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

Qualitative properties of positive solutions
of a mixed order nonlinear Schrödinger equation

Serena Dipierro Department of Mathematics and Statistics, University of Western Australia, 35 Stirling Highway, WA 6009 Crawley, Australia [email protected] Xifeng Su School of Mathematical Sciences, Laboratory of Mathematics and Complex Systems (Ministry of Education)
Beijing Normal University, No. 19, XinJieKouWai St., HaiDian District, Beijing 100875, P. R. China
[email protected], [email protected]
Enrico Valdinoci Department of Mathematics and Statistics, University of Western Australia, 35 Stirling Highway, WA 6009 Crawley, Australia [email protected]  and  Jiwen Zhang School of Mathematical Sciences, Laboratory of Mathematics and Complex Systems (Ministry of Education)
Beijing Normal University, No. 19, XinJieKouWai St., HaiDian District, Beijing 100875, P. R. China
Department of Mathematics and Statistics, University of Western Australia, 35 Stirling Highway, WA 6009 Crawley, Australia [email protected], [email protected]
Abstract.

In this paper, we deal with the following mixed local/nonlocal Schrödinger equation

{Δu+(Δ)su+u=upin n,u>0in n,lim|x|+u(x)=0,\left\{\begin{array}[]{ll}-\Delta u+(-\Delta)^{s}u+u=u^{p}\quad\hbox{in $\mathbb{R}^{n}$,}\\ u>0\quad\hbox{in $\mathbb{R}^{n}$,}\\ \lim\limits_{|x|\to+\infty}u(x)=0,\\ \end{array}\right.

where n2n\geqslant 2, s(0,1)s\in(0,1) and p(1,n+2n2)p\in\left(1,\frac{n+2}{n-2}\right).

The existence of positive solutions for the above problem is proved, relying on some new regularity results. In addition, we study the power-type decay and the radial symmetry properties of such solutions.

The methods make use also of some basic properties of the heat kernel and the Bessel kernel associated with the operator Δ+(Δ)s-\Delta+(-\Delta)^{s}: in this context, we provide self-contained proofs of these results based on Fourier analysis techniques.

Key words and phrases:
Mixed order operators, regularity theory, power-type decay, heat kernel.
2020 Mathematics Subject Classification:
35A08, 35B06, 35B09, 35B40, 35J10.

1. Introduction

In this paper, we are concerned with qualitative properties of solutions to the following mixed local/nonlocal Schrödinger equation

(1.1) Δu+(Δ)su+u=upin n,-\Delta u+(-\Delta)^{s}u+u=u^{p}\quad\hbox{in $\mathbb{R}^{n}$,}

satisfying u>0u>0 in n\mathbb{R}^{n} and

lim|x|+u(x)=0.\lim\limits_{|x|\to+\infty}u(x)=0.

Here above, for any s(0,1)s\in(0,1), the fractional Laplacian is defined as

(Δ)su(x)=cn,sP.V.nu(x)u(y)|xy|n+2s𝑑y,(-\Delta)^{s}u(x)=c_{n,s}\,{\text{P}.V.}\int_{\mathbb{R}^{n}}\frac{u(x)-u(y)}{|x-y|^{n+2s}}\,dy,

where cn,s>0c_{n,s}>0 is a suitable normalization constant, whose explicit value does not play a role here, and P.V.{\text{P}.V.} means that the integral is taken in the Cauchy principal value sense.

Moreover, we suppose that n2n\geqslant 2 and

(1.2) p(1,n+2n2) if n>2andp(1,+) if n=2.p\in\left(1,\frac{n+2}{n-2}\right)\;{\mbox{ if }}n>2\qquad{\mbox{and}}\qquad p\in(1,+\infty)\;{\mbox{ if }}n=2.

We recall that, on the one hand, the Schrödinger equation is the fundamental equation of physics for describing quantum mechanics. Feynman and Hibbs [31] formulated the non-relativistic quantum mechanics as a path integral over the Brownian paths, and this background leads to standard (non-fractional) Schrödinger equation. In this setting, during the last 30 years there have been important contributions to the analysis of the classical nonlinear Schrödinger equation, see e.g. [32, 42, 36, 43, 41, 48].

On the other hand, a nonlocal version of the Schrödinger equation has been introduced by Laskin, see [37, 38, 39]. Laskin constructed the fractional path integral and formulated the space fractional quantum mechanics on the basis of the Lévy flights. In the recent years, the fractional Schrödinger equation has been studied, under different perspectives, by many authors, see in particular [17, 30, 33, 29, 26, 25, 34] and the references therein.

In this framework (see e.g. the appendix in [25]) the classical and the fractional Schrödinger equations arise from a Feynman path integral over “all possible histories of the system”, subject to an action functional obtained as the superposition of a diffusive operator and a potential term: specifically, when the diffusive operator is Gaussian (i.e., as produced by classical Brownian motions), this procedure returns the classical Schrödinger equation, while when the diffusive process is 2s2s-stable for some s(0,1)s\in(0,1) (e.g., as induced by Lévy flights), one obtains the fractional Schrödinger equation corresponding to the fractional Laplacian (Δ)s(-\Delta)^{s}.

In view of this construction, it is also interesting to consider the case in which the diffusive operator in the action functional of the Feynman path integral is of “mixed type”, i.e., rather than possessing a given scaling invariance, it is obtained as the overlapping of two different diffusive operators, say a Gaussian and a 2s2s-stable one (or, similarly, that the diffusive operator acts alternately in a Gaussian fashion and in a 2s2s-stable way for “infinitesimal times” with respect to the time scale considered). This superposition of stochastic processes in the quantum action functional give rise precisely to the equation proposed in (1.1).

Our goal in this paper is to investigate existence and qualitative properties of solutions to (1.1). More precisely, we will obtain results of this type:

  • We prove the existence of nonnegative weak solutions by exploiting Ekeland’s variational principle (see Theorems 1.1 below).

  • We then obtain Hölder continuity and C1,αC^{1,\alpha}-regularity of weak solutions based on the LpL^{p}-theory for the Laplacian and some basic properties of the kernel of the mixed Bessel potential (see Theorems 1.2 and  1.3 below). To overcome the difficulties caused by the fact that the equation involves operators of different orders and therefore is not scale-invariant, we develop a “piecewise” argument presented in Appendices A-B which is new in the literature.

  • Based on the C1,αC^{1,\alpha}-regularity result, we next establish a C2,αC^{2,\alpha}-regularity result by combining a suitable truncation method and a covering argument (see Theorem 1.4 below).

  • Finally, we discuss qualitative properties of classical solutions, such as positivity, power-type decay at infinity, and radial symmetry (see Theorem 1.5 below).

We now state the main results of this paper.

Theorem 1.1 (Existence of nonnegative weak solutions).

There exists a nontrivial and nonnegative weak solution uu of (1.1).

The result above will be established by exploiting the Mountain Pass geometry of the functional associated with (1.1). Additionally, we will show that the energy of the solution in Theorem 1.1 is precisely at the Mountain Pass level, see Corollary 2.6.

We next establish some regularity results which provide useful auxiliary tools in the analysis of the equation under consideration. For this, we notice that the pseudo-differential operator Δ+(Δ)s-\Delta+(-\Delta)^{s} is the infinitesimal generator of a stochastic process XX, where XX is the mixture of the Brownian motion and an independent symmetric 2s2s-stable Lévy process, and, as such, can be characterized using the Fourier Transform {\mathcal{F}} as

(Δu+(Δ)su)(ξ)=(|ξ|2+|ξ|2s)u(ξ).{\mathcal{F}}\Big{(}-\Delta u+(-\Delta)^{s}u\Big{)}(\xi)=\Big{(}|\xi|^{2}+|\xi|^{2s}\Big{)}\,{\mathcal{F}}u(\xi).

Hence, we can define a heat kernel (x,t)\mathcal{H}(x,t) associated with the above operator as

(1.3) (x,t):=net(|ξ|2+|ξ|2s)+2πixξ𝑑ξ\mathcal{H}(x,t):=\int_{\mathbb{R}^{n}}e^{-t(|\xi|^{2}+|\xi|^{2s})+2\pi ix\cdot\xi}\,d\xi

for t>0t>0 and xnx\in\mathbb{R}^{n}.

In this context, the heat kernel (x,t)\mathcal{H}(x,t) may be viewed as a transition density of the Lévy process XX and, relying on probabilistic methods, some properties about the transition density of XX have been established: for example, Song and Vondraček [45] obtained upper and lower bounds on the transition density of XX by comparing the transition densities of the Brownian motion and the 2s2s-stable process.

On a related note, Bogdan, Grzywny, and Ryznar [21] obtained general bounds on the transition density of a pure-jump Lévy process based on Laplace transform arguments (by adapting this method appropriately, one can find a sharp upper bound on the transition density of XX, but no lower bound since the Lévy-Khintchine exponent of the operator Δ+(Δ)s-\Delta+(-\Delta)^{s} is not strictly between 0 and 22).

Furthermore, Cygan, Grzywny and Trojan [24] provided the asymptotic behaviors of the transition density of a Lévy process whose Lévy-Khintchine exponent is strictly and regularly varying between the indexes 0 and 22 (from these methods, since in our framework the index of the stochastic process at zero is 2s2s and at infinity is 22, only the asymptotic formula for the tail transition density of XX can be obtained).

Additionally, the Lévy process XX may be considered as a subordinate Brownian motion, and thus, with the aid of suitable Tauberian Theorems developed by [18], the asymptotic behaviors of the Green function of XX near zero and infinity were established by Rao, Song and Vondraček [44].

In this paper we will revisit these results, framing them in the setting needed for our purposes, and provide a new set of proofs making use of Fourier analysis techniques (see Theorem 3.1). These properties will serve as an essential step to investigate some basic features of the Bessel kernel 𝒦\mathcal{K}, defined as

(1.4) 𝒦(x):=0+et(x,t)𝑑t.\mathcal{K}(x):=\int_{0}^{+\infty}e^{-t}\,\mathcal{H}(x,t)\,dt.

In turn, these properties of 𝒦\mathcal{K} (see Theorem 3.2) will be pivotal to establish a series of regularity results for the mixed operator which can be described as follows.

The first regularity result that we present deals with the Hölder continuity of weak solutions of problem (1.1):

Theorem 1.2 (Hölder continuity).

Let uu be a weak solution of (1.1).

Then, uC0,μ(n)u\in C^{0,\mu}(\mathbb{R}^{n}) for some μ(0,1)\mu\in(0,1) and

uC0,μ(n)C,\|u\|_{C^{0,\mu}(\mathbb{R}^{n})}\leqslant C,

for some C>0C>0, depending only on nn, ss, pp and uH1(n)\|u\|_{H^{1}(\mathbb{R}^{n})}.

We also establish a C1,αC^{1,\alpha}-regularity result:

Theorem 1.3 (C1,αC^{1,\alpha}-regularity).

Let uu be a bounded weak solution of (1.1).

Then, uC1,α(n)u\in C^{1,\alpha}(\mathbb{R}^{n}) for any α(0,1)\alpha\in(0,1) and

uC1,α(n)c(uL(n)+uL(n)p),\|u\|_{C^{1,\alpha}(\mathbb{R}^{n})}\leqslant c\,\left(\|u\|_{L^{\infty}(\mathbb{R}^{n})}+\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}\right),

for some c>0c>0, depending only on nn, ss and pp.

Moreover,

lim|x|+u(x)=0.\lim_{|x|\to+\infty}u(x)=0.

From Theorem 1.3, we will also deduce the C2,αC^{2,\alpha}-regularity result below.

Theorem 1.4 (C2,αC^{2,\alpha}-regularity).

Let uu be a bounded weak solution of (1.1).

Then, uC2,α(n)u\in C^{2,\alpha}(\mathbb{R}^{n}) for any α(0,1)\alpha\in(0,1) and

uC2,α(n)Cmax{uL(n)p2,uL(n)},\|u\|_{C^{2,\alpha}(\mathbb{R}^{n})}\leqslant C\max\left\{\|u\|^{p^{2}}_{L^{\infty}(\mathbb{R}^{n})},\|u\|_{L^{\infty}(\mathbb{R}^{n})}\right\},

for some C>0C>0, depending only on nn, ss, pp and α\alpha.

Finally, we obtain the power-type decay at infinity and the radial symmetry of classical positive solutions to (1.1) by comparison arguments and the method of moving plane, respectively:

Theorem 1.5 (Qualitative properties of positive solutions).

The problem in (1.1) admits a classical, positive, radially symmetric solution uu.

Furthermore, uu has a power-type decay at infinity, that is, there exist constants 0<C1C20<C_{1}\leqslant C_{2} such that, for all |x|1|x|\geqslant 1,

C1|x|n+2su(x)C2|x|n+2s.\frac{C_{1}}{|x|^{n+2s}}\leqslant u(x)\leqslant\frac{C_{2}}{|x|^{n+2s}}.

The rest of this paper is organized as follows. In Section 2, we introduce the functional framework that we work in, and then obtain the existence of nonnegative weak solutions as given by Theorem 1.1.

Section 3 is devoted to the proofs of the regularity results in Theorems 1.2, 1.3 and 1.4.

In Section 4, we establish Theorem 1.5.

Several properties of the heat kernel (1.3) and the Bessel kernel (1.4) are established in Appendices A and B.

2. Existence of weak solutions

The aim of this section is to establish Theorem 1.1. For this, we will first introduce the functional setting and provide some basic definitions. The proof of Theorem 1.1 will then occupy the forthcoming Section 2.2.

2.1. Functional framework

Let n2n\geqslant 2 and s(0,1)s\in(0,1).

As usual, the norm in Lp(n)L^{p}(\mathbb{R}^{n}) will be denoted by uLp(n)\|u\|_{L^{p}(\mathbb{R}^{n})} and the norm in H1(n)H^{1}(\mathbb{R}^{n}) will be denoted by uH1(n)=uL2(n)+uL2(n)\|u\|_{H^{1}(\mathbb{R}^{n})}=\|u\|_{L^{2}(\mathbb{R}^{n})}+\|\nabla u\|_{L^{2}(\mathbb{R}^{n})}.

We also recall the so-called Gagliardo (semi)norm of uu, defined as

[u]Hs(n):=(2n|u(x)u(y)|2|xy|n+2s𝑑x𝑑y)12.[u]_{H^{s}(\mathbb{R}^{n})}:=\left(\iint_{\mathbb{R}^{2n}}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy\right)^{\frac{1}{2}}.

We point out that the Gagliardo (semi)norm is controlled by the H1H^{1}-norm, according to the following observation:

Lemma 2.1.

Let s(0,1)s\in(0,1).

Then, there exists a constant C>0C>0, depending only on nn and ss, such that, for all uH1(n)u\in H^{1}(\mathbb{R}^{n}),

[u]Hs(n)2C(uL2(n)2+uL2(n)2).[u]^{2}_{H^{s}(\mathbb{R}^{n})}\leqslant C\Big{(}\|u\|^{2}_{L^{2}(\mathbb{R}^{n})}+\|\nabla u\|^{2}_{L^{2}(\mathbb{R}^{n})}\Big{)}.
Proof.

Utilizing the Hölder inequality, we see that

[u]Hs(n)2=2n|u(x)u(y)|2|xy|n+2s𝑑x𝑑y=n|z|<1|u(x+z)u(x)|2|z|n+2s𝑑x𝑑z+n|z|1|u(x+z)u(x)|2|z|n+2s𝑑x𝑑zn|z|<1|01u(x+tz)z𝑑t|2|z|n+2s𝑑x𝑑z+2n|z|1|u(x+z)|2+|u(x)|2|z|n+2s𝑑x𝑑zuL2(n)2|z|<1dz|z|n+2s2+4uL2(n)2|z|1dz|z|n+2sC(uL2(n)2+uL2(n)2),\begin{split}[u]^{2}_{H^{s}(\mathbb{R}^{n})}&=\iint_{\mathbb{R}^{2n}}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy\\ &=\int_{\mathbb{R}^{n}}\int_{|z|<1}\frac{|u(x+z)-u(x)|^{2}}{|z|^{n+2s}}\,dx\,dz+\int_{\mathbb{R}^{n}}\int_{|z|\geqslant 1}\frac{|u(x+z)-u(x)|^{2}}{|z|^{n+2s}}\,dx\,dz\\ &\leqslant\int_{\mathbb{R}^{n}}\int_{|z|<1}\frac{|\int_{0}^{1}\nabla u(x+tz)z\,dt|^{2}}{|z|^{n+2s}}\,dx\,dz+2\int_{\mathbb{R}^{n}}\int_{|z|\geqslant 1}\frac{|u(x+z)|^{2}+|u(x)|^{2}}{|z|^{n+2s}}\,dx\,dz\\ &\leqslant\|\nabla u\|^{2}_{L^{2}(\mathbb{R}^{n})}\int_{|z|<1}\frac{dz}{|z|^{n+2s-2}}+4\|u\|^{2}_{L^{2}(\mathbb{R}^{n})}\int_{|z|\geqslant 1}\frac{dz}{|z|^{n+2s}}\\ &\leqslant C\left(\|u\|^{2}_{L^{2}(\mathbb{R}^{n})}+\|\nabla u\|^{2}_{L^{2}(\mathbb{R}^{n})}\right),\end{split}

as desired. ∎

In light of Lemma 2.1, given s(0,1)s\in(0,1), we will equip H1(n)H^{1}(\mathbb{R}^{n}) with the equivalent norm

us:=(uL2(n)2+uL2(n)2+[u]Hs(n)2)12\|u\|_{s}:=\Big{(}\|u\|^{2}_{L^{2}(\mathbb{R}^{n})}+\|\nabla u\|^{2}_{L^{2}(\mathbb{R}^{n})}+[u]^{2}_{H^{s}(\mathbb{R}^{n})}\Big{)}^{\frac{1}{2}}

and the scalar product

u,vs:=n(u(x)v(x)+u(x)v(x))𝑑x+2n(u(x)u(y))(v(x)v(y))|xy|n+2s𝑑x𝑑y.\langle u,v\rangle_{s}:=\int_{\mathbb{R}^{n}}\Big{(}u(x)v(x)+\nabla u(x)\cdot\nabla v(x)\Big{)}\,dx+\iint_{\mathbb{R}^{2n}}\frac{(u(x)-u(y))(v(x)-v(y))}{|x-y|^{n+2s}}\,dx\,dy.
Definition 2.2 (Weak solution).

We say that uH1(n)u\in H^{1}(\mathbb{R}^{n}) is a weak solution of (1.1) if, for all vH1(n)v\in H^{1}(\mathbb{R}^{n}),

n(u(x)v(x)+u(x)v(x))𝑑x+2n(u(x)u(y)(v(x)v(y))|xy|n+2s𝑑x𝑑y=n|u(x)|p1u(x)v(x)𝑑x.\int_{\mathbb{R}^{n}}\Big{(}u(x)v(x)+\nabla u(x)\cdot\nabla v(x)\Big{)}\,dx+\iint_{\mathbb{R}^{2n}}\frac{(u(x)-u(y)(v(x)-v(y))}{|x-y|^{n+2s}}\,dx\,dy=\int_{\mathbb{R}^{n}}|u(x)|^{p-1}u(x)v(x)\,dx.

We consider the functional :H1(n)\mathscr{F}:H^{1}(\mathbb{R}^{n})\to\mathbb{R} defined as

(u):=12us21p+1n|u(x)|p+1𝑑x=12n(u2(x)+|u(x)|2)𝑑x+122n|u(x)u(y)|2|xy|n+2s𝑑x𝑑y1p+1n|u(x)|p+1𝑑x.\begin{split}\mathscr{F}(u):=\;&\frac{1}{2}\|u\|^{2}_{s}-\frac{1}{p+1}\int_{\mathbb{R}^{n}}|u(x)|^{p+1}\,dx\\ =\;&\frac{1}{2}\int_{\mathbb{R}^{n}}\Big{(}u^{2}(x)+|\nabla u(x)|^{2}\Big{)}\,dx+\frac{1}{2}\iint_{\mathbb{R}^{2n}}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy-\frac{1}{p+1}\int_{\mathbb{R}^{n}}|u(x)|^{p+1}\,dx.\end{split}

We point out that, in light of Lemma 2.1 and the classical Sobolev embedding, the functional \mathscr{F} is finite for all uH1(n)u\in H^{1}(\mathbb{R}^{n}).

Furthermore, we have that C1(H1(n))\mathscr{F}\in C^{1}(H^{1}(\mathbb{R}^{n})) and, for all φH1(n)\varphi\in H^{1}(\mathbb{R}^{n}),

(u),φ\displaystyle\left\langle\nabla\mathscr{F}(u),\varphi\right\rangle
=\displaystyle= u,φsn|u(x)|p1u(x)φ(x)𝑑x\displaystyle\langle u,\varphi\rangle_{s}-\int_{\mathbb{R}^{n}}|u(x)|^{p-1}u(x)\varphi(x)\,dx
=\displaystyle= n(u(x)φ(x)+u(x)φ(x))𝑑x+2n(u(x)u(y))(φ(x)φ(y))|xy|n+2s𝑑x𝑑yn|u(x)|p1u(x)φ(x)𝑑x.\displaystyle\int_{\mathbb{R}^{n}}\Big{(}u(x)\varphi(x)+\nabla u(x)\cdot\nabla\varphi(x)\Big{)}\,dx+\iint_{\mathbb{R}^{2n}}\frac{(u(x)-u(y))(\varphi(x)-\varphi(y))}{|x-y|^{n+2s}}\,dx\,dy-\int_{\mathbb{R}^{n}}|u(x)|^{p-1}u(x)\varphi(x)\,dx.

As a consequence, being a weak solution of (1.1) according to Definition 2.2 is equivalent to being a critical point of the functional \mathscr{F}.

Now we introduce an auxiliary problem. For this, we use the notation u+:=max{u,0}u^{+}:=\max\left\{u,0\right\} and u:=max{u,0}u^{-}:=\max\left\{-u,0\right\}, so that u=u+uu=u^{+}-u^{-}, and we consider the following problem

(2.1) {Δu+(Δ)su+u=(u+)pin nu0in n,lim|x|+u(x)=0.\left\{\begin{array}[]{ll}-\Delta u+(-\Delta)^{s}u+u=(u^{+})^{p}\quad\text{in }\mathbb{R}^{n}\\ u\geqslant 0\quad\hbox{in $\mathbb{R}^{n}$,}\\ \lim\limits_{|x|\to+\infty}u(x)=0.\\ \end{array}\right.

The corresponding functional +:H1(n)\mathscr{F}^{+}:H^{1}(\mathbb{R}^{n})\to\mathbb{R} is given by

(2.2) +(u):=12us21p+1n(u+)p+1(x)𝑑x=12n(u2(x)+|u(x)|2)𝑑x+122n|u(x)u(y)|2|xy|n+2s𝑑x𝑑y1p+1n(u+)p+1(x)𝑑x.\begin{split}\mathscr{F}^{+}(u):=\;&\frac{1}{2}\|u\|^{2}_{s}-\frac{1}{p+1}\int_{\mathbb{R}^{n}}(u^{+})^{p+1}(x)\,dx\\ =\;&\frac{1}{2}\int_{\mathbb{R}^{n}}\Big{(}u^{2}(x)+|\nabla u(x)|^{2}\Big{)}\,dx+\frac{1}{2}\iint_{\mathbb{R}^{2n}}\frac{|u(x)-u(y)|^{2}}{|x-y|^{n+2s}}\,dx\,dy-\frac{1}{p+1}\int_{\mathbb{R}^{n}}(u^{+})^{p+1}(x)\,dx.\end{split}

We point out that +C1(H1(n))\mathscr{F}^{+}\in C^{1}(H^{1}(\mathbb{R}^{n})) and, for all φH1(n)\varphi\in H^{1}(\mathbb{R}^{n}),

+(u),φ=u,φsn(u+)p(x)φ(x)𝑑x=n(u(x)φ(x)+u(x)φ(x))𝑑x+2n(u(x)u(y))(φ(x)φ(y))|xy|n+2s𝑑x𝑑yn(u+)p(x)φ(x)𝑑x.\begin{split}&\left\langle\nabla\mathscr{F}^{+}(u),\varphi\right\rangle\\ =\;&\langle u,\varphi\rangle_{s}-\int_{\mathbb{R}^{n}}(u^{+})^{p}(x)\varphi(x)\,dx\\ =\;&\int_{\mathbb{R}^{n}}\Big{(}u(x)\varphi(x)+\nabla u(x)\cdot\nabla\varphi(x)\Big{)}\,dx+\iint_{\mathbb{R}^{2n}}\frac{(u(x)-u(y))(\varphi(x)-\varphi(y))}{|x-y|^{n+2s}}\,dx\,dy-\int_{\mathbb{R}^{n}}(u^{+})^{p}(x)\varphi(x)\,dx.\end{split}

Accordingly, being a weak solution of (2.1) is equivalent to being a critical point of +(u)\mathscr{F}^{+}(u).

The role of this auxiliary functional is made explicit by the following observation:

Lemma 2.3.

Let uu be a critical point for +\mathscr{F}^{+}.

Then,

(2.3) u0u\geqslant 0 a.e. in n\mathbb{R}^{n}.

Moreover,

(2.4) uu is a critical point for \mathscr{F}.
Proof.

We observe that uH1(n)u^{-}\in H^{1}(\mathbb{R}^{n}) and therefore

0\displaystyle 0 =\displaystyle= +(u),u\displaystyle\left\langle\nabla\mathscr{F}^{+}(u),u^{-}\right\rangle
=\displaystyle= n(u(x)u(x)+u(x)u(x))𝑑x+2n(u(x)u(y))(u(x)u(y))|xy|n+2s𝑑x𝑑y\displaystyle\int_{\mathbb{R}^{n}}\Big{(}u(x)u^{-}(x)+\nabla u(x)\cdot\nabla u^{-}(x)\Big{)}\,dx+\iint_{\mathbb{R}^{2n}}\frac{(u(x)-u(y))(u^{-}(x)-u^{-}(y))}{|x-y|^{n+2s}}\,dx\,dy
n(u+)p(x)u(x)𝑑x\displaystyle\qquad-\int_{\mathbb{R}^{n}}(u^{+})^{p}(x)u^{-}(x)\,dx
=\displaystyle= n((u)2(x)+|u(x)|2)𝑑x2n(u(x)u(y))2|xy|n+2s𝑑x𝑑y\displaystyle-\int_{\mathbb{R}^{n}}\Big{(}(u^{-})^{2}(x)+|\nabla u^{-}(x)|^{2}\Big{)}\,dx-\iint_{\mathbb{R}^{2n}}\frac{(u^{-}(x)-u^{-}(y))^{2}}{|x-y|^{n+2s}}\,dx\,dy
2nu+(x)u(y)+u(x)u+(y)|xy|n+2s𝑑x𝑑y\displaystyle\qquad-\iint_{\mathbb{R}^{2n}}\frac{u^{+}(x)u^{-}(y)+u^{-}(x)u^{+}(y)}{|x-y|^{n+2s}}\,dx\,dy
=\displaystyle= us22nu+(x)u(y)+u+(y)u(x)|xy|n+2s𝑑x𝑑yus2.\displaystyle-\|u^{-}\|^{2}_{s}-\iint_{\mathbb{R}^{2n}}\frac{u^{+}(x)u^{-}(y)+u^{+}(y)u^{-}(x)}{|x-y|^{n+2s}}\,dx\,dy\leqslant-\|u^{-}\|_{s}^{2}.

This implies that u=0u^{-}=0 a.e. in n\mathbb{R}^{n}, thus establishing (2.3).

As a consequence of (2.3), recalling also the definitions of the functionals \mathscr{F} and +\mathscr{F}^{+}, we obtain (2.4). ∎

In view of Lemma 2.3, we will now focus on critical points for the functional +.\mathscr{F}^{+}. Let

Γ:={γC([0,1],H1(n)) s.t. γ(0)=0 and +(γ(1))<0}\Gamma:=\left\{\gamma\in C([0,1],H^{1}(\mathbb{R}^{n}))\;{\mbox{ s.t. }}\;\gamma(0)=0\;{\mbox{ and }}\;\mathscr{F}^{+}(\gamma(1))<0\right\}

and define

(2.5) c:=infγΓsupt[0,1]+(γ(t)).c:=\inf\limits_{\gamma\in\Gamma}\sup\limits_{t\in[0,1]}\mathscr{F}^{+}(\gamma(t)).

We make the following observation:

Lemma 2.4.

We have that Γ\Gamma\not=\varnothing and

(2.6) c>0.c>0.
Proof.

Since p>1p>1, we have that +\mathscr{F^{+}} is unbounded from below and therefore Γ\Gamma\not=\varnothing.

Moreover, given uH1(n)u\in H^{1}(\mathbb{R}^{n}) with u0u\not\equiv 0, by the Sobolev inequality one has that

+(u)=12us21p+1uLp+1(n)p+1us2(12c~p+1usp1),\mathscr{F^{+}}(u)=\frac{1}{2}\|u\|^{2}_{s}-\frac{1}{p+1}\|u\|_{L^{p+1}({\mathbb{R}}^{n})}^{p+1}\geqslant\|u\|^{2}_{s}\left(\frac{1}{2}-\frac{\widetilde{c}}{p+1}\|u\|_{s}^{p-1}\right),

where c~>0\widetilde{c}>0 depends on nn and pp.

Hence, if us\|u\|_{s} is sufficiently small, say us(0,(p+14c~)1p1)\|u\|_{s}\in\left(0,\left(\frac{p+1}{4\widetilde{c}}\right)^{\frac{1}{p-1}}\right), we deduce that

+(u)14us2>0,\mathscr{F^{+}}(u)\geqslant\frac{1}{4}\|u\|_{s}^{2}>0,

which gives (2.6). ∎

We remark that 0 is a local minimum of +\mathscr{F}^{+} rather than a global minimum. Thus the functional +\mathscr{F}^{+} fulfills the geometric hypothesis of the Mountain Pass Theorem, but the Palais-Smale condition is not necessarily satisfied. To overcome this issue, we will make use of the following result, established in [23], which allows us to prove that +\mathscr{F}^{+} (and therefore \mathscr{F}) has a positive critical value, thus entailing Theorem 1.1.

Lemma 2.5.

([23, Lemma 2.18]) Let n2n\geqslant 2. Assume that {uk}\left\{u_{k}\right\} is bounded in H1(n)H^{1}(\mathbb{R}^{n}) and that there exists R>0R>0 such that

lim infk+supynBR(y)|uk(x)|2𝑑x=0.\liminf_{k\to+\infty}\sup\limits_{y\in\mathbb{R}^{n}}\int_{B_{R}(y)}|u_{k}(x)|^{2}\,dx=0.

Then, uk0u_{k}\to 0 in Lq(n)L^{q}(\mathbb{R}^{n}) for all q(2,2)q\in(2,2^{*}).

2.2. Existence of weak solutions

With the preliminary work done so far, we can now complete the proof of Theorem 1.1.

Proof of Theorem 1.1.

We recall the definition of the auxiliary functional +\mathscr{F}^{+} in (2.2). In light of Lemma 2.3, in order to find a nonnegative weak solution of (1.1), we focus on finding a nontrivial critical point for the auxiliary functional +\mathscr{F}^{+}.

We claim that

(2.7) the functional +\mathscr{F}^{+} has a nontrivial critical point.

To prove this, we use the Ekeland’s variational principle (see e.g. [40]) and obtain that there exists a sequence uku_{k} such that, as k+k\to+\infty,

(2.8) +(uk)cand+(uk)0,\mathscr{F}^{+}(u_{k})\to c\qquad\text{and}\qquad\nabla\mathscr{F^{+}}(u_{k})\to 0,

where cc is given in (2.5).

We notice that, from (2.8), for kk large enough,

(2.9) c+1+(uk)1p+1+(uk),uk=(121p+1)uks2.c+1\geqslant\mathscr{F}^{+}(u_{k})-\frac{1}{p+1}\left\langle\nabla\mathscr{F}^{+}(u_{k}),u_{k}\right\rangle=\left(\frac{1}{2}-\frac{1}{p+1}\right)\|u_{k}\|^{2}_{s}.

It follows that {uk}\left\{u_{k}\right\} is bounded in H1(n)H^{1}(\mathbb{R}^{n}), and therefore, up to a subsequence, {uk}\left\{u_{k}\right\} converges to some uH1(n)u\in H^{1}(\mathbb{R}^{n}) weakly in H1(n)H^{1}(\mathbb{R}^{n}) and strongly in Llocq(n)L^{q}_{loc}(\mathbb{R}^{n}) for all q(1,2)q\in(1,2^{*}).

