This work presents some improvements on related papers that investigate certain overdetermined problems associated to elliptic quasilinear operators. Our model operator is the p-Laplacian. Under suitable structural conditions, and assuming that a solution exists, we show that the domain of the problem is a ball centered at the origin. Furthermore we discuss a convenient form of comparison principle for this kind of problems.
Citation: Antonio Greco, Francesco Pisanu. Improvements on overdetermined problems associated to the p-Laplacian[J]. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022017
[1] | Chao Xia, Jiabin Yin . Two overdetermined problems for anisotropic $ p $-Laplacian. Mathematics in Engineering, 2022, 4(2): 1-18. doi: 10.3934/mine.2022015 |
[2] | Michiaki Onodera . Linear stability analysis of overdetermined problems with non-constant data. Mathematics in Engineering, 2023, 5(3): 1-18. doi: 10.3934/mine.2023048 |
[3] | Luigi Provenzano, Alessandro Savo . Isoparametric foliations and the Pompeiu property. Mathematics in Engineering, 2023, 5(2): 1-27. doi: 10.3934/mine.2023031 |
[4] | Filippo Gazzola, Gianmarco Sperone . Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040 |
[5] | Rolando Magnanini, Giorgio Poggesi . Interpolating estimates with applications to some quantitative symmetry results. Mathematics in Engineering, 2023, 5(1): 1-21. doi: 10.3934/mine.2023002 |
[6] | Bruno Bianchini, Giulio Colombo, Marco Magliaro, Luciano Mari, Patrizia Pucci, Marco Rigoli . Recent rigidity results for graphs with prescribed mean curvature. Mathematics in Engineering, 2021, 3(5): 1-48. doi: 10.3934/mine.2021039 |
[7] | 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 |
[8] | Peter Bella, Mathias Schäffner . Local boundedness for $ p $-Laplacian with degenerate coefficients. Mathematics in Engineering, 2023, 5(5): 1-20. doi: 10.3934/mine.2023081 |
[9] | Fernando Charro, Byungjae Son, Peiyong Wang . The Gelfand problem for the Infinity Laplacian. Mathematics in Engineering, 2023, 5(2): 1-28. doi: 10.3934/mine.2023022 |
[10] | María Medina, Pablo Ochoa . Equivalence of solutions for non-homogeneous $ p(x) $-Laplace equations. Mathematics in Engineering, 2023, 5(2): 1-19. doi: 10.3934/mine.2023044 |
This work presents some improvements on related papers that investigate certain overdetermined problems associated to elliptic quasilinear operators. Our model operator is the p-Laplacian. Under suitable structural conditions, and assuming that a solution exists, we show that the domain of the problem is a ball centered at the origin. Furthermore we discuss a convenient form of comparison principle for this kind of problems.
This paper deals with the equation
−Δpu=f(|x|,u) | (1.1) |
where Δp denotes the p-Laplace operator Δpu=div(|Du|p−2Du) with p∈(1,+∞). The function f(r,y) is defined for all (r,y)∈(0,+∞)×[0,+∞), it is positive whenever y>0, and satisfies the following conditions (labels correspond to [4]).
(H1) For almost every r∈(0,+∞), the function y↦f(r,y) is continuous with respect to y∈[0,+∞). Furthermore, for every y∈[0,+∞) and r0>0, the function r↦f(r,y) belongs to L∞((0,r0)).
(H2) For a.e. r∈(0,+∞) the function y↦f(r,y)/yp−1 is strictly decreasing with respect to y∈(0,+∞).
(H3) For every bounded interval (0,r0) there exists a constant C(r0) such that f(r,y)≤C(r0)(yp−1+1) for a.e. r∈(0,r0) and for all y∈[0,+∞).
