on Ω (6.7)

    with u0=u_ and u1 given by (6.5). Let us show that for all k0

    u_ukuk+1¯uin Ω, (6.8)

    that is, the sequence {uk}k0 is non-decreasing and uniformly bounded.

    We prove (6.8) by induction. First, notice that (6.6) proves the case when k=0. Assume (6.8) holds true for k1 and let us prove that ukuk+1. Since uk+1 is, by definition, a viscosity supersolution of (6.7), we have Δuk+10 and |uk+1|ΛeukΛeuk1 in the viscosity sense by the induction hypothesis. Therefore, uk+1 is a viscosity solution of

    min{|uk+1|Λeuk1,Δuk+1}0in Ω.

    By definition, we have min{|uk|Λeuk1,Δuk}=0 and ukuk+1 follows by comparison, see Lemma 2.3 (notice that euk1 is bounded, positive, and continuous, since the -superharmonicity of uk1 imply its Lipschitz continuity, see [35]).

    To prove that uk+1¯u, we show that

    min{|uk+1|eΛ,Δuk+1}0in Ω

    and use comparison for Eq (6.3) (see Lemma 2.3). Therefore, let x0Ω and ϕC2(Ω) such that uk+1ϕ attains a local maximum at x0. Assume that Δϕ(x0)>0 since we are done otherwise. Then, from (6.7), (6.4), (5.1), and the induction hypothesis we get

    |ϕ(x0)|Λeuk(x0)Λe¯ueΛ.

    4) We have obtained a non-decreasing sequence {uk}k0, uniformly bounded by u_ and ¯u given by (6.4). Therefore, we can pass to the limit in the viscosity sense in the same way as in Proposition 5.1 and get a viscosity solution uΛ of problem (6.1) as intended. It is also clear that the solution uΛ we just found is minimal for every Λ(0,Λmax], because any solution of (6.1) could be taken as ¯u in the iteration (note that the function uΛ does not depend on ¯u). Moreover, by (6.4) and (6.8), we have

    Λdist(x,Ω)uΛ(x)eΛdist(x,Ω)in Ω.

    Therefore, by Corollary 4.3, for every Λ(0,Λmax), uΛ is the only solution of (6.1) with u<1.

    5) Let us prove that the branch of minimal solutions is non-decreasing, i.e., uΛuΥ whenever 0<Λ<Υ<Λmax. To this aim, let us just observe that we can repeat the above construction taking ¯u=uΥ and keeping u_(x)=Λdist(x,Ω) as before. In this way, we recover the minimal solution uΛ with the estimate uΛuΥ<1.

    We conclude by showing that the branch of minimal solutions is a continuum. Arguing again as in the proof of Proposition 5.1, we see that, for every Λ(0,Λmax), the uniform limits

    ˆuΛ=limΥΛ+uΥ,andˇuΛ=limΥΛuΥ

    are both viscosity solutions of (6.1) with max{ˆuΛ,ˇuΛ}<1. Therefore ˆuΛˇuΛ by Corollary 4.3, as desired.

    This section shows that uniform limits of appropriately scaled, minimal solutions of

    {Δpu=λeuin ΩRnu=0on Ω (7.1)

    converge to the minimal solutions of the limit problem (6.1), found in Section 6. Observe that the fact that the limit solution is minimal is nontrivial; in principle, a limit solution could be different from the minimal one. Here is where we use the uniqueness results from Section 4. We prove the following.

    Theorem 7.1. Let Λ(0,Λmax), and {λp}p be a sequence such that

    limpλ1/ppp=Λ.

    For each λp, consider uλp,p, the minimal solution of (7.1) for λ=λp. Then,

    uλp,ppuΛ,uniformlyas p,

    where uΛ is the minimal solution of the limit problem (6.1).

    We devote the rest of the section to the proof of Theorem 7.1. In order to obtain estimates that allow us to pass to the limit, we provide an explicit construction of the branch of minimal solutions of (7.1). Although these are rather classic facts, see [20,21], some of our results appear to be new. Additionally, we provide a modified, more streamlined, and systematic construction that exhibits the dependences on p at each step, which is necessary in order to pass to the limit.