As a result, for any φC0(n)\varphi\in C^{\infty}_{0}(\mathbb{R}^{n}),

(2.10) 0=limk++(uk),φ=limk+(uk,φsn(uk+)pφ𝑑x)=u,φsn(u+)pφ𝑑x=+(u),φ,\begin{split}&0=\lim_{k\to+\infty}\left\langle\nabla\mathscr{F}^{+}(u_{k}),\varphi\right\rangle=\lim_{k\to+\infty}\left(\left\langle u_{k},\varphi\right\rangle_{s}-\int_{\mathbb{R}^{n}}(u_{k}^{+})^{p}\varphi\,dx\right)\\ &\qquad=\left\langle u,\varphi\right\rangle_{s}-\int_{\mathbb{R}^{n}}(u^{+})^{p}\varphi\,dx=\left\langle\nabla\mathscr{F}^{+}(u),\varphi\right\rangle,\end{split}

which entails that +(u)=0\nabla\mathscr{F}^{+}(u)=0.

Now, for every R>0R>0, we set

α:=lim infk+supynBR(y)|uk(x)|2𝑑x\alpha:=\liminf_{k\to+\infty}\sup_{y\in\mathbb{R}^{n}}\int_{B_{R}(y)}|u_{k}(x)|^{2}\,dx

and we claim that

(2.11) α(0,+).\alpha\in(0,+\infty).

Indeed, clearly α[0,+]\alpha\in[0,+\infty]. Moreover, since uku_{k} converges strongly in Lloc2(n)L^{2}_{loc}(\mathbb{R}^{n}) to some uH1(n)u\in H^{1}(\mathbb{R}^{n}), we have that α[0,+)\alpha\in[0,+\infty).

Hence, it remains to show that α>0\alpha>0. For this, we argue by contradiction and suppose that α=0\alpha=0. Then, we deduce from Lemma 2.5 that uk0u_{k}\to 0~{} in Lp+1(n)L^{p+1}(\mathbb{R}^{n}). As a consequence of this, and recalling (2.8), we have that

c=limk+(+(uk)12+(uk),uk)=(121p+1)limk+n(uk+)p+1(x)𝑑x=(121p+1)limk+uk+Lp+1(n)p+1=0,\begin{split}&c=\lim_{k\to+\infty}\left(\mathscr{F}^{+}(u_{k})-\frac{1}{2}\left\langle\nabla\mathscr{F}^{+}(u_{k}),u_{k}\right\rangle\right)=\left(\frac{1}{2}-\frac{1}{p+1}\right)\lim_{k\to+\infty}\int_{\mathbb{R}^{n}}(u_{k}^{+})^{p+1}(x)\,dx\\ &\qquad=\left(\frac{1}{2}-\frac{1}{p+1}\right)\lim_{k\to+\infty}\|u_{k}^{+}\|^{p+1}_{L^{p+1}(\mathbb{R}^{n})}=0,\end{split}

in contradiction with (2.6). This establishes (2.11).

It now follows from (2.11) that we have that there exists a sequence ykny_{k}\in\mathbb{R}^{n} such that, for all kk\in\mathbb{N} (and possibly taking a subsequence of uku_{k}),

(2.12) BR(yk)|uk(x)|2𝑑x>α2.\int_{B_{R}(y_{k})}|u_{k}(x)|^{2}\,dx>\frac{\alpha}{2}.

In light of (2.12), we define

(2.13) u¯k(x):=uk(x+yk)\bar{u}_{k}(x):=u_{k}(x+y_{k})

and we see that u¯ks=uks\|\bar{u}_{k}\|_{s}=\|{u}_{k}\|_{s}. Therefore,

+(u¯k)=12uks21p+1n(uk+)p+1(x)𝑑x=+(uk)\mathscr{F^{+}}(\bar{u}_{k})=\frac{1}{2}\|u_{k}\|_{s}^{2}-\frac{1}{p+1}\int_{\mathbb{R}^{n}}(u_{k}^{+})^{p+1}(x)\,dx=\mathscr{F^{+}}({u}_{k})

and, for all φH1(n)\varphi\in H^{1}(\mathbb{R}^{n}),

+(u¯k),φ=u¯k,φsn(u¯k+)p(x)φ(x)𝑑x\displaystyle\left\langle\nabla\mathscr{F}^{+}(\bar{u}_{k}),\varphi\right\rangle=\left\langle\bar{u}_{k},\varphi\right\rangle_{s}-\int_{\mathbb{R}^{n}}(\bar{u}_{k}^{+})^{p}(x)\varphi(x)\,dx
=uk,φ(yk)sn(uk+)p(x)φ(xyk)dx=+(uk),φ(yk),\displaystyle\qquad=\left\langle{u}_{k},\varphi(\cdot-y_{k})\right\rangle_{s}-\int_{\mathbb{R}^{n}}({u}_{k}^{+})^{p}(x)\varphi(x-y_{k})\,dx=\left\langle\nabla\mathscr{F}^{+}({u}_{k}),\varphi(\cdot-y_{k})\right\rangle,

and thus, in view of (2.8), we have that, as k+k\to+\infty,

(2.14) +(u¯k)cand+(u¯k)0.\mathscr{F}^{+}(\bar{u}_{k})\to c\qquad{\mbox{and}}\qquad\nabla\mathscr{F^{+}}(\bar{u}_{k})\to 0.

Now, using (2.9) with uku_{k} replaced by u¯k\bar{u}_{k}, we conclude that there exists u¯H1(n)\bar{u}\in H^{1}(\mathbb{R}^{n}) such that

(2.15) u¯k\bar{u}_{k} converges to u¯\bar{u} weakly in H1(n)H^{1}(\mathbb{R}^{n}) and strongly in Llocq(n)L^{q}_{loc}(\mathbb{R}^{n}) for all q(1,2)q\in(1,2^{*}).

Moreover, by (2.10) with uku_{k} replaced by u¯k\bar{u}_{k}, we deduce that +(u¯)=0\nabla\mathscr{F}^{+}(\bar{u})=0, and hence u¯\bar{u} is a critical point for +\mathscr{F}^{+}. Accordingly, to complete the proof of the claim in (2.7), it only remains to check that

(2.16) u¯0.\bar{u}\not\equiv 0.

To this end, we recall (2.12) and change variable z:=xykz:=x-y_{k} to get that

α2<BR(yk)|uk(x)|2𝑑x=BR(yk)|u¯k(xyk)|2𝑑x=BR|u¯k(z)|2𝑑z\displaystyle\frac{\alpha}{2}<\int_{B_{R}(y_{k})}|u_{k}(x)|^{2}\,dx=\int_{B_{R}(y_{k})}|\bar{u}_{k}(x-y_{k})|^{2}\,dx=\int_{B_{R}}|\bar{u}_{k}(z)|^{2}\,dz

Since u¯k\bar{u}_{k} converges to u¯\bar{u} strongly in Lloc2(n)L^{2}_{loc}(\mathbb{R}^{n}), this implies that

α2BR|u¯(z)|2𝑑z,\frac{\alpha}{2}\leqslant\int_{B_{R}}|\bar{u}(z)|^{2}\,dz,

which gives (2.16), thus completing the proof of (2.7). ∎

In particular, we can show that the nontrivial critical point for +\mathscr{F}^{+} found in the proof of Theorem 1.1 is at the critical level cc given by (2.5):

Corollary 2.6.

Let u¯\bar{u} be the nontrivial critical point for +\mathscr{F}^{+}, as given by (2.7), and let cc be as in (2.5).

Then,

(2.17) +(u¯)=c.\mathscr{F}^{+}(\bar{u})=c.
Proof.

We first check that

(2.18) +(u¯)c.\mathscr{F^{+}}(\bar{u})\geqslant c.

Indeed, we observe that

0=+(u¯),u¯=u¯s2u¯+Lp+1(n)p+1,0=\left\langle\nabla\mathscr{F}^{+}(\bar{u}),\bar{u}\right\rangle=\|\bar{u}\|^{2}_{s}-\|\bar{u}^{+}\|^{p+1}_{L^{p+1}(\mathbb{R}^{n})},

and therefore, in light of (2.16), we deduce that u¯+0\bar{u}^{+}\not\equiv 0.

Moreover, for every t>0t>0,

+(tu¯)=t22u¯s2tp+1p+1n(u¯+)p+1(x)𝑑x,\displaystyle\mathscr{F^{+}}(t\bar{u})=\frac{t^{2}}{2}\|\bar{u}\|^{2}_{s}-\frac{t^{p+1}}{p+1}\int_{\mathbb{R}^{n}}(\bar{u}^{+})^{p+1}(x)\,dx,

and thus, since p+1>2p+1>2, we have that +(tu¯)\mathscr{F^{+}}(t\bar{u})\to-\infty as t+t\to+\infty. Hence, there exists some T>0T>0 such that +(Tu¯)<0\mathscr{F^{+}}(T\bar{u})<0.

As a consequence, we consider the path γu¯(t):=tTu¯\gamma_{\bar{u}}(t):=tT\bar{u} for all t[0,1]t\in[0,1] and notice that γu¯Γ\gamma_{\bar{u}}\in\Gamma. Thus,

(2.19) csupt[0,1]+(γu¯(t))=supt[0,1]+(tTu¯).c\leqslant\sup_{t\in[0,1]}\mathscr{F^{+}}(\gamma_{\bar{u}}(t))=\sup_{t\in[0,1]}\mathscr{F^{+}}(tT\bar{u}).

In addition, let us define hu(t):=+(tu¯)h_{u}(t):=\mathscr{F^{+}}(t\bar{u}) for all t>0t>0. In this way, we compute

hu¯(t)=+(tu¯),u¯=tu¯s2tpu¯+Lp+1(n)p+1h_{\bar{u}}^{\prime}(t)=\left\langle\nabla\mathscr{F^{+}}(t\bar{u}),\bar{u}\right\rangle=t\|\bar{u}\|^{2}_{s}-t^{p}\|\bar{u}^{+}\|^{p+1}_{L^{p+1}(\mathbb{R}^{n})}

and we see that hu¯h_{\bar{u}} has a unique maximum tu¯>0t_{\bar{u}}>0.

Thus, we have that +(tu¯u¯),tu¯u¯=0\left\langle\nabla\mathscr{F^{+}}(t_{\bar{u}}\bar{u}),t_{\bar{u}}\bar{u}\right\rangle=0, which implies that tu¯=1t_{\bar{u}}=1. Accordingly, we have that +(u¯)+(tu¯)\mathscr{F^{+}}(\bar{u})\geqslant\mathscr{F^{+}}(t\bar{u}) for all t>0t>0. By combining this information with (2.19), we obtain (2.18).

We now check that

(2.20) +(u¯)c.\mathscr{F^{+}}(\bar{u})\leqslant c.

For this, we recall the definition of the sequence u¯k\bar{u}_{k} in (2.13) and we observe that, for every r>0r>0,

+(u¯k)12+(u¯k),u¯k=(121p+1)n(u¯k+)p+1(x)𝑑x(121p+1)Br(u¯k+)p+1(x)𝑑x.\mathscr{F}^{+}(\bar{u}_{k})-\frac{1}{2}\left\langle\nabla\mathscr{F}^{+}(\bar{u}_{k}),\bar{u}_{k}\right\rangle=\left(\frac{1}{2}-\frac{1}{p+1}\right)\int_{\mathbb{R}^{n}}(\bar{u}_{k}^{+})^{p+1}(x)\,dx\geqslant\left(\frac{1}{2}-\frac{1}{p+1}\right)\int_{B_{r}}(\bar{u}_{k}^{+})^{p+1}(x)\,dx.

Taking the limit as k+k\to+\infty and recalling (2.14) and (2.15), we get that

c(121p+1)Br(u¯+)p+1(x)𝑑x.c\geqslant\left(\frac{1}{2}-\frac{1}{p+1}\right)\int_{B_{r}}(\bar{u}^{+})^{p+1}(x)\,dx.

Since this holds true for all r>0r>0, we conclude that

c(121p+1)n(u¯+)p+1(x)𝑑x=+(u¯)12+(u¯),u¯=+(u¯),c\geqslant\left(\frac{1}{2}-\frac{1}{p+1}\right)\int_{\mathbb{R}^{n}}(\bar{u}^{+})^{p+1}(x)\,dx=\mathscr{F}^{+}(\bar{u})-\frac{1}{2}\left\langle\nabla\mathscr{F}^{+}(\bar{u}),\bar{u}\right\rangle=\mathscr{F}^{+}(\bar{u}),

which gives (2.20).

From (2.18) and (2.20), we obtain (2.17), as desired. ∎

3. Regularity of weak solutions

In this section we focus on the Hölder continuity, C1,αC^{1,\alpha} and C2,αC^{2,\alpha}-regularity results for weak solutions of (1.1) amd prove Theorems 1.2, 1.3 and 1.4. These regularity estimates will be the basis for the qualitative analysis we carry out in the next section.

To this end, we start with some basic properties of the heat kernel (1.3) and the Bessel kernel (1.4) which are essential for proving Theorems 1.2 and 1.3, as summarized by the following two results.

Theorem 3.1 (Properties of the heat kernel).

Let n1n\geqslant 1 and s(0,1)s\in(0,1). Let \mathcal{H} be as defined in (1.3).

Then,

  • \mathcal{H} is nonnegative, radially symmetric and nonincreasing with respect to r=|x|r=|x|.

  • There exist positive constants C1C_{1} and C2C_{2} such that

    (x,t)C1{t|x|n+2sts|x|n+2s}{tn2stn2}\displaystyle\mathcal{H}(x,t)\leqslant C_{1}\left\{\frac{t}{|x|^{n+2s}}\vee\frac{t^{s}}{|x|^{n+2s}}\right\}\wedge\left\{t^{-\frac{n}{2s}}\wedge t^{-\frac{n}{2}}\right\}
    and (x,t)C2{t|x|n+2sif  1<t<|x|2s,eπ|x|2ttn2if |x|2<t<|x|2s<1,\displaystyle\mathcal{H}(x,t)\geqslant C_{2}\begin{cases}\frac{t}{|x|^{n+2s}}\qquad&\mbox{if }\;1<t<|x|^{2s},\\ e^{\frac{\pi|x|^{2}}{t}}t^{-\frac{n}{2}}\qquad&\mbox{if }\;|x|^{2}<t<|x|^{2s}<1,\end{cases}

    where ab:=min{a,b}a\wedge b:=\min\left\{a,b\right\} and ab:=max{a,b}a\vee b:=\max\left\{a,b\right\}.

Theorem 3.2 (Properties of the Bessel kernel).

Let n1n\geqslant 1 and s(0,1)s\in(0,1). Let 𝒦\mathcal{K} be as defined in (1.4).

Then,

  • (a)

    𝒦\mathcal{K} is positive, radially symmetric, smooth in n{0}\mathbb{R}^{n}\setminus\left\{0\right\} and nonincreasing with respect to r=|x|r=|x|.

  • (b)

    There exist positive constants C3C_{3} and C4C_{4} such that, if |x|1|x|\geqslant 1,

    C3|x|n+2s𝒦(x)1C3|x|n+2s\frac{C_{3}}{|x|^{n+2s}}\leqslant\mathcal{K}(x)\leqslant\frac{1}{C_{3}|x|^{n+2s}}

    and, if |x|1|x|\leqslant 1,

    C4|x|n2𝒦(x)1C4{|x|2n if n3,1+|ln|x|| if n=2,1 if n=1.\frac{C_{4}}{|x|^{n-2}}\leqslant\mathcal{K}(x)\leqslant\frac{1}{C_{4}}\begin{cases}{|x|^{2-n}}\qquad&\text{ if }{n\geqslant 3},\\ 1+|\ln|x||&\text{ if }n=2,\\ 1&\text{ if }n=1.\end{cases}
  • (c)

    There exists a positive constant C5C_{5} such that, if |x|1|x|\geqslant 1,

    |𝒦(x)|C5|x|n+2s+1and|D2𝒦(x)|C5|x|n+2s+2.|\nabla\mathcal{K}(x)|\leqslant\frac{C_{5}}{|x|^{n+2s+1}}\qquad\text{and}\qquad|D^{2}\mathcal{K}(x)|\leqslant\frac{C_{5}}{|x|^{n+2s+2}}.
  • (d)

    If n3n\geqslant 3, then 𝒦Lq(n)\mathcal{K}\in L^{q}(\mathbb{R}^{n}) for all q[1,nn2)q\in[1,\frac{n}{n-2}). If n=1n=1, 22, then 𝒦Lq(n)\mathcal{K}\in L^{q}(\mathbb{R}^{n}) for all q[1,+)q\in[1,+\infty).

  • (e)

    If q[1,+)q\in[1,+\infty) and fLq(n)f\in L^{q}(\mathbb{R}^{n}), then the function u:=𝒦fu:=\mathcal{K}\ast f is a solution of

    Δu+(Δ)su+u=f in n.-\Delta u+(-\Delta)^{s}u+u=f\quad\text{ in }\mathbb{R}^{n}.

For the reader’s convenience, we provide the detailed proofs of Theorems 3.1 and 3.2 by using Fourier analysis techniques in Appendices A and B.

3.1. Hölder continuity of weak solutions

We devote this section to the proof of Theorem 1.2, which will achieved by combining LpL^{p}-theory and Theorem 3.2.

3.1.1. W2,pW^{2,p}-regularity of weak solutions

We shall explore the W2,pW^{2,p}-regularity theory for weak solutions of linear equations, which is a pivotal step towards the proof of the Hölder continuity for the nonlinear equation, which will be obtained combining the LpL^{p}-theory and a localization trick.

Lemma 3.3.

Let p1p\geqslant 1 and fLp(n)f\in L^{p}(\mathbb{R}^{n}). Let uu be a solution of

Δu+(Δ)su+u=fin n.-\Delta u+(-\Delta)^{s}u+u=f\qquad\text{in }\mathbb{R}^{n}.

Then, uW2,p(n)u\in W^{2,p}(\mathbb{R}^{n}) and

uW2,p(n)CfLp(n),\|u\|_{W^{2,p}(\mathbb{R}^{n})}\leqslant C\|f\|_{L^{p}(\mathbb{R}^{n})},

for some constant C>0C>0 depending on nn, ss and pp.

Proof.

The gist of the proof relies on checking that uu satisfies an equation of the type Δu+u=g-\Delta u+u=g, for some function gg, in order to apply the classical Calderón–Zygmund regularity theory to this equation. For this, the core of the argument will be to check that gLp(n)g\in L^{p}(\mathbb{R}^{n}) (and that gLp(n)CfLp(n)\|g\|_{L^{p}(\mathbb{R}^{n})}\leqslant C\|f\|_{L^{p}(\mathbb{R}^{n})}).

The technical details of the proof go as follows. From [46, Theorem 3, page 135], we know that we can identify the Sobolev space W2,p(n)W^{2,p}(\mathbb{R}^{n}) with the space

(3.1) 𝒲2,p:={uLp(n) s.t. 1((1+|ξ|2)u^)Lp(n)}.\mathcal{W}^{2,p}:=\left\{u\in L^{p}(\mathbb{R}^{n})\;{\mbox{ s.t. }}\;\mathcal{F}^{-1}\Big{(}(1+|\xi|^{2})\hat{u}\Big{)}\in L^{p}(\mathbb{R}^{n})\right\}.

In light of this, it suffices to show that u𝒲2,pu\in\mathcal{W}^{2,p}.

We first claim that

(3.2) uLp(n).u\in L^{p}(\mathbb{R}^{n}).

To this end, we point out that, in the distributional sense,

u=1(11+|ξ|2+|ξ|2s)f=𝒦f,u=\mathcal{F}^{-1}\left(\frac{1}{1+|\xi|^{2}+|\xi|^{2s}}\right)\ast f=\mathcal{K}\ast f,

where the kernel 𝒦\mathcal{K} is given in (1.4). Since 𝒦L1(n)\mathcal{K}\in L^{1}(\mathbb{R}^{n}) (thanks to (b) of Theorem 3.2), the Young’s convolution inequality gives that

upCn,sfp\|u\|_{p}\leqslant C_{n,s}\|f\|_{p}

for some constant Cn,s>0C_{n,s}>0. This estaslishes (3.2).

Moreover, we remark that uu satisfies

(1+|ξ|2)u^=1+|ξ|21+|ξ|2+|ξ|2sf^=g^,\displaystyle\Big{(}1+|\xi|^{2}\Big{)}\hat{u}=\frac{1+|\xi|^{2}}{1+|\xi|^{2}+|\xi|^{2s}}\hat{f}=\hat{g},
where g:=(δ0+𝒦1(1+|ξ|2s1+|ξ|2+|ξ|2s))f.\displaystyle g:=\left(\delta_{0}+\mathcal{K}-\mathcal{F}^{-1}\left(\frac{1+|\xi|^{2s}}{1+|\xi|^{2}+|\xi|^{2s}}\right)\right)\ast f.

Here above and in what follows δ0\delta_{0} denotes the Dirac’s delta at 0.

Now we define

ϕ(ξ)\displaystyle\phi(\xi) :=\displaystyle:= 1+|ξ|2s1+|ξ|2+|ξ|2s=11+|ξ|21+|ξ|2s\displaystyle\frac{1+|\xi|^{2s}}{1+|\xi|^{2}+|\xi|^{2s}}=\frac{1}{1+\frac{|\xi|^{2}}{1+|\xi|^{2s}}}
and (x)\displaystyle{\mbox{and }}\quad\mathcal{I}(x) :=\displaystyle:= 1ϕ(x)=0+etnet|ξ|21+|ξ|2se2πixξ𝑑ξ𝑑t.\displaystyle\mathcal{F}^{-1}\phi(x)=\int_{0}^{+\infty}e^{-t}\int_{\mathbb{R}^{n}}e^{-\frac{t|\xi|^{2}}{1+|\xi|^{2s}}}e^{2\pi ix\cdot\xi}\,d\xi\,dt.

We claim that

(3.3) L1(n).\mathcal{I}\in L^{1}(\mathbb{R}^{n}).

Once the claim in (3.3) will be established, the proof of Lemma 3.3 can be completed as follows. One notices that g=(δ0+𝒦)f=f+𝒦ffg=(\delta_{0}+\mathcal{K}-\mathcal{I})\ast f=f+\mathcal{K}\ast f-\mathcal{I}\ast f, and therefore

gLp(n)=f+𝒦ffLp(n)(1+𝒦L1(n)+L1(n))fLp(n).\|g\|_{L^{p}(\mathbb{R}^{n})}=\|f+\mathcal{K}\ast f-\mathcal{I}\ast f\|_{L^{p}(\mathbb{R}^{n})}\leqslant\left(1+\|\mathcal{K}\|_{L^{1}(\mathbb{R}^{n})}+\|\mathcal{I}\|_{L^{1}(\mathbb{R}^{n})}\right)\|f\|_{L^{p}(\mathbb{R}^{n})}.

From this, one concludes that u𝒲2,pu\in\mathcal{W}^{2,p} and that uW2,p(n)CfLp(n)\|u\|_{W^{2,p}(\mathbb{R}^{n})}\leqslant C\|f\|_{L^{p}(\mathbb{R}^{n})}, as desired.

Hence, to complete the proof of Lemma 3.3, we now focus on the proof of (3.3). For this, we define, for every κ\kappa, t1t_{1}, t2>0t_{2}>0 and xnx\in\mathbb{R}^{n},

𝒥(x,t1,κ,t2):=net1|ξ|2κ+t2|ξ|2se2πixξ𝑑ξ\mathcal{J}(x,t_{1},\kappa,t_{2}):=\int_{\mathbb{R}^{n}}e^{-\frac{t_{1}|\xi|^{2}}{\kappa+t_{2}|\xi|^{2s}}}e^{2\pi ix\cdot\xi}\,d\xi

and we observe that

(3.4) 𝒥(x,t,1,1)=tn2𝒥(t12x,1,1,ts)=tn22s𝒥(t122sx,1,ts1s,1).\mathcal{J}(x,t,1,1)=t^{-\frac{n}{2}}\mathcal{J}\left(t^{-\frac{1}{2}}x,1,1,t^{-s}\right)=t^{-\frac{n}{2-2s}}\mathcal{J}\left(t^{-\frac{1}{2-2s}}x,1,t^{\frac{s}{1-s}},1\right).

Also, for any t>0t>0, we notice that

(3.5) |𝒥(x,t,1,1)|Cn,s(tn22stn2).\begin{split}|\mathcal{J}(x,t,1,1)|&\leqslant C_{n,s}\left(t^{-\frac{n}{2-2s}}\vee t^{-\frac{n}{2}}\right).\end{split}

We split \mathcal{I} as (x)=1(x)+2(x)\mathcal{I}(x)=\mathcal{I}_{1}(x)+\mathcal{I}_{2}(x), where

1(x):=1+et𝒥(x,t,1,1)𝑑tand2(x):=01et𝒥(x,t,1,1)𝑑t.\mathcal{I}_{1}(x):=\int_{1}^{+\infty}e^{-t}\mathcal{J}(x,t,1,1)\,dt\qquad{\mbox{and}}\qquad\mathcal{I}_{2}(x):=\int_{0}^{1}e^{-t}\mathcal{J}(x,t,1,1)\,dt.

We will now prove in Step 1 that 1L1(n)\mathcal{I}_{1}\in L^{1}(\mathbb{R}^{n}) and in Step 2 that 2L1(n)\mathcal{I}_{2}\in L^{1}(\mathbb{R}^{n}). Thus, Step 1 and Step 2 will give the desired claim in (3.3).

Step 1. We prove that 1L1(n)\mathcal{I}_{1}\in L^{1}(\mathbb{R}^{n}) by exploiting the rescaling property (3.4). To this end, we observe that ts(0,1)t^{-s}\in(0,1) for every t(1,+)t\in(1,+\infty). Thus, we pick η(0,1)\eta\in(0,1) and use the Fourier Inversion Theorem for the radial function 𝒥(x,1,1,η)\mathcal{J}(x,1,1,\eta) (see e.g. [20, Chapter II]). In this way, we find that

(3.6) 𝒥(x,1,1,η)=(2π)n2|x|n210+er21+ηr2srn2Jn21(|x|r)𝑑r=(2π)n2|x|n0+et2|x|2+ηt2s|x|22stn2Jn21(t)𝑑t=(2π)n2|x|n0+et2|x|2+ηt2s|x|22s(2t1+n2|x|2+ηt2s|x|22s2sηt1+2s+n2|x|22s(|x|2+ηt2s|x|22s)2)Jn2(t)𝑑t,\begin{split}\mathcal{J}(x,1,1,\eta)&=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{\frac{n}{2}-1}}\int_{0}^{+\infty}e^{-\frac{r^{2}}{1+\eta r^{2s}}}r^{\frac{n}{2}}J_{\frac{n}{2}-1}(|x|r)\,dr\\ &=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{n}}\int_{0}^{+\infty}e^{-\frac{t^{2}}{|x|^{2}+\eta t^{2s}|x|^{2-2s}}}t^{\frac{n}{2}}J_{\frac{n}{2}-1}(t)\,dt\\ &=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{n}}\int_{0}^{+\infty}e^{-\frac{t^{2}}{|x|^{2}+\eta t^{2s}|x|^{2-2s}}}\left(\frac{2t^{1+\frac{n}{2}}}{|x|^{2}+\eta t^{2s}|x|^{2-2s}}-\frac{2s\eta t^{1+2s+\frac{n}{2}}|x|^{2-2s}}{\left(|x|^{2}+\eta t^{2s}|x|^{2-2s}\right)^{2}}\right)J_{\frac{n}{2}}(t)\,dt,\end{split}

where JvJ_{v} denotes the Bessel function of first kind of order vv.

We claim that

(3.7) lim|x|+supη(0,1)|x|n+22s𝒥(x,1,1,η)=0.\lim\limits_{|x|\to+\infty}\sup_{\eta\in(0,1)}|x|^{n+2-2s}\mathcal{J}(x,1,1,\eta)=0.

Indeed, from (3.6) it follows that

|x|n+22s𝒥(x,1,1,η)=(2π)n2 Re 0+et2|x|2+ηt2s|x|22s(2t1+n2|x|2s+ηt2s2sηt1+2s+n2(|x|2s+ηt2s)2)Hn21(t)𝑑t|x|^{n+2-2s}\mathcal{J}(x,1,1,\eta)={(2\pi)^{\frac{n}{2}}}\text{ Re }\int_{0}^{+\infty}e^{-\frac{t^{2}}{|x|^{2}+\eta t^{2s}|x|^{2-2s}}}\left(\frac{2t^{1+\frac{n}{2}}}{|x|^{2s}+\eta t^{2s}}-\frac{2s\eta t^{1+2s+\frac{n}{2}}}{\left(|x|^{2s}+\eta t^{2s}\right)^{2}}\right)H^{1}_{\frac{n}{2}}(t)\,dt

where Hn2(1)(z)H^{(1)}_{\frac{n}{2}}(z) is the Bessel function of the third kind and Re AA denotes the real part of AA.

We consider the straight line

L1:={z:argz=π6}L_{1}:=\left\{z\in\mathbb{C}:\arg z=\frac{\pi}{6}\right\}

and, using the Residue Theorem, we find that

|x|n+22s𝒥(x,1,1,η)=2(2π)n2 Re L1ez2|x|2+ηz2s|x|22s(z1+n2|x|2s+ηz2ssηz1+2s+n2(|x|2s+ηz2s)2)Hn21(z)𝑑z.|x|^{n+2-2s}\mathcal{J}(x,1,1,\eta)=2{(2\pi)^{\frac{n}{2}}}\text{ Re }\int_{L_{1}}e^{-\frac{z^{2}}{|x|^{2}+\eta z^{2s}|x|^{2-2s}}}\left(\frac{z^{1+\frac{n}{2}}}{|x|^{2s}+\eta z^{2s}}-\frac{s\eta z^{1+2s+\frac{n}{2}}}{\left(|x|^{2s}+\eta z^{2s}\right)^{2}}\right)H^{1}_{\frac{n}{2}}(z)\,dz.

Furthermore, we observe that, for any η(0,1)\eta\in(0,1) and |x|>1|x|>1,

(3.8) |L1ez2|x|2+ηz2s|x|22s(z1+n2|x|2s+ηz2ssηz1+2s+n2(|x|2s+ηz2s)2)Hn21(z)𝑑z|=|0+er2eiπ3|x|2+ηr2seisπ3|x|22s((reiπ6)1+n2|x|2s+η(reiπ6)2ssη(reiπ6)1+2s+n2(|x|2s+η(reiπ6)2s)2)Hn2(1)(reiπ6)eiπ6𝑑r|0+(r1+n2(|x|2s+2r2s)|x|4s+sr1+2s+n2(|x|4s+4r4s)|x|8s)|Hn2(1)(reiπ6)|𝑑r.\begin{split}&\left|\int_{L_{1}}e^{-\frac{z^{2}}{|x|^{2}+\eta z^{2s}|x|^{2-2s}}}\left(\frac{z^{1+\frac{n}{2}}}{|x|^{2s}+\eta z^{2s}}-\frac{s\eta z^{1+2s+\frac{n}{2}}}{\left(|x|^{2s}+\eta z^{2s}\right)^{2}}\right)H^{1}_{\frac{n}{2}}(z)\,dz\right|\\ =&\left|\int_{0}^{+\infty}e^{-\frac{r^{2}e^{i\frac{\pi}{3}}}{|x|^{2}+\eta r^{2s}e^{i\frac{s\pi}{3}}|x|^{2-2s}}}\left(\frac{(re^{i\frac{\pi}{6}})^{1+\frac{n}{2}}}{|x|^{2s}+\eta(re^{i\frac{\pi}{6}})^{2s}}-\frac{s\eta(re^{i\frac{\pi}{6}})^{1+2s+\frac{n}{2}}}{\left(|x|^{2s}+\eta(re^{i\frac{\pi}{6}})^{2s}\right)^{2}}\right)H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)e^{i\frac{\pi}{6}}\,dr\right|\\ \leqslant&\int_{0}^{+\infty}\left(\frac{r^{1+\frac{n}{2}}(|x|^{2s}+2r^{2s})}{|x|^{4s}}+\frac{sr^{1+2s+\frac{n}{2}}(|x|^{4s}+4r^{4s})}{|x|^{8s}}\right)\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr.\\ \end{split}

Employing (LABEL:estimate_of_H), we see that

(3.9) |Hn2(1)(reiπ6)|=|ieinπ4+eireiπ6et+et2en2t𝑑t|20+er4etent2𝑑t.\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|=\left|-ie^{-i\frac{n\pi}{4}}\int_{-\infty}^{+\infty}e^{ire^{i\frac{\pi}{6}}\frac{e^{t}+e^{-t}}{2}}e^{-\frac{n}{2}t}\,dt\right|\leqslant 2\int_{0}^{+\infty}e^{-\frac{r}{4}e^{t}}e^{\frac{nt}{2}}\,dt.

