Citation: Mouhamed Moustapha Fall, Veronica Felli, Alberto Ferrero, Alassane Niang. Asymptotic expansions and unique continuation at Dirichlet-Neumann boundary junctions for planar elliptic equations[J]. Mathematics in Engineering, 2019, 1(1): 84-117. doi: 10.3934/Mine.2018.1.84
[1] | Alessandra De Luca, Veronica Felli . Unique continuation from the edge of a crack. Mathematics in Engineering, 2021, 3(3): 1-40. doi: 10.3934/mine.2021023 |
[2] | Elena Beretta, M. Cristina Cerutti, Luca Ratti . Lipschitz stable determination of small conductivity inclusions in a semilinear equation from boundary data. Mathematics in Engineering, 2021, 3(1): 1-10. doi: 10.3934/mine.2021003 |
[3] | Yves Achdou, Ziad Kobeissi . Mean field games of controls: Finite difference approximations. Mathematics in Engineering, 2021, 3(3): 1-35. doi: 10.3934/mine.2021024 |
[4] | Andrea Manzoni, Alfio Quarteroni, Sandro Salsa . A saddle point approach to an optimal boundary control problem for steady Navier-Stokes equations. Mathematics in Engineering, 2019, 1(2): 252-280. doi: 10.3934/mine.2019.2.252 |
[5] | Catharine W. K. Lo, José Francisco Rodrigues . On an anisotropic fractional Stefan-type problem with Dirichlet boundary conditions. Mathematics in Engineering, 2023, 5(3): 1-38. doi: 10.3934/mine.2023047 |
[6] | Youchan Kim, Seungjin Ryu, Pilsoo Shin . Approximation of elliptic and parabolic equations with Dirichlet boundary conditions. Mathematics in Engineering, 2023, 5(4): 1-43. doi: 10.3934/mine.2023079 |
[7] | 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 |
[8] | François Murat, Alessio Porretta . The ergodic limit for weak solutions of elliptic equations with Neumann boundary condition. Mathematics in Engineering, 2021, 3(4): 1-20. doi: 10.3934/mine.2021031 |
[9] | Plamen Stefanov . Conditionally stable unique continuation and applications to thermoacoustic tomography. Mathematics in Engineering, 2019, 1(4): 789-799. doi: 10.3934/mine.2019.4.789 |
[10] | 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 |
The present paper deals with elliptic equations in planar domains with mixed boundary conditions and aims at proving asymptotic expansions and unique continuation properties for solutions near boundary points where a transition from Dirichlet to Neumann boundary conditions occurs.
A great attention has been devoted to the problem of unique continuation for solutions to partial differential equations starting from the paper by Carleman [5], whose approach was based on some weighted a priori inequalities. An alternative approach to unique continuation was developed by Garofalo and Lin [14] for elliptic equations in divergence form with variable coefficients, via local doubling properties and Almgren monotonicity formula; we also quote [18] for quantitative uniqueness obtained by monotonicity methods.
The monotonicity approach has the advantage of giving not only unique continuation but also precise asymptotics of solutions near a fixed point, via a suitable combination of monotonicity methods with blow-up analysis, as done in [9,10,11,12,13]. The method based on doubling properties and Almgren monotonicity formula has also been successfully applied to treat the problem of unique continuation from the boundary in [1,2,9,19,27] under homogeneous Dirichlet conditions and in [26] under homogeneous Neumann conditions. Furthermore, in [9] a sharp asymptotic description of the behaviour of solutions at conical boundary points was given through a fine blow-up analysis. In the present paper, we extend the procedure developed in [9,10,11,12,13] to the case of mixed Dirichlet/Neumann boundary conditions, providing sharp asymptotic estimates for solutions near the Dirichlet-Neumann junction and, as a consequence, unique continuation properties. In addition, comparing our result with the aforementioned papers, here we also provide an estimate of the remainder term in the difference between the solution and its asymptotic profile.
Let Ω be an open subset of R2 with Lipschitz boundary. Let Γn⊂∂Ω and Γd⊂∂Ω be two nonconstant curves (open in ∂Ω) such that ¯Γn∩¯Γd={P} for some P∈∂Ω. We are interested in regularity of weak solutions u∈H1(Ω) to the mixed boundary value problem
{−Δu=f(x)u,in Ω,∂νu=g(x)u,on Γn,u=0,on Γd, | (1.1) |
with f∈L∞(Ω) and g∈C1(¯Γn), see Section 2 for the weak formulation. Our aim is to prove unique continuation properties from the Dirichlet-Neumann junction {P}=¯Γn∩¯Γd and sharp asymptotics of nontrivial solutions near P provided ∂Ω is of class C2,γ in a neighborhood of P. We mention that some regularity results for solutions to second-order elliptic problems with mixed Dirichlet-Neumann type boundary conditions were obtained in [16,25], see also the references therein.
Some interest in the derivation of asymptotic expansions for solutions to planar mixed boundary value problems at Dirichlet-Neumann junctions arises in the study of crack problems, see e.g. [6,20]. Indeed, if we consider an elliptic equation in a planar domain with a crack and prescribe Neumann conditions on the crack and Dirichlet conditions on the rest of the boundary, in the case of the crack end-point belonging to the boundary of the domain we are led to consider a problem of the type described above in a neighborhood of the crack's tip (which corresponds to the Dirichlet-Neumann junction). We recall (see e.g. [6]) that, in crack problems, the coefficients of the asymptotic expansion of solutions near the crack's tip are related to the so called stress intensity factor.
In order to get a precise asymptotic expansion of u at point P∈¯Γn∩¯Γd, we will need to assume that ∂Ω is of class C2,δ near P. The asymptotic profile of the solution will be given by the function
Fk(rcosθ,rsinθ)=r2k−12cos(2k−12θ),r>0, θ∈(0,π), | (1.2) |
for some k∈N∖{0}. We note that Fk∈H1loc(R2) and solves the equation
{ΔFk=0,in R2+,Fk(x1,0)=0, for x1<0,∂x2Fk(x1,0)=0, for x1>0, | (1.3) |
where here and in the following R2+:={(x1,x2)∈R2:x2>0}.
The main result of the present paper provides an evaluation of the local behavior of weak solutions u∈H1(Ω) to (1.1) at the boundary point where the boundary conditions change. In order to simplify the statement and without losing generality, we can fix the cartesian axes in such a way that the following assumptions on Ω⊂R2 are satisfied. Here and in the remaining of this paper, Γn,Γd⊂∂Ω are nonconstant curves (open as subsets of ∂Ω) such that ¯Γn∩¯Γd={0} with 0∈∂Ω.
(ⅰ) The domain Ω is of class C2,δ in a neighborhood of 0, for some δ>0.
(ⅱ) The unit vector e1:=(1,0) is tangent to ∂Ω at 0 and pointed towards Γn. Moreover, the exterior unit normal vector to ∂Ω at 0 is (0,−1).
We are now in position to state the main result of the present paper.
Theorem 1.1. We assume that Ω satisfies the assumptions (ⅰ)-(ⅱ) above. Let u∈H1(Ω) be a nontrivial weak solution to (1.1), with f∈L∞(Ω) and g∈C1(¯Γn). Then, there exist k0∈N∖{0}, β∈R∖{0} and r>0 such that, for every ϱ∈(0,1/2), there exists C>0 such that
|u(x)−βFk0(φ(x))|≤C|x|2k0−12+ϱ,foreveryx∈Ω∩¯B+r. | (1.4) |
Here, the function φ:¯Ω∩Br0→¯R2+ is a conformal map of class C2, for some r0>0 only depending on Ω.
Remark 1.2. Here and in the sequel, we identify R2 with the complex plane C; hence, by a conformal map on an open set U⊂R2 we mean a holomorphic function with complex derivative everywhere non-zero on U. We notice that, if Ω satisfies (ⅰ)-(ⅱ) and φ:¯Ω∩Br0→¯R2+ is conformal, then Dφ(0)=αId and φ′(0)=α for some real α>0, where Dφ denotes the jacobian matrix of φ and φ′ denotes the complex derivative of φ.
As a direct consequence of Theorem 1.1, we derive the following Hopf-type lemma.
Corollary 1.3. Under the same assumptions as in Theorem 1.1, let u∈H1(Ω) be a non-trivial weak solution to (1.1), with u≥0. Then
(ⅰ) for every t∈[0,π),
limr→0u(rcost,rsint)r1/2=βα1/2cos(t2)>0, |
where α=φ′(0)>0 and φ is as in Theorem 1.1;
(ⅱ) for every cone C⊂R2 satisfying (1,0)∈C and (−1,0)∈R2∖¯C, we have
lim infx→0x∈¯Ω∩Cu(x)|x|1/2>0. |
A further relevant byproduct of our asymptotic analysis is the following unique continuation principle, whose proof follows directly from Theorem 1.1.
Corollary 1.4. Under the same assumptions as in Theorem 1.1, let u∈H1(Ω) be a weak solution to (1.1) such that u(x)=O(|x|n) as x∈Ω, |x|→0, for any n∈N. Then u≡0.
We observe that Theorem 1.1 provides a sharp asymptotic expansion (and consequently a unique continuation principle) at the boundary for 12-fractional elliptic equations in dimension 1. Indeed, if v∈H1/2(R) weakly solves
{(−Δ)1/2v=g(x)v,in (0,R),v=0,in R∖(0,R), |
for some g∈C1([0,R]), then its harmonic extension V∈H1loc(¯R2+) weakly solves
{−ΔV=0,in R2+,∂νV=g(x)V,on (0,R)×{0},V=0,on (R∖(0,R))×{0}, | (1.5) |
see [4]. Theorem 1.1 and Corollary 1.4 apply to (1.5). Hence, V (and in particular its restriction v) satisfies expansion (1.4) and a strong unique continuation principle from 0 (i.e. from a boundary point of the domain of v). We mention that unique continuation principles from interior points for fractional elliptic equations were established in [8].
We do not know if the C2,δ regularity on Ω and C1 regularity of the boundary potential g in Theorem 1.1 can be weakened in order to obtain a unique continuation property. On the other hand, we can conclude that a regularity assumption on the boundary is crucial for excluding the presence of logarithms in the asymptotic expansion at the junction. Indeed, in Section 8 we produce an example of a harmonic function on a domain with a C1-boundary which is not of class C2,δ, satisfying null Dirichlet boundary conditions on a portion of the boundary and null Neumann boundary conditions on the other portion, but exhibiting dominant logarithmic terms in its asymptotic expansion.
The proof of Theorem 1.1 combines the use of an Almgren type monotonicity formula, blow-up analysis and sharp regularity estimates. Indeed regularity estimates yield the expansion of u near zero as follows:
‖u−k0∑k=1ak(r)Fk∘φ‖L∞(Br)≤Cr2k0−12+ϱ, | (1.6) |
for every ϱ∈(0,1/2), for some C>0, k0≥1 and where ak(r)=⟨u,Fk∘φ⟩L2(Br)‖Fj∘φ‖2L2(Br). Now, if u is nontrivial, a blow-up analysis combined with Almgren type monotonicity formula allows to depict a k0≥1 for which ak0(r)→β≠0 and ak(r)→0 for every k<k0 as r→0. The proof of (1.6) uses also a blow-up analysis argument inspired by Serra [24], see also [22,23].
Remark 1.5. The extension of our results to higher dimensions are the object of current investigation. First of all, the implementation of the monotonicity argument for Dirichlet-Neumann problems exhibits substantial additional difficulties due to the positive dimension of the junction set and some role played by the geometry of the domain. Moreover, further technical difficulties appear in higher dimension since, in such a situation, we can no more make use of conformal transformations like the ones employed in Section 2 which are based on the Riemann mapping Theorem.
Remark 1.6. For the sake of simplicity of the exposition, in the present paper we considered an elliptic problem with the Laplacian and a linear term with a bounded potential; a possible extension to more general elliptic problems with variable coefficients and first order terms could be obtained with a more sophisticated monotonicity approach like in [9].
The paper is organized as follows. In Section 2 we introduce an auxiliary equivalent problem obtained by a conformal diffeomorphic deformation straightening B1∩∂Ω near 0 and state Theorem 2.1 giving the sharp asymptotic behaviour of its solutions. Section 3 contains some Hardy-Poincaré type inequalities for H1-functions vanishing on a portion of the boundary of half-balls. In Section 4 we develop an Almgren type monotonicity formula for the auxiliary problem which yields good energy estimates for rescaled solutions thus allowing the fine blow-up analysis performed in Section 5 and hence the proof of Theorem 2.1. Section 7 contains the proof of the main Theorem 1.1, which is based on Theorem 2.1 and on some regularity and approximation results established in Section 6. Finally, Section 8 is devoted to the construction of an example of a solution with logarithmic dominant term in a domain violating the C2,δ-regularity assumption.
For every R>0 let BR={(x1,x2)∈R2:x21+x22<R2} and B+R={(x1,x2)∈BR:x2>0}. Since ∂Ω is of class C2,δ near zero, we can find r0>0 such that Γ:=∂Ω∩Br0 is a C2,δ curve. Here and in the following, we let B be a C2,δ simply connected open bounded set such that B⊂Ω and ∂B∩∂Ω=Γ. For some functions
f∈L∞(B)andg∈C1(¯Γn), | (2.1) |
let u∈H1(B) be a solution to
{−Δu=f(x)u,in B,∂νu=g(x)u,on Γn,u=0,on Γd. | (2.2) |
We introduce the space H10,Γd(B) as the closure in H1(B) of the subspace
C∞0,Γd(B):={u∈C∞(¯B):u=0 on Γd∩∂B}. |
We say that u∈H1(B) is a weak solution to (2.2) if
{u∈H10,Γd(B),∫B∇u(x)∇v(x)dx=∫Bf(x)u(x)v(x)dx+∫Γnguvdsfor any v∈C∞0,∂B∖Γn(B) |
where C∞0,∂B∖Γn(B)={u∈C∞(¯B):u=0 on ∂B∖Γn}. Since B is of class C2,δ, in view of the Riemann mapping Theorem and [17,Theorem 5.2.4], there exists a conformal map ˆφ:¯B→¯B1 which is of class C2. Let N=ˆφ(0)∈∂B1 and let S be its antipodal. We then consider the map ˜φ:R2∖{S}→R2∖{¯S} given by ˜φ(z):=2¯z−S|z−S|2+¯S, where, for every z∈R2≃C, ¯z denotes the complex conjugate of z. This map is conformal and ˜φ(N)=0. In addition ˜φ(¯B1∖{S})⊂¯P where P is the half plane not containing ¯S whose boundary is the line passing through the origin orthogonal to ¯S.
