This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Landis-type conjecture for the half-Laplacian

Pu-Zhao Kow Department of Mathematics and Statistics, P.O. Box 35 (MaD), FI-40014 University of Jyväskylä, Finland. [email protected]  and  Jenn-Nan Wang Institute of Applied Mathematical Sciences, NCTS, National Taiwan University, Taipei 106, Taiwan. [email protected]
Abstract.

In this paper, we study the Landis-type conjecture, i.e., unique continuation property from infinity, of the fractional Schrödinger equation with drift and potential terms. We show that if any solution of the equation decays at a certain exponential rate, then it must be trivial. The main ingredients of our proof are the Caffarelli-Silvestre extension and Armitage’s Liouville-type theorem.

Key words and phrases:
Unique continuation property, Landis conjecture, half-Laplacian, Caffarelli-Silvestre extension, Liouville-type theorem
2020 Mathematics Subject Classification:
Primary: 35A02, 35B40, 35R11. Secondary: 35J05, 35J15

1. Introduction

In this paper, we consider the following equation with the half Laplacian

(1.1) (Δ)12u+𝐛(𝐱)u+q(𝐱)u=0inn,(-\Delta)^{\frac{1}{2}}u+{\bf b}({\bf x})\cdot\nabla u+q({\bf x})u=0\quad\text{in}\;\;\mathbb{R}^{n},

where n1n\geq 1. Our aim is to investigate the minimal decay rate of nontrivial solutions of (1.1). In other words, we consider the unique continuation property from infinity of (1.1). This problem is closely related to the conjecture proposed by Landis in the 60’s [KL88]. Landis conjectured that, if uu is a solution to the classical Schrödinger equation

(1.2) Δu+q(𝐱)u=0inn,-\Delta u+q({\bf x})u=0\quad\text{in}\;\;\mathbb{R}^{n},

with a bounded potential qq, satisfying the decay estimate

|u(𝐱)|exp(C|𝐱|1+),|u({\bf x})|\leq\exp(-C|{\bf x}|^{1+}),

then u0u\equiv 0. Landis’ conjecture was disproved by Meshkov [Mes92], who constructed a complex-valued potential qL(n)q\in L^{\infty}(\mathbb{R}^{n}) and a nontrivial solution uu of (1.2) such that

|u(𝐱)|exp(C|𝐱|43).|u({\bf x})|\leq\exp(-C|{\bf x}|^{\frac{4}{3}}).

In the same work, Meshkov showed that if

|u(𝐱)|exp(C|𝐱|43+),|u({\bf x})|\leq\exp(-C|{\bf x}|^{\frac{4}{3}+}),

then u0u\equiv 0. Based on a suitable Carleman estimate, a quantitative version of Meshkov’s result was established in [BK05], see also [CS99, Dav14, DZ18, DZ19, KL19, LUW11, LW14] for related results. We also refer to [Zhu18, Theorem 2] for some decay estimates at infinity for higher order elliptic equations.

In view of Meshkov’s example, Kenig modified Landis’ original conjecture and asked that whether the Landis’ conjecture holds true for real-valued potentials qq in [Ken06]. The real version of Landis’ conjecture in the plane was resolved recently in [LMNN20]. We also refer to [Dav20, DKW17, DKW20, KSW15] for the early development of the real version of Landis’ conjecture.

For the fractional Schrödinger equation, the Landis-type conjecture was studied in [RW19]. The main theme of this paper is to extend the results in [RW19] to the fractional Schrödinger equation with the half Laplacian (1.1). Previously, the authors in [KW19] proved some partial results for the fractional Schrödinger equation

(1.3) ((Δ)s+b(𝐱)𝐱+q(𝐱))u=0inn,((-\Delta)^{s}+b({\bf x}){\bf x}\cdot\nabla+q({\bf x}))u=0\quad\text{in}\;\;\mathbb{R}^{n},

where s(0,1)s\in(0,1) and bb, qq are scalar-valued functions. The main tools used in [KW19] are the Caffarelli-Silvestre extension and the Carleman estimate. The particular form of the drift coefficient in (1.3) is due to the applicability of the Carleman estimate. It turns out when s=12s=\frac{1}{2}, i.e., the case of half Laplacian, we can treat a general vector-valued drift coefficient 𝐛(𝐱){\bf b}({\bf x}) in (1.1). The underlying reason is that the Caffarelli-Silvestre extension solution of (Δ)12u=0(-\Delta)^{\frac{1}{2}}u=0 in n{\mathbb{R}}^{n} is a harmonic function in +n+1{\mathbb{R}}_{+}^{n+1}. Inspired by this observation, we show that if both 𝐛{\bf b} and qq are differentiable, then any nontrivial solution of (1.1) can not decay exponentially at infinity. The detailed statement is described in the following theorem.

Theorem 1.1.
Assume that there exists a constant Λ>0\Lambda>0 such that
(1.4a) qL(n)+qL(n)+𝐛L(n)Λ\|q\|_{L^{\infty}(\mathbb{R}^{n})}+\|\nabla q\|_{L^{\infty}(\mathbb{R}^{n})}+\|\nabla{\bf b}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\Lambda
and, furthermore, there exists an ϵ>0\epsilon>0, depending only on nn, such that
(1.4b) 𝐛L(n)ϵ.\|{\bf b}\|_{L^{\infty}(\mathbb{R}^{n})}\leq\epsilon.
Let uW2,p(n)u\in W^{2,p}(\mathbb{R}^{n}) for some integer p>np>n be a solution to (1.1) such that
(1.4c) |u(𝐱)|Λeλ|𝐱||u({\bf x})|\leq\Lambda e^{-\lambda|{\bf x}|}
for some λ>0\lambda>0, then u0u\equiv 0.
Remark 1.2.

Note that both (Δ)12u(-\Delta)^{\frac{1}{2}}u and u\nabla u are first orders. In view of the LpL^{p} estimate of the Riesz transform (2.1), (2.2), the assumption (1.4b) and the regularity requirement of uu are imposed to ensure that the non-local operator (Δ)12u(-\Delta)^{\frac{1}{2}}u is the dominated term in (1.1).

It is interesting to compare Theorem 1.1 with [RW19, Theorem 1]. Assume that uHs(n)u\in H^{s}(\mathbb{R}^{n}) is a solution to

(1.5) (Δ)su+q(𝐱)u=0inn(-\Delta)^{s}u+q({\bf x})u=0\quad\text{in}\,\,\mathbb{R}^{n}

such that |q(𝐱)|1|q({\bf x})|\leq 1 and |𝐱q(𝐱)|1|{\bf x}\cdot\nabla q({\bf x})|\leq 1. If

(1.6) ne|𝐱|α|u|2𝖽𝐱<for someα>1,\int_{\mathbb{R}^{n}}e^{|{\bf x}|^{\alpha}}|u|^{2}\,\mathsf{d}{\bf x}<\infty\quad\text{for some}\,\,\alpha>1,

then u0u\equiv 0. Therefore, for s=12s=\frac{1}{2}, Theorem 1.1 extends their results by slightly relaxing the condition on qq and also adding a drift term. Another key improvement is that the exponential decay rate eλ|𝐱|e^{-\lambda|{\bf x}|} is sharper than (1.6).

The proof of Theorem 1.1 consists of two steps. Inspired by [RW19], we first pass the boundary decay (1.4c) to the bulk decay of the Caffarelli-Silvestre extension solution (harmonic function) in the extended space n×(0,)\mathbb{R}^{n}\times(0,\infty). In the second step, we apply the Liouville-type theorem (Theorem 6.1) to the harmonic function. It is noted that we do not use any Carleman estimate here. On the other hand, using the harmonic function in the unit ball v0(𝐳):=(e1/𝐳α){v}_{0}({\bf z}):=\Re(e^{-1/{\bf z}^{\alpha}}), 𝐳{\bf z}\in{\mathbb{C}}, 0<α<10<\alpha<1 (see [Jin93]), it is not difficult to construct an example to show the optimality of the Liouville-type theorem. In view of this example, we believe that the decay assumption (1.4c) is optimal.