Under assumptions (H1), (H2) and (H3), J. I. Díaz and J. E. Saa in the fundamental paper [4] proved uniqueness of the bounded, weak solution u∈W1,p0(Ω) to the Dirichlet problem
{−Δpu=f(|x|,u), u≥0, u≢0inΩ;u(x)=0forx∈∂Ω, | (1.2) |
where Ω is a bounded, smooth, open subset of RN, N≥2. To be precise, the function f in [4] is allowed to depend on (x,u), and the assumptions stated there are less demanding because they are tailored on the specific set Ω. In the present paper, instead, the set Ω is an unknown of the problem: thus, conditions (H1) and (H3) above correspond to the requirement that [4,(H1) and (H3)] hold in every bounded Ω. For the same reason we require
limy→0+f(r,y)yp−1=+∞and limy→+∞f(r,y)yp−1=0fora.e.r∈(0,+∞). | (1.3) |
The last assumptions, together with (H1), (H2) and (H3), ensure the existence of a bounded weak solution u∈W1,p0(Ω) of problem (1.2) by [4,Théorème 2] (see also [1]). Such a solution is in fact positive in Ω and (if the domain is sufficiently smooth) belongs to the Hölder class C1,α(¯Ω) ([4], p. 522, last paragraph). Here, however, we start from the assumption that problem (1.6) is solvable in the class C1(¯Ω), so that the boundary conditions are intended in the classical sense: this implies that Ω is a domain of class C1 (see Lemma A.2). Concerning the regularity of f, we need two assumptions. First we require that
fislocallyuniformlyHöldercontinuous | (1.4) |
i.e., f(r,y) is uniformly Hölder continuous in every compact subset of (0,+∞)×[0,+∞). Clearly, the first part of condition (H1) is an immediate consequence of (1.4). If the weak solution u of problem (1.2) has a non-vanishing gradient at some point x∈Ω, then the operator Δpu is non-degenerate, and if, furthermore, x≠0, then by (1.4) u belongs to the Hölder class C2,α in a neighborhood of x (see, for instance, [5,Theorem 15.9]). The interior smoothness of u is required by Lemma 2.2 (a boundary-point lemma). In order to apply the lemma, we also need that for every R1>0 there exist δ∈(0,R1) and L∈R such that
f(r,y1)−f(r,y2)y1−y2≥L | (1.5) |
for every r∈(R1−δ,R1) and 0<y1<y2<δ. Assuming that Ω contains the origin, and denoting by q:(0,+∞)→(0,+∞) a prescribed function, we consider the overdetermined problem
{−Δpu=f(|x|,u), u>0inΩ;u(x)=0, |Du(x)|=q(|x|)forx∈∂Ω. | (1.6) |
Contrary to what one may expect, counterexamples show that problem (1.6) may well be solvable even though the domain Ω is not radially symmetric: see, for instance, [6,pp. 488–489] and [8,Section 5]. The purpose of the present paper is to find conditions on f,q such that (1.6) is solvable only if Ω is a ball centered at the origin. More precisely, we remove the restrictions that f(r,y) is monotone non-increasing with respect to r or y, which were imposed in [6] and [7], respectively. In general, given f and q, we introduce the function F(r,ρ,λ) by letting
F(r,ρ,λ)=(q(r))p−1r f(ρr,λrq(r)) |
for r,ρ,λ>0, and we assume that for every ρ,λ
F(r,ρ,λ)ismonotonenon−decreasingwithrespecttor. | (1.7) |
In Section 4 we demonstrate some special cases when condition (1.7) holds true. Our main result is the following:
Theorem 1.1. Let Ω be a bounded, connected open set in RN, N≥2, containing the origin. Suppose that the functions f and q are such that (H1), (H2) and (H3) hold, together with (1.3), (1.4), (1.5) and (1.7). If there exists a weak solution u∈C1(¯Ω) of the overdetermined problem (1.6), then Ω is a ball centered at the origin.
Similarly to [6,7,8], the result is achieved by means of a comparison with two radial functions. However, the supersolution is chosen following an idea in [9]: see Section 5 for the proof of the theorem. The result is new even in the case when p=2, i.e., Δp=Δ: an example is given by the sublinear Hénon problem
{−Δu=|x|MruMy, u>0inΩ;u=0, |Du(x)|=|x|1+μon∂Ω; | (1.8) |
with constants μ,Mr,My≥0 satisfying
μ−Mr2+μ≥My. | (1.9) |
See Theorem 4.3 for details. A further advancement lies in the fact that the strict monotonicity of F(r,ρ,λ) in r is not required. For instance, we have:
Corollary 1.2. Let Ω be a bounded, connected open set in RN, N≥2, containing the origin. Suppose that (H1) and (H3) hold, together with (1.3), (1.4) and (1.5). Suppose, further, that f(r,y) is non-increasing in r>0, and there exists a constant ε0∈(0,p−1] such that for every r∈(0,+∞) the function y↦f(r,y)/yp−1−ε0 is monotone non-increasing with respect to y∈(0,+∞). Let q be such that
theratioq(r)rpε0−1isnon−decreasinginr. | (1.10) |
If there exists a weak solution u∈C1(¯Ω) of the overdetermined problem (1.6), then Ω is a ball centered at the origin.
Corollary 1.2 improves [6,Theorem 1.2] because the monotonicity in (1.10) is intended in the broad sense. To achieve this, we develop in the next section a convenient boundary-point lemma. The lemma yields a sharper comparison between the solution u and the radial solution used in the proof of Theorem 1.1, which are regular enough by virtue of assumption (1.4).
Roughly speaking, the (weak) comparison principle asserts that if a subsolution u and a supersolution v of (1.1) satisfy u≤v on ∂Ω, then the same inequality holds a.e. in Ω. The strong comparison principle, instead, under the same assumption asserts that either u<v a.e. in Ω, or u and v are the same function. Note that the term strong comparison principle is sometimes used to refer to a class of theorems ensuring that u<v in Ω provided that the weaker inequality u≤v holds in the whole domain Ω (and u≢v): see for instance [3,Theorem 1.4] and the results in [16,17].
Due to the nonlinearity of the p-Laplacian, the difference u−v is not a subsolution, in general (unless p=2), and therefore the comparison principle, far from being a straightforward consequence of the maximum principle, is a self-standing and interesting task.
Several comparison principles for the p-Laplacian are found in the literature: for instance, the case when f vanishes identically is considered in [10,11]. For f=f(x) see [12]. For f depending on (x,y) and monotone in y let us mention [2,Proposition 2.3 (b)], as well as [15,Proposition 2.1], where f(x,y) is required to be non-increasing in y. The case when f(x,y)=a(x)yp−1 is considered in [8,Theorem 7.1]. More general nonlinearities (still non-decreasing in y) are admitted in [2,Theorem 2.1 and Proposition 2.3 (a)] provided that u=v=0 on the boundary.