    First, we show that problem (7.1) has a minimal solution up to a certain, explicit ˇλp.

    Proposition 7.2. Let ΩRn be a bounded domain and p>n. Then, problem (7.1) has a minimal solution uλ,p(x) for every λ(0,ˇλp], where

    ˇλp=(p1evp)p1 (7.2)

    and vp is given by (2.2). Moreover,

    1) For every λˇλp, we have the estimate

    λ1p1vp(x)uλ,p(x)eλ1p1vp(x)in ¯Ω. (7.3)

    2) For every λˇλp, the minimal solution uλ,p is the only solution of (7.1) with up1.

    3) The branch of minimal solutions is non-decreasing, in the sense that if 0<λ<μˇλp, then uλ,puμ,p in Ω.

    The uniqueness result in part 2 of Proposition 7.2 appears to be new. For the proof, we use the following comparison principle, an adaptation of [1,Lemma 4.1] to problems that are proper (in the sense of [14]) only for "small" sub- and supersolutions.

    Lemma 7.3. Let p>1 and f:RR be a non-negative continuous function for which there exists c(0,] such that

    f(t)tp1isnonincreasingforall t(0,c).

    Assume that u,vW1,p0(Ω)C1(Ω) are positive in Ω, max{u,v}c and

    Δpuf(u)andΔpvf(v)in Ω.

    Then uv in Ω.

    We omit the proof of the lemma since it is a straightforward modification of [1,Lemma 4.1] (note that c= in [1]). We proceed now with the proof of Proposition 7.2.

    Proof of Proposition 7.2. 1) Consider u_ and ¯u, the respective solutions of

    {Δpu_=λin Ωu_=0on Ω

    and

    {Δp¯u=λep1in Ω¯u=0on Ω.

    By the weak comparison principle for the p-Laplacian, we have that 0u_¯u in ¯Ω. Define now u1, solution of

    {Δpu1=λeu_in Ωu1=0on Ω. (7.4)

    We clearly have Δpu1λ=Δpu_. On the other hand, we find u_=λ1p1vp by rescaling, which together with (7.2) yields

    Δpu1λeu_λe(ˇλp)1/(p1)vp=λe(p1)/eΔp¯u.

    Then, by the weak comparison principle we have u_u1¯u in ¯Ω.

    2) Now, for each k1 define uk+1, solution of

    {Δpuk+1=λeukin Ωuk+1=0on Ω

    with u1 defined by (7.4). Let us show by induction that

    u_ukuk+1¯uin ¯Ω

    for all k1. It is easy to see that u_ukuk+1 by comparison. To prove uk+1¯u, notice that the induction hypothesis, the rescaling ¯u=λ1p1evp, and (7.2) yield

    Δpuk+1=λeukλe¯uλe(ˇλp)1/(p1)evp=λep1=Δp¯u.

    Then, uk+1¯u follows by comparison.

    3) We have obtained an increasing sequence {uk}k0, uniformly bounded by u_ and ¯u. Therefore, we can pass to the limit and get a solution uλ,p that satisfies the bounds (7.3). It is also clear that uλ,p is minimal, because any solution of (7.1) could be taken as ¯u in the iterative scheme (note that each uk does not depend on ¯u). Similarly, we see that the branch of minimal solutions is non-decreasing, since whenever λ<μ, we can take ¯u=uμ,p in the construction of uλ,p and obtain uλ,puμ,p.

    4) Finally, let us denote f(t)=λet. It is elementary to see that f(t)/tp1 is non-increasing for 0<t<p1. Moreover, by (7.2) and (7.3) we have that uλ,pp1. Therefore, we can apply Lemma 7.3 with c=p1 and conclude that uλ,p is the only solution of (7.1) with up1 for every λ(0,ˇλp].

    The next result states that problem (7.1) has no solution for large λ; that is, there is a value ˆλp>0 such that (7.1) has no weak solution with λ>ˆλp.