By combining (LABEL:H11) with (3.9), one obtains that, if η(0,1)\eta\in(0,1) and |x|>1|x|>1,

|L1ez2|x|2+ηz2s|x|22s(z1+n2|x|2s+ηz2ssηz1+2s+n2(|x|2s+ηz2s)2)Hn2(1)(z)𝑑z|2cn,s(|x|2s+|x|4s)\left|\int_{L_{1}}e^{-\frac{z^{2}}{|x|^{2}+\eta z^{2s}|x|^{2-2s}}}\left(\frac{z^{1+\frac{n}{2}}}{|x|^{2s}+\eta z^{2s}}-\frac{s\eta z^{1+2s+\frac{n}{2}}}{\left(|x|^{2s}+\eta z^{2s}\right)^{2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz\right|\leqslant 2c_{n,s}\left(|x|^{-2s}+|x|^{-4s}\right)

for some constant cn,sc_{n,s}. As a result, for any ε>0\varepsilon>0, there exists M>0M>0 independent of η\eta such that, for every |x|>M|x|>M and η(0,1)\eta\in(0,1),

|L1ez2|x|2+ηz2s|x|22s(z1+n2|x|2s+ηz2ssηz1+2s+n2(|x|2s+ηz2s)2)Hn2(1)(z)𝑑z|<ε2(2π)n2\left|\int_{L_{1}}e^{-\frac{z^{2}}{|x|^{2}+\eta z^{2s}|x|^{2-2s}}}\left(\frac{z^{1+\frac{n}{2}}}{|x|^{2s}+\eta z^{2s}}-\frac{s\eta z^{1+2s+\frac{n}{2}}}{\left(|x|^{2s}+\eta z^{2s}\right)^{2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz\right|<\frac{\varepsilon}{2(2\pi)^{\frac{n}{2}}}

which establishes (3.7), as desired.

By combining (3.4) with  (3.7), we know that there exists M>0M>0 depending on nn and ss such that, when t>1t>1 and |x|>Mt12|x|>Mt^{\frac{1}{2}},

0|𝒥(x,t,1,1)|<t1s|x|n+22s.0\leqslant\left|\mathcal{J}(x,t,1,1)\right|<\frac{t^{1-s}}{|x|^{n+2-2s}}.

Owing to this and (3.5), we conclude that, for every |x|>M|x|>M,

1(x)=1+et𝒥(x,t,1,1)𝑑t1|xM|2ett1s|x|n+22s𝑑t+Cn,s|xM|2+ettn2𝑑tc11|x|n+22s,\mathcal{I}_{1}(x)=\int_{1}^{+\infty}e^{-t}\mathcal{J}(x,t,1,1)\,dt\leqslant\int_{1}^{\left|\frac{x}{M}\right|^{2}}e^{-t}\frac{t^{1-s}}{|x|^{n+2-2s}}\,dt+C_{n,s}\int_{\left|\frac{x}{M}\right|^{2}}^{+\infty}e^{-t}t^{-\frac{n}{2}}\,dt\leqslant c_{1}\frac{1}{|x|^{n+2-2s}},

and, for every |x|M|x|\leqslant M,

1(x)=1+et𝒥(x,t,1,1)𝑑tCn,s|xM|2+ettn2𝑑tc21|x|n2+2s.\mathcal{I}_{1}(x)=\int_{1}^{+\infty}e^{-t}\mathcal{J}(x,t,1,1)\,dt\leqslant C_{n,s}\int_{\left|\frac{x}{M}\right|^{2}}^{+\infty}e^{-t}t^{-\frac{n}{2}}\,dt\leqslant c_{2}\frac{1}{|x|^{n-2+2s}}.

From the last two formulas we obtain that 1L1(n)\mathcal{I}_{1}\in L^{1}(\mathbb{R}^{n}), as desired.

Step 2. We now prove that 2L1(n)\mathcal{I}_{2}\in L^{1}(\mathbb{R}^{n}). This will be a byproduct of the rescaling property (3.4). The full argument goes as follows. Since t(0,1)t\in(0,1), one has that ts1s(0,1)t^{\frac{s}{1-s}}\in(0,1). We thus use again the Fourier Inversion Theorem for the radial function 𝒥(x,1,η,1)\mathcal{J}(x,1,\eta,1), finding that

𝒥(x,1,η,1)=ne|ξ|2η+|ξ|2se2πixξ𝑑ξ=(2π)n2|x|n0+et2η|x|2+t2s|x|22s(2t1+n2η|x|2+t2s|x|22s2st1+2s+n2|x|22s(η|x|2+t2s|x|22s)2)Jn2(t)𝑑t.\begin{split}\mathcal{J}(x,1,\eta,1)&=\int_{\mathbb{R}^{n}}e^{-\frac{|\xi|^{2}}{\eta+|\xi|^{2s}}}e^{2\pi ix\cdot\xi}\,d\xi\\ &=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{n}}\int_{0}^{+\infty}e^{-\frac{t^{2}}{\eta|x|^{2}+t^{2s}|x|^{2-2s}}}\left(\frac{2t^{1+\frac{n}{2}}}{\eta|x|^{2}+t^{2s}|x|^{2-2s}}-\frac{2st^{1+2s+\frac{n}{2}}|x|^{2-2s}}{\left(\eta|x|^{2}+t^{2s}|x|^{2-2s}\right)^{2}}\right)J_{\frac{n}{2}}(t)\,dt.\end{split}

We now claim that

(3.10) lim|x|+supη(0,1)|x|n+1s𝒥(x,1,η,1)=0.\lim\limits_{|x|\to+\infty}\sup_{\eta\in(0,1)}|x|^{n+1-s}\mathcal{J}(x,1,\eta,1)=0.

Indeed, using a similar argument as in the proof of Step 1, for any η(0,1)\eta\in(0,1) and |x|>1|x|>1, one has

||x|n+1s𝒥(x,1,η,1)|2(2π)n2|L1ez2η|x|2+z2s|x|22s(z1+n2η|x|1+s+z2s|x|1ssz1+2s+n2(η|x|1+3s2+z2s|x|1s2)2)Hn21(z)𝑑z|2(2π)n20+(6r1+n22s|x|1s+17sr12s+n2|x|1s)|Hn2(1)(reiπ6)|𝑑r.\begin{split}\left||x|^{n+1-s}\mathcal{J}(x,1,\eta,1)\right|&\leqslant 2(2\pi)^{\frac{n}{2}}\left|\int_{L_{1}}e^{-\frac{z^{2}}{\eta|x|^{2}+z^{2s}|x|^{2-2s}}}\left(\frac{z^{1+\frac{n}{2}}}{\eta|x|^{1+s}+z^{2s}|x|^{1-s}}-\frac{sz^{1+2s+\frac{n}{2}}}{\left(\eta|x|^{\frac{1+3s}{2}}+z^{2s}|x|^{\frac{1-s}{2}}\right)^{2}}\right)H^{1}_{\frac{n}{2}}(z)\,dz\right|\\ &\leqslant 2(2\pi)^{\frac{n}{2}}\int_{0}^{+\infty}\left(\frac{6r^{1+\frac{n}{2}-2s}}{|x|^{1-s}}+\frac{17sr^{1-2s+\frac{n}{2}}}{|x|^{1-s}}\right)\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr.\end{split}

Owing to (3.9), for any η(0,1)\eta\in(0,1) and |x|>1|x|>1, we infer that

|L1ez2η|x|2+z2s|x|22s(z1+n2η|x|2s+z2s|x|1ssz1+2s+n2(η|x|1+3s2+z2s|x|1s2)2)Hn21(z)𝑑z|cn,s|x|s1.\left|\int_{L_{1}}e^{-\frac{z^{2}}{\eta|x|^{2}+z^{2s}|x|^{2-2s}}}\left(\frac{z^{1+\frac{n}{2}}}{\eta|x|^{2s}+z^{2s}|x|^{1-s}}-\frac{sz^{1+2s+\frac{n}{2}}}{\left(\eta|x|^{\frac{1+3s}{2}}+z^{2s}|x|^{\frac{1-s}{2}}\right)^{2}}\right)H^{1}_{\frac{n}{2}}(z)\,dz\right|\leqslant c_{n,s}|x|^{s-1}.

Hence, for any ε>0\varepsilon>0, there exists M>0M>0 independent of η\eta such that, for every |x|>M|x|>M and η(0,1)\eta\in(0,1),

|L1ez2η|x|2+z2s|x|22s(z1+n2η|x|2s+z2s|x|1ssz1+2s+n2(η|x|1+3s2+z2s|x|1s2)2)Hn21(z)𝑑z|<ε2(2π)n2,\left|\int_{L_{1}}e^{-\frac{z^{2}}{\eta|x|^{2}+z^{2s}|x|^{2-2s}}}\left(\frac{z^{1+\frac{n}{2}}}{\eta|x|^{2s}+z^{2s}|x|^{1-s}}-\frac{sz^{1+2s+\frac{n}{2}}}{\left(\eta|x|^{\frac{1+3s}{2}}+z^{2s}|x|^{\frac{1-s}{2}}\right)^{2}}\right)H^{1}_{\frac{n}{2}}(z)\,dz\right|<\frac{\varepsilon}{2(2\pi)^{\frac{n}{2}}},

which establishes (3.10).

From (3.4) and (3.10), we can find M>0M>0, depending only on nn and ss, such that, if |x|>Mt122s|x|>Mt^{\frac{1}{2-2s}} and t(0,1)t\in(0,1),

0|𝒥(x,t,1,1)|<t12|x|n+1s.0\leqslant\left|\mathcal{J}(x,t,1,1)\right|<\frac{t^{\frac{1}{2}}}{|x|^{n+1-s}}.

Owing to this and (3.5), we find that, for every |x|>M|x|>M,

2(x)=01et𝒥(x,t,1,1)𝑑t0|xM|22sett12|x|n+1s𝑑tc11|x|n+1s,\mathcal{I}_{2}(x)=\int_{0}^{1}e^{-t}\mathcal{J}(x,t,1,1)\,dt\leqslant\int_{0}^{\left|\frac{x}{M}\right|^{2-2s}}e^{-t}\frac{t^{\frac{1}{2}}}{|x|^{n+1-s}}\,dt\leqslant c_{1}\frac{1}{|x|^{n+1-s}},

and, for every |x|M|x|\leqslant M,

2(x)=01et𝒥(x,t,1,1)𝑑t0|xM|22sett12|x|n+1s𝑑t+Cn,s|xM|22s1ettn22s𝑑tc21|x|n1+s.\mathcal{I}_{2}(x)=\int_{0}^{1}e^{-t}\mathcal{J}(x,t,1,1)\,dt\leqslant\int_{0}^{\left|\frac{x}{M}\right|^{2-2s}}e^{-t}\frac{t^{\frac{1}{2}}}{|x|^{n+1-s}}\,dt+C_{n,s}\int_{\left|\frac{x}{M}\right|^{2-2s}}^{1}e^{-t}t^{-\frac{n}{2-2s}}\,dt\leqslant c_{2}\frac{1}{|x|^{n-1+s}}.

Combining the last two formulas, one deduces that 2L1(n)\mathcal{I}_{2}\in L^{1}(\mathbb{R}^{n}) as well. ∎

3.1.2. C0,μC^{0,\mu}-regularity of weak solutions

We dedicate this part to show the C0,μC^{0,\mu}-regularity of weak solutions based on Lemma 3.3 and the usual iteration technique.

For this, we first make the following observation:

Lemma 3.4.

For all t(,n2)t\in\left(-\infty,\frac{n}{2}\right), let

ψ(t):=ntn2t.\psi(t):=\frac{nt}{n-2t}.

Then, for every t0(n2n2p,n2)t_{0}\in\left(\frac{n}{2}-\frac{n}{2p},\frac{n}{2}\right) there exists j0=j0(t0)j_{0}=j_{0}(t_{0})\in\mathbb{N} with j01j_{0}\geqslant 1 such that

ψψj0 times(t0)n2>ψψj times(t0)for all j<j0.\stackrel{{\scriptstyle{j_{0}{\mbox{ times}}}}}{{\psi\circ\dots\circ\psi}}(t_{0})\geqslant\frac{n}{2}>\stackrel{{\scriptstyle{j{\mbox{ times}}}}}{{\psi\circ\dots\circ\psi}}(t_{0})\qquad{\mbox{for all~{}$j<j_{0}$}}.
Proof.

Denote by

ψj(t0):=ψψj times(t0)\psi_{j}(t_{0}):=\stackrel{{\scriptstyle{j{\mbox{ times}}}}}{{\psi\circ\dots\circ\psi}}(t_{0})

and suppose by contradiction that, for all j1j\geqslant 1,

ψj(t0)<n2.\psi_{j}(t_{0})<\frac{n}{2}.

Also, notice that ψ(t)t\psi(t)\geqslant t, and therefore, for all j1j\geqslant 1, we have that ψj+1(t0)ψj(t0)\psi_{j+1}(t_{0})\geqslant\psi_{j}(t_{0}). Hence, the following limit exists

:=limj+ψj(t0)\ell:=\lim_{j\to+\infty}\psi_{j}(t_{0})

and moreover, by construction, [t0,n2]\ell\in\left[t_{0},\frac{n}{2}\right].

In addition,

ψj+1(t0)=nψj(t0)n2ψj(t0)\psi_{j+1}(t_{0})=\frac{n\,\psi_{j}(t_{0})}{n-2\psi_{j}(t_{0})}

and thus, taking the limit in jj,

=nn2.\ell=\frac{n\ell}{n-2\ell}.

Solving for \ell, we find that =0\ell=0, which gives the desired contradiction. ∎

Proof of Theorem 1.2.

We consider a sequence of cut off functions ϕjC0(n)\phi_{j}\in C^{\infty}_{0}(\mathbb{R}^{n}) for any j1j\geqslant 1 satisfying

(3.11) ϕj1 in B1/22j2,supp(ϕj)B1/22j3and0ϕj1 in n.\phi_{j}\equiv 1\text{ in }B_{1/2^{2j-2}},\qquad\text{supp}(\phi_{j})\subset B_{1/2^{2j-3}}\qquad{\mbox{and}}\qquad 0\leqslant\phi_{j}\leqslant 1\text{ in }\mathbb{R}^{n}.

Let uju_{j} solve

(3.12) Δuj+(Δ)suj+uj=ϕjupin n.-\Delta u_{j}+(-\Delta)^{s}u_{j}+u_{j}=\phi_{j}u^{p}\qquad\text{in }\mathbb{R}^{n}.

Then

Δ(uuj)+(Δ)s(uuj)+(uuj)=(1ϕj)upin n.-\Delta(u-u_{j})+(-\Delta)^{s}(u-u_{j})+(u-u_{j})=(1-\phi_{j})u^{p}\qquad\text{in }\mathbb{R}^{n}.

Moreover,

uuj=𝒦((1ϕj)up)u-u_{j}=\mathcal{K}\ast\Big{(}(1-\phi_{j})u^{p}\Big{)}

where 𝒦\mathcal{K} is given by (1.4).

Let now q0q_{0} be either equal to the critical exponent 2nn2\frac{2n}{n-2} if n>2n>2 or any real number in (p,+)(p,+\infty) if n=2n=2. Let also θ:=q0/(q0p)>1\theta:=q_{0}/(q_{0}-p)>1.

Moreover, we set γ1:=q0/p\gamma_{1}:=q_{0}/p. We observe that γ1>n2n2p\gamma_{1}>\frac{n}{2}-\frac{n}{2p}, thanks to (1.2). Hence, in light of Lemma 3.4, we can define

j:={0 if γ1n2,j0(γ1) if γ1<n2,j_{\star}:=\begin{dcases}0&{\mbox{ if }}\gamma_{1}\geqslant\frac{n}{2},\\ j_{0}(\gamma_{1})&{\mbox{ if }}\gamma_{1}<\frac{n}{2},\end{dcases}

and, for every j{0,,j}j\in\{0,\dots,j_{\star}\},

γj+1:=ψψj times(γ1).\gamma_{j+1}:=\stackrel{{\scriptstyle{j{\mbox{ times}}}}}{{\psi\circ\dots\circ\psi}}(\gamma_{1}).

In this way, we have that

(3.13) γj+1n2.\gamma_{j_{\star}+1}\geqslant\frac{n}{2}.

Also, for every j{0,,j}j\in\{0,\dots,j_{\star}\}, we set qj:=pγj+1q_{j}:=p\gamma_{j+1}.

In this setting, we claim that, for every j{1,,j+1}j\in\{1,\dots,j_{\star}+1\},

(3.14) ujW2,γj(n)Cj(uH1(n)pj+uH1(n)p).\|u_{j}\|_{W^{2,\gamma_{j}}(\mathbb{R}^{n})}\leqslant C_{j}\,\Big{(}\|u\|^{p^{j}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\Big{)}.

The proof of this claim is by induction over jj.

We first check that (3.14) holds true when j=1j=1. For this sake, we notice that γ1=q0/p\gamma_{1}=q_{0}/p, and therefore, since uLq0(n)u\in L^{q_{0}}(\mathbb{R}^{n}), one has that ϕ1upLγ1(n)\phi_{1}u^{p}\in L^{\gamma_{1}}(\mathbb{R}^{n}). From this and Lemma 3.3, applied to the equation for u1u_{1} in (3.12), it then follows that u1W2,γ1(n)u_{1}\in W^{2,\gamma_{1}}(\mathbb{R}^{n}) and

(3.15) u1W2,γ1(n)Cϕ1upLγ1(n)=Cϕ1upLγ1(B2)CuLq0(B2)pCuH1(n)p,\|u_{1}\|_{W^{2,\gamma_{1}}(\mathbb{R}^{n})}\leqslant C\|\phi_{1}u^{p}\|_{L^{\gamma_{1}}(\mathbb{R}^{n})}=C\|\phi_{1}u^{p}\|_{L^{\gamma_{1}}(B_{2})}\leqslant C\|u\|^{p}_{L^{q_{0}}(B_{2})}\leqslant C\|u\|^{p}_{H^{1}(\mathbb{R}^{n})},

up to renaming C>0C>0. The last inequality in (3.15) uses the Sobolev Embedding (see also [22, Corollary 9.11] when n=2n=2). This establishes (3.14) when j=1j=1.

Now we suppose that (3.14) holds true for all the indexes in {1,,j}\{1,\dots,j\} with jjj\leqslant j_{\star} and we prove it for the index j+1j+1. For this, we observe that, since γj<n/2\gamma_{j}<n/2 by construction, we have that qj=nγjn2γjq_{j}=\frac{n\gamma_{j}}{n-2\gamma_{j}} and γj+1=qj/p\gamma_{j+1}=q_{j}/p.

Moreover, owing to the Hölder inequality (used here with exponents q0/pq_{0}/p and θ\theta) and the smoothness of 𝒦\mathcal{K} away from the origin and its decay (recall (a) and (b) of Theorem 3.2), we have that, for every xB1/22j1x\in B_{1/2^{2j-1}},

|u(x)uj(x)|=|𝒦((1ϕj)up)(x)|𝒦Lθ(nB1/22j2)uLq0(n)pCuLq0(n)p.|u(x)-u_{j}(x)|=\Big{|}\mathcal{K}\ast\Big{(}(1-\phi_{j})u^{p}\Big{)}(x)\Big{|}\leqslant\|\mathcal{K}\|_{L^{\theta}(\mathbb{R}^{n}\setminus B_{1/2^{2j-2}})}\|u\|^{p}_{{L^{q_{0}}(\mathbb{R}^{n})}}\leqslant C\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}.

This and the triangle inequality give that

uLqj(B1/22j1)CuLq0(n)p+ujLqj(B1/22j1).\|u\|_{L^{q_{j}}(B_{1/2^{2j-1}})}\leqslant C\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}+\|u_{j}\|_{L^{q_{j}}(B_{1/2^{2j-1}})}.

From this and the Sobolev Embedding, we deduce that

uLqj(B1/22j1)C(uLq0(n)p+ujW2,γj(n)),\|u\|_{L^{q_{j}}(B_{1/2^{2j-1}})}\leqslant C\left(\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}+\|u_{j}\|_{W^{2,\gamma_{j}}(\mathbb{R}^{n})}\right),

up to renaming C>0C>0, depending on nn, ss, pp and jj. Hence, using the inductive assumption,

uLqj(B1/22j1)C(uLq0(n)p+uH1(n)pj+uH1(n)p).\|u\|_{L^{q_{j}}(B_{1/2^{2j-1}})}\leqslant C\left(\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}+\|u\|^{p^{j}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right).

Using again the Sobolev Embedding and renaming CC once more,

uLqj(B1/22j1)C(uH1(n)pj+uH1(n)p).\|u\|_{L^{q_{j}}(B_{1/2^{2j-1}})}\leqslant C\left(\|u\|^{p^{j}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right).

Hence, we have that ϕj+1upLγj+1(n)\phi_{j+1}u^{p}\in L^{\gamma_{j+1}}(\mathbb{R}^{n}) and

ϕj+1upLγj+1(n)CuLqj(B1/(22j1))pC(uH1(n)pj+uH1(n)p)p\displaystyle\|\phi_{j+1}u^{p}\|_{L^{\gamma_{j+1}}(\mathbb{R}^{n})}\leqslant C\|u\|^{p}_{L^{q_{j}}(B_{1/(2^{2j-1})})}\leqslant C\left(\|u\|^{p^{j}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right)^{p}
C(uH1(n)pj+1+uH1(n)p2)C(uH1(n)pj+1+uH1(n)p).\displaystyle\qquad\qquad\leqslant C\left(\|u\|^{p^{j+1}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p^{2}}_{H^{1}(\mathbb{R}^{n})}\right)\leqslant C\left(\|u\|^{p^{j+1}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right).

We can therefore use Lemma 3.3 for the equation (3.12) for uj+1u_{j+1}. In this way, we obtain that uj+1W2,γj+1(n)u_{j+1}\in W^{2,\gamma_{j+1}}(\mathbb{R}^{n}) and

uj+1W2,γj+1(n)Cϕj+1upLγj+1(n)C(uH1(n)pj+1+uH1(n)p),\|u_{j+1}\|_{W^{2,\gamma_{j+1}}(\mathbb{R}^{n})}\leqslant C\|\phi_{j+1}u^{p}\|_{L^{\gamma_{j+1}}(\mathbb{R}^{n})}\leqslant C\left(\|u\|^{p^{j+1}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right),

for some constant C>0C>0, depending on nn, ss, pp and jj. This completes the proof of the claim in (3.14).

Using (3.14) with j+1j_{\star}+1, we find that

(3.16) uj+1W2,γj+1(n)C(uH1(n)pj+1+uH1(n)p).\|u_{j_{\star}+1}\|_{W^{2,\gamma_{j_{\star}+1}}(\mathbb{R}^{n})}\leqslant C\,\Big{(}\|u\|^{p^{{j_{\star}+1}}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\Big{)}.

Now, in view of (3.13), we distinguish two cases, either γj+1>n/2\gamma_{j_{\star}+1}>{n}/{2} or γj+1=n/2\gamma_{j_{\star}+1}={n}/{2}.

If γj+1>n/2\gamma_{j_{\star}+1}>{n}/{2} we use the Sobolev Embedding Theorem, by choosing

μ(0,min{1,2nγj0+1}),\mu\in\left(0,\,\min\left\{1,2-\frac{n}{\gamma_{j_{0}+1}}\right\}\right),

and we see that uj+1C0,μ(n)u_{j_{\star}+1}\in C^{0,\mu}(\mathbb{R}^{n}). In particular, from (3.16), it follows that

(3.17) uj+1C0,μ(n)C(uH1(n)pj+1+uH1(n)p).\|u_{j_{\star}+1}\|_{C^{0,\mu}(\mathbb{R}^{n})}\leqslant C\left(\|u\|^{p^{j_{\star}+1}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right).

Moreover, using the smoothness of 𝒦\mathcal{K} away from the origin, when |x|<1/22j+1|x|<{1}/{2^{2j_{\star}+1}} one has that

|(uuj+1)|n|𝒦(xy)||(1ϕj+1(y))up(y)|𝑑y𝒦Lθ(nB1/(22j+1))uLq0(n)pCuLq0(n)p.\begin{split}&|\nabla(u-u_{j_{\star}+1})|\leqslant\int_{\mathbb{R}^{n}}|\nabla\mathcal{K}(x-y)||(1-\phi_{{j_{\star}+1}}(y))u^{p}(y)|\,dy\\ &\qquad\leqslant\|\mathcal{\nabla K}\|_{L^{\theta}(\mathbb{R}^{n}\setminus B_{{1}/({2^{2j_{\star}+1}})})}\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}\leqslant C\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}.\end{split}

By combining this and (3.17), we deduce that, for any xx, yB122j+1y\in B_{\frac{1}{2^{2j_{\star}+1}}} with xyx\neq y,

|u(x)u(y)||xy|μ=|(uuj+1)(x)(uuj+1)(y)+uj+1(x)uj+1(y)||xy|μsupξB122j+1|(uuj+1)(ξ)||xy|1μ+uj+1C0,μ(n)C(uH1(n)pj+1+uH1(n)p)\begin{split}\frac{|u(x)-u(y)|}{|x-y|^{\mu}}&=\frac{|(u-u_{j_{\star}+1})(x)-(u-u_{j_{\star}+1})(y)+u_{j_{\star}+1}(x)-u_{j_{\star}+1}(y)|}{|x-y|^{\mu}}\\ &\leqslant\sup\limits_{\xi\in B_{\frac{1}{2^{2j_{\star}+1}}}}|\nabla(u-u_{j_{\star}+1})(\xi)||x-y|^{1-\mu}+\|u_{j_{\star}+1}\|_{C^{0,\mu}(\mathbb{R}^{n})}\\ &\leqslant C\left(\|u\|^{p^{j_{\star}+1}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right)\end{split}

and

|u(x)||u(x)uj+1(x)|+|uj+1(x)|C(uH1(n)pj+1+uH1(n)p).|u(x)|\leqslant|u(x)-u_{j_{\star}+1}(x)|+|u_{j_{\star}+1}(x)|\leqslant C\left(\|u\|^{p^{j_{\star}+1}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right).

The ball B122j+1B_{\frac{1}{2^{2j_{\star}+1}}} is centred at the origin, but we may arbitrarily move it around n\mathbb{R}^{n}. Covering n\mathbb{R}^{n} with these balls, we conclude that uC0,μ(n)u\in C^{0,\mu}(\mathbb{R}^{n}) for some μ(0,1)\mu\in(0,1).

This completes the desired result when γj+1>n/2\gamma_{j_{\star}+1}>{n}/{2} and we now focus on the case γj+1=n/2\gamma_{j_{\star}+1}={n}/{2}.

Recalling (3.1) and using [46, Theorem 3, page 135], we see that uj+1𝒲2,n2u_{j_{\star}+1}\in\mathcal{W}^{2,\frac{n}{2}}. Furthermore, owing to [46, formula (40), page 135], we know that uj+1𝒲1+45,n2u_{j_{\star}+1}\in\mathcal{W}^{1+\frac{4}{5},\frac{n}{2}}, where

𝒲1+45,n2:={uLn2(n) s.t. 1((1+|ξ|2)1+4/52u^)Ln2(n)}.\mathcal{W}^{1+\frac{4}{5},\frac{n}{2}}:=\left\{u\in L^{\frac{n}{2}}(\mathbb{R}^{n})\;{\mbox{ s.t. }}\;\mathcal{F}^{-1}\left((1+|\xi|^{2})^{\frac{1+4/5}{2}}\hat{u}\right)\in L^{\frac{n}{2}}(\mathbb{R}^{n})\right\}.

Thus, from [30, Theorem 3.2], it follows that uj+1L5n(n)u_{j_{\star}+1}\in L^{5n}(\mathbb{R}^{n}) and

(3.18) uj+1L5n(n)Cuj+1W2,n2(n),\|u_{j_{\star}+1}\|_{L^{5n}(\mathbb{R}^{n})}\leqslant C\|u_{j_{\star}+1}\|_{W^{2,\frac{n}{2}}(\mathbb{R}^{n})},

for some constant C>0C>0.

Using the decay properties of 𝒦\mathcal{K} as given by Theorem 3.2, we obtain that, for any xB1/(22j+1)x\in B_{1/(2^{2j_{\star}+1})},

(3.19) |uuj+1|=|𝒦((1ϕj0+1)up)|𝒦Lθ(nB1/(22j+1))uLq0(n)pCuLq0(n)p.|u-u_{j_{\star}+1}|=\left|\mathcal{K}\ast\Big{(}(1-\phi_{j_{0}+1})u^{p}\Big{)}\right|\leqslant\|\mathcal{K}\|_{L^{\theta}(\mathbb{R}^{n}\setminus B_{1/(2^{2j_{\star}+1})})}\|u\|^{p}_{{L^{q_{0}}(\mathbb{R}^{n})}}\leqslant C\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}.

By combining (3.16), (3.18) and (3.19), we have that

uL5n(B1/(22j+1))C(uH1(n)pj+1+uH1(n)p).\|u\|_{L^{5n}(B_{1/(2^{2j_{\star}+1})})}\leqslant C\left(\|u\|^{{p}^{j_{\star}+1}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right).

This gives that ϕj+2upL5n/p(n)\phi_{j_{\star}+2}u^{p}\in L^{5n/p}(\mathbb{R}^{n}) and therefore, by Lemma 3.3, we deduce that uj+2W2,5np(n)u_{j_{\star}+2}\in W^{2,\frac{5n}{p}}(\mathbb{R}^{n}). Since p<n+2n2p<\frac{n+2}{n-2}, we observe that 5np>n2\frac{5n}{p}>\frac{n}{2}. Hence, using the Sobolev Embedding Theorem, we have that uj+2C0,μ(n)u_{j_{\star}+2}\in C^{0,\mu}(\mathbb{R}^{n}), with

μ(0min{1,2p5}),\mu\in\left(0\,\min\left\{1,2-\frac{p}{5}\right\}\right),

and

(3.20) uj+2C0,μ(n)Cuj+2W2,5n/p(n)C(uH1(n)pj+2+uH1(n)p).\|u_{j_{\star}+2}\|_{C^{0,\mu}(\mathbb{R}^{n})}\leqslant C\|u_{j_{\star}+2}\|_{W^{2,5n/p}(\mathbb{R}^{n})}\leqslant C\left(\|u\|^{p^{j_{\star}+2}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right).

Moreover, by the smoothness of 𝒦\mathcal{K}, for every |x|<1/22j+3|x|<{1}/{2^{2j_{\star}+3}}, we have that

|(uuj+2)|𝒦Lθ(nB1/(22j+3))uLq0(n)pCuLq0(n)p.|\nabla(u-u_{j_{\star}+2})|\leqslant\|\mathcal{\nabla K}\|_{L^{\theta}(\mathbb{R}^{n}\setminus B_{{1}/({2^{2j_{\star}+3}})})}\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}\leqslant C\|u\|^{p}_{L^{q_{0}}(\mathbb{R}^{n})}.

By combining this and (3.20), we have that, for any xx, yB122j+3y\in B_{\frac{1}{2^{2j_{\star}+3}}} with xyx\neq y,

|u(x)u(y)||xy|μC(uH1(n)pj+2+uH1(n)p)\frac{|u(x)-u(y)|}{|x-y|^{\mu}}\leqslant C\left(\|u\|^{p^{j_{\star}+2}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right)

and

|u(x)||u(x)uj+2(x)|+|uj+2(x)|C(uH1(n)pj+1+uH1(n)p).|u(x)|\leqslant|u(x)-u_{j_{\star}+2}(x)|+|u_{j_{\star}+2}(x)|\leqslant C\left(\|u\|^{p^{j_{\star}+1}}_{H^{1}(\mathbb{R}^{n})}+\|u\|^{p}_{H^{1}(\mathbb{R}^{n})}\right).

The ball B122j0+3B_{\frac{1}{2^{2j_{0}+3}}} is centred at the origin, but we may arbitrarily move it around n\mathbb{R}^{n}. Covering n\mathbb{R}^{n} with these balls, we conclude that uC0,μ(n)u\in C^{0,\mu}(\mathbb{R}^{n}) for some μ(0,1)\mu\in(0,1). This completes the proof of Theorem 1.2. ∎

3.2. C1,αC^{1,\alpha}-regularity of weak solutions

We aim here to establish the C1,αC^{1,\alpha}-regularity of weak solutions of (1.1) based on the LpL^{p}-theory and the smoothness of the kernel 𝒦\mathcal{K} defined by (1.4).

Proof of Theorem 1.3.

We consider the cut off function ϕ1C0(n)\phi_{1}\in C^{\infty}_{0}(\mathbb{R}^{n}) satisfying the properties in (3.11) with j=1j=1. Let u1u_{1} solve

Δu1+(Δ)su1+u1=ϕ1upin n.-\Delta u_{1}+(-\Delta)^{s}u_{1}+u_{1}=\phi_{1}u^{p}\qquad\text{in }\mathbb{R}^{n}.

Then

Δ(uu1)+(Δ)s(uu1)+(uu1)=(1ϕ1)upin n.-\Delta(u-u_{1})+(-\Delta)^{s}(u-u_{1})+(u-u_{1})=(1-\phi_{1})u^{p}\qquad\text{in }\mathbb{R}^{n}.

Moreover, we deduce that

uu1=𝒦((1ϕ1)up).u-u_{1}=\mathcal{K}\ast\Big{(}(1-\phi_{1})u^{p}\Big{)}.

We recall that uL(n)u\in L^{\infty}(\mathbb{R}^{n}), thanks to Theorem 1.2.