We make use of a comparison principle that holds regardless of the monotonicity of f in y, and does not require that u=v=0 on ∂Ω.
Theorem 2.1 (Comparison principle). Let Ω be an open set in RN, N≥2, possibly non-smooth and unbounded. Let u,v∈W1,p(Ω) be a weak subsolution and a weak supersolution, respectively, of equation (1.1) in Ω, where f satisfies (H1), (H2) and (H3). Assume that u,v>0 a.e. in Ω, the ratio u/v belongs to L∞(Ω), and (u−v)+∈W1,p0(Ω). Then u≤v a.e. in Ω.
Proof. The claim is an application of [6,Theorem 1.1 (2)], which holds for f=f(x,y), to the special case when f=f(|x|,y).
Note that if Ω is bounded, and if u,v are smooth up to the boundary, positive in Ω, satisfy u≤v on ∂Ω, and have non-vanishing gradients Du,Dv on ∂Ω, then the ratio u/v is bounded as required in Theorem 2.1: see Lemma A.1 for details. In order to prove Theorem 1.1 we also need a boundary-point lemma involving the outward derivatives of a smooth subsolution u and a smooth supersolution v with non-vanishing gradients:
Lemma 2.2 (Boundary-point lemma). Let G be a connected, bounded open set in RN, N≥2, satisfying the interior sphere condition at z1∈∂G, i.e., there exists a ball B⊂G such that z1∈∂B. Let u,v∈C2(G)∩C1(¯G) be such that u≤v in ¯G and Du(x),Dv(x)≠0 in ¯G, as well as u(z1)=v(z1). Assume
Δpv+f(x,v(x))≤Δpu+f(x,u(x))pointwiseinG, | (2.1) |
where f:G×(a,b)→R satisfies a Lipschitz condition in y from below, i.e. there exists a constant L such that
f(x,y1)−f(x,y2)y1−y2≥L | (2.2) |
for all x∈G and a<y1<y2<b. Here (a,b) is an interval such that u(x),u(y)∈(a,b) for all x∈G. Suppose, further, that either u or v belongs to the class C2(¯G). Then either u=v in G or
∂u∂ν(z1)>∂v∂ν(z1), |
where ν denotes the outward derivative to the ball B at z1.
Proof. Following [5,Theorem 10.1], we derive an inequality satisfied by the difference w=u−v∈C2(G)∩C1(¯G). However, in the present case assumption (ⅲ) of [5] (monotonicity of f with respect to y) is not in effect: this difficulty is overcome because w≤0 in the whole of ¯G by assumption. To be more specific, for ξ=(ξ1,…,ξN)∈RN∖{0} and for i,j=1,…,N define
aij(ξ)=|ξ|p−2δij+(p−2)|ξ|p−4ξiξj, |
where δij is Kronecker's delta, and notice that aij(ξ) is continuously differentiable in the punctured space RN∖{0}. Since u,v∈C2(G) and Du,Dv≠0, the p-Laplacian may be rewritten as Δpu=aij(Du(x))uij(x), where uij denotes the second derivative of u with respect to xixj and the summation over repeated indices is understood. Similarly, we may write Δpv=aij(Dv(x))vij(x), and by (2.1) we obtain
aij(Dv(x))wij(x)+(ahk(Du(x))−ahk(Dv(x)))uhk(x)+f(x,u(x))−f(x,v(x))≥0, |
where the summation over i,j and h,k=1,…,N is understood. Letting wi=∂w/∂xi, we have (ui(x)−vi(x))wi(x)=|Du(x)−Dv(x)|2, and therefore the inequality above may be rewritten as
aij(Dv(x))wij(x)+bi(x)wi(x)+c(x)w(x)≥0 inG, | (2.3) |
where the coefficients bi and c are defined as follows:
bi(x)={ahk(Du(x))−ahk(Dv(x))|Du(x)−Dv(x)|ui(x)−vi(x)|Du(x)−Dv(x)|uhk(x),Du(x)≠Dv(x); 0,Du(x)=Dv(x), |
c(x)={f(x,u(x))−f(x,v(x))u(x)−v(x),u(x)≠v(x); 0,u(x)=v(x). |
The coefficients aij(Dv(x)) in (2.3) are bounded in ¯G because Dv(x) is continuous in ¯G by assumption, and keeps away from zero. Assuming that u∈C2(¯G), let us check that the coefficients bi are also bounded in ¯G. Since the derivatives wi=ui−vi satisfy |wi(x)|≤|Dw(x)|, the ratio
ui(x)−vi(x)|Du(x)−Dv(x)| |
is clearly bounded. Hence suppose, contrary to the claim, that there exists a sequence of points xn∈¯G converging to some z1∈¯G and such that Du(xn)≠Dv(xn) for every n≥1 and
limn→+∞|ahk(Du(xn))−ahk(Dv(xn))||Du(xn)−Dv(xn)|=+∞. | (2.4) |
Since Du(x),Dv(x) are bounded in ¯G and keep away from zero, the numerator keeps bounded, and therefore we must have Du(z1)=Dv(z1)=ξ0≠0. Letting R=12|ξ0|, we have Du(xn),Dv(xn)∈BR(ξ0) for n large, and therefore the whole segment joining Du(xn) and Dv(xn) is included in BR(ξ0): hence it does not intersect the origin. We note in passing that the last assertion does not follow from the only fact that Du(x),Dv(x) keep far from zero (think to the case when Du=−Dv). Letting