    Proposition 7.4 ([20,Theorem 2.1]). Problem (7.1) does not have a solution for λ>ˆλp, where

    ˆλp=λ1(p,Ω)max{1,(p1e)p1}. (7.5)

    At this point we can define

    λmax,p=sup{λ>0 : problem (7.1) has a solution}. (7.6)

    In the next result we show that λmax,p is well-defined, find its asymptotic behavior as p, and complete the construction of the branch of minimal solutions.

    Proposition 7.5. Let ΩRn be a bounded domain and p>n. Then, λmax,p given by (7.6) is well-defined (in the sense that it is positive and finite). Moreover, (7.1) has a minimal solution uλ,p(x) for every λ(0,λmax,p) and no solution for λ>λmax,p. In addition,

    ˇλpλmax,pˆλp, (7.7)

    where ˇλp and ˆλp are respectively given by (7.2), (7.5), and

    limpλ1/pmax,pp=Λmax

    for Λmax defined by (5.1).

    Proof. By Propositions 7.2 and 7.4, we have that 0<ˇλpλmax,pˆλp<. Moreover, although we do not know λmax,p explicitly, (7.2), (7.5), and (7.7), along with Proposition 2.2 and Lemma 2.5 provide its asymptotic behavior, namely,

    limpλ1/pmax,pp=limpˇλ1/ppp=limpˆλ1/ppp=e1Λ1(Ω)=Λmax.

    Let us now complete the construction of the branch of minimal solutions. Since λmax,p< we can take μ arbitrarily close to λmax,p and uμ solution of

    {Δpuμ=μeuμin Ω,uμ=0on Ω.

    Then, for every λ(ˇλp,μ] we can produce a minimal solution as in Proposition 7.2, taking ¯u=uμ in the iteration.

    We are now ready to prove Theorem 7.1.

    Proof of Theorem 7.1. 1) We have that

    {Δpuλp,p=λpeuλp,pin Ωuλp,p=0on Ω.

    Multiplying the equation by uλp,p and integrating by parts, we get

    Ω|uλp,p(x)|pdx=λpΩuλp,p(x)euλp,p(x)dx.

    Let us fix p>n+1. Then, for every x,yΩ, there exists a positive constant C independent of p (see [10,Lemma 3.3]) such that

    |uλp,p(x)uλp,p(y)||xy|1nn+1C(Ω|uλp,p|n+1dx)1/(n+1)C|Ω|1n+11p(Ω|uλp,p|pdx)1/p=C|Ω|1n+11p(λpΩuλp,peuλp,pdx)1/pC|Ω|1n+1(λpuλp,peuλp,p)1/p. (7.8)

    Let us now find estimates for uλp,p.

    2) Consider ˇλp, given by (7.2). Since

    limpˇλ1/ppp=Λmax>Λ=limpλ1/ppp,

    there exists p0 such that λp<ˇλp for all pp0. Then, by estimate (7.3), we have

    λ1p1ppvp(x)uλp,p(x)pλ1p1ppevp(x)in ¯Ω,

    where vp is given by (2.2). Take ϵ>0 such that (1+ϵ)Λ<Λmax. By Proposition 2.2, we know that vpdist(,Ω) uniformly as p and we deduce that

    uλp,pp(1+ϵ)Λedist(,Ω)=(1+ϵ)ΛΛ1max<1 (7.9)

    for p large enough. Then, from (7.8) and the Arzelà-Ascoli theorem, we find that there exists a subsequence p and a limit function uΛ such that

    uλp,ppuΛ,uniformly as p.

    3) By Proposition 3.1, we have that uΛ is a viscosity solution of the limit problem (6.1). Additionally, from estimate (7.9) we deduce uΛΛΛ1max<1, and then Theorem 6.1 implies that uΛ must be the minimal solution of the limit problem (6.1). Therefore, the whole sequence uλp,p converges, and not only a subsequence, which concludes the proof.