Then the map ˜φ∘ˆφ is a conformal map which is of class C2 from a neighborhood of the origin ¯B∩Br into ¯P for some r>0. It is now clear that there exists a rotation R and a real number R>0 such that, letting UR:=φ−1(B+R), the map φ:=R∘˜φ∘ˆφ:¯UR→¯B+R is an invertible conformal map of class C2 with inverse φ−1:¯B+R→¯UR of class C2. Moreover φ(0)=0.
Since φ is a conformal diffeomorphism, in view of Remark1.2 we have that, under the assumptions of Theorem 1.1,
Dφ(0)=αId,with α=φ′(0)>0, | (2.3) |
being φ′(0) the complex derivative of φ at 0, which turns out to be real because of the assumption that (1,0) is tangent to ∂Ω at 0 and strictly positive because of the assumption that the exterior unit normal vector to ∂Ω at 0 is (0,−1). In addition, (2.3) implies that, if R is chosen sufficiently small, φ−1((−R,0)×{0})⊂Γd and φ−1((0,R)×{0})⊂Γn.
Therefore letting w=u∘φ−1:B+R→R and Ψ:=φ−1, we then have that w∈H1(B+R) solves
{−Δw(z)=p(z)w(z),in B+R,∂νw(x1,0)=q(x1)w(x1,0),x1∈(0,R),w=0,on (−R,0)×{0}, | (2.4) |
with
p(z)=|Ψ′(z)|2f(Ψ(z)),q(x1)=(g(Ψ(x1,0))|Ψ′(x1,0)|. |
It is plain that p∈L∞(B+R) and q∈C1([0,R)). Here and in the following, for every r>0, we define
Γrn:=(0,r)×{0} and Γrd:=(−r,0)×{0}. | (2.5) |
The following theorem describes the behaviour of w at 0 in terms of the limit of the Almgren quotient associated to w, which is defined as
N(r)=∫B+r|∇w|2dz−∫B+rpw2dz−∫r0q(x)w2(x,0)dx∫π0w2(rcost,rsint)dt. |
In Section 4 we will prove that N is well defined in the interval (0,R0) for some R0>0.
Theorem 2.1. Let w be a nontrivial solution to (2.4). Then there exists k0∈N, k0≥1, such that
limr→0+N(r)=2k0−12. | (2.6) |
Furthermore
τ−2k0−12w(τz)→β|z|2k0−12cos(2k0−12Argz)as τ→0+ |
strongly in H1(B+r) for all r>0 and in C0,μloc(¯R2+∖{0}) for every μ∈(0,1), where β≠0 and
β=2π∫π0R−2k0−12w(Rcoss,Rsins)cos(2k0−12s)ds+2π∫π0[∫R0t−k0+3/2−R1−2k0tk0+1/22k0−1p(tcoss,tsins)w(tcoss,tsins)dt]cos(2k0−12s)ds+2π∫R0t1/2−k0−R1−2k0tk0−1/22k0−1q(t)w(t,0)dt. | (2.7) |
In particular
τ−2k0−12w(τcost,τsint)→βcos(2k0−12t)in C0,μ([0,π])as τ→0+. | (2.8) |
The proof of Theorem 2.1 is based on the study of the monotonicity properties of the Almgren function N and on a fine blow-up analysis which will be performed in Sections 4 and 5.
In the description of the asymptotic behavior at the Dirichlet-Neumann junction of solutions to equation (2.4) a crucial role is played by eigenvalues and eigenfunctions of the angular component of the principal part of the operator.
Let us consider the eigenvalue problem
{−ψ"=λψ,in [0,π],ψ′(0)=0,ψ(π)=0. | (3.1) |
It is easy to verify that (3.1) admits the sequence of (all simple) eigenvalues
λk=14(2k−1)2,k∈N, k≥1, |
with corresponding eigenfunctions
ψk(t)=cos(2k−12t),k∈N, k≥1. |
It is well known that the normalized eigenfunctions
{√2πcos(2k−12t)}k≥1 | (3.2) |
form an orthonormal basis of the space L2(0,π). Furthermore, the first eigenvalue λ1=14 can be characterized as
λ1=14=minψ∈H1(0,π)∖{0}ψ(π)=0∫π0|ψ′(t)|2dt∫π0|ψ(t)|2dt. | (3.3) |
For every r>0, we let (recall (2.5) for the definition of Γrd)
Hr={w∈H1(B+r):w=0 onΓrd}. |
As a consequence of (3.3) we obtain the following Hardy-Poincaré inequality in Hr.
Lemma 3.1. For every r>0 and w∈Hr, we have that
∫B+r|∇w(z)|2dz≥14∫B+r|w(z)|2|z|2dz. |
Proof. Let w∈C∞(¯B+r) with w=0 on ¯Γrd=[−r,0]×{0}. Then, in view of (3.3),
∫B+r|∇w(z)|2dz=∫r0∫π0ρ(|∂∂ρ(w(ρcost,ρsint))|2+1ρ2|∂∂t(w(ρcost,ρsint))|2)dtdρ≥∫r01ρ(∫π0|∂∂t(w(ρcost,ρsint))|2dt)dρ≥14∫r01ρ(∫π0|w(ρcost,ρsint)|2dt)dρ=14∫B+r|w(z)|2|z|2dz. |
We conclude by density, recalling that the space of smooth functions vanishing on [−r,0]×{0} is dense in Hr, see e.g. [7].
Lemma 3.2. For every r>0 and w∈Hr, we have that x−11w2(x1,0)∈L1(0,r) and
∫r0w2(x1,0)x1dx1≤π∫B+r|∇w(z)|2dz. |
Proof. Let w∈C∞(¯B+r) with w=0 on [−r,0]×{0}. Then for any 0<x1<r
|w(x1,0)|=|∫π0ddtw(x1cost,x1sint)dt|=|∫π0x1∇w(x1cost,x1sint)⋅(−sint,cost)dt|≤√π√∫π0x21|∇w(x1cost,x1sint)|2dt. |
It follows that
∫r0w2(x1,0)x1dx1≤π∫r0∫π0x1|∇w(x1cost,x1sint)|2dtdx1=π∫B+r|∇w(z)|2dz. |
We conclude by density.
Let w∈H1(B+R) be a non trivial solution to (2.4). For every r∈(0,R] we define
D(r)=∫B+r|∇w|2dz−∫B+rpw2dz−∫r0q(x1)w2(x1,0)dx1 | (4.1) |
and
H(r)=1r∫S+rw2ds=∫π0w2(rcost,rsint)dt, | (4.2) |
where S+r:={(x1,x2):x21+x22=r2 and x2>0}.
In order to differentiate the functions D and H, the following Pohozaev type identity is needed.
Theorem 4.1. Let w solve (2.4). Then for a.e. r∈(0,R) we have
r2∫S+r|∇w|2ds=r∫S+r|∂w∂ν|2ds−12∫r0(q(x1)+x1q′(x1))w2(x1,0)dx1+r2q(r)w2(r,0)+∫B+rpwz⋅∇wdz | (4.3) |
and
∫B+r|∇w|2dz=∫B+rpw2dz+∫S+r∂w∂νwds+∫r0q(x1)w2(x1,0)dx1. | (4.4) |
Proof. We observe that, by elliptic regularity theory, w∈H2(B+r∖B+ε) for all 0<ε<r<R. Furthermore, the fact that w has null trace on ΓRd implies that ∂w∂x1 has null trace on ΓRd. Then, testing (2.4) with z⋅∇w and integrating over B+r∖B+ε, we obtain that
r2∫S+r|∇w|2ds−ε2∫S+ε|∇w|2ds=∫B+r∖B+εpwz⋅∇wdz+r∫S+r|∂w∂ν|2ds−ε∫S+ε|∂w∂ν|2ds+∫rεq(x1)w(x1,0)x1∂w∂x1(x1,0)dx1. | (4.5) |
An integration by parts, which can be easily justified by an approximation argument, yields that
∫rεq(x1)w(x1,0)x1∂w∂x1(x1,0)dx1=r2q(r)w2(r,0)−ε2q(ε)w2(ε,0)−12∫rε(q+x1q′)w2(x1,0)dx1. | (4.6) |
We observe that there exists a sequence εn→0+ such that
limn→∞[εnw2(εn,0)+εn∫S+εn|∇w|2ds]=0. |
Indeed, if no such sequence exists, there would exist ε0>0 such that
w2(r,0)+∫S+r|∇w|2ds≥Crfor all r∈(0,ε0),for some C>0; |
integration of the above inequality on (0,ε0) would then contradict the fact that w∈H1(B+R) and, by trace embedding, w∈L2(Γε0n). Then, passing to the limit in (4.5) and (4.6) with ε=εn yields (4.3). Finally (4.4) follows by testing (2.4) with w and integrating by parts in B+r.
In the following lemma we compute the derivative of the function H.
Lemma 4.2. H∈W1,1loc(0,R) and
H′(r)=2∫π0w(rcost,rsint)∂w∂ν(rcost,rsint)dt=2r∫S+rw∂w∂νds, | (4.7) |
in a distributional sense and for a.e. r∈(0,R), and
H′(r)=2rD(r),fora.e.r∈(0,R). | (4.8) |
Proof. Let ϕ∈C∞c(0,R). Since w,∇w∈L2(B+R) and w∈C1(B+R), using twice Fubini's Theorem we obtain that
∫R0H(r)ϕ′(r)dr=∫R0(∫π0w2(rcost,rsint)dt)ϕ′(r)dr=∫π0(∫R0w2(rcost,rsint)ϕ′(r)dr)dt=−∫π0(∫R0ddr(w2(rcost,rsint))ϕ(r)dr)dt=−∫π0(∫R0(2w(rcost,rsint)∂w∂ν(rcost,rsint))ϕ(r)dr)dt=−∫R0(∫π0(2w(rcost,rsint)∂w∂ν(rcost,rsint))dt)ϕ(r)dr |
thus proving (4.7). Identity (4.8) follows directly from (4.7) and (4.4).
Let us now study the regularity of the function D.
Lemma 4.3. The function D defined in (4.1) belongs to W1,1(0,R) and
D′(r)=2∫S+r|∂w∂ν|2ds−1r∫r0(q(x1)+x1q′(x1))w2(x1,0)dx1+2r∫B+rpwz⋅∇wdz−∫S+rpw2ds | (4.9) |
in a distributional sense and for a.e. r∈(0,R).
Proof. From the fact that w∈H1(B+R) and w|ΓRn∈L2(ΓRn), we deduce that D belongs to W1,1(0,R) and
D′(r)=∫S+r|∇w|2ds−∫S+rpw2ds−q(r)w2(r,0) | (4.10) |
for a.e. r∈(0,R) and in the distributional sense.
The conclusion follows combining (4.10) and (4.3).
Lemma 4.4. There exists R0∈(0,R) such that H(r)>0 for any r∈(0,R0).
Proof. Let R0∈(0,R) be such that
4‖p‖L∞(B+R)R20+π‖q‖L∞(ΓRn)R0<1. | (4.11) |
Assume by contradiction that there exists r0∈(0,R0) such that H(r0)=0, so that w=0 a.e. on S+r0. From (4.4) it follows that
∫B+r0|∇w|2dz−∫B+r0pw2dz−∫r00q(x1)w2(x1,0)dx1=0. |
From Lemmas 3.1 and 3.2, we get
0=∫B+r0|∇w|2dz−∫B+r0pw2dz−∫r00q(x1)w2(x1,0)dx1≥[1−4‖p‖L∞(B+R)r20−π‖q‖L∞(ΓRn)r0]∫B+r0|∇w|2dz, |
which, together with (4.11) and Lemma 3.1, implies w≡0 in B+r0. From classical unique continuation principles for second order elliptic equations with locally bounded coefficients (see e.g. [28]) we can conclude that w=0 a.e. in B+R, a contradiction.
Thanks to Lemma 4.4, the frequency function
N:(0,R0)→R,N(r)=D(r)H(r), | (4.12) |
is well defined. Using Lemmas 4.2 and 4.3, we now compute the derivative of N.
Lemma 4.5. The function N defined in (4.12) belongs to W1,1loc(0,R0) and
N′(r)=ν1(r)+ν2(r) | (4.13) |
in a distributional sense and for a.e. r∈(0,R0), where
ν1(r)=2r[(∫S+r|∂w∂ν|2ds)⋅(∫S+rw2ds)−(∫S+rw∂w∂νds)2](∫S+rw2ds)2 | (4.14) |
and
ν2(r)=−∫r0(q(x)+xq′(x))w2(x,0)dx∫S+rw2ds+2∫B+rpwz⋅∇wdz∫S+rw2ds−r∫S+rpw2ds∫S+rw2ds. | (4.15) |
Proof. From Lemmas 4.2, 4.4, and 4.3, it follows that N∈W1,1loc(0,R0). From (4.8) we deduce that
N′(r)=D′(r)H(r)−D(r)H′(r)(H(r))2=D′(r)H(r)−12r(H′(r))2(H(r))2 |
and the proof of the lemma easily follows from (4.7) and (4.9).
We now prove that N(r) admits a finite limit as r→0+.
Lemma 4.6. There exists γ∈[0,+∞) such that limr→0+N(r)=γ.