Mhk=maxξ∈¯BR(ξ0)|Dahk(ξ)|, |
by the mean value theorem we have
|ahk(Du(xn))−ahk(Dv(xn))|≤Mhk|Du(xn)−Dv(xn)|, |
contradicting (2.4). Thus, in order to conclude that the coefficients bi(x) are bounded in ¯G, it suffices to recall that uhk(x) is continuous in ¯G by assumption. In the case when v∈C2(¯G), instead, a similar argument leads to the linear inequality
aij(Du(x))wij(x)+˜bi(x)wi(x)+c(x)w(x)≥0 inG, |
with bounded coefficients ˜bi. In both cases, the boundedness of c(x) from below in G follows from (2.2). The sign of c(x) is not prescribed by assumption. However, since w(x)≤0, we have −c−(x)w(x)≥c(x)w(x), where −c−(x)=min{c(x),0}≤0. Hence, from (2.3) we deduce
aij(Dv(x))wij(x)+bi(x)wi(x)−c−(x)w(x)≥0 inG |
and the conclusion follows from the Hopf boundary-point lemma: see, for instance, [14,Corollary 2.8.5] (the Hopf lemma is also found in [13,Theorem 8,p. 67], but the assumptions on the boundedness of the coefficients need to be recovered from Theorem 6 and the subsequent remarks).
In the following lemma we summarize the properties of the radial solutions of problem (1.2) that are needed in the proof of Theorem 1.1.
Lemma 3.1. Let Ω=BR(0) for some R>0, and let f satisfy (H1), (H2), (H3), (1.3) and (1.4). Then the unique solution u=uR of problem (1.2) possesses the following properties:
1) uR is radially symmetric;
2) for every ε∈(0,R) there exists α∈(0,1) such that uR belongs to the Hölder class C1,α(¯BR(0))∩C2,α(¯BR(0)∖Bε(0));
3) DuR≠0 on ∂BR(0).
Proof. As mentioned in the Introduction, uniqueness follows from (H1), (H2) and (H3). Using (1.3), we also have the existence of a (positive) weak solution u=uR∈C1,α(¯BR(0)) for each R>0. Since problem (1.2) is invariant under rotations about the origin, uniqueness implies that uR is a radial function (Claim 1), and we may write uR(x)=vR(|x|) for a convenient function vR(r). Using the divergence theorem in the ball Br(0) for any r∈(0,R], we obtain
−|v′R(r)|p−2v′R(r)=∫r0f(ρ,vR(ρ))dρ, | (3.1) |
which immediately implies the last claim: in fact, it turns out that DuR(x)=0 if and only if x=0. Claim 2 follows from (1.4) and (3.1).
Before proving Theorem 1.1 in the next section, we show some special cases satisfying (1.7). In order to compare Theorem 1.1 with Theorem 1.2 of [6], we prove
Proposition 4.1. Assume that f(r,y) is non-negative and non-increasing in r>0. Suppose, further, that there exists ε0∈(0,p−1] such that (1.10) holds, and the ratio f(r,y)/yp−1−ε0 is non-increasing in y. Then (1.7) is satisfied.
Proof. Fix ρ,λ>0 and denote by ζ(r) the function ζ(r)=q(r)r1−p/ε0, which is non-decreasing by (1.10). Since rq(r)=ζ(r)rp/ε0, the variables t=ρ/r and y=λ/(rq(r)) decrease as r increases. Keeping this in mind, we write
f(ρr,λrq(r))=λp−1−ε0(rq(r))p−1−ε01yp−1−ε0f(t,y) |
and we observe that (q(r))p−1/r=(ζ(r))ε0(rq(r))p−1−ε0. Hence
(q(r))p−1rf(ρr,λrq(r))=λp−1−ε0(ζ(r))ε0f(t,y)/yp−1−ε0 |
and (1.7) follows.
Now let us take [7,Theorem 1.1] into consideration. To this purpose we consider f(r,y) non-negative and non-increasing in y>0, and we assume that there exists μ∈[−p,+∞) such that both
(q(r))p−1r1+μandrμf(1r,y)arenon−decreasingwithrespecttor>0. | (4.1) |
Note that (4.1) follows from (1.4) and (1.5) in [7]: more precisely, letting μ=−1−(p−1)σ, condition (1.4) in [7] is equivalent to the strict monotonicity of (q(r))p−1/r1+μ. Note, further, that if (H1) is in effect, then μ must belong to the interval [0,+∞) in order that the second expression in (4.1) is non-trivial and non-decreasing in r: indeed, if we take μ<0 and let r→+∞, then f(1/r,y) keeps bounded by (H1) and therefore rμf(1/r,y)→0. This and the monotonicity imply that the non-negative function f must vanish identically, but then problem (1.6) is unsolvable. Thus, the only case compatible with (H1) is when μ≥0, which corresponds to σ≤−1/(p−1): this was not mentioned in [7].