    This section proves that, under certain geometric assumptions on the domain Ω, it is possible to compute an explicit curve of solutions. Moreover, we establish a further non-existence result with the aid of this curve of solutions. To this aim, we consider the ridge set of Ω,

    R={xΩ:dist(x,Ω) is not differentiable at x}={xΩ: x1,x2Ω, x1x2, s.t. |xx1|=|xx2|=dist(x,Ω)}

    and its subset M, the set of maximal distance to the boundary,

    M={xΩ:dist(x,Ω)=dist(,Ω)}.

    We have proved in Theorem 6.1 the existence of minimal solutions for the limit problem (1.5), as well as several non-existence results in Propositions 3.2 and 5.1. These results hold for general bounded domains Ω. In this section, we find a second solution to the limit problem (1.5) under the additional assumption MR. Furthermore, both solutions lie on an explicit curve of solutions (see Figure 1). Some examples of domains satisfying MR are the ball, the annulus, and the stadium (convex hull of two balls of the same radius). A square or an ellipse does not verify the condition.

    Figure 1.  Curve of explicit solutions Λ1(Ω)uΛΛeuΛ=0 in Theorem 8.1 and regions of non-existence derived from Proposition 5.1, Theorem 8.4, and the uniqueness result in Theorem 6.1.

    We have the following result.

    Theorem 8.1. Let Λ>0 and Λmax given by (5.1). Assume that ΩRn is a bounded domain that satisfies MR. Let us consider solutions of the form

    u(x)=αdist(x,Ω),α>0 (8.1)

    for the problem

    {min{|u(x)|Λeu(x),Δu(x)}=0in Ω,u=0on Ω. (8.2)

    Then, problem (8.2)

    i) Has two solutions of the form (8.1) if 0<Λ<Λmax, corresponding to the two roots of

    αΛeαdist(,Ω)=0. (8.3)

    ii) Has one solution of the form (8.1) for Λ=Λmax, with α=dist(,Ω)1.

    iii) Has no solutions for Λ>Λmax, and only the trivial solution for Λ=0.

    Remark 8.2. By Theorem 6.1, for 0<Λ<Λmax the solution of the form (8.1) with smallest α is the minimal solution of (8.2).

    Proof. First of all, we are going to check that

    Δu(x)=0in ΩR

    in the viscosity sense. Let ϕC2 and x0ΩR such that uϕ has a local maximum at x0. We can assume u(x0)=ϕ(x0) and ϕ(x0)0. A Taylor expansion, and the fact that ϕ touches u from above at x0 yield

    Δϕ(x0)|ϕ(x0)|2+o(1)1ϵ2(2u(x0)maxyBϵ(x0)u(y)minyBϵ(x0)u(y))

    as ϵ0. From (8.1) we have that

    maxyBϵ(x0)u(y)=u(x0)+αϵ,minyBϵ(x0)u(y)=u(x0)αϵ

    and we deduce that u is -subharmonic in ΩR. The proof that it is also -superharmonic is analogous. Hence, we need make sure that

    |u(x)|Λeu(x)0xΩR.

    Indeed, we find that

    |u(x)|Λeu(x)=αΛeαdist(x,Ω)

    (recall that xR and the derivatives are classical). Since we can choose points xRM arbitrarily close to M, we find the necessary condition

    αΛeαdist(,Ω)0. (8.4)

    Next, we turn our attention to the ridge set R. First, observe that cones as in (8.1) are always supersolutions of (8.2) in the ridge set, since they cannot be touched from below with C2 functions at those points. Hence, we only have to consider the subsolution case. So, let x0R and ϕC2 such that uϕ has a local maximum point at x0. We aim to prove that

    min{|ϕ(x0)|Λeu(x0),Δϕ(x0)}0. (8.5)

    It is well-known (see for instance [27,Lemma 6.10]) that

    min{|u(x)|α,Δu(x)}=0

    in the viscosity sense. Thus, by definition of viscosity subsolution we have that either |ϕ(x0)|α or Δϕ(x0)0. In the latter case, (8.5) holds and there is nothing to prove. Thus, we can assume in the sequel that Δϕ(x0)>0 and |ϕ(x0)|α. Then, since x0RM, we have u(x0)=αdist(,Ω) and

    |ϕ(x0)|Λeu(x0)αΛeαdist(,Ω).