Proof. From Lemmas 3.1 and 3.2 it follows that
D(r)≥[1−4‖p‖L∞(B+R)r2−π‖q‖L∞(ΓRn)r]∫B+r|∇w|2dz, |
hence there exist ˉr∈(0,R0) and C1>0 such that
D(r)≥C1∫B+r|∇w|2dz,for all r∈(0,ˉr). |
In particular
N(r)≥0,for all r∈(0,ˉr). | (4.16) |
Moreover, using again Lemmas 3.1 and 3.2 we can estimate ν2 in (0,ˉr) as follows
|ν2(r)|≤‖q+xq′‖L∞(ΓRn)πr∫B+r|∇w|2dz∫S+rw2ds+‖p‖L∞(B+R)r(1+4r2)∫B+r|∇w(z)|2dz∫S+rw2ds+r‖p‖L∞(B+R)≤1C1(‖q+xq′‖L∞(ΓRn)π+‖p‖L∞(B+R)(1+4ˉr2))N(r)+ˉr‖p‖L∞(B+R). | (4.17) |
Since ν1≥0 by Schwarz's inequality, from Lemma 4.5 and the above estimate it follows that there exists C2>0 such that
N′(r)≥−C2(N(r)+1)for all r∈(0,ˉr), | (4.18) |
which implies that
ddr(eC2r(1+N(r)))≥0. |
It follows that the limit of r↦eC2r(1+N(r)) as r→0+ exists and is finite; hence the function N has a finite limit γ as r→0+. From (4.16) we deduce that γ≥0.
The function H defined in (4.2) can be estimated as follows.
Lemma 4.7. Let γ:=limr→0+N(r) be as in Lemma 4.6. Then
H(r)=O(r2γ)asr→0+. | (4.19) |
Moreover, for any σ>0,
r2γ+σ=O(H(r))asr→0+. | (4.20) |
Proof. From Lemma 4.6 we have that
Nis bounded in a neighborhood of0, | (4.21) |
hence from (4.18) it follows that N′≥−C3 for some positive constant C3 in a neighborhood of 0. Then
N(r)−γ=∫r0N′(ρ)dρ≥−C3r | (4.22) |
in a neighborhood of 0. From (4.8), (4.12), and (4.22) we deduce that, in a neighborhood of 0,
H′(r)H(r)=2N(r)r≥2γr−2C3, |
which, after integration, yields (4.19).
Since γ=limr→0+N(r), for any σ>0 there exists rσ>0 such that N(r)<γ+σ/2 for any r∈(0,rσ) and hence H′(r)H(r)=2N(r)r<2γ+σr for all r∈(0,rσ). By integration we obtain (4.20).
Lemma 5.1. Let w∈H1(B+R) be a non trivial solution to (2.4). Let γ:=limr→0+N(r) be as in Lemma 4.6. Then there exists k0∈N, k0≥1, such that
γ=2k0−12. |
Furthermore, for every sequence τn→0+, there exist a subsequence {τnk}k∈N such that
w(τnkz)√H(τnk)→˜w(z) | (5.1) |
strongly in H1(B+r) and in C0,μloc(¯B+r∖{0}) for every μ∈(0,1) and all r∈(0,1), where
˜w(rcost,rsint)=±√2πr2k0−12cos(2k0−12t),forallr∈(0,1)andt∈[0,π]. | (5.2) |
Proof. Let us set
wτ(z)=w(τz)√H(τ). | (5.3) |
We notice that, for all τ∈(0,R), wτ∈H1 and ∫S+1|wτ|2ds=∫π0|wτ(cost,sint)|2dt=1. Moreover, by scaling and (4.21),
∫B+1(|∇wτ(z)|2−τ2p(τz)|wτ(z)|2)dz−τ∫10q(τx)|wτ(x,0)|2dx=N(τ)=O(1) | (5.4) |
as τ→0+, whereas from Lemmas 3.1 and 3.2 it follows that
N(τ)≥1H(τ)[1−4‖p‖L∞(B+R)τ2−π‖q‖L∞(ΓRn)τ]∫B+τ|∇w|2dz=[1−4‖p‖L∞(B+R)τ2−π‖q‖L∞(ΓRn)τ]∫B+1|∇wτ|2dz | (5.5) |
for every τ∈(0,R0), being R0 as in (4.11). From (5.4), (5.5), and Lemma 3.1 we deduce that
{wτ}τ∈(0,R0)is bounded in H1(B+1). | (5.6) |
Therefore, for any given sequence τn→0+, there exists a subsequence τnk→0+ such that wτnk⇀˜w weakly in H1(B+1) for some ˜w∈H1(B+1). Due to compactness of trace embeddings, we have that ˜w=0 on Γ1d and
∫S+1|˜w|2ds=1. | (5.7) |
In particular ˜w≢. For every small \tau\in (0, R_0), w^\tau satisfies
\begin{equation}\label{eqlam} \begin{cases} -\Delta w^\tau = \tau^2p(\tau z)w^\tau , &\text{in }B_1^+, \\ \partial_\nu w^\tau = \tau q(\tau x_1, 0) w^\tau , &\text{on }\Gamma ^1_n, \\ w^\tau = 0, &\text{on }\Gamma ^1_d, \end{cases} \end{equation} | (5.8) |
in a weak sense, i.e.
\int_{B_1^+} \nabla w^\tau(z)\cdot\nabla \varphi(z)\, dz = \tau^2 \int_{B_1^+} p(\tau z)w^\tau(z)\varphi(z)\, dz+ \tau\int_0^1 q(\tau x)w^\tau(x, 0)\varphi(x, 0)\, dx |
for all \varphi \in H^1(B_1^+) s.t. \varphi = 0 on S_1^+\cup\Gamma _d^1. From weak convergence w^{\tau _{n_k}}\rightharpoonup \widetilde w in H^1(B_1^+), we can pass to the limit in (5.8) along the sequence \tau _{n_k} and obtain that \widetilde w weakly solves
\begin{equation}\label{eq:extended_limit} \begin{cases} -\Delta \widetilde w = 0, &\text{in }B_1^+, \\ \partial_\nu \widetilde w = 0, &\text{on }\Gamma ^1_n, \\ \widetilde w = 0, &\text{on }\Gamma ^1_d. \end{cases} \end{equation} | (5.9) |
From (5.6) it follows that \{\tau q(\tau x) w^\tau(x, 0)\}_{\tau\in(0, R_0)} is bounded in H^{1/2}(\Gamma ^1_n). Then, by elliptic regularity theory, for every 0 < r_1 < r_2 < 1 we have that \{w^\tau\}_{\tau\in(0, R_0)} is bounded in H^2(B^+_{r_2}\setminus \overline{B^+_{r_2}}). From compactness of trace embeddings we have that, up to passing to a further subsequence, \frac{\partial w^{\tau_{n_k}}}{\partial\nu}\to \frac{\partial \widetilde w}{\partial\nu} in L^2(S_r^+) for every r\in (0, 1). Testing equation (5.8) for \tau = \tau_{n_k} with w^{\tau} on B_r^+ we obtain that
\begin{align*} \int_{B_r^+} |\nabla w^{\tau_{n_k}}(z)|^2\, dz& = \int_{S_r^+}\frac{\partial w^{\tau_{n_k}}}{\partial\nu}w^{\tau_{n_k}}\, ds\\ &\qquad+ \tau_{n_k}^2 \int_{B_r^+} p(\tau_{n_k} z)|w^{\tau_{n_k}}(z)|^2\, dz+ \tau_{n_k}\int_0^r q(\tau_{n_k} x)|w^{\tau_{n_k}}(x, 0)|^2\, dx\\ &\mathop{\rightarrow}\limits_{k\to+\infty}\int_{S_r^+}\frac{\partial \widetilde w}{\partial\nu}\widetilde w\, ds = \int_{B_r^+} |\nabla \widetilde w (z)|^2\, dz, \end{align*} |
thus proving that \|w^{\tau _{n_k}}\|_{H^1(B_r^+)}\to \|\widetilde w\|_{H^1(B_r^+)} for all r\in(0, 1), and hence
w^{\tau _{n_k}}\to \widetilde w\quad\text{in }H^1(B_r^+) | (5.10) |
for every r\in (0, 1). Furthermore, by compact Sobolev embeddings, we also have that, up to extracting a further subsequence,
w^{\tau _{n_k}}\to \widetilde w \quad\text{in }C^{0, \mu}_{\rm loc}(\overline{B_r^+}\setminus\{0\}), |
for every r\in (0, 1) and \mu\in(0, 1).
For any r\in (0, 1) and k\in \mathbb{N}, let us define the functions
\begin{align*} &D_k(r) = \int_{B_r^+} |\nabla w^{\tau_{n_k}}|^2\, dz- \tau_{n_k}^2\int_{B_r^+} p(\tau_{n_k} z)|w^{\tau_{n_k}}(z)|^2dz -\tau_{n_k}\int_0^r q(\tau_{n_k} x)|w^{\tau_{n_k}} (x, 0)|^2\, dx, \\ &H_k(r) = \frac{1}{r}\int_{S_r^+}|w^{\tau_{n_k}}|^2 \, ds, \end{align*} |
and {\mathcal N}_k(r): = \frac{D_k(r)}{H_k(r)}. Direct calculations yield that {\mathcal N}_k(r) = {\mathcal N}(\tau_{n_k}r) for all r\in (0, 1). From (5.10) it follows that, for any fixed r\in (0, 1),
D_k(r)\to \widetilde D(r): = \int_{B_r^+} |\nabla \widetilde w|^2\, dz\quad\text{and}\quad H_k(r)\to \widetilde D(r): = \frac{1}{r}\int_{S_r^+}|\widetilde w|^2 \, ds. |
From classical unique continuation principles for harmonic functions it follows that \widetilde D(r)>0 and \widetilde H(r)>0 for all r\in(0, 1) (indeed \widetilde D(r) = 0 or \widetilde H(r) = 0 for some r\in(0, 1) would imply that \widetilde w \equiv 0 in B_r^+ and, by unique continuation, \widetilde w \equiv 0 in B_1^+, a contradiction). Hence, by Lemma 4.6,
\widetilde {\mathcal N}(r) = \frac{\widetilde D(r)}{\widetilde H(r)} = \lim\limits_{k\to \infty}{\mathcal N}_k(r) = \lim\limits_{k\to \infty} {\mathcal N}(\tau_{n_k}r) = \gamma | (5.11) |
for all r\in (0, 1). Therefore \widetilde {\mathcal N} is constant in (0, 1) and hence \widetilde {\mathcal N}'(r) = 0 for any r\in (0, 1). By (5.9) and Lemma 4.5 with p\equiv 0 and q\equiv0, we obtain
\left(\int_{S_r^+} \left|\frac{\partial \widetilde w}{\partial \nu}\right|^2 ds\right) \cdot \left(\int_{S_r^+} \widetilde w^2\, ds\right)-\left( \int_{S_r^+} \widetilde w\frac{\partial \widetilde w}{\partial \nu}\, ds\right)^{\!2} = 0 \quad \text{for all } r\in (0, 1), |
which implies that \widetilde w and \frac{\partial \widetilde w}{\partial \nu} are parallel as vectors in L^2(S_r^+). Hence there exists \eta = \eta(r) such that \frac{\partial \widetilde w}{\partial \nu}(r\cos t, r\sin t) = \eta(r) \widetilde w(r\cos t, r\sin t) for all r\in(0, 1) and t\in[0, \pi]. It follows that
\widetilde w(r\cos t, r\sin t) = \varphi(r) \psi(t), \quad r\in(0, 1), \ t\in [0, \pi], | (5.12) |
where \varphi(r) = e^{\int_1^r \eta(s)ds} and \psi(t) = \widetilde w(\cos t, \sin t). From (5.7) we have that \int_0^\pi \psi^2 = 1. From (5.9) and (5.12) we can conclude that
\begin{cases} \varphi"(r)\psi(t)+ \frac1r\varphi'(r)\psi(t)+\frac1{r^2}\varphi(r)\psi"(t) = 0, &r\in(0, 1), \quad t\in[0, 1], \\ \psi(\pi) = 0, \\ \psi'(0) = 0. \end{cases} |
Taking r fixed, we deduce that \psi is necessarily an eigenfunction of the eigenvalue problem (3.1). Then there exists k_0\in\mathbb{N}\setminus\{0\} such that \psi(t) = \pm\sqrt{\frac2\pi}\cos(\frac{2k_0-1}{2}t) and \varphi(r) solves the equation
\varphi"(r)+\frac{1}r\varphi'-\frac{(2k_0-1)^2}{4r^2}\varphi(r) = 0. |
Hence \varphi(r) is of the form
\varphi(r) = c_1 r^{\frac{2k_0-1}2}+c_2 r^{-\frac{2k_0-1}2} |
for some c_1, c_2\in\mathbb{R}. Since the function r^{-\frac{2k_0-1}2}\psi(t)\notin H^1(B_1^+), we deduce that necessarily c_2 = 0 and \varphi(r) = c_1 r^{\frac{2k_0-1}2}. Moreover, from \varphi(1) = 1, we obtain that c_1 = 1 and then
\widetilde w(r\cos t, r\sin t) = \pm\sqrt{\frac2\pi}\, r^{\frac{2k_0-1}2} \cos\bigg(\frac{2k_0-1}{2}t\bigg) , \quad \text{for all }r\in (0, 1)\text{ and }t\in[0, \pi]. | (5.13) |
From (5.13) it follows that
\widetilde H(r) = \int_0^\pi \widetilde w^2(r\cos t, r\sin t)\, dt = r^{2k_0-1}. |
Hence, in view of (4.8),
\gamma = \widetilde {\mathcal N}(r) = \frac r2 \frac{\widetilde H'(r)}{\widetilde H(r)} = \frac r2(2k_0-1)\frac{r^{2k_0-2}}{r^{2k_0-1}} = \frac{2k_0-1}{2}. |
The proof of the lemma is thereby complete.
We observe that at this stage of our analysis we cannot exclude that the limit function \widetilde w found in Lemma 5.1 depends on the subsequence. In order to prove that the convergence in (5.1) actually holds as \tau\to 0^+ we need to univocally identify the limit profile \widetilde w.
Lemma 5.2. Let w\not\equiv 0 satisfy (2.4), H be defined in (4.2), and \gamma: = \lim_{r\rightarrow 0^+} {\mathcal N}(r) be as in Lemma 4.6. Then the limit \lim_{r\to0^+}r^{-2\gamma}H(r) exists and it is finite.
