• Non ci sono risultati.

The purpose of the present paper is to describe the asymptotic behavior of solutions to the following class of fractional elliptic semilinear equations with singular homogeneous potentials

N/A
N/A
Protected

Academic year: 2021

Condividi "The purpose of the present paper is to describe the asymptotic behavior of solutions to the following class of fractional elliptic semilinear equations with singular homogeneous potentials"

Copied!
31
0
0

Testo completo

(1)

UNIQUE CONTINUATION PROPERTY AND LOCAL ASYMPTOTICS OF SOLUTIONS TO FRACTIONAL ELLIPTIC EQUATIONS

MOUHAMED MOUSTAPHA FALL AND VERONICA FELLI

Abstract. Asymptotics of solutions to fractional elliptic equations with Hardy type potentials is studied in this paper. By using an Almgren type monotonicity formula, separation of variables, and blow-up arguments, we describe the exact behavior near the singularity of solutions to linear and semilinear fractional elliptic equations with a homogeneous singular potential related to the fractional Hardy inequality. As a consequence we obtain unique continuation properties for fractional elliptic equations.

1. Introduction

The purpose of the present paper is to describe the asymptotic behavior of solutions to the following class of fractional elliptic semilinear equations with singular homogeneous potentials

(1) (−∆) s u(x) − λ

|x| 2s u(x) = h(x)u(x) + f (x, u(x)), in Ω, where Ω ⊂ R N is a bounded domain containing the origin,

N > 2s, s ∈ (0, 1), λ < Λ N,s := 2 2s Γ 2 N +2s 4  Γ 2 N −2s 4  , (2)

h ∈ C 1 (Ω \ {0}), |h(x)| + |x · ∇h(x)| 6 C h |x| −2s+ε as |x| → 0, (3)

( f ∈ C 1 (Ω × R), t 7→ F (x, t) ∈ C 1 (Ω × R),

|f (x, t)t| + |f t 0 (x, t)t 2 | + |∇ x F (x, t) · x| 6 C f |t| p for a.e. x ∈ Ω and all t ∈ R, (4)

where 2 < p 6 2 (s) = N −2s 2N , F (x, t) = R t

0 f (x, r) dr, C f , C h , ε > 0 are positive constants inde- pendent of x ∈ Ω and t ∈ R, ∇ x F denotes the gradient of F with respect to the x variable, and f t 0 (x, t) = ∂f ∂t (x, t).

We recall that for any ϕ ∈ C c (R N ) and s ∈ (0, 1), the fractional Laplacian (−∆) s ϕ is defined as

(5) (−∆) s ϕ(x) = C(N, s) P.V.

Z

R

N

ϕ(x) − ϕ(y)

|x − y| N +2s dy = C(N, s) lim

ρ→0

+

Z

|x−y|>ρ

ϕ(x) − ϕ(y)

|x − y| N +2s dy where P.V. indicates that the integral is meant in the principal value sense and

C(N, s) = π

N2

2 2s Γ N +2s 2 

Γ(2 − s) s(1 − s).

The Dirichlet form associated to (−∆) s on C c (R N ) is given by (6) (u, v) D

s,2

(R

N

) = C(N, s)

2 Z

R

2N

(u(x) − u(y))(v(x) − v(y))

|x − y| N +2s dx dy = Z

R

N

|ξ| 2s b v(ξ) u(ξ) dξ, b

Date: Revised version, June 15, 2013.

M. M. Fall is supported by the Alexander von Humboldt foundation. V. Felli was partially supported by the PRIN2009 grant “Critical Point Theory and Perturbative Methods for Nonlinear Differential Equations”.

2010 Mathematics Subject Classification. 35R11, 35B40, 35J60, 35J75.

Keywords. Fractional elliptic equations, Caffarelli-Silvestre extension, Hardy inequality, Unique continuation property.

1

(2)

where u denotes the unitary Fourier transform of u. It defines a scalar product thanks to (8) or b (9) below. From now on we define D s,2 (R N ) as the completion on C c (R N ) with respect to the norm induced by the scalar product (6).

By a weak solution to (1) we mean a function u ∈ D s,2 (R N ) such that (7) (u, ϕ) D

s,2

(R

N

) =

Z

 λ

|x| 2s u(x) + h(x)u(x) + f (x, u(x))



ϕ(x) dx, for all ϕ ∈ C c (Ω).

We notice that the right hand side of (7) is well defined in view of assumptions (3)–(4), the Hardy-Littlewood-Sobolev inequality

(8) S N,s kuk 2 L

2∗(s)

(R

N

) 6 kuk 2 D

s,2

(R

N

) , and the following Hardy inequality, due to Herbst in [19] (see also [27]),

Λ N,s

Z

R

N

u 2 (x)

|x| 2s dx 6 Z

R

N

|ξ| 2s u b 2 (ξ) dξ = kuk 2 D

s,2

(R

N

) , for all u ∈ D s,2 (R N ).

(9)

It should also be remarked that, in (7), we allow p = 2 (s) and that u is not prescribed outside Ω.

One of the aim of this paper is to give the precise behavior of a solution u to (1). The rate and the shape of u are given by the the eigenvalues and the eigenfunctions of the following eigenvalue problem

− div S

N

1−2s 1S

N

ψ) = µ θ 1 1−2s ψ, in S N + ,

− lim θ

1

→0

+

θ 1 1−2sS

N

ψ · e 1 = κ s λψ, on ∂S N + , (10)

where

(11) κ s = Γ(1 − s)

2 2s−1 Γ(s) and

S N + = {(θ 1 , θ 2 , . . . , θ N +1 ) ∈ S N : θ 1 > 0} = n

z

|z| : z ∈ R N +1 , z · e 1 > 0 o ,

with e 1 = (1, 0, . . . , 0); we refer to section 2.1 for a variational formulation of (10). From classical spectral theory (see section 2.1 for the details) problem (10) admits a diverging sequence of real eigenvalues with finite multiplicity

µ 1 (λ) 6 µ 2 (λ) 6 · · · 6 µ k (λ) 6 · · · Moreover µ 1 (λ) > − N −2s 2  2

, see Lemma 2.2. We notice that if λ = 0 then µ k (0) > 0 for all k.

Our first main result is the following asymptotics of solutions at the singularity.

Theorem 1.1. Let u ∈ D s,2 (R N ) be a nontrivial solution to (1) in a bounded domain Ω ⊂ R N containing the origin as in (7) with s, λ, h and f satisfying assumptions (2), (3) and (4). Then there exists an eigenvalue µ k

0

(λ) of (10) and an eigenfunction ψ associated to µ k

0

(λ) such that (12) τ

2s−N2

q (

2s−N2

)

2

k0

(λ)

u(τ x) → |x|

N −2s2

+

q (

2s−N2

)

2

k0

(λ)

ψ  0, x

|x|



as τ → 0 + , in C loc 1,α (B 1 0 \ {0}) for some α ∈ (0, 1), where B 1 0 := {x ∈ R N : |x| < 1}, and, in particular, (13) τ

2s−N2

q (

2s−N2

)

2

k0

(λ)

u (τ θ 0 ) → ψ (0, θ 0 ) in C 1,α (S N −1 ) as τ → 0 + , where S N −1 = ∂S N + .