Proposition 4.2. If f(r,y) is non-negative and non-increasing in y>0, and if q(r) and f(r,y) satisfy (4.1) for some μ≥−p, then condition (1.7) holds true.
Proof. Denoting by ψ(r) the monotone non-decreasing function given by ψ(r)=(q(r))p−1/r1+μ, we may write
(q(r))p−1r=ψ(r)rμ. | (4.2) |
Hence, for every ρ>0 we have
(q(r))p−1rf(ρr,y)=ψ(r)rμf(ρr,y)=ψ(r)ρμsμf(1s,y), |
where s=r/ρ. Since the right-hand side is non-decreasing by (4.1), we deduce that
(q(r))p−1rf(ρr,y)isnon−decreasinginr>0. | (4.3) |
To conclude the proof, we rewrite (4.2) as (rq(r))p−1=ψ(r)rμ+p, which implies that rq(r) is non-decreasing (because μ+p≥0): recalling that f(r,y) is non-increasing in y, this and (4.3) imply (1.7).
When the functions q(r) and f(r,y) are differentiable, a sufficient condition in order that (1.7) holds is given by the next theorem, which is applicable to the example in (1.8) and (1.9). Before proceeding further, recall that for every M∈R and for every positive, differentiable function g(r) the condition (g(r)/rM)′≤0 for r>0 is equivalent to
rg′(r)g(r)≤M. |
We will use several expressions like that in the sequel.
Theorem 4.3. Let q(r) be a positive, differentiable function of r>0 and define ϕ(r)=(q(r))p−1/r. Suppose there exists a constant μ such that
infr>0rϕ′(r)ϕ(r)≥μ∈(−p,+∞). | (4.4) |
Furthermore, let f(r,y) be positive and differentiable for every r,y>0, and denote fr=∂f/∂r and fy=∂f/∂y, for shortness. Suppose there exist constants Mr,My such that
supr,y>0rfr(r,y)f(r,y)≤Mr∈(−p,+∞), | (4.5) |
supr,y>0yfy(r,y)f(r,y)≤My∈(−∞,+∞). | (4.6) |
If
μ−Mrp+μ≥Myp−1 | (4.7) |
then (1.7) holds true.
Proof. Let us check that
ddr{ϕ(r)f(ρr,λrq(r))}≥0 |
for every ρ,λ>0. Letting t=ρ/r and y=λ/(rq(r)), and since f is differentiable, the inequality above may be rewritten as
rϕ′(r)ϕ(r)≥tfr(t,y)f(t,y)+yfy(t,y)f(t,y)(1+rq′(r)q(r)). | (4.8) |
Using the identity
1+rϕ′(r)ϕ(r)=(p−1)rq′(r)q(r) | (4.9) |
together with assumption (4.4), we see
1+rq′(r)q(r)=1p−1(p+rϕ′(r)ϕ(r))>0. |
Hence, in view of assumptions (4.5) and (4.6), it is enough to ensure
rϕ′(r)ϕ(r)≥Mr+Myp−1(p+rϕ′(r)ϕ(r)) |
in order that (4.8) holds. Of course, the last inequality is equivalent to
rϕ′(r)ϕ(r)−Mrp+rϕ′(r)ϕ(r)≥Myp−1. |
To complete the proof, observe that the rational function (x−Mr)/(p+x) is strictly increasing in the variable x>−p (because Mr+p>0), hence assumption (4.7) implies the claim.
Remark 4.4. By virtue of (4.9), we may define ω=(1+μ)/(p−1) and rewrite assumptions (4.4) and (4.7), respectively, as
infr>0rq′(r)q(r)≥ω∈(−1,+∞),(p−1)ω−1−Mrω+1≥My. |
Remark 4.5. Assumption (H1) implies Mr,My≥0. Indeed, if Mr<0 then the ratio ϑ(r,y)=f(r,y)/rMr is strictly decreasing in r, and therefore
limr→0+f(r,y)=limr→0+rMrϑ(r,y)=+∞, |
which is in contrast with (H1). A similar argument shows that My≥0. Assumption (H2) immediately implies My≤p−1. Finally, if Mr,My≥0 and (4.7) is in effect, then obviously μ≥0.
Example 4.6. Theorem 4.3 is applicable, for instance, to the case when p=2, q(r)=r1+μ, and f(r,y)=rMryMy, with constants μ,Mr,My≥0 satisfying (1.9). In this case, the assumptions in [6] are not satisfied (when Mr>0), nor hold the assumptions in [7] (if My>0).
