• Non ci sono risultati.

Sharp Convergence Rate of Eigenvalues in a Domain with a Shrinking Tube

N/A
N/A
Protected

Academic year: 2021

Condividi "Sharp Convergence Rate of Eigenvalues in a Domain with a Shrinking Tube"

Copied!
36
0
0

Testo completo

(1)

Sharp Convergence Rate of Eigenvalues in a Domain with a

Shrinking Tube

Veronica Felli

Roberto Ognibene

September 9, 2018

Abstract

In this paper we consider a class of singularly perturbed domains, obtained by attaching a cylindrical tube to a fixed bounded region and letting its section shrink to zero. We use an Almgren-type monotonicity formula to evaluate the sharp convergence rate of perturbed simple eigenvalues, via Courant-Fischer Min-Max characterization and blow-up analysis for scaled eigenfunctions.

Keywords. Singularly perturbed domains, asymptotics of eigenvalues, monotonicity formula. MSC classification. Primary: 35P20; Secondary: 35P15, 35J25.

1

Introduction and Main Results

The purpose of this work is to investigate the behaviour of the eigenvalues of the Dirichlet-Laplacian in a class of singularly perturbed domains: in particular we are interested in the sharp convergence rate of the eigenvalue variation, i.e. in the evaluation of the leading term in its asymptotic expan-sion. The perturbation consists in attaching a cylindrical tube to a fixed domain and letting the section of the tube shrink.

Let N ≥ 2 and Ω ⊆ RN be open, bounded and connected. Suppose that 0 ∈ ∂Ω and that ∂Ω is flat in a neighbourhood of the origin, namely

∃ Rmax> 1 such that M := {(x1, . . . , xN) ∈ RN: x1= 0, |x| ≤ Rmax} ⊆ ∂Ω. (1)

Using the following notation for the positive half-space, half-balls and half-spheres RN+ := {(x1, . . . , xN) ∈ RN: x1> 0},

Br+:= {x ∈ RN+: |x| < r}, Sr+:= {x ∈ RN+: |x| = r}, we can suppose, without losing generality, that

B+R

max⊆ Ω ⊆ R

N +.

Let Σ ⊂⊂ M be open, connected and containing the origin 0. For simplicity of exposition we assume that ∂Σ is of class C2; although this regularity assumption can be relaxed, see Remark 3.7. Moreover we assume, for sake of simplicity, that the radius of Σ in RN −1 is 1, i.e.

max

x∈∂Σ|x| = 1.

Dipartimento di Scienza dei Materiali, Universit`a di Milano–Bicocca, Via Cozzi 55, 20125 Milano, Italy, veronica.felli@unimib.it

(2)

x2, . . . , xN

x1

0

(a) Unperturbed Domain

x2, . . . , xN x1 Σ Ω T (b) Perturbed Domain Figure 1

Finally we assume that Σ is starshaped with respect to 0, i.e.

x · ν ≥ 0 for all x ∈ ∂Σ, (2)

where ν denotes the exterior unit normal vector to ∂Σ. Let  ∈ R, with 0 <  ≤ 1, and let T:= (−1, 0] × Σ be a cylindrical tube with section Σ = {x : x ∈ Σ}. Let us denote by

Ω= Ω ∪ T (3)

the perturbed domain (see Figure 1).

Let p ∈ L∞(RN) be a weight function such that p ≥ 0 a.e. and p 6≡ 0 in Ω. For any open, bounded set ω ⊆ RN, we consider the weighted Dirichlet eigenvalue problem for the Laplacian on ω

(

−∆ϕ = λpϕ, in ω,

ϕ = 0, on ∂ω. (Eω)

By classical spectral theory we have that, if p 6≡ 0 in ω, there exists a sequence of positive eigen-values of (Eω)

0 < λ1(ω) < λ2(ω) ≤ λ3(ω) ≤ . . .

repeated according to their multiplicity. We denote by (λn)n := (λn(Ω))nthe sequence of

eigenval-ues of the unperturbed problem (EΩ), and by (ϕn)na corresponding sequence of eigenfunctions such

thatR Ωp|ϕn| 2 dx = 1 andR Ωpϕnϕmdx = 0 if n 6= m. Similarly, we denote by (λ  n)n := (λn(Ω))n

and (ϕn)n the sequences of eigenvalues and eigenfunctions of the perturbed problem (EΩ), such

thatRp|ϕn|2dx = 1 and

R

Ωpϕ 

nϕmdx = 0 if n 6= m.

Let j ∈ N be such that

λj is simple. (4)

Assumption (4) is not so restrictive: indeed, the simplicity of all eigenvalues is a generic property with respect to perturbations of the domain , see [33, 37].

Classical results (see for instance [13, 19]) ensure the continuity with respect to our domain perturbation, i.e. λ

j is simple for  small and

λj −→ λj, as  → 0. (5)

Furthermore, for every  we can choose the eigenfunction ϕ

j in such a way that

ϕj−→ ϕj, in H01(Ω

1), as  → 0, (6)

where the functions are trivially extended in Ω1 outside their domains.

(3)

The problem of convergence of eigenvalues and eigenfunctions of the Dirichlet-Laplacian with respect to perturbations of the domain has been widely studied in the past. For instance, for general perturbations that cover the shrinking tube, in [9] the authors investigated the stability of the spectrum with respect to general scalar products, while [12] dealt with the convergence of solutions of a nonlinear eigenvalue problem (see also [17, 18]). Within an extensive literature, we mention [26], [27] and [15] as detailed surveys. In [20, 14] bounds for the rate of convergence have been found; furthermore, in [36] the framework is pretty similar to ours and the author proved an estimate of the type λj− λj = O(

a), where a depends only on the distance between λ j and

its neighbours. We mention also that asymptotic expansions of the eigenvalues of the Dirichlet Laplacian in domains with a thin shrinking cylindrical protrusion of finite length were obtained in [23], see also [6] for a related problem in a two-dimensional domain with thin shoots; we notice that Theorem 4.1 in [23] provides the exact vanishing rate of the eigenvalue variation λj− λj only

when ∇ϕj(0) 6= 0, but it does not say what is the leading term in the expansion when ∇ϕj(0) = 0.

For what concerns Neumann boundary conditions, among many others, we cite [7, 8, 28, 29, 35], which take into account singular perturbations like the shrinking tube.

A motivation for the interest in studying the spectral behaviour of the Laplacian on thin branching domains comes from physics: for instance, it occurs in the theory of quantum graphs, which models the propagation of waves in quasi one-dimensional structures, like quantum wires, narrow waveguides, photonic crystals, blood vessels, etc. (see e.g. [11, 31] and reference therein). Moreover, this topic is also related with engineering problems, such as elasticity and multi-structure problems, as well explained in surveys [16, 34].

The starting points of this work are [23] and [2, 3, 22]. On the one hand, the present paper aims at providing a criterion for selecting the leading term in the asymptotic expansion given in [23], based on the vanishing order of the limit eigenfunction at the junction; on the other hand, it improves and generalizes some results of [2]. We note that [2] (as well as many of the aforementioned articles) deals with dumbbell domains in which the tubular handle is vanishing. However, from the point of view of both the expected results and the technical approach, our method does not require substantial adaptions to treat also the dumbbell case; hence for the sake of simplicity of exposition, in the present paper we consider only perturbations of type (3).

In order to state our main results, we first need to recall some known facts. Let us consider the eigenvalue problem for the standard Laplacian on the (N − 1)-dimensional unit sphere

− ∆SN −1ψ = µψ in SN −1. (7)

It is well known that the eigenvalues of (7) are µk = k(k + N − 2), for k = 0, 1, . . . and that their

multiplicities are (see [10])

mk = k + N − 2 k  +k + N − 3 k − 1  .

If Ek denotes the eigenspace of the eigenvalue µk, then Lk≥0Ek = L2(SN −1). Furthermore it

is known that the elements of Ek are spherical harmonics, i.e. homogeneous polynomials (of N

variables) of degree k. We are interested in eigenfunctions of (7) that vanish on {x1 = 0}, so let

us call

Ek0:= {ψ ∈ Ek: ψ(0, θ2, . . . , θN) = 0} .

It is well known (see e.g. [21, Th. 1.3]) that the local behaviour at 0 ∈ ∂Ω of eigenfunctions of (EΩ) can be described in term of spherical harmonics vanishing on {x1 = 0}. In particular there

exist k ∈ N, k ≥ 1, and Ψ ∈ E0 k, Ψ 6= 0, such that r−kϕj(rθ) → Ψ in C1,τ(S1+), as r → 0 +, (8) r1−k∇ϕj(rθ) → ∇Ψ in C0,τ(S1+, R N), as r → 0+, (9)

for all τ ∈ (0, 1). Furthermore the asymptotic homogeneity order k can be characterized as the limit of an Almgren frequency function (see [5]), i.e.

(4)

Hereafter we will denote

ψk(rθ) := rkΨ(θ), r ≥ 0, θ ∈ S+1. (10)

The exact asymptotic estimate of the eigenvalue variation we are going to prove involves a nonzero constant mk(Σ) which admits the following variational characterization. Let us consider the

func-tional J : D1,2(Π) −→ R, J (u) :=1 2 Z Π |∇u|2dx − Z Σ u∂ψk ∂x1 dS, where T1−:= (−∞, 0] × Σ, Π := T1−∪ RN+,

and, for any open set ω ⊆ RN, D1,2(ω) denotes the completion of the space C

c (ω) with respect

to the L2 norm of the gradient (see Section 2 for further details). In dimension 2, we will always

deal with spaces D1,2(ω) with ω such that RN \ ω contains a half-line; in this case D1,2(ω) can be

characterized as a concrete functional space thanks to the validity of a Hardy inequality also in dimension 2, see Theorem 9.1.

By standard minimization methods, one can prove that J is bounded from below and that the infimum

mk(Σ) := inf

u∈D1,2(Π)J (u) (11)

is attained by some wk. Moreover

mk(Σ) = − 1 2 Z Π |∇wk| 2 dx = −1 2 Z Σ ∂ψk ∂ν wkdS < 0, (12)

see [22]. With this framework in mind we are able to state our first (and main) result.

Theorem 1.1. Under assumptions (1), (2) and (4), let k denote the vanishing order of the un-perturbed eigenfunction ϕj as in (8)–(9). Then

lim →0 λj− λj N +2k−2 = Ck(Σ), where Ck(Σ) = −2mk(Σ) > 0 (13) and mk(Σ) is defined in (11).

We recall that, for N ≥ 3, an asymptotic expansion for the eigenvalue variation is constructed using the concordance method in [23, Theorem 4.1], but explicit formulas are given only for the first perturbed coefficient, which turns out to be a multiple of |∇ϕj(0)|2; in dimension N = 2,

[23, Theorem 10.1] performs a more detailed asymptotic analysis with the computation of all the coefficients. Hence, for N ≥ 3, [23] finds outs what is the leading term in the asymptotic expansion only when ∇ϕj(0) 6= 0. We emphasize that, differently from [23], Theorem 1.1 detects the exact

vanishing rate of λj− λj also when ∇ϕj(0) = 0 and N ≥ 3; more precisely it establishes a direct

correspondence between the order of the infinitesimal λj− λjand the number k, which is the order

of vanishing of ϕj at the junction point 0.