When 𝐛0{\bf b}\equiv 0, the following theorem can be found in [Kow21, Theorem 1.1.9], which was obtained using similar ideas as in the proof of Theorem 1.1.

Theorem 1.3.

Let qL(n)q\in L^{\infty}(\mathbb{R}^{n}) (not necessarily differentiable) satisfy

qL(n)Λ.\|q\|_{L^{\infty}(\mathbb{R}^{n})}\leq\Lambda.

If uH12(n)u\in H^{\frac{1}{2}}(\mathbb{R}^{n}) is a solution to (1.1) with 𝐛0{\bf b}\equiv 0 such that

(1.7) ne|𝐱||u|2𝖽𝐱<,\int_{\mathbb{R}^{n}}e^{|{\bf x}|}|u|^{2}\,\mathsf{d}{\bf x}<\infty,

then u0u\equiv 0.

Remark 1.4.

Theorem 1.3 is an immediate consequence of [RW19, Proposition 2.2] and Theorem 6.1 (without using Proposition 5.1). Therefore we only need (1.7), namely, (1.4c) is unnecessary when 𝐛0{\bf b}\equiv 0.

It is interesting to compare this result with [RW19, Theorem 2]. There, it was proved that if uHs(n)u\in H^{s}(\mathbb{R}^{n}) solves (1.5) with |q(x)|1|q(x)|\leq 1 and

(1.8) ne|𝐱|α|u|2𝖽𝐱<for someα>4s4s1,\int_{\mathbb{R}^{n}}e^{|{\bf x}|^{\alpha}}|u|^{2}\,\mathsf{d}{\bf x}<\infty\quad\text{for some}\,\,\alpha>\frac{4s}{4s-1},

then u0u\equiv 0. When s=12s=\frac{1}{2}, (1.8) becomes

ne|𝐱|α|u|2𝖽𝐱<\int_{\mathbb{R}^{n}}e^{|{\bf x}|^{\alpha}}|u|^{2}\,\mathsf{d}{\bf x}<\infty

for α>2\alpha>2, which is clearly stronger than (1.7). On the other hand, Theorem 1.3 holds regardless whether qq is real-valued or complex-valued.

This paper is organized as follows. In Section 2, we will study the decaying behavior of u\nabla u. In Section 3, we localize the nonlocal operator (Δ)12(-\Delta)^{\frac{1}{2}} by the Caffarelli-Silvestre extension. In Section 4, we derive some useful estimates about the Caffarelli-Silvestre extension u~\tilde{u} of the solution uu, which is harmonic. In Section 5, we obtain the decay rate of u~\tilde{u} from that of uu. Finally, we prove Theorem 1.1 in Section 6 by Armitage’s Liouville-type theorem. Furthermore, we provide another proof of this Liouville-type theorem in Appendix A.

2. Decay of the gradient

Let 1<p<1<p<\infty. For each uLp(n)u\in L^{p}(\mathbb{R}^{n}), let ψ\psi satsify (Δ)12ψ=u(-\Delta)^{\frac{1}{2}}\psi=u and let 𝐮:=ψ{\bf u}:=\nabla\psi. Using the LpL^{p}-boundedness of the Riesz transform [Ste16] (see also [BG13]), we can show that

(2.1) 𝐮Lp(n)C(n,p)uLp(n).\|{\bf u}\|_{L^{p}(\mathbb{R}^{n})}\leq C(n,p)\|u\|_{L^{p}(\mathbb{R}^{n})}.

We remark that this estimate is also used in the proof of [CCW01, Theorem 2.1]. Note that we can formally write 𝐮=(Δ)12u{\bf u}=\nabla(-\Delta)^{-\frac{1}{2}}u. Plugging (Δ)12ψ=u(-\Delta)^{\frac{1}{2}}\psi=u and 𝐮=ψ{\bf u}=\nabla\psi into (2.1) implies

(2.2) ψLp(n)C(n,p)(Δ)12ψLp(n).\|\nabla\psi\|_{L^{p}(\mathbb{R}^{n})}\leq C(n,p)\|(-\Delta)^{\frac{1}{2}}\psi\|_{L^{p}(\mathbb{R}^{n})}.

Thanks to (2.2), we can obtain the following lemma.

Lemma 2.1.

Let 2p<2\leq p<\infty be an integer. Assume that (1.4a) and (1.4b) hold. Let uW2,p(n)u\in W^{2,p}(\mathbb{R}^{n}) be a solution to (1.1) such that the decay assumption (1.4c) holds, then (Δ)12u(-\Delta)^{\frac{1}{2}}u satisfies

(2.3) neλ2|𝐱||(Δ)12u|2𝖽𝐱+neλ2|𝐱||(Δ)12u|p𝖽𝐱C\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|(-\Delta)^{\frac{1}{2}}u|^{2}\,\mathsf{d}{\bf x}+\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|(-\Delta)^{\frac{1}{2}}u|^{p}\,\mathsf{d}{\bf x}\leq C

for some positive constant C=C(n,p,λ,Λ)C=C(n,p,\lambda,\Lambda).

Proof.

We first estimate the LpL^{p}-norm of u\nabla u. Taking LpL^{p}-norm on (1.1) and using (2.2), (1.4b), (1.4c), we have

uLp(n)\displaystyle\|\nabla u\|_{L^{p}(\mathbb{R}^{n})} C(n,p)(Δ)12uLp(n)\displaystyle\leq C(n,p)\|(-\Delta)^{\frac{1}{2}}u\|_{L^{p}(\mathbb{R}^{n})}
C(n,p)(𝐛L(n)uLp(n)+qL(n)uLp(n))\displaystyle\leq C(n,p)\bigg{(}\|{\bf b}\|_{L^{\infty}(\mathbb{R}^{n})}\|\nabla u\|_{L^{p}(\mathbb{R}^{n})}+\|q\|_{L^{\infty}(\mathbb{R}^{n})}\|u\|_{L^{p}(\mathbb{R}^{n})}\bigg{)}
ϵC(n,p)uLp(n)+C(n,p,λ,Λ).\displaystyle\leq\epsilon C(n,p)\|\nabla u\|_{L^{p}(\mathbb{R}^{n})}+C(n,p,\lambda,\Lambda).

Choosing ϵ=(2C(n,p))1\epsilon=(2C(n,p))^{-1} in the estimate above gives

(2.4) uLp(n)C(n,p,λ,Λ).\|\nabla u\|_{L^{p}(\mathbb{R}^{n})}\leq C(n,p,\lambda,\Lambda).

