Skip to content
BY 4.0 license Open Access Published by De Gruyter September 20, 2019

A-priori bounds for quasilinear problems in critical dimension

  • Giulio Romani EMAIL logo

Abstract

We establish uniform a-priori bounds for solutions of the quasilinear problems

ΔNu=f(u)in Ω,u=0on Ω,

where Ω ⊂ ℝN is a bounded smooth convex domain and f is positive, superlinear and subcritical in the sense of the Trudinger-Moser inequality. The typical growth of f is thus exponential. Finally, a generalisation of the result for nonhomogeneous nonlinearities is given. Using a blow-up approach, this paper completes the results in [1, 2], extending the class of nonlinearities for which the uniform a-priori bound applies.

MSC 2010: 35J92; 35B45; 35B44

1 Introduction and main results

The study of a-priori bounds for solutions of elliptic boundary value problems, that is, establishing the existence of a positive constant C such that ∥uL(Ω)C for all solutions u, is an interesting and important issue. Indeed, on the one hand it is a key point to show existence of solutions by means of the degree theory. Moreover, a large number of different techniques have been developed to get such results for problems of the kind

Δmu=f(x,u)in Ω,u=0on Ω, (1)

where Ω ⊂ ℝN is typically a smooth bounded domain, Δmu := div(|∇u|m–2u) and f : Ω × ℝ → ℝ+ has a subcritical growth in the second variable. Here subcriticality is meant in the sense of the Sobolev embeddings.

When m = 2, namely for second-order semilinear problems, uniform a-priori bounds were firstly obtained for a (non-optimal) class of subcritical nonlinearities by Brezis and Turner in [3] by means of Hardy-Sobolev inequalities. Some years later, Gidas and Spruck in [4] and de Figueiredo, Lions and Nussbaum in [5] improved this result, applying respectively a blow-up method and a moving-planes technique together with a Pohožaev identity. The main assumption therein was that the growth at ∞ of f is controlled by a suitable power with exponent 2* – 1 = N+2N2 , 2* being the critical Sobolev exponent. We also point out the more recent work of Castro and Pardo [6] which enlarges the class of nonlinearities involved.

Regarding quasilinear equations, similar results have been achieved for 1 < m < N by Azizieh and Clément [7], Ruiz [8], Dall’Aglio et al. [9], Lorca and Ubilla [10] and Zou [11]. In these works, the nonlinearity may depend also on x and on the gradient, nonetheless its growth at infinity with respect to u should be less than a subcritical power. Originally, most of these results concerned the case 1 < m ≤ 2 due to some symmetry and monotonicity arguments for solutions to m-laplacian equations which were at that time available only for that range of m; however, those techniques have been later extended also to the case m > 2 in the papers [12, 13]. See also the recent work of Damascelli and Pardo [2], which further extends the class of nonlinearities for which the a-priori bound applies, in the spirit of [5, 6].

When, on the other hand, we consider the limiting case m = N, the Sobolev space W01,N (Ω) compactly embeds in all Lebesgue spaces and the well-known Trudinger-Moser inequality shows that the maximal growth for a function g such that ∫Ωg(u)dx < + ∞ for any u W01,N (Ω) is g(t)=exp{αNtNN1}, the constant αN > 0 being explicit. Therefore, we can consider problems of kind (1) with suitable exponential nonlinearities. Nevertheless, a-priori bounds may be found only up to the threshold tet, as the work [14] of Brezis and Merle shows for m = N = 2 (see also Section 5). In addition, the authors proved local a-priori bounds when the nonlinearity is of kind f(x, t) = V(x)et. This, together with the boundary analysis [5], yields the desired a-priori bound for convex domains. The results in [14] have been later on extended in the quasilinear setting by Aguilar Crespo and Peral Alonso in [15] and we refer also to [16] for concentration compactness issues in this direction.

The boundary value problem (1) with m = N and general subcritical nonlinearities (in the sense of the Trudinger-Moser inequality) has been studied by Lorca, Ruf and Ubilla in [1], where the authors considered superlinear growths either controlled by the map tetα for some α ∈ (0, 1), or which are comparable to tet. To prove a uniform a-priori bound in the first alternative, techniques involving Orlicz spaces are applied, while for the second the authors adapt the strategy in [14].

The goal of our work is to fill the gap between these growths. Indeed, Passalacqua improved the results of [1] in his Ph.D. Thesis [17] by means of a subtle modification of the argument therein, but the gap was still not completely filled and this seems to be out of reach with those techniques. Here, instead, we apply a blow-up method inspired by [18] to deduce the desired estimates in the interior of the domain. Moreover, our results complete the analysis in [2] for the case m = N and, further, to the best of our knowledge, they seem to be new even for the semilinear problem when N = 2.

We note that a similar approach has been also applied by Mancini and the author in [19] to study higher-order problems.

Throughout all the paper, Ω is a bounded and strictly convex domain in ℝN with C1,α boundary and the nonlinearity f satisfies the following conditions:

  1. f ≥ 0, fC1(ℝ) and f′ ≥ 0 on [M, +∞) for some M ≥ 0;

  2. there exists a positive constant d such that lim inft+f(t)tN1+d=+;

  3. there exists limt+f(t)f(t):=β[0,+).

The second assumption is standard (cf. [1, 2]) and in the linear case N = 2 is equivalent to say that f is slightly more than superlinear. Assumption (A3) is a control on the growth at ∞ of f. Indeed, Gronwall Lemma implies that for any ε > 0 there are constants Cε, Dε > 0 such that

min{0,Dεe(βε)tCε}f(t)Dεe(β+ε)t+Cε. (2)

We refer to [19, Lemma 1.1] for a detailed proof.

In the blow-up analysis, it will be useful to distinguish between the following cases.

Definition 1.1

Let f satisfy (A1)-(A3). We say that f is subcritical if

limt+f(t)f(t)=0.