The proof of Theorem 1.1 is based on lower and upper bounds for the difference λj− λjcarried

out using the Min-Max Courant-Fischer characterization of the eigenvalues, see Section 6. To obtain the exact asymptotics for the eigenvalue variation it is crucial to sharply control the energy of perturbed eigenfunctions in neighbourhoods of the junction with radius of order . The sharpness of our energy estimates is related to the identification of a nontrivial limit profile for blow-up of scaled eigenfunctions, as stated in the following theorem.

Theorem 1.2. Under the same assumptions of Theorem 1.1, let ϕ

j be chosen as in (6). Then

−kϕj(x) → Φ(x) as  → 0, in H1(T

1 ∪ B +

R) for all R > 1, where Φ := wk+ ψk, being wk the minimizer for (11) and ψk the

(5)

As mentioned before, Theorem 1.1 generalizes and improves [2, Th. 1.1]: indeed, in [2] the weight p was assumed to vanish in a neighbourhood of the junction Σ and only the case of vanishing order k = 1 for the unperturbed eigenfunction ϕj was considered. Furthermore the dimension

N = 2 was not included in [2]. As in [2], a fundamental tool for the proof of the energy estimates needed to study the local behaviour of eigenfunctions is an Almgren-type monotonicity formula, which was first introduced by Almgren [5] and then used by Garofalo and Lin [24] to study unique continuation properties for elliptic partial differential equations.

In the particular case treated in [2, 22], precise pointwise estimates from above and from below for the perturbed eigenfunction and its gradient were directly obtained via comparison and maximum principles: indeed, if the limit eigenfunction has minimal vanishing order at the origin and the weight vanishes around the junction, then such eigenfunction has a fixed sign and is harmonic in a neighbourhood of 0. These estimates were used in [22] to get rid of a remainder term in the derivative of the Almgren quotient for the perturbed problem, however they are not available in the more general framework of the present paper. Nevertheless, under the geometric assumption (2) on the tube section we succeed in proving that the remainder term has a positive sign, thus obtaining the monotonicity formula, see Proposition 5.9. We also point out that the 2-dimensional case requires the proof of an ad hoc Hardy type inequality for functions vanishing on a fixed half-line, see (117).

We observe that in [2, 22] the limit of the blow-up family −kϕ

j(x) was recognized by its

frequency at infinity, which must be necessarily equal to the minimal one, i.e. 1, in the particular case k = 1. In the general case k ≥ 1, the monotonicity argument implies that the frequency of the limit profile is less than or equal to k, and this seems to be not enough for a univocal identification. To overcome this difficulty, we use here an argument inspired by [1] and based on a local inversion result giving an energy control for the difference between the blow-up eigenfunction and a k-homogeneous profile, see Corollary 7.4.

The paper is organized as follows. After some preliminary results in Section 2, in Section 3 we prove a Pohozaev-type identity, which is combined with the Poincar´e inequalities of Section 4 to develop a monotonicity argument in Section 5. From the monotonicity formula established in Corollary 5.10, we derive some local energy estimates which allow us to deduce sharp upper and lower bounds for the eigenvalue variation in Section 6. In Section 7 we perform a blow-up analysis for scaled eigenfunctions from which we deduce first Theorem 1.2 and then, in Section 8, our main result Theorem 1.1. Finally, in the appendix we recall an Hardy type inequality in dimension 2 for functions vanishing on half-lines and an abstract lemma on maxima of quadratic forms.

2

Preliminaries and Notation

In this section we introduce some basic definitions and notation which will be useful in the rest of the paper. We start fixing some notation:

Ωr:= T∪ B+r,  ∈ (0, 1), r ∈ (, Rmax), Cr:= ∂Br+\ S + r, r ∈ (0, Rmax), Πr:= T1−∪ B + r, r > 1.

For any measurable set ω ⊆ RN, we denote as |ω| its N -dimensional Lebesgue measure. For any R ≥ 2, we will denote as ηR a cut-off function satisfying

ηR∈ C∞(Π), ηR(x) = ( 1, for x ∈ Π \ ΠR, 0, for x ∈ ΠR/2, |ηR(x)| ≤ 1, |∇ηR(x)| ≤ 4/R for all x ∈ Π. (14)

We now recall a well known quantitative result about the first eigenvalue of the Dirichlet-Laplacian on bounded domains.

Theorem 2.1 (Faber-Krahn Inequality). Let ω ⊆ RN be open and bounded and let λD

(6)

the first eigenvalue of the Dirichlet-Laplacian on ω. Then λD1(ω) ≥

λD1(B1)|B1| 2/N

|ω|2/N ,

where B1 denotes the N -dimensional ball centered at the origin and with radius 1.

Moreover, if we denote CN := 1 |B1| 2/N λD 1(B1) , (15)

and we combine the previous Theorem with the usual Poincar´e Inequality, we have that Z

ω

|u|2dx ≤ CN|ω|2/N

Z

ω

|∇u|2dx for all u ∈ H01(ω). (16)

2.1

The Space H

R

For R > 1 let us define the function space HRas the completion of Cc∞(ΠR∪ SR+) with respect to

the norm induced by the scalar product (u, v)HR:=

Z

ΠR

∇u · ∇v dx.

Since ΠR is bounded in at least 1 direction, the Poincar´e Inequality holds. Hence HR,→ H1(ΠR)

continuously and we have the following characterization

HR:=u ∈ H1(ΠR) : u = 0 on ∂ΠR\ SR+ .

Moreover, when N ≥ 3, the classical Sobolev inequality implies that HR,→ L2

(ΠR) continuously,

where 2∗= N −22N .

2.2

Limit Profiles

In this section we introduce some limit profiles that will appear in the blow-up analysis of scaled eigenfunctions. We recall the following result from [22, Lemma 2.4].

Proposition 2.2. For every ψ ∈ C2

(RN

+) ∩ C1(RN+) such that

( −∆ψ = 0, in RN +,

ψ = 0, on ∂RN+, there exists a unique Φ = Φ(ψ) : Π → R such that

Φ ∈ HR, for all R > 1, (17) ( −∆Φ = 0, in Π, Φ = 0, on ∂Π, (18) Z Π |∇(ψ − Φ)|2dx < +∞. (19)

Hereafter we will denote

Φ := Φ(ψk) (20)

where ψk is the function defined in (10). As observed in [22] we have that

Φ = (

ψk+ wk, in RN+,

wk, in Π \ RN+,

(21)

where wk is the function realizing the minimum mk(Σ) in (11). We observe that, in the particular

case N = 2, the function Φ corresponds to the function Xk introduced in [23, Sections 10-11]. By

(7)

Lemma 2.3. For every R > 1 there exists a unique function UR∈ HR solution to the following minimization problem min u∈HR Z ΠR |∇u|2dx : u = ψk on SR+  . Moreover it weakly solves

     −∆UR= 0, in ΠR, UR= 0, on ∂ΠR\ SR+, UR= ψk, on SR+.

Lemma 2.4. For every r > 1 we have

UR−→ Φ in Hr, as R → +∞.

Proof. We can assume R > max{r, 2}. The function UR− Φ satisfies, in a weak sense,

     −∆(UR− Φ) = 0, in ΠR, UR− Φ = 0, on ∂ΠR\ SR+, UR− Φ = ψk− Φ, on SR+,

and then it is the least energy function among those having these boundary conditions. Let η = ηR∈ C∞(Π) be the cut-off function defined in (14). Then

Z Πr |∇(UR− Φ)|2dx ≤ Z ΠR |∇(UR− Φ)|2dx ≤ Z ΠR |∇(η(ψk− Φ))|2dx ≤ ≤ 2 Z ΠR |∇η|2|ψk− Φ| 2 dx + 2 Z Π |η|2|∇(ψk− Φ)| 2 dx ≤ ≤ 32 R2 Z ΠR\ΠR/2 |ψk− Φ| 2 dx + 2 Z Π−ΠR/2 |∇(ψk− Φ)| 2 dx ≤ ≤ 32 Z Π\ΠR/2 |ψk− Φ| 2 |x|2 dx + 2 Z Π−ΠR/2 |∇(ψk− Φ)| 2 dx −→ 0

thanks to (19) and Hardy’s Inequality. In the case N = 2 we use the fact that 1 + |x|2≤ 2|x|2for |x| ≥ 1 and the 2-dimensional Hardy’s Inequality (117).

Using again the Dirichlet principle, we construct also the limit profile ZR as follows.

Lemma 2.5. For every R > 1 there exists a unique function ZR ∈ H1(BR+) solution to the

following minimization problem

min u∈H1(B+ R) ( Z BR+ |∇u|2dx : u = 0 on CR, u = Φ on SR+ ) .

Moreover it weakly solves

     −∆ZR= 0, in BR+, ZR= 0, on CR, ZR= Φ, on SR+.

3

A Pohozaev-Type Inequality

The purpose of this section is to prove the following inequality.

Proposition 3.1. There exists ˜, ˜r > 0, with 0 < ˜ < ˜r ≤ Rmax, such that, for  ∈ (0, ˜], r ∈ (, ˜r]

(8)

h(s)

s δ

δ

(a) The function h

x2, . . . , xN x1 r Ω r,δ Qδ Sr+

(b) The Approximating Domain

Figure 2

We observe that solutions to problems of type (

−∆u = f, in Ωr, u = 0, on ∂Ωr, with f ∈ L2(Ω

r), in general do not belong to H2(Ωr) since ∂Ωr is only Lipschitz continuous and

doesn’t verify a uniform exterior ball condition (which ensures L2-integrability of second order

derivatives, see [4]). But, along a proof of a Pohozaev Identity, one tests the equation with the function ∇ϕ

i· x, which could fail to be H1 in our case. To overcome this difficulty we implement

an approximation process.

3.1

Approximating Domains

Let  ∈ (0, 1] and r ∈ (, Rmax]: in order to remove the concave edge Γ := ∂(Σ), we will

approximate the domain Ω

r with a family of starshaped domains Ωr,δ (with 0 ≤ δ < r − ) such

that

Ωr,0= Ωr, Ωr,δ1 ⊂ Ω

r,δ2 for all 0 ≤ δ1≤ δ2< r − ,

and such that every Ω

r,δ verifies the uniform exterior ball condition. In particular we will define

Qδ such that

Ωr,δ= Ωr∪ Qδ.

For δ > 0 small we define a “δ-enlargement” of Σ:

Σδ:=x ∈ RN −1\ Σ : dist(x, Γ) < δ .