Now, owing to the Hölder inequality and the smoothness of 𝒦\mathcal{K} away from the origin and its decay (recall Theorem 3.2), we have that, for every xB1/2x\in B_{{1}/{2}},

|u(x)u1(x)|=|𝒦((1ϕ1)up)(x)|𝒦L1(nB1/2)uL(n)pcn,s,puL(n)p,|u(x)-u_{1}(x)|=\Big{|}\mathcal{K}\ast\Big{(}(1-\phi_{1})u^{p}\Big{)}(x)\Big{|}\leqslant\|\mathcal{K}\|_{L^{1}(\mathbb{R}^{n}\setminus B_{1/2})}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}\leqslant c_{n,s,p}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})},
|(uu1)(x)|n|𝒦(xy)||(1ϕ1(y))up(y)|𝑑y𝒦L1(nB1/2)uL(n)pcn,s,puL(n)p|\nabla(u-u_{1})(x)|\leqslant\int_{\mathbb{R}^{n}}|\nabla\mathcal{K}(x-y)||(1-\phi_{1}(y))u^{p}(y)|\,dy\leqslant\|\mathcal{\nabla K}\|_{L^{1}(\mathbb{R}^{n}\setminus B_{1/2})}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}\leqslant c_{n,s,p}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}

and

(3.21) |D2(uu1)(x)|n|D2K(xy)||(1ϕ1(y))up(y)|𝑑yD2KL1(nB1/2)uL(n)pcn,suL(n)p.|D^{2}(u-u_{1})(x)|\leqslant\int_{\mathbb{R}^{n}}|{D^{2}K}(x-y)||(1-\phi_{1}(y))u^{p}(y)|\,dy\leqslant\|{D^{2}K}\|_{L^{1}(\mathbb{R}^{n}\setminus B_{1/2})}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}\leqslant c_{n,s}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}.

In the light of these estimates, we can focus on the regularity of u1u_{1}. For this, we observe that, for any q1q\geqslant 1,

(3.22) ϕ1upLq(n)=(n|ϕ1(x)up(x)|q𝑑x)1/q(πn2n2Γ(n2))1/quL(n)p(1+πn2n2Γ(n2))uL(n)p.\|\phi_{1}u^{p}\|_{L^{q}(\mathbb{R}^{n})}=\left(\int_{\mathbb{R}^{n}}|\phi_{1}(x)u^{p}(x)|^{q}\,dx\right)^{1/q}\leqslant\left(\frac{\pi^{\frac{n}{2}}}{\frac{n}{2}\Gamma(\frac{n}{2})}\right)^{1/q}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}\leqslant\left(1+\frac{\pi^{\frac{n}{2}}}{\frac{n}{2}\Gamma(\frac{n}{2})}\right)\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}.

That is, ϕ1upLq(n)\phi_{1}u^{p}\in L^{q}(\mathbb{R}^{n}) for any q1q\geqslant 1. From Lemma 3.3, it then follows that u1W2,q(n)u_{1}\in W^{2,q}(\mathbb{R}^{n}) for any q1q\geqslant 1, and in particular one can take any q>nq>n. Accordingly, from the Sobolev Embedding, we deduce that u1C1,1nq(n)u_{1}\in C^{1,1-\frac{n}{q}}(\mathbb{R}^{n}).

Thus, denoting by α:=1nq\alpha:=1-\frac{n}{q}, from Lemma 3.3 and (3.22), it follows that

u1C1,α(n)Cnu1W2,q(n)Cn,sϕ1upLq(n)Cn,suL(n)p\|u_{1}\|_{C^{1,\alpha}(\mathbb{R}^{n})}\leqslant C_{n}\|u_{1}\|_{W^{2,q}(\mathbb{R}^{n})}\leqslant C_{n,s}\|\phi_{1}u^{p}\|_{L^{q}(\mathbb{R}^{n})}\leqslant C_{n,s}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}

for some constant Cn,sC_{n,s} independent of α\alpha.

From this and (3.21), we deduce that for any xx, yB1/2y\in B_{1/2} with xyx\neq y,

|Du(x)Du(y)||xy|α=|D(uu1)(x)D(uu1)(y)+Du1(x)Du1(y)||xy|αsupξB1/4|D2(uu1)(ξ)||xy|1α+u1C1,α(n)cn,s,puL(n)p.\begin{split}\frac{|Du(x)-Du(y)|}{|x-y|^{\alpha}}&=\frac{|D(u-u_{1})(x)-D(u-u_{1})(y)+Du_{1}(x)-Du_{1}(y)|}{|x-y|^{\alpha}}\\ &\leqslant\sup\limits_{\xi\in B_{1/4}}|D^{2}(u-u_{1})(\xi)||x-y|^{1-\alpha}+\|u_{1}\|_{C^{1,\alpha}(\mathbb{R}^{n})}\\ &\leqslant c_{n,s,p}\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}.\end{split}

Therefore, we have that uC1,α(B1/2¯)u\in C^{1,\alpha}(\overline{B_{1/2}}) for any α(0,1)\alpha\in(0,1).

We point out that the ball B1/2B_{1/2} is centred at the origin, but we may arbitrarily move it around n\mathbb{R}^{n}. Covering n\mathbb{R}^{n} with these balls, we conclude that uC1,α(n)u\in C^{1,\alpha}(\mathbb{R}^{n}) for any α(0,1)\alpha\in(0,1). This completes the proof of Theorem 1.3. ∎

3.3. C2,αC^{2,\alpha}-regularity of weak solutions

The goal of this section is to establish the C2,αC^{2,\alpha}-regularity result for solutions of problem (1.1), as stated in Theorem 1.4. For this, we combine a suitable truncation argument for the solution uu with the C1,αC^{1,\alpha}-regularity argument.

We point out that Theorem 1.4 can be obtained by appropriately modifying [47, Theorem 1.6]. For the convenience of the reader, we sketch the proof in the following subsections.

To begin with, we introduce some notations. For α(0,1)\alpha\in(0,1), kk\in{\mathbb{N}}, x0B3/4x_{0}\in B_{3/4} and R(0,120)R\in\left(0,\frac{1}{20}\right), we denote the interior norms as follows:

[u]α;BR(x0)\displaystyle[u]_{\alpha;B_{R}(x_{0})} :=\displaystyle:= supx,yBR(x0)|u(x)u(y)||xy|α,\displaystyle\sup\limits_{x,y\in B_{R}(x_{0})}\frac{|u(x)-u(y)|}{|x-y|^{\alpha}},
|u|k;BR(x0)\displaystyle|u|^{\prime}_{k;B_{R}(x_{0})} :=\displaystyle:= j=0kRjDjuL(BR(x0))\displaystyle\sum\limits_{j=0}\limits^{k}R^{j}\|D^{j}u\|_{L^{\infty}(B_{R}(x_{0}))}
and |u|k,α;BR(x0)\displaystyle{\mbox{and }}\quad|u|^{\prime}_{k,\alpha;B_{R}(x_{0})} :=\displaystyle:= |u|k;BR(x0)+Rk+α[Dku]α;BR(x0).\displaystyle|u|^{\prime}_{k;B_{R}(x_{0})}+R^{k+\alpha}[D^{k}u]_{\alpha;B_{R}(x_{0})}.

3.3.1. A mollifier technique and a truncation argument

Let uH1(n)u\in H^{1}(\mathbb{R}^{n}) solve (1.1), and let ηε\eta_{\varepsilon} be a standard mollifier. For every xnx\in\mathbb{R}^{n}, R(0,120)R\in\left(0,\frac{1}{20}\right) and ε(0,R)\varepsilon\in(0,R), we denote by

uε(x):=(ηεu)(x)=|y|εηε(y)u(xy)𝑑y.u_{\varepsilon}(x):=(\eta_{\varepsilon}\ast u)(x)=\int_{|y|\leqslant\varepsilon}\eta_{\varepsilon}(y)u(x-y)\,dy.

Also, we set

gε:=ηε(u+up).g_{\varepsilon}:=\eta_{\varepsilon}\ast(-u+u^{p}).

Then, we have that

(3.23) Δuε+(Δ)suε=gε in B1.-\Delta u_{\varepsilon}+(-\Delta)^{s}u_{\varepsilon}=g_{\varepsilon}\qquad\text{ in }B_{1}.

Moreover, the following regularity estimates on uεu_{\varepsilon} and gεg_{\varepsilon} follow as a direct consequence of their definitions:

Lemma 3.5.

Let uL(n)u\in L^{\infty}(\mathbb{R}^{n}). Then, uεL(n)u_{\varepsilon}\in L^{\infty}(\mathbb{R}^{n}) and

uεL(n)uL(n).\|u_{\varepsilon}\|_{L^{\infty}(\mathbb{R}^{n})}\leqslant\|u\|_{L^{\infty}(\mathbb{R}^{n})}.

If in addition uC1(n)u\in C^{1}(\mathbb{R}^{n}), then, for every yny\in\mathbb{R}^{n},

gεCα(B1(y)¯)cp(uC1(n)+uC1(n)p).\|g_{\varepsilon}\|_{C^{\alpha}(\overline{B_{1}(y)})}\leqslant c_{p}\left(\|u\|_{C^{1}(\mathbb{R}^{n})}+\|u\|_{C^{1}(\mathbb{R}^{n})}^{p}\right).

We now use a cut off argument for uεu_{\varepsilon} to get the a C2,αC^{2,\alpha}-estimate for uεu_{\varepsilon}. Consider a cut off function ϕC0(n)\phi\in C^{\infty}_{0}(\mathbb{R}^{n}) satisfying

ϕ1 in B3/2,supp(ϕ)B2and0ϕ1 in n\phi\equiv 1\text{ in }B_{{3}/{2}},\qquad\text{supp}(\phi)\subset B_{2}\qquad{\mbox{and}}\qquad 0\leqslant\phi\leqslant 1\text{ in }\mathbb{R}^{n}

and let

ϕR(x):=ϕ(xx0R).\phi^{R}(x):=\phi\left(\frac{x-x_{0}}{R}\right).

We point out that

B4R(x0)B1 and supp(ϕR)B2R(x0).B_{4R}(x_{0})\subset B_{1}\quad{\mbox{ and }}\quad\text{supp}(\phi^{R})\subset B_{2R}(x_{0}).

With this notation, one obtains the following result:

Lemma 3.6.

([47, Lemma 5.4]) Let α(0,1)\alpha\in(0,1) and gClocα(B1)g\in C^{\alpha}_{\rm loc}(B_{1}). Let uC2,α(B1¯)L(n)u\in C^{2,\alpha}(\overline{B_{1}})\cap L^{\infty}(\mathbb{R}^{n}) be a solution of

Δu+(Δ)su=g in B1.-\Delta u+(-\Delta)^{s}u=g\qquad\text{ in }B_{1}.

Then, there exists ψCα(BR(x0))\psi\in C^{\alpha}(B_{R}(x_{0})) such that v:=ϕRuv:=\phi^{R}u satisfies

Δv+(Δ)sv=ψ in B1.-\Delta v+(-\Delta)^{s}v=\psi\qquad\text{ in }B_{1}.

In particular,

(3.24) R2|ψ|0,α;BR(x0)Cn,s(R2|g|0,α;BR(x0)+uL(n)),R^{2}|\psi|^{\prime}_{0,\alpha;B_{R}(x_{0})}\leqslant C_{n,s}\left(R^{2}|g|^{\prime}_{0,\alpha;B_{R}(x_{0})}+\|u\|_{L^{\infty}(\mathbb{R}^{n})}\right),

for some positive constant Cn,sC_{n,s}.

As a consequence of Lemma 3.6, setting vε:=ϕRuεv_{\varepsilon}:=\phi^{R}u_{\varepsilon}, we have that there exists ψεCα(BR(x0))\psi_{\varepsilon}\in C^{\alpha}(B_{R}(x_{0})) such that vεv_{\varepsilon} satisfies

Δvε+(Δ)svε=ψε in B1.-\Delta v_{\varepsilon}+(-\Delta)^{s}v_{\varepsilon}=\psi_{\varepsilon}\qquad\text{ in }B_{1}.

In particular, employing (3.24), one finds that

(3.25) R2|ψε|0,α;BR(x0)Cn,s(R2|gε|0,α;BR(x0)+uεL(n)).\begin{split}R^{2}|\psi_{\varepsilon}|^{\prime}_{0,\alpha;B_{R}(x_{0})}&\leqslant C_{n,s}\left(R^{2}|g_{\varepsilon}|^{\prime}_{0,\alpha;B_{R}(x_{0})}+\|u_{\varepsilon}\|_{L^{\infty}(\mathbb{R}^{n})}\right).\end{split}

We also observe that, since vεC0(B2R(x0))v_{\varepsilon}\in C^{\infty}_{0}(B_{2R}(x_{0})), for all δ>0\delta>0, there exists Cδ>0C_{\delta}>0 such that

R2|(Δ)svε|0,α;BR(x0)=R2(Δ)svεL(BR(x0))+R2+α[(Δ)svε]α;BR(x0)Cn,s|uε|2,α0;B2R(x0)δ|uε|2,α;B2R(x0)+CδuεL(B2R(x0)),\begin{split}&R^{2}|(-\Delta)^{s}v_{\varepsilon}|^{\prime}_{0,\alpha;B_{R}(x_{0})}=R^{2}\|(-\Delta)^{s}v_{\varepsilon}\|_{L^{\infty}(B_{R}(x_{0}))}+R^{2+\alpha}[(-\Delta)^{s}v_{\varepsilon}]_{\alpha;B_{R}(x_{0})}\\ &\qquad\leqslant C_{n,s}|u_{\varepsilon}|^{\prime}_{2,\alpha_{0};B_{2R}(x_{0})}\leqslant\delta|u_{\varepsilon}|^{\prime}_{2,\alpha;B_{2R}(x_{0})}+C_{\delta}\|u_{\varepsilon}\|_{L^{\infty}(B_{2R}(x_{0}))},\end{split}

where

α0:={0α<22s,α(1s)α22s.\alpha_{0}:=\begin{cases}0\qquad&\alpha<2-2s,\\ \alpha-(1-s)&\alpha\geqslant 2-2s.\end{cases}

Thus, combining this with [35, Theorem 4.6], one deduces that for all δ>0\delta>0 there exists Cδ>0C_{\delta}>0 such that

|vε|2,α;BR/2(x0)C(vεL(BR(x0))+R2(|ψε|0,α;BR(x0)+|(Δ)svε|0,α;BR(x0)))C(vεL(BR(x0))+R2|ψε|0,α;BR(x0)+δ|uε|2,α;B2R(x0)+CδuεL(B2R(x0))).\begin{split}|v_{\varepsilon}|^{\prime}_{2,\alpha;B_{R/2}(x_{0})}&\leqslant C\left(\|v_{\varepsilon}\|_{L^{\infty}(B_{R}(x_{0}))}+R^{2}\left(|\psi_{\varepsilon}|^{\prime}_{0,\alpha;B_{R}(x_{0})}+|(-\Delta)^{s}v_{\varepsilon}|^{\prime}_{0,\alpha;B_{R}(x_{0})}\right)\right)\\ &\leqslant C\bigg{(}\|v_{\varepsilon}\|_{L^{\infty}(B_{R}(x_{0}))}+R^{2}|\psi_{\varepsilon}|^{\prime}_{0,\alpha;B_{R}(x_{0})}+\delta|u_{\varepsilon}|^{\prime}_{2,\alpha;B_{2R}(x_{0})}+C_{\delta}\|u_{\varepsilon}\|_{L^{\infty}(B_{2R}(x_{0}))}\bigg{)}.\end{split}

Therefore, for every x0B3/4x_{0}\in B_{{3}/{4}}, recalling the definition of vεv_{\varepsilon} and exploting (3.25) and Lemma 3.5, we conclude that for all δ>0\delta>0, there exists CδC_{\delta} such that

(3.26) |uε|2,α;BR/2(x0)=|vε|2,α;BR/2(x0)C(R2|ψε|0,α;BR(x0)+δ|uε|2,α;B2R(x0)+CδuL(B2R(x0)))C(R2|gε|0,α;BR(x0)+uεL(n)+δ|uε|2,α;B2R(x0)+CδuL(B2R(x0)))C(uC1(n)+uC1(n)p+uL(n)+δ|uε|2,α;B2R(x0)+CδuL(B2R(x0)))C(uC1(n)+uC1(n)p+CδuL(n)+δ|uε|2,α;B2R(x0)),\begin{split}&|u_{\varepsilon}|^{\prime}_{2,\alpha;B_{R/2}(x_{0})}=|v_{\varepsilon}|^{\prime}_{2,\alpha;B_{R/2}(x_{0})}\\ &\qquad\leqslant C\left(R^{2}|\psi_{\varepsilon}|^{\prime}_{0,\alpha;B_{R}(x_{0})}+\delta|u_{\varepsilon}|^{\prime}_{2,\alpha;B_{2R}(x_{0})}+C_{\delta}\|u\|_{L^{\infty}(B_{2R}(x_{0}))}\right)\\ &\qquad\leqslant C\left(R^{2}|g_{\varepsilon}|^{\prime}_{0,\alpha;B_{R}(x_{0})}+\|u_{\varepsilon}\|_{L^{\infty}(\mathbb{R}^{n})}+\delta|u_{\varepsilon}|^{\prime}_{2,\alpha;B_{2R}(x_{0})}+C_{\delta}\|u\|_{L^{\infty}(B_{2R}(x_{0}))}\right)\\ &\qquad\leqslant C\left(\|u\|_{C^{1}(\mathbb{R}^{n})}+\|u\|_{C^{1}(\mathbb{R}^{n})}^{p}+\|u\|_{L^{\infty}(\mathbb{R}^{n})}+\delta|u_{\varepsilon}|^{\prime}_{2,\alpha;B_{2R}(x_{0})}+C_{\delta}\|u\|_{L^{\infty}(B_{2R}(x_{0}))}\right)\\ &\qquad\leqslant C\left(\|u\|_{C^{1}(\mathbb{R}^{n})}+\|u\|_{C^{1}(\mathbb{R}^{n})}^{p}+C_{\delta}\|u\|_{L^{\infty}(\mathbb{R}^{n})}+\delta|u_{\varepsilon}|^{\prime}_{2,\alpha;B_{2R}(x_{0})}\right),\end{split}

for some C>0C>0, depending on nn, ss, pp and α\alpha.

3.3.2. Interior C2,αC^{2,\alpha}-regularity

The estimate in (3.26), coupled with the following statement, will allow us to obtain that the C2,αC^{2,\alpha}-norm of uεu_{\varepsilon} is bounded uniformly in some ball with respect to ε\varepsilon and thus use Arzelà-Ascoli Theorem to complete the proof of Theorem 1.4. The technical details go as follows:

Proposition 3.7.

([47, Proposition 4.3]) Let yny\in\mathbb{R}^{n}, d>0d>0 and uC2,α(Bd(y))u\in C^{2,\alpha}(B_{d}(y)). Suppose that, for any δ>0\delta>0, there exists Λδ>0\Lambda_{\delta}>0 such that, for any xBd(y)x\in B_{d}(y) and any r(0,d|xy|]r\in(0,d-|x-y|], we have that

(3.27) |u|2,α;Br/8(x)Λδ+δ|u|2,α;Br/2(x).|u|^{\prime}_{2,\alpha;B_{r/8}(x)}\leqslant\Lambda_{\delta}+\delta|u|^{\prime}_{2,\alpha;B_{r/2}(x)}.

Then, there exist constants δ0\delta_{0}, C>0C>0, depending only on nn, α\alpha and dd, such that

uC2,α(Bd/8(y))CΛδ0.\|u\|_{C^{2,\alpha}(B_{d/8}(y))}\leqslant C\Lambda_{\delta_{0}}.
Proof of Theorem 1.4.

We will use Proposition 3.7 with d:=1/10d:=1/10, so that, for every yB1/2y\in B_{1/2},

Bd(y)B3/4andd/4<120.B_{d}(y)\subset B_{3/4}\quad\text{and}\quad d/4<\frac{1}{20}.

Moreover, we notice that the estimate in (3.26) tells us that formula (3.27) is verified in our setting with uu replaced by uεu_{\varepsilon}, r:=4Rr:=4R and

Λδ:=uC1(n)+uC1(n)p+CδuL(n).\Lambda_{\delta}:=\|u\|_{C^{1}(\mathbb{R}^{n})}+\|u\|_{C^{1}(\mathbb{R}^{n})}^{p}+C_{\delta}\|u\|_{L^{\infty}(\mathbb{R}^{n})}.

Therefore, we are in a position of exploiting Proposition 3.7, thus obtaining that, for every yB1/2y\in B_{1/2},

uεC2,α(B1/80(y)¯)C(uC1(n)+uC1(n)p+CδuL(n)).\|u_{\varepsilon}\|_{C^{2,\alpha}(\overline{B_{1/80}(y)})}\leqslant C\left(\|u\|_{C^{1}(\mathbb{R}^{n})}+\|u\|_{C^{1}(\mathbb{R}^{n})}^{p}+C_{\delta}\|u\|_{L^{\infty}(\mathbb{R}^{n})}\right).

From the Arzelà-Ascoli Theorem, we obtain that uC2,α(B1/80(y)¯)u\in C^{2,\alpha}(\overline{B_{1/80}(y)}), for every yB1/2y\in B_{1/2}, and

uC2,α(B1/80(y)¯)C(uC1(n)+uC1(n)p+uL(n)).\|u\|_{C^{2,\alpha}(\overline{B_{1/80}(y)})}\leqslant C\left(\|u\|_{C^{1}(\mathbb{R}^{n})}+\|u\|_{C^{1}(\mathbb{R}^{n})}^{p}+\|u\|_{L^{\infty}(\mathbb{R}^{n})}\right).

Hence, a covering argument and Theorem 1.3, give that

uC2,α(B1/2¯)C(uC1(n)+uC1(n)p+uL(n))C(uL(n)+uL(n)p+(uL(n)+uL(n)p)p+uL(n)),\begin{split}\|u\|_{C^{2,\alpha}(\overline{B_{1/2}})}&\leqslant C\left(\|u\|_{C^{1}(\mathbb{R}^{n})}+\|u\|_{C^{1}(\mathbb{R}^{n})}^{p}+\|u\|_{L^{\infty}(\mathbb{R}^{n})}\right)\\ &\leqslant C\left(\|u\|_{L^{\infty}(\mathbb{R}^{n})}+\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}+\left(\|u\|_{L^{\infty}(\mathbb{R}^{n})}+\|u\|^{p}_{L^{\infty}(\mathbb{R}^{n})}\right)^{p}+\|u\|_{L^{\infty}(\mathbb{R}^{n})}\right),\end{split}

where the constant C>0C>0 depends on nn, ss, α\alpha and pp.

The ball B1B_{1} is centered at the origin, but we may arbitrarily move it around n\mathbb{R}^{n}. Covering n\mathbb{R}^{n} with these balls, we obtain the desired result. ∎

4. Qualitative properties of positive solutions

In this section, we are concerned with the positivity, the decay at infinity and the radial symmetry of classical solutions of (1.1), as stated in Theorem 1.5.

4.1. Power-type decay of classical positive solutions

In this part, we shall apply the Maximum Principle and comparison arguments to obtain the decay at infinity of classical positive solutions of (1.1).

4.1.1. Existence of classical positive solutions

We devote this part to establish the existence of classical positive solutions.

Theorem 4.1.

Let n2n\geqslant 2 and s(0,1)s\in(0,1).

Then, problem (1.1) has a classical solution, which satisfies u>0u>0 in n\mathbb{R}^{n}. Moreover,

(4.1) lim|x|+u(x)=0.\lim_{|x|\to+\infty}u(x)=0.
Proof.

From Theorems 1.1 and 1.4, we have that there exists a classical nonnegative solution of (1.1).

Hence, we can now focus on the proving that u>0u>0 in n\mathbb{R}^{n}. For this, we argue by contradiction and we assume that there exists a global minimum point x0nx_{0}\in\mathbb{R}^{n} at which u(x0)=0u(x_{0})=0. Accordingly, we have that Δu(x0)0\Delta u(x_{0})\geqslant 0 and (Δ)su(x0)<0(-\Delta)^{s}u(x_{0})<0. As a result, we deduce from (1.1) that

0=up(x0)=Δu(x0)+(Δ)su(x0)+u(x0)<0,0=u^{p}(x_{0})=-\Delta u(x_{0})+(-\Delta)^{s}u(x_{0})+u(x_{0})<0,

which is a contradiction.

Finally, from Theorem 1.3, we also have (4.1). ∎

4.1.2. Power-type decay of classical positive solutions

In this part, we exploit suitable barriers constructed using the Bessel kernel 𝒦\mathcal{K} to establish the following result:

Theorem 4.2.

Let n2n\geqslant 2 and s(0,1)s\in(0,1). Let uu be a positive classical solution of (1.1).

Then, there exist constants 0<C1C20<C_{1}\leqslant C_{2} such that, for every |x|1|x|\geqslant 1,

C1|x|n+2su(x)C2|x|n+2s.\frac{C_{1}}{|x|^{n+2s}}\leqslant u(x)\leqslant\frac{C_{2}}{|x|^{n+2s}}.

To prove this result, we start with the following two lemmata, which construct suitable subsolutions and supersolutions. We start with the analysis of the subsolution:

Lemma 4.3.

There exists a function ωC1,α(n)\omega\in C^{1,\alpha}(\mathbb{R}^{n}) satisfying

(4.2) {Δω+(Δ)sω+ω=0in nB1,ω>0in n,lim|x|+ω(x)=0\begin{cases}-\Delta\omega+(-\Delta)^{s}\omega+\omega=0\quad\text{in }{\mathbb{R}}^{n}\setminus B_{1},\\ \omega>0\quad\text{in }{\mathbb{R}}^{n},\\ \displaystyle\lim_{|x|\to+\infty}\omega(x)=0\end{cases}

and, for every |x|>1|x|>1,

(4.3) ω(x)c1|x|n+2s,\omega(x)\geqslant\frac{c_{1}}{|x|^{n+2s}},

for some constant c1>0c_{1}>0.

Proof.

We consider the function ω:=𝒦χB1/2\omega:=\mathcal{K}\ast\chi_{B_{1/2}}, where χB1/2\chi_{B_{1/2}} is the characteristic function of the ball B1/2B_{1/2}.

It is immediate to check that ω>0\omega>0 solves

Δω+(Δ)sω+ω=χB1/2in n,-\Delta\omega+(-\Delta)^{s}\omega+\omega=\chi_{B_{1/2}}\quad\text{in }{\mathbb{R}}^{n},

and therefore the equation in (4.2) is satisfied.

Also, in light of the smoothness of 𝒦\mathcal{K} (see e.g. (c) of Theorem 3.2), we have that ωC2(nB1)\omega\in C^{2}(\mathbb{R}^{n}\setminus B_{1}).

In particular, since χB1/2Lq(n)\chi_{B_{1/2}}\in L^{q}(\mathbb{R}^{n}) for all q1q\geqslant 1, we have that ωW2,q(n)\omega\in W^{2,q}(\mathbb{R}^{n}) for all q1q\geqslant 1 (by Lemma 3.3), which implies that ωC1,α(n)\omega\in C^{1,\alpha}(\mathbb{R}^{n}) for any α(0,1)\alpha\in(0,1). As a result, the limit in (4.2) is also satisfied.

We now check (4.3). Using the lower bound on 𝒦\mathcal{K} (see e.g. (b) of Theorem 3.2), for any |x|>1|x|>1, one has that

ω(x)=B1/2𝒦(xy)𝑑ycB1/2dy|xy|n+2scB1/2dy(|x|+|y|)n+2sc|x|n+2s,\omega(x)=\int_{B_{1/2}}\mathcal{K}(x-y)\,dy\geqslant c\int_{B_{1/2}}\frac{dy}{|x-y|^{n+2s}}\geqslant c\int_{B_{1/2}}\frac{dy}{(|x|+|y|)^{n+2s}}\geqslant\frac{c}{|x|^{n+2s}},

up to renaming c>0c>0. As a consequence of this, we obtain (4.3), as desired. ∎

The supersolution is constructed exploting Theorem 3.2 with a parameter a>0a>0 in place of 11:

Lemma 4.4.

Let n1n\geqslant 1 and s(0,1)s\in(0,1). Let a>0a>0 and

𝒦a:=1(1a+|ξ|2+|ξ|2s).{\mathcal{K}_{a}}:=\mathcal{F}^{-1}\left(\frac{1}{a+|\xi|^{2}+|\xi|^{2s}}\right).

Then,

  • (a)

    𝒦a\mathcal{K}_{a} is positive, radially symmetric, smooth in n{0}\mathbb{R}^{n}\setminus\left\{0\right\} and nonincreasing with respect to r=|x|r=|x|.

  • (b)

    There exist positive constants C1C_{1} and C2C_{2} such that, if |x|1|x|\geqslant 1,

    C1|x|n+2s𝒦a(x)1C1|x|n+2s\frac{C_{1}}{|x|^{n+2s}}\leqslant\mathcal{K}_{a}(x)\leqslant\frac{1}{C_{1}|x|^{n+2s}}

    and, if |x|1|x|\leqslant 1,

    C2|x|n2𝒦a(x)1C2{|x|2n if n3,1+|ln|x|| if n=2,1+|x| if n=1.\frac{C_{2}}{|x|^{n-2}}\leqslant\mathcal{K}_{a}(x)\leqslant\frac{1}{C_{2}}\begin{cases}|x|^{2-n}\quad&{\mbox{ if }}n\geqslant 3,\\ 1+|\ln|x||&{\mbox{ if }}n=2,\\ 1+|x|&{\mbox{ if }}n=1.\end{cases}
  • (c)

    There exists a positive constant CC such that, if |x|1|x|\geqslant 1,

    |𝒦a(x)|C|x|n+2s+1and|D2𝒦a(x)|C|x|n+2s+2.|\nabla\mathcal{K}_{a}(x)|\leqslant\frac{C}{|x|^{n+2s+1}}\qquad\mbox{and}\qquad|D^{2}\mathcal{K}_{a}(x)|\leqslant\frac{C}{|x|^{n+2s+2}}.
  • (d)

    If n3n\geqslant 3, then 𝒦aLq(n)\mathcal{K}_{a}\in L^{q}(\mathbb{R}^{n}) for all q[1,nn2)q\in[1,\frac{n}{n-2}). If n=1n=1, 22, then 𝒦aLq(n)\mathcal{K}_{a}\in L^{q}(\mathbb{R}^{n}) for all q[1,+)q\in[1,+\infty).

  • (e)

    There exists a positive constant CC such that, if |x|1|x|\geqslant 1,

    𝒦a(x)C|x|n+2s.\mathcal{K}_{a}(x)\geqslant\frac{C}{|x|^{n+2s}}.

Exploting the kernel given by Lemma 4.4, we are able to construct the supersolution as follows:

Lemma 4.5.

There exists vC1,α(n)v\in C^{1,\alpha}(\mathbb{R}^{n}) satisfying

{Δv+(Δ)sv+12v=0in nB1,v>0in n,lim|x|+v(x)=0.\begin{cases}-\Delta v+(-\Delta)^{s}v+\frac{1}{2}v=0\qquad\text{in }{\mathbb{R}}^{n}\setminus B_{1},\\ v>0\quad\text{in }{\mathbb{R}}^{n},\\ \displaystyle\lim_{|x|\to+\infty}v(x)=0.\end{cases}

Also, when |x|>1|x|>1 we have that

0<v(x)c2|x|n+2s,0<v(x)\leqslant\frac{c_{2}}{|x|^{n+2s}},

for some constant c2>0c_{2}>0.

Proof.

We consider the function v:=𝒦1/2χB1/2v:=\mathcal{K}_{1/2}\ast\chi_{B_{1/2}}, where 𝒦1/2\mathcal{K}_{1/2} is defined in Lemma 4.4 with a:=1/2a:=1/2.

Thus, v>0v>0 and it satisfies

Δv+(Δ)sv+12v=χB1/2in n.-\Delta v+(-\Delta)^{s}v+\frac{1}{2}v=\chi_{B_{1/2}}\qquad\text{in }\mathbb{R}^{n}.

Moreover, from Lemma 4.4, when |x|>1|x|>1 we have that

v(x)=B1/2𝒦1/2(xy)𝑑ycB1/2dy|xy|n+2scB1/2dy(|x|1/2)n+2sc|x|n+2s,v(x)=\int_{B_{1/2}}\mathcal{K}_{1/2}(x-y)\,dy\leqslant c\int_{B_{1/2}}\frac{dy}{|x-y|^{n+2s}}\leqslant c\int_{B_{1/2}}\frac{dy}{(|x|-1/2)^{n+2s}}\leqslant\frac{c}{|x|^{n+2s}},

for some positive constant cc, that may change from step to step.

Furthermore, using the smoothness of 𝒦1/2\mathcal{K}_{1/2} (see e.g. (a) of Lemma 4.4), we have that vC2(nB1)v\in C^{2}(\mathbb{R}^{n}\setminus B_{1}). We claim that vC1,α(n)v\in C^{1,\alpha}(\mathbb{R}^{n}) for any α(0,1)\alpha\in(0,1).