On the other hand, we say that f is critical if

limt+f(t)f(t)(0,+).

Remark 1

In view of (2), an example of a subcritical nonlinearity f(t) = logτ(1 + t)tp etα with τ ≥ 0, p ≥ 1 and α ∈ [0, 1), and for a critical nonlinearity is f(t)=eγt(t+1)q with γ > 0. Notice that the second nonlinearity was excluded by the previous results [1, 17] e.g. when γ = 1, q ≥ 1.

Although this distinction will be relevant in the proof, the critical case being more delicate, our main result applies for both classes of nonlinearities.

Theorem 1.1

Let Ω ⊂ ℝN, N ≥ 2, be a bounded strictly convex domain of class C1,α for some α ∈ (0, 1) and let f satisfy (A1)-(A3). Then, there exists a constant C > 0 such that all weak solutions u W01,N (Ω) of the problem

ΔNu=f(u)inΩ,u=0onΩ, (3)

satisfyuL(Ω)C.

We recall that u W01,N (Ω) is a weak solution of problem (3) if

Ω|u|N2uφ=Ωf(u)φ,for allφC0(Ω¯).

Notice that, unlike [1, 2], we are not prescribing that our solutions are in L(Ω). Indeed, by [17, Proposition 6.2.2], every W01,N (Ω) weak solution of problem (3) is classical, namely it belongs to C1,γ(Ω) for some γ ∈ (0, 1) (see also [9, Proposition 3.1]).

Remark 2

By the regularity theory of quasilinear problems by [20, 21] the C1,α assumptions on ∂Ω are sufficient to guarantee that the uniform a-priori estimate in Theorem 1.1 is actually in C1,α̃(Ω) for some α̃ ∈ (0, 1).

Remark 3

We may set our problem in the framework of entropy solutions and obtain the same result. Indeed, one may prove that, under conditions (A1)-(A3), such solutions are indeed weak and in turn classical. For further details, we refer to Section 5, in particular to Remark 6.

Once the uniform a-priori bound of Theorem 1.1 is established, the existence of a weak solution for problem (3) may be proved by means of the topological degree theory, provided a control of the behaviour of f in 0 is imposed.

Proposition 1.2

In addition of the assumptions of Theorem 1.1, suppose that

lim supt0+f(t)tN1<λ1,

where λ1 denotes the first eigenvalue ofΔN. Then, problem (3) admits a positive weak solution.

We will not give the proof of Theorem 1.2, being rather standard once Theorem 1.1 is established, addressing the interested reader to [2, Theorem 1.5].

This paper is organized as follows: in Section 2 we collect some auxiliary results as well as boundary and energy estimates obtained in [1]; Section 3 contains the proof of Theorem 1.1 distinguishing between the subcritical and the critical case. In Section 4 we give a generalisation of our result for a class of nonhomogeneous problems and finally in Section 5 we provide an example of nonlinearities which, although being subcritical in the sense of Trudinger-Moser, do not satisfy assumption (A3) and for which one can find unbounded solutions.

2 Preliminary estimates

In this section, we state some known results which will be used in the sequel. We begin with a result by Serrin, improved by the regularity theory from [20]. Next, we state a Harnack-type inequality.

Lemma 2.1

([22], Theorem 2). Let u be a weak solution ofΔNu = h in B2R(0) ⊂ Ω. Then,

uC1,α(BR(0))C(uLN(B2R(0))+KR), (4)

for a suitable α ∈ (0, 1) and positive constants C(α, R, hLNNε(Ω) ) and K(R, hLNNε(Ω) ), for some ε ∈ (0, 1].

Lemma 2.2

([22], Theorem 6). Let u ≥ 0 be a solution ofΔNu = h in B3R(0) ⊂ Ω. Then

maxBR(0)(u)C(minBR(0)(u)+K),

where the constants C, K depending on R and on hLNNε(Ω) , ε ∈ (0, 1].

The next result is taken from [1, Propositions 2.1 - 3.1] and concerns the behaviour near the boundary of solutions of (3) and their energy.

Proposition 2.3

There exist positive constants r, C0 such that every solution u W01,N (Ω) of (3) verifies

u(x)C0and|u(x)|C0,xΩr, (5)

where Ωr := {xΩ | d(x, ∂Ω) ≤ r}, and

Ωf(u)Λ. (6)

Roughly speaking, the proof of the first inequality in (5) is obtained similarly to the analogous estimate in the second-order case (see [5]) by means of a Picone inequality originally proved in [23] (see also [2, Theorem 2.5]). This implies the second inequality in (5) by standard regularity estimates. Finally, (6) follows from (5) testing the equation with a suitable function with vanishing gradient outside Ωr.

We point out that the uniform estimates (5)-(6) do not rely on assumption (A3), so they can be deduced for any superlinear growth (in the sense of assumption (A2)).

3 Uniform bounds inside the domain

The argument is based on a blow-up method inspired by [18] and developed in [19]: supposing the existence of an unbounded sequence of solutions, we define a rescaled sequence which locally converges to a solution of a Liouville’s equation on ℝN. Then, we have to distinguish two cases according to the growth of f. If f is subcritical, the right-hand side of the limit equation is constant and this readily yields the contradiction. In the critical framework, the Liouville’s equation is still nonlinear, so a more accurate analysis is needed: we will see that the energy concentrates around the blow-up points with a threshold which is too large for the energy bound of Proposition 2.3 to hold.