Let h ∈ C∞((0, +∞)) ∩ C([0, +∞)) such that h(0) = δ, h(s) = 0 for s ∈ [δ, +∞), h0(s) < 0 for all s ∈ (0, δ) and h−1(s) = h(s) for s ∈ (0, δ). We define G : Σδ ⊂ RN −1→ R as

G(z) := −h(d(z)), where d(z) = dist(z, Γ). Now, let

Qδ =x ∈ RN: (0, x2, . . . , xN) ∈ Σδ and G(x2, . . . , xN) < x1≤ 0 ,

see Figure 2.

For what concerns the regularity we observe that, since Γ is of class C2, then also d and the

(9)

Remark 3.2. We point out that Ω

r,δ is starshaped. Indeed, first, if x ∈ Cr\ (Σδ∪ Σ), trivially

x · ν(x) = 0. If x ∈ {x1= −} ∩ ∂Ωr,δ, then x · ν(x) = −x1=  > 0. Third, if x ∈ (−, −δ) × Γ,

then ν(x) = ν0(x0), where x0 = (x2, . . . , xN) and ν0(x0) is the exterior unit normal of Γ; thus

x · ν(x) = x0· ν0(x0) ≥ 0 by (2). Finally, let x ∈ {(G(x0), x0) : x0∈ Σδ}: in this case we have that

ν(x) = (−1, −h 0(d(x0))∇d(x0)) k(−1, −h0(d(x0))∇d(x0))k, and so x · ν(x) = −G(x 0) − h0(d(x0))∇d(x0) · x0 k(−1, −h0(d(x0))∇d(x0))k ≥ 0,

where we used the properties of the functions G and h and the fact that, since 0 ∈ Σ and Σ is starshaped, ∂x∂d

ixi≥ 0 for any i = 2, . . . , N , see [25, Proof of Lemma 14.16].

3.2

Approximating Problems

For α ∈ (0, 1), let us fix ˜r, ˜ > 0, with 0 < ˜ < ˜r ≤ Rmax, such that

Ωr,δ ≤  1 − α CNλjkpk∞ N2 for all  ∈ (0, ˜), r ∈ (, ˜r), δ ∈ (0, r − ), (23) where CN has been defined in (15).

For fixed i ∈ {1, . . . , j},  ∈ (0, ˜], r ∈ (, ˜r] and for all δ ∈ (0, r − ), let us consider the problem      −∆u = λ ipu, in Dδ, u = 0, on ∂Dδ\ Sr+, u = ϕi, on Sr+, (24)

where, for simplicity of notation, in this section we call Dδ := Ωr,δ; we also denote δ0:= r − .

Theorem 3.3. There exists a unique uδ ∈ H1(Dδ) solution to problem (24). Moreover the family

{uδ}δ∈(0,δ0) is bounded in H

1(D

δ0) with respect to δ.

Proof. We observe that, if we extend ϕ

i to zero in Dδ0\ Ω



r and we let v = u − ϕi, then problem

(24) is equivalent to ( −∆v = λipv + F, in Dδ, v = 0, on ∂Dδ, (25) where F := λ

ipϕi+ ∆ϕi ∈ H−1(Dδ). Existence and uniqueness of a solution vδ ∈ H01(Dδ) to (25)

easily comes from Lax-Milgram Theorem. Indeed, the bilinear form a(v, w) :=

Z

(∇v · ∇w − λipvw) dx, v, w ∈ H01(Dδ)

is coercive, since, by (16) and (23), we have a(v, v) = Z Dδ  |∇v|2− λ ip|v| 2 dx ≥1 − CNλjkpk∞|Dδ| 2 N Z Dδ |∇v|2dx ≥ α Z Dδ |∇v|2dx. (26)

From Lax-Milgram Theorem we also know that k∇vδkL2(D

δ0)= k∇vδkL2(Dδ)≤

kF kH−1(D δ)

α ,

where vδ has been trivially extended in Dδ0 \ Dδ. One can easily prove that kF kH−1(D

δ)= O(1)

as δ → 0. Then uδ := vδ + ϕi is the unique solution to (24) and {uδ}δ∈(0,δ0) is bounded in

(10)

Theorem 3.4. If uδ ∈ H1(Dδ) is the unique solution to (24), then

uδ −→ ϕi in H 1(D

δ0), as δ → 0.

Proof. Since {uδ}δ∈(0,δ0) is bounded in H

1(D

δ0), then there exists U ∈ H

1(D

δ0) such that, up to

a subsequence,

uδ * U weakly in H1(Dδ0), as δ → 0.

Actually U ∈ H1(Ωr) and moreover it weakly solves

     −∆U = λipU, in Ω  r, U = 0, on ∂Ωr\ S+r, U = ϕi, on Sr+. From the coercivity obtained in (26) we deduce that U = ϕ

i.

In order to prove strong convergence in H1(D

δ0), we notice that Z Dδ0 |∇uδ|2dx = λi Z Dδ0 pu2δdx + Z Sr+ ϕi∂uδ ∂ν dS. (27)

From the compactness of the embedding H1(D

δ0) ,→ L 2(D δ0) we have that uδ → ϕ  i in L2(Dδ0) and so RD δ0p|uδ| 2 dx → RD δ0p|ϕ  i| 2

dx. Moreover, from the equation (24), we have that ∇uδ * ∇ϕi

weakly in H(div, Dδ0) as δ → 0. Hence classical trace theorems for vector functions yield

Z Sr+ ∂uδ ∂ν ϕ  idS → Z Sr+ ∂ϕ i ∂ν ϕ  idS, as δ → 0.

Therefore, from (27) and from the equation satisfied by ϕ

i, we conclude that, along a subsequence,

lim δ→0 Z Dδ0 |∇uδ|2dx = λi Z Dδ0 p|ϕi|2dx + Z S+r ∂ϕi ∂ν ϕ  idS = Z Ω r |∇ϕ i| 2 dx = Z Dδ0 |∇ϕ i| 2 dx.

Thanks to Urysohn’s Subsequence Principle the proof is thereby complete. Theorem 3.5. Let uδ ∈ H1(Dδ) be the unique solution to (24). Then

∇uδ −→ ∇ϕi in L 2(S+

r), as δ → 0.

Proof. By classical elliptic regularity theory, it is easy to prove that an odd reflection of uδ through

the hyperplane {x1 = 0} in a neighbourhood of {x : |x| = r} converges to ϕi in H

2, as δ → 0.

Hence the conclusion follows by trace embeddings.

Theorem 3.6. Let uδ be the unique solution to (24). Then the following identity holds

Z S+ r |∇uδ| 2 dS − N − 2 r Z Dδ |∇uδ| 2 dx =1 r Z ∂Dδ\S+r  ∂uδ ∂ν 2 x · ν dS + 2 Z Sr+  ∂uδ ∂ν 2 dS +2λ  i r Z Dδ p uδ∇uδ· x dx.

Proof. We first observe that, by classical regularity theory, uδ ∈ H2(Dδ) since Dδ verifies an

exterior ball condition. Let us now test equation (24) with the function ∇uδ · x ∈ H1(Dδ).

Integrating by parts and using the following identity ∇uδ· ∇(∇uδ· x) = 1 2div  |∇uδ| 2 x−N − 2 2 |∇uδ| 2

(11)

Proof of Proposition 3.1. Let uδ be the unique solution to (24). From Theorem 3.6 and Remark 3.2 we know that Z Sr+ |∇uδ| 2 dS −N − 2 r Z Dδ |∇uδ| 2 dx ≥ 2 Z Sr+  ∂uδ ∂ν 2 dS +2λ  i r Z Dδ p uδ∇uδ· x dx.

Now, thanks to Theorems 3.4 and 3.5, we can pass to the limit as δ → 0 in the above inequality, thus obtaining (22).

Remark 3.7. We observe that the assumption of C2-regularity for ∂Σ can be relaxed; indeed, if

Σ is less regular (e.g. if ∂Σ is Lipschitz continuous), we can approximate Σ with a class of C2

-regular domains (Σ)β and start the procedure of Section 3 from the domain (Σ)β.

4

Poincar´

e-Type Inequalities

In this section we consider the following spaces for  ∈ (0, 1] and r > : V(Br+) :=u ∈ H

1(B+

r) : u = 0 on Cr\ Σ , V0(Br+) := {u ∈ H 1(B+

r) : u = 0 on Cr}.

We point out that, for 0 ≤ 1≤ 2< r,

H01(B+r) ⊆ V0(B+r) ⊆ V1(B + r) ⊆ V2(B + r) ⊆ H 1(B+ r). (28)

Lemma 4.1 (Poincar´e-Type Inequality). Let r > 0. Then, for every u ∈ H1(Br+), the following

inequality holds N − 1 r2 Z B+r |u|2dx ≤ Z B+r |∇u|2dx +1 r Z Sr+ |u|2dS. (29)

Proof. Integrating the equality div(u2x) = 2u∇u · x + N u2 over B+

r and recalling the elementary

inequality 0 ≤ (u + ∇u · x)2= |u|2+ |∇u · x|2+ 2u∇u · x, we obtain that Z ∂Br+ |u|2x · ν dS = Z B+r  2u∇u · x + N |u|2dx ≥ − Z B+r  |u|2+ |∇u · x|2dx + N Z Br+ |u|2dx. Since x · ν = 0 on Crand |x| ≤ r, then

r Z S+ r |u|2dS ≥ −r2 Z B+ r |∇u|2dx + (N − 1) Z B+ r |u|2dx. Reorganizing the terms and dividing by r2yields the thesis.

Lemma 4.2. For 0 ≤ σ < 1 the infimum

mσ= inf u∈Vσ(B1+) u6=0 R B+ 1 |∇u|2dx R S+ 1 |u|2dS

is achieved. Moreover mσ > 0, the map σ 7→ mσ is non-increasing in [0, 1) and continuous in 0

and m0= 1.

Proof. For u ∈ Vσ(B1+), let us denote

F (u) := R B+1|∇u| 2 dx R S1+|u| 2 dS . Let {un}n ⊆ Vσ(B+1) be a minimizing sequence such that

R

S1+|un| 2

dS = 1. From (29) it follows that {un}nis bounded in H1(B1+) and so there exists ˜u ∈ H

1(B+

1) such that, up to a subsequence,

un * ˜u in H1(B1+). Taking into account the compact embedding H 1(B+

1) ,→ L 2(∂B+

(12)

that R

S1+|˜u| 2

dS = 1 and then ˜u 6= 0. Moreover ˜u = 0 on Cr\ σΣ, since {un} do; in particular

˜

u ∈ Vσ(B1+). By weak lower semicontinuity, we have that

Z B+1 |∇˜u|2dx ≤ lim inf n→+∞ Z B1+ |∇un| 2 dx = mσ.

Then mσ = F (˜u), i.e. ˜u attains the infimum mσ. Trivially mσ > 0, due to the null boundary

conditions on Cr\ σΣ. The monotonicity of the map σ 7→ mσ follows from (28).