Indeed, we point out that χB1/2Lq(n)\chi_{B_{1/2}}\in L^{q}(\mathbb{R}^{n}) for all q1q\geqslant 1 and also we see that

(12+|ξ|2+|ξ|2s)(v)(ξ)=(χB1/2)(ξ).\left(\frac{1}{2}+|\xi|^{2}+|\xi|^{2s}\right)\mathcal{F}(v)(\xi)=\mathcal{F}({\chi_{B_{1/2}}})(\xi).

Then,

(1+|ξ|2+|ξ|2s)(v)(ξ)=1+|ξ|2+|ξ|2s(1/2+|ξ|2+|ξ|2s)(χB1/2)(ξ)=g^,\Big{(}1+|\xi|^{2}+|\xi|^{2s}\Big{)}\mathcal{F}(v)(\xi)=\frac{1+|\xi|^{2}+|\xi|^{2s}}{(1/2+|\xi|^{2}+|\xi|^{2s})}\mathcal{F}({\chi_{B_{1/2}}})(\xi)=\hat{g},

where g:=(δ+12𝒦1/2)χB1/2g:=\left(\delta+\frac{1}{2}\mathcal{K}_{1/2}\right)\ast{\chi_{B_{1/2}}}.

Since 𝒦1/2L1(n)\mathcal{K}_{1/2}\in L^{1}(\mathbb{R}^{n}), we have that gLq(n)g\in L^{q}(\mathbb{R}^{n}) for all q1q\geqslant 1 and

Δv+(Δ)sv+v=gin n.-\Delta v+(-\Delta)^{s}v+v=g\qquad\text{in }\mathbb{R}^{n}.

These facts and Lemma 3.3 imply that vW2,q(n)v\in W^{2,q}(\mathbb{R}^{n}) for all q1q\geqslant 1. By the Sobolev Embedding, we obtain that vC1,α(n)v\in C^{1,\alpha}(\mathbb{R}^{n}) for any α(0,1)\alpha\in(0,1) and also

lim|x|+v(x)=0.\lim_{|x|\to+\infty}v(x)=0.\qed
Proof of Theorem 4.2.

We consider the function ω\omega given by Lemma 4.3. Utilizing the positivity and continuity of uu and ω\omega in n\mathbb{R}^{n}, we can find a constant β>0\beta>0 such that h:=uβω>0h:=u-\beta\omega>0 in B1¯\overline{B_{1}}.

Furthermore, from Theorem 4.1 and Lemma 4.3, we see that

{Δh+(Δ)shh in nB1,lim|x|+h(x)=0.\begin{cases}-\Delta h+(-\Delta)^{s}h\geqslant-h\quad\text{ in }{\mathbb{R}}^{n}\setminus B_{1},\\ \displaystyle\lim_{|x|\to+\infty}h(x)=0.\end{cases}

We claim that

(4.4) h(x)0 for all |x|>1.h(x)\geqslant 0\qquad{\mbox{ for all~{}$|x|>1$.}}

Indeed, for the sake of contradiction, we assume that there exists a global strictly negative minimum point x0nB1x_{0}\in\mathbb{R}^{n}\setminus B_{1}. Since hC2(nB1)h\in C^{2}(\mathbb{R}^{n}\setminus B_{1}), we have Δh(x0)0\Delta h(x_{0})\geqslant 0 and (Δ)sh(x0)0(-\Delta)^{s}h(x_{0})\leqslant 0, and thus

0<h(x0)Δh(x0)+(Δ)sh(x0)0.0<-h(x_{0})\leqslant-\Delta h(x_{0})+(-\Delta)^{s}h(x_{0})\leqslant 0.

This contradiction implies (4.4).

From (4.4), we have that, for every |x|>1|x|>1,

u(x)βω(x)c1|x|n+2s.u(x)\geqslant\beta\omega(x)\geqslant\frac{c_{1}}{|x|^{n+2s}}.

This establishes the bound from below in Theorem 4.2.

We now focus on the bound from above. Owing to (4.1), we can find some R>1R>1 such that

Δu+(Δ)su+12u0in nBR.-\Delta u+(-\Delta)^{s}u+\frac{1}{2}u\leqslant 0\qquad\text{in }\mathbb{R}^{n}\setminus B_{R}.

In this case, we make use of the function vv given by Lemma 4.5. From the positivity and continuity of uu and vv in n\mathbb{R}^{n}, there exists γ>0\gamma>0 such that g:=vγu>0g:=v-\gamma u>0 in BR¯\overline{B_{R}}.

In view of Theorem 4.1 and Lemma 4.5, one has that

{Δg+(Δ)sg12g in nBR,lim|x|+g(x)=0.\begin{cases}-\Delta g+(-\Delta)^{s}g\geqslant-\frac{1}{2}g\quad\text{ in }{\mathbb{R}}^{n}\setminus B_{R},\\ \displaystyle\lim_{|x|\to+\infty}g(x)=0.\end{cases}

We claim that

(4.5) g(x)0 for all |x|>R.g(x)\geqslant 0\qquad{\mbox{ for all~{}$|x|>R$.}}

For this, by contradiction, we assume that there exists a global strictly negative minimum point x1nBRx_{1}\in\mathbb{R}^{n}\setminus B_{R}. Since gC2(nB1)g\in C^{2}(\mathbb{R}^{n}\setminus B_{1}), we have that Δg(x1)0\Delta g(x_{1})\geqslant 0 and (Δ)sg(x1)0(-\Delta)^{s}g(x_{1})\leqslant 0, and hence

0<12g(x1)Δg(x1)+(Δ)sg(x1)0.0<-\frac{1}{2}g(x_{1})\leqslant-\Delta g(x_{1})+(-\Delta)^{s}g(x_{1})\leqslant 0.

This contradiction gives (4.5).

In turn, (4.5) implies that, for all |x|>R|x|>R,

u(x)γv(x)c2|x|n+2s.u(x)\leqslant\gamma v(x)\leqslant\frac{c_{2}}{|x|^{n+2s}}.

Thus, the continuity of uu allows us to complete the proof of the bound from above in Theorem 4.2. ∎

4.2. Radial symmetry of positive solutions

Our aim is now to prove the radial symmetry of the positive solution found in Theorem 4.1. The main statement of this section is the following:

Theorem 4.6.

Let n2n\geqslant 2 and s(0,1)s\in(0,1).

Then, all classical positive solutions of (1.1) are radially symmetric.

To establish Theorem 4.6, we will exploit the moving planes method. To begin with, we recall some notation. For any λ\lambda\in\mathbb{R}, we set

Σλ\displaystyle\Sigma_{\lambda} :=\displaystyle:= {xn s.t. x1>λ}\displaystyle\left\{x\in\mathbb{R}^{n}\;{\mbox{ s.t. }}\;x_{1}>\lambda\right\}
and Tλ\displaystyle{\mbox{and }}\qquad T_{\lambda} :=\displaystyle:= {xn s.t. x1=λ}.\displaystyle\left\{x\in\mathbb{R}^{n}\;{\mbox{ s.t. }}\;x_{1}=\lambda\right\}.

Moreover, for any x=(x1,x2,,xn)nx=(x_{1},x_{2},\dots,x_{n})\in\mathbb{R}^{n}, we set xλ:=(2λx1,x2,,xn)x^{\lambda}:=\left(2\lambda-x_{1},x_{2},\dots,x_{n}\right) and uλ(x):=u(xλ)u_{\lambda}(x):=u(x^{\lambda}).

With this notation, we first establish the following:

Lemma 4.7.

If uuλu\geqslant u_{\lambda} and uuλu\not\equiv u_{\lambda} in Σλ\Sigma_{\lambda}, then u>uλu>u_{\lambda} in Σλ\Sigma_{\lambda}.

Proof.

Suppose by contradiction that there exists x0Σλx_{0}\in\Sigma_{\lambda} such that u(x0)=uλ(x0)u(x_{0})=u_{\lambda}(x_{0}). In the light of Theorem 1.4, we know that uC2(n)u\in C^{2}(\mathbb{R}^{n}). We thus have Δ(u(x0)uλ(x0))0-\Delta(u(x_{0})-u_{\lambda}(x_{0}))\leqslant 0, and therefore

0=Δ(u(x0)uλ(x0))+(Δ)s(u(x0)uλ(x0))+(u(x0)uλ(x0))=n(u(y)uλ(y))|x0y|n+2s𝑑y=Σλ(u(y)uλ(y))|x0y|n+2s𝑑y+nΣλ(u(y)uλ(y))|x0y|n+2s𝑑yΣλ(u(y)uλ(y))|x0y|n+2s𝑑y+Σλ(uλ(y)u(y))|x0yλ|n+2s𝑑y=Σλ(u(y)uλ(y))(1|x0yλ|n+2s1|x0y|n+2s)𝑑y0,\begin{split}0&=-\Delta(u(x_{0})-u_{\lambda}(x_{0}))+(-\Delta)^{s}(u(x_{0})-u_{\lambda}(x_{0}))+(u(x_{0})-u_{\lambda}(x_{0}))\\ &=\int_{\mathbb{R}^{n}}\frac{-(u(y)-u_{\lambda}(y))}{|x_{0}-y|^{n+2s}}\,dy\\ &=\int_{\Sigma_{\lambda}}\frac{-(u(y)-u_{\lambda}(y))}{|x_{0}-y|^{n+2s}}\,dy+\int_{\mathbb{R}^{n}\setminus\Sigma_{\lambda}}\frac{-(u(y)-u_{\lambda}(y))}{|x_{0}-y|^{n+2s}}\,dy\\ &\leqslant\int_{\Sigma_{\lambda}}\frac{-(u(y)-u_{\lambda}(y))}{|x_{0}-y|^{n+2s}}\,dy+\int_{\Sigma_{\lambda}}\frac{-(u_{\lambda}(y)-u(y))}{|x_{0}-y^{\lambda}|^{n+2s}}\,dy\\ &=\int_{\Sigma_{\lambda}}\big{(}u(y)-u_{\lambda}(y)\big{)}\left(\frac{1}{|x_{0}-y^{\lambda}|^{n+2s}}-\frac{1}{|x_{0}-y|^{n+2s}}\right)\,dy\\ &\leqslant 0,\end{split}

which implies that uuλu\equiv u_{\lambda} in Σλ\Sigma_{\lambda}. This is in contradiction with the assumption and therefore the desired result is established. ∎

Now, we define

λ0:=sup{λ s.t. u(x)>uλ(x) for all xΣλ}.\lambda_{0}:=\sup\Big{\{}\lambda\in\mathbb{R}\;{\mbox{ s.t. }}\;u(x)>u_{\lambda}(x)\text{ for all }x\in\Sigma_{\lambda}\Big{\}}.
Lemma 4.8.

We have that λ0>\lambda_{0}>-\infty.

To check this, we will need the following estimate:

Lemma 4.9.

Let q>r>n2q>r>\frac{n}{2}. Then, for any measurable set Ωn\Omega\subset\mathbb{R}^{n}, there exists c>0c>0, depending on nn and ss, such that, for all measurable functions g:Ωg:\Omega\to\mathbb{R},

Ω𝒦(y)g(y)dyLq(Ω)cgLr(Ω).\left\|\int_{\Omega}\mathcal{K}(\cdot-y)g(y)\,dy\right\|_{L^{q}(\Omega)}\leqslant c\,\|g\|_{L^{r}(\Omega)}.

The proof of Lemma 4.9 follows the same line as the one of [30, Lemma 5.2]. We provide here the details for the facility of the reader, since some exponents need to be adjusted to our setting.

Proof of Lemma 4.9.

Notice that if gLr(Ω)\|g\|_{L^{r}(\Omega)} is unbounded, the claim is obviously true. Hence, we can suppose that gLr(Ω)<+\|g\|_{L^{r}(\Omega)}<+\infty.

From (d) of Theorem 3.2 we know that if n3n\geqslant 3, then 𝒦Lγ(n)\mathcal{K}\in L^{\gamma}(\mathbb{R}^{n}) for γ[1,nn2)\gamma\in[1,\frac{n}{n-2}) and if n=2n=2, then 𝒦Lγ(n)\mathcal{K}\in L^{\gamma}(\mathbb{R}^{n}) for γ[1,+)\gamma\in[1,+\infty).

We point out that, since r>n/2r>n/2, setting γ:=rr1\gamma:=\frac{r}{r-1}, we see that γ(1,nn2)\gamma\in\left(1,\frac{n}{n-2}\right) when n3n\geqslant 3. Thus, using the Hölder inequality with exponents γ\gamma and rr, we have that, for every xΩx\in\Omega,

|Ω𝒦(xy)g(y)𝑑y|𝒦Lγ(n)gLr(Ω)CgLr(Ω).\left|\int_{\Omega}\mathcal{K}(x-y)g(y)\,dy\right|\leqslant\|\mathcal{K}\|_{L^{\gamma}(\mathbb{R}^{n})}\|g\|_{L^{r}(\Omega)}\leqslant C\|g\|_{L^{r}(\Omega)}.

In addition, by Young’s convolution inequality,

Ω𝒦(y)g(y)dyLr(Ω)𝒦L1(n)gLr(Ω)CgLr(Ω).\left\|\int_{\Omega}\mathcal{K}(\cdot-y)g(y)\,dy\right\|_{L^{r}(\Omega)}\leqslant\|\mathcal{K}\|_{L^{1}(\mathbb{R}^{n})}\|g\|_{L^{r}(\Omega)}\leqslant C\|g\|_{L^{r}(\Omega)}.

As a result, we conclude that

Ω𝒦(y)g(y)dyLq(Ω)=(Ω(Ω𝒦(xy)g(y)𝑑y)qr(Ω𝒦(xy)g(y)𝑑y)r𝑑x)1/qΩ𝒦(xy)g(y)𝑑yL(Ω)1r/qΩ𝒦(xy)g(y)𝑑yLr(Ω)r/qCgLr(Ω),\begin{split}&\left\|\int_{\Omega}\mathcal{K}(\cdot-y)g(y)\,dy\right\|_{L^{q}(\Omega)}\\ =\;&\left(\int_{\Omega}\left(\int_{\Omega}\mathcal{K}(x-y)g(y)\,dy\right)^{q-r}\left(\int_{\Omega}\mathcal{K}(x-y)g(y)\,dy\right)^{r}\,dx\right)^{1/q}\\ \leqslant\;&\left\|\int_{\Omega}\mathcal{K}(x-y)g(y)\,dy\right\|^{1-r/q}_{L^{\infty}(\Omega)}\,\left\|\int_{\Omega}\mathcal{K}(x-y)g(y)\,dy\right\|^{r/q}_{L^{r}(\Omega)}\\ \leqslant\;&C\|g\|_{L^{r}(\Omega)},\end{split}

as desired. ∎

We will also need the following intermediate estimate:

Lemma 4.10.

Let q>max{n,2/(p1)}q>\max\{n,2/(p-1)\}. Let

(4.6) Σλ:={xΣλ s.t. u(x)<uλ(x)}.\Sigma_{\lambda}^{-}:=\left\{x\in\Sigma_{\lambda}\;{\mbox{ s.t. }}\;u(x)<u_{\lambda}(x)\right\}.

Then,

(4.7) uλuLq(Σλ)cuλLq(p1)(Σλ)p1uλuLq(Σλ).\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}\leqslant c\|u_{\lambda}\|^{p-1}_{L^{q(p-1)}(\Sigma_{\lambda}^{-})}\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}.
Proof.

Owing to the fact that 𝒦\mathcal{K} is radial symmetric (recall (a) of Theorem 3.2), we deduce that

u(x)\displaystyle u(x) =\displaystyle= n𝒦(xy)up(y)𝑑y=Σλ𝒦(xy)up(y)𝑑y+Σλ𝒦(xλy)uλp(y)𝑑y\displaystyle\int_{\mathbb{R}^{n}}\mathcal{K}(x-y)u^{p}(y)\,dy=\int_{\Sigma_{\lambda}}\mathcal{K}(x-y)u^{p}(y)\,dy+\int_{\Sigma_{\lambda}}\mathcal{K}(x^{\lambda}-y)u^{p}_{\lambda}(y)\,dy
and uλ(x)\displaystyle{\mbox{and }}\qquad u_{\lambda}(x) =\displaystyle= n𝒦(xy)uλp(y)𝑑y=Σλ𝒦(xy)uλp(y)𝑑y+Σλ𝒦(xλy)up(y)𝑑y.\displaystyle\int_{\mathbb{R}^{n}}\mathcal{K}(x-y)u_{\lambda}^{p}(y)\,dy=\int_{\Sigma_{\lambda}}\mathcal{K}(x-y)u^{p}_{\lambda}(y)\,dy+\int_{\Sigma_{\lambda}}\mathcal{K}(x^{\lambda}-y)u^{p}(y)\,dy.

Moreover, we point out that |yxλ|>|yx||y-x^{\lambda}|>|y-x| for every xx, yΣλy\in\Sigma_{\lambda}. Collecting these pieces of information and recalling that 𝒦\mathcal{K} is nonincreasing with respect to r=|x|r=|x|, we conclude that, for every xΣλx\in\Sigma_{\lambda},

uλ(x)u(x)=Σλ(𝒦(xy)𝒦(xλy))(uλp(y)up(y))𝑑yΣλ(𝒦(xy)𝒦(xλy))(uλp(y)up(y))𝑑yΣλ𝒦(xy)(uλp(y)up(y))𝑑y.\begin{split}u_{\lambda}(x)-u(x)&=\int_{\Sigma_{\lambda}}\left(\mathcal{K}(x-y)-\mathcal{K}(x^{\lambda}-y)\right)\left(u^{p}_{\lambda}(y)-u^{p}(y)\right)\,dy\\ &\leqslant\int_{\Sigma^{-}_{\lambda}}\left(\mathcal{K}(x-y)-\mathcal{K}(x^{\lambda}-y)\right)\left(u^{p}_{\lambda}(y)-u^{p}(y)\right)\,dy\\ &\leqslant\int_{\Sigma^{-}_{\lambda}}\mathcal{K}(x-y)\left(u^{p}_{\lambda}(y)-u^{p}(y)\right)\,dy.\end{split}

Therefore, by the positivity of 𝒦\mathcal{K} and the Mean Value Theorem, we derive that

|uλ(x)u(x)|Σλ𝒦(xy)|uλp(y)up(y)|𝑑ypΣλ𝒦(xy)uλp1(y)|uλ(y)u(y)|𝑑y.|u_{\lambda}(x)-u(x)|\leqslant\int_{\Sigma^{-}_{\lambda}}\mathcal{K}(x-y)\left|u^{p}_{\lambda}(y)-u^{p}(y)\right|\,dy\leqslant p\int_{\Sigma^{-}_{\lambda}}\mathcal{K}(x-y)u^{p-1}_{\lambda}(y)\left|u_{\lambda}(y)-u(y)\right|\,dy.

Since q>max{n,2/(p1)}q>\max\{n,2/(p-1)\}, we can use Lemma 4.9 with r:=q/2r:=q/2, Ω:=Σλ\Omega:=\Sigma_{\lambda}^{-} and g:=uλp1(uλu)g:=u^{p-1}_{\lambda}(u_{\lambda}-u) and we find that

uλuLq(Σλ)pΣλ𝒦(y)uλp1(y)|uλ(y)u(y)|dyLq(Σλ)cuλp1(uλu)Lq/2(Σλ),\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}\leqslant p\left\|\int_{\Sigma^{-}_{\lambda}}\mathcal{K}(\cdot-y)u^{p-1}_{\lambda}(y)\left|u_{\lambda}(y)-u(y)\right|\,dy\right\|_{L^{q}(\Sigma_{\lambda}^{-})}\leqslant c\left\|u^{p-1}_{\lambda}(u_{\lambda}-u)\right\|_{L^{q/2}(\Sigma_{\lambda}^{-})},

for some c>0c>0 depending on nn, ss and pp.

From this and the Hölder inequality, we obtain (4.7).

We point out that, for all m[2,+)m\in[2,+\infty),

uLm(n)m=n|u(x)|m𝑑xuL(n)m2n|u(x)|2𝑑x<+,\displaystyle\|u\|_{L^{m}(\mathbb{R}^{n})}^{m}=\int_{\mathbb{R}^{n}}|u(x)|^{m}\,dx\leqslant\|u\|_{L^{\infty}(\mathbb{R}^{n})}^{m-2}\int_{\mathbb{R}^{n}}|u(x)|^{2}\,dx<+\infty,

thanks to Theorem 1.2. This gives that the norms in (4.7) are finite. ∎

Proof of Lemma 4.8.

We recall the definition of Σλ\Sigma_{\lambda}^{-} in (4.6) and we observe that, for every m[2,+)m\in[2,+\infty),

limλuλLm(Σλ)m=limλΣλ|uλ(x)|m𝑑x\displaystyle\lim_{\lambda\to-\infty}\|u_{\lambda}\|^{m}_{L^{m}(\Sigma_{\lambda}^{-})}=\lim_{\lambda\to-\infty}\int_{\Sigma_{\lambda}^{-}}|u_{\lambda}(x)|^{m}\,dx
=limλ{x1>λ}{u(xλ)>u(x)}|u(xλ)|m𝑑xlimλ{y1<λ}|u(y)|m𝑑y=0.\displaystyle\qquad=\lim_{\lambda\to-\infty}\int_{\{x_{1}>\lambda\}\cap\{u(x^{\lambda})>u(x)\}}|u(x^{\lambda})|^{m}\,dx\leqslant\lim_{\lambda\to-\infty}\int_{\{y_{1}<\lambda\}}|u(y)|^{m}\,dy=0.

Thus, there exists λ\lambda^{-}\in\mathbb{R} sufficiently negative such that cuλLq(p1)(Σλ)p11/2c\|u_{\lambda}\|^{p-1}_{L^{q(p-1)}(\Sigma_{\lambda}^{-})}\leqslant 1/2 for all λλ\lambda\leqslant\lambda^{-}. Therefore, using this information into (4.7), we find that, for all λλ\lambda\leqslant\lambda^{-},

uλuLq(Σλ)cuλLq(p1)(Σλ)p1uλuLq(Σλ)12uλuLq(Σλ).\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}\leqslant c\|u_{\lambda}\|^{p-1}_{L^{q(p-1)}(\Sigma_{\lambda}^{-})}\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}\leqslant\frac{1}{2}\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}.

This implies that, for all λλ\lambda\leqslant\lambda^{-}, we have that |Σλ|=0|\Sigma_{\lambda}^{-}|=0, and therefore, for all xΣλx\in\Sigma_{\lambda}, we see that u(x)uλ(x)u(x)\geqslant u_{\lambda}(x).

Now, from Lemma 4.7 it follows that, for all λλ\lambda\leqslant\lambda^{-},

(4.8) either u>uλu>u_{\lambda} in Σλ\Sigma_{\lambda} or uuλu\equiv u_{\lambda} in n\mathbb{R}^{n}.

We claim that

(4.9) if uuλ1uλ2u\equiv u_{\lambda_{1}}\equiv u_{\lambda_{2}} in n\mathbb{R}^{n} for some λ1\lambda_{1}, λ2\lambda_{2}\in\mathbb{R}, then λ1=λ2\lambda_{1}=\lambda_{2}.

Indeed, suppose that λ2>λ1\lambda_{2}>\lambda_{1} and let T:=2(λ2λ1)T:=2(\lambda_{2}-\lambda_{1}). Then, for all xnx\in\mathbb{R}^{n},

u(x+Te1)=u((x+Te1)λ2)=u(2λ2e1(x+Te1))=u(2λ1e1x)=u(x).u(x+Te_{1})=u((x+Te_{1})^{\lambda_{2}})=u(2\lambda_{2}e_{1}-(x+Te_{1}))=u(2\lambda_{1}e_{1}-x)=u(x).

That is, uu is periodic in the first coordinate of period TT, violating the fact that uu is non-trivial and with finite L2L^{2}-norm. This establishes (4.9).

From (4.8) and (4.9), we thereby conclude that u>uλu>u_{\lambda} in Σλ\Sigma_{\lambda}, for all λλ\lambda\leqslant\lambda^{-} (up to taking λ\lambda^{-} smaller if needed). This implies that λ0λ\lambda_{0}\geqslant\lambda^{-}, as desired. ∎

Lemma 4.11.

We have that λ0<+\lambda_{0}<+\infty.

Proof.

Suppose by contradiction that there exists a sequence λk+\lambda_{k}\to+\infty such that u(x)>uλk(x)u(x)>u_{\lambda_{k}}(x) for all xΣλkx\in\Sigma_{\lambda_{k}}. Then, thanks to (4.1),

0=limk+u(2λke1)limk+u((2λke1)λk)=u(0).0=\lim_{k\to+\infty}u(2\lambda_{k}e_{1})\geqslant\lim_{k\to+\infty}u((2\lambda_{k}e_{1})^{\lambda_{k}})=u(0).

This is a contradiction with the fact that u>0u>0 in n\mathbb{R}^{n}, in light of Theorem 4.1. ∎

In light of Lemma 4.8, we have that

{λ s.t. u(x)>uλ(x) for all xΣλ}\Big{\{}\lambda\in\mathbb{R}\;{\mbox{ s.t. }}\;u(x)>u_{\lambda}(x)\text{ for all }x\in\Sigma_{\lambda}\Big{\}}\neq\varnothing

and moreover

(4.10) u(x)uλ0(x)u(x)\geqslant u_{\lambda_{0}}(x) for all xΣλ0x\in\Sigma_{\lambda_{0}}.

With this preliminary work, we can prove the following:

Lemma 4.12.

We have that uλ0(x)=u(x)u_{\lambda_{0}}(x)=u(x) for all xΣλ0x\in\Sigma_{\lambda_{0}}.

Proof.

Suppose by contradiction that uuλ0u\not\equiv u_{\lambda_{0}} in Σλ0\Sigma_{\lambda_{0}}. Then, thanks to (4.10) and Lemma 4.7, we have that

(4.11) u>uλ0 in Σλ0.u>u_{\lambda_{0}}\qquad{\mbox{ in }}\Sigma_{\lambda_{0}}.

We claim that there exists ε>0\varepsilon>0, depending on nn, ss, pp and uu, such that

(4.12) uuλu\geqslant u_{\lambda} in Σλ\Sigma_{\lambda}, for all λ[λ0,λ0+ε)\lambda\in[\lambda_{0},\lambda_{0}+\varepsilon).

Indeed, we recall the definition of Σλ\Sigma_{\lambda}^{-} in (4.6) and that, for any q>max{n,2/(p1)}q>\max\{n,2/(p-1)\}, the estimate in (4.7) holds true, namely

(4.13) uλuLq(Σλ)cuλLq(p1)(Σλ)p1uλuLq(Σλ).\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}\leqslant c\|u_{\lambda}\|^{p-1}_{L^{q(p-1)}(\Sigma_{\lambda}^{-})}\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}.

Now we pick R0>0R_{0}>0 large enough such that, for all λ[λ0,λ0+1]\lambda\in[\lambda_{0},\lambda_{0}+1],

(4.14) cuλLq(p1)(Σλ(nBR0))p113.c\|u_{\lambda}\|^{p-1}_{L^{q(p-1)}(\Sigma_{\lambda}^{-}\cap(\mathbb{R}^{n}\setminus B_{R_{0}}))}\leqslant\frac{1}{3}.

Moreover, from (4.11) we deduce that, for ε>0\varepsilon>0 small enough, we can find cε>0c_{\varepsilon}>0 such that

u(x)uλ0(x)cεfor every xΣλ0+εBR0.u(x)-u_{\lambda_{0}}(x)\geqslant c_{\varepsilon}\qquad\text{for every }x\in\Sigma_{\lambda_{0}+\varepsilon}\cap B_{R_{0}}.

In addition, by the continuity of uu, up to taking ε\varepsilon smaller, we have that, for all λ[λ0,λ0+ε)\lambda\in[\lambda_{0},\lambda_{0}+\varepsilon),

u(x)uλ(x)for every xΣλ0+εBR0.u(x)\geqslant u_{\lambda}(x)\qquad\text{for every }x\in\Sigma_{\lambda_{0}+\varepsilon}\cap B_{R_{0}}.

This implies that, for all λ[λ0,λ0+δε)\lambda\in[\lambda_{0},\lambda_{0}+\delta_{\varepsilon}),

ΣλBR0(ΣλΣλ0+ε)BR0.\Sigma_{\lambda}^{-}\cap B_{R_{0}}\subset(\Sigma_{\lambda}\setminus\Sigma_{\lambda_{0}+\varepsilon})\cap B_{R_{0}}.

Thus, for all λ[λ0,λ0+ε)\lambda\in[\lambda_{0},\lambda_{0}+\varepsilon), we have that

|ΣλBR0||(ΣλΣλ0+ε)BR0|cnR0n1ε.|\Sigma_{\lambda}^{-}\cap B_{R_{0}}|\leqslant|(\Sigma_{\lambda}\setminus\Sigma_{\lambda_{0}+\varepsilon})\cap B_{R_{0}}|\leqslant c_{n}R_{0}^{n-1}\varepsilon.

As a result, up to taking ε\varepsilon even smaller, we find that, for all λ[λ0,λ0+δε0)\lambda\in[\lambda_{0},\lambda_{0}+\delta_{\varepsilon_{0}}),

cuλLq(p1)(ΣλBR0)p113.c\|u_{\lambda}\|^{p-1}_{L^{q(p-1)}(\Sigma_{\lambda}^{-}\cap B_{R_{0}})}\leqslant\frac{1}{3}.

As a consequence of this and (4.14), for all λ[λ0,λ0+δε0)\lambda\in[\lambda_{0},\lambda_{0}+\delta_{\varepsilon_{0}}),

cuλLq(p1)(Σλ)p123.c\|u_{\lambda}\|^{p-1}_{L^{q(p-1)}(\Sigma_{\lambda}^{-})}\leqslant\frac{2}{3}.

Pluggin this information into (4.13), we thereby obtain that

uλuLq(Σλ)23uλuLq(Σλ).\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}\leqslant\frac{2}{3}\|u_{\lambda}-u\|_{L^{q}(\Sigma_{\lambda}^{-})}.

This, in turn, implies that |Σλ|=0|\Sigma_{\lambda}^{-}|=0 for all λ[λ0,λ0+δε0)\lambda\in[\lambda_{0},\lambda_{0}+\delta_{\varepsilon_{0}}), which gives the desired claim in (4.12).

Owing to (4.12) and Lemma 4.7, we conclude that, for all λ[λ0,λ0+δε0)\lambda\in[\lambda_{0},\lambda_{0}+\delta_{\varepsilon_{0}}), either u>uλu>u_{\lambda} in Σλ\Sigma_{\lambda} or uuλu\equiv u_{\lambda} in n\mathbb{R}^{n}. This and (4.9) give that the second possibility cannot occur, and therefore, for all λ[λ0,λ0+δε0)\lambda\in[\lambda_{0},\lambda_{0}+\delta_{\varepsilon_{0}}), we have that u>uλu>u_{\lambda} in Σλ\Sigma_{\lambda}.

This is a contradiction with the definition of λ0\lambda_{0} and thus the desired result is established. ∎

Proof of Theorem 4.6.

By translation, we may suppose that λ0=0\lambda_{0}=0. Hence, we have that uu is symmetric with respect to the x1x_{1}-axis, i.e. u(x1,x)=u(x1,x)u(x_{1},x^{\prime})=u(-x_{1},x^{\prime}). Using the same approach in any arbitrary direction, we obtain that uu is radially symmetric (see e.g. [27, Lemma 8.2.1] for full details of this standard argument). ∎

Proof of Theorem 1.5.

Theorem 1.5 now follows from Theorems 4.1, 4.2 and 4.6. ∎

Appendix A Properties of the heat kernel

In this section, we focus on the fundamental properties of the heat kernel introduced in (1.3), with the aim of establishing Theorem 3.1. For this, we observe that the main step to establish Theorem 3.1 is to investigate the inverse Fourier transform and the asymptotic behavior of \mathcal{H} by scaling techniques. We thus need to overcome the additional difficulties caused by the fact that the operator that we take into account is not scale invariant, and therefore our analysis cannot rely entirely on either the purely classical or the purely fractional counterparts, as given in [19, 30].

To start with, let us define a “two-scales” function for s(0,1)s\in(0,1), t1t_{1}, t2>0t_{2}>0 and xnx\in\mathbb{R}^{n} as

(A.1) (x,t1,t2):=ne(t1|ξ|2s+t2|ξ|2)e2πixξ𝑑ξ,\mathcal{H}(x,t_{1},t_{2}):=\int_{\mathbb{R}^{n}}e^{-(t_{1}|\xi|^{2s}+t_{2}|\xi|^{2})}e^{2\pi ix\cdot\xi}\,d\xi,

which satisfies the following rescaling properties

(A.2) (x,t,t)=tn2s(t12sx,1,t11s)=tn2(t12x,t1s,1).\mathcal{H}(x,t,t)=t^{-\frac{n}{2s}}\mathcal{H}(t^{-\frac{1}{2s}}x,1,t^{1-\frac{1}{s}})=t^{-\frac{n}{2}}\mathcal{H}(t^{-\frac{1}{2}}x,t^{1-s},1).