Let (uk)k be a sequence of solutions of (3) and suppose there exist points (xk)kΩ such that

uk(xk)=ukL(Ω)=:Mk+. (7)

Since Ω is bounded, then, up to a subsequence, xkxΩ, where d(x, ∂Ω) > 0 by Proposition 2.3. Moreover, we define vk : Ωk → ℝ as

vk(x):=uk(xk+μkx)Mk, (8)

where Ωk:=Ωxkμk and

μk:=1(f(Mk))1/N. (9)

Notice that μk → 0 by (A1)-(A2) and this implies Ωk ↗ ℝN, since xΩ. Moreover, each vk satisfies (in a weak sense) the equation

ΔNvk(x)=f(uk(xk+μkx))f(Mk)[0,1],xΩk. (10)

Indeed, for any test function φ C0 (Ω) there holds

Ωk|vk|N2vkφ=ΩkμkN1|uk|N2(xk+μkx)uk(xk+μkx)φ(x)dx=μkN1Ω|uk|N2(y)uk(y)(φ)(yxkμk)μkdyμkN=Ωf(uk(y))φ(yxkμk)dy=Ωkf(uk(xk+μkx))φ(x)μkNdx=Ωkf(uk(xk+μkx))f(Mk)φ(x)dx.

Moreover, fixing R > 0 and looking at the behaviour of the sequence (vk)k in BR(0), applying the Harnack inequality given by Lemma 2.2 to wk := –vk ≥ 0, we find

vkL(BR(0))=maxBR(0)|vk|=maxBR(0)wkC(R)(minBR(0)wk+K(R))=K~(R). (11)

Therefore, by the local estimate provided by Lemma 2.1, we infer that, up to a subsequence, vkv in Cloc1,α (ℝN). We claim that v is a weak solutions of the Liouville’s equation

ΔNv=eβvinRN, (12)

where β is defined in (A3). Indeed, the equation (10) can be rewritten as

ΔNvk(x)=exp{logf(uk(xk+μkx))logf(Mk)}

and thus, by a first-order Taylor expansion of log ∘ f around Mk, one finds

ΔNvk(x)=ef(zk(x))f(zk(x))(uk(xk+μkx)Mk)=ef(zk(x))f(zk(x))vk(x), (13)

where zk(x) := Mk + θ(x)(uk(xk + μkx) – Mk) = Mk + θ(x) vk(x), θ(x) ∈ (0, 1). Since vkv uniformly on compact sets and Mk → +∞, then zk(x) → +∞ uniformly on compact sets, so (12) follows by taking the limits as k → +∞ in (13).

Consequently, we split our analysis according to whether f is subcritical or critical in the sense of Definition 1.1.

3.1 The subcritical case (β = 0)

In this case the equation (12) satisfied by the limit function v reduces to

ΔNv=1inRN. (14)

Recalling the energy estimate (6) of Proposition 2.3, it is sufficient to integrate (14) on ℝN to get the desired contradiction, which proves Theorem 1.1 for such nonlinearities:

+=RNdx=RNlimk+ef(zk(x))f(zk(x))vk(x)χΩk(x)dx=RNlimk+elog(f(uk(xk+μkx)))log(f(Mk))χΩk(x)dxlim infk+Ωkf(uk(xk+μkx))f(Mk)dx=[y=xk+μkx]lim infk+Ωf(uk(y))dyΛ. (15)

3.2 The critical case (β > 0)

With the same steps as in (15), one actually gets

RNeβvdxΛ. (16)

Therefore, we can characterize the limit profile v by standard computations from a Liouville-type result by Esposito [24, Theorem 1.1] as

v(x)=Nβlog1+βN1CN|x|N1N1, (17)

where CN:=NN2N1N1. Furthermore, we claim that the energy concentrates around the blow-up points sequence (xk)k, namely there exists a constant θ > 0 such that

limR+lim infk+BRμk(xk)f(uk(y))dyθ>0, (18)

where θ is characterized by

θ=NωNβN1N2N1N1, (19)

with ωN denoting the volume of the unit ball in ℝN. Indeed,

0<θ:=RNeβv=limR+BR(0)limk+e(f(zk(x))f(zk(x)))vk(x)dxlimR+lim infk+BR(0)f(uk(xk+μkx)f(Mk)dx=limR+lim infk+BRμk(xk)f(uk(y))dy.

We can quantify the constant θ as follows. Let us define the auxiliary function w := βv + (N – 1) log β. Then, it is easy to show that w satisfies –ΔNw = ew and ∫N ew < ∞. Therefore, by [24, Theorem 1.1], we have

θ=RNeβv=1βN1RNew=NωNβN1N2N1N1.

In other words, if we apply the blow-up technique around a maximum point, we see that we can characterize the limit profile and we get the concentration of the energy as in (18)-(19). This is indeed what happens around any blow-up point of the sequence (uk)k, meaning points belonging to the set

S:={yΩ¯|(yj)jΩ|yjy,ukj(yj)+ as j+}.

Lemma 3.1

From the sequence (uk)k one can extract a subsequence still denoted by (uk)k for which the following holds.

There are P ∈ ℕ and sequences (xk,i)k for 1 ≤ iP such that limk→+∞ uk(xk,i) = +∞ for any i and, setting

vk,i(x):=uk(xk,i+μk,ix)uk(xk,i),μk,i:=(f(uk(xk,i)))1/N,

and x(i) = limk→+∞ xk,i, we have

  1. limk+|xk,ixk,j|μk,i=+ for 1 ≤ ijP;

  2. vk,iv in Cloc1,γ (ℝN), for 1 ≤ iP, where v is defined in (17) and (18)-(19) still hold;

  3. supk inf1≤iP|xxk,i|N f(uk(x)) ≤ C for every xΩ.

Proof

The proof is inspired by [18, Claims 5-7]. We say that the property 𝓗p holds whenever there exist p sequences (xk,i)k for 1 ≤ ip such that uk(xk,i) → +∞ as k → +∞ and such that (i) and (ii) hold. From our previous investigation, we know that 𝓗1 holds. Our aim is to prove that the number of such sequences is finite by showing the following: if 𝓗p holds then either 𝓗p+1 holds too or there exists a constant C > 0 such that (iii) holds. Indeed, if so, supposing by contradiction that 𝓗p holds for any p, then we would be able to find a sequence of disjoint balls Bk,i(xk,i) such that (18) holds, and thus by Lemma 2.3,

ΛΩf(uk(x))dxi=1pBRμk,i(xk,i)f(uk(x))dx=i=1pBRμk,i(xk,i)f(uk(x))dxpθ,

an upper bound for p, a contradiction. As a consequence, 𝓗p does not hold for any p and, by the claimed alternative, the proof is completed.