Now we have to prove continuity. Let σn → 0+. For every n there exists ˜un ∈ Vσn(B

+ 1) such that Z S+1 |˜un| 2 dS = 1, Z B1+ |∇˜un| 2 dx = mσn.

Then, since mσn ≤ m0 for all n, we have that {˜un}n is bounded in H

1(B+

1). Thus there exists

u0∈ H1(B+1) sucht that, up to a subsequence, ˜un* u0 weakly in H1(B1+). So, first

Z S+ 1 |u0| 2 dS = 1 and u0= 0 on Cr. Furthermore m0≤ Z B1+ |∇u0| 2

dx ≤ lim inf mσn≤ lim sup mσn≤ m0.

Then, along the subsequence, m0 = limn→+∞mσn. Thanks to Urysohn’s Subsequence Principle

we may conclude that m0= limσ→0+mσ.

Finally we prove that m0= 1. Since the function u0achieving m0is harmonic in B1+, thanks to

classical monotonicity arguments (we refer to [21] for further details) we can say that the function

r 7−→ M (r) := r R B+r|∇u0| 2 dx R S+r|u0| 2 dS

is non-decreasing and that there exists k ∈ N, k ≥ 1, such that limr→0+M (r) = k. Hence M (r) ≥ 1

for every r ≥ 0. Then

m0= R B1+|∇u0| 2 dx R S+1|u0| 2 dS = M (1) ≥ 1.

Furthermore the function v(x) = x1 belongs to V0(B1+) and F (v) = 1; hence m0= 1.

Corollary 4.3. Let  ∈ (0, 1] and r > . Then m/r r Z Sr+ |u|2dS ≤ Z Br+

|∇u|2dx for all u ∈ V(B+r).

Moreover, for every ρ ∈ (0, 1) there exists µρ> 1 such that, if  < µrρ, then

1 − ρ r Z Sr+ |u|2dS ≤ Z B+r

|∇u|2dx for all u ∈ V(Br+). (30)

Proof. If we let σ = /r in the previous Lemma, we have that m/r Z S+ 1 |u|2dS ≤ Z B+ 1

|∇u|2dx for all u ∈ V/r(B1+).

The first inequality follows by the change of variables y = rx, while the second one trivially comes from the continuity of mσ in 0.

From (29) and Corollary 4.3 one can easily prove the following corollary.

Corollary 4.4. For every ρ ∈ (0, 1) there exists µρ> 1 such that, for every r > 0 and  < µrρ,

(13)

5

Monotonicity Formula

For any 0 <  < r ≤ Rmax, λ > 0, ϕ ∈ H1(Ωr), let us introduce the functions

E(ϕ, r, λ, ) := 1 rN −2 Z Ω r  |∇ϕ|2− λp|ϕ|2dx and H(ϕ, r) := 1 rN −1 Z S+ r |ϕ|2dS. Moreover we define the Almgren type frequency function

N (ϕ, r, λ, ) := E(ϕ, r, λ, ) H(ϕ, r) .

Lemma 5.1 (Integration on the Tube). There exists a constant κ = κ(N, Σ) > 0, depending only on N and |Σ|, such that, for every  ∈ (0, 1] and for every u ∈ H1(T

) such that u = 0 on ∂T\ Σ, Z T |u|2dx ≤ κ2(N −1)/N Z T |∇u|2dx.

Proof. Let ˜T = T∪ ς(T), where ς is the reflection through the hyperplane {x1 = 0}, and let ˜u

be the even extension of u on ˜T. Since ˜u ∈ H01( ˜T), thanks to (16) we have that

Z T |u|2dx =1 2 Z ˜ T |˜u|2dx ≤CN 2 | ˜T| 2/N Z ˜ T |∇˜u|2dx = CN22/N|Σ| 2/N 2(N −1)/N Z T |∇u|2dx. Hence we can conclude the proof letting κ = CN22/N|Σ|2/N.

Lemma 5.2. There exists 1∈ (0, 1], R1> 0, with 0 < 1< R1≤ Rmax, such that

H(ϕi, r) > 0 for all  ∈ (0, 1], for all r ∈ (, R1], for all i = 1, . . . , j.

Proof. Suppose by contradiction that for every n there exists n ∈ (0, 1], rn ∈ (n, Rmax] and

in ∈ {1, . . . , j} such that rn → 0 and H(ϕinn, rn) = 0. Let us denote νn := λ

n

in, ξn := ϕ

n

in and

Ωn := Ωrnn. From this it follows that ξn= 0 on S

+ rn and that Z Ωn |∇ξn| 2 dx = νn Z Ωn p|ξn| 2 dx ≤ λjkpk∞ Z Ωn |ξn| 2 dx. Using Lemma 5.1 when integrating on the tube we obtain

Z Tn |ξn|2dx ≤ κ2(N −1)/Nn Z Ωn |∇ξn|2dx.

Moreover (29) says that

Z B+rn |ξn| 2 dx ≤ r 2 n N − 1 Z Ωn |∇ξn| 2 dx. Then we have that

Z Ωn |∇ξn|2dx ≤ λjkpk  κ2(N −1)/Nn + r 2 n N − 1  Z Ωn |∇ξn|2dx.

Thus ξn ≡ 0 in Ωn, provided n is sufficiently large. Thanks to classical unique continuation

properties for elliptic equations it follows that ξn= 0 in Ωn, which is a contradiction.

Lemma 5.3. Let R2= min (  N − 1 λjkpk 1/2 , Rmax ) .

For every r ∈ (0, R2] there exist cr> 0 and r∈ (0, 1], with r< r, such that

(14)

Proof. We will prove the lemma for a fixed i ∈ {1, . . . , j} and take as cr the minimum among the

constants found for each i. Suppose by contradiction that for a certain r ∈ (0, R2] and for every n

(large enough) there exists n∈ (0, 1/n) such that

H(ϕn

i , r) <

1

n. (32)

We first note that, since n→ 0, then λin → λi (see [19]). Moreover

Z Ω1 |∇ϕn i | dx = Z Ωn |∇ϕn i | dx = λ n i Z Ωn p|ϕn i | 2 = λn i ≤ λj.

Hence there exists v ∈ H1

0(Ω1) such that ϕ n

i * v weakly in H 1

0(Ω1) along a subsequence. Note

that actually v ∈ H1 0(Ω) and v 6≡ 0 in Ω, since R Ωp|v| 2 dx = 1. Moreover ϕn i → v strongly in L2(Ω1) and in L2(S+

r), so that (32) implies that v = 0 on Sr+. This tells us that v weakly solves

( −∆v = λipv, in B+r,

v = 0, on ∂Br+. Testing the above equation with v we obtain thatR

B+ r|∇v| 2 dx = λiRB+ r p|v| 2

dx and then, thanks to (29) and the fact that λi≤ λj and r ≤ R2,

0 = Z B+r  |∇v|2− λip|v|2  dx ≥ N − 1 r2 − λjkpk∞  Z Br+ |v|2dx ≥ N − 1 R2 2 − λjkpk  Z Br+ |v|2dx.

Due to the initial choice of R2 we have that this last factor is positive; then v = 0 in B+r.

This, together with classical unique continuation principles, implies that v = 0 in Ω, which is a contradiction.

Let ˜ and ˜r be the constants found in Proposition 3.1.

Proposition 5.4. Let i ∈ {1, . . . , j},  ∈ (0, ˜] and r ∈ (, ˜r]. Then dE dr(ϕ  i, r, λ  i, ) ≥ 1 rN −2 " 2 Z Sr+  ∂ϕ i ∂ν 2 dS+ +2λ  i r Z Ω r p ϕi∇ϕ i· x dx − λ  i Z Sr+ p|ϕi|2dS + N − 2 r λ  i Z Ω r p|ϕi|2dx # (33) and dH dr(ϕ  i, r) = 2 rN −1 Z S+ r ϕi∂ϕ  i ∂ν dS = 2 rE(ϕ  i, r, λ  i, ). (34)

Proof. We compute the derivative dE dr = 2 − N rN −1 Z Ω r  |∇ϕ i| 2 − λ ip|ϕ  i| 2 dx + 1 rN −2 Z S+ r  |∇ϕ i| 2 − λ ip|ϕ  i| 2 dS.

Then, thanks to (22), we obtain (33). The proof of (34) follows from direct computations, the equation satisfied by ϕ

i and integration by parts.

Lemma 5.5. Let ρ ∈ (0, 1/2], µρ be as in Corollary 4.3,  ∈ (0, 1] and r ∈ (, Rmax]. If µρ< r,

then Z Ω r |u|2dx ≤ K,r1 Z Ω r |∇u|2dx for any u ∈ H1(Ω

r) such that u = 0 on ∂Ωr\ Sr+, where

(15)

Proof. Thanks to Lemma 5.1 we have an estimate about the integral on the tube, i.e. Z T |u|2dx ≤ κ2(N −1)/N Z T |∇u|2dx ≤ κ2(N −1)/N Z Ω r |∇u|2dx. On the other hand, by (31) we have that

Z Br+ |u|2dx ≤ r 2 N − 1  1 + 1 1 − ρ  Z B+r |∇u|2dx ≤ 3r 2 N − 1 Z Ω r |∇u|2dx. The conclusion follows by adding the two parts.

Lemma 5.6. Let ρ ∈ (0, 1/2), µρ be as in Corollary 4.3,  ∈ (0, 1] and r ∈ (, Rmax]. If µρ< r,

then Z Ω r |u∇u · x| dx ≤ K2 ,r Z Ω r |∇u|2dx for any u ∈ H1(Ω

r) such that u = 0 on ∂Ωr\ Sr+, where

K,r2 =√2κ(N −1)/N+ r 3 N − 1r 2 and κ is as in Lemma 5.1.

Proof. First we consider the integral over T: thanks to Cauchy-Schwarz Inequality and Lemma

5.1 we know that Z T |u∇u · x| dx dx ≤√2κ(N −1)/N Z Ω r |∇u|2dx. From the Cauchy-Schwarz Inequality and (31) it follows that

Z Br+ |u∇u · x| dx ≤ r 3 N − 1r 2 Z Ω r |∇u|2dx. Adding the two parts we conclude the proof.

Corollary 5.7. Let ρ ∈ (0, 1/2), µρ be as in Corollary 4.3,  ∈ (0, 1], r ∈ (, Rmax]. If µρ < r

then Z Ω r  |∇u|2− λ ip|u| 2 dx ≥ 1 − λikpkK,r1  Z Ω r |∇u|2dx (35) for any u ∈ H1(Ω

r) such that u = 0 on ∂Ωr\ Sr+ and for all i ∈ {1, . . . , j}. Furthermore there

exists r0≤ Rmax such that, for every r,  satisfying µρ< r ≤ r0, we have

Z Ω r |∇u|2dx ≤ 2 Z Ω r  |∇u|2− λ ip|u| 2 dx

for any u ∈ H1(Ωr) such that u = 0 on ∂Ωr\ Sr+ and for all i ∈ {1, . . . , j}.