Next, we estimate the LpL^{p}-norm of 2u\nabla^{2}u. Differentiating (1.1) yields

(2.5) (Δ)12ju+𝐛(𝐱)(ju)+j𝐛(𝐱)u+q(𝐱)ju+jq(𝐱)u=0(-\Delta)^{\frac{1}{2}}\partial_{j}u+{\bf b}({\bf x})\cdot\nabla(\partial_{j}u)+\partial_{j}{\bf b}({\bf x})\cdot\nabla u+q({\bf x})\partial_{j}u+\partial_{j}q({\bf x})u=0

for each j=1,,nj=1,\cdots,n. Taking the LpL^{p}-norm of (2.5), we have

(ju)Lp(n)\displaystyle\|\nabla(\partial_{j}u)\|_{L^{p}(\mathbb{R}^{n})} C(n,p)(Δ)12juLp(n)\displaystyle\leq C(n,p)\|(-\Delta)^{\frac{1}{2}}\partial_{j}u\|_{L^{p}(\mathbb{R}^{n})}
C(n,p)(𝐛L(n)(ju)Lp(n)+𝐛L(n)uLp(n)\displaystyle\leq C(n,p)\bigg{(}\|{\bf b}\|_{L^{\infty}(\mathbb{R}^{n})}\|\nabla(\partial_{j}u)\|_{L^{p}(\mathbb{R}^{n})}+\|\nabla{\bf b}\|_{L^{\infty}(\mathbb{R}^{n})}\|\nabla u\|_{L^{p}(\mathbb{R}^{n})}
+qL(n)uLp(n)+qL(n)uLp(n))\displaystyle\quad+\|q\|_{L^{\infty}(\mathbb{R}^{n})}\|\nabla u\|_{L^{p}(\mathbb{R}^{n})}+\|\nabla q\|_{L^{\infty}(\mathbb{R}^{n})}\|u\|_{L^{p}(\mathbb{R}^{n})}\bigg{)}
C(n,p)(ϵ(ju)Lp(n)+ΛuLp(n)+ΛuLp(n))\displaystyle\leq C(n,p)\bigg{(}\epsilon\|\nabla(\partial_{j}u)\|_{L^{p}(\mathbb{R}^{n})}+\Lambda\|\nabla u\|_{L^{p}(\mathbb{R}^{n})}+\Lambda\|u\|_{L^{p}(\mathbb{R}^{n})}\bigg{)}
12(ju)Lp(n)+C(n,p,λ,Λ),\displaystyle\leq\frac{1}{2}\|\nabla(\partial_{j}u)\|_{L^{p}(\mathbb{R}^{n})}+C(n,p,\lambda,\Lambda),

and hence

(2.6) 2uLp(n)C(n,p,λ,Λ).\|\nabla^{2}u\|_{L^{p}(\mathbb{R}^{n})}\leq C(n,p,\lambda,\Lambda).

Hence, it follows from the Sobolev embedding that uL(n)C(n,p,λ,Λ)\|\nabla u\|_{L^{\infty}({\mathbb{R}}^{n})}\leq C(n,p,\lambda,\Lambda).

Now we would like to derive the L2L^{2}-decay of u\nabla u. Combining (2.4) and (2.6), it is easy to see that

neλ2|𝐱||ju|2𝖽𝐱=neλ2|𝐱|(ju)(ju)𝖽𝐱\displaystyle\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|^{2}\,\mathsf{d}{\bf x}=\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}(\partial_{j}u)(\partial_{j}u)\,\mathsf{d}{\bf x}
=λ2nxj|𝐱|eλ2|𝐱|uju𝖽𝐱neλ2|𝐱|(j2u)u𝖽𝐱\displaystyle=-\frac{\lambda}{2}\int_{\mathbb{R}^{n}}\frac{x_{j}}{|{\bf x}|}e^{\frac{\lambda}{2}|{\bf x}|}u\partial_{j}u\,\mathsf{d}{\bf x}-\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}(\partial_{j}^{2}u)u\,\mathsf{d}{\bf x}
λ2neλ2|𝐱||u||ju|𝖽𝐱+neλ2|𝐱||u||j2u|𝖽𝐱\displaystyle\leq\frac{\lambda}{2}\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|u||\partial_{j}u|\,\mathsf{d}{\bf x}+\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|u||\partial_{j}^{2}u|\,\mathsf{d}{\bf x}
λΛ2neλ2|𝐱||ju|𝖽𝐱+Λneλ2|𝐱||j2u|𝖽𝐱(by (1.4c))\displaystyle\leq\frac{\lambda\Lambda}{2}\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|\,\mathsf{d}{\bf x}+\Lambda\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}|{\bf x}|}|\partial_{j}^{2}u|\,\mathsf{d}{\bf x}\quad\text{(by \eqref{eq:decay-assumption})}
λΛ2(neλ2p|𝐱|𝖽𝐱)1pjuLp(n)+Λ(neλ2p|𝐱|𝖽𝐱)1pj2uLp(n)\displaystyle\leq\frac{\lambda\Lambda}{2}\bigg{(}\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}p^{\prime}|{\bf x}|}\,\mathsf{d}{\bf x}\bigg{)}^{\frac{1}{p^{\prime}}}\|\partial_{j}u\|_{L^{p}(\mathbb{R}^{n})}+\Lambda\bigg{(}\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}p^{\prime}|{\bf x}|}\,\mathsf{d}{\bf x}\bigg{)}^{\frac{1}{p^{\prime}}}\|\partial_{j}^{2}u\|_{L^{p}(\mathbb{R}^{n})}
C(n,p,λ,Λ)(where p is the conjugate exponent of p),\displaystyle\leq C(n,p,\lambda,\Lambda)\quad(\mbox{where $p^{\prime}$ is the conjugate exponent of $p$}),

that is, we obtain

(2.7) neλ2|𝐱||u|2𝖽𝐱C(n,p,λ,Λ).\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|\nabla u|^{2}\,\mathsf{d}{\bf x}\leq C(n,p,\lambda,\Lambda).

We now continue to obtain the L2L^{2}-decay of (Δ)12u(-\Delta)^{\frac{1}{2}}u. In view of (1.1) and using (1.4a), (1.4b), (1.4c), (2.7), we have

neλ2|𝐱||(Δ)12u|2𝖽𝐱\displaystyle\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|(-\Delta)^{\frac{1}{2}}u|^{2}\,\mathsf{d}{\bf x}
neλ2|𝐱||𝐛(𝐱)u|2𝖽𝐱+neλ2|𝐱||q(𝐱)u|2𝖽𝐱\displaystyle\leq\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|{\bf b}({\bf x})\cdot\nabla u|^{2}\,\mathsf{d}{\bf x}+\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|q({\bf x})u|^{2}\,\mathsf{d}{\bf x}
𝐛L(n)2neλ2|𝐱||u|2𝖽𝐱+qL(n)2neλ2|𝐱||u|2𝖽𝐱\displaystyle\leq\|{\bf b}\|_{L^{\infty}(\mathbb{R}^{n})}^{2}\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|\nabla u|^{2}\,\mathsf{d}{\bf x}+\|q\|_{L^{\infty}(\mathbb{R}^{n})}^{2}\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|u|^{2}\,\mathsf{d}{\bf x}
(2.8) C(n,λ,Λ).\displaystyle\leq C(n,\lambda,\Lambda).

Here we may choose a smaller ϵ\epsilon if necessary.

Our next task is to derive the LpL^{p}-decay of u\nabla u. First of all, let pp be odd. We then have

neλ2|𝐱||ju|p𝖽𝐱=neλ2|𝐱||ju|(ju)p1𝖽𝐱\displaystyle\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|^{p}\,\mathsf{d}{\bf x}=\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|(\partial_{j}u)^{p-1}\,\mathsf{d}{\bf x}
={ju0}eλ2|𝐱||ju|(ju)p2(ju)𝖽𝐱\displaystyle=\int_{\{\partial_{j}u\neq 0\}}e^{\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|(\partial_{j}u)^{p-2}(\partial_{j}u)\,\mathsf{d}{\bf x}
=λ2nxj|𝐱|eλ2|𝐱||ju|(ju)p2u𝖽𝐱{ju0}eλ2|𝐱|ju|ju|(j2u)(ju)p2u𝖽𝐱\displaystyle=-\frac{\lambda}{2}\int_{\mathbb{R}^{n}}\frac{x_{j}}{|{\bf x}|}e^{\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|(\partial_{j}u)^{p-2}u\,\mathsf{d}{\bf x}-\int_{\{\partial_{j}u\neq 0\}}e^{\frac{\lambda}{2}|{\bf x}|}\frac{\partial_{j}u}{|\partial_{j}u|}(\partial_{j}^{2}u)(\partial_{j}u)^{p-2}u\,\mathsf{d}{\bf x}
(p2)neλ2|𝐱||ju|(ju)p3(j2u)u𝖽𝐱\displaystyle\quad-(p-2)\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|(\partial_{j}u)^{p-3}(\partial_{j}^{2}u)u\,\mathsf{d}{\bf x}
λΛ2neλ2|𝐱||ju|p1𝖽𝐱+Λ(p1)neλ2|𝐱||ju|p2|j2u|𝖽𝐱(by (1.4c))\displaystyle\leq\frac{\lambda\Lambda}{2}\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|^{p-1}\,\mathsf{d}{\bf x}+\Lambda(p-1)\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|^{p-2}|\partial_{j}^{2}u|\,\mathsf{d}{\bf x}\quad\text{(by \eqref{eq:decay-assumption})}
(2.9) C(n,p,λ,Λ)+Λ(p1)neλ2|𝐱||ju|p2|j2u|𝖽𝐱(using (2.4)).\displaystyle\leq C(n,p,\lambda,\Lambda)+\Lambda(p-1)\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|^{p-2}|\partial_{j}^{2}u|\,\mathsf{d}{\bf x}\quad\text{(using \eqref{eq:Lp-gradient-bound})}.