Now the claim has to be proved. Suppose that 𝓗p for some p ∈ ℕ but not (iii). Hence, defining wk(x) := inf1≤ip|xxk,i|N f(uk(x)), one can find a sequence of points (yk)k such that wk(yk) = ∥wk ↗ +∞. We show that for these points (yk)k together with (xk,i)k for 1 ≤ iP, 𝓗p+1 holds. First we get uk(yk) → +∞, since

+wk(yk)(diam(Ω))Nf(uk(yk))

and by the local boundedness of f on [0, +∞). Moreover, defining

u~k(x):=uk(yk+γkx)uk(yk),γk:=(f(uk(yk)))1/N,

first we have

inf1ip|ykxk,i|γk=inf1ip|ykxk,i|f(uk(yk))1/N=wk(yk)1/N+. (20)

Then, suppose by contradiction that |ykxk,i| = O(μk,i), that is yk = xk,i + μk,iθk,i, with |θk,i| ≤ C. We have

wk(yk)|ykxk,i|Nf(uk(yk))=μk,iN|θk,i|Nf(uk(yk))=|θk,i|Nf(uk(yk))f(uk(xk,i))=|θk,i|Nef(zk,i(θk,i))f(zk,i(θk,i))vk,i(θk,i)|θ,i|Neβv(θ,i)=|θ,i|N(1+βCN11N|θ,i|NN1)N<+,

a contradiction, so (i) is proved.

Now, it remains to prove that uk has the same blow-up behaviour also for the sequence (yk)k. To this aim, we first show that, for any R > 0

f(uk(yk+γkx))f(uk(yk))C (21)

for some C = C(R) > 0 and for any xBR(0). Notice that this is not obvious as in (10), since (yk)k do not have to be necessarily maximum points for (uk)k. Rewriting in terms of the sequence (uk)k the inequality wk(yk + γkx) ≤ wk(yk), which holds for any xBR(0) for k large enough, one finds

f(uk(yk+γkx))f(uk(yk))(inf1ip|ykxk,i|inf1ip|yk+γkxxk,i|)N. (22)

Fix now ε ∈ (0, 1). By (20), there exists = (R, ε) > 0 such that for any k > we have kε|ykxk,i|. Therefore |yk + γkxxk,i| ≥ |ykxk,i| – γkR ≥ (1 – ε)|ykxk,i| and so (21) follows from (22) with C = (1 – ε)N. In order to complete the proof of the local compactness of the sequence (ũk)k and consequently get (ii), we need to have its local boundedness in the LN norm according to Lemma 2.1. This time, unlike (11), before applying the Harnack inequality of Lemma 2.2, we need to know that our sequence is uniformly bounded from above, which is not immediate. So, suppose by contradiction that for any sequence of positive constants (Ck)k ↗ +∞ there exist points (k)kBR(0) such that ũk(k) ≥ Ck holds, thus, since f is increasing,

f(uk(yk+γkx~k))f(Ck+uk(yk)).

Together with (21), this implies

f(Ck+uk(yk))Cf(uk(yk)). (23)

Choosing now Ck = euk(yk)2 and setting tk := uk(yk), (23) rereads as

f(etk2+tk)Cf(tk). (24)

Then, by superlinearity of f and by the fact that f(s) ≤ eγs + D for some γ, D > 0 for any s ∈ ℝ+ by (A3), we have

etk2+tkC¯f(etk2+tk)Cf(tk)eγtk+D,

which yields a contradiction since tk ↗ +∞. Hence, M(R) := maxBR(0) ũk ∈ [0, +∞) and we can consider the function Uk := M(R) – ũk. We have Uk ≥ 0, minBR(0) Uk = 0 and furthermore Uk satisfies

ΔNUk=f(uk(yk+γkx))f(uk(yk))[C(R),0],

where the bound from below follows from (21). Therefore, by Lemma 2.2, we get

|M(R)+minBR(0)u~k|=maxBR(0)|Uk|=maxBR(0)UkC(R)[minBR(0)Uk+K(R)]=C(R)K(R),

which in turn implies

|maxBR(0)u~k|minBR(0)u~k||C~(R). (25)

Therefore, either maxBR(0)ũk ≥ |minBR(0)ũk|, and in this case we easily infer ∥ũkL(BR(0))C(R), or we deduce from (25) that

|minBR(0)u~k|maxBR(0)u~kC~(R).

This readily implies minBR(0)ũk ≥ –C(R), so again we find ∥ũkL(BR(0))C(R).

Applying Lemma 2.1, we infer that (ũk)k is locally compact in Cloc1,α (ℝN) and, with similar steps as in (13)-(17), we finally infer (ii). Consequently, the property 𝓗p is satisfied and the proof is concluded. □

Corollary 3.2

The set of blow-up points is finite.

Proof

We show the finiteness of S by proving that S = {x(i), 1 ≤ iP}, where the points x(i) are defined as in Lemma 3.1. Indeed, let ∉ {x(i), 1 ≤ iP} for which one can find a sequence k such that, up to a subsequence, uk(k) → +∞. We have inf1≤iP, k∈ℕ|kx(i)| ≥ > 0 and, by Proposition 2.3, d(k, ∂Ω) ≥ η > 0 for some constants , η. Then, by (iii) of Lemma 3.1 and the superlinearity of f, we infer

Cinf1iP|x¯kxk,i|Nf(uk(x¯k))d¯N2(uk(x¯k)C),

a contradiction. □

Now we are in the position to prove Theorem 1.1.