Proof. In view of (4.19) it is sufficient to prove that the limit exists. By (4.2), (4.8), and Lemma 4.6 we have that
\frac{d}{dr} \frac{H(r)}{r^{2\gamma}} = 2r^{-2\gamma-1} (D(r)-\gamma H(r)) = 2r^{-2\gamma-1} H(r) \int_0^r {\mathcal N}'(\rho) d\rho, |
and then, by integration over (r, R_0),
\frac{H(R_0)}{R_0^{2\gamma}}- \frac{H(r)}{r^{2\gamma}} = 2\int_r^{R_0} \frac{H(\rho)}{\rho^{2\gamma+1}} \left( \int_0^\rho \nu_1(t) dt \right) d\rho +2\int_r^{R_0} \frac{H(\rho)}{\rho^{2\gamma+1}} \left( \int_0^\rho \nu_2(t) dt \right) d\rho | (5.14) |
where \nu_1 and \nu_2 are as in (4.14) and (4.15). Since, by Schwarz's inequality, \nu_1\geq 0, we have that \lim_{r\to 0^+} \int_r^{R_0} \rho^{-2\gamma-1} H(\rho) \left(\int_0^\rho \nu_1(t) dt \right) d\rho exists. On the other hand, from Lemma 4.6 \mathcal N is bounded and hence from (4.17) we deduce that \nu_2 is bounded close to 0^+. Hence, in view of (4.19), the function \rho\mapsto \rho^{-2\gamma-1} H(\rho) \left(\int_0^\rho \nu_2(t) dt \right) is bounded and hence integrable near 0. We conclude that both terms at the right hand side of (5.14) admit a limit as r\to 0^+ thus completing the proof.
The following lemma provides some pointwise estimate for solutions to (2.4).
Lemma 5.3. Let w\in H^1(B_R^+) be a nontrivial solution to (2.4). Then there exist C_4, C_5>0 and \bar r\in(0, R_0) such that
(ⅰ) \sup_{S_{r}^+}|w|^2\leq \frac{C_4}{r}\int_{S_r^+}|w(z)|^2\, ds for every 0 < r < \bar r,
(ⅱ) |w(z)|\leq C_5|z|^\gamma for all z\in B_{\bar r}^+, with \gamma as in Lemma 4.6.
Proof. We first notice that (ⅱ) follows directly from (ⅰ) and (4.19). In order to prove (ⅰ), we argue by contradiction and assume that there exists a sequence \tau_n\to 0^+ such that
\sup\limits_{t\in[0, \pi]}\Big|w\Big(\frac{\tau_n}2\cos t, \frac{\tau_n}2\sin t\Big)\Big|^2 \gt n H \Big (\frac{\tau_n}2 \Big) |
with H as in (4.2), i.e., defining w^\tau as in (5.3)
\sup\limits_{x\in S_{1/2}^+}|w^{\tau_n}(z)|^2 \gt 2n\int_{S_{1/2}^+}|w^{\tau_n}(z)|^2 ds. | (5.15) |
From Lemma 5.1, there exists a subsequence \tau_{n_k} such that w^{\tau_{n_k}}\to\widetilde w in C^{0}(S_{1/2}^+) with \widetilde w being as in (5.2), hence passing to the limit in (5.15) a contradiction arises.
To obtain a sharp asymptotics of H(r) as r\to 0^+, it remains to prove that \lim_{r\to 0^+} r^{-2\gamma} H(r) is strictly positive.
Lemma 5.4. Under the same assumptions as in Lemmas 5.2 and 5.3, we have that
\lim\limits_{r\to0^+}r^{-2\gamma}H(r) \gt 0. |
Proof. From Lemma 5.1 there exists k_0\in\mathbb{N}, k_0\geq 1 such that \gamma = \frac{2k_0-1}{2}. Let us expand w as
w(r\cos t, r\sin t) = \sum\limits_{k = 1}^\infty\varphi_k(r) \cos\left(\tfrac{2k-1}{2}t\right) | (5.16) |
where
\varphi_k(r) = \frac2\pi\int_{0}^\pi w(r\cos t, r\sin t) \cos\left(\tfrac{2k-1}{2}t\right)\, dt. | (5.17) |
The Parseval identity yields
H(r) = \frac\pi2 \sum\limits_{k = 1}^{\infty}\varphi_k^2(r), \quad\text{for all }0 \lt r\leq R. | (5.18) |
From (4.19) and (5.18) it follows that, for all k\geq1,
\varphi_k(r) = O(r^{\gamma})\quad\text{as }r\to0^+. | (5.19) |
Let \eta\in C^\infty_c(0, R). Testing (2.4) with the function \eta(r)\cos\left(\frac{2k-1}2 t\right), by (5.16) we obtain
\begin{align} \notag& \frac \pi 2\int_0^R r\varphi_k'(r)\eta'(r)\, dr+\frac \pi 2\int_0^R \tfrac{(2k-1)^2}4 \frac 1r \varphi_k(r)\eta(r)\, dr = \int_0^R q(r)w(r, 0)\eta(r)\, dr \\ & \label{eq:form-var}\qquad \qquad +\int_0^R r\eta(r)\left(\int_0^\pi p(r\cos t, r\sin t)w(r\cos t, r\sin t)\cos\left(\tfrac{2k-1}2\, t\right)\, dt\right)dr \, . \end{align} | (5.20) |
Integrating by parts in the first in integral on the left hand side of (5.20) and exploiting the fact that \eta\in C^\infty_c(0, R) is an arbitrary test function, we infer
-\varphi_k"(r)-\frac{1}{r}\varphi_k'(r)+ \frac14(2k-1)^2 \frac{\varphi_k(r)}{r^2} = \zeta_k(r), \quad\text{in }(0, R), |
where
\zeta_k(r) = \frac2{\pi r}\, q(r) w(r, 0)+\frac2\pi\int_0^\pi p(r\cos t, r\sin t) w(r\cos t, r\sin t) \cos\left(\tfrac{2k-1}{2}t\right)\, dt. | (5.21) |
Then, by a direct calculation, there exist c_1^k, c_2^k\in\mathbb{R} such that
\varphi_k(r) = r^{\frac{2k-1}{2}} \bigg(c_1^k+\int_r^R\frac{t^{\frac{1-2k}{2}+1}}{2k-1} \zeta_k(t)\, dt\bigg)+ r^{\frac{1-2k}{2}} \bigg(c_2^k+\int_r^R\frac{t^{\frac{2k-1}{2}+1}}{1-2k} \zeta_k(t)\, dt\bigg). | (5.22) |
From Lemma 5.3 it follows that
\zeta_{k_0}(r) = O\left(r^{\frac{2k_0-1}{2}-1}\right)\quad\text{as }r\to 0^+, | (5.23) |
and hence the functions
t\mapsto t^{\frac{1-2k_0}{2}+1} \zeta_{k_0}(t) \quad\text{and}\quad t\mapsto t^{\frac{2k_0-1}{2}+1}\zeta_{k_0}(t) |
belong to L^1(0, R). Hence
r^{\frac{2{k_0}-1}{2}} \bigg(c_1^{k_0}+\int_r^R\frac{t^{\frac{1-2{k_0}}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt\bigg) = o(r^{\frac{1-2{k_0}}{2}})\quad\text{as }r\to0^+, |
and then, by (5.19), there must be
c_2^{k_0} = \int_0^R\frac{t^{\frac{2{k_0}-1}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt. |
From (5.23), we then deduce that
r^{\frac{1-2k_0}{2}} \bigg(c_2^{k_0}+\int_r^R\frac{t^{\frac{2{k_0}-1}{2}+1}}{1-2{k_0}} \zeta_{k_0}(t)\, dt\bigg) = r^{\frac{1-2k_0}{2}} \int_0^r\frac{t^{\frac{2{k_0}-1}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt = O(r^{k_0+\frac12}) | (5.24) |
as r\to0^+. From (5.22) and (5.24), we obtain that
\varphi_{k_0}(r) = r^{\frac{2{k_0}-1}{2}} \bigg(c_1^{k_0}+\int_r^R\frac{t^{\frac{1-2{k_0}}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt +O(r)\bigg)\quad\text{as }r\to0^+. | (5.25) |
Let us assume by contradiction that \lim_{r \to 0^+} r^{-2\gamma}H(r) = 0. Then (5.18) would imply that
\lim\limits_{r\to0^+}r^{-\frac{2k_0-1}{2}}\varphi_{k_0}(r) = 0, |
and hence, in view of (5.25), we would have that
c_1^{k_0}+\int_0^R\frac{t^{\frac{1-2{k_0}}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt = 0, |
which, together with (5.23), implies
r^{\frac{2{k_0}-1}{2}} \bigg(c_1^{k_0}+\int_r^R\frac{t^{\frac{1-2{k_0}}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt \bigg) = r^{\frac{2{k_0}-1}{2}} \int_0^r\frac{t^{\frac{1-2{k_0}}{2}+1}}{1-2{k_0}} \zeta_{k_0}(t)\, dt = O(r^{\frac12+k_0}) | (5.26) |
as r\to0^+. From (5.25) and (5.26), we conclude that \varphi_{k_0}(r) = O(r^{\frac12+k_0}) as r\to0^+, namely,
\sqrt{H(\tau)} \int_0^{\pi}w^\tau(\cos t, \sin t)\cos\left(\tfrac{2k_0-1}{2}t\right)\, dt = O(\tau^{\frac12+k_0})\quad\text{as }\tau\to0^+, |
where w^\tau is defined in (5.3). From (4.20), there exists C>0 such that \sqrt{H(\tau)}\geq C\tau^{\gamma+\frac12} for \tau small, and therefore
\int_0^{\pi}w^\tau(\cos t, \sin t)\cos\left(\tfrac{2k_0-1}{2}t\right)\, dt = O(\tau^{\frac12})\quad\text{as }\tau\to0^+. | (5.27) |
From Lemma 5.1, for every sequence \tau_n\to0^+, there exist a subsequence \{\tau_{n_k}\}_{k\in\mathbb{N}} such that
w^{\tau_{n_k}}(\cos t, \sin t)\to \pm\sqrt{\frac2\pi}\, \cos\bigg(\frac{2k_0-1}{2}t\bigg) \quad\text{in } L^2(0, \pi). | (5.28) |
From (5.27) and (5.28), we infer that
0 = \lim\limits_{k\to+\infty} \int_0^{\pi}w^{\tau_{n_k}} (\cos t, \sin t)\cos\left(\tfrac{2k_0-1}{2}t\right)\, dt = \pm\sqrt{\frac2\pi}\, \int_0^{\pi}\cos^2\left(\tfrac{2k_0-1}{2}t\right)\, dt = \pm\sqrt{\frac\pi2}, |
thus reaching a contradiction.
Proof of Theorem 2.1. Identity (2.6) follows from Lemma 5.1, thus there exists k_0\in \mathbb{N}, k_0\geq 1, such that \gamma = \lim_{r\to 0^+}{\mathcal N}(r) = \frac{2k_0-1}{2}.
Let \{\tau_n\}_{n\in\mathbb{N}}\subset (0, +\infty) be such that \lim_{n\to+\infty}\tau_n = 0. Then, from Lemmas 5.1 and 5.4, scaling and a diagonal argument, there exists a subsequence \{\tau_{n_k}\}_{k\in\mathbb{N}} and \beta\neq0 such that
\frac{w(\tau_{n_k}z)}{\tau_{n_k}^\gamma}\to \beta |z|^{\frac{2k_0-1}2} \cos\bigg(\tfrac{2k_0-1}{2}\mathop{\rm Arg} z\bigg) | (5.29) |
strongly in H^1(B_r^+) for all r>0 and in C^{0, \mu}_{\rm loc}(\overline{\mathbb{R}^2_+}\setminus\{0\}) for every \mu\in (0, 1). In particular
\tau_{n_k}^{-\gamma}w(\tau_{n_k}(\cos t, \sin t))\to \beta \cos\bigg(\frac{2k_0-1}{2}t\bigg) | (5.30) |
in C^{0, \mu}([0, \pi]). To prove that the above converge occurs as \tau\to0^+ and not only along subsequences, we are going to show that \beta depends neither on the sequence \{\tau_n\}_{n\in\mathbb{N}} nor on its subsequence \{\tau_{n_k}\}_{k\in\mathbb{N}}.
Defining \varphi_{k_0} and \zeta_{k_0} as in (5.17) and (5.21), from (5.30) it follows that
\frac{\varphi_{k_0}(\tau_{n_k})}{\tau_{n_k}^{\gamma}} = \frac2\pi\int_{0}^\pi\frac{w(\tau_{n_k}\cos t, \tau_{n_k}\sin t)}{\tau_{n_k}^{\gamma}} \cos\left(\tfrac{2k_0-1}{2}t\right)\, dt\to \frac2\pi\, \beta\int_{0}^\pi\cos^2\left(\tfrac{2k_0-1}{2}t\right)\, dt = \beta | (5.31) |
as k\to+\infty. On the other hand, from (5.22), (5.24), and (5.25) we know that that
\begin{align} \notag \varphi_{k_0}(\tau)& = \tau^{\frac{2k_0-1}{2}} \bigg(c_1^{k_0}+\int_\tau^R\frac{t^{\frac{1-2k_0}{2}+1}}{2k_0-1} \zeta_{k_0}(t)\, dt\bigg)+\tau^{\frac{1-2k_0}{2}} \int_0^\tau\frac{t^{\frac{2{k_0}-1}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt\\ \label{eq:14}& = \tau^{\frac{2{k_0}-1}{2}} \bigg(c_1^{k_0}+\int_\tau^R\frac{t^{\frac{1-2{k_0}}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt +O(\tau)\bigg)\quad\text{as }\tau\to0^+. \end{align} | (5.32) |
Choosing \tau = R in the first line of (5.32), we obtain
c_1^{k_0} = R^{-\frac{2k_0-1}{2}}\varphi_{k_0}(R)-R^{1-2k_0} \int_0^R\frac{t^{\frac{2{k_0}-1}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt . |
Hence, from the second line of (5.32), we obtain that
\tau^{-\gamma}\varphi_{k_0}(\tau)\to R^{-\frac{2k_0-1}{2}}\varphi_{k_0}(R)-R^{1-2k_0} \int_0^R\frac{t^{\frac{2{k_0}-1}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt+\int_0^R\frac{t^{\frac{1-2{k_0}}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt, |
as \tau\to0^+. Then, from (5.31) we deduce that
\beta = R^{-\frac{2k_0-1}{2}}\varphi_{k_0}(R)-R^{1-2k_0} \int_0^R\frac{t^{\frac{2{k_0}-1}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt+\int_0^R\frac{t^{\frac{1-2{k_0}}{2}+1}}{2{k_0}-1} \zeta_{k_0}(t)\, dt. | (5.33) |
In particular \beta depends neither on the sequence \{\tau_n\}_{n\in\mathbb{N}} nor on its subsequence \{\tau_{n_k}\}_{k\in\mathbb{N}}, thus implying that the convergence in (5.29) actually holds as \tau\to 0^+ and proving the theorem. We observe that (2.7) follows by replacing (5.17) and (5.21) into (5.33).