Note that

neλ2|𝐱||ju|p2|j2u|𝖽𝐱\displaystyle\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|^{p-2}|\partial_{j}^{2}u|\,\mathsf{d}{\bf x}
(neλ2r1|𝐱|𝖽𝐱)1r1(n|ju|r2(p2)𝖽𝐱)1r2(n|j2u|r3𝖽𝐱)1r3,\displaystyle\leq\bigg{(}\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}r_{1}|{\bf x}|}\,\mathsf{d}{\bf x}\bigg{)}^{\frac{1}{r_{1}}}\bigg{(}\int_{\mathbb{R}^{n}}|\partial_{j}u|^{r_{2}(p-2)}\,\mathsf{d}{\bf x}\bigg{)}^{\frac{1}{r_{2}}}\bigg{(}\int_{\mathbb{R}^{n}}|\partial_{j}^{2}u|^{r_{3}}\,\mathsf{d}{\bf x}\bigg{)}^{\frac{1}{r_{3}}},

where 1<r1,r2,r3<1<r_{1},r_{2},r_{3}<\infty satisfy

1r1+1r2+1r3=1.\frac{1}{r_{1}}+\frac{1}{r_{2}}+\frac{1}{r_{3}}=1.

Since we consider odd p3p\geq 3, we can choose r1=pr_{1}=p, r2=pp2r_{2}=\frac{p}{p-2} and r3=pr_{3}=p. Hence, we obtain from (2.4) and (2.6) that

(2.10) neλ2|𝐱||ju|p2|j2u|𝖽𝐱C(n,p,λ,Λ).\int_{\mathbb{R}^{n}}e^{-\frac{\lambda}{2}|{\bf x}|}|\partial_{j}u|^{p-2}|\partial_{j}^{2}u|\,\mathsf{d}{\bf x}\leq C(n,p,\lambda,\Lambda).

Combining (2.9) and (2.10) gives

(2.11) neλ2|𝐱||u|p𝖽𝐱C(n,p,λ,Λ).\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|\nabla u|^{p}\,\mathsf{d}{\bf x}\leq C(n,p,\lambda,\Lambda).

When pp is even, estimate (2.11) follows from the same argument above by noting |ju|p=(ju)p|\partial_{j}u|^{p}=(\partial_{j}u)^{p}.

Finally, we estimate the LpL^{p}-decay of (Δ)12u(-\Delta)^{\frac{1}{2}}u. Using the equation (1.1) and by (1.4a), (1.4b), (1.4c), (2.11)), we have

neλ2|𝐱||(Δ)12u|p𝖽𝐱\displaystyle\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|(-\Delta)^{\frac{1}{2}}u|^{p}\,\mathsf{d}{\bf x}
C(neλ2|𝐱||𝐛(𝐱)u|p𝖽𝐱+neλ2|𝐱||q(𝐱)u|p𝖽𝐱)\displaystyle\leq C\bigg{(}\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|{\bf b}({\bf x})\cdot\nabla u|^{p}\,\mathsf{d}{\bf x}+\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|q({\bf x})u|^{p}\,\mathsf{d}{\bf x}\bigg{)}
C(𝐛L(n)pneλ2|𝐱||u|p𝖽𝐱+qL(n)pneλ2|𝐱||u|p𝖽𝐱)\displaystyle\leq C\bigg{(}\|{\bf b}\|_{L^{\infty}(\mathbb{R}^{n})}^{p}\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|\nabla u|^{p}\,\mathsf{d}{\bf x}+\|q\|_{L^{\infty}(\mathbb{R}^{n})}^{p}\int_{\mathbb{R}^{n}}e^{\frac{\lambda}{2}|{\bf x}|}|u|^{p}\,\mathsf{d}{\bf x}\bigg{)}
(2.12) C(n,p,λ,Λ).\displaystyle\leq C(n,p,\lambda,\Lambda).

Consequently, (2.3) is a direct consequence of (2.8) and (2.12). ∎

3. Caffarelli-Silvestre extension

In this section, we briefly discuss the Caffarelli-Silvestre extension [CS07]. We also refer to [GFR19, Appendix A] for higher order fractional Laplacian.

Let +n+1:=n×+={𝐱=(𝐱,xn+1)𝐱n,xn+1>0}\mathbb{R}_{+}^{n+1}:=\mathbb{R}^{n}\times\mathbb{R}_{+}=\begin{Bmatrix}\begin{array}[]{l|l}{\bf x}=({\bf x}^{\prime},x_{n+1})&{\bf x}^{\prime}\in\mathbb{R}^{n},x_{n+1}>0\end{array}\end{Bmatrix} and 𝐱0=(𝐱,0)n×{0}{\bf x}_{0}=({\bf x}^{\prime},0)\in\mathbb{R}^{n}\times\{0\}. For R>0R>0, we denote

BR+(𝐱0)\displaystyle B_{R}^{+}({\bf x}_{0}) :={𝐱+n+1|𝐱𝐱0|R},\displaystyle:=\begin{Bmatrix}\begin{array}[]{l|l}{\bf x}\in\mathbb{R}_{+}^{n+1}&|{\bf x}-{\bf x}_{0}|\leq R\end{array}\end{Bmatrix},
BR(𝐱0)\displaystyle B_{R}^{\prime}({\bf x}_{0}) :={𝐱n×{0}|𝐱𝐱0|R}.\displaystyle:=\begin{Bmatrix}\begin{array}[]{l|l}{\bf x}\in\mathbb{R}^{n}\times\{0\}&|{\bf x}-{\bf x}_{0}|\leq R\end{array}\end{Bmatrix}.

To simplify the notations, we also denote BR+:=BR+(0)B_{R}^{+}:=B_{R}^{+}(0) and BR:=BR(0)B_{R}^{\prime}:=B_{R}^{\prime}(0). We define two Sobolev spaces

H˙1(+n+1)\displaystyle\dot{H}^{1}(\mathbb{R}_{+}^{n+1}) :={v:+n+1+n+1|v|2𝖽𝐱<},\displaystyle:=\begin{Bmatrix}\begin{array}[]{l|l}v:\mathbb{R}_{+}^{n+1}\rightarrow\mathbb{R}&\int_{\mathbb{R}_{+}^{n+1}}|\nabla v|^{2}\,\mathsf{d}{\bf x}<\infty\end{array}\end{Bmatrix},
Hloc1(+n+1)\displaystyle H_{{\rm loc}}^{1}(\mathbb{R}_{+}^{n+1}) :={vH˙1(+n+1)n×(0,r)|v|2𝖽𝐱< for some constant r>0}.\displaystyle:=\begin{Bmatrix}\begin{array}[]{l|l}v\in\dot{H}^{1}(\mathbb{R}_{+}^{n+1})&\int_{\mathbb{R}^{n}\times(0,r)}|v|^{2}\,\mathsf{d}{\bf x}<\infty\text{ for some constant }r>0\end{array}\end{Bmatrix}.