    Recalling (8.4), we discover that the only possibility is that (8.3) holds. The rest of the proof is devoted to study the number of positive solutions of equation (8.3).

    Consider ΦΛ(α)=Λeαdist(,Ω)α. It is elementary to show that ΦΛ is convex and has a global minimum at

    αmin=dist(,Ω)1log(Λdist(,Ω)).

    This minimum value is

    minαRΦΛ(α)=ΦΛ(αmin)=dist(,Ω)1(1+log(Λdist(,Ω))).

    Whenever this minimum is strictly positive, Eq (8.3) has no solution. This happens when Λ>Λmax (in fact, Proposition 5.1 gives a stronger result in this case). Furthermore, notice that if Λ=0, then necessarily α=0. These facts amount to (iii). When the minimum equals 0, that is, when Λ=Λmax, then there exists a unique solution with α=dist(,Ω)1. This is part (ii). And finally, for part (i), notice that when the minimum is strictly negative (0<Λ<Λmax), equation (8.3) has two roots.

    Remark 8.3. Theorem 8.1 yields the following implicit curve of cone solutions

    Λ1(Ω)uΛΛeuΛ=0,

    where Λ1(Ω)=dist(,Ω)1 is the first -eigenvalue, see [30]. The same curve was deduced heuristically by Lions in the context of the Gelfand problem for the Laplacian in [36,p. 465,item (h) and Remark 2.4]. Unfortunately, Lions uses this example to caution against the heuristic reasoning since the bifurcation diagram is of corkscrew-type for dimensions 3n9. One could wonder why we do not see a similar situation in Theorem 8.1. However, according to [16,Lemma 2.3], the corresponding corkscrew-type diagram for the p-Laplacian in the radial case occurs in the range

    p<n<p(p+3)p1,

    which cannot happen as p.

    The following result shows that we can enlarge the region of nonexistence of solutions for certain domains by taking advantage of the curve of explicit solutions.

    Theorem 8.4. Let Ω be a bounded domain such that MR, and assume M is Lipschitz connected. Then, for every Λ>0, the only solutions of the problem

    {min{|uΛ(x)|ΛeuΛ(x),ΔuΛ(x)}=0in Ω,uΛ=0on Ω (8.6)

    satisfying

    Λ1(Ω)uΛΛeuΛ0, (8.7)

    are the explicit solutions found in Theorem 8.1, which satisfy (8.7) with an equality.

    The idea of the proof of Theorem 8.4 is to show that any solution uΛ satisfying (8.7) must necessarily be a cone and therefore belong to the curve of solutions given by Theorem 8.1. First, we show that solutions of (8.6) that satisfy (8.7) must lie below a cone with their same height.

    Lemma 8.5. Let Ω be a bounded domain and uΛ be a viscosity solution of (8.6) satisfying (8.7). Then,

    uΛuΛdist(,Ω)dist(,Ω)in Ω.

    Proof. It is enough to prove that

    min{|uΛ(x)|Λ1(Ω)uΛ,ΔuΛ(x)}0in Ω (8.8)

    in the viscosity sense. Then one gets uΛ(x)uΛdist(,Ω)1dist(x,Ω) in Ω by comparison (Lemma 2.3), and the result follows.

    To prove (8.8), let ϕC2 such that uΛϕ has a maximum at x0Ω. As uΛ is a viscosity solution of (8.6), it satisfies

    min{|ϕ(x0)|ΛeuΛ(x0),Δϕ(x0)}0in Ω.

    If Δϕ(x0)0 we are done, so assume Δϕ(x0)>0 and |ϕ(x0)|ΛeuΛ(x0)0. Using (8.7), we have

    |ϕ(x0)|Λ1(Ω)uΛΛeuΛ(x0)Λ1(Ω)uΛ0,

    and then

    min{|ϕ(x0)|Λ1(Ω)uΛ,Δϕ(x0)}0in Ω

    as desired.

    Remark 8.6. Lemma 8.5 holds for any bounded domain Ω without the assumption MR.