Proof. The first statement (35) easily comes from Lemma 5.5. Besides, if we choose r0 ≤ Rmax

such that

K,r1 ≤ Kr10,r0 ≤

1 2λjkpk

, from (35), we can conclude the proof.

Lemma 5.8. Let ρ ∈ (0, 1/2) and µρ be as in Corollary 4.3. Let R1 and 1 be as in Lemma 5.2.

Then there exists τ > 0 depending only on N , λj, kpk∞ and |Σ| such that, for every  ∈ (0, 1],

r1, r2, with 0 < µρ < r1≤ r2≤ min{1, R1}, we have that

(16)

Proof. With the notation E(r) = E(ϕ i, r, λi, ), H(r) = H(ϕi, r) and dH dr(ϕ  i, r) = H0(r), we have

that, from Proposition 5.4 and Corollary 5.7 H0(r) = 2 rE(r) = 2 rN −1 Z Ω r  |∇ϕ i| 2 − λ ip|ϕ  i| 2 dx ≥ 2 rN −1(1 − λ  ikpk∞K 1 ,r) Z Ω r |∇ϕ i| 2 dx

for all µρ< r ≤ min{1, R1}. Hence, since λi ≤ λj, thanks to (30)

H0(r) ≥ 2 rN −1(1 − λjkpk∞K 1 ,r) 1 − ρ r Z Sr+ |ϕ i| 2 dS = 2(1 − ρ) r (1 − λjkpk∞K 1 ,r)H(r). So we have H0(r) H(r) ≥ 2(1 − ρ) r h 1 − τ12(N −1)/N− τ2r2 i

where τ1= λjkpkκ and τ2= λjkpkN −13 . Since  < r and r ≤ 1, if τ0= τ1+ τ2, then

(log H(r))0 =H 0(r) H(r) ≥ 2(1 − ρ) r h 1 − τ0r2(N −1)/N i ≥2(1 − ρ) r − 2τ0r 1−2 N.

Integrating from r1 to r2and letting τ := τ0N/(N − 1), we obtain

logH(ϕ  i, r2) H(ϕ i, r1) ≥ 2(1 − ρ) logr2 r1 − τ (r22(N −1)/N− r2(N −2)/N1 ) ≥ 2(1 − ρ) logr2 r1 − τ R12(N −1)/N. Taking the exponentials yields the thesis.

Hereafter let R0 := min{1, R1, R2, r0} where R1, R2, r0 are defined in Lemma 5.2, Lemma 5.3

and Corollary 5.7 respectively. Moreover let 0= min{1, ˜, 1} where ˜, 1are defined in Proposition

3.1 and Lemma 5.2 respectively.

Proposition 5.9. Let ρ ∈ (0, 1/2) and µρ be as in Corollary 4.3. Then, for every r ∈ (0, R0],

 ∈ (0, 0] such that 0 < µρ< r ≤ R0 dN dr (ϕ  i, r, λ  i, ) ≥ −f (r)N (ϕ  i, r, λ  i, ) for all i ∈ {1, . . . , j}, where f (r) = c1r + c2r(N −2)/N + c3r−1/N

and cn’s are positive constants depending only on ρ, kpk∞, λj, the dimension N and the geometry

of the problem (in particular on Ω and on |Σ|N −1). Proof. With the usual notation

dN dr (ϕ  i, r, λ  i, ) =: N 0(r), dE dr(ϕ  i, r, λ  i, ) := E 0(r), dH dr(ϕ  i, r) := H 0(r),

from Proposition 5.4 we have that

N0(r) ≥ 1 H2 2 r2N −3 (" Z Sr+  ∂ϕ i ∂ν 2 dS  Z S+r |ϕi| 2 dS  − Z S+r ϕi∂ϕ  i ∂ν dS 2# + " λ i r Z Ω r p ϕi∇ϕ i· x dx − λ i 2 Z Sr+ p|ϕi|2dS +N − 2 2r λ  i Z Ω r p|ϕi|2dx # Z S+r |ϕ i| 2 dS ) .

By Cauchy-Schwarz Inequality we have that

(17)

Thanks to Lemmas 5.5, 5.6, Corollary 4.3 and Corollary 5.7 we can say that N0(r) ≥ −2λ  ikpk∞ R Sr+|ϕ  i| 2  K,r2 +(N − 2) 2 K 1 ,r+ r2 2(1 − ρ)  Z Ω r |∇ϕ i| 2 dx ≥ − 4λ  ikpk∞ rN −1H(r)r N −2E(r)  K,r2 +(N − 2) 2 K 1 ,r+ r 2  . Taking into account that Kn

,r< Kr,rn , n = 1, 2, we have

N0(r) ≥ −c

1r + c2r(N −2)/N + c3r−1/N



N (r) = −f (r)N (r) by some constants c1, c2, c3> 0 independent of r and .

Corollary 5.10. Let ρ ∈ (0,12) and µρ be as in Corollary 4.3. Then, for every µ > µρ, r ∈ (0, R0],

and  ∈ (0, 0] such that µ ≤ r ≤ R0, we have that

N (ϕ i, r, λ  i, ) ≤ e RR0 r f (t) dtN (ϕ i, R0, λi, ).

Proof. Form Proposition 5.9 it follows that e−RrR0f (t) dtN (r)0 ≥ 0, which, by integration over

(r, R0), yields the conclusion.

5.1

Energy Estimates

Proposition 5.11. Let ρ ∈ (0, 1/2). Then there exists Kρ> 0 such that, for every R ≥ Kρ and

for every i ∈ {1, . . . , j}, we have Z Ω R |∇ϕ i| 2 dx = O(N −2H(ϕi, Kρ)) as  → 0+, (36) Z Ω R |ϕ i| 2 dx = O(N −N2H(ϕ i, Kρ)) as  → 0+, (37) Z S+R |ϕ i| 2 dS = O(N −1H(ϕi, Kρ)) as  → 0+. (38)

Proof. For ρ ∈ (0, 1/2) let us consider µρ as in Corollary 4.3, 0= min{1, ˜, 1}, R0 as in Lemma

5.3 and let Kρ > max{µρ, R0/0, R0/R0}. From Corollary 5.10 we deduce that, if R ≥ Kρ and

R < R0 N (ϕ i, R, λi, ) ≤ e RR0 R f (t) dtN (ϕ i, R0, λi, ). (39)

Now let us analyze the frequency function N at radius R0:

E(ϕi, R0, λi, ) = 1 RN −20 Z Ω R0  |∇ϕi| 2 − λi|ϕ  i| 2 dx ≤ 1 RN −20 Z Ω |∇ϕi| 2 dx ≤ λj RN −20 . Moreover, thanks to Lemma 5.3

H(ϕi, R0) ≥ cR0.

Thus we have that

N (ϕi, R0, λi, ) ≤ λj cR0R N −2 0 . (40) Then, from (39) Z Ω R  |∇ϕ i| 2 − λ i|ϕi| 2 dx ≤ const H(ϕi, R)(R)N −2. (41)

(18)

Now let Kρ ≤ r ≤ R0. Then, from Proposition 5.4 H0(ϕ i, r) H(ϕ i, r) = 2 rN (ϕ  i, r, λ  i, )

and from Corollary 5.10 and (40)

H0(ϕi, r) H(ϕ

i, r)

≤ C

r. (43)

Now, integrating the previous inequality from Kρ to R, we obtain

log H(ϕ  i, R) H(ϕ i, Kρ) ≤ C log R Kρ , hence H(ϕ

i, R) ≤ constρ,RH(ϕi, Kρ), i.e. H(ϕi, R) = O(H(ϕi, Kρ)) as  → 0. Then (36)

follows from (42), whereas (38) is a direct consequence of the previous estimate and definition of H. Finally, thanks to Lemma 5.5 and (36), we have

Z Ω R |ϕ i| 2 dx ≤ (c12(N −1)/N + c2(R)2) Z Ω R |∇ϕ i| dx = O(N − 2 NH(ϕ i, Kρ)), as  → 0, thus proving (37).

Proposition 5.12. Let ρ ∈ (0, 1/2) and Kρ be as in Proposition 5.11. Then there exists Cρ > 0

such that, for every R ≥ Kρ, for every  ∈ (0, 0] such that R ≤ R0, and for every i ∈ {1, . . . , j}

we have Z Ω R |∇ϕ i| 2 dx ≤ Cρ(R)N −2ρ, Z Ω R |ϕi| 2 dx ≤ Cρ(R)N +2−2ρ−2/N, Z SR+ |ϕ i| 2 dx ≤ Cρ(R)N +1−2ρ.

Proof. From Lemma 5.8 we know that

H(ϕi, R) ≤ expτ R12(N −1)/N R R0

2(1−ρ)

H(ϕi, R0) (44)

and, from (30), we have H(ϕi, R0) = 1 RN −10 Z S+ R0 |ϕ i| 2 dS ≤ 1 RN −20 (1 − ρ) Z Ω R0 |∇ϕ i| 2 dx ≤ λj RN −20 (1 − ρ). (45) Combining (44) and (45) with Proposition 5.11 (in particular estimate (42) in the proof) we can deduce all the claims.

As a consequence of Proposition 5.12 we can say that

H(ϕi, Kρ) = O(2−2ρ) as  → 0. (46)

As a byproduct of the proof of Proposition 5.11 we obtain the following result.

Corollary 5.13. Let ρ ∈ (0, 1/2) and Kρ be as in Proposition 5.11. Then there exist ¯C, q > 0

such that, if  ∈ (0, 0] and Kρ < R0,

(19)

Proof. If we integrate (43) over (Kρ, R0) and take the exponentials, we obtain H(ϕj, R0) H(ϕ j, Kρ) ≤ R0 Kρ q ,

denoting by q the constant C in (43). Then Lemma 5.3 implies that H(ϕj, Kρ) ≥ cR0

 Kρ

R0

q . Hence the claim is proved with ¯C := cR0(Kρ/R0)

q.

6

Estimates on the Difference λ

j

− λ

j

6.1

Upper Bound

For any i ∈ {0, . . . , j}, R > 1 and  ∈ (0, 1], with R ≤ Rmax, let us consider the following

minimization problem min ( Z BR+ |∇u|2dx : u ∈ H1(B+R), u = 0 on CR, u = ϕi on S + R ) . One can prove that this problem has a unique solution vint

i,R,, which weakly solves

     −∆vint i,R, = 0, in B + R, vi,R,int = 0, on CR, vi,R,int = ϕi, on SR+.