Proof of Theorem 1.1. Part Ⅰ. Assume that there exists a weak solution u∈C1(¯Ω) of the overdetermined problem (1.6). Then Ω is a domain of class C1 by Lemma A.2. Define
R1=minx∈∂Ω|x|,R2=maxx∈∂Ω|x|, |
and let u1 be the positive solution of problem (1.2) in the ball B1=BR1(0). Since u≥0 in ¯Ω, we have 0=u1≤u on ∂B1. In order to apply the comparison principle (Theorem 2.1) in the ball B1 we need to check that u1/u∈L∞(B1): this follows from Lemma A.1 because u vanishes at a boundary point z1∈∂B1 if and only if z1∈∂Ω: but then |Du(z1)|=q(|z1|)>0, hence the assumptions of the lemma are satisfied. By the comparison principle we have u1≤u in B1, and therefore for each z1∈∂B1∩∂Ω we may write
|Du1(z1)|≤|Du(z1)|. | (5.1) |
Part Ⅱ. We claim that the equality holds in (5.1) if and only if Ω=B1. To prove this, suppose that |Du1(z1)|=|Du(z1)|>0 at some z1∈∂B1∩∂Ω. By continuity, there exists ε∈(0,R1) such that if we define G=B1∩Bε(z1) then we have Du1,Du≠0 in ¯G. By (1.5), and by reducing ε if necessary, we may assume that for every x∈G the pairs (|x|,u1(x)) and (|x|,u(x)) belong to the set (R1−δ,R1)×(0,δ) where f satisfies a Lipschitz condition w.r.t. y from below. Hence by the boundary-point lemma (Lemma 2.2) we have u1=u in G, which implies u=0 on ¯G∩∂B1. This shows that the set of all z∈∂B1 such that u(z)=0 is a relatively open subset of ∂B1. Since such a set is obviously closed, it follows that u=0 on ∂B1, hence ∂B1⊂∂Ω. Finally, since Ω is connected, we must have Ω=B1. Thus, the equality holds in (5.1) if and only if Ω=B1. To complete the proof of the theorem, it is enough to verify that |Du1(z1)|=|Du(z1)|.
Part Ⅲ. Let B2=BR2(0), for shortness, and define
a=R2q(R2)R1q(R1). |
With such a value of the parameter a, the function v:B2→R given by v(x)=au1(R1x/R2) satisfies
|Dv|∂B2q(R2)=|Du1|∂B1q(R1)≤1, | (5.2) |
where the last inequality is a consequence of (5.1). Let us check that v is a supersolution of (1.1). By a straightforward computation we obtain
−Δpv(x)=−(q(R2)q(R2))p−1R1R2Δpu1(R1x/R2)=(q(R2)q(R2))p−1R1R2f(ρR2,λR2q(R2)), |
where we have put ρ=R1x and λ=R1q(R1)v(x). Now, using assumption (1.7), it follows that −Δpv(x)≥f(|x|,v(x)), hence v is a supersolution of (1.1), as claimed.
Conclusion. Since v≥0 in ¯B2⊃∂Ω, and u=0 on ∂Ω, using Theorem 2.1 we find u≤v in Ω. Therefore at each z2∈∂B2∩∂Ω we have |Du(z2)|≤|Dv(z2)|, hence
1≤|Dv|∂B2q(R2). |
By comparing the last inequality with (5.2) we see that |Du1|∂B1=q(R1), hence the equality holds in (5.1), and the theorem follows.
Proof of Corollary 1.2. To prove the corollary it suffices to observe that its assumptions imply (1.7) by virtue of Proposition 4.1. Note that assumption (H2), although not mentioned in the statement, is silently in effect as a consequence of the monotonicity of the ratio f(r,y)/yp−1−ε0.
The first author is a member of the Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM) and is partially supported by the research project Evolutive and stationary Partial Differential Equations with a focus on bio-mathematics, funded by Fondazione di Sardegna (2019).
The authors declare no conflict of interest.
In order to apply Theorem 2.1 (comparison principle) we need to know that u/v∈L∞(Ω). It was observed in [4,p. 522,last two lines] that the ratio u/v is bounded in a bounded domain Ω in the case when u and v are positive in Ω, vanish along the boundary, and have a positive inward derivative ∂u/∂ν on ∂Ω (which is equivalent to having a non-vanishing gradient there).
The assumption u=v=0 on ∂Ω can be turned into 0≤u≤v on ∂Ω and the conclusion continues to hold: this is used in the proof of Theorem 1.1, and it was also used in [6] (see the three lines following (17) on p. 405). Let us give a precise statement and proof.
Recall that a domain Ω⊂RN with nonempty boundary is of class C1 when ∂Ω is locally the graph of a continuously differentiable function. Following [5,p. 10], we write u∈C1(¯Ω) if u∈C1(Ω)∩C0(¯Ω) and each derivative ui:Ω→R is the restriction to Ω of a continuous function (still denoted by ui) defined on ¯Ω. In such a case, the gradient Du=(u1,…,uN) is defined in the closure ¯Ω. In this section we represent x∈RN as x=(x′,xN), where x′=(x1,…,xN−1)∈RN−1, and we denote by eN the N-th element of the canonical base of RN. Furthermore we let Dx′u=(u1,…,uN−1). By a cylindrical neighborhood of the origin we mean the set Ur={(x′,xN):|x′|,|xN|<r} for some r>0.
Lemma A.1. Let Ω be a bounded domain of class C1 in RN, N≥2, and let u,v∈C1(¯Ω) be positive in Ω and satisfy 0≤u≤v on ∂Ω. Suppose that for any boundary point z∈∂Ω where v(z)=0, the gradient Dv(z) does not vanish. Then the ratio u/v is bounded in Ω.