Proof of Theorem 1.1

By (7) S is nonempty. Let x0S and r small such that SBr(x0) = {x0}. In this way, ∥ukL(K) < C for any KBr(x0) ∖ {x0} and thus ukũ in Cloc1,α (Br(x0) ∖ {x0}) by Lemma 2.1. Moreover, by the energy bound of Proposition 2.3, the sequence (fk)k := (f(uk))k is bounded in L1, thus fkμ in 𝓜(Br(x0)) ∩ Lloc (Br(x0) ∖ {x0}) where μ is a positive measure which is singular at x0. Hence, we can decompose μ as

μ=A(x)dx+a0δx0, (26)

where 0 ≤ AL1(Br(x0)), a0 is a positive constant and δx0 denotes the Dirac distribution centered at x0. We first claim that a0θ with θ as in (18)-(19). Indeed, for any t ∈ (0, r) we have

Bt(x0)dμ=limk+Bt(x0)f(uk(x))dxlim infk+BRμk(xk)f(uk(x))dxθ (27)

by (18), where here xk is a blow-up sequence converging to x0. As t is arbitrary, we thus find a0θ.

Let now w be the distributional solution of

ΔNw=a0δx0in Br(x0),w=0on Br(x0). (28)

Then by [25, Theorem 2.1] wC1,α(Br(x0) ∖ {x0) and has the explicit form

w(x)=a0NωN1N1logr|xx0|.

We claim that

wu~inBr(x0). (29)

If so, choosing ε(0,βN), in view of (2) on the one hand by Proposition 2.3 we get

Br(x0)e(βε)wBr(x0)e(βε)u~lim infk+Br(x0)e(βε)ukclim infk+Br(x0)f(uk)+dC(Λ).

On the other hand, by means of the explicit expressions for w and θ:

Br(x0)e(βε)w=Br(x0)r|xx0|(βε)a0NωN1N1dxBr(x0)r|xx0|(βε)θNωN1N1dxBr(x0)r|xx0|(βε)βN2N1dx>Br(x0)r|xx0|Ndx=+,

a contradiction. Therefore the proof of Theorem 1.1 is completed once we prove (29). Here, we follow the strategy in [1]. For any j ∈ ℕ, let us define the maps Bj : → ℝ+ by

Bj(s)=0for s<0,sfor 0sj,jfor s>j,

and the functions zk(j) := Bj(wuk) ≥ 0. Then, it is easy to see that zk(j) (x0) = j for any k, that zk(j)|Br(x0)=0 and, moreover, zk(j)W01,N(BR(x0)). Furthermore, as Bj is continuous, one has zk(j) z(j) pointwise, where

z(j)(s)=Bj(wu~)for xx0,jfor x=x0.

Notice that zk(j) may be extended to 0 in ΩBr(x0). Since w and uk solve respectively problems (28) and (3), then

Br(x0)|w|N2w|u~|N2u~zk(j)=a0zk(j)(x0)Ωf(uk)zk(j)=a0jBr(x0)f(uk)zk(j).

Recalling f(uk) ⇀ μ and (26), using the generalized Fatou’s Lemma with measures (see e.g. [26, Chapter 11, Proposition 17]), one infers

lim infk+Br(x0)f(uk)zk(j)dxBr(x0)z(j)dμ{x0}z(j)dμa0j,

and thus

Br(x0)|w|N2w|u~|N2u~z(j)0,

that is,

Br(x0){0wu~j}|w|N2w|u~|N2u~wu~0,

which holds for any choice of j ∈ ℕ. Then, the well-known inequality

dN|XY|N|X|N2X|Y|N2Y,XY,

for any XY ∈ ℝN, the constant dN > 0 depending only on N (see [27, Proposition 4.6]), implies

dNBr(x0)|(wu~)+|N0,

which finally proves our claim wũ ≤ 0 in Br(x0), since (wũ)+ ≤ 0 on ∂Br(x0). □

4 Generalization to nonhomogeneous problems

So far, we studied the homogeneous quasilinear problem (3), which allows a clearer exposition and a more direct comparison with the results in [1, 2]. However, a similar analysis can be carried out also in the nonhomogeneous setting, provided some conditions at infinity and near the boundary are fulfilled. More precisely, we may consider the problem

ΔNu=h(x,u)in Ω,u=0on Ω, (30)

where h : Ω × ℝ → ℝ+ ∪ {0} is a Carathéodory function satisfying the following conditions:

  1. hL(Ω × [0, τ]) for all τ ∈ ℝ+ and h(x, t) > 0 for any xΩ and t > 0;

  2. there exist fC1([0, +∞)) satisfying (A1)-(A3) and 0 < aL(Ω) ∩ C(Ω) such that

    limt+h(x,t)a(x)f(t)=1uniformly inΩ;
  3. there exist , δ̄ > 0 such that

    • h(⋅, t) ∈ C1(Ω) for all t ≥ 0 and ht (x, t) ≥ 0 in Ω × ℝ+;

    • xh(x, t) ⋅ Ψ ≤ 0 for all xΩr, t ≥ 0 and unit vectors Ψ such that |Ψn(x)| ≤ δ̄.

We recall Ωr := {xΩ | d(x, ∂Ω) ≤ r}.

Remark 4

In (H2) we assume a > 0 only in the interior of Ω, but it may vanish on the boundary. Moreover, assumption (H3) is imposed to uniformly control solutions near the boundary by a moving-planes technique, prescribing in broad terms that h should be decreasing in x along suitable outer directions and increasing in t in a suitable neighbourhood of ∂Ω.