Given any μ\mu\in\mathbb{R} and uHμ(n)u\in H^{\mu}(\mathbb{R}^{n}). Following from [GFR19, Lemma A.1], there exists u~𝒞(+n+1)\tilde{u}\in\mathcal{C}^{\infty}(\mathbb{R}_{+}^{n+1}) such that

(3.1) {Δu~=0in+n+1,limxn+10u~(,xn+1)uHμ(n)=0,\begin{cases}\Delta\tilde{u}=0\quad\text{in}\;\;\,\mathbb{R}_{+}^{n+1},\\ \displaystyle{\lim_{x_{n+1}\rightarrow 0}}\|\tilde{u}(\cdot,x_{n+1})-u\|_{H^{\mu}(\mathbb{R}^{n})}=0,\end{cases}

where =(,n+1)=(1,,n,n+1)\nabla=(\nabla^{\prime},\partial_{n+1})=(\partial_{1},\cdots,\partial_{n},\partial_{n+1}), and the half Laplacian is equivalent to the Dirichlet-to-Neumann map of the extension problem (3.1):

(3.2) limxn+10n+1u~(,xn+1)+(Δ)12uHμ1(n)=0\lim_{x_{n+1}\rightarrow 0}\|\partial_{n+1}\tilde{u}(\cdot,x_{n+1})+(-\Delta)^{\frac{1}{2}}u\|_{H^{\mu-1}(\mathbb{R}^{n})}=0

(see [GFR19, (A.3)]). In particular, when μ=12\mu=\frac{1}{2}, it follows from [GFR19, Corollary A.2] that

u~H˙1(+n+1)CuH12(n)for some positive constant C.\|\tilde{u}\|_{\dot{H}^{1}(\mathbb{R}_{+}^{n+1})}\leq C\|u\|_{H^{\frac{1}{2}}(\mathbb{R}^{n})}\quad\text{for some positive constant }C.

In view of this observation, if uH1(n)u\in H^{1}(\mathbb{R}^{n}) and both 𝐛,q{\bf b},q are bounded, we can reformulate (1.1) as the following local elliptic equation:

(3.3) {Δu~=0in+n+1,u~(𝐱,0)=u(𝐱)onn(inH1(n)-sense),limxn+10n+1u~(𝐱)=𝐛(𝐱)u+q(𝐱)uonn(inL2(n)-sense).\begin{cases}\Delta\tilde{u}=0&\text{in}\;\;\mathbb{R}_{+}^{n+1},\\ \tilde{u}({\bf x}^{\prime},0)=u({\bf x}^{\prime})&\text{on}\;\;\mathbb{R}^{n}\quad(\text{in}\;\;H^{1}(\mathbb{R}^{n})\text{-sense}),\\ {\displaystyle\lim_{x_{n+1}\rightarrow 0}}\partial_{n+1}\tilde{u}({\bf x})={\bf b}({\bf x}^{\prime})\cdot\nabla^{\prime}u+q({\bf x}^{\prime})u&\text{on}\;\;\mathbb{R}^{n}\quad(\text{in}\;\;L^{2}(\mathbb{R}^{n})\text{-sense}).\end{cases}

Since uH1(n)dom((Δ)12)u\in H^{1}(\mathbb{R}^{n})\equiv{\rm dom}\,((-\Delta)^{\frac{1}{2}}), from [Sti10, page 48–49], we have that u~Hloc1(+n+1)\tilde{u}\in H_{{\rm loc}}^{1}(\mathbb{R}_{+}^{n+1}) and

(3.4) u~(,xn+1)L2(n)uL2(n).\|\tilde{u}(\bullet,x_{n+1})\|_{L^{2}(\mathbb{R}^{n})}\leq\|u\|_{L^{2}(\mathbb{R}^{n})}.

4. Some estimates related to the extension problem

The following lemma is a special case of [RW19, Equation (19)] (see also [KW19, Lemma 3.2]).

Lemma 4.1.

Let u~𝒞(+n+1)\tilde{u}\in\mathcal{C}^{\infty}(\mathbb{R}_{+}^{n+1}) be a solution to (3.1). Then the following estimate holds for any 𝐱0n×{0}{\bf x}_{0}\in\mathbb{R}^{n}\times\{0\}:

u~L2(BcR+(𝐱0))\displaystyle\|\tilde{u}\|_{L^{2}(B_{{c}R}^{+}({\bf x}_{0}))}\leq C(u~L2(B16R+(𝐱0))+R12uL2(B16R(𝐱0)))α\displaystyle C\bigg{(}\|\tilde{u}\|_{L^{2}(B_{16R}^{+}({\bf x}_{0}))}+R^{\frac{1}{2}}\|u\|_{L^{2}(B_{16R}^{\prime}({\bf x}_{0}))}\bigg{)}^{\alpha}
(4.1) ×(R32limxn+10n+1u~L2(B16R(𝐱0))+R12uL2(B16R(𝐱0)))1α\displaystyle\times\bigg{(}R^{\frac{3}{2}}\bigg{\|}\lim_{x_{n+1}\rightarrow 0}\partial_{n+1}\tilde{u}\bigg{\|}_{L^{2}(B_{16R}^{\prime}({\bf x}_{0}))}+R^{\frac{1}{2}}\|u\|_{L^{2}(B_{16R}^{\prime}({\bf x}_{0}))}\bigg{)}^{1-\alpha}

for some positive constants C=C(n)C=C(n), α=α(n)(0,1)\alpha=\alpha(n)\in(0,1) and c=c(n)(0,1){c}={c}(n)\in(0,1), all of them are independent of RR and 𝐱0{\bf x}_{0}.

By choosing σ=12\sigma=\frac{1}{2}, ν=2\nu=2 and a(𝐱)0a({\bf x}^{\prime})\equiv 0 in [JLX14, Proposition 2.6(i)], we obtain the following version of De Giorgi-Nash-Moser type theorem.

Lemma 4.2.

Let u~𝒞(+n+1)\tilde{u}\in\mathcal{C}^{\infty}(\mathbb{R}_{+}^{n+1}) be satisfy (3.1) and p>np>n. There exists a constant C=C(n,p)>0C=C(n,p)>0 such that

(4.2) u~L(B14+)C[u~L2(B1+)+limxn+10n+1u~Lp(B1)].\|\tilde{u}\|_{L^{\infty}(B_{\frac{1}{4}}^{+})}\leq C\bigg{[}\|\tilde{u}\|_{L^{2}(B_{1}^{+})}+\bigg{\|}\lim_{x_{n+1}\rightarrow 0}\partial_{n+1}\tilde{u}\bigg{\|}_{L^{p}(B_{1}^{\prime})}\bigg{]}.

Combining (4.1) and (4.2), together with some suitable scaling, we can obtain the following lemma.

Lemma 4.3.