Proof. Suppose, contrary to the claim, that there exists a sequence of points xk∈Ω such that u(xk)/v(xk)→+∞. Without loss of generality we may assume that xk converges to z∈¯Ω. Since the ratio u(x)/v(x) is continuous (and finite) in Ω, we must have z∈∂Ω, and u(z)=v(z)=0. Since Ω is a C1-domain, after a convenient translation and rotation of the coordinate frame we may further assume that z=0, and there exists a cylindrical neighborhood Ur of the origin such that the intersection Ur∩∂Ω is the graph of a continuously differentiable function xN=f(x′) satisfying Dx′f(0)=0. We may also assume that the intersection Ur∩Ω lies above the graph of f, i.e., if (x′,xN)∈Ur∩Ω then xN>f(x). Since v(0)=0≤v(x) for all x∈¯Ω, the directional derivative vξ(0)=ξ⋅Dv(0) satisfies vξ(0)≥0 for every direction ξ such that ξ⋅eN≥0: this implies Dx′v(0)=0 and therefore we may let λ0=vN(0)>0 because Dv(0)≠0. By shrinking the neighborhood Ur if necessary, we may achieve that vN(x)>12λ0 in Ur∩Ω. Furthermore we define C=supΩuN(x)<+∞. Finally, recall that u≤v on ∂Ω. By the fundamental theorem of calculus, for every x=(x′,xN)∈Ur∩Ω we have
u(x)v(x)≤u(x′,f(x′))+C(xN−f(x′))v(x′,f(x′))+12λ0(xN−f(x′))≤v(x′,f(x′))+C(xN−f(x′))v(x′,f(x′))+12λ0(xN−f(x′))=ρ(x)+Cρ(x)+12λ0≤max{1,2Cλ0} |
where
ρ(x)=v(x′,f(x′))xN−f(x′)≥0. |
This contradicts the assumption that u/v is unbounded, and the lemma follows.
The domain Ω is required to belong to the smoothness class C1 in order that the preceding lemma holds: let us prove that the regularity of the domain, which is not mentioned explicitly in the statement of Theorem 1.1, follows from the other assumptions.
Lemma A.2. Consider an open, proper subset Ω⊂RN, N≥2. If there exists u∈C1(¯Ω), u>0 in Ω, such that u(z)=0 and Du(z)≠0 for every z∈∂Ω, then Ω is a domain of class C1. Furthermore Du(z) has the direction of the inward normal to ∂Ω at z.
Proof. Part Ⅰ: Definitions. Let us fix a boundary point z0∈∂Ω. Without loss of generality, we may assume that z0=0 and Du(0)=λ0eN with some λ0>0. By the continuity of Du at 0, for every positive ε<λ0/2 there exists a cylindrical neighborhood Ur of the origin, with a convenient r=r(ε)>0, such that |Dx′u|,|uN−λ0|<ε in Ur∩¯Ω. In particular, uN>λ0−ε>0 in Ur∩¯Ω. Let ϑ be the unique solution of the equation tanϑ=(λ0−ε)/ε in the interval (π4,π2). For every ¯x∈RN we define the upper cone with vertex in ¯x and half-opening ϑ by V+ϑ(¯x)={(x′,xN):|x′−¯x′|<(xN−¯xN)tanϑ}. The corresponding lower cone is V−ϑ(¯x)={(x′,xN):|x′−¯x′|<(¯xN−xN)tanϑ}. Clearly, x∈V+ϑ(¯x) if and only if ¯x∈V−ϑ(x). Since ϑ>π4, for every x′∈RN−1 satisfying |x′|<r the vertical line ℓx′={(x′,t):t∈R} intersects both Ur∩V+ϑ(0) and Ur∩V−ϑ(0), hence
ℓx′∩Ur∩V+ϑ(0)≠∅,ℓx′∩Ur∩V−ϑ(0)≠∅. | (9.1) |
Part Ⅱ. Take any point ¯x=(¯x′,¯xN)∈Ur∩Ω. We claim that Ur∩V+ϑ(¯x)⊂Ω. Indeed, since the intersection Ur∩Ω is an open, nonempty subset of RN, then the segment S described by (¯x′,xN), when ¯x′ is kept fixed and xN>¯xN is let vary, is included in Ur∩Ω provided that xN−¯xN is small. Furthermore the value of u(¯x′,xN) is easily estimated by means of the fundamental theorem of calculus, and for xN>¯xN we may write
u(¯x′,xN)=u(¯x′,¯xN)+∫xN¯xNuN(¯x′,t)dt>u(¯x′,¯xN)+(λ0−ε)(xN−¯xN). |
Thus, u(¯x′,xN) stays positive along S and increases with xN, and therefore the segment S can be extended by allowing xN to range in the whole interval (¯xN,r) without hitting the boundary ∂Ω, where u vanishes.
Now, integrating along the orthogonal directions to eN, we find that Ur∩V+ϑ(¯x)⊂Ω, as claimed, and for every x=(x′,xN)∈Ur∩V+ϑ(¯x) we can estimate u(x) in terms of u(¯x) as follows:
u(x)−u(¯x)=u(x)−u(¯x′,xN)+u(¯x′,xN)−u(¯x)=∫10(x′−¯x′)⋅Dx′u(tx′+(1−t)¯x′)dt+∫xN¯xNuN(¯x′,t)dt>−ε|x′−¯x′|+(λ0−ε)(xN−¯xN)>0. |
Part Ⅲ. Observe that for every z∈Ur∩∂Ω we have Ur∩V+ϑ(z)⊂Ω because there exists a sequence of points ¯xk∈Ur∩Ω converging to z, and we may apply Part Ⅱ to each ¯xk. Furthermore, no point ¯x∈Ur∩V−ϑ(z) can be in Ω, for otherwise we would have z∈Ur∩V+ϑ(¯x)⊂Ω, a contradiction. Choosing z=0, and by (9.1), we see that for every x′ satisfying |x′|<r there exists xN∈(−r,r) such that the point (x′,xN) belongs to Ur∩∂Ω. Choosing z=(x′,xN) we see that the value of xN is unique: in other terms, the intersection Ur∩∂Ω is the graph of a function xN=f(x′). The argument above also shows that f satisfies a Lipschitz condition with constant L(ε)=cotϑ=ε/(λ0−ε). Since ε is arbitrary (and Ur shrinks to the origin, in general, when ε→0), it follows that f(x′) is differentiable at x′=0, and we have Dx′f(0)=0. This implies that ∂Ω has an inward normal ν at 0, which has the same direction as Du(0). Since the choice of the boundary point z0∈∂Ω is arbitrary, f(x′) must be differentiable at every x′ such that |x′|<r=r(ε), and the gradient Dx′f(x′) satisfies |Dx′f(x′)|≤L(ε). Since L(ε)→0 when ε→0, we have that Dx′f(x′) is continuous at x′=0, and the lemma follows.
[1] |
M. Belloni, B. Kawohl, A direct uniqueness proof for equations involving the p-Laplace operator, Manuscripta Math., 109 (2002), 229–231. doi: 10.1007/s00229-002-0305-9
![]() |
[2] | M. Cuesta, P. Takáč, A strong comparison principle for positive solutions of degenerate elliptic equations, Differ. Integral Equ., 13 (2000), 721–746. |
[3] | L. Damascelli, B. Sciunzi, Harnack inequalities, maximum and comparison principles, and regularity of positive solutions of m-Laplace equations, Calc. Var., 25 (2005), 139–159. |
[4] | J. I. Díaz, J. E. Saa, Existence et unicité de solutions positives pour certaines équations elliptiques quasilinéaires, C. R. Acad. Sci. Paris Sér. I Math., 305 (1987), 521–524. |
[5] | D. Gilbarg, N. S. Trudinger, Elliptic partial differential equations of second order, 2 Eds., Springer-Verlag, 1998. |
[6] |
A. Greco, Comparison principle and constrained radial symmetry for the subdiffusive p-Laplacian, Publ. Mat., 58 (2014), 485–498. doi: 10.5565/PUBLMAT_58214_24
![]() |
[7] |
A. Greco, Constrained radial symmetry for monotone elliptic quasilinear operators, J. Anal. Math., 121 (2013), 223–234. doi: 10.1007/s11854-013-0033-y
![]() |
[8] | A. Greco, Symmetry around the origin for some overdetermined problems, Adv. Math. Sci. Appl., 13 (2003), 387–399. |
[9] |
A. Greco, V. Mascia, Non-local sublinear problems: existence, comparison, and radial symmetry, Discrete Contin. Dyn. Syst., 39 (2019), 503–519. doi: 10.3934/dcds.2019021
![]() |
[10] | J. Heinonen, T. Kilpeläinen, O. Martio, Nonlinear potential theory of degenerate elliptic equations, New York: The Clarendon Press, Oxford University Press, 1993. |
[11] | P. Lindqvist, Notes on the stationary p-Laplace equation, Springer, 2019. |
[12] |
M. Lucia, S. Prashanth, Strong comparison principle for solutions of quasilinear equations, P. Am. Math. Soc., 132 (2003), 1005–1011. doi: 10.1090/S0002-9939-03-07285-X
![]() |
[13] | M. H. Protter, H. F. Weinberger, Maximum principles in differential equations, Springer-Verlag, 1984. |
[14] | P. Pucci, J. Serrin, The maximum principle, Birkhäuser, 2007. |
[15] |
P. Pucci, B. Sciunzi, J. Serrin, Partial and full symmetry of solutions of quasilinear elliptic equations, via the Comparison Principle, Contemp. Math., 446 (2007), 437–444. doi: 10.1090/conm/446/08643
![]() |
[16] |
P. Roselli, B. Sciunzi, A strong comparison principle for the p-Laplacian, P. Am. Math. Soc., 135 (2007), 3217–3224. doi: 10.1090/S0002-9939-07-08847-8
![]() |
[17] |
B. Sciunzi, Regularity and comparison principles for p-Laplace equations with vanishing source term, Commun. Contemp. Math., 16 (2014), 1450013. doi: 10.1142/S0219199714500138
![]() |
1. | Antonio Greco, Benyam Mebrate, An overdetermined problem related to the Finsler p$p$‐Laplacian, 2024, 70, 0025-5793, 10.1112/mtk.12267 | |
2. | Tynysbek Kalmenov, Nurbek Kakharman, An overdetermined problem for elliptic equations, 2024, 9, 2473-6988, 20627, 10.3934/math.20241002 |