We notice that if t(1,+)t\in(1,+\infty), then t11s(0,1)t^{1-\frac{1}{s}}\in(0,1), while if t(0,1)t\in(0,1), then t1s(0,1)t^{1-s}\in(0,1). As a consequence of this, we shall take 2s2s-scaling for t(1,+)t\in(1,+\infty) and take 22-scaling for t(0,1)t\in(0,1) in order to discuss the properties of the heat kernel \mathcal{H}.

A.1. Nonnegativity of heat kernel

We perform some auxiliary analysis on the kernel (x,t)\mathcal{H}(x,t) defined in (1.3). For this sake, we recall a result contained in [30, Lemma A.2].

Lemma A.1.

If ff and gL1(n)g\in L^{1}(\mathbb{R}^{n}) are radially symmetric, nonnegative and decreasing in r=|x|r=|x|, then fgf\ast g is radially symmetric and decreasing in rr.

Then, we have:

Lemma A.2.

Let n1n\geqslant 1 and s(0,1)s\in(0,1). Let \mathcal{H} be defined as in (1.3).

Then, \mathcal{H} is nonnegative, radially symmetric, and nonincreasing with respect to r=|x|r=|x|.

Proof.

We point out that, being the Fourier transform of a radially symmetric function, (x,t)\mathcal{H}(x,t) is radially symmetric in xx. To prove the nonnegativity of \mathcal{H}, we adapt the arguments in [30].

For all s(0,1]s\in(0,1], we define the radially symmetric nonnegative functions fsf_{s} as

fs(x):=As(1|x|n+2sχnB1(x)+χB1(x)),f_{s}(x):=A_{s}\left(\frac{1}{|x|^{n+2s}}\,\chi_{\mathbb{R}^{n}\setminus B_{1}}(x)+\chi_{B_{1}}(x)\right),

and, for any a>0a>0,

ga(x):=A1an2πn2eπ2a|x|2,g^{a}(x):=A_{1}a^{-\frac{n}{2}}\pi^{\frac{n}{2}}e^{-\frac{\pi^{2}}{a}|x|^{2}},

where AsA_{s} and A1A_{1} are such that nfs(x)𝑑x=1\int_{\mathbb{R}^{n}}f_{s}(x)\,dx=1 and nga(x)𝑑x=1\int_{\mathbb{R}^{n}}g^{a}(x)\,dx=1, respectively.

In this way, we have that

(fs)(ξ)=1+n(cos(2πξx)1)fs(x)𝑑x=1+As|ξ|2s|y||ξ|cos(2πy1)1|y|n+2s𝑑y+As|ξ|n|y||ξ|cos(2πy1)1dy.\begin{split}\mathcal{F}(f_{s})(\xi)&=1+\int_{\mathbb{R}^{n}}\left(\cos(2\pi\xi\cdot x)-1\right)f_{s}(x)\,dx\\ &=1+A_{s}|\xi|^{2s}\int_{|y|\geqslant|\xi|}\frac{\cos\left(2\pi y_{1}\right)-1}{|y|^{n+2s}}\,dy+A_{s}|\xi|^{-n}\int_{|y|\leqslant|\xi|}{\cos\left(2\pi y_{1}\right)-1}\,dy.\end{split}

Let also

c:=Asncos(2πy1)1|y|n+2s𝑑y.c:=-A_{s}\int_{\mathbb{R}^{n}}\frac{\cos\left(2\pi y_{1}\right)-1}{|y|^{n+2s}}\,dy.

Then, it is immediate to check that

(fs)(ξ)=1c|ξ|2s(1+ω(ξ)),\mathcal{F}(f_{s})(\xi)=1-c|\xi|^{2s}(1+\omega(\xi)),

where ω(ξ)0\omega(\xi)\to 0 if ξ0\xi\to 0.

We now define, for all kk\in\mathbb{N},

fk(x):=kn2s(fsfsfs)(k12sx),f_{k}(x):=k^{\frac{n}{2s}}(f_{s}\ast f_{s}\ast\cdots\ast f_{s})(k^{\frac{1}{2s}}x),

the convolution product being taken kk times.

By the properties of the Fourier transform with respect to the convolution product, one has that, for all ξn\xi\in\mathbb{R}^{n},

(A.3) (fk)(ξ)=((fs)(ξk1/(2s)))k=(1c|ξ|2sk(1+ω(ξk1/(2s))))k.\mathcal{F}(f_{k})(\xi)=\left(\mathcal{F}(f_{s})\left(\frac{\xi}{k^{1/(2s)}}\right)\right)^{k}=\left(1-c\frac{|\xi|^{2s}}{k}\left(1+\omega\left(\frac{\xi}{k^{1/(2s)}}\right)\right)\right)^{k}.

We define, for all kk\in\mathbb{N},

gka(x):=fkga(x)g_{k}^{a}(x):=f_{k}\ast g^{a}(x)

and, by combining (A.3) and the fact that (ga)(ξ)=ea|ξ|2\mathcal{F}(g^{a})(\xi)=e^{-a|\xi|^{2}}, we see that

(gka)(ξ)=(1c|ξ|2sk(1+ω(ξk1/(2s))))kea|ξ|2.\mathcal{F}(g_{k}^{a})(\xi)=\left(1-c\frac{|\xi|^{2s}}{k}\left(1+\omega\left(\frac{\xi}{k^{1/(2s)}}\right)\right)\right)^{k}e^{-a|\xi|^{2}}.

We note that the right-hand-side of the above equation converges to ec|ξ|2sa|ξ|2e^{-c|\xi|^{2s}-a|\xi|^{2}} pointwise as k+k\to+\infty. Furthermore, since, for all ξn\xi\in\mathbb{R}^{n},

|(gka)(ξ)|gkaL1(n)=1,|\mathcal{F}(g_{k}^{a})(\xi)|\leqslant\|g_{k}^{a}\|_{L^{1}(\mathbb{R}^{n})}=1,

this convergence also holds in 𝒮(n)\mathcal{S}^{\prime}(\mathbb{R}^{n}), that is, as k+k\to+\infty,

(gka)(ξ)ec|ξ|2sa|ξ|2in 𝒮(n).\mathcal{F}(g_{k}^{a})(\xi)\to e^{-c|\xi|^{2s}-a|\xi|^{2}}\qquad\text{in }\mathcal{S}^{\prime}(\mathbb{R}^{n}).

As a result, taking the inverse Fourier transform, we see that gkag_{k}^{a} converges in 𝒮(n)\mathcal{S}^{\prime}(\mathbb{R}^{n}) to (x,c,a)\mathcal{H}(x,c,a) for any a>0a>0.

Since gkag_{k}^{a} is nonnegative for all kk and a>0a>0, we deduce the nonnegativity of (x,c,a)\mathcal{H}(x,c,a).

By scaling and exploiting the continuity of (x,c,a)\mathcal{H}(x,c,a) with respect to xx, one derives the nonnegativity of (x,t,t)\mathcal{H}(x,t,t) for all t>0t>0.

The monotonicity of the heat kernel \mathcal{H} can be also deduced by the fact that fsf_{s} and gag^{a} are nonincreasing in r=|x|r=|x| and Lemma A.1. ∎

A.2. Asymptotic formulae for the heat kernel

To obtain the bounds on \mathcal{H}, we point out two asymptotic formulae for (x,t1,t2)\mathcal{H}(x,t_{1},t_{2}) as defined in (A.1). These asymptotics are described in the next Lemmata A.3 and A.4.

Lemma A.3.

Let n1n\geqslant 1, s(0,1)s\in(0,1), η(0,1)\eta\in(0,1) and (x,1,η)\mathcal{H}(x,1,\eta) be as defined by (A.1).

Then, for any ε>0\varepsilon>0, there exists M>0M>0 independent of η\eta such that, for every |x|>M|x|>M and η(0,1)\eta\in(0,1),

||x|n+2s(x,1,η)2n+2sπn21ssin(πs)Γ(n2+s)Γ(s)|<ε.\left||x|^{n+2s}\mathcal{H}(x,1,\eta)-2^{n+2s}\pi^{\frac{n}{2}-1}s\sin(\pi s)\Gamma\left(\frac{n}{2}+s\right)\Gamma(s)\right|<\varepsilon.

That is,

lim|x|+|x|n+2s(x,1,η)=2n+2sπn21ssin(πs)Γ(n2+s)Γ(s),\lim\limits_{|x|\to+\infty}|x|^{n+2s}\mathcal{H}(x,1,\eta)=2^{n+2s}\pi^{\frac{n}{2}-1}s\sin(\pi s)\Gamma\left(\frac{n}{2}+s\right)\Gamma(s),

and the limit is uniform with respect to η(0,1)\eta\in(0,1).

Proof.

From the definition of (x,t1,t2)\mathcal{H}(x,t_{1},t_{2}), one has that

(x,1,η)=ne|ξ|2sη|ξ|2e2πixξ𝑑ξ.\mathcal{H}(x,1,\eta)=\int_{\mathbb{R}^{n}}e^{-|\xi|^{2s}-\eta|\xi|^{2}}e^{2\pi ix\cdot\xi}\,d\xi.

By using the Fourier Inversion Theorem for radial functions, see e.g. [20, Chapter II], and the Bessel functions in [28], it is immediate to obtain the 11-dimensional integral representation of (x,1,η)\mathcal{H}(x,1,\eta), namely that

(A.4) (x,1,η)=(2π)n2|x|n210+er2sηr2rn2Jn21(|x|r)𝑑r,\mathcal{H}(x,1,\eta)=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{\frac{n}{2}-1}}\int_{0}^{+\infty}e^{-r^{2s}-\eta r^{2}}r^{\frac{n}{2}}J_{\frac{n}{2}-1}(|x|r)\,dr,

where Jn21J_{\frac{n}{2}-1} denotes the Bessel function of first kind of order n21\frac{n}{2}-1.

We also recall that, for all zz\in\mathbb{C},

ddz(zn2Jn2(z))=zn2Jn21(z)\frac{d}{dz}\left(z^{\frac{n}{2}}J_{\frac{n}{2}}(z)\right)=z^{\frac{n}{2}}J_{\frac{n}{2}-1}(z)

and

Jn2(z)2πzcos(znπ4π4)when |z|+ and argz<π2.J_{\frac{n}{2}}(z)\to\sqrt{\frac{2}{\pi z}}\cos\left(z-\frac{n\pi}{4}-\frac{\pi}{4}\right)\qquad{\mbox{when~{}$|z|\to+\infty\;$ and~{}$\;\arg z<\frac{\pi}{2}$.}}

Consequently,

(A.5) (x,1,η)=(2π)n2|x|n210+er2sηr2rn2Jn21(|x|r)𝑑r=(2π)n2|x|n0+et2s|x|2sηt2|x|2ddt(tn2Jn2(t))𝑑t=(2π)n2|x|net2s|x|2sηt2|x|2tn2Jn2(t)|t=0t=++(2π)n2|x|n0+et2s|x|2sηt2|x|2(2stn2+2s1|x|2s+η2tn2+1|x|2)Jn2(t)𝑑t=(2π)n2|x|n0+et2s|x|2sηt2|x|2(2stn2+2s1|x|2s+η2tn2+1|x|2)Jn2(t)𝑑t.\begin{split}\mathcal{H}(x,1,\eta)&=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{\frac{n}{2}-1}}\int_{0}^{+\infty}e^{-r^{2s}-\eta r^{2}}r^{\frac{n}{2}}J_{\frac{n}{2}-1}(|x|r)\,dr\\ &=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{n}}\int_{0}^{+\infty}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}\frac{d}{dt}\left(t^{\frac{n}{2}}J_{\frac{n}{2}}(t)\right)\,dt\\ &={(2\pi)^{\frac{n}{2}}}{|x|^{-n}}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}t^{\frac{n}{2}}J_{\frac{n}{2}}(t)\bigg{|}_{t=0}^{t=+\infty}\\ &\qquad+{(2\pi)^{\frac{n}{2}}}{|x|^{-n}}\int_{0}^{+\infty}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}\left(\frac{2st^{\frac{n}{2}+2s-1}}{|x|^{2s}}+\eta\frac{2t^{\frac{n}{2}+1}}{|x|^{2}}\right)J_{\frac{n}{2}}(t)\,dt\\ &={(2\pi)^{\frac{n}{2}}}{|x|^{-n}}\int_{0}^{+\infty}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}\left(\frac{2st^{\frac{n}{2}+2s-1}}{|x|^{2s}}+\eta\frac{2t^{\frac{n}{2}+1}}{|x|^{2}}\right)J_{\frac{n}{2}}(t)\,dt.\end{split}

As a result,

(A.6) |x|n+2s(x,1,η)=(2π)n20+et2s|x|2sηt2|x|2(2stn2+2s1+2ηtn2+1|x|2s2)Jn2(t)𝑑t=(2π)n2Re0+et2s|x|2sηt2|x|2(2stn2+2s1+2ηtn2+1|x|2s2)Hn2(1)(t)𝑑t,\begin{split}|x|^{n+2s}\mathcal{H}(x,1,\eta)&={(2\pi)^{\frac{n}{2}}}\int_{0}^{+\infty}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}\left({2st^{\frac{n}{2}+2s-1}}+2\eta{t^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)J_{\frac{n}{2}}(t)\,dt\\ &={(2\pi)^{\frac{n}{2}}}\text{Re}\int_{0}^{+\infty}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}\left({2st^{\frac{n}{2}+2s-1}}+2\eta{t^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(t)\,dt,\end{split}

where Hn2(1)(z)H^{(1)}_{\frac{n}{2}}(z) is the Bessel function of the third kind and Re AA denotes the real part of AA.

We now choose a straight line L1L_{1} running from 0 to ++\infty in the upper half-plane and making a π6\frac{\pi}{6} angle with the positive real axis, that is

(A.7) L1:={z:argz=π6}.L_{1}:=\left\{z\in\mathbb{C}:\arg z=\frac{\pi}{6}\right\}.

Furthermore, applying the Residue Theorem, we find that

(A.8) 0+et2s|x|2sηt2|x|2(2stn2+2s1+2ηtn2+1|x|2s2)Hn2(1)(t)𝑑t=L1ez2s|x|2sηz2|x|2(2szn2+2s1+2ηzn2+1|x|2s2)Hn2(1)(z)𝑑zlimR+CRez2s|x|2sηz2|x|2(2szn2+2s1+2ηzn2+1|x|2s2)Hn2(1)(z)𝑑z,\begin{split}&\int_{0}^{+\infty}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}\left({2st^{\frac{n}{2}+2s-1}}+2\eta{t^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(t)\,dt\\ =&\int_{L_{1}}e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}\left({2sz^{\frac{n}{2}+2s-1}}+2\eta{z^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz\\ &\quad-\lim\limits_{R\to+\infty}\int_{C_{R}}e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}\left({2sz^{\frac{n}{2}+2s-1}}+2\eta{z^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz,\end{split}

where CRC_{R} denotes the arc

CR:={z s.t. z=Reiθ, with θ(0,π6)}.C_{R}:=\left\{z\in\mathbb{C}\;{\mbox{ s.t. }}\;z=Re^{i\theta},\;{\mbox{ with }}\theta\in\left(0,\frac{\pi}{6}\right)\right\}.

We claim that, for every |x|>1|x|>1 and η(0,1)\eta\in(0,1),

(A.9) limR+CRez2s|x|2sηz2|x|2(2szn2+2s1+2ηzn2+1|x|2s2)Hn2(1)(z)𝑑z=0.\lim\limits_{R\to+\infty}\int_{C_{R}}e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}\left({2sz^{\frac{n}{2}+2s-1}}+2\eta{z^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz=0.

Indeed, since zCRz\in C_{R}, one has that

(A.10) |CRez2s|x|2sηz2|x|2(2szn2+2s1+2ηzn2+1|x|2s2)Hn2(1)(z)𝑑z|=|0π6eR2sei2sθ|x|2sηR2ei2θ|x|2(2s(Reiθ)n2+2s1+2η(Reiθ)n2+1|x|2s2)Hn2(1)(Reiθ)iReiθ𝑑θ|0π6eR2scos2sθ|x|2sηR2cos2θ|x|2(2sRn2+2s+2ηRn2+2|x|2s2)|Hn2(1)(Reiθ)|𝑑θeR2s2|x|2sηR22|x|2(2sRn2+2s+2ηRn2+2|x|2s2)0π6|Hn2(1)(Reiθ)|𝑑θ.\begin{split}&\left|\int_{C_{R}}e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}\left({2sz^{\frac{n}{2}+2s-1}}+2\eta{z^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz\right|\\ =\;&\left|\int_{0}^{\frac{\pi}{6}}e^{-\frac{R^{2s}e^{i2s\theta}}{|x|^{2s}}-\eta\frac{R^{2}e^{i2\theta}}{|x|^{2}}}\left({2s(Re^{i\theta})^{\frac{n}{2}+2s-1}}+2\eta{(Re^{i\theta})^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(Re^{i\theta})iRe^{i\theta}\,d\theta\right|\\ \leqslant\;&\int_{0}^{\frac{\pi}{6}}e^{-\frac{R^{2s}\cos 2s\theta}{|x|^{2s}}-\eta\frac{R^{2}\cos 2\theta}{|x|^{2}}}\left({2sR^{\frac{n}{2}+2s}}+2\eta{R^{\frac{n}{2}+2}}{|x|^{2s-2}}\right)\left|H^{(1)}_{\frac{n}{2}}(Re^{i\theta})\right|\,d\theta\\ \leqslant\;&e^{-\frac{R^{2s}}{2|x|^{2s}}-\eta\frac{R^{2}}{2|x|^{2}}}\left({2sR^{\frac{n}{2}+2s}}+2\eta{R^{\frac{n}{2}+2}}{|x|^{2s-2}}\right)\int_{0}^{\frac{\pi}{6}}\left|H^{(1)}_{\frac{n}{2}}(Re^{i\theta})\right|\,d\theta.\end{split}

Also, from the fact that

Hn2(1)(z)2πzei(znπ4π4)when |z|+ and argz<π2,H^{(1)}_{\frac{n}{2}}(z)\to\sqrt{\frac{2}{\pi z}}e^{i(z-\frac{n\pi}{4}-\frac{\pi}{4})}\qquad{\mbox{when~{}$|z|\to+\infty\;$ and~{}$\;\arg z<\frac{\pi}{2}$,}}

one obtains that

(A.11) limR+|Hn2(1)(Reiθ)|limR+|2πReiθ|=0.\lim_{R\to+\infty}\left|H^{(1)}_{\frac{n}{2}}(Re^{i\theta})\right|\leqslant\lim_{R\to+\infty}\left|\sqrt{\frac{2}{\pi Re^{i\theta}}}\right|=0.

By combining (A.11) with (LABEL:step_1_estimate_of_H1), we conclude that

limR+|CRez2s|x|2sηz2|x|2(2szn2+2s1+2ηzn2+1|x|2s2)Hn2(1)(z)𝑑z|=0,\lim_{R\to+\infty}\left|\int_{C_{R}}e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}\left({2sz^{\frac{n}{2}+2s-1}}+2\eta{z^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz\right|=0,

which implies (A.9), as desired.

As a consequence, recalling (A.6), and inserting (A.9) into (LABEL:L1), we find that

(A.12) |x|n+2s(x,1,η)=(2π)n2ReL1ez2s|x|2sηz2|x|2(2szn2+2s1+2ηzn2+1|x|2s2)Hn2(1)(z)𝑑z=(2π)n2Re0+e(reiπ6)2s|x|2sη(reiπ6)2|x|2(2s(reiπ6)n2+2s1+2η(reiπ6)n2+1|x|2s2)Hn2(1)(reiπ6)eiπ6𝑑r.\begin{split}&|x|^{n+2s}\mathcal{H}(x,1,\eta)={(2\pi)^{\frac{n}{2}}}\text{Re}\int_{L_{1}}e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}\left({2sz^{\frac{n}{2}+2s-1}}+2\eta{z^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz\\ =\;&(2\pi)^{\frac{n}{2}}\text{Re}\int_{0}^{+\infty}e^{-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}\left({2s\left(re^{i\frac{\pi}{6}}\right)^{\frac{n}{2}+2s-1}}+2\eta{\left(re^{i\frac{\pi}{6}}\right)^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)e^{i\frac{\pi}{6}}\,dr.\end{split}

Now we claim that for any ε>0\varepsilon>0 there exists M>0M>0 independent of η\eta such that, for every |x|>M|x|>M and η(0,1)\eta\in(0,1),

||x|n+2s(x,1,η)(2π)n2ReL12szn2+2s1Hn2(1)(z)𝑑z|<ε.\left||x|^{n+2s}\mathcal{H}(x,1,\eta)-(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz\right|<\varepsilon.

Indeed, we first make use of the properties of Hn2(1)(z)H^{(1)}_{\frac{n}{2}}\left(z\right) on L1L_{1} defined in (A.7) to prove that the integral in the right-hand-side of (LABEL:x_H(x,1,eta)) is uniformly bounded with respect to |x|>1|x|>1 and η(0,1)\eta\in(0,1).

Let zL1z\in L_{1}, from the definition of Hn2(1)(z)H^{(1)}_{\frac{n}{2}}\left(z\right) (see e.g. [28, page 21]), one has

(A.13) |Hn2(1)(z)|=|Hn2(1)(reiπ6)|=|ieinπ4+eireiπ6et+et2ent2𝑑t|0+er4(et+et)(ent2+ent2)𝑑t20+er4etent2𝑑t.\begin{split}&\left|H^{(1)}_{\frac{n}{2}}\left(z\right)\right|=\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|=\left|-ie^{-i\frac{n\pi}{4}}\int_{-\infty}^{+\infty}e^{ire^{i\frac{\pi}{6}}\frac{e^{t}+e^{-t}}{2}}e^{-\frac{nt}{2}}\,dt\right|\\ &\qquad\leqslant\int_{0}^{+\infty}e^{-\frac{r}{4}\left(e^{t}+e^{-t}\right)}\left(e^{\frac{nt}{2}}+e^{\frac{-nt}{2}}\right)\,dt\leqslant 2\int_{0}^{+\infty}e^{-\frac{r}{4}e^{t}}e^{\frac{nt}{2}}\,dt.\end{split}

Furthermore, we derive that, for every η(0,1)\eta\in(0,1) and |x|>1|x|>1,

(A.14) |0+e(reiπ6)2s|x|2sη(reiπ6)2|x|2(2s(reiπ6)n2+2s1+2η(reiπ6)n2+1|x|2s2)Hn2(1)(reiπ6)eiπ6𝑑r|0+(2srn2+2s1+2rn2+1)|Hn2(1)(reiπ6)|𝑑r4s0+ent20+er4etrn2+2s1𝑑r𝑑t+40+ent20+er4etrn2+1𝑑r𝑑ts4n2+2s+1Γ(n2+2s)0+e2st𝑑t+4n2+2Γ(n2+2)0+e2t𝑑t2n+4s+1Γ(n2+2s)+2n+3Γ(n2+2).\begin{split}&\left|\int_{0}^{+\infty}e^{-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}\left({2s\left(re^{i\frac{\pi}{6}}\right)^{\frac{n}{2}+2s-1}}+2\eta{\left(re^{i\frac{\pi}{6}}\right)^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)e^{i\frac{\pi}{6}}\,dr\right|\\ \leqslant\;&\int_{0}^{+\infty}\left({2sr^{\frac{n}{2}+2s-1}}+2r^{\frac{n}{2}+1}\right)\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ \leqslant\;&4s\int_{0}^{+\infty}e^{\frac{nt}{2}}\int_{0}^{+\infty}e^{-\frac{r}{4}e^{t}}{r^{\frac{n}{2}+2s-1}}\,dr\,dt+4\int_{0}^{+\infty}e^{\frac{nt}{2}}\int_{0}^{+\infty}e^{-\frac{r}{4}e^{t}}{r^{\frac{n}{2}+1}}\,dr\,dt\\ \leqslant\;&s4^{\frac{n}{2}+2s+1}\Gamma(\frac{n}{2}+2s)\int_{0}^{+\infty}e^{-2st}\,dt+4^{\frac{n}{2}+2}\Gamma(\frac{n}{2}+2)\int_{0}^{+\infty}e^{-2t}\,dt\\ \leqslant\;&2^{n+4s+1}\Gamma(\frac{n}{2}+2s)+2^{n+3}\Gamma(\frac{n}{2}+2).\end{split}

By combining (LABEL:x_H(x,1,eta)) with (A.14), one has

(A.15) ||x|n+2s(x,1,η)(2π)n2ReL12szn2+2s1Hn2(1)(z)𝑑z|=|(2π)n2ReL1(ez2s|x|2sηz2|x|21)2szn2+2s1Hn2(1)(z)+ez2s|x|2sηz2|x|22ηzn2+1|x|2s2Hn2(1)(z)dz||(2π)n2ReL1(ez2s|x|2sηz2|x|21)2szn2+2s1Hn2(1)(z)𝑑z|+(2π)n22n+3Γ(n2+2)|x|2s2=:A1+A2.\begin{split}&\left||x|^{n+2s}\mathcal{H}(x,1,\eta)-(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz\right|\\ =\;&\left|{(2\pi)^{\frac{n}{2}}}\text{Re}\int_{L_{1}}\left(e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}-1\right){2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}(z)+e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}2\eta{z^{\frac{n}{2}+1}}{|x|^{2s-2}}H^{(1)}_{\frac{n}{2}}(z)\,dz\right|\\ \leqslant\;&\left|{(2\pi)^{\frac{n}{2}}}\text{Re}\int_{L_{1}}\left(e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}-1\right){2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}(z)\,dz\right|+(2\pi)^{\frac{n}{2}}2^{n+3}\Gamma\left(\frac{n}{2}+2\right)|x|^{2s-2}\\ =:\;&A_{1}+A_{2}.\end{split}

Let us estimate the term A1A_{1}. For some R>0R>0 large enough, one has

(A.16) A1(2π)n20+|e(reiπ6)2s|x|2sη(reiπ6)2|x|21|2srn2+2s1|Hn2(1)(reiπ6)|𝑑r(2π)n20R|e(reiπ6)2s|x|2sη(reiπ6)2|x|21|2srn2+2s1|Hn2(1)(reiπ6)|𝑑r+(2π)n2R+|e(reiπ6)2s|x|2sη(reiπ6)2|x|21|2srn2+2s1|Hn2(1)(reiπ6)|𝑑r,=:B1R+B2R.\begin{split}A_{1}&\leqslant(2\pi)^{\frac{n}{2}}\int_{0}^{+\infty}\left|e^{-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}-1\right|{2sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ &\leqslant(2\pi)^{\frac{n}{2}}\int_{0}^{R}\left|e^{-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}-1\right|{2sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ &\qquad+(2\pi)^{\frac{n}{2}}\int_{R}^{+\infty}\left|e^{-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}-1\right|{2sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr,\\ &=:B_{1}^{R}+B_{2}^{R}.\end{split}

From (A.14), we observe that

B2R(2π)n2R+2srn2+2s1|Hn2(1)(reiπ6)|𝑑r(2π)n24s0+ent2R+er4etrn2+2s1𝑑r𝑑t(2π)n24s0+ent2R+er4(t+1)rn2+2s1𝑑r𝑑t(2π)n24s0+ent2Rt4R+er4rn2+2s1𝑑r𝑑t(2π)n2s2n+4s+4Γ(n2+2s)R2n.\begin{split}B_{2}^{R}&\leqslant(2\pi)^{\frac{n}{2}}\int_{R}^{+\infty}{2sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\leqslant(2\pi)^{\frac{n}{2}}4s\int_{0}^{+\infty}e^{\frac{nt}{2}}\int_{R}^{+\infty}e^{-\frac{r}{4}e^{t}}{r^{\frac{n}{2}+2s-1}}\,dr\,dt\\ &\leqslant(2\pi)^{\frac{n}{2}}4s\int_{0}^{+\infty}e^{\frac{nt}{2}}\int_{R}^{+\infty}e^{-\frac{r}{4}(t+1)}{r^{\frac{n}{2}+2s-1}}\,dr\,dt\leqslant(2\pi)^{\frac{n}{2}}4s\int_{0}^{+\infty}e^{\frac{nt}{2}-\frac{Rt}{4}}\int_{R}^{+\infty}e^{-\frac{r}{4}}{r^{\frac{n}{2}+2s-1}}\,dr\,dt\\ &\leqslant\frac{(2\pi)^{\frac{n}{2}}s2^{n+4s+4}\Gamma(\frac{n}{2}+2s)}{R-2n}.\end{split}

As a result, for any ε>0\varepsilon>0, taking

(A.17) R0:=4(2π)n2s2n+4s+4Γ(n2+2s)ε+2n,R_{0}:=\frac{4(2\pi)^{\frac{n}{2}}s2^{n+4s+4}\Gamma(\frac{n}{2}+2s)}{\varepsilon}+2n,

one has that

(A.18) B2R0ε4.B_{2}^{R_{0}}\leqslant\frac{\varepsilon}{4}.

We now turn to estimating B1R0B_{1}^{R_{0}}. For this, we take R0R_{0} as in (A.17) and we remark that

B1R0(2π)n20R0(|er2scossπ3|x|2sηr2cosπ3|x|2cos(r2scossπ3|x|2sηr2cosπ3|x|2)1|+|sin(r2scossπ3|x|2sηr2cosπ3|x|2)|)2srn2+2s1|H(1)n2(reiπ6)|dr.\begin{split}B_{1}^{R_{0}}&\leqslant(2\pi)^{\frac{n}{2}}\int_{0}^{R_{0}}\bigg{(}\left|e^{-\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\eta\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}}\cos\left(-\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\eta\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}\right)-1\right|\\ &\qquad+\left|\sin\left(-\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\eta\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}\right)\right|\bigg{)}{2sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr.\end{split}

Note also that, for any r<R0r<R_{0}, |x|>R02|x|>R_{0}^{2} large enough and η(0,1)\eta\in(0,1),

(A.19) |er2scossπ3|x|2sηr2cosπ3|x|2cos(r2scossπ3|x|2sηr2cosπ3|x|2)1|+|sin(r2scossπ3|x|2s+ηr22|x|2)|1eR02scossπ3|x|2sR022|x|2cos(R02s|x|2s+R022|x|2)+sin(R02s|x|2s+R022|x|2).\begin{split}&\left|e^{-\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\eta\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}}\cos\left(-\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\eta\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}\right)-1\right|+\left|\sin\left(\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}+\eta\frac{r^{2}}{2|x|^{2}}\right)\right|\\ \leqslant\;&1-e^{-\frac{R_{0}^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{R_{0}^{2}}{2|x|^{2}}}\cos\left(\frac{R_{0}^{2s}}{|x|^{2s}}+\frac{R_{0}^{2}}{2|x|^{2}}\right)+\sin\left(\frac{R_{0}^{2s}}{|x|^{2s}}+\frac{R_{0}^{2}}{2|x|^{2}}\right).\end{split}

By combining (A.14) with  (LABEL:R_0_estimate), we see that for every ε>0\varepsilon>0, there exists TR0,ε>0T_{R_{0},\varepsilon}>0 such that, if |x|>TR0,ε|x|>T_{R_{0},\varepsilon},

1eR02scossπ3|x|2sR022|x|2cos(R02s|x|2s+R022|x|2)+sin(R02s|x|2s+R022|x|2)ε2n+4s+3(2π)n2Γ(n2+2s),1-e^{-\frac{R_{0}^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{R_{0}^{2}}{2|x|^{2}}}\cos\left(\frac{R_{0}^{2s}}{|x|^{2s}}+\frac{R_{0}^{2}}{2|x|^{2}}\right)+\sin\left(\frac{R_{0}^{2s}}{|x|^{2s}}+\frac{R_{0}^{2}}{2|x|^{2}}\right)\leqslant\frac{\varepsilon}{2^{n+4s+3}(2\pi)^{\frac{n}{2}}\Gamma(\frac{n}{2}+2s)},

which implies that

(A.20) B1R0<ε4.B_{1}^{R_{0}}<\frac{\varepsilon}{4}.

Recalling (A.16), by combining (A.18) with (A.20), we deduce that, for every ε>0\varepsilon>0, there exists M1>0M_{1}>0, depending only on nn, ss and ε\varepsilon, such that, for every |x|>M1|x|>M_{1},

A1ε2.A_{1}\leqslant\frac{\varepsilon}{2}.

Regarding A2A_{2}, from (LABEL:A1_A2), we see that, for every ε>0\varepsilon>0, there exists M2>0M_{2}>0, depending only on nn, ss and ε\varepsilon, such that, for every |x|>M2|x|>M_{2},

A2ε2.A_{2}\leqslant\frac{\varepsilon}{2}.

We take M:=max{M1,M2}M:=\max\left\{M_{1},M_{2}\right\} and we use (LABEL:A1_A2) to find that, for every |x|>M|x|>M,

||x|n+2s(x,1,η)(2π)n2ReL12szn2+2s1Hn2(1)(z)𝑑z|<ε.\left||x|^{n+2s}\mathcal{H}(x,1,\eta)-(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz\right|<\varepsilon.