Let u~𝒞(+n+1)\tilde{u}\in\mathcal{C}^{\infty}(\mathbb{R}_{+}^{n+1}) be a solution to (3.1) and p>np>n. Then the following inequality holds for all 𝐱0n×{0}{\bf x}_{0}\in\mathbb{R}^{n}\times\{0\} and R1R\geq 1:

u~L(BcR+(𝐱0))\displaystyle\|\tilde{u}\|_{L^{\infty}(B_{cR}^{+}({\bf x}_{0}))} C(u~L2(B16R+(𝐱0))+R12uL2(B16R(𝐱0)))α\displaystyle\leq C\bigg{(}\|\tilde{u}\|_{L^{2}(B_{16R}^{+}({\bf x}_{0}))}+R^{\frac{1}{2}}\|u\|_{L^{2}(B_{16R}^{\prime}({\bf x}_{0}))}\bigg{)}^{\alpha}
×(R32(Δ)12uL2(B16R(𝐱0))+R12uL2(B16R(𝐱0)))1α\displaystyle\quad\times\bigg{(}R^{\frac{3}{2}}\|(-\Delta)^{\frac{1}{2}}u\|_{L^{2}(B_{16R}^{\prime}({\bf x}_{0}))}+R^{\frac{1}{2}}\|u\|_{L^{2}(B_{16R}^{\prime}({\bf x}_{0}))}\bigg{)}^{1-\alpha}
(4.3) +R32(Δ)12uLp(BR(𝐱0))\displaystyle\quad+R^{\frac{3}{2}}\|(-\Delta)^{\frac{1}{2}}u\|_{L^{p}(B_{R}^{\prime}({\bf x}_{0}))}

for some positive constants C=C(n,p)C=C(n,p), α=α(n)(0,1)\alpha=\alpha(n)\in(0,1) and c=c(n)(0,1){c}={c}(n)\in(0,1), all of them are independent of RR and 𝐱0{\bf x}_{0}.

Proof.

Without loss of generality, it suffices to take 𝐱0=0{\bf x}_{0}=0. Let v~(𝐱)=u~(R𝐱)\tilde{v}({\bf x})=\tilde{u}(R{\bf x}) and let v(𝐱)=u(R𝐱)v({\bf x}^{\prime})=u(R{\bf x}^{\prime}), we observe that

{Δv~=0in+n+1,v~(𝐱,0)=v(𝐱)on𝐱n.\begin{cases}\Delta\tilde{v}=0&\text{in}\;\;\mathbb{R}_{+}^{n+1},\\ \tilde{v}({\bf x}^{\prime},0)=v({\bf x}^{\prime})&\text{on}\;\;{\bf x}^{\prime}\in\mathbb{R}^{n}.\end{cases}

From (4.1) and (4.2), it follows that

v~L(Bc+)\displaystyle\|\tilde{v}\|_{L^{\infty}(B_{c}^{+})} C(v~L2(B16+)+vL2(B16))α(limxn+10n+1v~L2(B16)+vL2(B16))1α\displaystyle\leq C\bigg{(}\|\tilde{v}\|_{L^{2}(B_{16}^{+})}+\|v\|_{L^{2}(B_{16}^{\prime})}\bigg{)}^{\alpha}\bigg{(}\bigg{\|}\lim_{x_{n+1}\rightarrow 0}\partial_{n+1}\tilde{v}\bigg{\|}_{L^{2}(B_{16}^{\prime})}+\|v\|_{L^{2}(B_{16}^{\prime})}\bigg{)}^{1-\alpha}
(4.4) +limxn+10n+1v~Lp(B1).\displaystyle\quad+\bigg{\|}\lim_{x_{n+1}\rightarrow 0}\partial_{n+1}\tilde{v}\bigg{\|}_{L^{p}(B_{1}^{\prime})}.

Note that

v~L(Bc+)\displaystyle\|\tilde{v}\|_{L^{\infty}(B_{c}^{+})} =u~L(BcR+),\displaystyle=\|\tilde{u}\|_{L^{\infty}(B_{cR}^{+})},
v~L2(B16+)\displaystyle\|\tilde{v}\|_{L^{2}(B_{16}^{+})} =Rn+12u~L2(B16R+),\displaystyle=R^{-\frac{n+1}{2}}\|\tilde{u}\|_{L^{2}(B_{16R}^{+})},
vL2(B16)\displaystyle\|v\|_{L^{2}(B_{16}^{\prime})} =Rn2uL2(B16R),\displaystyle=R^{-\frac{n}{2}}\|u\|_{L^{2}(B_{16R}^{\prime})},
limxn+10n+1v~Lp(B1)\displaystyle\bigg{\|}\lim_{x_{n+1}\rightarrow 0}\partial_{n+1}\tilde{v}\bigg{\|}_{L^{p}(B_{1}^{\prime})} =R1nplimxn+10n+1u~Lp(BR),p2.\displaystyle=R^{1-\frac{n}{p}}\bigg{\|}\lim_{x_{n+1}\rightarrow 0}\partial_{n+1}\tilde{u}\bigg{\|}_{L^{p}(B_{R}^{\prime})},\quad p\geq 2.

Hence, (4.4) becomes

u~L(BcR+)\displaystyle\|\tilde{u}\|_{L^{\infty}(B_{cR}^{+})} CRn2Rα2(u~L2(B16R+)+R12uL2(B16R))α\displaystyle\leq CR^{-\frac{n}{2}}R^{-\frac{\alpha}{2}}\bigg{(}\|\tilde{u}\|_{L^{2}(B_{16R}^{+})}+R^{\frac{1}{2}}\|u\|_{L^{2}(B_{16R}^{\prime})}\bigg{)}^{\alpha}
×R12(1α)(R32limxn+10n+1u~L2(B16R)+R12uL2(B16R))1α\displaystyle\quad\times R^{-\frac{1}{2}(1-\alpha)}\bigg{(}R^{\frac{3}{2}}\bigg{\|}\lim_{x_{n+1}\rightarrow 0}\partial_{n+1}\tilde{u}\bigg{\|}_{L^{2}(B_{16R}^{\prime})}+R^{\frac{1}{2}}\|u\|_{L^{2}(B_{16R}^{\prime})}\bigg{)}^{1-\alpha}
+R12npR32limxn+10n+1u~Lp(BR).\displaystyle\quad+R^{-\frac{1}{2}-\frac{n}{p}}R^{\frac{3}{2}}\bigg{\|}\lim_{x_{n+1}\rightarrow 0}\partial_{n+1}\tilde{u}\bigg{\|}_{L^{p}(B_{R}^{\prime})}.

Since R1R\geq 1, (4.3) follows immediately. ∎

5. Boundary decay to bulk decay

In this section, we will establish that the boundary decay implies the bulk decay.

Proposition 5.1.

Assume that (1.4a) and (1.4b) are satisfied. Let uW2,p(n)u\in W^{2,p}(\mathbb{R}^{n}) for some integer n<p<n<p<\infty be a solution to (1.1) and the decay assumption (1.4c) holds. Then

(5.1) |u~(𝐱)|Cec|𝐱|for 𝐱=(𝐱,xn+1)+n+1.|\tilde{u}({\bf x})|\leq Ce^{-c|{\bf x}|}\quad\text{for }\;\;{\bf x}=({\bf x}^{\prime},x_{n+1})\in\mathbb{R}_{+}^{n+1}.
Proof.

Given any R1R\geq 1, choosing 𝐱0n×{0}{\bf x}_{0}\in\mathbb{R}^{n}\times\{0\} with |𝐱0|=32R|{\bf x}_{0}|=32R. By (1.4c), we have

(5.2) uL2(B16R(𝐱0))CecR.\|u\|_{L^{2}(B_{16R}^{\prime}({\bf x}_{0}))}\leq Ce^{-cR}.

Furthermore, (3.4) yields

(5.3) u~L2(B16R+(𝐱0))u~L2(n×(0,16R))4R12uL2(n)C(n,λ,Λ)R12.\|\tilde{u}\|_{L^{2}(B_{16R}^{+}({\bf x}_{0}))}\leq\|\tilde{u}\|_{L^{2}(\mathbb{R}^{n}\times(0,16R))}\leq 4R^{\frac{1}{2}}\|u\|_{L^{2}(\mathbb{R}^{n})}\leq C(n,\lambda,\Lambda)R^{\frac{1}{2}}.

Plugging (2.3), (5.2) and (5.3) into (4.3) implies

u~L(BcR+(𝐱0))CecR.\|\tilde{u}\|_{L^{\infty}(B_{cR}^{+}({\bf x}_{0}))}\leq Ce^{-cR}.

Following the chain of balls argument described in [RW19, Proposition 2.2, Step 2], we finally conclude our result. ∎

6. Proof of Theorem 1.1

Recall the following Liouville-type theorem in [Arm85, Theorem B].