In this section, we prove some regularity and approximation results, which will be used to estimate the Hölder norm of the difference between a solution u to (1.1) and its asymptotic profile \beta F_{k_0}.
Proposition 6.1. Let f\in L^\infty (B_4^+), g\in L^\infty (\Gamma _n^4) and let v\in H^1(B_4)\cap L^\infty(B_4^+) solve
\begin{equation}\label{eq:v-solv-app-1} \begin{cases} -\Delta v = f, &in \;B_4^+, \\ \partial_\nu v = g, &on \;\Gamma ^4_n, \\ v = 0, &on \;\Gamma ^4_d. \end{cases} \end{equation} | (6.1) |
Then, for every \varepsilon >0, there exists a constant C>0 (independent of v, f, and g) such that
\|v\|_{C^{1/2-\varepsilon } (\overline {B_2^+})}\leq C \left( \|f\|_{L^\infty(B_4^+)}+ \|g\|_{L^\infty(\Gamma ^4_n)}+ \|v\|_{L^\infty(B_4^+)} \right) . |
Proof. In the sequel we denote as C>0 a positive constant independent of v, f, and g which may vary from line to line. We consider a C^2 domain \Omega ' such that B_3^+\subset \Omega '\subset B_4^+ and \Gamma ^3_n\cup \Gamma ^3_d\subset \partial \Omega '. We define the functions (obtained uniquely by minimization arguments) v_1 \in H^1 (\Omega ') satisfying
\begin{equation}\label{eq:v1-solv-app} \begin{cases} -\Delta v_1 = f, &\text{in } \Omega ' , \\ \partial_\nu v_1 = 0, &\text{ on $\Gamma ^{{3}}_n$ } , \\ v_1 = 0, & \text{ on $\partial\Omega '\setminus \Gamma ^3_n$ }, \end{cases} \end{equation} | (6.2) |
and \tilde{ v}_2\in H^{1/2}(\mathbb{R}) satisfying
\begin{cases} (-\Delta )^{\frac{1}{2}} \tilde {v}_2 = g, &\text{in } (0, 4) , \\ \tilde {v}_2 = 0, & \text{on}\; \mathbb{R} \setminus (0, 4) . \end{cases} |
Therefore by (fractional) elliptic regularity theory (see e.g. [21,Proposition 1.1]), we deduce that
\|\tilde {v}_2 \|_{C^{1/2}({\mathbb{R}})} \leq C \| g\|_{L^\infty(\Gamma _n^4)}. | (6.3) |
Consider the Poisson kernel P(x_1, x_2) = \frac 1\pi x_2 |x|^{-2} with respect to the half-space \mathbb{R}^2_+, see [4,Section 2.4]. We define
v_2(x_1, x_2) = (P(\cdot, x_2)\star \tilde {v}_2)( x_1) = \frac 1\pi x_2\int_{\mathbb{R}}\frac{\tilde {v}_2(t)}{x_2^2+(x_1-t)^2 }\, dt = \frac 1\pi \int_{\mathbb{R} }\frac{\tilde {v}_2( x_1-r x_2)}{1+r^2 }\, dr |
where with the symbol \star we denoted the convolution product with respect to the first variable. One can check that v_2\in H^1_{\rm loc}(\overline{\mathbb{R}^2_+}) (see for example [3,Subsection 2.1]) and
\begin{cases} -\Delta v_2 = 0, &\text{in } \; \mathbb{R}^2_+ , \\ \partial_\nu v_2 = g, & \text{on}\; \Gamma ^4_n , \\ v_2 = 0, & \text{on}\; \mathbb{R} \setminus (0, 4). \end{cases} | (6.4) |
It is easy to see that
\|v_2\|_{L^\infty(\mathbb{R}^2_+)}\leq C\|\tilde {v}_2\|_{L^\infty(\mathbb{R})}. |
Moreover by (6.3), for x, y\in \overline {\mathbb{R}^2_+} we get
|v_2(x)- v_2(y)|\leq C \| g\|_{L^\infty(\Gamma _n^4)}|x-y|^{1/2}\int_{\mathbb{R} }\frac{\max(1, |r|^{1/2})}{1+r^2 }\, dr\leq C \| g\|_{L^\infty(\Gamma _n^4)}|x-y|^{1/2}. |
It follows that
\|v_2\|_{C^{1/2}(\overline {\mathbb{R}^2_+})} \leq C \| g\|_{L^\infty(\Gamma _n^4)} . | (6.5) |
By [25,Theorem 1] and continuous embeddings of Besov spaces into Hölder spaces, we get
\|v_1\|_{C^{1/2-\varepsilon }(\overline{\Omega '})}^2 \leq C \|v_1\|_{H^1(\Omega ')} \left( \| f\|_{L^\infty(B_4^+)} + \|v_1\|_{H^1(\Omega ')} \right) . |
Multiplying (6.2) by v_1, integrating by parts and using Young's inequality, we get
C \| v_1\|_{L^2(\Omega ')}^2 \leq \|\nabla v_1\|_{L^2(\Omega ')}^2\leq \| v_1\|_{L^2(\Omega ')} \| f\|_{L^2(B_4^+)} \leq \varepsilon \| v_1\|_{L^2(\Omega ')}^2+ C_\varepsilon \| f\|_{L^\infty(B_4^+)}^2, |
where in the first estimate we have used the Poincaré inequality for functions vanishing on a portion of the boundary. We then conclude that
\|v_1\|_{C^{1/2-\varepsilon }(\overline {\Omega '})} \leq C \| f\|_{L^\infty(B_4^+)}. | (6.6) |
Now, thanks to (6.1), (6.2) and (6.4), the function V: = v-(v_1+v_2)\in H^1(\Omega ') solves the equation
\begin{cases} -\Delta V = 0, &\text{in }\Omega ', \\ \partial_\nu V = 0, &\text{on }\Gamma ^3_n, \\ V = 0, &\text{on }\Gamma ^3_d. \end{cases} | (6.7) |
By elliptic regularity theory, we have that
\|V\|_{C^{2}(\overline { B_{5/2}^+\setminus B_1^+} )} \le C \|V\|_{H^1(B_r^+)} | (6.8) |
where r is a fixed radius satisfying \frac 52 < r < 3 and C>0 is independent of V. Let \eta a radial cutoff function compactly supported in B_3 satisfying \eta\equiv 1 in B_r; testing (6.7) with \eta V, we infer that \|V\|_{H^1(B_r^+)}\le C\|V\|_{L^2(\Omega ')} for some constant C>0 independent of V. Hence by (6.8) we obtain
\label{eq:est-for-V} \|V\|_{C^{2}(\overline { B_{5/2}^+\setminus B_1^+} )} \leq C \|V\|_{L^\infty(\Omega ')}. | (6.9) |
Let {\tilde \eta }\in C^\infty_c(B_{5/2}) be a radial function, with {\tilde \eta }\equiv 1 on B_2. Then the function {\tilde V }: = {\tilde \eta } V\in H^1(\mathbb{R}^2_+) solves
\begin{cases} -\Delta {\tilde V } = - V\Delta {\tilde \eta }- 2\nabla V \cdot \nabla {\tilde \eta }, &\text{in }\mathbb{R}^2_+, \\ \partial_\nu {\tilde V }(x_1, 0) = 0, & x_1\in (0, +\infty), \\ {\tilde V }(x_1, 0) = 0, & x_1\in (-\infty, 0) . \end{cases} |
Then by [25,Theorem 1], the arguments above, (6.9), (6.5) and (6.6), we deduce that
\| v-(v_1+v_2) \|_{C^{1/2-\varepsilon }(\overline {B_2^+})} \leq \|{\tilde V } \|_{C^{1/2-\varepsilon }(\overline {\mathbb{R}^2_+})} \leq C \|V\|_{L^\infty(\Omega ')} \leq C \left( \|f\|_{L^\infty(B_4^+)}+ \|g\|_{L^\infty(\Gamma ^4_n)}+ \|v\|_{L^\infty(B_4^+)} \right) . |
This, combined again with (6.5) and (6.6) completes the proof.
Recalling (1.2), for every k\in \mathbb{N} with k\geq1, we consider the finite dimensional linear subspace of L^2(B_r^+), given by
\mathcal{S}_k: = \left\{\sum\limits_{j = 1}^k a_j F_j\, :\, (a_1, \ldots , a_k)\in \mathbb{R}^k \right\}. |
For every r>0, k\geq1, and u\in L^2(B_r^+), we let
F_{k, r}^u: = \textrm{Argmin}_{F\in \mathcal{S}_k}\int_{B_r^+}(u(x)-F(x))^2\, dx |
be the L^2(B_r^+)-projection of u on \mathcal{S}_k, so that
\min\limits_{F\in \mathcal{S}_k}\int_{B_r^+}(u(x)-F(x))^2\, dx = \int_{B_r^+}(u(x)-F_{k, r}^u(x))^2\, dx |
and
\int_{B_r^+}(u(x)-F_{k, r}^u(x))F(x)\, dx = 0, \quad\textrm{for all} \;F\in\mathcal{S}_k. | (6.10) |
Next, we estimate the L^\infty norm of the difference between a solution of a mixed boundary value problem on B_1^+ and its projection on \mathcal{S}_k.
Proposition 6.2. Let u\in H^1(B_1^+)\cap L^\infty(\mathbb{R}^2_+) solve
\begin{cases} -\Delta u = f, &in \;B_1^+, \\ \partial_\nu u = g, &on \;\Gamma ^1_n, \\ u = 0, &on \;\Gamma ^1_d, \end{cases} | (6.11) |
where, for some k\in \mathbb{N}\setminus\{0\} and \overline {C} >0,
|f(x)| \leq \overline {C} | x|^{\max(\gamma _k-\frac{3}{2}, 0)}, \quad for \;every\; x\in B_1^+, \\ |g(x_1)|\leq \overline {C} | x_1|^{\max({\gamma _k-\frac{1}{2}}, 0)}, \quad for \; every\; x_1\in (0, 1), |
and \gamma_k = \frac{2k-1}{2}. Then, for every \alpha \in (0, 1/2), we have that
\sup\limits_{r \gt 0}r^{-\gamma_k-\alpha }\|u-F_{k, r}^u \|_{L^\infty(B_r^+)} \lt \infty. |
Proof. In the sequel, C>0 stands for a positive constant, only depending on \alpha, \overline {C} and k, which may vary from line to line. Assume by contradiction that, there exists \alpha \in(0, 1/2) such that
\sup\limits_{r \gt 0}r^{-\gamma _k-\alpha }\|u-F_{k, r}^u \|_{L^\infty(B_r^+)} = \infty. |
We consider the nonincreasing function
\Theta(r): = \sup\limits_{\overline {r} \gt r} \overline {r} ^{-\gamma _k-\alpha } \| u- F_{k, \overline {r} }^u\|_{L^\infty(B_{\overline {r} }^+)}. |
It is clear from our assumption that
\Theta(r) \nearrow +\infty \qquad\textrm{ as} \;r\to 0. |
Then there exists a sequence r_n \to 0 such that
r^{-\gamma _k-\alpha }_n\| u- F_{k, r_n}^u\|_{L^\infty(B_{r_n}^+)}\geq \frac{\Theta(r_n)}{2}. |
We define
v_n(x): = r^{-\gamma _k-\alpha}_n \frac{ u(r_n x)- F_{k, r_n}^u(r_n x) }{\Theta(r_n)} , |
so that
\| v_n\|_{L^\infty(B_1^+)} \geq \frac{1}{2}. | (6.12) |
Moreover, by a change of variable in (6.10), we get
\int_{B_1^+}v_n(x)F(x)\, dx = 0 \qquad\textrm{ for every} \;F\in \mathcal{S}_k. | (6.13) |
Claim: For R = 2^m and r>0, we have
\frac{1}{r^{\gamma _k+\alpha}\Theta(r)}\| F^u_{k, r R}-F^u_{k, r}\|_{L^\infty(B_{ rR}^+)}\leq C R^{\gamma _k+\alpha}. | (6.14) |
Indeed, by definition, for every \overline {r} > r>0, we have
\| u- F_{k, \overline {r} }^u\|_{L^\infty(B_{\overline {r} }^+)}\leq \overline {r} ^{\gamma _k+\alpha } \Theta(r) |
and thus, using the monotonicity of \Theta, for every x\in B_r^+ we get
| F_{k, 2 r}^u(x) - F_{k, r}^u(x)| \leq \| u- F_{k, 2 r}^u\|_{L^\infty(B_{ 2r}^+)}+ \| u- F_{k, r}^u\|_{L^\infty(B_{ r}^+)}\leq 2^{1+\gamma_k+\alpha }r^{\gamma_k+\alpha } \Theta(r)\leq C r^{\gamma_k+\alpha } \Theta(r). | (6.15) |
Letting F_{k, r}^u = \sum_{j = 1}^k a_j(r)F_j and \gamma_j = \frac{2j-1}{2}, by taking the L^2(B_r^+)-norm in (6.15), we get
| a_j(2r)-a_j(r)| r^{\gamma_j}\leq C r^{\gamma_k+\alpha } \Theta(r) \qquad\textrm{ for every}\; r \gt 0. |
Then
\begin{align*} \frac{1}{r^{\gamma_k+\alpha }\Theta(r)}\| F^u_{k, r 2^m}-F^u_{k, r}\|_{L^\infty(B_{ r 2^m}^+ )}& \leq \frac{1}{r^{\gamma_k+\alpha }\Theta(r)} \sum\limits_{j = 1}^{k} | a_j(r 2^m)-a_j(r)| (r2^m)^{\gamma_j} \\ & \leq \frac{1}{r^{\gamma_k+\alpha }\Theta(r)} \sum\limits_{j = 1}^{k} \sum\limits_{i = 1}^{m} | a_j(2^i r) - a_j(2^{i-1} r)| (r 2^m)^{\gamma_j} \\ &\leq \frac{C}{r^{\gamma_k+\alpha }\Theta(r)} \sum\limits_{j = 1}^{k} \sum\limits_{i = 1}^{m} 2^{\gamma_j m} 2^{(\gamma_k-\gamma_j+\alpha )(i-1)}r^{\gamma_k +\alpha } \Theta(2^{i-1}r)\\ &\leq {C} \sum\limits_{j = 1}^{k} \sum\limits_{i = 1}^{m} 2^{\gamma_j m} 2^{(\gamma_k-\gamma_j+\alpha )(i-1)}\leq {C} \sum\limits_{j = 1}^{k} 2^{\gamma_j m} 2^{(\gamma_k-\gamma_j+\alpha ) m} \\ &\leq C 2^{m(\gamma_k+\alpha )} . \end{align*} |
This proves the claim.