Remark 5

From (H1)-(H2) it follows that for each ε > 0 there exists a constant dε ≥ 0 such that

(1ε)a(x)f(t)dεh(x,t)(1+ε)a(x)f(t)+dεfor allt0,xΩ. (31)

Theorem 4.1

Let Ω ⊂ ℝN, N ≥ 2, be a bounded strictly convex domain of class C1,α for some α ∈ (0, 1) and let h satisfy (H1)-(H3). Then, there exists a constant C > 0 such that

uL(Ω)C

for all weak solutions u W01,N (Ω) of problem (3).

The proof of Theorem 4.1 mainly follows the arguments of the previous sections, so we sketch here only the remarkable modifications.

Proof

First, we want to prevent blow-up near the boundary, so we adapt in our context the results of Proposition 2.3. With the same argument as in [1, Proposition 2.1] and taking (31) into account, one infers that solutions are bounded in Llocd (Ω), namely

ΩudΦ1NC, (32)

where d is defined in (A2), C is independent of u and Φ1 W01,N (Ω) is the first (positive) eigenfunction of –ΔN on Ω. In order to prove that (32) yields a uniform bound near ∂Ω, we have to show that here our solutions are decreasing with respect to outer directions. We apply a moving-planes technique in the spirit of [28, Lemma 3.2]. Let us first fix some notation. For any direction ν, set

Tλν:={xRN|xν=λ}andΩλν:={xΩ|xν<λ},

the latter being nonempty for λ > a(ν) := infxΩ xν. Moreover, for any xΩ, we denote by xλν the symmetric point with respect to the hyperplane Tλν , namely

xλν=Rλν(x):=x+2(λxν)ν.

Similarly, we define uλν:=Rλν(u), which is well-defined on (Ωλν):=Rλν(Ωλν). Our aim is to compare u and uλν near the boundary in the case ν = n(x) (the outer normal) and λa(ν) small, in order to conclude that u is decreasing along these directions in a small neighborhood of ∂Ω. First notice that by convexity of Ω, as long as λa(ν) is small, we have (Ωλν) Ω and Ωλν(Ωλν)Ωr¯. Therefore, for such λ, in Ωλν one has

ΔNuλν(x)(ΔNu(x))=h(xλν,uλν)h(x,u)(H3)h(x,uλν)h(x,u)=M(x)(uλνu),

where

M(x)=ht(x,η(x,λ))(H3)0

for some real number η(x, λ) lying between uλν (x) and u(x) by the mean value theorem, hence ML( Ωλν ). Therefore, uλν et u satisfy

ΔNuλνM(x)uλνΔNuM(x)uin Ωλν,uλνuon Ωλν,

and the weak comparison principle in small domains [12, Theorem 1.3] yields uλν u in Ωλν . We point out that the mentioned result is originally stated for M positive constant and for homogeneous nonlinearities, but it can be easily adapted to our setting (here assumption (H1) is required). By arbitrariness of λ and xΩ, one deduces that u is decreasing along the outer normal direction and then, by a compactness argument, that there exist δ ∈ (0, δ̄] and r ∈ (0, ] independent of u such that ∇u(x) ⋅ Ψ ≤ 0 for every xΩr and for every direction Ψ such that |Ψn(x)| < δ. The boundedness of u and ∇u in a neighbourhood of the boundary now follows by a standard argument by [5], and moreover one infers

Ωh(x,u)dxΛ,

with Λ independent of u. We refer to [1, Propositions 2.1-3.1] for the details.

Let us now address to the blow-up analysis. We shall see that the argument carried out in Section 3 applies to the problem (30) with only minor modifications. In particular, the rescaled functions vk defined in (8)-(9) turn out to be weak solutions of

ΔNvk(x)=h(xk+μkx,uk(xk+μkx))f(Mk),xΩk:=Ωxkμk,

and one shows that vkv in Cloc1,α (ℝN) which satisfies

ΔNv=a(x)eβvinRN.

We recall that d(x, ∂Ω) > 0, as we have already excluded boundary blow-up, thus a(x) > 0 by (H2).

In the subcritical case, the same argument as in Section 3.1 holds and, in the critical framework, Lemma 3.1 easily adapts, showing that near blow-up points the energy concentrates:

limR+lim infk+BRμk(xk)h(y,uk(y))dyθ>0,

with the same constant θ defined in (19). Then, the conclusion of the proof is analogous to the one for the homogeneous case.□

5 A counterexample

As briefly mentioned in the Introduction, our assumption (A3) can be interpreted as a control from above for the growth at ∞ of the nonlinearity f by a suitable power of et. In the spirit of Brezis-Merle [14, Example 2], we show an example of nonlinearities which are still subcritical in the sense of the Trudinger-Moser inequality without satisfying (A3) and for which one may find a positive potential a(x) such that the problem (30) admits unbounded solutions, in a suitable distributional sense.

To this aim, let f(t) = etα with α(1,NN1). Notice that f clearly satisfies (A1) and (A2), it is subcritical in the sense of Trudinger-Moser, but the condition (A3) is not fulfilled. Moreover, fix parameters δ:=(α1)N+1αN>0 and γ:=α1α(0,1) and consider the family of functions

ϕβ(t)=t+βtγδlogt,

where β > 0 is a positive parameter that will be chosen later. Notice that ϕβ(t) > tδ log t and ϕβ(t)>1dt. Hence, defining

uβ(r):=ϕβ(l(r))1α,r=|x|,

where l(r):=log(1r), there exists ρ0 > 0 such that for any r ∈ (0, ρ0) there holds uβ(r) > 0 and uβ (r) < 0. In the sequel, we adapt the strategy in [19, §6].