Theorem 6.1.

Suppose that Δu~=0\Delta\tilde{u}=0 in +n+1\mathbb{R}_{+}^{n+1}. If u~\tilde{u} satisfies the decay property (5.1), then u~0\tilde{u}\equiv 0.

It is obvious that Theorem 1.1 is an easy consequence of Proposition 5.1 and Theorem 6.1. We now say a few words about the proof of Theorem 1.3. As Proposition 5.1, the boundary decay (1.7) implies the bulk decay (5.1). In the case of 𝐛0{\bf b}\equiv 0, the proof of Proposition 5.1 remains true when qq is bounded.

To make the paper self-contained, we will give another proof of Theorem 6.1 in Appendix A.

Appendix A Proof of Theorem 6.1

First of all, we introduce a mapping from the ball to the upper half-space, and back, which preserves the Laplacian. For convenience, we define

𝐱:=𝐱|𝐱|2for 𝐱n+1{0},{\bf x}^{*}:=\frac{{\bf x}}{|{\bf x}|^{2}}\quad\text{for }\,\,{\bf x}\in\mathbb{R}^{n+1}\setminus\{0\},

see e.g. [ABR01]. Let 𝐬=(0,,0,1){\bf s}=(0,\cdots,0,-1) be the south pole of the unit sphere 𝒮n\mathcal{S}^{n}, and we define

Φ(𝐳):=2(𝐳𝐬)+𝐬=(2𝐳,1|𝐳|2)|𝐳|2+(1+zn+1)2\Phi({\bf z}):=2({\bf z}-{\bf s})^{*}+{\bf s}=\frac{(2{\bf z}^{\prime},1-|{\bf z}|^{2})}{|{\bf z}^{\prime}|^{2}+(1+z_{n+1})^{2}}

for all 𝐳=(𝐳,zn+1)n+1{𝐬}{\bf z}=({\bf z}^{\prime},z_{n+1})\in\mathbb{R}^{n+1}\setminus\{{\bf s}\}. It is easy to see that Φ2=Id\Phi^{2}={\rm Id}.

Let B1(0)B_{1}(0) be the unit ball in n+1\mathbb{R}^{n+1}. The following lemma can be found in [ABR01].

Lemma A.1.

The mapping Φ:n+1{𝐬}n+1{𝐬}\Phi:\mathbb{R}^{n+1}\setminus\{{\bf s}\}\rightarrow\mathbb{R}^{n+1}\setminus\{{\bf s}\} is injective. Furthermore, it maps B1(0)B_{1}(0) onto +n+1\mathbb{R}_{+}^{n+1}, and maps +n+1\mathbb{R}_{+}^{n+1} onto B1(0)B_{1}(0). It also maps 𝒮n{𝐬}\mathcal{S}^{n}\setminus\{{\bf s}\} onto n{\mathbb{R}}^{n} and maps n{\mathbb{R}}^{n} onto 𝒮n{𝐬}\mathcal{S}^{n}\setminus\{{\bf s}\}.

Given any function ww defined on a domain Ω\Omega in n+1{𝐬}\mathbb{R}^{n+1}\setminus\{{\bf s}\}. The Kelvin transform 𝒦[w]\mathcal{K}[w] of ww is defined by

(A.1) 𝒦[w](𝐳):=2n12|𝐳𝐬|1nw(Φ(𝐳))for all𝐳Φ(Ω).\mathcal{K}[w]({\bf z}):=2^{\frac{n-1}{2}}|{\bf z}-{\bf s}|^{1-n}w(\Phi({\bf z}))\quad\text{for all}\,\,{\bf z}\in\Phi(\Omega).

The following lemma can be found in [ABR01], which exhibits a crucial property of the Kelvin transform.

Lemma A.2.

Let Ω\Omega be any domain in n+1{𝐬}\mathbb{R}^{n+1}\setminus\{{\bf s}\}. Then uu is harmonic on Ω\Omega if and only if 𝒦[u]\mathcal{K}[u] is harmonic on Φ(Ω)\Phi(\Omega).

Now, we are ready to prove Theorem 6.1.

Proof of Theorem 6.1.

To begin, it is not hard to compute

|Φ(𝐳)|\displaystyle|\Phi({\bf z})| =4|𝐳|2+(|𝐳|21)2|𝐳|2+(1+zn+1)2=|2|𝐳|+i((zn+1)2+|𝐳|21)(zn+11)2+|𝐳|2|\displaystyle=\frac{\sqrt{4|{\bf z}^{\prime}|^{2}+(|{\bf z}|^{2}-1)^{2}}}{|{\bf z}^{\prime}|^{2}+(1+z_{n+1})^{2}}=\bigg{|}\frac{2|{\bf z}^{\prime}|+i((-z_{n+1})^{2}+|{\bf z}^{\prime}|^{2}-1)}{(-z_{n+1}-1)^{2}+|{\bf z}^{\prime}|^{2}}\bigg{|}
=|(zn+1+1)+i|𝐳|(zn+11)+i|𝐳||.\displaystyle=\bigg{|}\frac{(-z_{n+1}+1)+i|{\bf z}^{\prime}|}{(-z_{n+1}-1)+i|{\bf z}^{\prime}|}\bigg{|}.

The decay assumption (5.1) implies that for 𝐳{\bf z} near the south pole 𝐬{\bf s},

|𝒦[u~](𝐳)|\displaystyle|\mathcal{K}[\tilde{u}]({\bf z})| =2n12|𝐳𝐬|1n|u~(Φ(𝐳))|C|𝐳𝐬|1nec|Φ(𝐳)|\displaystyle=2^{\frac{n-1}{2}}|{\bf z}-{\bf s}|^{1-n}|\tilde{u}(\Phi({\bf z}))|\leq C|{\bf z}-{\bf s}|^{1-n}e^{-c|\Phi({\bf z})|}
=C|𝐳𝐬|1nexp(c|(zn+1+1)+i|𝐳|(zn+11)+i|𝐳||)\displaystyle=C|{\bf z}-{\bf s}|^{1-n}\exp\bigg{(}-c\bigg{|}\frac{(-z_{n+1}+1)+i|{\bf z}^{\prime}|}{(-z_{n+1}-1)+i|{\bf z}^{\prime}|}\bigg{|}\bigg{)}
C|𝐳𝐬|1nexp(c1|𝐳𝐬|)\displaystyle\leq C|{\bf z}-{\bf s}|^{1-n}\exp\bigg{(}-c\frac{1}{|{\bf z}-{\bf s}|}\bigg{)}
Cexp(c1|𝐳𝐬|).\displaystyle\approx C\exp\bigg{(}-c\frac{1}{|{\bf z}-{\bf s}|}\bigg{)}.

From Lemma A.2, we know that 𝒦[u~]\mathcal{K}[\tilde{u}] is harmonic on B1(0)B_{1}(0). By [Jin93, Theorem 1], we obtain that 𝒦[u~]0\mathcal{K}[\tilde{u}]\equiv 0. In view of (A.1) and Lemma A.1, we then conclude that u~0\tilde{u}\equiv 0 in Φ(B1)=+n+1\Phi(B_{1})=\mathbb{R}_{+}^{n+1}. ∎

Acknowledgments

Kow is partially supported by the Academy of Finland (Centre of Excellence in Inverse Modelling and Imaging, 312121) and by the European Research Council under Horizon 2020 (ERC CoG 770924). Wang is partially supported by MOST 108-2115-M-002-002-MY3 and MOST 109-2115-M-002-001-MY3.