From the definition of \Theta and (6.14), for R = 2^m\geq 1, we have
\begin{align*} \sup\limits_{x\in B_R^+}|v_n(x)|& = \frac{1}{r^{\gamma_k+\alpha }_n\Theta(r_n)}\| u-F^u_{k, r_n}\|_{L^\infty(B_{r_n R}^+)}\\ &\leq \frac{1}{r^{\gamma_k+\alpha }_n\Theta(r_n)}\| u-F^u_{k, r_n R}\|_{L^\infty(B_{r_n R}^+)}+ \frac{1}{r^{\gamma_k+\alpha }_n\Theta(r_n)}\| F^u_{k, r_n R}-F^u_{k, r_n}\|_{L^\infty(B_{ r_n R}^+)}\\ &\leq \frac{ 1}{ r^{\gamma_k+\alpha }_n \Theta(r_n) } (r_nR)^{\gamma_k+\alpha } \Theta( r_n) + CR^{\gamma_k+\alpha }\\ &\leq CR^{\gamma_k+\alpha }. \end{align*} |
Consequently, letting R\geq 1 and m_0\in \mathbb{N} be the smallest integer such that 2^{m_0}\geq R, we obtain that
\sup\limits_{x\in B_R^+}|v_n(x)|\leq \sup\limits_{x\in B_{2^{m_0}}^+}|v_n(x)|\leq C2^{m_0(\gamma_k+\alpha)}\leq C(2R)^{\gamma_k+\alpha}\leq CR^{\gamma_k+\alpha }, | (6.16) |
with C being a positive constant independent of R. Thanks to (1.3) and (6.11), it is plain that
\begin{cases} -\Delta v_n = \frac{r_n^{2-\gamma_k-\alpha }}{\Theta(r_n)} f (r_n \cdot ), &\text{in }B_{{1}/{r_n}}^+, \\[5pt] \partial_\nu v_n = \frac{r_n^{1-\gamma_k-\alpha }}{\Theta(r_n)} g (r_n \cdot ), &\text{on }\Gamma ^{{1}/{r_n}}_n, \\[5pt] v_n = 0, &\text{on }\Gamma ^{{1}/{r_n}}_d. \end{cases} |
By assumption, we have that \frac{r_n^{2-\gamma_k-\alpha }}{\Theta(r_n)} f (r_n x) and \frac{r_n^{1-\gamma_k-\alpha }}{\Theta(r_n)} g (r_n x_1) are bounded in L^\infty(B_M^+) and L^\infty(\Gamma ^M_n) respectively, for every M>0. Hence, by Proposition 6.1 and (6.16), we have that v_n is bounded in C^{\delta }(\overline{B_M^+}) for every M>0 and \delta \in(0, 1/2). Furthermore, it is easy to verify that v_n is bounded in H^1(B_M^+) for every M>0. Then, for every M>0 and \delta \in(0, 1/2), v_n converges in C^{\delta }(\overline{B_M^+}) (and weakly in H^1(B_M^+)) to some v\in C^{\delta }_{\rm loc}(\overline {\mathbb{R}^2_+})\cap H^1_{\rm loc}(\mathbb{R}^2_+) satisfying
\begin{cases} -\Delta v = 0, &\text{in } \mathbb{R}^2_+, \\ \partial_\nu v = 0, &\text{on }\Gamma ^{\infty}_n, \\ v = 0, &\text{on }\Gamma ^{\infty}_d, \end{cases} |
and by (6.16), for every R>1,
\|v\|_{L^\infty(B_R^+)}\leq C R^{\gamma_k+\alpha }. |
By Lemma 6.3 (below), we deduce that necessarily
v\in \mathcal{S}_k. |
This clearly yields a contradiction when passing to the limit in (6.12) and (6.13).
The following Liouville type result was used in the proof of Proposition 6.2.
Lemma 6.3 (Liouville theorem). Let v\in C(\overline {\mathbb{R}^2_+})\cap H^1_{\rm loc}(\mathbb{R}^2_+) satisfy
\begin{cases} -\Delta v = 0, &\text{in } \mathbb{R}^2_+, \\ \partial_\nu v = 0, &\text{on }\Gamma ^{\infty}_n, \\ v = 0, &\text{on }\Gamma ^{\infty}_d, \end{cases} |
and, for some \alpha \in (0, 1/2) and C>0,
\|v\|_{L^\infty(B_R^+)}\leq C\, R^{\gamma_k+\alpha }\quad for \;every\; R \gt 1, | (6.17) |
where \gamma_k = \frac{2k-1}{2}, k\in\mathbb{N}\setminus\{0\}. Then
v\in \mathcal{S}_k. |
Proof. Arguing as in the proof of Lemma 5.4, we expand v in Fourier series with respect to the orthonormal basis of L^2(0, \pi) given in (3.2) as
v(r\cos t, r\sin t) = \sum\limits_{j = 1}^\infty\varphi_j(r) \cos\left(\tfrac{2j-1}{2}t\right) |
where \varphi_j(r) = \frac2\pi\int_{0}^\pi v(r\cos t, r\sin t) \cos\left(\tfrac{2j-1}{2}t\right)\, dt. From assumption (6.17) and the Parseval identity we have that
\frac\pi2 \sum\limits_{j = 1}^{\infty}\varphi_j^2(r) = \int_0^\pi v^2(r\cos t, r\sin t)\, dt\leq\pi C^2 r^{2(\gamma_k+\alpha)}, \quad\text{for all }r \gt 1. |
It follows that
|\varphi_j(r)|\leq {\rm const\, }r^{\gamma_k+\alpha}\quad\text{for all }j\geq1\text{ and }r \gt 1, | (6.18) |
for some {\rm const\, }>0 independent of j and r.
From the equation satisfied by v it follows that the functions \varphi_j satisfy
-\varphi_j"(r)-\frac{1}{r}\varphi_j'(r)+ \frac14(2j-1)^2 \frac{\varphi_j(r)}{r^2} = 0, \quad\text{in }(0, +\infty), |
and then, for all j\geq1, there exist c_1^j, c_2^j\in\mathbb{R} such that
\varphi_j(r) = c_1^j r^{\frac{2j-1}{2}}+ c_2^j r^{\frac{1-2j}{2}}\quad\text{for all }r \gt 0. |
The fact v is continous and v(0) = 0 implies that \varphi_j(r) = o(1) as r\to 0^+. As a consequence we have that c_2^j = 0 for all j\geq1. On the other hand (6.18) implies that c_1^j = 0 for all j>k. Therefore we conclude that
v(r\cos t, r\sin t) = \sum\limits_{j = 1}^k c_1^j r^{\frac{2j-1}{2}} \cos\left(\tfrac{2j-1}{2}t\right) = \sum\limits_{j = 1}^k c_1^j F_j(r\cos t, r\sin t), |
i.e. v\in \mathcal{S}_k.
We are now in position to prove Theorem 1.1.
Proof of Theorem 1.1. Let w = u\circ\varphi^{-1}, with \varphi:\overline{\mathcal U_R}\to \overline{B_R^+} being the conformal map constructed in Section 2. Let \gamma = \frac{2k_0-1}{2}, with k_0 being as in Theorem 2.1. We define (recalling (5.3))
\widetilde w^\tau(z): = \tau^{-\gamma}w(\tau z) = \tau^{-\gamma}\sqrt{H( \tau )} w^\tau (z) . |
From Theorem 2.1 we have that there exists \beta\neq0 such that \widetilde w^\tau\to \beta F_{k_0} in H^1(B_r^+) for all r>0 and in C^{0, \mu}_{\rm loc}(\overline{\mathbb{R}^2_+}\setminus\{0\}) for every \mu\in(0, 1).
Claim 1: We have
w(y) = \beta F_{k_0}(y)+o(|y|^\gamma ) \qquad\textrm{ as}\; |y|\to 0 \;\; \textrm{and }\; y\in {B_R^+}. | (7.1) |
If this does not hold true then there exists a sequence of points y_m\in (B_R^+\cup \Gamma _n^R)\setminus \{0\} and C>0 such that y_m\to 0 and
|y_m|^{-\gamma} | w(y_m)-\beta F_{k_0}(y_m)| = | {\tilde w}^{\tau _m}( z_m)-\beta F_{k_0}(z_m)|\geq C \gt 0, |
where \tau _m = |y_m| and z_m = \frac{y_m}{|y_m|}. If m is large enough, we get a contradiction with (2.8). This proves (7.1) as claimed.
Let \varrho\in(0, 1/2) and let p and q be the functions introduced in (2.8). This proves (7.1), by the fact that p\in L^\infty(B_R^+) and q\in C^1([0, R)), and by Proposition 6.2 applied to w, we have that, for every r\in (0, R),
|w(x)-F_{k_0, r}^w(x)| \leq C r^{\gamma +\varrho}, \quad\textrm{for every}\; x\in B_r^+, | (7.2) |
for some positive constant C>0 independent of r, which could vary from line to line in the sequel.
From (7.1) and (7.2) we deduce that
\sup\limits_{x\in B_r^+} r^{-\gamma}|\beta F_{k_0}(x)-F_{k_0, r}^w(x)|\to 0, \quad\textrm{as}\; r \to 0^+. | (7.3) |
Claim 2: We have
|\beta F_{k_0}(x)-F_{k_0, r}^w(x)| \leq C r^{\gamma +\varrho}, \quad\textrm{for every} \;x\in B_r^+. | (7.4) |
Once this claim is proved, then according to (7.2), we can easily deduce that for any r\in (0, R)
|w(x)- \beta F_{k_0}(x) |\leq |w(x)- F_{k_0, r}^w(x) |+|F_{k_0, r}^w- \beta F_{k_0}(x) | \leq C r^{\gamma +\varrho}, \quad\textrm{for every} \;x\in B_r^+. |
In particular,
|w(x)- \beta F_{k_0}(x) | \leq C |x|^{\gamma +\varrho}, \quad\textrm{for every}\; x\in B_{R}^+ |
which finishes the proof of Theorem 1.1.
Let us now prove Claim 2. Writing F_{k_0, r}^w(x) = \sum_{j = 1}^{k_0}a_j(r) F_j(x) , by (7.3) we have that
| \beta -a_{k_0}(r) |\to 0, \quad\text{as }r\to0^+ . |
Moreover by taking the L^2(B^+_r)-norms in (7.3), we find that
(a_{k_0}(r)-\beta )^2 r^{2\gamma +2}+\sum\limits_{j = 1}^{k_0-1} a_j^2(r)r^{2\gamma_j+2} \leq C r^{2\gamma + 2}, \quad\textrm{for every}\; R \gt r \gt 0, |
with \gamma_j = \frac{2j-1}{2}. This yields, for j = 1, \ldots , k_0-1,
|a_j(r)|\leq C r^{\gamma -\gamma_j}\to 0 \qquad\textrm{ as} \;r\to 0. | (7.6) |
From (7.2), we get, for every x\in B_r^+ and R>r>0,
\bigg| w(x)- \sum\limits_{j = 1}^{k_0}a_j(r) F_j(x)\bigg|\leq C r^{\varrho+\gamma }. |
Hence, for every x\in B_{r/2}^+, we have that
\bigg| \sum\limits_{j = 1}^{k_0}(a_j(r)-a_j(2^{-1}r)) F_j(x) \bigg|\leq | F_{k_0, r}^w(x) - w (x) |+ | F_{k_0, 2^{-1}r}^w(x) - w(x) |\leq C r^{\varrho+\gamma }. |
Taking the L^2(B^+_{r/2})-norms in the previous inequality, we find that, for every r\in (0, R)
\sum\limits_{j = 1}^{k_0}| a_{j}(r) - a_{j}(2^{-1}r)| r^{\gamma_j} \leq C r^{\gamma +\varrho}. |
This implies that
| a_{j}(r) - a_{j}(2^{-1}r)| \leq C r^{\varrho+\gamma -\gamma_j}\qquad\textrm{for all}\; 1\leq j\leq k_0 \textrm{ and }\;r\in(0, R). |
From this, (7.5) and (7.6), we obtain
|\beta - a_{k_0}(r)| r^{-\varrho}+ \sum\limits_{j = 1}^{k_0-1} |a_j(r)| r^{-\varrho-\gamma +\gamma_j}\leq \sum\limits_{j = 1}^{k_0} \sum\limits_{i = 0}^\infty |a_{j}(r 2^{-i-1} )- a_{j}(r 2^{-i})|r^{-\varrho-\gamma +\gamma_j} \leq C \sum\limits_{i = 0}^\infty 2^{-i\varrho}. |
This implies that, for every x\in B_r^+,
|\beta F_{k_0}(x)-F_{k_0, r}^w(x)|\leq |\beta - a_{k_0}(r)| r^{\gamma }+ \sum\limits_{j = 1}^{k_0-1} |a_j(r)| r^{ \gamma_j} \leq C r^{\gamma +\varrho}. |
That is (7.4) as claimed.