That is,

(A.21) lim|x|+|x|n+2s(x,1,η)=(2π)n2ReL12szn2+2s1Hn2(1)(z)𝑑z,\lim\limits_{|x|\to+\infty}|x|^{n+2s}\mathcal{H}(x,1,\eta)=(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz,

where the limit is uniform with respect to η\eta.

Applying the Residue Theorem again to the region composed of L1L_{1} and the straight line

(A.22) L2:={z s.t. argz=π2},L_{2}:=\left\{z\in\mathbb{C}\;{\mbox{ s.t. }}\;\arg z=\frac{\pi}{2}\right\},

we obtain that

L12szn2+2s1Hn2(1)(z)𝑑z=L22szn2+2s1Hn2(1)(z)𝑑zlimR+CR2szn2+2s1Hn2(1)(z)𝑑z,\int_{L_{1}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz=\int_{L_{2}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz-\lim\limits_{R\to+\infty}\int_{C_{R}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz,

where

CR:={z s.t. z=Reiθ, with θ(π6,π2)}.C_{R}:=\left\{z\in\mathbb{C}\;{\mbox{ s.t. }}z=Re^{i\theta},\;{\mbox{ with }}\theta\in\left(\frac{\pi}{6},\frac{\pi}{2}\right)\right\}.

We now claim that

limR+CR2szn2+2s1Hn2(1)(z)𝑑z=0.\lim\limits_{R\to+\infty}\int_{C_{R}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz=0.

Indeed, for R>0R>0 large enough, we have that

limR+|CR2szn2+2s1Hn2(1)(z)𝑑z|=limR+|π6π2iReiθ2s(Reiθ)n2+2s1Hn2(1)(Reiθ)𝑑θ|limR+π6π22sRn2+2s20+eR4etent2𝑑t𝑑θlimR+4πs3Rn2+2s0+eR4(t+1)ent2𝑑t=0.\begin{split}\lim_{R\to+\infty}\left|\int_{C_{R}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz\right|&=\lim_{R\to+\infty}\left|\int_{\frac{\pi}{6}}^{\frac{\pi}{2}}iRe^{i\theta}2s\left(Re^{i\theta}\right)^{\frac{n}{2}+2s-1}H^{(1)}_{\frac{n}{2}}\left(Re^{i\theta}\right)\,d\theta\right|\\ &\leqslant\lim_{R\to+\infty}\int_{\frac{\pi}{6}}^{\frac{\pi}{2}}2sR^{\frac{n}{2}+2s}2\int_{0}^{+\infty}e^{-\frac{R}{4}e^{t}}e^{\frac{nt}{2}}\,dt\,d\theta\\ &\leqslant\lim_{R\to+\infty}\frac{4\pi s}{3}R^{\frac{n}{2}+2s}\int_{0}^{+\infty}e^{-\frac{R}{4}(t+1)}e^{\frac{nt}{2}}\,dt=0.\end{split}

Hence, recalling (A.21) and exploiting the modified Bessel function of the third kind K(x)K(x) (see e.g. [28, page 5]), one finds that

lim|x|+|x|n+2s(x,1,η)=(2π)n2ReL22szn2+2s1Hn2(1)(z)𝑑z=(2π)n2Re0+2srn2+2s1eiπn4eiπsHn2(1)(ir)𝑑r=(2π)n2Re0+2srn2+2s1(i)eiπs2πKn2(r)𝑑r=2n2+2πn21ssin(πs)0+rn2+2s1Kn2(r)𝑑r=2n+2sπn21ssin(πs)Γ(n2+s)Γ(s).\begin{split}\lim\limits_{|x|\to+\infty}|x|^{n+2s}\mathcal{H}(x,1,\eta)&=(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{2}}{2sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz\\ &=(2\pi)^{\frac{n}{2}}\text{Re}\int_{0}^{+\infty}2sr^{\frac{n}{2}+2s-1}e^{i\frac{\pi n}{4}}e^{i\pi s}H^{(1)}_{\frac{n}{2}}\left(ir\right)\,dr\\ &=(2\pi)^{\frac{n}{2}}\text{Re}\int_{0}^{+\infty}2sr^{\frac{n}{2}+2s-1}(-i)e^{i\pi s}\frac{2}{\pi}K_{\frac{n}{2}}(r)\,dr\\ &=2^{\frac{n}{2}+2}\pi^{\frac{n}{2}-1}s\sin(\pi s)\int_{0}^{+\infty}r^{\frac{n}{2}+2s-1}K_{\frac{n}{2}}(r)\,dr\\ &=2^{n+2s}\pi^{\frac{n}{2}-1}s\sin(\pi s)\Gamma\left(\frac{n}{2}+s\right)\Gamma(s).\end{split}

This last integral is evaluated on [28, page 51], allowing us to complete the proof of Lemma A.3. ∎

Lemma A.4.

Let n1n\geqslant 1, s(0,1)s\in(0,1), η(0,1)\eta\in(0,1) and (x,η,1)\mathcal{H}(x,\eta,1) be as defined by (A.1).

Then, for any ε>0\varepsilon>0, there exists M>0M>0 independent of η\eta such that, for every |x|>M|x|>M and η(0,1)\eta\in(0,1),

||x|n+2s(x,η,1)2n+2sπn21ηssin(πs)Γ(n2+s)Γ(s)|<ε.\left||x|^{n+2s}\mathcal{H}(x,\eta,1)-2^{n+2s}\pi^{\frac{n}{2}-1}\eta s\sin(\pi s)\Gamma\left(\frac{n}{2}+s\right)\Gamma(s)\right|<\varepsilon.

That is

(A.23) lim|x|+|x|n+2s(x,η,1)=2n+2sπn21ηssin(πs)Γ(n2+s)Γ(s),\lim\limits_{|x|\to+\infty}|x|^{n+2s}\mathcal{H}(x,\eta,1)=2^{n+2s}\pi^{\frac{n}{2}-1}\eta s\sin(\pi s)\Gamma\left(\frac{n}{2}+s\right)\Gamma(s),

and the limit is uniform with respect to η(0,1)\eta\in(0,1).

Proof.

As in the proof of Lemma A.3, first of all, the 1-dimensional integral representation of (x,η,1)\mathcal{H}(x,\eta,1) follows from (A.5), that is

(x,η,1)=(2π)n2|x|n0+eηt2s|x|2st2|x|2(η2stn2+2s1|x|2s+2tn2+1|x|2)Jn2(t)𝑑t.\mathcal{H}(x,\eta,1)={(2\pi)^{\frac{n}{2}}}{|x|^{-n}}\int_{0}^{+\infty}e^{-\eta\frac{t^{2s}}{|x|^{2s}}-\frac{t^{2}}{|x|^{2}}}\left(\eta\frac{2st^{\frac{n}{2}+2s-1}}{|x|^{2s}}+\frac{2t^{\frac{n}{2}+1}}{|x|^{2}}\right)J_{\frac{n}{2}}(t)\,dt.

It thus follows that

|x|n+2s(x,η,1)=(2π)n20+eηt2s|x|2st2|x|2(2ηstn2+2s1+2tn2+1|x|2s2)Jn2(t)𝑑t=(2π)n2Re0+eηt2s|x|2st2|x|2(2ηstn2+2s1+2tn2+1|x|2s2)Hn2(1)(t)𝑑t.\begin{split}|x|^{n+2s}\mathcal{H}(x,\eta,1)&={(2\pi)^{\frac{n}{2}}}\int_{0}^{+\infty}e^{-\eta\frac{t^{2s}}{|x|^{2s}}-\frac{t^{2}}{|x|^{2}}}\left({2\eta st^{\frac{n}{2}+2s-1}}+2{t^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)J_{\frac{n}{2}}(t)\,dt\\ &={(2\pi)^{\frac{n}{2}}}\text{Re}\int_{0}^{+\infty}e^{-\eta\frac{t^{2s}}{|x|^{2s}}-\frac{t^{2}}{|x|^{2}}}\left({2\eta st^{\frac{n}{2}+2s-1}}+2{t^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(t)\,dt.\end{split}

Exploiting the Residue Theorem once again, and referring to (LABEL:L1) and (A.9), we infer that

|x|n+2s(x,η,1)=(2π)n2ReL1eηz2s|x|2sz2|x|2(2ηszn2+2s1+2zn2+1|x|2s2)Hn2(1)(z)𝑑z=(2π)n2Re0+eη(reiπ6)2s|x|2s(reiπ6)2|x|2(2ηs(reiπ6)n2+2s1+2(reiπ6)n2+1|x|2s2)Hn2(1)(reiπ6)eiπ6𝑑r.\begin{split}&|x|^{n+2s}\mathcal{H}(x,\eta,1)={(2\pi)^{\frac{n}{2}}}\text{Re}\int_{L_{1}}e^{-\eta\frac{z^{2s}}{|x|^{2s}}-\frac{z^{2}}{|x|^{2}}}\left({2\eta sz^{\frac{n}{2}+2s-1}}+2{z^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}(z)\,dz\\ =\;&(2\pi)^{\frac{n}{2}}\text{Re}\int_{0}^{+\infty}e^{-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}\left({2\eta s\left(re^{i\frac{\pi}{6}}\right)^{\frac{n}{2}+2s-1}}+2{\left(re^{i\frac{\pi}{6}}\right)^{\frac{n}{2}+1}}{|x|^{2s-2}}\right)H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)e^{i\frac{\pi}{6}}\,dr.\end{split}

Moreover, owing to (A.14), we observe that, for every η(0,1)\eta\in(0,1),

(A.24) ||x|n+2s(x,η,1)(2π)n2ReL12ηszn2+2s1Hn2(1)(z)𝑑z|=|(2π)n2ReL1(eηz2s|x|2sz2|x|21)2ηszn2+2s1Hn2(1)(z)+eηz2s|x|2sz2|x|22zn2+1|x|2s2Hn2(1)(z)dz||(2π)n2ReL1(eηz2s|x|2sz2|x|21)2ηszn2+2s1Hn2(1)(z)𝑑z|+(2π)n22n+3Γ(n2+2)|x|2s2=:Q1+Q2.\begin{split}&\left||x|^{n+2s}\mathcal{H}(x,\eta,1)-(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{2\eta sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz\right|\\ =\;&\left|{(2\pi)^{\frac{n}{2}}}\text{Re}\int_{L_{1}}\left(e^{-\eta\frac{z^{2s}}{|x|^{2s}}-\frac{z^{2}}{|x|^{2}}}-1\right){2\eta sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}(z)+e^{-\eta\frac{z^{2s}}{|x|^{2s}}-\frac{z^{2}}{|x|^{2}}}2{z^{\frac{n}{2}+1}}{|x|^{2s-2}}H^{(1)}_{\frac{n}{2}}(z)\,dz\right|\\ \leqslant\;&\left|{(2\pi)^{\frac{n}{2}}}\text{Re}\int_{L_{1}}\left(e^{-\eta\frac{z^{2s}}{|x|^{2s}}-\frac{z^{2}}{|x|^{2}}}-1\right){2\eta sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}(z)\,dz\right|+(2\pi)^{\frac{n}{2}}2^{n+3}\Gamma\left(\frac{n}{2}+2\right)|x|^{2s-2}\\ =:\;&Q_{1}+Q_{2}.\end{split}

We now focus on the term Q1Q_{1}. For some R>0R>0 large enough, one deduces that

(A.25) Q1(2π)n20+|eη(reiπ6)2s|x|2s(reiπ6)2|x|21|2ηsrn2+2s1|Hn2(1)(reiπ6)|𝑑r=(2π)n20R|eη(reiπ6)2s|x|2s(reiπ6)2|x|21|2ηsrn2+2s1|Hn2(1)(reiπ6)|𝑑r+(2π)n2R+|eη(reiπ6)2s|x|2s(reiπ6)2|x|21|2ηsrn2+2s1|Hn2(1)(reiπ6)|𝑑r=:Y1R+Y2R.\begin{split}Q_{1}&\leqslant(2\pi)^{\frac{n}{2}}\int_{0}^{+\infty}\left|e^{-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}-1\right|{2\eta sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ &=(2\pi)^{\frac{n}{2}}\int_{0}^{R}\left|e^{-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}-1\right|{2\eta sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ &\qquad+(2\pi)^{\frac{n}{2}}\int_{R}^{+\infty}\left|e^{-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}-1\right|{2\eta sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ &=:Y_{1}^{R}+Y_{2}^{R}.\end{split}

As for Y2RY_{2}^{R}, we notice that, utilizing (A.14), for every η(0,1)\eta\in(0,1),

Y2R(2π)n2R+2srn2+2s1|Hn2(1)(reiπ6)|𝑑r(2π)n2s2n+4s+4Γ(n2+2s)R2n.Y_{2}^{R}\leqslant(2\pi)^{\frac{n}{2}}\int_{R}^{+\infty}{2sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\leqslant\frac{(2\pi)^{\frac{n}{2}}s2^{n+4s+4}\Gamma(\frac{n}{2}+2s)}{R-2n}.

Therefore, for any ε>0\varepsilon>0, picking

(A.26) R1:=4(2π)n2s2n+4s+4Γ(n2+2s)ε+2n,R_{1}:=\frac{4(2\pi)^{\frac{n}{2}}s2^{n+4s+4}\Gamma(\frac{n}{2}+2s)}{\varepsilon}+2n,

one finds that

(A.27) Y2R1ε4.Y_{2}^{R_{1}}\leqslant\frac{\varepsilon}{4}.

As regards Y1R1Y_{1}^{R_{1}}, if R1R_{1} is as defined in (A.26) we have that

Y1R1(2π)n20R1(|eηr2scossπ3|x|2sr2cosπ3|x|2cos(ηr2scossπ3|x|2sr2cosπ3|x|2)1|+|sin(ηr2scossπ3|x|2sr2cosπ3|x|2)|2srn2+2s1|Hn2(1)(reiπ6)|dr.\begin{split}Y_{1}^{R_{1}}&\leqslant(2\pi)^{\frac{n}{2}}\int_{0}^{R_{1}}\bigg{(}\left|e^{-\eta\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}}\cos\left(-\eta\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}\right)-1\right|\\ &\qquad+\left|\sin\left(-\eta\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}\right)\right|{2sr^{\frac{n}{2}+2s-1}}\left|H^{(1)}_{\frac{n}{2}}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr.\end{split}

We notice that, for any r<R1r<R_{1}, |x|>R12|x|>R_{1}^{2} large enough and η(0,1)\eta\in(0,1),

(A.28) |eηr2scossπ3|x|2sr2cosπ3|x|2cos(ηr2scossπ3|x|2sr2cosπ3|x|2)1|+|sin(ηr2scossπ3|x|2s+r22|x|2)|1eR12scossπ3|x|2sR122|x|2cos(R12s|x|2s+R122|x|2)+sin(R12s|x|2s+R122|x|2),\begin{split}&\left|e^{-\eta\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}}\cos\left(-\eta\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}\right)-1\right|+\left|\sin\left(\eta\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}+\frac{r^{2}}{2|x|^{2}}\right)\right|\\ \leqslant\;&1-e^{-\frac{R_{1}^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{R_{1}^{2}}{2|x|^{2}}}\cos\left(\frac{R_{1}^{2s}}{|x|^{2s}}+\frac{R_{1}^{2}}{2|x|^{2}}\right)+\sin\left(\frac{R_{1}^{2s}}{|x|^{2s}}+\frac{R_{1}^{2}}{2|x|^{2}}\right),\end{split}

Utilizing (LABEL:R_1_estimate), we obtain that for every ε>0\varepsilon>0, there exists TR1,ε>0T_{R_{1},\varepsilon}>0 such that, for every |x|>TR1,ε|x|>T_{R_{1},\varepsilon},

1eR12scossπ3|x|2sR122|x|2cos(R12s|x|2s+R122|x|2)+sin(R12s|x|2s+R122|x|2)ε2n+4s+3(2π)n2Γ(n2+2s).1-e^{-\frac{R_{1}^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{R_{1}^{2}}{2|x|^{2}}}\cos\left(\frac{R_{1}^{2s}}{|x|^{2s}}+\frac{R_{1}^{2}}{2|x|^{2}}\right)+\sin\left(\frac{R_{1}^{2s}}{|x|^{2s}}+\frac{R_{1}^{2}}{2|x|^{2}}\right)\leqslant\frac{\varepsilon}{2^{n+4s+3}(2\pi)^{\frac{n}{2}}\Gamma(\frac{n}{2}+2s)}.

As a consequence of this and (A.14), one obtains that

(A.29) Y1R1<ε4.Y_{1}^{R_{1}}<\frac{\varepsilon}{4}.

Recalling (A.25), by combining (A.27) with (A.29), we deduce that for every ε>0\varepsilon>0, there exists M3>0M_{3}>0, depending only on nn, ss and ε\varepsilon, such that, for every |x|>M3|x|>M_{3},

Q1ε2.Q_{1}\leqslant\frac{\varepsilon}{2}.

Regarding Q2Q_{2}, thanks to (LABEL:Q1_Q2), we see that for every ε>0\varepsilon>0, there exists M4>0M_{4}>0, depending only on nn, ss and ε\varepsilon, such that, for every |x|>M4|x|>M_{4},

Q2ε2.Q_{2}\leqslant\frac{\varepsilon}{2}.

Taking M:=max{M3,M4}M:=\max\left\{M_{3},M_{4}\right\}, it follows from (LABEL:Q1_Q2) that, for every |x|>M|x|>M and η(0,1)\eta\in(0,1),

||x|n+2s(x,η,1)(2π)n2ReL12ηszn2+2s1Hn2(1)(z)𝑑z|<ε.\left||x|^{n+2s}\mathcal{H}(x,\eta,1)-(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{2\eta sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz\right|<\varepsilon.

That is,

lim|x|+|x|n+2s(x,η,1)=(2π)n2ReL12ηszn2+2s1Hn2(1)(z)𝑑z,\lim\limits_{|x|\to+\infty}|x|^{n+2s}\mathcal{H}(x,\eta,1)=(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{2\eta sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz,

and the limit is uniform with respect to η\eta.

As a consequence of this, by using the modified Bessel function of the third kind K(x)K(x) again (see e.g. [28, page 5]), it is not difficult to check that

lim|x|+|x|n+2s(x,η,1)=(2π)n2ReL22ηszn2+2s1Hn2(1)(z)𝑑z=2n+2sπn21ηssin(πs)Γ(n2+s)Γ(s).\begin{split}\lim\limits_{|x|\to+\infty}|x|^{n+2s}\mathcal{H}(x,\eta,1)&=(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{2}}{2\eta sz^{\frac{n}{2}+2s-1}}H^{(1)}_{\frac{n}{2}}\left(z\right)\,dz\\ &=2^{n+2s}\pi^{\frac{n}{2}-1}\eta s\sin(\pi s)\Gamma\left(\frac{n}{2}+s\right)\Gamma(s).\end{split}

Thus, we obtain (A.23), as desired. ∎

A.3. Bounds on the heat kernel

In this section, we establish the following result related to the upper and lower bounds on the heat kernel \mathcal{H} by using the two asymptotic formulae in Lemmata A.3 and A.4.

Lemma A.5.

Let n1n\geqslant 1 and s(0,1)s\in(0,1). Let \mathcal{H} be as defined in (1.3).

Then, there exist positive constants C1C_{1} and C2C_{2} such that

(A.30) (x,t)C1{t|x|n+2sts|x|n+2s}{tn2stn2};\mathcal{H}(x,t)\leqslant C_{1}\left\{\frac{t}{|x|^{n+2s}}\vee\frac{t^{s}}{|x|^{n+2s}}\right\}\wedge\left\{t^{-\frac{n}{2s}}\wedge t^{-\frac{n}{2}}\right\};
(A.31) and (x,t)C2{t|x|n+2sif  1<t<|x|2s;eπ|x|2ttn2if |x|2<t<|x|2s<1,\text{and }\quad\mathcal{H}(x,t)\geqslant C_{2}\begin{cases}\frac{t}{|x|^{n+2s}}\qquad&\mbox{if }\;1<t<|x|^{2s};\\ e^{\frac{\pi|x|^{2}}{t}}t^{-\frac{n}{2}}\qquad&\mbox{if }\;|x|^{2}<t<|x|^{2s}<1,\end{cases}

where ab:=min{a,b}a\wedge b:=\min\left\{a,b\right\} and ab:=max{a,b}a\vee b:=\max\left\{a,b\right\}.

Proof.

We observe that, by the definition of \mathcal{H} in (1.3), for every t>0t>0,

(A.32) (x,t)c1{tn2tn2s}\mathcal{H}(x,t)\leqslant c_{1}\left\{t^{-\frac{n}{2}}\wedge t^{-\frac{n}{2s}}\right\}

for some constant c1c_{1} depending only on nn and ss.

In addition, in the light of Lemma A.3, we know that there exists M>0M>0 independent of η\eta such that, for all η(0,1)\eta\in(0,1) and |x|>M|x|>M,

(A.33) α2<|x|n+2s(x,1,η)<2α,\frac{\alpha}{2}<|x|^{n+2s}\mathcal{H}(x,1,\eta)<2\alpha,

where α=2n+2sπn21ssin(πs)Γ(n2+s)Γ(s)\alpha=2^{n+2s}\pi^{\frac{n}{2}-1}s\sin(\pi s)\Gamma(\frac{n}{2}+s)\Gamma(s).

Furthermore, for every t(1,+)t\in(1,+\infty), one has that t11s(0,1)t^{1-\frac{1}{s}}\in(0,1). Hence, by combining (A.2) with (A.33), one concludes that

α2<|t12sx|n+2s(t12sx,1,t11s)<2αif |x|>Mt12s.\frac{\alpha}{2}<|t^{-\frac{1}{2s}}x|^{n+2s}\mathcal{H}(t^{-\frac{1}{2s}}x,1,t^{1-\frac{1}{s}})<2\alpha\quad\text{if }|x|>Mt^{\frac{1}{2s}}.

That is,

αtn2s+12|x|n+2s<(t12sx,1,t11s)<2αtn2s+1|x|n+2sif |x|>Mt12s.\frac{\alpha t^{\frac{n}{2s}+1}}{2|x|^{n+2s}}<\mathcal{H}(t^{-\frac{1}{2s}}x,1,t^{1-\frac{1}{s}})<\frac{2\alpha t^{\frac{n}{2s}+1}}{|x|^{n+2s}}\quad\text{if }|x|>Mt^{\frac{1}{2s}}.

Exploiting (A.2), one sees that, if t>1t>1 and |x|>Mt12s|x|>Mt^{\frac{1}{2s}},

(A.34) αt2|x|n+2s<(x,t)<2αt|x|n+2s.\frac{\alpha t}{2|x|^{n+2s}}<\mathcal{H}(x,t)<\frac{2\alpha t}{|x|^{n+2s}}.

Similarly, for every t(0,1)t\in(0,1), we have that t1s(0,1)t^{1-s}\in(0,1). In this way, using the nonnegativity of \mathcal{H}, by combining (A.2) with Lemma A.4, we know that there exists M>0M>0 independent of η\eta such that, for every η(0,1)\eta\in(0,1), t(0,1)t\in(0,1) and |x|>Mt12|x|>Mt^{\frac{1}{2}},

(A.35) 0(x,t)<2αts|x|n+2s.0\leqslant\mathcal{H}(x,t)<\frac{2\alpha t^{s}}{|x|^{n+2s}}.

Gathering (A.32), (A.34) and (A.35), we obtain (A.30), as desired.

Let us now estimate (A.31). We first claim that for any ε>0\varepsilon>0, there exists δ(0,1)\delta\in(0,1) such that, for every |x|<δ|x|<\delta and |x|2t|x|2s|x|^{2}\leqslant t\leqslant|x|^{2s},

(A.36) |(x,t)eπ|x|2ttn2ωnΓ(n2)2|<ε.\left|\frac{\mathcal{H}(x,t)}{e^{-\pi\frac{|x|^{2}}{t}}t^{-\frac{n}{2}}}-\frac{\omega_{n}\Gamma(\frac{n}{2})}{2}\right|<\varepsilon.

Indeed, from the definition of \mathcal{H} it follows that

(x,t)=net(|ξ|2s+|ξ|2)+2πixξ𝑑ξ=eπ|x|2tne(t12ξiπxt12)2et|ξ|2s𝑑ξ=eπ|x|2ttn2ne|yiπxt12|2et1s|y|2s𝑑y.\begin{split}&\mathcal{H}(x,t)=\int_{\mathbb{R}^{n}}e^{-t(|\xi|^{2s}+|\xi|^{2})+2\pi ix\cdot\xi}\,d\xi=e^{-\frac{\pi|x|^{2}}{t}}\int_{\mathbb{R}^{n}}e^{-\left(t^{\frac{1}{2}}\xi-i\pi xt^{-\frac{1}{2}}\right)^{2}}e^{-t|\xi|^{2s}}\,d\xi\\ &\qquad=e^{-\frac{\pi|x|^{2}}{t}}t^{-\frac{n}{2}}\int_{\mathbb{R}^{n}}e^{-|y-i\pi xt^{-\frac{1}{2}}|^{2}}e^{-t^{1-s}|y|^{2s}}\,dy.\end{split}

Accordingly, recalling the fact that |x|2t|x|2s|x|^{2}\leqslant t\leqslant|x|^{2s} and taking δ=(εωnΓ(n2+s)eπ2)12s(1s)\delta=\left(\frac{\varepsilon}{\omega_{n}\Gamma(\frac{n}{2}+s)e^{\pi^{2}}}\right)^{\frac{1}{2s(1-s)}}, one derives that

|(x,t)eπ|x|2ttn2ne|yiπxt12|2𝑑y|=|ne|yiπxt12|2(1et1s|y|2s)𝑑y|ne|y|2+π2t1s|y|2s𝑑yωnΓ(n2+s)t1seπ2ωnΓ(n2+s)|x|2s(1s)eπ2ε.\begin{split}&\left|\frac{\mathcal{H}(x,t)}{e^{-\pi\frac{|x|^{2}}{t}}t^{-\frac{n}{2}}}-\int_{\mathbb{R}^{n}}e^{-|y-i\pi xt^{-\frac{1}{2}}|^{2}}\,dy\right|=\left|\int_{\mathbb{R}^{n}}e^{-|y-i\pi xt^{-\frac{1}{2}}|^{2}}\left(1-e^{-t^{1-s}|y|^{2s}}\right)\,dy\right|\\ &\qquad\qquad\leqslant\int_{\mathbb{R}^{n}}e^{-|y|^{2}+\pi^{2}}t^{1-s}|y|^{2s}\,dy\leqslant\omega_{n}\Gamma\left(\frac{n}{2}+s\right)t^{1-s}e^{\pi^{2}}\\ &\qquad\qquad\leqslant\omega_{n}\Gamma\left(\frac{n}{2}+s\right)|x|^{2s(1-s)}e^{\pi^{2}}\leqslant\varepsilon.\end{split}

Subsequently, applying the Residue Theorem, we observe that

ne|yiπxt12|2𝑑y=niπxt12e|z|2𝑑z=ne|z|2𝑑z=ωnΓ(n2)2\int_{\mathbb{R}^{n}}e^{-|y-i\pi xt^{-\frac{1}{2}}|^{2}}\,dy=\int_{\mathbb{R}^{n}-i\pi xt^{-\frac{1}{2}}}e^{-|z|^{2}}\,dz=\int_{\mathbb{R}^{n}}e^{-|z|^{2}}\,dz=\frac{\omega_{n}\Gamma(\frac{n}{2})}{2}

and we obtain (A.36), as desired.

Taking now ε:=ωnΓ(n2)4\varepsilon:=\frac{\omega_{n}\Gamma(\frac{n}{2})}{4} in (A.36) and utilizing (A.34), we establish the desired result in (A.31). ∎

Appendix B Properties of the Bessel kernel

The main aim of this section is to prove Theorem 3.2 by exploiting the properties of the heat kernel \mathcal{H} defined in (1.3).

To start with, in view of the definition of 𝒦\mathcal{K}, we observe that, from the nonnegativity, radial symmetry and monotonicity of (,t)\mathcal{H}(\cdot,t) for all t>0t>0, it follows that the kernel 𝒦\mathcal{K} is nonnegative, radially symmetric and nonincreasing in r=|x|r=|x|.

B.1. Decay of the Bessel kernel

In this part, our goal is to employ the lower and upper bounds on the heat kernel \mathcal{H} (see e.g. (A.30) and (A.31)) to obtain the decay of 𝒦\mathcal{K}, as given by the following Lemmata B.1 and B.2.

Lemma B.1.

Let n1n\geqslant 1 and s(0,1)s\in(0,1). Let 𝒦\mathcal{K} be as defined in (1.4).

Then, if |x|1|x|\geqslant 1,

(B.1) 𝒦(x)c1|x|n+2s\mathcal{K}(x)\leqslant\frac{c_{1}}{|x|^{n+2s}}

and, if |x|1|x|\leqslant 1,

𝒦(x)c2{|x|2n if n3,1+|ln|x|| if n=2,1 if n=1,\mathcal{K}(x)\leqslant c_{2}\begin{cases}|x|^{2-n}\qquad&\text{ if }n\geqslant 3,\\ 1+|\ln|x||&\text{ if }n=2,\\ 1&\text{ if }n=1,\end{cases}

for some positive constants c1c_{1} and c2c_{2} depending on nn and s.s.

Proof.

From (A.30), we derive that

0(x,t)1C1{tn2st|x|n+2s}for all t>1,\displaystyle 0\leqslant\mathcal{H}(x,t)\leqslant\frac{1}{C_{1}}\left\{t^{-\frac{n}{2s}}\wedge\frac{t}{|x|^{n+2s}}\right\}\qquad{\mbox{for all~{}$t>1$,}}
and 0(x,t)1C1{tn2ts|x|n+2s}for all t(0,1).\displaystyle 0\leqslant\mathcal{H}(x,t)\leqslant\frac{1}{C_{1}}\left\{t^{-\frac{n}{2}}\wedge\frac{t^{s}}{|x|^{n+2s}}\right\}\qquad{\mbox{for all~{}$t\in(0,1)$.}}

From these formulas we conclude that, if |x|>1|x|>1,

(B.2) 𝒦(x)=1+et(x,t)𝑑t+01et(x,t)𝑑t1C1(1|x|2sett|x|n+2s𝑑t+|x|2s+ettn2s𝑑t+01etts|x|n+2s𝑑t)2Γ(2)+Γ(s+1)C1|x|n+2s\begin{split}\mathcal{K}(x)&=\int_{1}^{+\infty}e^{-t}\mathcal{H}(x,t)\,dt+\int_{0}^{1}e^{-t}\mathcal{H}(x,t)\,dt\\ &\leqslant\frac{1}{C_{1}}\left(\int_{1}^{\left|{x}\right|^{2s}}e^{-t}\frac{t}{|x|^{n+2s}}\,dt+\int_{\left|{x}\right|^{2s}}^{+\infty}e^{-t}t^{-\frac{n}{2s}}\,dt+\int_{0}^{1}e^{-t}\frac{t^{s}}{|x|^{n+2s}}\,dt\right)\\ &\leqslant\frac{2\Gamma(2)+\Gamma(s+1)}{C_{1}|x|^{n+2s}}\end{split}

and, if |x|1|x|\leqslant 1,

(B.3) 𝒦(x)=1+et(x,t)𝑑t+01et(x,t)𝑑t1C1(1+ettn2s𝑑t+0|x|2etts|x|n+2s𝑑t+|x|21ettn2𝑑t){n+2C1(n2)|x|n2 if n34(1+|ln|x||) if n=21+Γ(12)+|x| if n=1.\begin{split}\mathcal{K}(x)&=\int_{1}^{+\infty}e^{-t}\mathcal{H}(x,t)\,dt+\int_{0}^{1}e^{-t}\mathcal{H}(x,t)\,dt\\ &\leqslant\frac{1}{C_{1}}\left(\int_{1}^{+\infty}e^{-t}t^{-\frac{n}{2s}}\,dt+\int_{0}^{\left|{x}\right|^{2}}e^{-t}\frac{t^{s}}{|x|^{n+2s}}\,dt+\int_{\left|{x}\right|^{2}}^{1}e^{-t}t^{-\frac{n}{2}}\,dt\right)\\ &\leqslant\begin{cases}\frac{n+2}{C_{1}(n-2)|x|^{n-2}}\qquad\qquad&{\mbox{ if }}n\geqslant 3\\ 4(1+|\ln|x||)&{\mbox{ if }}n=2\\ 1+\Gamma\left(\frac{1}{2}\right)+|x|&{\mbox{ if }}n=1.\end{cases}\end{split}

By combining (B.2) with (B.3), we obtain (B.1). ∎

We shall make use of the nonnegativity of 𝒦\mathcal{K} and (A.31) to prove the strict positivity of 𝒦\mathcal{K} and a lower bound on the behaviour of 𝒦\mathcal{K}.

Lemma B.2.

Let n1n\geqslant 1 and s(0,1)s\in(0,1). Let 𝒦\mathcal{K} be as in (1.4).