    Next, we recall the following result from [43,Theorem 2.4,(i)], which is a crucial point in the proof of Theorem 8.4.

    Lemma 8.7. Let Ω be a bounded domain such that M is Lipschitz connected. If u is -superharmonic (see [34,35]) then,

    {xΩ: u(x)=uL(Ω)}M.

    Now, we can complete the proof of Theorem 8.4.

    Proof of Theorem 8.4. Consider uΛ solution of (8.6) satisfying (8.7). Notice that

    v(x)=uΛdist(,Ω)dist(,Ω)

    is the unique (see [25]) viscosity solution of the problem

    {Δv(x)=0in ΩMv(x)=uΛon Mv(x)=0on Ω. (8.9)

    Since uΛ is -superharmonic, it is also a viscosity supersolution of (8.9) by Lemma 8.7. Then, we get vuΛ by comparison (see [25]), and Lemma 8.5 yields uΛv. That is, uΛ is of the form (8.1). Since all the solutions of (8.6) of the form (8.1) are given by Theorem 8.1, we find that there are no solutions with

    Λ1(Ω)uΛΛeuΛ>0.

    Furthermore, if Λ1(Ω)uΛΛeuΛ=0, then uΛ must be one of the explicit solutions in Theorem 8.1.

    F. C. partially supported by a Wayne State University 2021-22 University Research Grant, and grants MTM2017-84214-C2-1-P and PID2019-110712GB-I100 funded by MCIN/AEI/10.13039/ 501100011033 and by "ERDF A way of making Europe".

    The authors declare no conflict of interest.