Proposition 6.1. Let ρ ∈ (0, 1/2) and Kρ be as in Proposition 5.11. Then

Z BR+ ∇vint i,R, 2 dx = O(N −2H(ϕi, Kρ)) as  → 0+, (48) Z BR+ vi,R,int 2 dx = O(NH(ϕi, Kρ)) as  → 0+, (49) Z S+R vinti,R, 2 dx = O(N −1H(ϕi, Kρ)) as  → 0+. (50)

for all R ≥ 2 and for any i = 1, . . . , j. Moreover there exists ˆCρ such that, if R ≥ max{2, Kρ} and

 < R0/R, Z BR+ ∇vint i,R, 2 dx ≤ ˆCρ(R)N −2ρ, (51) Z BR+ vi,R,int 2 dx ≤ ˆCρ(R)N +2−2ρ, (52) Z SR+ vi,R,int 2 dx ≤ ˆCρ(R)N +1−2ρ. (53)

Proof. Proving (50) is trivial due to (38), since vint i,R,= ϕ

 i on S

+

R. Let η = ηR(·), with ηRdefined

(20)

where the last step comes from (31). Combining this inequality with (36) we obtain (48). More-over (31) and (48) yield (49). Finally estimates (51)–(53) follow from the above argument and Proposition 5.12.

Now let us define, for all i ∈ {1, . . . , j}, for all R > 1 and  ∈ (0, 1] such that R ≤ Rmax,

vi,R, := ( vi,R,int , in BR+, ϕi, in Ω \ BR+, (54) and ZR(x) := v int j,R,(x) q H(ϕ j, Kρ) , ϕ˜(x) := ϕ  j(x) q H(ϕ j, Kρ) . (55)

It is easy to prove that the family of functions {v1,R,, . . . , vj,R,} is linearly independent in H01(Ω).

Lemma 6.2. For all R ≥ max{2, Kρ}, we have that, as  → 0+,

Z Ω |∇vj,R,| 2 dx = λj+ N −2H(ϕj, Kρ) Z B+ R |∇Z R| 2 dx − Z ΠR |∇ ˜ϕ|2dx ! , (56) Z Ω |∇vi,R,| 2

dx = λi+ O(N −2ρ) for all i ∈ {1, . . . , j}, (57)

Z Ω ∇vi,R,· ∇vj,R,dx = O  N −1−ρqH(ϕ j, Kρ)  for all i ∈ {1, . . . , j − 1}, (58) Z Ω

∇vi,R,· ∇vn,R,dx = O(N −2ρ) for all i, n ∈ {1, . . . , j}, i 6= n, (59)

Z Ω p|vj,R,| 2 dx = 1 + O(N −2/NH(ϕj, Kρ)), (60) Z Ω p|vi,R,| 2

dx = 1 + O(N +2−2ρ−2/N) for all i ∈ {1, . . . , j}, (61) Z Ω pvi,R,vj,R,dx = O  N +1−ρ−2/NqH(ϕ j, Kρ)  for all i ∈ {1, . . . , j − 1}, (62) Z Ω

pvi,R,vn,R,dx = O(N +2−2ρ−2/N) for all i, n ∈ {1, . . . , j}, i 6= n, (63)

where, in (56), ˜ϕ has been trivially extended in Π

R outside its domain.

Proof. We will only prove the first part of the estimates, i.e. (56), (57), (58), (59), since the second part is completely analogous. To prove (56) we observe that, by scaling,

Z Ω |∇vj,R,| 2 dx = Z B+ R ∇vint j,R, 2 dx + Z Ω ∇ϕ j 2 dx − Z Ω R ∇ϕ j 2 dx = λj+ N −2H(ϕj, Kρ) Z BR+ |∇Z R| 2 dx − Z ΠR |∇ ˜ϕ|2dx ! .

Thanks to Propositions 5.12 and 6.1 we have that Z Ω |∇vi,R,| 2 dx = Z B+ R ∇vint i,R, 2 dx + Z Ω |∇ϕ i| 2 dx − Z Ω R |∇ϕ i| 2 dx = λi + O(N −2ρ),

(21)

as  → 0+, thus proving (58). Similarly, for i 6= n Z Ω ∇vi,R,· ∇vn,R,= Z B+ R ∇vint i,R,∇vintn,R,dx + Z Ω ∇ϕ i· ∇ϕndx − Z Ω R ∇ϕ i· ∇ϕndx = O(N −2ρ), as  → 0+, which provides (59).

We construct a basis {ˆv1,R,, . . . , ˆvj,R,} of the space span {v1,R,, . . . , vj,R,} such that

Z Ω p ˆvn,R,ˆvm,R,dx = 0 for n 6= m, by defining ˆ vj,R,= vj,R,, vˆi,R, = vi,R,− j X n=i+1

di,nvˆn,R,, for all i = 1, . . . , j − 1 ,

where di,n= R Ωp vi,R,ˆvn,R,dx R Ωp|ˆvn,R,| 2 dx .

Using the estimates established in Lemma 6.2, one can prove the following Z Ω |∇ˆvj,R,| 2 dx = λj+ N −2H(ϕj, Kρ) Z B+R |∇Z R| 2 dx − Z ΠR |∇ ˜ϕ|2dx ! , (64) Z Ω |∇ˆvi,R,| 2

dx = λi+ O(N −2ρ) for all i ∈ {1, . . . , j}, (65) Z Ω ∇ˆvj,R,· ∇ˆvi,R,dx = O(N −1−ρ q H(ϕ j, Kρ)) for all i ∈ {1, . . . , j − 1}, (66) Z Ω

∇ˆvi,R,· ∇ˆvm,R,dx = O(N −2ρ) for all i, m ∈ {1, . . . , j}, i 6= m, (67)

Z Ω p|ˆvj,R,| 2 dx = 1 + O(N −2/NH(ϕj, Kρ)), (68) Z Ω p|ˆvi,R,| 2

dx = 1 + O(N +2−2ρ−2/N) for all i ∈ {1, . . . , j}, (69) as  → 0.

Proposition 6.3. Let ρ ∈ (0, 1/2), Kρ as defined in Proposition 5.11 and R ≥ Kρ. For  < R0/R

there exists fR() such that

λj− λj ≤  N −2H(ϕ j, Kρ)(fR() + o(1)) as  → 0 and fR() = Z B+R |∇Z R| 2 dx − Z ΠR |∇ ˜ϕ|2dx, where ˜ϕ has been trivially extended in Π

R outside its domain.

Proof. By the Courant-Fischer Min-Max characterization of eigenvalues

λj = min      max α1,...,αj∈R Pj i=1|αi|2=1 R Ω ∇  Pj i=1αiui  2 dx R Ωp Pj i=1αiui 2 dx : {u1, . . . , uj} ⊆ H01(Ω) linearly independent      .

(22)

we obtain that λj− λj ≤ max α1,...,αj∈R Pj i=1|αi|2=1 Z Ω ∇  j X i=1 αiv˜i,R,  2 dx − λj= max α1,...,αj∈R Pj i=1|αi|2=1 j X i,n=1 Mi,n αiαn where Mi,n = R Ω∇ˆvi,R,· ∇ˆvn,R,dx R Ωp|ˆvi,R,| 2 dx 1/2R Ωp|ˆvn,R,| 2 dx 1/2 − λ  jδ n i, with δn

i denoting the usual Kronecker delta, i.e. δin = 0 for i 6= n and δin = 1 for i = n. From

estimates (64)–(69) one can derive the following estimates Mj,j() = N −2H(ϕj, Kρ)(fR() + O(2−2/N)), Mi,j() = O  N −1−ρqH(ϕ j, Kρ) 

and Mi,i() = λi− λj+ o(1) for all i < j,

Mi,n() = O(N −2ρ) for all i, n < j, i 6= n,

as  → 0. Moreover, from Corollary 5.13, we know that H(ϕj, Kρ) ≥ ¯Cq for some ¯C, q > 0.

Therefore, taking also into account (46) and the fact that fR() = O(1) as  → 0 in view of (36)

and (48), the hypotheses of Lemma 9.3 are satisfied with σ() = N −2H(ϕj, Kρ), µ() = fR() + o(1), α =

N

2 − ρ, M > (2ρ − 2 + q) 2 N − 2ρ. The proof is thereby complete.

6.2

Lower Bound

For any R > 1 and  ∈ (0, 1], with R ≤ Rmax, let us consider the following minimization problem

min ( Z Ω R |∇u|2dx : u ∈ H1(ΩR), u = 0 on ∂ΩR\ S + R, u = ϕj on SR+ ) . (70)

One can prove that this problem has a unique solution wint

j,R,, which weakly verifies

     −∆wint j,R,= 0, in Ω  R, wj,R,int = 0, on ∂ΩR\ S+ R, wj,R,int = ϕj, on SR+. Let us define wj,R,:= ( wintj,R,, in ΩR, ϕj, in Ω \ BR+.

Lemma 6.4. There exists ˜C > 0 such that, for all i ∈ {1, . . . , j − 1}, for all R > 1 and  ∈ (0, 1], with R ≤ Rmax, Z B+R |∇ϕi| 2 dx ≤ ˜C(R)N, Z BR+ |ϕi| 2 dx ≤ ˜C(R)N +2, Z SR+ |ϕi| 2 dx ≤ ˜C(R)N +1, and Z B+R |∇ϕj| 2 dx ≤ ˜C(R)N +2k−2, Z BR+ |ϕj| 2 dx ≤ ˜C(R)N +2k, Z S+R |ϕj| 2 dx ≤ ˜C(R)N +2k−1.

(23)

Lemma 6.5. There exists ˆC > 0 such that, for all R > 1 and  ∈ (0, 1], with R ≤ Rmax, Z Ω R ∇wint j,R, 2 dx ≤ ˆC(R)N +2k−2, (71) Z SR+ wj,R,int 2 dx ≤ ˆC(R)N +2k−1. (72)

Furthermore, for all R > µ1/2 and  ∈ (0, 1], with R ≤ Rmax,

Z Ω R wj,R,int 2 dx ≤ ˆC(R)N +2k−2/N. (73)

Proof. (72) is trivial since wint

j,R, = ϕj on SR+. Also (71) is simple since ϕj is an admissible test

function for (70). Finally (73) comes from (71), (72) and Lemma 5.5. Now let us define, for all R > 1 and  ∈ (0, 1] such that R ≤ Rmax,

UR(x) =

wintj,R,(x)

k , W

(x) = ϕj(x)

k . (74)

From (8), we easily deduce that

W−→ ψk in H1(BR+) as  → 0, for all R > 0,

where ψk has been defined in (10).

Lemma 6.6. We have that

UR −→ UR in HR as  → 0, for all R > 1,

where UR is defined in Lemma 2.3.

Proof. From Lemma 6.5 and from the definition of U

R we know that Z ΠR |∇U R| 2 dx = O(1) as  → 0. where U

R has been trivially extended in ΠR outside its domain. So there exists V = VR ∈ HR

such that, along a sequence  = n→ 0,

UR * V weakly in HR as  = n→ 0.