First, using an implicit function argument as in [19, Lemma 6.2], one can show (up to a smaller value of ρ0) that for any ρ ∈ (0, ρ0) there exists β=β(ρ)l(ρ)1α as ρ → 0 such that uβ(ρ)(ρ)β(ρ)α=0 on ∂Bρ(0). With this choice, one has uβ(ρ)(r)l(r)1α as r → 0. Then, for r ∈ (0, ρ) we compute

ΔNuβ(ρ)(r)=ddrrN1|uβ(ρ)(r)|N2uβ(ρ)(r)=N1αNrN(11α)uβ(ρ)N1αN(r)ϕβ(ρ)(l(r))Nuβ(ρ)(1α)(N1)(r)ϕβ(ρ)(l(r))N2ϕβ(ρ)(l(r)). (33)

Using the relations 12ϕβ(ρ)1 and ϕβ(ρ)δl(r)2 for r sufficiently small, one infers from (33) that there exists ρ1 < ρ0 such that for any ρ ∈ (0, ρ1) and r ∈ (0, ρ) one has –ΔNuβ(ρ)(r) > 0.

Let us now fix ρ ∈ (0, ρ1) and define

w(r):=N1α(uβ(ρ)(r)β(ρ)α). (34)

We have that w > 0 solves pointwise

ΔNw=a(x)ewαin Bρ(0){0},w=0on Bρ(0),

with a(x) := –ewα ΔNw > 0. Now we claim that aL(Bρ(0)) but wL(Bρ(0)). Indeed,

wα(r)=Nuβ(ρ)α(r)(1β(ρ)αuβ(ρ)(r))α=Nuβ(ρ)α(r)Nβ(ρ)uβ(ρ)α1(r)+o(1)=Nl(r)+Nβ(ρ)l(r)γNδlogl(r)Nβ(ρ)l(r)α1α+o(1)

as r → 0. Recalling now that γ = α1α , we get

wα(r)=Nl(r)Nδlogl(r)+o(1)

as r → 0. Therefore,

a(x)=N1αΔNuβ(ρ)ewα=N1α(N1)(α1)αN+1rNl(r)N1αNαrNl(r)Nδ(1+o(1))=N1α(N1)(α1)αN+1+o(1).

as r → 0, by our choice of δ. Hence aL(Bρ(0)). However, w is not bounded near 0 as it inherits the same behaviour of uβ(ρ).

Now we want to prove that w is an entropy solution of the problem (30). This kind of solutions has been introduced in the context of quasilinear elliptic problems in [29] in order to weaken the notion of weak solution for problems with L1-data. Here we recall the precise meaning.

Definition 5.1

We say that u T01,N (Ω) if u is measurable and Tku W01,N (Ω) for any k > 0, where Tku is defined as the truncation of u at level k, namely

Tk(s)=kfor s<k,sfor ksk,kfor s>k.

Definition 5.2

A function u T01,N (Ω) is called an entropy solution of problem (30) if for any test function φ C0 (Ω) there holds

{|uφ|<k}|u|N2u(uφ)dxΩTk(uφ)h(x,u(x))dx. (35)

We start proving Tkw W01,N (Bρ(0)) for any k > 0 and w defined in (34). Indeed, Tkw is radial and

Tkw(r)=w(r)for r[rk,ρ],w(rk)for r[0,rk),

for some suitable rk ∈ (0, ρ), so it is easy to see that

rkρ|dwdr|rN1drrkρw(r)N(1α)(Nγ(1logr))Nrdr<Ck.

In order to show that u satisfies (35), let φ C0 (Bρ(0)) and notice that Tk(wφ) ∈ W01,N (Bρ(0)). By our choice of a, the problem (30) is pointwise satisfied, thus we may test it in Bρ(0) by Tk(wφ):

Bρ(0)ΔNwTk(wφ)dx=Bρ(0)a(x)ewαTk(wφ)dx.

Note that the right-hand side is well-defined as we proved that aL(Bρ(0)) and ewαL1(Bρ(0)). Integrating by parts the left-hand side and recalling Tk(wφ)|∂Bρ(0) = 0, we have

Bρ(0)|w|N2wTk(wφ)dx=Bρ(0)a(x)ewαTk(wφ)dx

which is the case of the equality in (35).

We have therefore showed that, although w is an entropy solution of the problem

ΔNw=a(x)ewαin Bρ(0),w=0on Bρ(0),

however, wL(Bρ(0)), as w(r) ∼ l(r)1α as r → 0.

Remark 6

This counterexample shows that the class of the nonlinearities considered by assumption (A3) is sharp in order to have the property of (uniform) boundedness within the class of entropy solutions. Indeed, coming back to problem (30) under assumption (A3) and looking for entropy solutions in the sense of Definition 5.2, since this time (A3) implies a control of the nonlinearity by a suitable power of et, say p, one has

Ω|h(x,u(x))|αdxCΩeαpu<+,

where the last inequality is due to [15, Corollary 1.7]. Therefore entropy solutions are classical, and our analysis and thus Theorem 1.1 apply.

References

[1] Lorca S., Ruf B., Ubilla P. A priori bounds for superlinear problems involving the N-Laplacian. J. Differential Equations 246 (2009), no. 5, 2039-2054.10.1016/j.jde.2008.10.002Search in Google Scholar

[2] Damascelli L., Pardo R. A priori estimates for some elliptic equations involving the p-Laplacian. Nonlinear Anal. Real World Appl. 41 (2018), 475-496.10.1016/j.nonrwa.2017.11.003Search in Google Scholar

[3] Brezis H., Turner R.E.L. On a class of superlinear elliptic problems. Comm. Partial Differential Equations 2 (1977), no. 6, 601-614.10.1080/03605307708820041Search in Google Scholar

[4] Gidas B., Spruck J. A priori bounds for positive solutions of nonlinear elliptic equations, Comm. Partial Differential Equations 6 (1981), no. 8, 883-901.10.1080/03605308108820196Search in Google Scholar

[5] de Figueiredo D.G., Lions P.-L., Nussbaum R.D. A priori estimates and existence of positive solutions of semilinear elliptic equations. J. Math. Pures Appl. (9) 61 (1982), no. 1, 41-63.Search in Google Scholar

[6] Castro A., Pardo R. A priori bounds for positive solutions of subcritical elliptic equations. Rev. Mat. Complut. 28 (2015), no. 3, 715-731.10.1007/s13163-015-0180-zSearch in Google Scholar

[7] Azizieh C., Clément P. A priori estimates and continuation methods for positive solutions of p-Laplace equations. J. Differential Equations 179 (2002), no. 1, 213-245.10.1006/jdeq.2001.4029Search in Google Scholar

[8] Ruiz D. A priori estimates and existence of positive solutions for strongly nonlinear problems. J. Differential Equations 199 (2004), no. 1, 96-114.10.1016/j.jde.2003.10.021Search in Google Scholar

[9] Dall’Aglio A., De Cicco V., Giachetti D., Puel J.-P. Existence of bounded solutions for nonlinear elliptic equations in unbounded domains. NoDEA Nonlinear Differential Equations Appl. 11 (2004), no. 4, 431-450.10.1007/s00030-004-1070-0Search in Google Scholar

[10] Lorca S., Ubilla P. A priori estimate for a quasilinear problem depending on the gradient. J. Math. Anal. Appl. 367 (2010), no. 1, 69-74.10.1016/j.jmaa.2009.12.030Search in Google Scholar

[11] Zou H.H. A priori estimates and existence for quasi-linear elliptic equations. Calc. Var. Partial Differential Equations 33 (2008), no. 4, 417-437.10.1007/s00526-008-0168-3Search in Google Scholar

[12] Damascelli L., Sciunzi B. Regularity, monotonicity and symmetry of positive solutions of m-Laplace equations. J. Differential Equations 206 (2004), no. 2, 483-515.10.1016/j.jde.2004.05.012Search in Google Scholar

[13] Damascelli L., Sciunzi B. Harnack inequalities, maximum and comparison principles, and regularity of positive solutions of m-Laplace equations. Calc. Var. Partial Differential Equations 25 (2006), no. 2, 139-159.10.1007/s00526-005-0337-6Search in Google Scholar

[14] Brezis H., Merle F. Uniform estimates and blow-up behavior for solutions ofΔu = V(x)eu in two dimensions. Comm. Partial Differential Equations 16 (1991), no. 8-9, 1223-1253.10.1080/03605309108820797Search in Google Scholar

[15] Aguilar Crespo J.A., Peral Alonso I. Blow-up behavior for solutions ofΔNu = V(x)eu in bounded domains inN. Nonlinear Anal. 29 (1997), no. 4, 365-384.10.1016/S0362-546X(96)00042-9Search in Google Scholar

[16] Esposito P., Morlando F. On a quasilinear mean field equation with an exponential nonlinearity. J. Math. Pures Appl. (9) 104 (2015), no. 2, 354-382.10.1016/j.matpur.2015.03.001Search in Google Scholar

[17] Passalacqua T. Some applications of functional inequalities to semilinear elliptic equations Ph.D. Thesis, Università degli Studi di Milano (2015).Search in Google Scholar

[18] Robert F., Wei J.-C. Asymptotic behavior of a fourth order mean field equation with Dirichlet boundary condition. Indiana Univ. Math. J. 57 (2008), no. 5, 2039-2060.10.1512/iumj.2008.57.3324Search in Google Scholar

[19] Mancini G., Romani G. Uniform bounds for higher-order semilinear problems in conformal dimension. arXiv:1710.05354v310.1016/j.na.2019.111717Search in Google Scholar

[20] Lieberman G.M. Boundary regularity for solutions of degenerate elliptic equations. Nonlinear Anal. 12 (1988), no. 11, 1203-1219.10.1016/0362-546X(88)90053-3Search in Google Scholar

[21] Tolksdorf P. Regularity for a more general class of quasilinear elliptic equations. J. Differential Equations 51 (1984), no. 1, 126-150.10.1016/0022-0396(84)90105-0Search in Google Scholar

[22] Serrin J. Local behavior of solutions of quasi-linear equations. Acta Math. 111 (1964) 247-302.10.1007/BF02391014Search in Google Scholar

[23] Allegretto W., Huang Y.X. A Picone’s identity for the p-Laplacian and applications. Nonlinear Anal. 32 (1998), no. 7, 819-830.10.1016/S0362-546X(97)00530-0Search in Google Scholar

[24] Esposito P. A classification result for the quasi-linear Liouville equation. Ann. Inst. H. Poincaré Anal. Non Linéaire 35 (2018), no. 3, 781-801.10.1016/j.anihpc.2017.08.002Search in Google Scholar

[25] Kichenassamy S., Véron L. Singular solutions of the p-Laplace equation. Math. Ann. 275 (1986), no. 4, 599-615.10.1007/BF01459140Search in Google Scholar

[26] Royden H.L. Real Analysis. 2nd edition. Macmillan. 1968 xii+349 pp.Search in Google Scholar

[27] Ren X., Wei J.-C. Counting peaks of solutions to some quasilinear elliptic equations with large exponents. J. Differential Equations 117 (1995), no. 1, 28-55.10.1006/jdeq.1995.1047Search in Google Scholar

[28] de Figueiredo D., do Ó J.M., Ruf B. Semilinear elliptic systems with exponential nonlinearities in two dimensions. Adv. Nonlinear Stud. 6 (2006), no. 2, 199-213.10.1515/ans-2006-0205Search in Google Scholar

[29] Bénilan P., Boccardo L., Gallouët T., Gariepy R., Pierre M., Vázquez J.L. An L1-theory of existence and uniqueness of solutions of nonlinear elliptic equations. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 22 (1995), no. 2, 241-273.Search in Google Scholar

Received: 2018-11-01
Accepted: 2019-03-15
Published Online: 2019-09-20

© 2020 Giulio Romani, published by De Gruyter

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

Downloaded on 27.5.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2020-0025/html
Scroll to top button