    [1] B. Abdellaoui, I. Peral, Existence and nonexistence results for quasilinear elliptic equations involving the p-Laplacian with a critical potential, Ann. Mat. Pura Appl. IV. Ser., 182 (2003), 247–270. https://doi.org/10.1007/s10231-002-0064-y doi: 10.1007/s10231-002-0064-y
    [2] A. Anane, J. L. Lions, Simplicitè et isolation de la premiere valeur propre du p-laplacien avec poids, C. R. Acad. Sci. Paris Sér. I Math., 305 (1987), 725–728.
    [3] T. Bhattacharya, E. DiBenedetto, J. Manfredi, Limit as p of Δpup=f and related extremal problems, Rend. Sem. Mat. Univ. Politec. Torino, 1989, 15–68.
    [4] M. Bocea, M. Mihǎilescu, Existence of nonnegative viscosity solutions for a class of problems involving the -Laplacian, Nonlinear Differ. Equ. Appl., 23 (2016), 11. https://doi.org/10.1007/s00030-016-0373-2 doi: 10.1007/s00030-016-0373-2
    [5] G. Bratu, Sur les équations intégrales non linéaires, Bulletin de la S. M. F., 42 (1914), 113–142.
    [6] H. Brezis, L. Oswald, Remarks on sublinear elliptic equations, Nonlinear Anal. Theor., 10 (1986), 55–64. https://doi.org/10.1016/0362-546X(86)90011-8 doi: 10.1016/0362-546X(86)90011-8
    [7] X. Cabré, M. Sanchón, Semi-stable and extremal solutions of reaction equations involving the p-Laplacian, Commun. Pure Appl. Anal., 6 (2007), 43–67. https://doi.org/10.3934/cpaa.2007.6.43 doi: 10.3934/cpaa.2007.6.43
    [8] S. Chandrasekhar, An introduction to the study of stellar structures, New York: Dover, 1957.
    [9] S. Chanillo, M. Kiessling, Surfaces with prescribed Gauss curvature, Duke Math. J., 105 (2000), 309–353. https://doi.org/10.1215/S0012-7094-00-10525-X doi: 10.1215/S0012-7094-00-10525-X
    [10] F. Charro, E. Parini, Limits as p of p-laplacian problems with a superdiffusive power-type nonlinearity: positive and sign-changing solutions, J. Math. Anal. Appl., 372 (2010), 629–644. https://doi.org/10.1016/j.jmaa.2010.07.005 doi: 10.1016/j.jmaa.2010.07.005
    [11] F. Charro, E. Parini, Limits as p of p-laplacian eigenvalue problems perturbed with a concave or convex term, Calc. Var., 46 (2013), 403–425. https://doi.org/10.1007/s00526-011-0487-7 doi: 10.1007/s00526-011-0487-7
    [12] F. Charro, I. Peral, Limit branch of solutions as p for a family of sub-diffusive problems related to the p-laplacian, Commun. Part. Diff. Eq., 32 (2007), 1965–1981. https://doi.org/10.1080/03605300701454792 doi: 10.1080/03605300701454792
    [13] F. Charro, I. Peral, Limits as p of p-Laplacian concave-convex problems, Nonlinear Anal. Theor., 75 (2012), 2637–2659. https://doi.org/10.1016/j.na.2011.11.009 doi: 10.1016/j.na.2011.11.009
    [14] M. G. Crandall, H. Ishii, P. L. Lions, User's guide to viscosity solutions of second order partial differential equations, Bull. Amer. Math. Soc., 27 (1992), 1–67. https://doi.org/10.1090/S0273-0979-1992-00266-5 doi: 10.1090/S0273-0979-1992-00266-5
    [15] J. Dávila, Singular solutions of semi-linear elliptic problems, In: Handbook of differential equations: stationary partial differential equations, Amsterdam: Elsevier/North-Holland, 2008, 83–176. https://doi.org/10.1016/S1874-5733(08)80019-8
    [16] M. Del Pino, J. Dolbeault, M. Musso, Multiple bubbling for the exponential nonlinearity in the slightly supercritical case, Commun. Pure Appl. Anal., 5 (2006), 463–482. https://doi.org/10.3934/cpaa.2006.5.463 doi: 10.3934/cpaa.2006.5.463
    [17] R. Emden, Gaskugeln: Anwendungen der mechanischen Wärmetheorie auf kosmologische und meteorologische Probleme (German), Germany: B. G. Teubner, 1907.
    [18] N. Fukagai, M. Ito, K. Narukawa, Limit as p of p-Laplace eigenvalue problems and L-inequality of the Poincare type, Differ. Integral Equ., 12 (1999), 183–206.
    [19] J. P. García Azorero, I. Peral, Existence and nonuniqueness for the p-laplacian: nonlinear eigenvalues, Commun. Part. Diff. Eq., 12 (1987), 1389–1430.
    [20] J. García Azorero, I. Peral Alonso, On a Emden-Fowler type equation, Nonlinear Anal. Theor., 18 (1992), 1085–1097. https://doi.org/10.1016/0362-546X(92)90197-M doi: 10.1016/0362-546X(92)90197-M
    [21] J. García Azorero, I. Peral Alonso, J. P. Puel, Quasilinear problems with exponential growth in the reaction term, Nonlinear Anal. Theor., 22 (1994), 481–498. https://doi.org/10.1016/0362-546X(94)90169-4 doi: 10.1016/0362-546X(94)90169-4