This means that

∇U R* ∇V weakly in L 2 R) as  = n→ 0. Since U R = W on S + R and W  → ψ

k in L2(S+R), then V satisfies (in a weak sense) the same

equation as UR, defined in Lemma 2.3. So, by uniqueness, V = UR. Since the limit V = UR

is the same for every subsequence, Urysohn’s Subsequence Principle implies that the convergence U

R* UR holds as  → 0 (not only along subsequences).

To prove strong convergence it is enough to show that kU

RkHR → kURkHR as  → 0. First

we notice that, trivially, −∆UR * −∆UR weakly in L2(ΠR): so, we have that ∇UR * ∇UR in

H(div, ΠR), thus ∂U R ∂ν * ∂UR ∂ν in  H001/2(SR+) ∗ as  → 0,

where (H001/2(SR+))∗ is the dual of the Lions-Magenes space H001/2(SR+). Then, since W → ψk in

(24)

It is easy to prove that the family of functions {ϕ1, ϕ2, . . . , ϕj−1, wj,R,} is linearly independent

in H1

0(Ω). As in the previous section, we construct a new basis of the space

span{ϕ1, ϕ2, . . . , ϕj−1, wj,R,} ⊆ H01(Ω )

by defining, for all i = 1, . . . , j − 1

ˆ wi,R, = ϕi and ˆ wj,R,= wj,R,− j−1 X i=1 ciϕi, where ci= Z Ω pwj,R,ϕidx.

In this way we have thatR

Ωp ˆwn,R,wˆm,R,dx = 0 if n 6= m.

Using the estimates established in Lemmas 6.4 and 6.5, one can prove the following Z Ω |∇ ˆwj,R,|2dx = λj+ N +2k−2 Z ΠR |∇U R| 2 dx − Z BR+ |∇W|2 dx + o(1) ! , (75) Z Ω

∇ ˆwj,R,· ∇ ˆwi,R,dx = O(N +k−1) for all i ∈ {1, . . . , j − 1}, (76)

Z

Ω

p| ˆwj,R,|2dx = 1 + O(N +2k−2/N), (77)

as  → 0.

Proposition 6.7. Let ρ ∈ (0, 1/2), Kρ as defined in Proposition 5.11 and R ≥ Kρ. For  < R0/R

there exists gR() such that

λj− λj ≤ N +2k−2(gR() + o(1)) as  → 0 and gR() = Z ΠR |∇U R| 2 dx − Z B+R |∇W|2 dx, where U

R has been trivially extended in ΠR outside its domain.

Proof. By the Courant-Fischer Min-Max characterization of eigenvalues

λj= min      max α1,...,αj∈R Pj i=1|αi|2=1 R Ω ∇ Pj i=1αiui  2 dx R Ωp P j i=1αiui 2 dx : {u1, . . . , uj} ⊆ H01(Ω ) linearly independent      .

(25)

with δn

i denoting the usual Kronecker delta. From estimates (75)–(77) it follows that

Lj,j = N +2k−2(gR() + o(1)), Li,j = O(

N +k−1) for all i < j,

Li,i= λi− λj for all i < j, Li,n= 0 for all i, n < j, i 6= n,

as  → 0. Therefore, taking into account that gR() = O(1) as  → 0 in view of Lemma 6.4 and

(71), the hypotheses of Lemma 9.3 are satisfied with

σ() = N +2k−2, µ() = gR() + o(1), α =

N

2 , M >

4(k − 1)

N .

The proof is thereby complete.

From the fact that W→ ψk in H1(BR+), as  → 0, for all R > 0, and from Lemma 6.6 we can

deduce the following result.

Lemma 6.8. For all R > 1 we have that

gR() −→ gR as  → 0, where gR:= Z ΠR |∇UR|2dx − Z BR+ |∇ψk|2dx. (78)

In order to compute the limit limR→∞gR, we introduce the functions

ζ(r) := Z S+ 1 Φ(rθ)Ψ(θ) dS(θ) for r ≥ 1, (79) χR(r) := Z S1+ UR(rθ)Ψ(θ) dS(θ) for 1 ≤ r ≤ R. (80) Moreover, we denote γN := Z S1+ |Ψ(θ)|2dS(θ). (81)

We immediatly notice, thanks to Lemma 2.4 and to the embedding H1(B+

1) ,→ L2(S + 1), that

ζ(1) = lim

R→+∞χR(1). (82)

Lemma 6.9. Let ζ be the function defined in (79), γN the constant defined in (81) and mk(Σ)

the one defined in (11). Then

ζ(1) = γN −

2mk(Σ)

N + 2k − 2. (83)

Proof. From the definition of Φ, given in (20), one can easily prove that ζ satisfies the following ODE rN +2k−1(r−kζ(r))00 = 0 in (1, +∞). This yields r−kζ(r) = ζ(1) + C1 − r −N −2k+2 N + 2k − 2 (84)

for some constant C ∈ R. Now we note that r−kζ(r) → γN as r → +∞. Indeed, since Φ = wk+ψk,

we can rewrite

ζ(r) = Z

S+1

wk(rθ)Ψ(θ) dS(θ) + γNrk.

By evaluating the vanishing order at 0 of the Kelvin transform of the restriction of the function wk on Π \ Π1, we can prove that

|wk(x)| ≤ const|x| 1−N

(26)

Hence, when r → +∞ r−kζ(r) − γN ≤ Z S+1 |wk(rθ)| rk |Ψ(θ)| dS(θ) ≤ const r 1−N −k→ 0.

Then we can find the constant C in (84), letting r → +∞; so we can rewrite ζ as

ζ(r) = γNrk+ (ζ(1) − γN)r−N −k+2 in (1, +∞). (85)

Taking the derivative leads to

ζ0(r) = kγNrk−1+ (N + k − 2)(γN− ζ(1))r−N −k+1

= (N + 2k − 2)γNrk−1−

(N + k − 2)ζ(r)

r .

(86)

Hence, taking into account the definition of ζ and evaluating its derivative at r = 1, we obtain Z

S+1

∂Φ

∂ν(θ)Ψ(θ) dS(θ) = (N + 2k − 2)γN − (N + k − 2)ζ(1). (87) Since −∆Φ = 0 in B+1, multiplying this equation by ψk and integrating by parts we obtain that

Z B1+ ∇Φ · ∇ψkdx = Z S+1 ∂Φ ∂νψkdS = Z S+1 ∂Φ ∂νΨ dS. (88)

Then, let us test the equation −∆ψk = 0 with Φ. From (12) and (21) it follows that

Z B+ 1 ∇ψk· ∇Φ dx = Z S+ 1 ∂ψk ∂ν Φ dS − Z Σ ∂ψk ∂x1 Φ dS = Z S+ 1 ∂ψk ∂ν Φ dS − Z Σ ∂ψk ∂x1 wkdS = Z S1+ ∂ψk ∂ν Φ dS + 2mk(Σ). (89)

Moreover we note that

∂ψk

∂ν (θ) = kΨ(θ) on S

+

1. (90)

Then, from (89) and (90) we obtain Z B+ 1 ∇ψk· ∇Φ dx = k Z S+ 1 ΦΨ dS + 2mk(Σ) = kζ(1) + 2mk(Σ). (91)

Finally, combining (87), (88) and (91) leads to the thesis.

Lemma 6.10. Let gR be as defined in (78) and mk(Σ) as in (11). Then limR→+∞gR= 2mk(Σ).

Proof. Integrating by parts we have that gR= Z S+R  ∂UR ∂ν − ∂ψk ∂ν  ψkdS.

If χR is the function defined in (80), then

χ0R(r) = Z

S+1

∂UR

∂ν (rθ)Ψ(θ) dS(θ) and, by a change of variable,

χ0R(r) = r1−N −k Z

Sr+

∂UR

∂ν ψkdS. (92)

By simple computations one can prove that χR solves

rN +2k−1(r−kχR(r))0 0

(27)

By integration, we arrive at

r−kχR(r) = χR(1) + C

1 − r−N −2k+2

N + 2k − 2 . (93)

From the fact that UR = RKΨ on S+R, we have that χR(R) = RkγN, and this allows us to know

the constant C. After some computations, the expression (93) then becomes as follows r−kχR(r) = χR(1) + (γN− χR(1)) 1 − r−N −2k+2 1 − R−N −2k+2, r ∈ (1, R). From (92), we get Z SR+ ∂UR ∂ν ψkdS = χ 0 R(R)R N +k−1 = [γN(N + k − 2) − χR(1)(N + 2k − 2)]R −N −k+1+ kγ NRk−1 1 − R−N −2k+2 R N +k−1 = γN(N + k − 2) − χR(1)(N + 2k − 2) + kγNR N +2k−2 1 − R−N −2k+2 . (94)

For what concerns the second part of gR, it is easy to see that

∂ψk ∂ν (rθ) = kr k−1Ψ(θ). Therefore Z S+R ∂ψk ∂ν ψkdS = Z S1+ kRN +2k−2|Ψ|2dS(θ) = kRN +2k−2γN. (95)

Finally, combining (82), (94), (95) and Lemma 6.9 and taking the limit when R → +∞ we reach the conclusion.

Combining Propositions 6.3 and 6.7 with Lemmas 6.8 and 6.10 we obtain the following upper-lower estimate for the eigenvalue variation.

Proposition 6.11. Let ρ ∈ (0, 1/2), Kρ as defined in Proposition 5.11 and mk(Σ) as in (11).

Then, for all R ≥ Kρ, we have that, as  → 0,

−2mk(Σ) + o(1) ≤ λj− λj N +2k−2 ≤ H(ϕ j, Kρ) 2k (fR() + o(1)).

Since −2mk(Σ) > 0, as a direct consequence of Proposition 6.11 we obtain the following estimate

from below for H(ϕ j, Kρ).

Corollary 6.12. We have that 2k

H(ϕ j, Kρ)

= O(1) as  → 0.

7

Blow-up Analysis

Let us introduce the functional

F : R × H01(Ω) −→ R × H −1(Ω) (λ, ϕ) 7−→ (kϕk2H1 0(Ω)− λj, −∆ϕ − λpϕ) where kϕk2 H1 0(Ω) =R Ω|∇ϕ| 2dx and H−1(Ω)h−∆ϕ − λpϕ, viH1 0(Ω)= Z Ω (∇ϕ · ∇v − λpϕv) dx.

From the assumptions we know that F (λj, ϕj) = (0, 0). Moreover, from the simplicity assumption

(28)

Lemma 7.1. The functional F is differentiable at (λj, ϕj) and its differential dF (λj, ϕj) : R × H01(Ω) −→ R × H −1(Ω) dF (λj, ϕj)(λ, ϕ) =  2 Z Ω ∇ϕj· ∇ϕ dx, −∆ϕ − λpϕj− λjpϕ  is invertible.

Lemma 7.2. Let ρ ∈ (0, 1/2), Kρ as defined in Proposition 5.11 and R ≥ Kρ. Then, when  → 0,

vj,R,−→ ϕj in H01(Ω),

where vj,R, is defined in (54).

Proof. First note that Z Ω |∇(vj,R,− ϕj)| 2 dx = Z Ω ∇(ϕj− ϕj) 2 dx − Z Ω R ∇(ϕ j− ϕj) 2 dx + Z BR+ ∇(vint j,R,− ϕj) 2 dx.

The first term tends to zero because of (6). For the second and the third term we can exploit the energy estimates in Proposition 5.12, Lemma 6.4 and Proposition 6.1 to conclude.

Lemma 7.3. Let ρ ∈ (0, 1/2), Kρ as defined in Proposition 5.11 and R ≥ Kρ. Then

kvj,R,− ϕjkH1 0(Ω) = ON/2−1qH(ϕ j, Kρ)  as  → 0 and, in particular, Z Ω\B+R ∇(ϕ j− ϕj) 2 dx = O N −2H(ϕj, Kρ) as  → 0. (96)

Proof. Taking into account Lemma 7.2 and (5), from the differentiability of the functional F it follows that F (λj, vj,R,) = dF (λj, ϕj)(λj− λj, vj,R,− ϕj) + o  λj− λj + kvj,R,− ϕjkH1 0(Ω) 

as  → 0. Now let us apply dF (λj, ϕj)−1 to both members and obtain

λj− λj + kvj,R,− ϕjkH1 0(Ω) ≤ dF (λj, ϕj)−1 L(R×H−1(Ω),R×H1 0(Ω)) F (λj, vj,R,) R×H−1(Ω)(1 + o(1)) and so kvj,R,− ϕjkH1 0(Ω)≤ C  kvj,R,k 2 H1 0(Ω)− λj + −∆vj,R,− λjpvj,R, H−1(Ω)  . (97)

Thanks to (56), Proposition 6.11, and the fact that fR() = O(1) as  → 0 in view of (36) and

(48), kvj,R,k 2 H1 0(Ω) − λj ≤ kvj,R,kH01(Ω)− λ  j + λj− λj = O(N −2H(ϕj, Kρ)). (98) Let u ∈ H1

0(Ω) be such that kukH1

(29)

So we have that Z Ω ∇vj,R,· ∇u dx − λj Z Ω pvj,R,u dx ≤ ≤ s Z B+R ∇v int j,R, 2 dx + λj Z Ω pϕju dx − Z Ω pvj,R,u dx  + s Z Ω R ∇ϕ j 2 dx. (99)

Now let us analyze the middle term Z Ω pϕju dx − Z Ω pvj,R,u dx = Z BR+ pϕju dx − Z BR+ pvj,R,int u dx ≤ ≤ const s Z B+ R ϕ j 2 dx + sZ B+ R |vint j,R,|2dx !

where we implicitly used the Poincar´e Inequality. Thanks to inequality (29) and to the energy estimates made in Proposition 5.11

Z BR+ ϕj 2 dx ≤ (R) 2 N − 1 Z B+R ∇ϕ j 2 dx + R N − 1 Z S+R |ϕ j| 2dx = O(NH(ϕ j, Kρ)) as  → 0.

Then, from (99), Proposition 5.12 and Proposition 6.1 we obtain that Z Ω ∇vj,R,· ∇u dx − λj Z Ω pvj,R,u dx = O  N/2−1qH(ϕ j, Kρ)  as  → 0 uniformly with respect to u ∈ H1

0(Ω) with kukH1 0(Ω)≤ 1 and hence −∆vj,R,− λjpvj,R, H−1(Ω)= O  N/2−1qH(ϕ j, Kρ)  as  → 0. (100)

The conclusion follows by combining (97), (98), and (100).

Corollary 7.4. Let ρ ∈ (0, 1/2), Kρ as defined in Proposition 5.11 and R ≥ Kρ. Then

Z 1 Ω\B + R ∇ ˜ϕ − kH(ϕ j, Kρ)−1/2∇W 2 dx = O(1), as  → 0,

where ˜ϕ is defined in (55) and W in (74), while 1Ω = {1x : x ∈ Ω}. Proof. It directly follows from a change of variables in (96).

The following Theorem provides a blow-up analysis for scaled eigenfunctions, which contains Theorem 1.2.

Theorem 7.5. Let ρ ∈ (0, 1/2) and Kρ as defined in Proposition 5.11. Then

(30)

Proof. Let n → 0. From Corollary 6.12 we deduce that, up to a subsequence, (n)k q H(ϕn j , Kρn) −→ c ≥ 0.

Since, in view of Proposition 5.11, { ˜ϕn} is bounded in H

R, by a diagonal process there exists ˜Φ,

with ˜Φ ∈ HR for all R > 2, and a subsequence (still denoted by n) such that

˜ ϕn * ˜Φ weakly in H R for all R > 2. (104) MoreoverRS+ Kρ | ˜ϕn|2dS = KN −1

ρ , hence, by compactness of trace embeddings,

Z

S+

| ˜Φ|2dS = KρN −1, (105)

thus implying that ˜Φ 6≡ 0.

Actually we can prove that the convergence in (104) is strong. Indeed, consider the equation solved by ˜ϕn:            −∆ ˜ϕn= ( n)2λjnp ˜ϕ n, in  1 n, 0 × Σ  ∪ BR+, ˜ ϕn= 0, on ∂ 1 n, 0 × Σ  ∪ BR+\ SR+, ˜ ϕn(x) = ϕ n j (nx) √ H(ϕnj ,Kρn) , on SR+. If we consider the restriction to BR+\ B+

R/2and the odd reflection through the hyperplane x1= 0,

we have that { ˜ϕn} is bounded in H2(B

R\ BR/2), where BR = {x ∈ RN: |x| < R}. Hence, up to a subsequence, ∂ ˜ϕ∂νn → ∂ ˜Φ ∂ν in L 2(S+ R) and therefore Z ΠR |∇ ˜ϕn|2dx = ( n)2λjn Z ΠR p| ˜ϕn|2dx + Z S+R ∂ ˜ϕn ∂ν ϕ˜ ndS → Z S+R ∂ ˜Φ ∂ν ˜ Φ dS = Z ΠR |∇ ˜Φ|2dx. Then we conclude that ˜ϕn→ ˜Φ strongly in H

R for all R > 2.

From Corollary 7.4 it follows that there exist c0 > 0 and n0∈ N such that, for all n ≥ n0 and

˜ R > R, Z B+˜ R\B + R ∇ ˜ϕ n− ( n)kH(ϕjn, Kρn)−1/2∇Wn 2 dx ≤ c0. Let us recall that Wn → ψ

k in H1(BR+˜) and (since the norms are equivalent) also in HR˜: so,

passing to the limit as n → ∞ in the above estimate, we obtain that Z B+ ˜ R\B + R |∇ ˜Φ − c∇ψk|2dx ≤ c0.

Since the constant c0 is independent on ˜R, we deduce that Z

Π

|∇ ˜Φ − c∇ψk|2dx < +∞. (106)

Moreover, the function ˜Φ satisfies the following equation ( −∆˜Φ = 0, in Π,

˜

Φ = 0, on ∂Π. (107)

We claim that c > 0. Otherwise, if c = 0 then, by (106) and (107), we could say that ˜Φ = 0, which would contradict (105).

From Proposition 2.2 we conclude that ˜Φ = c Φ and hence, in view of (105), c = Λ−1/2ρ . Since

the limit of the sequence { ˜ϕn} is the same for any choice of the subsequence, we conclude the

(31)

Corollary 7.6. For all R > 2 we have that ZR −→

1 pΛρ

ZR in H1(BR+) as  → 0.

Proof. From the definitions of the functions Z

R and ZR (in (55) and Lemma 2.5 respectively)

       −∆(pΛρZR − ZR) = 0, in BR+, pΛρZR − ZR= 0, on CR, pΛρZR − ZR=pΛρϕ˜− Φ, on SR+. So Z

R− ZR is the unique, least energy solution with these prescribed boundary conditions. Now,

let η = ηR be as defined in (14). We have that

Z BR+ ∇(pΛρZ  R− ZR) 2 dx ≤ Z B+R ∇(η(pΛρϕ˜ − Φ)) 2 dx ≤ ≤ 2 Z BR+ |∇η|2 pΛρϕ˜ − Φ 2 dx + 2 Z B+R η2 ∇(pΛρϕ˜ − Φ) 2 dx ≤ ≤ 32 R2 Z B+ R pΛρϕ˜ − Φ 2 dx + 2 Z B+ R ∇(pΛρϕ˜ − Φ) 2 dx → 0

as  → 0, thanks to (101) and to the embedding HR⊂ L2(ΠR). The conclusion follows taking into

account Poincar´e Inequality for functions vanishing on a portion of the boundary.

8

Proof of Theorem 1.1

Thanks to Theorem 7.5 and Corollary 7.6, we know that fR:= lim →0fR() = 1 Λρ Z B+ R |∇ZR| 2 dx − 1 Λρ Z ΠR |∇Φ|2dx.

Moreover, in view of Proposition 6.11 and (102), we have that, for any R > max{2, Kρ}

Ck(Σ) ≤ lim inf →0 λj− λj N +2k−2 ≤ lim sup →0 λj− λj N +2k−2 ≤ ΛρfR, (108)

where Ck(Σ) = −2mk(Σ) > 0. To complete the proof of our main result it is then enough to show

that

lim

R→+∞ΛρfR= Ck(Σ).

For every R > 2 let us define ξR(r) :=

Z

S1+

ZR(rθ)Ψ(θ) dS(θ) for 0 ≤ r ≤ R. (109)

Lemma 8.1. There holds Z S+R ∂(ZR− ψk) ∂ν (Φ − ψk) dS −→ 0 as R → +∞, (110) Z SR+ ∂(ψk− Φ) ∂ν (Φ − ψk) dS −→ 0 as R → +∞. (111)

Proof. In order to prove (110), we first take into account the equation solved by ZR− ψk, i.e.

Riferimenti

Documenti correlati

[r]

On the other hand, partitions with points of odd multiplicity can be obtained as nodal domains by minimizing a certain eigenvalue of an Aharonov-Bohm Hamiltonian with respect to

Finally, we remark that the sharp estimates obtained in [1] for a approaching an interior point along a general direction don’t make use of an explicit construction of the

Sections 4 and 5 contain a blow-up analysis (at the right and at the left junction respectively) leading to some uniform bounds of the frequency function which allow describing,

A detailed analysis in dimension 2 has been performed in [14], where Gadyl’shin investigated the case in which the perturbing set is a segment of length  → 0 + contained in

A deep insight into [17] highlights how the dependence on the subsequences in Theorem 1.1 is a priori given by two different facts: on one hand, the convergence to the limit profile

Comparing the above asymptotic expansion for the spectral fractional Laplacian with the ex- pansion derived in Theorem 1.7, we note that only in the case of the restricted

Furthermore, the new method also exhibited some advantages over the principal component regression which uses less of information contained in the sample since the number