We should mention that the above result is stated in such form for the sake of simplicity. In fact, for the eigenfunction ψ in (13), we obtain precisely its components in any basis of the eigenspace corresponding to µ k

0

(λ), see Theorem 4.1 (below). We also remark that µ 1 (λ) < 0 for λ > 0 and it is determined implicitly by the usual Gamma function, see Proposition 2.3.

As a particular case of Theorem 1.1, if λ = 0 we obtain that the convergence stated in 12 holds

in C 1,α (B 1 0 ).

(3)

Corollary 1.2. Let u ∈ D s,2 (R N ) be a nontrivial weak solution to (14) (−∆) s u(x) = h(x)u(x) + f (x, u(x)), in Ω,

in a bounded domain Ω ⊂ R N with s ∈ (0, 1), h ∈ C 1 (Ω), and f satisfying (4). Then, for every x 0 ∈ Ω, there exists an eigenvalue µ k

0

= µ k

0

(0) of problem (10) with λ = 0 and an eigenfunction ψ associated to µ k

0

such that

(15) τ

2s−N2

q (

2s−N2

)

2

k0

u(x 0 + τ (x − x 0 ))

→ |x − x 0 |

N −2s2

+

q (

2s−N2

)

2

k0

ψ 

0, x − x 0

|x − x 0 |



as τ → 0 + , in C 1,α ({x ∈ R N : x − x 0 ∈ B 1 0 }).

Our next result contains the so called strong unique continuation property which is a direct consequence of Theorem 1.1.

Theorem 1.3. Suppose that all the assumptions of Theorem 1.1 hold true. Let u be a solution to (1) in a bounded domain Ω ⊂ R N containing the origin. If u(x) = o(|x| n ) = o(1)|x| n as |x| → 0 for all n ∈ N, then u ≡ 0 in Ω.

From Theorem 1.2 a unique continuation principle related to sets of positive measures follows.

Theorem 1.4. Let u ∈ D s,2 (R N ) be a weak solution to (14) in a bounded domain Ω ⊂ R N with s ∈ (0, 1), h ∈ C 1 (Ω), and f satisfying (4). If u ≡ 0 on a set E ⊂ Ω of positive measure, then u ≡ 0 in Ω.

An interesting application of Theorem 1.4 is that the nodal sets of eigenfunctions for the frac- tional laplacian operator have zero Lebesgue measure. We should point out that this was a key assumption in [15], where the authors studied the existence of weak solutions to some non-local equations.

Recent research in the field of second order elliptic equations has devoted a great attention to the problem of unique continuation property in the presence of singular lower order terms, see e.g.

[11, 20, 22]. Two different kinds of approach have been developed to treat unique continuation:

a first one is due to Carleman [5] and is based on weighted priori inequalities, whereas a second one is due to Garofalo and Lin [17] and is based on local doubling properties proved by Almgren monotonicity formula. In the present paper we will follow the latter approach. Furthermore, in the spirit of [12, 13, 14], the combination of monotonicity methods with blow-up analysis will unable us to prove not only unique continuation but also the precise asymptotics of solutions stated in Theorem 1.1.

As far as unique continuation from sets of positive measures is concerned, we mention [6] where it was proved for second order elliptic operators by combining strong unique continuation property with the De Giorgi inequality. Since the validity of a the De Giorgi type inequality for the fractional problem (or even its extension, see (17)) seems to be hard to prove, we will base the proof of Theorem 1.4 directly on the asymptotic of solutions proved in Theorem 1.1.

To explain our argument of proving the asymptotic behavior, let us we write (7) as

(16) (−∆) s u = G(x, u) in Ω.

The proof of our results is based on the study of the Almgren frequency function at the origin 0:

“ratio of the local energy over mass near the origin”. Due to the non-locality of the Dirichlet form associated to (−∆) s , it is not clear how to set up an Almgren’s type frequency function using this energy as in the local case s = 1. A way out for this difficulty is to use the Caffarelli-Silvestre extension [4] which can be seen as a local version of (16). The Caffarelli-Silvestre extension of a solution u to (16) is a function w defined on

R N +1 + = {z = (t, x) : t ∈ (0, +∞), x ∈ R N }

satisfying w = u on Ω and solving in some weak sense (see Section 2 for more details) the boundary value problem

(17) (div(t 1−2s ∇w) = 0, in R N +1 + ,

− lim

t→0

+

t 1−2s ∂w ∂t = κ s G(x, w), on Ω.

(4)

Here and in the following, we write z = (t, x) ∈ R N +1 + with x ∈ R N and t > 0, and we identify R N with ∂R N +1 + , so that Ω is contained in ∂R N +1 + .

We then consider the Almgren’s frequency function N (r) = D(r)

H(r) where

D(r) = 1 r N −2s

 Z

B

r+

t 1−2s |∇w| 2 dt dx − κ s

Z

B

r0

G(x, w)w dx



, H(r) = 1 r N +1−2s

Z

S

r+

t 1−2s w 2 dS, being

B + r = {z = (t, x) ∈ R N +1 + : |z| < r}, B 0 r := {x ∈ R N : |x| < r}, S r + = {z = (t, x) ∈ R N +1 + : |z| = r},

and dS denoting the volume element on N -dimensional spheres.

As a first but nontrivial step, we prove that lim r→0 N (r) := γ exists and it is finite, see Lemma 3.14. Next, we make a blow-up analysis by zooming around the origin the solution w normalized also by √

H. More precisely, setting w τ (z) = √ w(τ z)

H(τ ) , we have that w τ converges, (in some H¨ older and Sobolev spaces) to w solving the limiting equation e

(div(t 1−2s ∇ w) = 0, e in B + 1 ,

− lim

t→0

+

t 1−2s ∂ ∂t w e = |z| κ

s2s

λ w, e on B 1 0 .

To obtain this, the fact that h is negligible with respect to the Hardy potential and f is at most critical with respect to the Sobolev exponent (see assumptions (3) and (4)) plays a crucial role; we refer to Lemma 4.2 for more details.

The main point is that the Almgren’s frequency for w is e e N (r) = lim τ →0 N (τ r) = γ, i.e. N e is constant; hence w and ∇ e w · e |z| z are proportional on L 2 (S + r ; t 1−2s ). As a consequence we ob- tain w(z) = ϕ(|z|)ψ e |z| z . By separating variables in polar coordinates, we obtain that ψ is an eigenfunction of (10) for some eigenvalue µ k

0

(λ). By the method of variation of constants and the fact that w has finite energy near the origin, we then prove that ϕ(r) is proportional to e r γ and that γ = 2s−N 2 +

q 2s−N 2

 2

+ µ k

0

(λ). We finally complete the proof by showing that lim r→0 r −2γ H(r) > 0, see Lemma 4.5.

We should mention that in the recent literature, a great attention has been addressed to non- local fractional diffusion and many papers have been devoted to the study of existence, non- existence, regularity and qualitative properties of solutions to elliptic equations associated to frac- tional Laplace type operators, see e.g. [2, 3, 4, 7, 9, 10, 24, 25] and references therein.

2. Preliminaries and notations Notation. We list below some notation used throughout the paper.

- S N = {z ∈ R N +1 : |z| = 1} is the unit N -dimensional sphere.

- S N + = {(θ 1 , θ 2 , . . . , θ N +1 ) ∈ S N : θ 1 > 0} := S N ∩ R N +1 + . - dS denotes the volume element on N -dimensional spheres.

- dS 0 denotes the volume element on (N − 1)-dimensional spheres.

Let D 1,2 (R N +1 + ; t 1−2s ) be the completion of C c (R N +1 + ) with respect to the norm kwk D

1,2

(R

N +1+

;t

1−2s

) =

 Z

R

N +1+

t 1−2s |∇w(t, x)| 2 dt dx

 1/2 .

We recall that there exists a well defined continuous trace map Tr : D 1,2 (R N +1 + ; t 1−2s ) → D s,2 (R N ),

see e.g. [2].

(5)

For every u ∈ D s,2 (R N ), let H(u) ∈ D 1,2 (R N +1 + ; t 1−2s ) be the unique solution to the minimiza- tion problem

(18) Z

R

N +1+

t 1−2s |∇H(u)| 2 dt dx

= min

 Z

R

N +1+

t 1−2s |∇v| 2 dt dx : v ∈ D 1,2 (R N +1 + ; t 1−2s ), Tr(v) = u

 .

By Caffarelli and Silvestre [4] we have that (19)

Z

R

N +1+

t 1−2s ∇H(u) · ∇ ϕ dt dx = κ e s (u, Tr ϕ) e D

s,2

(R

N

) for all ϕ ∈ D e 1,2 (R N +1 + ; t 1−2s ), where κ s is defined in (11).

We notice that combining (9), (18), (19) and (8), we obtain the following Hardy-trace inequality κ s Λ N,s

Z

R

N

(Tr v) 2 (x)

|x| 2s dx 6 Z

R

N +1+

t 1−2s |∇v| 2 dt dx, for all v ∈ D 1,2 (R N +1 + ; t 1−2s ) (20)

and also the Sobolev-trace inequality κ s S N,s k Tr vk 2 L

2∗(s)

(R

N

) 6

Z

R

N +1+

t 1−2s |∇v| 2 dt dx, for all v ∈ D 1,2 (R N +1 + ; t 1−2s ).

(21)

2.1. Separation of variables in the extension operator. For every R > 0, we define the space H 1 (B R + ; t 1−2s ) as the completion of C (B R + ) with respect to the norm

kwk H

1

(B

R+

;t

1−2s

) =

 Z

B

R+

t 1−2s 

|∇w(t, x)| 2 + w 2 (t, x)  dt dx

 1/2

.

By direct calculations, we can obtain the following lemma concerning separation of variables in the extension operator.

Lemma 2.1. If v ∈ H 1 (B R + ; t 1−2s ) is such that v(z) = f (r)ψ(θ) for a.e. z = (t, x) ∈ R N +1 + , with r = |z| < R and θ = |z| z ∈ S N , then

div(t 1−2s ∇v(z)) = 1

r N r N +1−2s f 0  0

θ 1 1−2s ψ(θ) + r −1−2s f (r) div S

N

1−2s 1S

N

ψ(θ))

in the distributional sense, where θ 1 = t r = θ · e 1 with e 1 = (1, 0, . . . , 0), div S

N

(respectively ∇ S

N

) denotes the Riemannian divergence (respectively gradient) on the unit sphere S N endowed with the standard metric.

Let us define H 1 (S N + ; θ 1 1−2s ) as the completion of C (S N + ) with respect to the norm kψk H

1

(S

N+

1−2s1

) =

 Z

S

N+

θ 1 1−2s |∇ S

N

ψ(θ)| 2 + ψ 2 (θ)dS

 1/2 . We also denote

L 2 (S N + ; θ 1−2s 1 ) := n

ψ : S N + → R measurable such that R

S

N+

θ 1 1−2s ψ 2 (θ) dS < +∞ o . The following trace inequality on the unit half-sphere S N + holds.

Lemma 2.2. There exists a well defined continuous trace operator H 1 (S N + ; θ 1 1−2s ) → L 2 (∂S N + ) = L 2 (S N −1 ).

Moreover for every ψ ∈ H 1 (S N + ; θ 1−2s 1 ) κ s Λ N,s

 Z

S

N −1

|ψ(θ 0 )| 2 dS 0



6  N − 2s 2

 2 Z

S

N+

θ 1−2s 1 |ψ(θ)| 2 dS + Z

S

N+

θ 1−2s 1 |∇ S

N

ψ(θ)| 2 dS.

where dS 0 denotes the volume element on the sphere S N −1 = ∂S N + = {(θ 1 , θ 0 ) ∈ S N + : θ 1 = 0}.

(6)

Proof. Let ψ ∈ C (S N + ) and f ∈ C c (0, +∞) with f 6= 0. Rewriting (20) for v(z) = f (r)ψ(θ), r = |z|, θ = |z| z , we obtain that

κ s Λ N,s

 Z +∞

0

r N −1−2s f 2 (r) dr

 Z

S

N −1

|ψ(0, θ 0 )| 2 dS 0



6

 Z +∞

0

r N +1−2s |f 0 (r)| 2 dr

 Z

S

N+

θ 1 1−2s |ψ(θ)| 2 dS



+

 Z +∞

0

r N −1−2s f 2 (r) dr

 Z

S

N+

θ 1 1−2s |∇ S

N

+

ψ(θ)| 2 dS

 , and hence, by optimality of the classical Hardy constant, see [18],

κ s Λ N,s

 Z

S

N −1

|ψ(0, θ 0 )| 2 dS 0



6

 Z

S

N+

θ 1−2s 1 |ψ(θ)| 2 dS

 inf

f ∈C

c

(0,+∞)

R +∞

0 r N +1−2s |f 0 (r)| 2 dr R +∞

0 r N −1−2s f 2 (r) dr + Z

S

N+

θ 1−2s 1 |∇ S

N

+

ψ(θ)| 2 dS

=  N − 2s 2

 2 Z

S

N+

θ 1 1−2s |ψ(θ)| 2 dS + Z

S

N+

θ 1 1−2s |∇ S

N

+

ψ(θ)| 2 dS.