    [22] I. M. Gel'fand, Some problems in the theory of quasilinear equations, Amer. Math. Soc. Trans., 29 (1963), 295–381.
    [23] J.-F. Grosjean, p-Laplace operator and diameter of manifolds, Ann. Glob. Anal. Geom., 28 (2005), 257–270. https://doi.org/10.1007/s10455-005-6637-4 doi: 10.1007/s10455-005-6637-4
    [24] J. Jacobsen, K. Schmitt, The Liouville-Bratu-Gelfand problem for radial operators, J. Differ. Equations, 184 (2002), 283–298. https://doi.org/10.1006/jdeq.2001.4151 doi: 10.1006/jdeq.2001.4151
    [25] R. Jensen, Uniqueness of Lipschitz extensions: Minimizing the sup norm of the gradient, Arch. Rational Mech. Anal., 123 (1993), 51–74. https://doi.org/10.1007/BF00386368 doi: 10.1007/BF00386368
    [26] D. D. Joseph, T. S. Lundgren, Quasilinear Dirichlet problems driven by positive sources, Arch. Rational Mech. Anal., 49 (1973), 241–269. https://doi.org/10.1007/BF00250508 doi: 10.1007/BF00250508
    [27] P. Juutinen, Minimization problems for Lipschitz functions via viscosity solutions, Helsinki: Suomalainen Tiedeakatemia, 1998.
    [28] P. Juutinen, Principal eigenvalue of a badly degenerate operator, J. Differ. Equations, 236 (2007), 532–550. https://doi.org/10.1016/j.jde.2007.01.020 doi: 10.1016/j.jde.2007.01.020
    [29] P. Juutinen, P. Lindqvist, On the higher eigenvalues for the -eigenvalue problem, Calc. Var., 23 (2005), 169–192. https://doi.org/10.1007/s00526-004-0295-4 doi: 10.1007/s00526-004-0295-4
    [30] P. Juutinen, P. Lindqvist, J. Manfredi, The -eigenvalue problem, Arch. Rational Mech. Anal., 148 (1999), 89–105. https://doi.org/10.1007/s002050050157 doi: 10.1007/s002050050157
    [31] B. Kawohl, On a family of torsional creep problems, J. Reine Angew. Math., 410 (1990), 1–22.
    [32] P. Lindqvist, On the equation div(|u|p2u)+λ|u|p2u=0, Proc. Amer. Math. Soc., 109 (1990), 157–164. https://doi.org/10.1090/S0002-9939-1990-1007505-7 doi: 10.1090/S0002-9939-1990-1007505-7
    [33] P. Lindqvist, Addendum: "On the equation div(|u|p2u)+λ|u|p2u=0'' [Proc. Amer. Math. Soc. 109 (1990), no. 1,157–164; MR1007505 (90h: 35088)], Proc. Amer. Math. Soc., 116 (1992), 583–584. https://doi.org/10.1090/S0002-9939-1992-1139483-6
    [34] P. Lindqvist, J. Manfredi, The Harnack inequality for -harmonic functions, Electron. J. Diff. Eqns., 5 (1995), 1–5.
    [35] P. Lindqvist, J. Manfredi, Note on -superharmonic functions, Revista Matemática de la Universidad Complutense de Madrid, 10 (1997), 471–480.
    [36] P. L. Lions, On the existence of positive solutions of semilinear elliptic equations, SIAM Rev., 24 (1982), 441–467. https://doi.org/10.1137/1024101 doi: 10.1137/1024101
    [37] J. Liouville, Sur l'équation aux différences partielles d2logλdudv±λ2a2=0, Journal de mathématiques pures et appliquées 1re série, 18 (1853), 71–72.
    [38] M. Mihǎilescu, D. Stancu-Dumitru, C. Varga, The convergence of nonnegative solutions for the family of problems Δpu=λeu as p, ESAIM: COCV, 24 (2018), 569–578. https://doi.org/10.1051/cocv/2017048 doi: 10.1051/cocv/2017048
    [39] I. Peral, Some results on quasilinear elliptic equations: growth versus shape, In: Proceedings of the second school of nonlinear functional analysis and applications to differential equations, Trieste: World Scientific, 1998,153–202.
    [40] O. W. Richardson, The emission of electricity from hot Bodies, India: Longmans, Green and Company, 1921.
    [41] M. Sanchón, Regularity of the extremal solution of some nonlinear elliptic problems involving the p-Laplacian, Potential Anal., 27 (2007), 217–224. https://doi.org/10.1007/s11118-007-9053-5 doi: 10.1007/s11118-007-9053-5
    [42] G. W. Walker, Some problems illustrating the forms of nebulae, Proc. R. Soc. Lond. A, 91 (1915), 410–420. https://doi.org/10.1098/rspa.1915.0032 doi: 10.1098/rspa.1915.0032
    [43] Y. Yu, Some properties of the ground states of the infinity Laplacian, Indiana Univ. Math. J., 56 (2007), 947–964.
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2153) PDF downloads(227) Cited by(0)

Figures and Tables

Figures(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog