• 沒有找到結果。

ON THE FRACTIONAL LANDIS CONJECTURE ANGKANA R ¨ULAND AND JENN-NAN WANG Abstract.

N/A
N/A
Protected

Academic year: 2022

Share "ON THE FRACTIONAL LANDIS CONJECTURE ANGKANA R ¨ULAND AND JENN-NAN WANG Abstract."

Copied!
27
0
0

加載中.... (立即查看全文)

全文

(1)

ON THE FRACTIONAL LANDIS CONJECTURE

ANGKANA R ¨ULAND AND JENN-NAN WANG

Abstract. In this paper we study a Landis-type conjecture for fractional Schr¨odinger equa- tions of fractional power s ∈ (0, 1) with potentials. We discuss both the cases of differentiable and non-differentiable potentials. On the one hand, it turns out for differentiable potentials with some a priori bounds, if a solution decays at a rate e−|x|1+, then this solution is triv- ial. On the other hand, for s ∈ (1/4, 1) and merely bounded non-differentiable potentials, if a solution decays at a rate e−|x|α with α > 4s/(4s − 1), then this solution must again be trivial. Remark that when s → 1, 4s/(4s − 1) → 4/3 which is the optimal exponent for the standard Laplacian. For the case of non-differential potentials and s ∈ (1/4, 1), we also derive a quantitative estimate mimicking the classical result by Bourgain and Kenig.

1. Introduction

In this work, we study a Landis-type conjecture for the fractional Schr¨odinger equation,

(1) ((−∆)s+ q)u = 0 in Rn

with s∈ (0, 1) and

(2) |q(x)| ≤ 1.

Roughly speaking, we are interested in the maximal vanishing rate of solutions to this equation at infinity. For s = 1, in Section 3.5 in [KL88] V.A. Kondratev and E.M. Landis conjectured that if |q(x)| ≤ 1 and |u(x)| ≤ C0 satisfies |u(x)| ≤ exp(−C|x|1+), then u ≡ 0. The Landis conjecture was disproved by Meshkov [Mes91], who constructed a potential q and a nontrivial uwith |u(x)| ≤ C exp(−C|x|43). He also showed that if |u(x)| ≤ C exp(−C|x|43+), then u≡ 0.

In their seminal work, Bourgain and Kenig [BK05] derived a quantitative form of Meshkov’s result in their resolution of Anderson localization for the Bernoulli model in higher dimensions.

It should be pointed out that in Meshkov’s counterexample both q and u are complex -valued functions. In other words, the exponent 4/3 is optimal in the complex case (which corresponds to the situation of systems). The proof in [BK05] is based on the Carleman method. In the spirit of the Carleman method, several extensions have been made in [CS99,Dav14,DZ17,DZ18,LW14], which also take singular drift coefficients and potentials into account.

In view of Meshkov’s counterexample, Kenig [Ken06] refined the Landis conjecture and asked whether this conjecture is true for real -valued potentials and solutions. In 2005, Kenig, Silvestre and the second author [KSW15] confirmed the Landis conjecture (in a quantitative form) when n = 2 and q ≥ 0. This result was later extended to the more general situation with ∆ being replaced by any second order elliptic operator [DKW17]. In the very recent preprint [DKW18], this is further improved by also allowing for (exponentially) small negative contributions in the potential. The Landis conjecture in the real case with n = 1 was recently studied by Rossi [Ros18].

The main theme of this paper is to investigate a Landis-type conjecture for fractional Schr¨odinger equations. We will consider both qualitative and quantitative estimates when the potentials are

Wang is supported in part by MOST 105-2115-M-002-014-MY3.

1

(2)

either differentiable or simply bounded. Similar to the original Landis conjecture (for the Laplace operator), we are concerned with the maximal decay rate of non-trivial solutions. We are espe- cially interested in understanding how the decay rate depends on the fractional power s∈ (0, 1).

Detailed statements of our results are described below.

1.1. Qualitative estimates. We first discuss the qualitative behaviour of solutions to (1). Here we will show that if the potential q satisfies an additional regularity estimate, then independently of the value of s∈ (0, 1) there are no super-exponentially decaying solutions to (1).

Theorem 1. Let s∈ (0, 1) and assume that u ∈ Hs(Rn) is a solution to (1) such that q satisfies q∈ C1(Rn), (2), and in addition

(3) |x · ∇q(x)| ≤ 1

holds. Suppose that u further satisfies the following decay behaviour: there exists α >1 such that ˆ

Rn

e|x|α|u|2dx≤ C < ∞.

(4)

Then u≡ 0.

For s = 1, a similar qualitative estimate as in Theorem 1 with a differentiable potential satisfying (2) and (3) was proved by Meshkov [Mes89]. Without the additional regularity result on q, it is still possible to prove a qualitative decay result. However, as our argument for this does not distinguish between the real and complex situation, the obtained decay deteriorates.

Theorem 2. Let s∈ (1/4, 1) and assume that u ∈ Hs(Rn, C) is a solution to (1). Suppose that (2) holds and u further satisfies the following decay behaviour: there exists α > 4s−14s such that

ˆ

Rn

e|x|α|u|2dx≤ C < ∞.

(5)

Then u≡ 0.

We emphasize that as s→ 1 in both of our main results, the identified critical decay exponents correspond to the ones from the case s = 1. Moreover, in the first result, Theorem1, the critical decay rate does not depend on the value of s ∈ (0, 1). It is thus natural to ask whether the derived decay exponents are optimal or rather an artifact of our argument. Let us comment on this. For real-valued problems (i.e. scalar equations), we expect that the exponential decay (independent of the value of s∈ (0, 1)) as the critical decay behaviour is sharp. Indeed, as in [KSW15] and [BK05] it is possible to relate the decay behaviour at infinity to the local maximal vanishing rate at zero (if growth conditions are assumed, which are necessary due to the global character of the problem). Analogous arguments as in the classical case s = 1 would lead to the conjecture that when considering

• the equation (1) with potentials of the size|q(x)| ≤ M (instead of |q(x)| ≤ 1),

• and solutions u(x) which satisfy the growth bounds kukL(Rn)≤ C0andkukL(B1)≥ 1, a lower bound of the form

(6) kukL(Br)≥ rCM

1 2s,

holds. Here C = C(C0) > 0 and r∈ (0, r0) for some sufficiently small constant r0 >0. Results of this flavour have been proved for eigenfunctions or equations with differentiable potentials (with dependences on the C1 norm of the potentials, that is, M in (6) is the size of kqkC1(Rn)) in [R¨ul17a,Zhu15] (on compact manifolds or bounded domains, respectively). For the spectral fractional Laplacian and its eigenfunctions on compact manifolds these dependences are indeed immediate consequences from the corresponding ones of the Laplacian.

(3)

We recall that∇q satisfies (3). Taking this expected quantitative maximal rate of vanishing (6) for granted and relying on a scaling argument as in [BK05,KSW15], i.e. considering solutions uR,x0(x) for x0∈ BR\ BR/2 of the correspondingly rescaled version of (1), then suggest global lower bounds of the type

|zinf0|=R sup

|z−z0|<1|u(z)| ≥ exp(−CR log(R)) for R  1,

where C = C(C0) > 0. Hence, this strongly suggests that for the class of potentials under con- sideration, Theorem1 is essentially sharp (possibly up to logarithmic contributions). However, we further believe that as in [KSW15], at least under sign conditions on the potential and in one dimension (which on the level of the Cafferelli-Silvestre extension corresponds to the two- dimensional setting from [KSW15] in which complex analysis tools are available), it might be possible to reduce the necessary regularity for q to Lregularity.

In view of Meshkov’s example [Mes91], in the case of Theorem2, at least the growth behaviour for s → 1 is expected to be optimal. As in the case s = 1, the Carleman estimate of Theorem 5 which lies at the core of the argument for Theorem2 is of perturbative character. It hence does not distinguish between the real vs the complex, i.e. the scalar vs the systems cases.

It would be interesting to investigate whether also for s ∈ (1/4, 1) there are Meshkov-type examples saturating the proposed exponents from Theorem2. We remark that the restriction to s∈ (1/4, 1) seems necessary as long as we only consider radial weight functions in our Carleman estimates (due to the subelliptic nature of these estimates). We seek to prove the results of Theorems1 and2by combining elliptic estimates with Carleman estimates.

1.2. Quantitative estimates. In this subsection, we present a quantitative version of Theorem 2. Note that here q∈ L(Rn) and that (2) is satisfied.

Theorem 3. Let s∈ (1/4, 1) and assume that u ∈ Hs(Rn, C) is a solution to (1). Suppose that u further satisfieskukL2({|x|<1})= 1 and that there exists a constant C0>0 such that

kukL(Rn)≤ C0.

Then there exists a constant C = C(n, s, C0) > 0 such that for R > 0 large

|xinf0|=RkukL({|x−x0|<1})≥ Ce−CR

4s 4s−1log R.

We remark that by virtue of the ellipticity of the problem, it does not really matter in which topology one works for the lower bounds. For instance, it would have equally been possible to derive similar results under the assumption that|u(0)| = 1.

This result is similar in flavour to the decay estimates in [BK05]. However both the arguments in the qualitative and the quantitative settings involve new intricacies and technical challenges due to the nonlocal character of the equations at hand. Since lower bound estimates for nonlocal equations pose serious difficulties, as in various other works on (quantitative) unique continuation [FF14,FF15,BG17,Gar17,R¨ul15,R¨ul17a,R¨ul17b,RS17], we opt for working with the Caffarelli- Silvestre extension [CS07], c.f. (7), (8) in Section2, instead of dealing with the nonlocal equation (1) directly. This allows us to investigate a local (degenerate) elliptic equation. It however comes at the expense of having to study this equation in n+1 instead of n dimensions. In the additional dimension, the control on the solution can hence only be derived through the equation. This implies that we always have to transfer information from the boundary to the bulk and vice versa. In the qualitative estimates, we for instance have to show that (exponential) decay on the boundary implies (exponential) decay in the bulk. Similarly, we have to transfer upper and lower bounds in the quantitative results from the boundary into the bulk and vice versa. This poses non-trivial challenges, which however are overcome by an ingredient which was used in [RS17]:

(4)

As one of our key tools which allows us to switch between the bulk and the boundary we rely on a boundary-bulk interpolation inequality, c.f. Proposition2.4.

1.3. Organization of the article. This paper is organized as follows. In Section 2, we will collect several preliminary results that are needed in the proofs of the main theorems. We then derive Carleman estimates for the fractional Laplacian (−∆)s with differentiable or non- differentiable potentials in Section3. The proofs of the qualitative estimates, Theorem 1and2, are given in Section4. Finally, in Section5, we will prove Theorem3.

2. Auxiliary Results

2.1. The Cafferelli-Silvestre extension. In the sequel, it will be convenient to localize the problem at hand. This will be achieved by means of the Caffarelli-Silvestre extension [CS07]

which allows us to address the problem at hand by relying on tools for local equations. To this end, for s∈ (0, 1) and u ∈ Hs(Rn) we consider a solution ˜u∈ ˙H1(Rn+1+ , x1−2sn+1) :={v : Rn+1+ → R :

´

Rn+1+

x1−2sn+1|∇v|2dx≤ C < ∞} of the degenerate elliptic equation (whose weight however still is in the Muckenhoupt class),

∇ · x1−2sn+1∇˜u = 0 in Rn+1+ ,

˜

u= u on Rn× {0}.

(7)

We recall that by the observations in [CS07], (−∆)su(x) = cn,s lim

xn+1→0x1−2sn+1n+1u(x)˜

for some constant cn,s 6= 0. In particular, the equation (1) can be reformulated as the local, degenerate elliptic equation

∇ · x1−2sn+1∇˜u = 0 in Rn+1+ ,

˜

u= u on Rn× {0}, cn,s lim

xn+1→0x1−2sn+1n+1u˜= qu on Rn× {0}.

(8)

This however comes at the expense of adding a new variable, in which we have to infer control by exploiting the equation.

When dealing with this equation, it will be convenient to also introduce the following notations for the underlying domains and the related weighted function spaces. For Ω⊂ Rn+1+ , x0∈ Rn+1+ , r, R >0, we denote

Br+(x0) :={x ∈ Rn+1+ : |x − x0| ≤ r}, B0r(x0) :={x ∈ Rn× {0} : |x − x0| ≤ r}, Br+:= Br+(0), Br0 := Br0(0),

A+r,R:={x ∈ Rn+1+ : r≤ |x| ≤ R}, A0r,R:={x ∈ Rn× {0} : r ≤ |x| ≤ R}, H1(Ω, x1−2sn+1) :={v : Ω ∩ Rn+1+ → R :

ˆ

Ω∩Rn+1+

x1−2sn+1(|v|2+|∇v|2)dx≤ C < ∞},

H1(S+n, θ1−2sn ) :={v : S+n → R : ˆ

S+n

θn1−2s(|v|2+|∇S+nv|2)dθ≤ C < ∞}, θn= xn+1

|x| . As an important elliptic estimate which we will be using frequently we recall Caccioppoli’s inequality.

(5)

Lemma 2.1. Let s ∈ (0, 1) and ˜u ∈ H1(B4r+, x1−2sn+1) be a solution to (7). Then, there exists C= C(n, s) > 0 such that

kxn+11−2s2 ∇˜ukL2(Br+)≤ C



r−1kxn+11−2s2 u˜kL2(B2r+)+kukL122(B2r0 )k limx

n+1→0x1−2sn+1n+1u˜kL122(B02r)

 . Proof. The proof follows as for instance in Lemma 4.5 in [RS17], where however the boundary terms are estimated by an L2− L2estimate instead of an Hs− H−sestimate.  2.2. Boundary decay implies bulk decay. In order to deal with the original nonlocal problem (1), we seek to apply methods which were developed for (quantitative) unique continuation results for the local equation (8). Hence, we first translate the decay behaviour that is valid on Rn to decay behaviour which also holds on Rn+1+ . To this end, we heavily rely on interior and boundary three balls estimates for the degenerate elliptic equation (8).

Proposition 2.2. Let s∈ (0, 1) and u ∈ Hs(Rn) be a solution to (1). Assume that (2) holds and there exist constants C, β≥ 1 such that

(9) ke|x|β/2ukL2(Rn)≤ C.

Then, there exist constants C1, c1 > 0 such that for all x = (x0, xn+1) ∈ Rn+1+ the Caffarelli- Silvestre extension u(x) satisfies˜

|˜u(x0, xn+1)| ≤ C1e−c1|(x0,xn+1)|β.

In order to infer the claimed interior decay, we rely on propagation of smallness estimates.

Here we make use of two types of propagation of smallness results: The first being an interior propagation of smallness while the second one is a boundary-bulk propagation of smallness estimate. In order to use tools from the quantitative analysis of elliptic equations, in the sequel, we view (1) in terms of its Caffarelli-Silvestre extension (8).

Proposition 2.3. Let s ∈ (0, 1) and ˜u ∈ H1(B+4, x1−2sn+1) be a solution to (7). Assume that r∈ (0, 1) and x0= (x00,5r)∈ B2+. Then, there exists α= α(n, s)∈ (0, 1) such that

k˜ukL(B+2r(x0)) ≤ Ck˜ukαL(Br+(x0))k˜uk1−αL(B4r+(x0)).

Proof. As (x0)n+1= 5r, this follows from a standard interior L2three balls estimate (c.f. Propo- sition 5.4 in [RS17]) together with L2− L estimates for uniformly elliptic equations.  Proposition 2.4. Let s ∈ (0, 1) and let ˜u ∈ H1(Rn+1+ , x1−2sn+1) be a solution to (8) with q ∈ L(Rn). Assume that x0∈ Rn× {0}. Then,

(a) there exist α = α(n, s)∈ (0, 1) and c = c(n, s) ∈ (0, 1) such that kxn+11−2s2 u˜kL2(B+cr(x0))

≤C

kxn+11−2s2 u˜kL2(B+16r(x0))+ r1−skukL2(B016r(x0))

α

×

×



rs+1k limx

n+1→0x1−2sn+1n+1u˜kL2(B16r0 (x0))+ r1−skukL2(B016r(x0))

1−α

+ C

kxn+11−2s2 u˜kL2(B16r+ (x0))+ r1−skukL2(B016r(x0))

1+s2s

×

×



rs+1k limx

n+1→0x1−2sn+1n+1u˜kL2(B16r0 (x0))+ r1−skukL2(B016r(x0))

1−s1+s .

(6)

(b) there exist α = α(n, s)∈ (0, 1) and c = c(n, s) ∈ (0, 1) such that

k˜ukL(B+cr 2

)≤Crn2 

rs−1kxn+11−2s2 u˜kL2(B16r+ )+kukL2(B016r)

α

r2skqukL2(B016r)+kukL2(B16r0 )

1−α

+ Crn2 

rs−1kxn+11−2s2 u˜kL2(B+16r)+kukL2(B016r)1+s2s

r2skqukL2(B016r)+kukL2(B016r)1−s1+s + Crn2rskqukL122(B16r0 )kukL122(B016r).

(10)

Proof. The proof relies on a splitting argument and the boundary-bulk interpolation estimates from Propositions 5.10-5.12 (also Proposition 5.6) in [RS17]. In order to infer the claim, we argue in two steps, first deriving a suitable L2 estimate and then upgrading this to an L estimate.

By scaling, it suffices to prove the estimate for r = 1. Without loss of generality, we can take x0= 0.

Step 1: The L2 estimate. For the L2 estimate we rely on Propositions 5.10-5.12 in [RS17].

Here we distinguish between the cases s∈ [12,1) and s∈ (0,12).

Step 1a: The case s∈ [12,1). In order to invoke the estimate from [RS17], we split our solution uinto two parts ˜u= u1+ u2. The function u1 deals with the Dirichlet data

∇ · x1−2sn+1∇u1= 0 in Rn+1+ , u1= ζu in Rn× {0}.

Here ζ∈ C0(B160 ) is a smooth cut-off function, which is equal to one in B80. We will estimate u1

by bounds on the Caffarelli-Silvestre extension. The function u2= ˜u− u1 in turn is admissible in Propositions 5.10-5.12 in [RS17], i.e., u2|B80 = 0.

We begin with the estimate for u1: Invoking Lemma 4.2 in [RS17], we obtain the bound

kxn+11−2s2 u1kL2(Rn+1+ )≤ CkζukHs−1(Rn)≤ CkζukL2(Rn)≤ CkukL2(B160 ). (11)

The estimate for u2 follows from Proposition 5.10 in [RS17]. The result assert that for each s∈ (12,1) there exists a constant c = c(s, n)∈ (0, 1) and α = α(s, n) ∈ (0, 1) such that

(12) kxn+11−2s2 u2kL2(B+c)≤ Ckxn+11−2s2 u2kαL2(B+2)k limx

n+1→0x1−2sn+1n+1u2k1−αL2(B20).

We modify this by interpolation in order to obtain an estimate where the normal derivative of u2 is measured in the H−2snorm. To this end, we note that for any w by interpolation and the characterization of the trace map (c.f. Step 1 in the proof of Proposition 5.11 in [RS17]):

kwkL2(Rn)≤ Ckwk

2s 1+s

H1−s(Rn)kwk

1−s 1+s

H−2s(Rn)

≤ C

kxn+12s−12 wkL2(Rn+1+ )+kxn+12s−12 ∇wkL2(Rn+1+ )

1+s2s

kwk

1−s 1+s

H−2s(Rn)

≤ C µ1−s

kxn+12s−12 wkL2(Rn+1+ )+kxn+12s−12 ∇wkL2(Rn+1+ )

+ µ−2skwkH−2s(Rn)

. (13)

(7)

Applying this to w = ηx1−2sn+1n+1u2, where η is a smooth, radial cut-off function which is equal to one on B2+ and vanishes outside of B4+ gives us

(14)

k limx

n+1→0x1−2sn+1n+1u2kL2(B02)

≤C µ1−s

kxn+11−2s2n+1u2kL2(B4+)+kxn+12s−12 ∇(ηx1−2sn+1n+1u2)kL2(Rn+1+ )



−2sk limx

n+1→0ηx1−2sn+1n+1u2kH−2s(Rn)

 .

Similar to the proof of Proposition 5.11 in [RS17], we now estimate each term on the right hand side of (14). The last term gives us

(15) k limx

n+1→0ηx1−2sn+1n+1u2kH−2s(Rn)≤ Ck limx

n+1→0x1−2sn+1n+1u2kH−2s(B08). Applying Caccioppoli’s inequality in Lemma2.1(with zero Dirichlet condition) implies (16) kxn+11−2s2n+1u2kL2(B4+)≤ Ckxn+11−2s2 u2kL2(B8+).

It remains to estimate the second term on the right hand side of (14). Indeed, for the resulting bulk term we have

(17)

kxn+12s−12 ∇(ηx1−2sn+1n+1u2)kL2(Rn+1+ )≤ kxn+11−2s2 (∇η)(∂n+1u2)kL2(Rn+1+ )+kxn+11−2s2 η∇0n+1u2kL2(Rn+1+ )

+kxn+12s−12 η∂n+1x1−2sn+1n+1u2kL2(Rn+1+ )

≤ kxn+11−2s2 (∂n+1u2)kL2(B+4)+kxn+11−2s20n+1u2kL2(B4+)

+kxn+11−2s2 η∆0u2kL2(Rn+1+ )

≤ Ckxn+11−2s2 u2kL2(B+8).

Here we first used the triangle inequality, then the support condition for η and the equation for u2 and finally applied Caccioppoli’s inequality (twice for the last two terms, noting that ∇0u2 solves a similar problem).

Substituting (15)-(17) into (14) and optimizing the resulting estimate in µ > 0 gives k limx

n+1→0x1−2sn+1n+1u2kL2(B02)≤ Ckxn+11−2s2 u2k

2s 1+s

L2(B+8)k limx

n+1→0x1−2sn+1n+1u2k

1−s 1+s

H−2s(B80). Inserting this into (12) leads to

(18) kxn+11−2s2 u2kL2(B+c)≤ Ckxn+11−2s2 u2kαL˜2(B+8)k limx

n+1→0x1−2sn+1n+1u2k1− ˜H−2sα (B08)

where ˜α = 1−s1+sα+1+s2s . By slight abuse of notation, in the sequel, we simply drop the tilde.

Combining the two bounds (11), (18) and k limx

n+1→0x1−2sn+1n+1u1kH−2s(B016)≤ k(−∆)su1kL2(Rn)≤ ku1kL2(Rn)≤ CkukL2(B160 ),

(8)

and applying the triangle inequality leads to kxn+11−2s2 u˜kL2(Bc+)

≤C

kxn+11−2s2 u˜kL2(B+16)+kukL2(B160 )α k limx

n+1→0x1−2sn+1n+1u˜kH−2s(B016)+kukL2(B160 )

1−α

≤C

kxn+11−2s2 u˜kL2(B+16)+kukL2(B160 )α k limx

n+1→0x1−2sn+1n+1u˜kL2(B160 )+kukL2(B016)

1−α

. (19)

This already implies the claim of (a). In order to exploit it for the proof of (b), we strengthen the estimate slightly. By Caccioppoli’s inequality of Lemma2.1 (now with an L2 estimate for the boundary contributions), we can further upgrade (19) to

kxn+11−2s2 u˜kL2(Bc+˜)+kxn+11−2s2 ∇˜ukL2(B+˜c)

≤C

kxn+11−2s2 u˜kL2(B16+)+kukL2(B160 )α k limx

n+1→0x1−2sn+1n+1u˜kL2(B016)+kukL2(B016)

1−α

+ Ck limx

n+1→0x1−2sn+1n+1u˜kL122(B160 )kukL122(B016)

(20)

with ˜c= c/2.

Step 1b: The case s ∈ (0,12). The case s ∈ (0, 1/2) is similar as the case discussed above and relies on a splitting strategy. As above, the estimate for u1 is a direct consequence of the boundary bulk estimates for the Caffarelli-Silvestre extension. Thus, the main remaining estimate is the derivation of the corresponding analogue of (18). As in the proof of (18) this follows an application of the corresponding L2result from [RS17] (Proposition 5.12) and interpolation. More precisely, Proposition 5.12 in [RS17] implies that for some c = c(n, s) > 0 and α = α(n, s)∈ (0, 1) we have

kxn+11−2s2 u2kL2(B+c)

≤ C(kxn+11−2s2 u2kαL2(B2+)k limx

n+1→0x1−2sn+1n+1u2k1−αL2(B02)+k limx

n+1→0x1−2sn+1n+1u2kL2(B20)).

In order to pass from this estimate which involves an L2 norm of the weighted Neumann data to an estimate which involves its H−2s norm, we apply the interpolation estimate (13) as in the case s∈ (12,1). With this estimate at hand, the analogues of (19) and (20) then follow by combining the estimates of the splitting argument as above. Note that (20) now becomes

kxn+11−2s2 u˜kL2(B+˜c)+kxn+11−2s2 ∇˜ukL2(Bc+˜)

≤C

kxn+11−2s2 u˜kL2(B+16)+kukL2(B016)

α k limx

n+1→0x1−2sn+1n+1˜ukL2(B160 )+kukL2(B160 )

1−α

+ C

kxn+11−2s2 u˜kL2(B16+)+kukL2(B016)1+s2s  k limx

n+1→0x1−2sn+1n+1u˜kL2(B160 )+kukL2(B016)

1−s1+s

+ Ck limx

n+1→0x1−2sn+1n+1u˜kL122(B016)kukL122(B160 ). (21)

Step 2: The L estimate. In order to pass from the L2-based bounds from step 1 to L based estimates, we rely on an estimate due to Jin, Li, Xiong [JLX11] (Proposition 2.4 (i), c.f.

(9)

also Proposition 3.2 in [FF14]), which states that under our conditions on B1/2+ it holds k˜ukL(B1/2+ )≤ C(kxn+11−2s2 u˜kL2(B+1)+kxn+11−2s2 ∇˜ukL2(B1+)).

(22)

Combining this with the estimate (21) and inserting the identity cn,s lim

xn+1→0x1−2sn+1n+1u˜ = qu entails

k˜ukL(B+˜c

2

)≤C

kxn+11−2s2 ˜ukL2(B16+)+kukL2(B160 )

α

kqukL2(B016)+kukL2(B016)

1−α

+ C

kxn+11−2s2 u˜kL2(B+16)+kukL2(B160 )1+s2s

kqukL2(B016)+kukL2(B160 )1−s1+s + CkqukL122(B160 )kukL122(B016).

(23)

Based on this, we can also derive a pure Lestimate. Indeed, using H¨older’s inequality together with the L2integrability of the function (0, 1)3 t 7→ t1−2s2 ∈ R for s ∈ (0, 1) on bounded domains results in

k˜ukL(B+c˜

2

)≤C(1 + kqkL(B160 ))1−α

k˜ukL(B+16)+kukL(B016)α

kuk1−αL(B160 )

+ C(1 +kqkL(B160 ))1−s1+s

k˜ukL(B+16)+kukL(B016)

1+s2s kuk

1−s 1+s

L(B016)

+ CkqkL12(B016)kukL(B016)

≤C(1 + kqkL(B160 ))1−s1+s

k˜ukL(B+16)+kukL(B160 )α

kuk1−αL(B160 )

+

k˜ukL(B+16)+kukL(B016)1+s2s kuk

1−s 1+s

L(B016)



+ CkqkL12(B016)kukL(B016). Here we recall that

1− α = 1− s

1 + s−1− s

1 + sα < 1− s 1 + s.

 Proof of Proposition2.2. Step 1: Ldecay. We first prove that the L2 bound in the statement of the proposition entails a similar L bound. In the sequel, we denote by ˜c, ˜Cgeneral positive constants which may depend on n, s and which are likely to change from line to line. Pick any R≥ 1 and x0∈ Rn× {0} with |x0| = 32R, (9) implies

kukL2(B16R0 (x0))≤ ˜Ce−˜cRβ. (24)

We next recall that by the Hs(Rn) boundedness of u and the properties of the Caffarelli-Silvestre extension, we have kxn+11−2s2 ∇˜ukL2(Rn+1+ ) ≤ C. Thus, the L2 boundedness of u and Poincar´e’s inequality then also yield a bound for kxn+11−2s2 u˜kL2(Rn×(0,C1)) for any C1 > 0. Combining the bound from Proposition2.4(a), the L2loc(Rn+1+ , x1−2sn+1) boundedness of ˜u, the fact thatkqkL ≤ 1 and (24), we infer that

kxn+11−2s2 u˜kL2(B+cR(x0)) ≤ ˜Ce−˜cRβ.

(10)

R 2R

−R

−2R

xn+1

x0

Figure 1. The chain of balls from the proof of Proposition2.2.

Here c > 0 denotes the constant from Proposition2.4. Finally, invoking a translated and rescaled version of (10) then also entails the bound

(25) k˜ukL(B+

c2 R 16

(x0))≤ ˜Ce−˜cRβ.

This yields a bound for ˜ufor|x| ≥ 32R. We may in particular use this for R = 10. Hence, only on the compact set BC+ with C = 320 an L bound has not yet been obtained. This however follows by applying a rescaled version of (22). Thus, ˜u ∈ L(Rn+1+ ). The estimate in the bounded region and the quantitative estimate in the unbounded annuli can finally be combined to infer that for all R > 0 and x0∈ Rn× {0}, |x0| = 32R we have

k˜ukL(B+

c2 R 16

(x0))≤ ˜Ce−˜cRβ.

In particular, by choosing R = 16cR˜ and keeping x0∈ Rn× {0}, |x0| = 32R, we also obtain the estimate

(26) k˜ukL(B+

c ˜R(x0)) ≤ ˜Ce−¯c ˜Rβ.

Step 2: Conclusion. With the bounds from Step 1, Propositions2.3and2.4at hand, the proof of Proposition2.2follows by a chain of balls argument. More precisely, for x = (x0, xn+1) with

|x| > 2 there exists a value R = 2k for k∈ N such that x ∈ A+R,2R:={x ∈ Rn+1+ : |x| ∈ (R, 2R]}.

This annulus can be covered by a finite union of balls and half balls:

A+R,2R

m1

[

j=1

Br+j(xj)∪

m2

[

k=1

BcR+((x0k,0))

with the property that these balls form a chain, i.e. there is sufficient overlap between these to iterate the following estimates (c.f. Figure1for an illustration of this), where c is the constant derived in Proposition 2.4. We explain this iteration more carefully. Starting with a half ball BcR+(¯x) with ¯x= (x01,0)∈ Rn× {0} we invoke (26) to infer that

k˜ukL(BcR+x))≤ Ce−aRβ, where a = a(n, s) > 0.

We now begin with the propagation of decay estimates into Rn+1+ along a chain of balls.

We first choose a ball Br+1(x1) ⊂ BcR+(¯x) with r1 = cR5 in such a way that a large part of Br+1(x1) ⊂ A+R,2R, B4r+1 ⊂ Rn+1+ and such that |B+2r1(x1)∆BcR+ (¯x)| ≥ c0R for some constant

(11)

c0= c0(n, s) > 0. In B4r+1(x1) we apply the three balls inequality of Proposition2.3. Therefore, in combination with the Lestimates from step 1, we obtain

k˜ukL(B2r1+ (x1))≤ Ck˜ukαL(B+

4r1(x1))k˜uk1−αL(B+r1(x1))

≤ Ck˜uk1−αL(BcR+x))

≤ Ce−a(1−α)Rβ.

We iterate this along our chain of balls B+rj(xj) with rj =cR5 , which allows us to eventually cover A+R,2R with a finite number of balls. In particular, it yields the decay estimate

k˜ukL(A+R,2R)≤ Ce−˜aRβ

for some positive constants C(m1, m2, n, s), ˜a(m1, m2, n, s). Finally, the analogous bounds also follow in arbitrary other dyadic annuli centered at zero by rescaling the previous estimate (the number of balls in the chain of balls estimate stays constant, since the size of the balls is also

rescaled). 

2.3. An interpolation inequality. We need another bulk-boundary interpolation estimate, which will play a relevant role in our Carleman inequalities in the next section. Although this was already introduced in [R¨ul15], we reprove it here for self-containedness.

Proposition 2.5. Let s∈ (0, 1) and u : S+n → R with u ∈ H1(S+n, θ1−2sn ). Then there exists a constant C = C(s, n) > 0 such that for all τ > 1

kukL2(Sn−1)≤ C(τ1−sn1−2s2 ukL2(S+n)+ τ−sn1−2s2SnukL2(S+n)).

Proof. We argue in two steps.

Step 1: Derivation of a whole space estimate. Let w∈ Hs(Rn)∩H1(Rn+1+ , x1−2sn+1). Then trace estimates in the space H1(Rn+1+ , x1−2sn+1) (c.f. for instance Lemma 4.4 in [RS17]) imply

(27) kwkL2(Rn)≤ C(kxn+11−2s2 wkL2(Rn+1+ )+kxn+11−2s2 ∇wkL2(Rn+1+ )).

Starting from (27), scaling x by τ−1 with τ > 0 (i.e., x→ τ−1x), we then obtain kwkL2(Rn)≤ C(τ1−skxn+11−2s2 wkL2(Rn+1+ )+ τ−skxn+11−2s2 ∇wkL2(Rn+1+ )).

(28)

Step 2: Conclusion. Considering u ∈ H1(S+n, θ1−2sn ), we first extend this function zero ho- mogeneously into a neighbourhood of S+n and multiply it by a cut-off function, i.e. we define w(x) := η(x)u(|x|x), where η(x) = 1 if|x| ∈ (1/2, 3/2) and η(x) = 0 if |x| ∈ (0, 1/4) ∪ (2, ∞). The resulting compactly supported function still satisfies w ∈ H1(Rn+1+ , x1−2sn+1) and further has the property that

kxn+11−2s2 ∇wkL2(Rn+1+ )≤ C(kθn1−2s2SnukL2(Sn+)+kθn1−2s2 ukL2(Sn+)), kxn+11−2s2 wkL2(Rn+1+ )≤ Ckθn1−2s2 ukL2(Sn+),

kukL2(Sn−1)≤ CkwkL2(Rn).

Inserting these into (28) and choosing τ ≥ τ0>1 (for some uniform τ0) then implies kukL2(Sn−1)≤ CkwkL2(Rn)≤ C(τ1−skxn+11−2s2 wkL2(Rn+1+ )+ τ−skxn+11−2s2 ∇wkL2(Rn+1))

≤ C(τ1−s(1 + τ−1)kθn1−2s2 ukL2(S+n)+ τ−sn1−2s2 ukL2(Sn+)).

參考文獻

相關文件

Numerical results are reported for some convex second-order cone programs (SOCPs) by solving the unconstrained minimization reformulation of the KKT optimality conditions,

Abstract Based on a class of smoothing approximations to projection function onto second-order cone, an approximate lower order penalty approach for solving second-order cone

Based on a class of smoothing approximations to projection function onto second-order cone, an approximate lower order penalty approach for solving second-order cone

[1] F. Goldfarb, Second-order cone programming, Math. Alzalg, M.Pirhaji, Elliptic cone optimization and prime-dual path-following algorithms, Optimization. Terlaky, Notes on duality

We point out that extending the concepts of r-convex and quasi-convex functions to the setting associated with second-order cone, which be- longs to symmetric cones, is not easy

Especially, the additional gauge symmetry are needed in order to have first order differential equations as equations

Abstract In this paper, we consider the smoothing Newton method for solving a type of absolute value equations associated with second order cone (SOCAVE for short), which.. 1

Abstract We investigate some properties related to the generalized Newton method for the Fischer-Burmeister (FB) function over second-order cones, which allows us to reformulate