Remark 7.1. (ⅰ) Since \varphi is conformal, we have that {\tilde F }: = F_{k_0}\circ\varphi satisfies {\tilde F }\in H^1 (\mathcal{U}_R) and solves the homogeneous equation
\begin{cases} \Delta {\tilde F } = 0, & \textrm{in}\; \mathcal{U}_R, \\ {\tilde F } = 0, & \textrm{on }\; \Gamma _d\cap\partial \mathcal{U}_R\\ \partial _{\nu }{\tilde F } = 0, & \textrm{on} \; \Gamma _n\cap\partial \mathcal{U}_R. \end{cases} | (7.7) |
(ⅱ) Let \Upsilon: U^+: = \mathcal{B}\cap U\to B_\rho^+ define a C^2 parametrization (e.g. given by a system of Fermi coordinates), for some open neighborhood U of 0, with \Upsilon(0) = 0, D \Upsilon(0) = Id, \Upsilon(\Gamma _n\cap U)\subset \Gamma _n^\rho and \Upsilon(\Gamma _d\cap U)\subset \Gamma _d^\rho. By Theorem 1.1, for every \varrho\in (0, 1/2), there exist C, \rho_0>0 such that
| u( \Upsilon^{-1}(y))-\beta \alpha ^{\frac{2k_0-1}{2}} F_{k_0}(y)|\leq C |y|^{ \frac{2k_0-1}{2}+ \varrho }, \quad \textrm{for every} \;y\in B_{\rho_0}^+, | (7.8) |
with \alpha>0 as in (2.3). Indeed, to see this, we first observe that (7.8) is equivalent to
| u(x)-\beta F_{k_0}(\alpha \Upsilon(x))|\leq c |x|^{\frac{2k_0-1}{2}+\varrho}, \quad \textrm{for every}\;x\in \Upsilon^{-1}(B_{\rho_0}^+), | (7.9) |
for some constant c>0. We then further note that
|D F_{k_0}(x)|\leq c |x|^{\frac{2k_0-1}{2}-1} |
and thus
\begin{align*} |F_{k_0}(\alpha \Upsilon(x))-F_{k_0}(\varphi (x))&|\leq c |x|^{\frac{2k_0-1}{2}-1}|\alpha \Upsilon(x) -\varphi (x)|\\ &\leq c |x|^{\frac{2k_0-1}{2}-1} |x|^2\\ &\leq c |x|^{\frac{2k_0-1}{2}+1}, \end{align*} |
in a neighborhood of 0, where c>0 is a positive constant independent of x possibly varying from line to line. This, together with (1.4) and the triangular inequality, gives (7.9).
Proof of Corollary 1.3. From Theorem 1.1 and (7.8) it follows that, if u\in H^1(\Omega) is a non-trivial solution to (1.1), then there exist k_0\in \mathbb{N}\setminus\{0\} and \beta\in\mathbb{R}\setminus\{0\} such that, for every t\in[0, \pi),
\lim\limits_{r\to 0}r^{-\frac{2k_0-1}{2}} u(r\cos t, r\sin t) = \beta\alpha^{\frac{2k_0-1}{2}}\cos\left(\tfrac{2k_0-1}{2}t\right). | (7.10) |
Therefore, if u\geq0, we have that necessarily k_0 = 1 so that statement (ⅰ) follows. Moreover, (7.8) implies that
u(r\cos t, r\sin t)\geq \beta\alpha^{1/2}r^{1/2}\cos\left(\tfrac{t}{2}\right) -Cr^{1/2+\varrho}, |
which easily provides statement (ⅱ).
Proof of Corollary 1.4. Let us assume by contradiction that u\not\equiv 0. Then, Theorem 1.1 and (7.8) imply that (7.10) holds for every t\in[0, \pi) and for some k_0\in \mathbb{N}\setminus\{0\} and \beta\in\mathbb{R}\setminus\{0\}. Taking n>\frac{2k_0-1}2, (7.10) contradicts the assumption that u(x) = O(|x|^n) as |x|\to 0.
In this section we show that the presence of a logarithmic term in the asymptotic expansion cannot be excluded without assuming enough regularity of the boundary.
Let us define in the Gauss plane the set
A: = \mathbb{C}\setminus\{x_1\in \mathbb{R}\subset \mathbb{C}:x_1\leq 0\} |
and the holomorphic function \eta:A\to\mathbb{C} defined as follows:
\eta(z): = \log r+i\theta \quad \text{for any}\; z = re^{i\theta }\in A, r \gt 0, \theta \in\left(-\pi, \pi\right). |
Let us consider the holomorphic function
v(z): = e^{2\eta(-iz)}\eta(-iz) \qquad \text{for any } z\in \mathbb{C}\setminus \{ix_2: x_2 \le 0\} |
and the set
\mathcal Z: = \{z\in \mathbb{C}\setminus \{ix_2: x_2 \le 0\}:\Im(v(z)) = 0\} . | (8.1) |
If z = re^{i\theta } with r>0, \theta \in \left(-\frac \pi2, \frac{3\pi}2\right)\setminus \{-\frac \pi 4, 0, \frac \pi 4, \frac{\pi}2, \frac{3\pi}4, \pi, \frac{5\pi}4\}, then z\in \mathcal Z if and only
r = \rho(\theta ): = \exp\left[-\bigg(\theta -\frac{\pi}2\bigg)\cot(2\theta ) \right] . | (8.2) |
For some fixed \sigma\in \left(0, \frac\pi 2\right), we define the curve \Gamma _+\subset \mathcal Z parametrized by
\Gamma _+: \begin{cases} x_1(\theta ) = \rho(\theta )\cos\theta \\ x_2(\theta ) = \rho(\theta )\sin\theta \end{cases} \qquad \theta \in (-\sigma, 0) \, . | (8.3) |
If we choose \sigma>0 sufficiently small then \Gamma _+ is the graph of a function h_+ defined in a open right neighborhood U_+ of 0. Moreover h_+ is a Lipschitz function in U_+, h_+\in C^2(U_+) and
\lim\limits_{x_1\to 0^+} \frac{h_+(x_1)}{x_1} = 0 \, , \quad \lim\limits_{x_1\to 0^+} h_+'(x_1) = 0 \, . | (8.4) |
Then we define the harmonic function
u(x_1, x_2): = -\Im(v(z)) \qquad \text{ for any } z = x_1+ix_2\in \mathbb{C}\setminus \{iy: y \le 0\} \, . | (8.5) |
In polar coordinates the function u reads
u(r, \theta ) = r^2 \left[(\log r) \sin(2\theta )+\left(\theta -\frac{\pi}2\right) \cos(2\theta )\right] . | (8.6) |
From (8.1–8.2) and (8.6) we deduce that u vanishes on \Gamma _+.
The next step is to find a curve \Gamma _- on which \frac{\partial u}{\partial\nu} = 0 where \nu = (\nu_1, \nu_2) is the unit normal to \Gamma _- satisfying \nu_2\le 0. We observe that
u(x_1, x_2) = x_1 x_2\log(x_1^2+x_2^2)+\left[\arctan\left(\frac{x_2}{x_1}\right)+\frac \pi 2\right](x_1^2-x_2^2) \quad \text{for any } x_1 \lt 0 , \, x_2\in \mathbb{R} \, . |
From direct computation we obtain
\frac{\partial u}{\partial x_1}(x_1, x_2) = x_2\log(x_1^2+x_2^2)+x_2 +2\left[\arctan\left(\frac{x_2}{x_1}\right)+\frac \pi 2\right]x_1 \, , \\ \frac{\partial u}{\partial x_2}(x_1, x_2) = x_1\log(x_1^2+x_2^2)+x_1 -2\left[\arctan\left(\frac{x_2}{x_1}\right)+\frac \pi 2\right]x_2 \, . |
We now define
H_1(x_1, x_2) = \frac{2\left[\arctan\left(\frac{x_2}{x_1}\right)+\frac \pi 2\right]x_1}{\log(x_1^2+x_2^2)} \quad \text{and} \quad H_2(x_1, x_2) = \frac{2\left[\arctan\left(\frac{x_2}{x_1}\right)+\frac \pi 2\right]x_2}{\log(x_1^2+x_2^2)} |
on the set B_1\cap \Pi_- where \Pi_-: = \{(x_1, x_2)\in\mathbb{R}^2:x_1 < 0\}. One can easily check that H_1, H_2 admit continuous extensions defined on B_1\cap \overline \Pi_- which we still denote by H_1 and H_2 respectively. We also observe that H_1, H_2\in C^1(B_1\cap \overline \Pi_-). Therefore H_1, H_2 may be extended also on the right of the x_2-axis up to restrict them to a disk of smaller radius. For example one may define
H_1(x_1, x_2): = 3H_1(-x_1, x_2)-2H_1(-2x_1, x_2) \quad \text{and} \quad H_2(x_1, x_2): = 3H_2(-x_1, x_2)-2H_2(-2x_1, x_2) |
for any (x_1, x_2)\in B_{1/2}\cap \Pi_+ where we put \Pi_+: = \{(x_1, x_2)\in\mathbb{R}^2:x_1>0\}. One may check that the new functions H_1, H_2 belong to C^1(B_{1/2}).
We can now define the functions V_1, V_2:B_{1/2}\to \mathbb{R} by
V_1(x_1, x_2): = \begin{cases} x_2+\frac{x_2}{\log(x_1^2+x_2^2)}+H_1(x_1, x_2), & \quad \text{if } (x_1, x_2)\neq (0, 0), \\ 0, & \quad \text{if } (x_1, x_2) = (0, 0) \, , \end{cases} \\ V_2(x_1, x_2): = \begin{cases} x_1+\frac{x_1}{\log(x_1^2+x_2^2)}-H_2(x_1, x_2), & \quad \text{if } (x_1, x_2)\neq (0, 0), \\ 0, & \quad \text{if } (x_1, x_2) = (0, 0) \, . \end{cases} |
One may verify that V_1, V_2\in C^1(B_{1/2}). Moreover we have
\frac{\partial V_1}{\partial x_1}(0, 0) = 0 \, , \quad \frac{\partial V_1}{\partial x_2}(0, 0) = 1 \, , \quad \frac{\partial V_2}{\partial x_1}(0, 0) = 1 \, , \quad \frac{\partial V_2}{\partial x_2}(0, 0) = 0 \, . |
Then we consider the dynamical system
\begin{cases} x_1'(t) = V_1(x_1(t), x_2(t)) \\ x_2'(t) = V_2(x_1(t), x_2(t)) \, . \end{cases} | (8.37) |
After linearization at (0, 0), by [15,Theorem Ⅸ.6.2] we deduce that the stable and unstable manifolds corresponding to the stationary point (0, 0) of (8.7), are respectively tangent to the eigenvectors (1, -1) and (1, 1) of the matrix DV(0, 0) where V is the vector field (V_1, V_2).
We define the curve \Gamma _- as the stable manifold of (8.7) at (0, 0) intersected with B_\varepsilon \cap \Pi_- where \varepsilon\in (0, \frac 12) can be chosen sufficiently small in such a way that \Gamma _- becomes the graph of a function h_- defined in a open left neighborhood U_- of 0. Combining the definitions of h_+ and h_- we can introduce a function h:U_+\cup U_-\cup \{0\}\to \mathbb{R} such that h\equiv h_+ on U_+, h\equiv h_- on U_- and h(0) = 0.
Then we introduce a positive number R sufficiently small and a domain \Omega \subseteq B_R such that
\Omega = \{(x_1, x_2)\in B_R:x_2 \gt h(x_1)\}. |
One can easily check that the function u defined in (8.5) belongs to H^1(\Omega). From the above construction, we deduce that u = 0 on \Gamma _+\cap \partial\Omega and \frac{\partial u}{\partial \nu} = 0 on \Gamma _-\cap \partial\Omega . We observe that \partial\Omega admits a corner at 0 of amplitude \frac{3\pi}4.
The presence of a logarithmic term in u can be explained since the C^{2, \delta }-regularity assumption is not satisfied from the right, i.e. h_{|U_+\cup \{0\}}\not\in C^{2, \delta }(U_+\cup \{0\}) for any \delta \in (0, 1). To see this, it is sufficient to study the behavior of h(x_1)-x_1 h'(x_1) in a right neighborhood of zero.
By (8.3) we know that \theta \in \big(-\frac{\pi}2, 0\big) and hence, if x_1 belongs to a sufficiently small right neighborhood of 0, by (8.2) we have
\frac{1}2 \log \big(x_1^2+(h_+(x_1))^2\big) \tan\left[2\arctan\left(\frac{h_+(x_1)}{x_1}\right)\right] +\arctan\left(\frac{h_+(x_1)}{x_1}\right)-\frac{\pi}2 = 0 . | (8.8) |
By (8.4) and (8.8) we have that, as x_1\to 0^+,
\tan \left[2\arctan\left(\frac{h_+(x_1)}{x_1}\right)\right] = -\frac{2\arctan\big(\frac{h_+(x_1)}{x_1}\big)-\pi}{\log \big(x_1^2+(h_+(x_1))^2\big)} = \frac{\pi}2 \frac{1}{\log x_1}+o\left(\frac{1}{\log x_1}\right) . | (8.9) |
Differentiating both sides of (8.8) and multiplying by x_1^2+(h_+(x_1))^2 we obtain the identity
\big(x_1+h_+(x_1)h_+'(x_1)\big) \, \tan\left[2\arctan\left(\frac{h_+(x_1)}{x_1}\right)\right] \\ \quad \quad \quad \quad \quad +\left\{ 1+\frac{\log \big(x_1^2+(h_+(x_1))^2\big)} {\cos^2\left[2\arctan\left(\frac{h_+(x_1)}{x_1}\right)\right]} \right\} \big(x_1h_+'(x_1)-h_+(x_1) \big) = 0 | (8.10) |
and hence (8.4) and (8.9) yield
x_1h_+'(x_1)-h_+(x_1)\sim -\frac{\pi}4 \, \frac{x_1}{\log^2 x_1} \qquad \text{as } x_1\to 0^+ . | (8.11) |
This shows that h_+ \not \in C^2(U_+\cup\{0\}) (and a fortiori cannot be extended to be of class C^{2, \delta }).