References

  • [Arm85] D. H. Armitage. Uniqueness theorems for harmonic functions in half-spaces. J. Math. Anal. Appl., 106(1):219–236, 1985. MR0780333, doi:10.1016/0022-247X(85)90145-3.
  • [ABR01] S. Axler, P. Bourdon, and W. Ramey. Harmonic function theory, Graduate Texts in Mathematics, volume 137. Springer-Verlag, New York, 2001. MR1184139, doi:10.1007/b97238.
  • [BG13] F. Baudoin and N. Garofalo. A note on the boundedness of riesz transform for some subelliptic operators. Int. Math. Res. Not. IMRN, 2013(2):398–421, 2013. MR3010694, doi:10.1093/imrn/rnr271, arXiv:1105.0467.
  • [BK05] J. Bourgain and C. E. Kenig. On localization in the continuous Anderson-Bernoulli model in higher dimension. Invent. Math., 161(2):389–426, 2005. MR2180453, doi:10.1007/s00222-004-0435-7.
  • [CS07] L. Caffarelli and L. Silvestre. An extension problem related to the fractional laplacian. Comm. Partial Differential Equations, 32(7–9):1245–1260, 2007. MR2354493, doi:10.1080/03605300600987306, arXiv:math/0608640.
  • [CCW01] P. Constantin, D. Cordoba, and J. Wu. On the critical dissipative quasi-geostrophic equation. Indiana Univ. Math. J., 50(special issue):97–107, 2001. Dedicated to Professors Ciprian Foias and Roger Temam, MR1855665, arXiv:math/0103040.
  • [CS99] J. Cruz-Sampedro. Unique continuation at infinity of solutions to Schrödinger equations with complex-valued potentials. Proc. Edinburgh Math. Soc. (2), 42(1):143–153, 1999. MR1669361, doi:10.1017/S0013091500020071.
  • [Dav14] B. Davey. Some quantitative unique continuation results for eigenfunctions of the magnetic Schrödinger operator. Comm. Partial Differential Equations, 39(5):876–945, 2014. MR3196190, doi:10.1080/03605302.2013.796380, arXiv:1209.5822.
  • [Dav20] B. Davey. On Landis’ conjecture in the plane for some equations with sign-changing potentials. Rev. Mat. Iberoam., 36(5):1571–1596, 2020. MR4161296, doi:10.4171/rmi/1176, arXiv:1809.01520.
  • [DKW17] B. Davey, C. Kenig, and J.-N. Wang. The Landis conjecture for variable coefficient second-order elliptic PDEs. Trans. Amer. Math. Soc., 369(11):8209–8237, 2017. MR3695859, doi:10.1090/tran/7073, arXiv:1510.04762.
  • [DKW20] B. Davey, C. Kenig, and J.-N. Wang. On Landis’ conjecture in the plane when the potential has an exponentially decaying negative part (reprint). St. Petersburg Math. J., 31(2):337–353, 2020. MR3937504, doi:10.1090/spmj/1600, arXiv:1808.09420.
  • [DZ18] B. Davey and J. Zhu. Quantitative uniqueness of solutions to second order elliptic equations with singular potentials in two dimensions. Calc. Var. Partial Differential Equations, 57(3):Paper No. 92, 27 pp., 2018. MR3800851, doi:10.1007/s00526-018-1345-7, arXiv:1704.00632.
  • [DZ19] B. Davey and J. Zhu. Quantitative uniqueness of solutions to second-order elliptic equations with singular lower order terms. Comm. Partial Differential Equations, 44(11):1217–1251, 2019. MR3995096, doi:10.1080/03605302.2019.1629957, arXiv:1702.04742.
  • [GFR19] M.-Á García-Ferrero and A. Rüland. Strong unique continuation for the higher order fractional Laplacian. Mathematics in Engineering, 1(4):715–774, 2019. MR4138565, doi:10.3934/mine.2019.4.715, arXiv:1902.09851.
  • [JLX14] T. Jin, Y. Li, and J. Xiong. On a fractional Nirenberg problem, part I: blow up analysis and compactness of solutions. J. Eur. Math. Soc. (JEMS), 16(6):1111–1171, 2014. MR3226738, doi:10.4171/JEMS/456, arXiv:1111.1332.
  • [Jin93] Z. R. Jin. A unique continuation property on the boundary for solutions of elliptic equations. Trans. Amer. Math. Soc., 336(2):639–653, 1993. MR1085944, doi:10.1090/S0002-9947-1993-1085944-3.
  • [Ken06] C. Kenig. Some recent quantitative unique continuation theorems. In Séminaire: Équations aux Dérivées Partielles (2005–2006), Exp. No. XX. École Polytech., Palaiseau, 2006. MR2276085.
  • [KSW15] C. Kenig, L. Silvestre, and J.-N. Wang. On Landis’ conjecture in the plane. Comm. Partial Differential Equations, 40(4):766–789, 2015. MR3299355, doi:10.1080/03605302.2014.978015, arXiv:1404.2496.
  • [KL88] V. A. Kondrat’ev and E. M. Landis. Qualitative theory of second order linear partial differential equations. In Partial differential equations 3 (Russian), Itogi Nauki i Tekhniki. Ser. Sovrem. Probl. Mat. Fund. Napr., volume 32, pages 99–215. Vsesoyuz. Inst. Nauchn. i Tekhn. Inform, Moscow, 1988. MR1133457, Mi intf115.
  • [Kow21] P.-Z. Kow. Unique Continuation Property of the Fractional Elliptic Operators and Applications. PhD thesis, Department of Mathematics, National Taiwan University, 2021. doi:10.6342/NTU202100933.
  • [KL19] P.-Z. Kow and C.-L. Lin. On decay rate of solutions for the stationary Navier-Stokes equation in an exterior domain. J. Differential Equations, 266(6):3279–3309, 2019. MR3912683, doi:10.1016/j.jde.2018.09.002.
  • [KW19] P.-Z. Kow and J.-N. Wang. Unique continuation from the infinity for the fractional laplacian with a drift term. preprint, 2019.
  • [LUW11] C.-L. Lin, G. Uhlmann, and J.-N. Wang. Asymptotic behavior of solutions of the stationary Navier-Stokes equations in an exterior domain. Indiana Univ. Math. J., 60(6):2093–2106, 2011. MR3008262, arXiv:1008.3953.
  • [LW14] C.-L. Lin and J.-N. Wang. Quantitative uniqueness estimates for the general second order elliptic equations. J. Funct. Anal., 266(8):5108–5125, 2014. MR3177332, doi:10.1016/j.jfa.2014.02.016, arXiv:1303.2189.
  • [LMNN20] A. Logunov, E. Malinnikova, N. Nadirashvili, and F. Nazarov. The Landis conjecture on exponential decay. arXiv preprint, 2020. arXiv:2007.07034.
  • [Mes92] V. Z. Meshkov. On the possible rate of decrease at infinity of the solutions of second-order partial differential equations. (English translation). Math. USSR-Sb., 72(2):343–361, 1992. MR1110071, doi:10.1070/SM1992v072n02ABEH001414.
  • [RW19] A. Rüland and J.-N. Wang. On the fractional Landis conjecture. J. Funct. Anal., 277(9):3236–3270, 2019. MR3997635, doi:10.1016/j.jfa.2019.05.026, arXiv:1809.04480.
  • [Ste16] E. M. Stein. Singular integrals and differentiability properties of functions (PMS-30), volume 30. Princeton university press, 2016. MR0290095, doi:10.1515/9781400883882.
  • [Sti10] P. R. Stinga. Fractional powers of second order partial differential operators, extension problem and regularity theory. PhD thesis, Universidad Autónoma de Madrid, Facultad de Ciencias, Departamento de Matemáticas, 2010. https://repositorio.uam.es.
  • [Zhu18] J. Zhu. Quantitative unique continuation of solutions to higher order elliptic equations with singular coefficients. Calc. Var. Partial Differential Equations, 57(2):Paper No. 58, 35 pp., 2018. MR3773807, doi:10.1007/s00526-018-1328-8, arXiv:1704.01446.