By density of C (S N + ) in H 1 (S N + ; θ 1 1−2s ), we obtain the conclusion.

In view of Lemma 2.1, in order to construct an orthonormal basis of L 2 (S N + ; θ 1 1−2s ) for expanding solutions to (17) in Fourier series, we are naturally lead to consider the eigenvalue problem (10), which admits the following variational formulation: we say that µ ∈ R is an eigenvalue of problem (10) if there exists ψ ∈ H 1 (S N + ; θ 1 1−2s ) \ {0} (called eigenfunction) such that

Q(ψ, υ) = µ Z

S

N+

θ 1−2s 1 ψ(θ)υ(θ) dS, for all υ ∈ H 1 (S N + ; θ 1−2s 1 ), where

Q : H 1 (S N + ; θ 1−2s 1 ) × H 1 (S N + ; θ 1 1−2s ) → R, Q(ψ, υ) =

Z

S

N+

θ 1−2s 1S

N

ψ(θ) · ∇υ(θ) dS − λκ s

Z

S

N −1

T ψ(θ 0 )T υ(θ 0 ) dS 0 .

By Lemma 2.2 the bilinear form Q is continuous and weakly coercive on H 1 (S N + ; θ 1 1−2s ). Moreover the belonging of the weight t 1−2s to the second Mackenhoupt class ensures that the embedding H 1 (S N + ; θ 1−2s 1 ) ,→,→ L 2 (S N + ; θ 1 1−2s ) is compact (see [8] for weighted embeddings with Mackenhoupt weights). Then, from classical spectral theory (see e.g. [23, Theorem 6.16]), problem (10) admits a diverging sequence of real eigenvalues with finite multiplicity µ 1 (λ) 6 µ 2 (λ) 6 · · · 6 µ k (λ) 6 · · · the first of which admits the variational characterization

(22) µ 1 (λ) = min

ψ∈H

1

(S

N+

1−2s1

)\{0}

Q(ψ, ψ) R

S

N+

θ 1 1−2s ψ 2 (θ) dS . Furthermore, in view of Lemma 2.2, we have that

(23) µ 1 (λ) > −  N − 2s

2

 2 .

To each k > 1, we associate an L 2 (S N + ; θ 1−2s 1 )-normalized eigenfunction ψ k ∈ H 1 (S N + ; θ 1−2s 1 ) \ {0}

corresponding to the k-th eigenvalue µ k (λ), i.e. satisfying (24) Q(ψ k , υ) = µ k (λ)

Z

S

N+

θ 1 1−2s ψ k (θ)υ(θ) dS, for all υ ∈ H 1 (S N + ; θ 1 1−2s ).

In the enumeration µ 1 (λ) 6 µ 2 (λ) 6 · · · 6 µ k (λ) 6 · · · we repeat each eigenvalue as many times as its multiplicity; thus exactly one eigenfunction ψ k corresponds to each index k ∈ N, k > 1. We can choose the functions ψ k in such a way that they form an orthonormal basis of L 2 (S N + ; θ 1−2s 1 ).

We can also determine µ 1 (λ) for λ ∈ (0, Λ N,s ), where Λ N,s is the fractional Hardy constant

defined in (2).

(7)

Proposition 2.3. For every α ∈ 0, N −2s 2 , we define λ(α) = 2 2s Γ N +2s+2α 4 

Γ N −2s−2α 4 

Γ N +2s−2α 4  Γ N −2s+2α 4  .

Then the mapping α 7→ λ(α) is continuous and decreasing. In addition we have that µ 1 (λ(α)) = α 2 −  N − 2s

2

 2

for all α ∈



0, N − 2s 2

 .

Proof. It was proved in [9, Lemma 3.1] that, for every α ∈ 0, N −2s 2 , there exists a positive continuous function Φ α : R N +1 + → R such that

(25)

 

 

div(t 1−2s ∇Φ α ) = 0 in R N +1 +

Φ α = |x|

2s−N2

on R N \ {0}

−t 1−2s ∂Φ ∂t

α

= κ s λ(α)|x| −2s Φ α on R N \ {0}, Moreover Φ α ∈ H 1 (B R + ; t 1−2s ) for every R > 0 and Φ α is scale invariant, i.e.

Φ α (τ z) = τ

2s−N2

Φ α (z), for all τ > 0, thus implying that, for all z ∈ R N +1 + ,

(26) c 1 |z|

2s−N2

6 Φ α (z) 6 c 2 |z|

2s−N2

,

for some positive constants c 1 , c 2 . It is also known (see for instance [16]) that the map α 7→ λ(α) is continuous and monotone decreasing.

We write

Φ α (z) =

X

k=1

Φ k α (|z|)ψ k (z/|z|), Φ k α (r) = Z

S

N+

ψ k (θ)Φ α (rθ)dS.

In particular, since ψ 1 > 0, by (26) we have, for every r ∈ (0, R), (27) c 0 1 |r|

2s−N2

6 Φ 1 α (r) 6 c 0 2 |r|

2s−N2

,

for some positive constants c 0 1 , c 0 2 . Using (25), we have, weakly, for every k > 1 and r ∈ (0, R), ( 1

r

N

r N +1−2sk α ) 0  0

θ 1−2s 1 ψ k (θ) + r −1−2s Φ k α div S

N

1−2s 1S

N

ψ k (θ)) = 0,

−Φ k α lim θ

1

→0

+

θ 1 1−2sS

N

ψ k (θ) · e 1 = κ s λ(α) ψ k (0, θ 0k α .

Testing the above equation with ψ 1 > 0 and using also the fact that Φ 1 α > 0, we obtain (Φ 1 α ) 00 + N + 1 − 2s

r (Φ 1 α ) 0 − µ 1 (λ(α)) r 2 Φ 1 α = 0 and hence Φ 1 α (r) is of the form

Φ 1 α (r) = c 3 r σ

+

+ c 4 r σ

for some c 3 , c 4 ∈ R, where

σ + = − N − 2s

2 +

s

 N − 2s 2

 2

+ µ 1 (λ(α)) and σ = − N − 2s

2 −

s

 N − 2s 2

 2

+ µ 1 (λ(α)).

Since the function |z| σ

k0

ψ 1 ( |z| z ) / ∈ H 1 (B 1 + ; t 1−2s ), we deduce that c 4 = 0 and thus for every r ∈ (0, R)

Φ α 1 (r) = c 3 r σ

+

. This together with (27) implies that

µ 1 (λ(α)) = α 2 −  N − 2s 2

 2

for all α ∈



0, N − 2s 2



,

as claimed.

(8)

2.2. Hardy type inequalities. From well-known weighted embedding inequalities and the fact that the weight t 1−2s belongs to the second Mackenhoupt class (see e.g. [8]), the embedding H 1 (B r + ; t 1−2s ) ,→ L 2 (B r + ; t 1−2s ) is compact. It can be also proved that both the trace operators (28) H 1 (B r + ; t 1−2s ) ,→,→ L 2 (S r + ; θ 1 1−2s ),

(29) H 1 (B r + ; t 1−2s ) ,→,→ L 2 (B r 0 ) are well defined and compact.

For sake of simplicity, in the following of this paper, we will often denote the trace of a function with the same letter as the function itself.

The following Hardy type inequality with boundary terms holds.

Lemma 2.4. For all r > 0 and w ∈ H 1 (B r + ; t 1−2s ), the following inequality holds

 N − 2s 2

 2 Z

B

+r

t 1−2s w 2 (z)

|z| 2 dz 6 Z

B

+r

t 1−2s



∇w(z) · z

|z|

 2

dz +  N − 2s 2r

 Z

S

r+

t 1−2s w 2 dS.

Proof. By scaling, it is enough to prove the stated inequality for r = 1. Let V (z) = |z|

2s−N2

, z ∈ R N +1 + \ {0}.

We notice that V satisfies

(30) − div(t 1−2s ∇V ) =  N − 2s 2

 2

t 1−2s |z| −2 V (z) in R N +1 + \ {0}.

Hence, letting w ∈ C (B 1 + ), multiplying (30) with w V

2

, and integrating over B 1 + \B δ + with δ ∈ (0, 1), we obtain

 N − 2s 2

 2 Z

B

+1

t 1−2s w 2 (z)

|z| 2 dz

= Z

B

1+

\B

δ+

t 1−2s ∇V (z) · ∇  w 2 V



(z) dz − Z

S

+1

t 1−2s (∇V · ν) w 2 V dS +

Z

S

δ+

t 1−2s (∇V · ν) w 2 V dS

= −(N − 2s) Z

B

+1

\B

δ+

t 1−2s w

|z|

 ∇w · z

|z|

 dz −  N − 2s 2

 2 Z

B

+1

\B

δ+

t 1−2s w 2

|z| 2 dz + N − 2s

2 Z

S

1+

t 1−2s w 2 dS − N − 2s 2

1 δ

Z

S

+δ

t 1−2s w 2 dS where ν(z) = |z| z . Since, by Schwarz’s inequality

(31) − (N − 2s) Z

B

1+

\B

δ+

t 1−2s w

|z|

 ∇w · z

|z|

 dz

6  N − 2s 2

 2 Z

B

+1

\B

+δ

t 1−2s w 2

|z| 2 dz + Z

B

+1

\B

+δ

t 1−2s 

∇w · z

|z|

 2 dz and

1 δ

Z

S

δ+

t 1−2s w 2 dS = O(δ N −2s ) = o(1) for δ → 0 + ,

letting δ → 0 we obtain the stated inequality for r = 1 and w ∈ C (B + 1 ). The conclusion follows by density of C (B 1 + ) in H 1 (B 1 + ; t 1−2s ).

Lemma 2.5. For every r > 0 and w ∈ H 1 (B r + ; t 1−2s ), the following inequality holds κ s Λ N,s

Z

B

0r

w 2

|x| 2s dx 6  N − 2s 2r

 Z

S

+r

t 1−2s w 2 dS + Z

B

r+

t 1−2s |∇w| 2 dz.

(9)

Proof. Let w ∈ C (B + r ). Then, passing to polar coordinates and using Lemmas 2.2 and 2.4, we obtain

κ s Λ N,s

Z

B

0r

w 2 (0, x)

|x| 2s dx = κ s Λ N,s

Z r 0

ρ N −1−2s

 Z

S

N −1

w 2 (0, ρθ 0 ) dS 0

 dρ

6 Z r

0

ρ N −1−2s

  N − 2s 2

 2 Z

S

N+

θ 1 1−2s |w(ρθ)| 2 dS + Z

S

N+

θ 1−2s 1 |∇ S

N

w(ρθ)| 2 dS

 dρ

=  N − 2s 2

 2 Z

B

r+

t 1−2s w 2

|z| 2 dz + Z r

0

ρ N −1−2s

 Z

S

N+

θ 1−2s 1 |∇ S

N

w(ρθ)| 2 dS

 dρ

6  N − 2s 2r

 Z

S

r+

t 1−2s w 2 dS

+ Z r

0

ρ N +1−2s

 Z

S

N+

θ 1−2s 1  1

ρ 2 |∇ S

N

w(ρθ)| 2 +

∂w

∂ρ (ρθ)

2  dS

 dρ

=  N − 2s 2r

 Z

S

r+

t 1−2s w 2 dS + Z

B

+r

t 1−2s |∇w| 2 dz.

The conclusion follows by density of C (B + r ) in H 1 (B r + ; t 1−2s ).

The following Sobolev type inequality with boundary terms holds.

Lemma 2.6. There exists e S N,s > 0 such that, for all r > 0 and w ∈ H 1 (B + r ; t 1−2s ),

 Z

B

0r

|w| 2

(s) dx



2∗ (s)2

6 e S N,S  N − 2s 2r

Z

S

r+

t 1−2s w 2 dS + Z

B

+r

t 1−2s |∇w| 2 dz

 .

Proof. By scaling, it is enough to prove the statement for r = 1. Let w ∈ C (B + 1 ) and denote by w its fractional Kelvin transform defined as e w(z) = |z| e −(N −2s) w |z| z

2

. Some computations (see also [10]) show that

Z

B

+1

t 1−2s |∇w| 2 dz + (N − 2s) Z

S

1+

t 1−2s w 2 dS = Z

R

N +1+

\B

+1

t 1−2s |∇ w| e 2 dz, (32)

Z

B

01

|w| 2

(s) dx = Z

R

N

\B

10

| w| e 2

dx.

(33)

In particular the function

v(z) =

( w(z), if z ∈ B 1 + , w(z), e if z ∈ R N +1 + \ B + 1 belongs to D 1,2 (R N +1 + ; t 1−2s ), so that, by (21) we have

(34)

 Z

R

N

|v| 2

(s) dx



2∗ (s)2

6 1

S N,s κ s

Z

R

N +1+

t 1−2s |∇v| 2 dz.

From (32), (33), and (34), it follows that

 2

Z

B

01

|w| 2

(s) dx



2∗ (s)2

6 2

S N,s κ s

 Z

B

1+

t 1−2s |∇w| 2 dz + N − 2s 2

Z

S

+1

t 1−2s w 2 dS



which, by density, yields the conclusion.

Combining Lemma 2.5 and Lemma 2.6 the following corollary follows.

Corollary 2.7. For all r > 0 and w ∈ H 1 (B + r ; t 1−2s ), the following inequalities hold (35)

Z

B

+r

t 1−2s |∇w| 2 dz − κ s λ Z

B

0r

w 2

|x| 2s dx + N − 2s 2r

Z

S

r+

t 1−2s w 2 dS

> κ s (Λ N,s − λ) Z

B

0r

w 2

|x| 2s dx

(10)

and (36)

Z

B

+r

t 1−2s |∇w| 2 dz − κ s λ Z

B

0r

w 2

|x| 2s dx + N − 2s 2r

Z

S

r+

t 1−2s w 2 dS

> Λ N,s − λ (1 + Λ N,s ) e S N,s

 Z

B

r0

|w| 2

(s) dx



2∗ (s)2

.

3. The Almgren type frequency function

Let R > 0 be such that B R 0 ⊂⊂ Ω and w ∈ H 1 (B R + ; t 1−2s ) be a nontrivial solution to

(37) (div(t 1−2s ∇w) = 0, in B R + ,

− lim t→0

+

t 1−2s ∂w ∂t (t, x) = κ s

 λ

|x|

2s

w + hw + f (x, w) 

, on B 0 R , in a weak sense, i.e., for all ϕ ∈ C c (B R + ∪ B R 0 ), we have that

(38)

Z

R

N +1+

t 1−2s ∇w · ∇ϕ dt dx = κ s

Z

B

0R

 λ

|x| 2s w + hw + f (x, w)

 ϕ dx, with s, λ, h, f as in assumptions (2), (3), and (4).

The main result of this section is the existence of the limit as r → 0 + of the Almgren’s frequency function (see [17] and [1]) associated to w

(39) N (r) = r

"

Z

B

+r

t 1−2s |∇w| 2 dt dx − κ s

Z

B

r0

 λ

|x| 2s w 2 + h(x)w 2 + f (x, w)w

 dx

#

Z

S

r+

t 1−2s w 2 dS

.

We notice that, by Lemma 2.5, w(0, ·) ∈ L 2 (B R 0 ; |x| −2s ) and so the L 1 (0, R)-function r 7→

Z

S

+r

t 1−2s |∇w| 2 dS, respectively r 7→

Z

∂B

0r

w 2

|x| 2s dS 0 , is the weak derivative of the W 1,1 (0, R)-function

r → Z

B

+r

t 1−2s |∇w| 2 dz, respectively r 7→

Z

B

r0

w 2

|x| 2s dx.

In particular, for a.e. r ∈ (0, R), ∂w ∂ν ∈ L 2 (S r + ; t 1−2s ), where ν = ν(z) is the unit outer normal vector ν(z) = |z| z .

Lemma 3.1. For a.e. r ∈ (0, R) and every ϕ ∈ C e (B r + ) Z

B

r+

t 1−2s ∇w · ∇ ϕ dz = e Z

S

r+

t 1−2s ∂w

∂ν ϕ dS + κ e s

Z

B

r0

 λ

|x| 2s w + hw + f (x, w)

 ϕ dx. e

Proof. It follows by testing (38) with ϕ(z)η e n (|z|) where η n (ρ) = 1 if ρ < r − n 1 , η n (r) = 0 if ρ > r, η(ρ) = n(r − ρ) is r − n 1 6 ρ 6 r, passing to the limit, and noticing that a.e. r ∈ (0, R) is a Lebesgue point for the L 1 (0, R)- function r 7→ R

S

r+

t 1−2s ∂w ∂ν ϕ dS. e

Lemma 3.2. (i) Let r > 0 and V ∈ L q (B 0 r ) for some q > N 2s . For every t 0 , r 0 > 0 such that [0, t 0 ) × B 0 r

0

b B r + ∪ B 0 r there exist positive constants A 1 > 0, α ∈ (0, 1) depending on t 0 , r 0 , r, kV k L

q

(B

r0

) such that for every v ∈ H 1 (B + r ; t 1−2s ) solving

( div(t 1−2s ∇v) = 0, in B r +

− lim t→0

+

t 1−2s v t = V (x)v, on B r 0 , we have that v ∈ C 0,α ([0, t 0 ) × B r 0

0

) and

(40) kvk C

0,α

([0,t

0

)×B

0r0

) 6 A 1 kvk H

1

(B

+r

;t

1−2s

) .

(11)

(ii) Let v ∈ H 1 (B r + ; t 1−2s ) ∩ C 0,α (B r + ) for some α ∈ (0, 1) be a function satisfying ( div(t 1−2s ∇v) = 0, in B + r

− lim t→0

+

t 1−2s v t = g(x, v), on B 0 r ,

where g ∈ C 1 (B r 0 × R) and |g(x, ρ)| 6 c(|ρ| + |ρ| p−1 ) for some 2 < p 6 2 (s) = N −2s 2N , c > 0, and every x ∈ B r 0 and ρ ∈ R. Let t 0 , r 0 > 0 such that [0, t 0 ) × B r 0

0

b B r + . Then there exist positive constants A 2 , β depending only on N , p, s, c, r, r 0 , t 0 , kvk H

1

(B

+r

;t

1−2s

) and kgk C

1

(B

0

r0

×[0,A

3

]) where A 3 = kvk C

0,α

([0,t

0

)×B

0

r0

) , with β ∈ (0, 1), such that (41) k∇ x vk C

0,β

([0,t

0

)×B

r00

) 6 A 2 ,

(42) kt 1−2s v t k C

0,β

([0,t

0

)×B

r00

) 6 A 2 .

Remark 3.3. The dependence of the constant A 2 in Lemma 3.2 on kvk H

1

(B

r+

;t

1−2s

) is continuous;

in particular we can take the same A 2 for a family of solutions which are uniformly bounded in H 1 (B r + ; t 1−2s ) ∩ C 0,α (B + r ).

Proof. Part (i) and (40) follows from [[21], Proposition 2.4].

To prove (ii), for h ∈ R N with |h| << 1, we set v h (t, x) = v(t,x+h)−v(t,x)

|h| for every x ∈ B r 0

0

/2 . Then we have

( div(t 1−2s ∇v h ) = 0, in B + r

0

/2

− lim t→0

+

t 1−2s v t h = c h (x)v h + b h , on B r 0

0

/2 , where

c h (x) = g(x, v(0, x + h)) − g(x, v(0, x))

v(0, x + h) − v(0, x) χ {v(0,x+h)6=v(0,x)} (x) with χ A being the characteristic function of a set A and

b h (x) = g(x + h, v(0, x + h)) − g(x, v(0, x + h))

|h| .

Let A 3 = kvk C

0,α

([0,t

0

)×B

r00

) . Then we have kc h k L

(B

0

r0/4

) + kb h k L

(B

0

r0/4

) 6 kg ρ k L

(B

0

r0/2

×[0,A

3

]) + k∇ x gk L

(B

0

r0/2

×[0,A

3

]),

for every small h. Applying once again [[21], Proposition 2.4], kv h k C

0,α

([0,t

0

/8)×B

0

r0/8

) 6 C(kv h k L

([0,t

0

/4)×B

0

r0/4

) + kb h k L

(B

0

r0/2

) ) 6 Ckv h k L

2

([0,t

0

/2)×B

0

r0/2

;t

1−2s

) + Ck∇gk L

(B

0

r0/2

×[0,A

3

])

6 Ck∇vk L

2

([0,t

0

)×B

r00

;t

1−2s

) + Ck∇gk L

(B

0

r0/2

×[0,A

3

])

6 A 2

for every small h. By the Arzel` a-Ascoli Theorem, passing to the limit as |h| → 0, we conclude that z 7→ ∇ x v(z) ∈ C([0, t 0 /8) × B r 0

0

/8 ) and estimate (41) holds for all β ∈ (0, α).

Since the map x 7→ g(x, v(0, x)) ∈ C 0,β (B 0 r

0

/8 ), estimate (42) follows from [[3], Lemma 4.5].

Lemma 3.4. Let w be a solution to (37) in the sense of (38). Then there exist p 0 > 2 (s) and R 0 ∈ (0, R) such that w ∈ L p

0

(B R +

0

).

Proof. By (3), there are δ > 0 and R δ ∈ (0, R) such that

(43) (λ + |x| 2s |h(x)|) 6 λ + δ < Λ N,s , for all x ∈ B R 0

δ

.

Let β > 1. For all L > 0, we define F L (τ ) = |τ | β if |τ | < L and F L (τ ) = βL β−1 |τ | + (1 − β)L β if

|τ | > L. Put G = 1

β F L F L 0 . It is easy to verify that, for all τ ∈ R,

(44) τ G L (τ ) 6 τ 2 G 0 L (τ ), τ G L (τ ) 6 (F L (τ )) 2 , (F L 0 (τ )) 2 6 βG 0 L (τ ).

(12)

Let η ∈ C c (B R +

δ

∪ B R 0

δ

) be a radial cut-off function such that η ≡ 1 in B + R

δ

/2 . It is clear that ζ := η 2 G L (w) ∈ H 1 (B + R ; t 1−2s ) and F L (w) ∈ H 1 (B + R ; t 1−2s ). Using ζ as a test in (37), from (43) and integration by parts, we have that

Z

B

+

t 1−2s η 2 |∇w| 2 G 0 L (w)dtdx − κ s (λ + δ) Z

B

0

|x| −2s η 2 wG L (w)dx

6 −2 Z

B

+

t 1−2s η∇w · ∇ηG L (w)dtdx

+ c Z

B

0

η 2 wG L (w)dx + c Z

B

0

η 2 |w| 2

(s)−2 wG L (w)dx,

for some positive c > 0 depending only on C f , s, p, N . By Young’s inequality and (44), we have that, for every σ > 0,

2(η∇w) · (w∇η) G L (w) w

6 σ

2 η 2 |∇w| 2 G 0 L (w) + 2

σ |∇η| 2 (F L (w)) 2 , hence we obtain that

 1 − σ

2

 Z

B

+

t 1−2s η 2 |∇w| 2 G 0 L (w) dt dx − κ s (λ + δ) Z

B

0

|x| −2s η 2 wG L (w) dx

6 c Z

B

0

η 2 (|w| 2

(s)−2 + 1)w G L (w) dx + 2 σ

Z

B

+

t 1−2s |∇η| 2 ψ 2 dt dx,

where we have set ψ = F L (w). Therefore by (44) 1

β

 1 − σ

2

 Z

B

+

t 1−2s η 2 |∇ψ| 2 dt dx − κ s (λ + δ) Z

B

0

|x| −2s (ηψ) 2 dx

6 c Z

B

0

(|w| 2

(s)−2 + 1)(ηψ) 2 dx + 2 σ

Z

B

+

t 1−2s |∇η| 2 ψ 2 dt dx.

Since |∇(ηψ)| 2 6 (1 + σ)η 2 |∇ψ| 2 + (1 + σ 12 |∇η| 2 , we have that 1

β(1 + σ)

 1 − σ

2

 Z

B

+

t 1−2s |∇(ηψ)| 2 dt dx − κ s (λ + δ) Z

B

0

|x| −2s (ηψ) 2 dx

6 c Z

B

0

(|w| 2

(s)−2 + 1)(ηψ) 2 dx + C(c, β, σ, R δ ) Z

B

+

t 1−2s ψ 2 dt dx

for some positive C(c, β, σ, R δ ) > 0 depending only on c, σ, R δ , and β. By H¨ older inequality and Lemma 2.6, we have that

Z

B

0

(|w| 2

(s)−2 + 1)(ηψ) 2 dx

6 e S N,s

 Z

B

0

|w| 2

(s) dx



2∗ (s)−22∗ (s)

+ |B R 0

δ

|

2sN

 Z

B

+

t 1−2s |∇(ηψ)| 2 dt dx.

We deduce that (45) A

Z

B

+

t 1−2s |∇(ηψ)| 2 dt dx − κ s (λ + δ) Z

B

0

|x| −2s (ηψ) 2 dx 6 const Z

B

+

t 1−2s ψ 2 dtdx, for some positive const > 0 depending only on C f , s, p, N, β, σ, R δ , where

A = 1

(1 + σ)β

 1 − σ

2

 − c e S N,s

 Z

B

0

|w| 2

(s) dx



2∗ (s)−22∗ (s)

+ |B R 0

δ

|

2sN

 .

(13)

From Hardy inequality (20), we have that A

Z

B

+

t 1−2s |∇(ηψ)| 2 dt dx − κ s (λ + δ) Z

B

0

|x| −2s (ηψ) 2 dx

>



A − λ + δ Λ N,s

 Z

B

+

t 1−2s |∇(ηψ)| 2 dt dx, and, by (43), we can choose β sufficiently close to 1 and σ, R δ sufficiently small such that

A − λ + δ Λ N,s

> 0.

Hence we have that C

Z

B

+

t 1−2s |∇(ηψ)| 2 dt dx 6 Z

B

+

t 1−2s ψ 2 dtdx,

for some constant C > 0 depending on f, h, s, p, N, β, ε, w, λ, δ. From Lemma 2.6 it follows that C e S −1 N,s

 Z

B

0

|ηF L (w)| 2

(s) dx



2∗ (s)2

6

Z

B

+

t 1−2s |w| dtdx for all L > 0.

Hence by taking the limit as L → +∞

(46) C e S N,s −1

 Z

B

0

Rδ /2

|w| β2

(s) dx



2∗ (s)2

6

Z

B

+

t 1−2s |w| dtdx.

The conclusion follows since β > 1 and H 1 (B + R ; t 1−2s ) ,→ L q (B + R ; t 1−2s ) for some q > 2, see for instance [[8], Theorem 1.2].

Remark 3.5. From Lemma 3.4, we deduce that, if w ∈ H 1 (B R + ; t 1−2s ) is a weak solution to (37), then |x| λ

2s

+ h + f (x,w) w ∈ L q loc (B R 0 \ {0}) for some 2s N < q 6 p−2 p

0

and hence, from Lemma 3.2, we conclude that w ∈ C loc 0,α (B r + \ {0}), ∇ x v ∈ C loc 0,β (B r + \ {0}), and t 1−2s v t ∈ C loc 0,β (B r + \ {0}) for all r ∈ (0, R) and some 0 < β < α < 1.

The following Pohozaev-type identity holds.

Theorem 3.6. Let w solves (38). Then for a.e. r ∈ (0, R) there holds (47) − N − 2s

2

 Z

B

r+

t 1−2s |∇w| 2 dz − κ s λ Z

B

0r

w 2

|x| 2s dx



+ r 2

 Z

S

r+

t 1−2s |∇w| 2 dS − κ s λ Z

∂B

r0

w 2

|x| 2s dS 0



= r Z

S

r+

t 1−2s

∂w

∂ν

2

dS − κ s 2

Z

B

0r

(N h + ∇h · x)w 2 dx + rκ s 2

Z

∂B

0r

hw 2 dS 0

+ rκ s Z

∂B

0r

F (x, w) dS 0 − κ s Z

B

r0

[∇ x F (x, w) · x + N F (x, w)] dx and

(48) Z

B

+r

t 1−2s |∇w| 2 dz − κ s λ Z

B

0r

w 2

|x| 2s dx

= Z

S

r+

t 1−2s ∂w

∂ν w dS + κ s Z

B

0r



hw 2 + f (x, w)w

 dx.

Proof. We write our problem in the form

( div(t 1−2s ∇w) = 0, in B R + ,

− lim t→0

+

t 1−2s w t = G(x, w), on B 0 R , where G ∈ C 1 (B R 0 \ {0} × R), G(x, %) = κ s λ

|x|

2s

% + h(x)% + f (x, %).

(14)

We have, on B R + , the formula

(49) div  1

2 t 1−2s |∇w| 2 z − t 1−2s (z · ∇w)∇w



= N − 2s

2 t 1−2s |∇w| 2 − (z · ∇w) div(t 1−2s ∇w).

Let ρ < r < R. Now we integrate by parts over the set O δ := (B + r \ B ρ + ) ∩ {(t, x), t > δ} with δ > 0. We have

N − 2s 2

Z

O

δ

t 1−2s |∇w(z)| 2 dz = − 1 2 δ 2−2s

Z

B

0

r2 −δ2

\B

0

ρ2 −δ2

|∇w| 2 (δ, x)dx

+ δ 2−2s Z

B √

0

r2 −δ2

\B

0

ρ2 −δ2

|w t | 2 (δ, x)dx

+ r 2

Z

S

+r

∩{t>δ}

t 1−2s |∇w| 2 dS − r Z

S

+r

∩{t>δ}

t 1−2s

∂w

∂ν

2

dS

− ρ 2 Z

S

+ρ

∩{t>δ}

t 1−2s |∇w| 2 dS + ρ Z

S

+ρ

∩{t>δ}

t 1−2s

∂w

∂ν

2

dS

+ Z

B

0

r2 −δ2

\B

0

ρ2 −δ2

(x · ∇ x w(δ, x)) δ 1−2s w t (δ, x) dx.

We now claim that there exists a sequence δ n → 0 such that

n→∞ lim

"

1 2 δ n 2−2s

Z

B

r0

|∇w| 2n , x)dx + δ 2−2s n Z

B

0r

|w t | 2n , x)dx

#

= 0.

If no such sequence exists, we would have

lim inf

δ→0

"

1 2 δ 2−2s

Z

B

0r

|∇w| 2 (δ, x)dx + δ 2−2s Z

B

r0

|w t | 2 (δ, x)dx

#

> C > 0

and thus there exists δ 0 > 0 such that 1

2 δ 2−2s Z

B

r0

|∇w| 2 (δ, x)dx + δ 2−2s Z

B

r0

|w t | 2 (δ, x)dx > C

2 for all δ ∈ (0, δ 0 ).

It follows that 1 2 δ 1−2s

Z

B

0r

|∇w| 2 (δ, x)dx + δ 1−2s Z

B

0r

|w t | 2 (δ, x)dx > C

2δ for all δ ∈ (0, δ 0 )

and so integrating the above inequality on (0, δ 0 ) we contradict the fact that w ∈ H 1 (B + R ; t 1−2s ).

Next, from the Dominated Convergence Theorem, Lemma 3.2 and Remark 3.5, we have that lim

δ→0

Z

B

0

r2 −δ2

\B

0

ρ2 −δ2

(x · ∇ x w(δ, x)) δ 1−2s w t (δ, x) dx = − Z

B

r0

\B

ρ0

(x · ∇ x w) G(x, w) dx.

We conclude that (replacing O δ with O δ

n

, for a sequence δ n → 0) that

(50) N − 2s 2

Z

B

+r

\B

+ρ

t 1−2s |∇w(z)| 2 dz = r 2

Z

S

r+

t 1−2s |∇w| 2 dS − r Z

S

+r

t 1−2s

∂w

∂ν

2

dS

− ρ 2 Z

S

ρ+

t 1−2s |∇w| 2 dS − ρ Z

S

+ρ

t 1−2s

∂w

∂ν

2

dS − Z

B

r0

\B

ρ0

(x · ∇ x w) G(x, w) dx.

Riferimenti

Documenti correlati

Postconditioning (PostC) can be obtained either with brief cycles of ischemia/reperfusion (I- PostC) or with a direct targeting of mitochondria with Diazoxide

The use of monotonicity methods to study unique con- tinuation properties of solutions to elliptic partial differential equations dates back to the pioneering paper by Garofalo and

Monotonicity methods were recently used in [7, 8, 9] to prove not only unique continuation but also precise asymptotics near singularities of solutions to linear and semilinear

On the other hand, it can be proven for example using Morse Theory in ordered Banach spaces (see [2]), that the equation admits a biradial sign-changing solution having biradial

Through an Almgren type monotonicity formula and separation of variables, we describe the exact asymptotics near the singularity of solutions to at most critical semilinear

Aharonov-Bohm magnetic potentials are associated to thin solenoids: if the radius of the solenoid tends to zero while the flux through it remains constant, then the particle is

Aharonov-Bohm magnetic potentials are associated to thin solenoids: if the radius of the solenoid tends to zero while the flux through it remains constant, then the particle is