We observe that the reason of the appearance of a logarithmic term is not due to the presence of a corner at 0; indeed we are going to construct a domain with C^1-boundary for which the same phenomenon occurs. In order to do this, it is sufficient to take the domain \Omega and the function u defined above and to apply a suitable deformation in order to remove the angle. We recall that \Omega exhibits a corner at 0 whose amplitude is \frac{3\pi}4.
For this reason, we define F:\mathbb{C}\setminus\{ix_2:x_2\le 0\}\to \mathbb{C} by
F(z): = r^{\frac 43} \, e^{i\frac 43 \theta } \quad \text{for any } z = re^{i\theta } \, , \ r \gt 0\, , \theta \in \left(-\frac{\pi}2, \frac{3\pi}2\right) . |
We observe that, up to shrink R if necessary, the map F:\Omega \to F(\Omega) is invertible so that we may define \widetilde\Omega : = F(\Omega) and \widetilde u:\widetilde \Omega \to \mathbb{R}, \widetilde u(y_1, y_2): = u(F^{-1}(y_1, y_2)) for any (y_1, y_2)\in \widetilde\Omega .
We also define the curves \widetilde\Gamma _+: = F(\Gamma _+) and \widetilde\Gamma _-: = F(\Gamma _-). Up to shrink R if necessary, we may assume that \widetilde\Gamma _+ and \widetilde\Gamma _- are respectively the graphs of two functions \widetilde h_+ and \widetilde h_-.
It is immediate to verify that \widetilde u = 0 on \widetilde\Gamma _+. We also prove that \frac{\partial\widetilde u}{\partial \nu} = 0 on \widetilde\Gamma _-. To avoid confusion with the notion of normal unit vectors to \Gamma _- and \widetilde\Gamma _- we denote them respectively with \nu_{\Gamma _-} and \nu_{\widetilde\Gamma _-}. Since \widetilde u is still harmonic, \frac{\partial u}{\partial \nu_{\Gamma _-}} = 0 on \Gamma _- and F is a conformal mapping, for any \widetilde \varphi\in C^\infty_c(\widetilde\Omega \cup \widetilde \Gamma _-), we have
\begin{align*} \int_{\widetilde\Gamma _-} \frac{\partial \widetilde u}{\partial\nu_{\widetilde \Gamma _-}}\widetilde \varphi\, ds & = \int_{\widetilde\Omega } \nabla\widetilde u(y)\nabla \widetilde \varphi(y)\, dy = \int_{\widetilde\Omega } [\nabla u(F^{-1}(y))(DF(F^{-1}(y)))^{-1}] \nabla \widetilde \varphi(y)\, dy \\[7pt] & = \int_{\Omega } \big[\nabla u(x)(DF(x))^{-1}\big] \nabla \widetilde \varphi(F(x))\, |\mathop{\rm det}(DF(x))| \, dx \\[7pt] & = \int_{\Omega } \big[\nabla u(x)(DF(x))^{-1}\big] \big[\nabla\varphi(x)(DF(x))^{-1}\big]\, |\mathop{\rm det}(DF(x))| \, dx \\[7pt] & = \int_{\Omega } \nabla u(x)\nabla\varphi(x) \, dx = \int_{\Gamma _-} \frac{\partial u}{\partial \nu_{\Gamma _-}}\varphi \, ds = 0 \end{align*} |
where we put \varphi(x) = \widetilde \varphi(F(x)). This proves that \frac{\partial\widetilde u}{\partial \nu_{\widetilde \Gamma _-}} = 0 on \widetilde \Gamma _-.
Finally we prove for \widetilde h_+ an estimate similar to (8.11).
From the definition of F it follows that \widetilde\Gamma _+ admits a representation in polar coordinates of the type
r = \widetilde\rho(\theta ): = \exp\left[-\left(\theta -\frac{2\pi}3\right)\cot\left(\frac{3\theta }2\right)\right] \, . | (8.12) |
Proceeding exactly as for (8.8)-(8.9) one can prove that
\frac{1}2 \log \big(x_1^2+(\widetilde h_+(x_1))^2\big) \tan\left[\frac 32\arctan\left(\frac{\widetilde h_+(x_1)}{x_1}\right)\right] +\arctan\left(\frac{\widetilde h_+(x_1)}{x_1}\right)-\frac{2\pi}3 = 0 \, . | (8.13) |
As we did for h_+, also for the function \widetilde h_+ one can prove that
\lim\limits_{x_1\to 0} \frac{\widetilde h_+(x_1)}{x_1} = 0 \, , \quad \lim\limits_{x_1\to 0^+} \widetilde h_+'(x_1) = 0 \, . | (8.14) |
By (8.14) we have
\tan \left[\frac 32\arctan\left(\frac{\widetilde h_+(x_1)}{x_1}\right)\right] = -\frac{2\arctan\big(\frac{\widetilde h_+(x_1)}{x_1}\big)-\frac{4\pi}3}{\log \big(x_1^2+(\widetilde h_+(x_1))^2\big)} = \frac{2\pi}3 \frac{1}{\log x_1}+o\left(\frac{1}{\log x_1}\right) \quad \text{as } x_1\to 0^+ \, . | (8.15) |
Differentiating both sides of (8.13) and multiplying by x_1^2+(\widetilde h_+(x_1))^2 we obtain the identity
\big(x_1+\widetilde h_+(x_1)\widetilde h_+'(x_1)\big) \, \tan\left[\frac 32\arctan\left(\frac{\widetilde h_+(x_1)}{x_1}\right)\right] \\ \quad \quad \quad \quad \quad +\left\{ 1+\frac{3\log \big(x_1^2+(\widetilde h_+(x_1))^2\big)} {4\cos^2\left[\frac 32\arctan\left(\frac{\widetilde h_+(x_1)}{x_1}\right)\right]} \right\} \big(x_1\widetilde h_+'(x_1)-\widetilde h_+(x_1) \big) = 0. | (8.16) |
By (8.14), (8.15) and (8.16), we obtain
x_1\widetilde h_+'(x_1)-\widetilde h_+(x_1)\sim -\frac{4\pi}9 \, \frac{x_1}{\log^2 x_1} \qquad \text{as } x_1\to 0^+ . | (8.17) |
The above arguments show that \partial\widetilde\Omega is of class C^1 but not of class C^{1, \delta } (and a fortiori not of class C^{2, \delta }).
M.M. Fall is supported by the Alexander von Humboldt foundation. V. Felli is partially supported by the PRIN2015 grant "Variational methods, with applications to problems in mathematical physics and geometry". A. Ferrero is partially supported by the PRIN2012 grant "Equazioni alle derivate parziali di tipo ellittico e parabolico: aspetti geometrici, disuguaglianze collegate, e applicazioni" and by the Progetto di Ateneo 2016 of the University of Piemonte Orientale "Metodi analitici, numerici e di simulazione per lo studio di equazioni differenziali a derivate parziali e applicazioni". A. Ferrero and V. Felli are partially supported by the INDAM-GNAMPA 2017 grant "Stabilitàe analisi spettrale per problemi alle derivate parziali".
All authors declare no conflicts of interest in this paper.
[1] |
Adolfsson V, Escauriaza L (1997) C1,α domains and unique continuation at the boundary. Comm Pure Appl Math 50: 935–969. doi: 10.1002/(SICI)1097-0312(199710)50:10<935::AID-CPA1>3.0.CO;2-H
![]() |
[2] | Adolfsson V, Escauriaza L, Kenig C (1995) Convex domains and unique continuation at the boundary. Rev Mat Iberoam 11: 513–525. |
[3] |
Brändle C, Colorado E, de Pablo A, et al. (2013) A concave-convex elliptic problem involving the fractional Laplacian. Proceedings of the Royal Society of Edinburgh Section A: Mathematics 143: 39–71. doi: 10.1017/S0308210511000175
![]() |
[4] |
Caffarelli L, Silvestre L (2007) An extension problem related to the fractional Laplacian. Commun Part Diff Eq 32: 1245–1260. doi: 10.1080/03605300600987306
![]() |
[5] | Carleman T (1939) Sur un problème d'unicité pur les systèmes d'équations aux dérivées partielles à deux variables indépendantes. Ark Mat Astr Fys 26: 9. |
[6] |
Dal Maso G, Orlando G, Toader R (2015) Laplace equation in a domain with a rectilinear crack: higher order derivatives of the energy with respect to the crack length. NoDEA Nonlinear Diff 22: 449–476. doi: 10.1007/s00030-014-0291-0
![]() |
[7] |
Doktor P, Zensek A (2006) The density of infinitely differentiable functions in Sobolev spaces with mixed boundary conditions. Appl Math 51: 517–547. doi: 10.1007/s10492-006-0019-5
![]() |
[8] |
Fall MM, Felli V (2014) Unique continuation property and local asymptotics of solutions to fractional elliptic equations. Commun Part Diff Eq 39: 354–397. doi: 10.1080/03605302.2013.825918
![]() |
[9] |
Felli V, Ferrero A (2013) Almgren-type monotonicity methods for the classification of behaviour at corners of solutions to semilinear elliptic equations. Proceedings of the Royal Society of Edinburgh Section A: Mathematics 143: 957–1019. doi: 10.1017/S0308210511001314
![]() |
[10] |
Felli V, Ferrero A (2014) On semilinear elliptic equations with borderline Hardy potentials. J Anal Math 123: 303–340. doi: 10.1007/s11854-014-0022-9
![]() |
[11] | Felli V, Ferrero A, Terracini S (2011) Asymptotic behavior of solutions to Schrödinger equations near an isolated singularity of the electromagnetic potential. J Eur Math Soc 13: 119–174. |
[12] |
Felli V, Ferrero A, Terracini S (2012) On the behavior at collisions of solutions to Schrödinger equations with many-particle and cylindrical potentials. Discrete Contin Dyn-A 32: 3895–3956. doi: 10.3934/dcds.2012.32.3895
![]() |
[13] |
Felli V, Ferrero A, Terracini S (2012) A note on local asymptotics of solutions to singular elliptic equations via monotonicity methods. Milan J Math 80: 203–226. doi: 10.1007/s00032-012-0174-y
![]() |
[14] |
Garofalo N, Lin FH (1986) Monotonicity properties of variational integrals, Ap weights and unique continuation. Indiana U Math J 35: 245–268. doi: 10.1512/iumj.1986.35.35015
![]() |
[15] | Hartman P (1964) Ordinary Differential Equations, Wiley, New York. |
[16] |
Kassmann M, Madych WR (2007) Difference quotients and elliptic mixed boundary value problems of second order. Indiana U Math J 56: 1047–1082. doi: 10.1512/iumj.2007.56.2836
![]() |
[17] | Krantz SG (2006) Geometric function theory. Explorations in complex analysis, Cornerstones, Birkhäuser Boston, Inc., Boston, MA. |
[18] |
Kukavica I (1998) Quantitative uniqueness for second-order elliptic operators. Duke Math J 91: 225–240. doi: 10.1215/S0012-7094-98-09111-6
![]() |
[19] |
Kukavica I, Nyström K (1998) Unique continuation on the boundary for Dini domains. Proc Amer Math Soc 126: 441–446. doi: 10.1090/S0002-9939-98-04065-9
![]() |
[20] |
Lazzaroni G, Toader R (2011) Energy release rate and stress intensity factor in antiplane elasticity. J Math Pures Appl 95: 565–584. doi: 10.1016/j.matpur.2011.01.001
![]() |
[21] |
Ros-Oton X, Serra J (2014) The Dirichlet problem for the fractional Laplacian: regularity up to the boundary. J Math Pures Appl 101: 275–302. doi: 10.1016/j.matpur.2013.06.003
![]() |
[22] |
Ros-Oton X, Serra J (2016) Regularity theory for general stable operators. J Differ Equations 260: 8675–8715. doi: 10.1016/j.jde.2016.02.033
![]() |
[23] |
Ros-Oton X, Serra J (2016) Boundary regularity for fully nonlinear integro-differential equations. Duke Math J 165: 2079–2154. doi: 10.1215/00127094-3476700
![]() |
[24] |
Serra J (2015) Regularity for fully nonlinear nonlocal parabolic equations with rough kernels. Calc Var Partial Dif 54: 615–629. doi: 10.1007/s00526-014-0798-6
![]() |
[25] |
Savaré G (1997) Regularity and perturbation results for mixed second order elliptic problems. Commun Part Diff Eq 22: 869–899. doi: 10.1080/03605309708821287
![]() |
[26] |
Tao X, Zhang S (2005) Boundary unique continuation theorems under zero Neumann boundary conditions. B Aust Math Soc 72: 67–85. doi: 10.1017/S0004972700034882
![]() |
[27] |
Tao X, Zhang S (2008) Weighted doubling properties and unique continuation theorems for the degenerate Schrödinger equations with singular potentials. J Math Anal Appl 339: 70–84. doi: 10.1016/j.jmaa.2007.06.042
![]() |
[28] | Wolff TH (1992) A property of measures in R^N and an application to unique continuation. Geom Funct Anal. |
1. | Veronica Felli, Alberto Ferrero, Unique continuation and classification of blow-up profiles for elliptic systems with Neumann boundary coupling and applications to higher order fractional equations, 2020, 196, 0362546X, 111826, 10.1016/j.na.2020.111826 | |
2. | L. Abatangelo, V. Felli, C. Léna, Eigenvalue variation under moving mixed Dirichlet–Neumann boundary conditions and applications, 2020, 26, 1292-8119, 39, 10.1051/cocv/2019022 | |
3. | Veronica Felli, Giovanni Siclari, Unique continuation from a crack’s tip under Neumann boundary conditions, 2022, 222, 0362546X, 113002, 10.1016/j.na.2022.113002 | |
4. | Alessandra De Luca, Veronica Felli, Stefano Vita, Strong unique continuation and local asymptotics at the boundary for fractional elliptic equations, 2022, 400, 00018708, 108279, 10.1016/j.aim.2022.108279 |