Then, 𝒦\mathcal{K} is positive and

𝒦(x)c3|x|n+2s if |x|>1and𝒦(x)c4|x|n2 if |x|1\mathcal{K}(x)\geqslant\frac{c_{3}}{|x|^{n+2s}}\quad\text{ if }|x|>1\qquad\text{and}\qquad\mathcal{K}(x)\geqslant\frac{c_{4}}{|x|^{n-2}}\quad\text{ if }|x|\leqslant 1

for some constants c3c_{3} and c4c_{4} depending only on nn and s.s.

Proof.

From (A.31), we know that

(x,t)C2t|x|n+2sif  1<t<|x|2s.\mathcal{H}(x,t)\geqslant C_{2}\frac{t}{|x|^{n+2s}}\qquad\mbox{if }\;1<t<|x|^{2s}.

Recalling the fact that \mathcal{H} is nonnegative, one deduces that, for any |x|>2|x|>2,

(B.4) 𝒦(x)=0+et(x,t)𝑑tC21|x|2set(x,t)𝑑tC212ett|x|n+2s𝑑te2|x|n+2s.\mathcal{K}(x)=\int_{0}^{+\infty}e^{-t}\mathcal{H}(x,t)\,dt\geqslant C_{2}\int_{1}^{\left|{x}\right|^{2s}}e^{-t}\mathcal{H}(x,t)\,dt\geqslant C_{2}\int_{1}^{2}e^{-t}\frac{t}{|x|^{n+2s}}\,dt\geqslant\frac{e^{-2}}{|x|^{n+2s}}.

Moreover, in light of (A.31), for every |x|2<t<|x|2s<1|x|^{2}<t<|x|^{2s}<1,

(x,t)>C2eπ|x|2ttn2.{\mathcal{H}(x,t)}>C_{2}e^{-\pi\frac{|x|^{2}}{t}}t^{-\frac{n}{2}}.

As a consequence, by the definition of 𝒦,\mathcal{K}, one has that, for every |x|<12|x|<\frac{1}{2},

(B.5) 𝒦(x)|x|2|x|2sC2eπ|x|2ttn2et𝑑te1C2|x|22s1eπy|x|n+2yn22𝑑ye1C2|x|n21222s1eπyyn22𝑑y=cn,s|x|n2.\begin{split}&\mathcal{K}(x)\geqslant\int_{|x|^{2}}^{|x|^{2s}}C_{2}e^{-\pi\frac{|x|^{2}}{t}}t^{-\frac{n}{2}}e^{-t}\,dt\geqslant e^{-1}C_{2}\int_{|x|^{2-2s}}^{1}e^{-\pi y}|x|^{-n+2}y^{\frac{n}{2}-2}\,dy\\ &\qquad\qquad\geqslant\frac{e^{-1}C_{2}}{|x|^{n-2}}\int_{{\frac{1}{2}}^{2-2s}}^{1}e^{-\pi y}y^{\frac{n}{2}-2}\,dy=\frac{c_{n,s}}{|x|^{n-2}}.\end{split}

In addition, we recall that 𝒦0\mathcal{K}\geqslant 0, and 𝒦\mathcal{K} is nonincreasing in r=|x|r=|x|. According to (B.4) and (LABEL:low_bound), one concludes that 𝒦(x)>0\mathcal{K}(x)>0 for every xnx\in\mathbb{R}^{n}. ∎

B.2. Smoothness of the Bessel kernel

In this section, we focus on the smoothness of the kernel 𝒦\mathcal{K} given by (1.4). These properties of 𝒦\mathcal{K} are useful in the proofs of our regularity results.

Lemma B.3.

Let n1n\geqslant 1 and s(0,1)s\in(0,1). Let 𝒦\mathcal{K} be as in (1.4).

Then, there exists C>0C>0, depending only on nn and ss, such that, for all |x|1|x|\geqslant 1,

|𝒦(x)|C|x|n+2s+1and|D2𝒦(x)|C|x|n+2s+2.|\nabla\mathcal{K}(x)|\leqslant\frac{C}{|x|^{n+2s+1}}\quad\text{and}\quad|D^{2}\mathcal{K}(x)|\leqslant\frac{C}{|x|^{n+2s+2}}.
Proof.

Since 𝒦\mathcal{K} is radially symmetric, with a slight abuse of notation we write 𝒦(x)=𝒦(r)\mathcal{K}(x)=\mathcal{K}(r), with r=|x|r=|x|. Furthermore, recalling the definition of 𝒦\mathcal{K} and employing (A.4), we infer that

𝒦(r)=(2π)n2rn210+0+et(1+τ2s+τ2)τn2Jn21(rτ)𝑑τ𝑑t.\mathcal{K}(r)=\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}-1}}\int_{0}^{+\infty}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)\,d\tau\,dt.

Taking the derivative of the expression above with respect to rr, one finds that

(B.6) 𝒦(r)=(2π)n2(n2)2r𝒦(r)+(2π)n2rn20+0+et(1+τ2s+τ2)τn2Jn21(rτ)τr𝑑τ𝑑t:=I1(r)+I2(r).\begin{split}\mathcal{K}^{\prime}(r)=&\frac{-(2\pi)^{\frac{n}{2}}(n-2)}{2r}\mathcal{K}(r)+\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}\tau^{\frac{n}{2}}J^{\prime}_{\frac{n}{2}-1}(r\tau)\tau r\,d\tau\,dt\\ :=&I_{1}(r)+I_{2}(r).\end{split}

Concerning I1I_{1}, owing to (B.1), one has that, for every r>1r>1,

|I1(r)|c1(2π)n2|n2|2rn+2s+1.|I_{1}(r)|\leqslant\frac{c_{1}(2\pi)^{\frac{n}{2}}|n-2|}{2r^{n+2s+1}}.

We now estimate I2I_{2}. For this, integrating by parts in τ\tau, we see that

I2(r)=(2π)n2rn20+et(1+τ2s+τ2)τn2+1Jn21(rτ)|τ=0τ=+dt(2π)n2rn20+0+et(1+τ2s+τ2)Jn21(rτ)((n2+1)τn2t(2sτ2s+2τ2)τn2)𝑑τ𝑑t=n+22r𝒦(r)+(2π)n2rn20+0+et(1+τ2s+τ2)τn2Jn21(rτ)(2tsτ2s+2tτ2)𝑑τ𝑑t.\begin{split}I_{2}(r)&=\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}\tau^{\frac{n}{2}+1}J_{\frac{n}{2}-1}(r\tau)\bigg{|}_{\tau=0}^{\tau=+\infty}\,dt\\ &\quad-\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}J_{\frac{n}{2}-1}(r\tau)\left(\left(\frac{n}{2}+1\right)\tau^{\frac{n}{2}}-t(2s\tau^{2s}+2\tau^{2})\tau^{\frac{n}{2}}\right)\,d\tau\,dt\\ &=-\frac{n+2}{2r}\mathcal{K}(r)+\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)\left(2ts\tau^{2s}+2t\tau^{2}\right)\,d\tau\,dt.\end{split}

Furthermore, integrating by parts in tt,

2s(2π)n2rn20+0+et(1+τ2s+τ2)τn2Jn21(rτ)tτ2s𝑑τ𝑑t=(2π)n2rn20+tet(1+τ2)etτ2sτ2sτn2Jn21(rτ)2sτ2s|t=0t=+dτ+2s(2π)n2rn20+0+etτ2sτ2s(et(1+τ2)tet(1+τ2)(1+τ2))τn2+2sJn21(rτ)𝑑τ𝑑t=2s(2π)n2rn20+0+etτ2s(et(1+τ2)tet(1+τ2)(1+τ2))τn2Jn21(rτ)𝑑τ𝑑t=2sr𝒦(r)2s(2π)n2rn20+0+tet(1+τ2s+τ2)(1+τ2)τn2Jn21(rτ)𝑑τ𝑑t=2sr𝒦(r)2sr0+tet(r,t,t)𝑑t2s(2π)n2rn20+0+et(1+τ2s+τ2)τn2Jn21(rτ)tτ2𝑑τ𝑑t.\begin{split}&2s\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)t\tau^{2s}\,d\tau\,dt\\ =\;&\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\frac{te^{-t(1+\tau^{2})}e^{-t\tau^{2s}}}{-\tau^{2s}}\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)2s\tau^{2s}\bigg{|}_{t=0}^{t=+\infty}\,d\tau\\ &\quad+2s\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}\frac{e^{-t\tau^{2s}}}{\tau^{2s}}\left(e^{-t(1+\tau^{2})}-te^{-t(1+\tau^{2})}(1+\tau^{2})\right)\tau^{\frac{n}{2}+2s}J_{\frac{n}{2}-1}(r\tau)\,d\tau\,dt\\ =\;&2s\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}{e^{-t\tau^{2s}}}\left(e^{-t(1+\tau^{2})}-te^{-t(1+\tau^{2})}(1+\tau^{2})\right)\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)\,d\tau\,dt\\ =\;&\frac{2s}{r}\mathcal{K}(r)-2s\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}te^{-t(1+\tau^{2s}+\tau^{2})}(1+\tau^{2})\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)\,d\tau\,dt\\ =\;&\frac{2s}{r}\mathcal{K}(r)-\frac{2s}{r}\int_{0}^{+\infty}te^{-t}\mathcal{H}(r,t,t)\,dt-2s\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)t\tau^{2}\,d\tau\,dt.\end{split}

As a consequence,

(B.7) (2π)n2rn20+0+et(1+τ2s+τ2)τn2Jn21(rτ)(2tsτ2s+2tτ2)𝑑τ𝑑t=(2s2)(2π)n2rn20+0+et(1+τ2s+τ2)τn2Jn21(rτ)tτ2s𝑑τ𝑑t+2r𝒦(r)2r0+tet(r,t,t)𝑑t=:(2s2)𝒢(r)+2r𝒦(r)2r0+tet(r,t,t)𝑑t.\begin{split}&\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)\left(2ts\tau^{2s}+2t\tau^{2}\right)\,d\tau\,dt\\ =\;&(2s-2)\frac{(2\pi)^{\frac{n}{2}}}{r^{\frac{n}{2}}}\int_{0}^{+\infty}\int_{0}^{+\infty}e^{-t(1+\tau^{2s}+\tau^{2})}\tau^{\frac{n}{2}}J_{\frac{n}{2}-1}(r\tau)t\tau^{2s}\,d\tau\,dt+\frac{2}{r}\mathcal{K}(r)-\frac{2}{r}\int_{0}^{+\infty}te^{-t}\mathcal{H}(r,t,t)\,dt\\ =:\;&(2s-2)\mathcal{G}(r)+\frac{2}{r}\mathcal{K}(r)-\frac{2}{r}\int_{0}^{+\infty}te^{-t}\mathcal{H}(r,t,t)\,dt.\end{split}

From (A.1) and (A.30), it follows that, for any r>1r>1,

(B.8) 0+tet(r,t,t)𝑑t=0+tet(r,t)𝑑tcrn+2s\int_{0}^{+\infty}te^{-t}\mathcal{H}(r,t,t)\,dt=\int_{0}^{+\infty}te^{-t}\mathcal{H}(r,t)\,dt\leqslant\frac{c}{r^{n+2s}}

where the constant cc depends only on nn and ss.

Let us now estimate the integral 𝒢(r)\mathcal{G}(r). For this, for every t1t_{1}, t2>0t_{2}>0 and xn{0}x\in\mathbb{R}^{n}\setminus\left\{0\right\} we set

Φ(x,t1,t2):=1|x|net1|ξ|2st2|ξ|2|ξ|2se2πixξ𝑑ξ.\Phi(x,t_{1},t_{2}):=\frac{1}{|x|}\int_{\mathbb{R}^{n}}e^{-t_{1}|\xi|^{2s}-t_{2}|\xi|^{2}}|\xi|^{2s}e^{2\pi ix\cdot\xi}\,d\xi.

Using again the Fourier Inversion Theorem for radial functions, we see that

(B.9) Φ(x,t1,t2)=(2π)n2|x|n20+et1r2st2r2rn2+2sJn21(|x|r)𝑑r,\Phi(x,t_{1},t_{2})=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{\frac{n}{2}}}\int_{0}^{+\infty}e^{-t_{1}r^{2s}-t_{2}r^{2}}r^{\frac{n}{2}+2s}J_{\frac{n}{2}-1}(|x|r)\,dr,

where Jn21J_{\frac{n}{2}-1} denotes the Bessel function of first kind of order n21\frac{n}{2}-1. Furthermore, we observe that Φ\Phi satisfies the following scaling properties:

(B.10) Φ(x,t,t)=tn2s1Φ(t12sx,1,t11s)=tn2sΦ(t12x,t1s,1).\Phi(x,t,t)=t^{-\frac{n}{2s}-1}\Phi(t^{-\frac{1}{2s}}x,1,t^{1-\frac{1}{s}})=t^{-\frac{n}{2}-s}\Phi(t^{-\frac{1}{2}}x,t^{1-s},1).

By the definition of 𝒢\mathcal{G} in (B.7), we see that

𝒢(r)=1+tetΦ(r,t,t)dt+01tetΦ(r,t,t)dt=:𝒢1(r)+𝒢2(r).\mathcal{G}(r)=\int_{1}^{+\infty}te^{-t}\Phi(r,t,t)\,dt+\int_{0}^{1}te^{-t}\Phi(r,t,t)\,dt=:\mathcal{G}_{1}(r)+\mathcal{G}_{2}(r).

We now focus on estimating 𝒢1\mathcal{G}_{1} and 𝒢2\mathcal{G}_{2}.

Step 1. We estimate 𝒢1\mathcal{G}_{1}. Let us start by calculating the value of lim|x|+|x|n+2s+1Φ(x,1,η)\lim\limits_{|x|\to+\infty}|x|^{n+2s+1}\Phi(x,1,\eta) for every η(0,1)\eta\in(0,1). From (B.9), we infer that

Φ(x,1,η)=(2π)n2|x|n+2s+10+et2s|x|2sηt2|x|2tn2+2sJn21(t)𝑑t.\Phi(x,1,\eta)=\frac{(2\pi)^{\frac{n}{2}}}{|x|^{n+2s+1}}\int_{0}^{+\infty}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}t^{\frac{n}{2}+2s}J_{\frac{n}{2}-1}(t)\,dt.

Thus, owing to the fact that Hn21(1)H^{(1)}_{\frac{n}{2}-1} is the Bessel function of the third kind, we have that

|x|n+2s+1Φ(x,1,η)=(2π)n2Re0+et2s|x|2sηt2|x|2tn2+2sHn21(1)(t)𝑑t.|x|^{n+2s+1}\Phi(x,1,\eta)={(2\pi)^{\frac{n}{2}}}\text{Re}\int_{0}^{+\infty}e^{-\frac{t^{2s}}{|x|^{2s}}-\eta\frac{t^{2}}{|x|^{2}}}t^{\frac{n}{2}+2s}H^{(1)}_{\frac{n}{2}-1}(t)\,dt.

Using again the Residue Theorem, we deduce that

|x|n+2s+1Φ(x,1,η)=(2π)n2ReL1ez2s|x|2sηz2|x|2zn2+2sHn21(1)(z)𝑑z\begin{split}|x|^{n+2s+1}\Phi(x,1,\eta)&={(2\pi)^{\frac{n}{2}}}\text{Re}\int_{L_{1}}e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}z^{\frac{n}{2}+2s}H^{(1)}_{\frac{n}{2}-1}(z)\,dz\end{split}

where L1L_{1} is defined in (A.7).

We now claim that for any ε>0\varepsilon>0 there exists M>0M>0 independent of η\eta such that, for every |x|>M|x|>M and η(0,1)\eta\in(0,1),

||x|n+2s+1Φ(x,1,η)(2π)n2ReL1zn2+2sHn21(1)(z)𝑑z|<ε.\left|\,|x|^{n+2s+1}\Phi(x,1,\eta)-(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{z^{\frac{n}{2}+2s}}H^{(1)}_{\frac{n}{2}-1}\left(z\right)\,dz\right|<\varepsilon.

For this purpose, we notice that, for R>0R>0 large enough,

||x|n+2s+1Φ(x,1,η)(2π)n2ReL1zn2+2sHn21(1)(z)𝑑z|=|(2π)n2ReL1(ez2s|x|2sηz2|x|21)zn2+2sHn21(1)(z)𝑑z|(2π)n20R|e(reiπ6)2s|x|2sη(reiπ6)2|x|21|rn2+2s|Hn21(1)(reiπ6)|𝑑r+(2π)n2R+|e(reiπ6)2s|x|2sη(reiπ6)2|x|21|rn2+2s|Hn21(1)(reiπ6)|𝑑r=:1R+2R.\begin{split}&\left|\,|x|^{n+2s+1}\Phi(x,1,\eta)-(2\pi)^{\frac{n}{2}}\text{Re}\int_{L^{1}}{z^{\frac{n}{2}+2s}}H^{(1)}_{\frac{n}{2}-1}\left(z\right)\,dz\right|\\ =\;&\left|{(2\pi)^{\frac{n}{2}}}\text{Re}\int_{L_{1}}\left(e^{-\frac{z^{2s}}{|x|^{2s}}-\eta\frac{z^{2}}{|x|^{2}}}-1\right){z^{\frac{n}{2}+2s}}H^{(1)}_{\frac{n}{2}-1}(z)\,dz\right|\\ \leqslant\;&(2\pi)^{\frac{n}{2}}\int_{0}^{R}\left|e^{-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}-1\right|{r^{\frac{n}{2}+2s}}\left|H^{(1)}_{\frac{n}{2}-1}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ &\qquad+(2\pi)^{\frac{n}{2}}\int_{R}^{+\infty}\left|e^{-\frac{\left(re^{i\frac{\pi}{6}}\right)^{2s}}{|x|^{2s}}-\eta\frac{\left(re^{i\frac{\pi}{6}}\right)^{2}}{|x|^{2}}}-1\right|{r^{\frac{n}{2}+2s}}\left|H^{(1)}_{\frac{n}{2}-1}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ =:\;&\mathcal{B}^{R}_{1}+\mathcal{B}^{R}_{2}.\end{split}

To estimate 2R\mathcal{B}_{2}^{R}, we note that, by the definition of Hn21(1)(z)H^{(1)}_{\frac{n}{2}-1}\left(z\right) (see e.g. [28, page 21]),

2R(2π)n2R+rn2+2s|Hn21(1)(reiπ6)|𝑑r(2π)n220+e(n21)tR+er4etrn2+2s𝑑r𝑑t(2π)n220+e(n21)tR+er4(t+1)rn2+2s𝑑r𝑑t(2π)n220+e(n21R4)tR+er4rn2+2s𝑑r𝑑t(2π)n22n+4s+5Γ(n2+2s+1)R2n+2.\begin{split}&\mathcal{B}^{R}_{2}\leqslant(2\pi)^{\frac{n}{2}}\int_{R}^{+\infty}{r^{\frac{n}{2}+2s}}\left|H^{(1)}_{\frac{n}{2}-1}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\leqslant(2\pi)^{\frac{n}{2}}2\int_{0}^{+\infty}e^{\left(\frac{n}{2}-1\right)t}\int_{R}^{+\infty}e^{-\frac{r}{4}e^{t}}{r^{\frac{n}{2}+2s}}\,dr\,dt\\ &\qquad\leqslant(2\pi)^{\frac{n}{2}}2\int_{0}^{+\infty}e^{\left(\frac{n}{2}-1\right)t}\int_{R}^{+\infty}e^{-\frac{r}{4}(t+1)}{r^{\frac{n}{2}+2s}}\,dr\,dt\\ &\qquad\leqslant(2\pi)^{\frac{n}{2}}2\int_{0}^{+\infty}e^{\left(\frac{n}{2}-1-\frac{R}{4}\right)t}\int_{R}^{+\infty}e^{-\frac{r}{4}}{r^{\frac{n}{2}+2s}}\,dr\,dt\leqslant\frac{(2\pi)^{\frac{n}{2}}2^{n+4s+5}\Gamma(\frac{n}{2}+2s+1)}{R-2n+2}.\end{split}

As a consequence, given ε>0\varepsilon>0, taking

(B.11) R3:=2(2π)n22n+4s+5Γ(n2+2s+1)ε+2n2,R_{3}:=\frac{2(2\pi)^{\frac{n}{2}}2^{n+4s+5}\Gamma(\frac{n}{2}+2s+1)}{\varepsilon}+2n-2,

one has that

(B.12) 2R3ε2.\mathcal{B}^{R_{3}}_{2}\leqslant\frac{\varepsilon}{2}.

We now consider 1R3\mathcal{B}^{R_{3}}_{1} where R3R_{3} is defined in (B.11). Recalling (A.14), we have that

1R3(2π)n20R3(|er2scossπ3|x|2sηr2cosπ3|x|2cos(r2scossπ3|x|2sηr2cosπ3|x|2)1|+|sin(r2scossπ3|x|2sηr2cosπ3|x|2)|)rn2+2s|H(1)n21(reiπ6)|dr(2π)n2Γ(n2+2s+1)2n+4s+2s+1(1eR32scossπ3|x|2sR322|x|2cos(R32s|x|2s+R322|x|2)+sin(R32s|x|2s+R322|x|2)).\begin{split}\mathcal{B}^{R_{3}}_{1}&\leqslant(2\pi)^{\frac{n}{2}}\int_{0}^{R_{3}}\bigg{(}\left|e^{-\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\eta\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}}\cos\left(-\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\eta\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}\right)-1\right|\\ &\qquad+\left|\sin\left(-\frac{r^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\eta\frac{r^{2}\cos\frac{\pi}{3}}{|x|^{2}}\right)\right|\bigg{)}{r^{\frac{n}{2}+2s}}\left|H^{(1)}_{\frac{n}{2}-1}\left(re^{i\frac{\pi}{6}}\right)\right|\,dr\\ &\leqslant\frac{(2\pi)^{\frac{n}{2}}\Gamma\left(\frac{n}{2}+2s+1\right)2^{n+4s+2}}{s+1}\left(1-e^{-\frac{R_{3}^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{R_{3}^{2}}{2|x|^{2}}}\cos\left(\frac{R_{3}^{2s}}{|x|^{2s}}+\frac{R_{3}^{2}}{2|x|^{2}}\right)+\sin\left(\frac{R_{3}^{2s}}{|x|^{2s}}+\frac{R_{3}^{2}}{2|x|^{2}}\right)\right).\end{split}

It follows that for any ε>0\varepsilon>0 there exists TR3,ε>0T_{R_{3},\varepsilon}>0 such that, for every |x|>TR3,ε|x|>T_{R_{3},\varepsilon},

1eR32scossπ3|x|2sR322|x|2cos(R32s|x|2s+R322|x|2)+sin(R32s|x|2s+R322|x|2)(s+1)ε2n+4s+3(2π)n2Γ(n2+2s+1),1-e^{-\frac{R_{3}^{2s}\cos\frac{s\pi}{3}}{|x|^{2s}}-\frac{R_{3}^{2}}{2|x|^{2}}}\cos\left(\frac{R_{3}^{2s}}{|x|^{2s}}+\frac{R_{3}^{2}}{2|x|^{2}}\right)+\sin\left(\frac{R_{3}^{2s}}{|x|^{2s}}+\frac{R_{3}^{2}}{2|x|^{2}}\right)\leqslant\frac{(s+1)\varepsilon}{2^{n+4s+3}(2\pi)^{\frac{n}{2}}\Gamma(\frac{n}{2}+2s+1)},

which implies that

1R3<ε2.\mathcal{B}^{R_{3}}_{1}<\frac{\varepsilon}{2}.

By combining this and (B.12), we deduce that for any ε>0\varepsilon>0 there exists M>0M>0, depending only on nn, ss and ε\varepsilon, such that, for every |x|>M|x|>M,

||x|n+2s+1Φ(x,1,η)(2π)n2ReL1zn2+2sHn21(1)(z)𝑑z|ε.\left|\,|x|^{n+2s+1}\Phi(x,1,\eta)-(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{z^{\frac{n}{2}+2s}}H^{(1)}_{\frac{n}{2}-1}\left(z\right)\,dz\right|\leqslant{\varepsilon}.

That is,

lim|x|+|x|n+2s+1Φ(x,1,η)=(2π)n2ReL1zn2+2s,Hn21(1)(z)dz\lim\limits_{|x|\to+\infty}|x|^{n+2s+1}\Phi(x,1,\eta)=(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{1}}{z^{\frac{n}{2}+2s}},H^{(1)}_{\frac{n}{2}-1}\left(z\right)\,dz

where the limit is independent of η\eta.

Moreover, applying the Residue Theorem again to the region composed of L1L^{1} and the straight line L2L_{2} given in (A.22), and using the modified Bessel function of the third kind K(x)K(x) (see e.g. [28, page 5]), it follows that

(B.13) lim|x|+|x|n+2s+1Φ(x,1,η)=(2π)n2ReL2zn2+2sHn21(1)(z)𝑑z=(2π)n2Re0+irn2+2seiπn4eiπsHn21(1)(ir)𝑑r=(2π)n2Re0+rn2+2sieiπs2πKn21(r)𝑑r=2n+2sπn21sin(πs)Γ(n2+s)Γ(s+1):=β.\begin{split}\lim\limits_{|x|\to+\infty}|x|^{n+2s+1}\Phi(x,1,\eta)&=(2\pi)^{\frac{n}{2}}\text{Re}\int_{L_{2}}{z^{\frac{n}{2}+2s}}H^{(1)}_{\frac{n}{2}-1}\left(z\right)\,dz\\ &=(2\pi)^{\frac{n}{2}}\text{Re}\int_{0}^{+\infty}ir^{\frac{n}{2}+2s}e^{i\frac{\pi n}{4}}e^{i\pi s}H^{(1)}_{\frac{n}{2}-1}\left(ir\right)\,dr\\ &=(2\pi)^{\frac{n}{2}}\text{Re}\int_{0}^{+\infty}r^{\frac{n}{2}+2s}ie^{i\pi s}\frac{2}{\pi}K_{\frac{n}{2}-1}(r)\,dr\\ &=-2^{n+2s}\pi^{\frac{n}{2}-1}\sin(\pi s)\Gamma\left(\frac{n}{2}+s\right)\Gamma(s+1):=-\beta.\end{split}

From this, we infer that there exists M1>0M_{1}>0, depending only on nn and ss, such that, for every |x|>M1|x|>M_{1},

β2<|x|n+2s|Φ(x,1,η)|<2β.\frac{\beta}{2}<|x|^{n+2s}|\Phi(x,1,\eta)|<2\beta.

From (B.10), we deduce that, if t>1t>1 and |x|>M1t12s|x|>M_{1}t^{\frac{1}{2s}},

(B.14) βt12s2|x|n+2s+1<|Φ(x,t,t)|<2βt12s|x|n+2s+1.\frac{\beta t^{\frac{1}{2s}}}{2|x|^{n+2s+1}}<|\Phi(x,t,t)|<\frac{2\beta t^{\frac{1}{2s}}}{|x|^{n+2s+1}}.

Moreover, according to the definition of Φ(x,t,t)\Phi(x,t,t), one has that

(B.15) 0|Φ(x,t,t)|Λn,s|x|1(tn2s1tn2s).0\leqslant|\Phi(x,t,t)|\leqslant\Lambda_{n,s}|x|^{-1}\left(t^{-\frac{n}{2s}-1}\wedge t^{-\frac{n}{2}-s}\right).

Owing to (B.14) and (B.15), we conclude that, for every |x|>M1|x|>M_{1},

(B.16) |𝒢1(x)|1+tet|Φ(x,t,t)|𝑑t1|xM1|2stet2βt12s|x|n+2s+1𝑑t+|xM1|2s+tettn2s1|x|1Λn,s𝑑t2βΓ(2+12s)|x|n+2s+1+Λn,sΓ(2)M1n+2s|x|n+2s+1\begin{split}|\mathcal{G}_{1}(x)|&\leqslant\int_{1}^{+\infty}te^{-t}|\Phi(x,t,t)|\,dt\leqslant\int_{1}^{\left|\frac{x}{M_{1}}\right|^{2s}}te^{-t}\frac{2\beta t^{\frac{1}{2s}}}{|x|^{n+2s+1}}\,dt+\int_{\left|\frac{x}{M_{1}}\right|^{2s}}^{+\infty}te^{-t}t^{-\frac{n}{2s}-1}|x|^{-1}\Lambda_{n,s}\,dt\\ &\leqslant\frac{2\beta\Gamma(2+\frac{1}{2s})}{|x|^{n+2s+1}}+\frac{\Lambda_{n,s}\Gamma(2)M_{1}^{n+2s}}{|x|^{n+2s+1}}\end{split}

and, for every 1<|x|M11<|x|\leqslant M_{1},

|𝒢1(x)|1+tet|Φ(x,t,t)|𝑑t|xM1|2s+tettn2s1Λn,s|x|1𝑑tΛn,sΓ(2)M1n+2s|x|n+2s+1.\begin{split}|\mathcal{G}_{1}(x)|\leqslant\int_{1}^{+\infty}te^{-t}|\Phi(x,t,t)|\,dt\leqslant\int_{\left|\frac{x}{M_{1}}\right|^{2s}}^{+\infty}te^{-t}t^{-\frac{n}{2s}-1}\Lambda_{n,s}|x|^{-1}\,dt\leqslant\frac{\Lambda_{n,s}\Gamma(2)M_{1}^{n+2s}}{|x|^{n+2s+1}}.\end{split}

As a consequence of this and (B.16), we have that, for every |x|>1|x|>1,

(B.17) |𝒢1(x)|1|x|n+2s+1(2βΓ(2+12s)+Λn,sΓ(2)M1n+2s).|\mathcal{G}_{1}(x)|\leqslant\frac{1}{|x|^{n+2s+1}}\left(2\beta\,\Gamma\left(2+\frac{1}{2s}\right)+\Lambda_{n,s}\Gamma(2)M_{1}^{n+2s}\right).

Step 2. We now estimate the 𝒢2\mathcal{G}_{2}. To this end, we remark that

lim|x|+|x|n+2s+1Φ(x,η,1)=2n+2sπn21sin(πs)Γ(n2+s)Γ(s+1)=β,\lim\limits_{|x|\to+\infty}|x|^{n+2s+1}\Phi(x,\eta,1)=-2^{n+2s}\pi^{\frac{n}{2}-1}\sin(\pi s)\Gamma\left(\frac{n}{2}+s\right)\Gamma(s+1)=-\beta,

where β\beta is defined in (B.13) and the limit is independent of η(0,1)\eta\in(0,1).

Recalling (B.10), we see that there exists M2>0M_{2}>0, depending only on nn and ss, such that, for all t(0,1)t\in(0,1) and |x|>M2t12|x|>M_{2}t^{\frac{1}{2}},

(B.18) βt122|x|n+2s+1<|Φ(x,t,t)|<2βt12|x|n+2s+1.\frac{\beta t^{\frac{1}{2}}}{2|x|^{n+2s+1}}<|\Phi(x,t,t)|<\frac{2\beta t^{\frac{1}{2}}}{|x|^{n+2s+1}}.

Owing to (B.15) and  (B.18), we conclude that, if 1<|x|<M21<|x|<M_{2},

(B.19) |𝒢2(x)|01tet|Φ(x,t,t)|𝑑t0|xM2|2tet2βt12|x|n+2s+1𝑑t+|xM2|21tettn2s|x|1Λn,s𝑑t2βΓ(52)|x|n+2s+1+Λn,sΓ(2)M2n+2s|x|n+2s+1\begin{split}&|\mathcal{G}_{2}(x)|\leqslant\int_{0}^{1}te^{-t}|\Phi(x,t,t)|\,dt\leqslant\int_{0}^{\left|\frac{x}{M_{2}}\right|^{2}}te^{-t}\frac{2\beta t^{\frac{1}{2}}}{|x|^{n+2s+1}}\,dt+\int_{\left|\frac{x}{M_{2}}\right|^{2}}^{1}te^{-t}t^{-\frac{n}{2}-s}|x|^{-1}\Lambda_{n,s}\,dt\\ &\qquad\qquad\leqslant\frac{2\beta\Gamma(\frac{5}{2})}{|x|^{n+2s+1}}+\frac{\Lambda_{n,s}\Gamma(2)M_{2}^{n+2s}}{|x|^{n+2s+1}}\end{split}

and, if |x|>M2|x|>M_{2},

|𝒢2(x)|01tet|Φ(x,t,t)|𝑑t0|xM2|2tet2βt12|x|n+2s+1𝑑t2βΓ(52)|x|n+2s+1.\begin{split}|\mathcal{G}_{2}(x)|\leqslant\int_{0}^{1}te^{-t}|\Phi(x,t,t)|\,dt\leqslant\int_{0}^{\left|\frac{x}{M_{2}}\right|^{2}}te^{-t}\frac{2\beta t^{\frac{1}{2}}}{|x|^{n+2s+1}}\,dt\leqslant\frac{2\beta\Gamma(\frac{5}{2})}{|x|^{n+2s+1}}.\end{split}

By (LABEL:phi_2_x>m), one sees that, for every |x|>1|x|>1,

(B.20) |𝒢2(x)|1|x|n+2s+1(2βΓ(52)+Λ2Γ(2)M2n+2s).|\mathcal{G}_{2}(x)|\leqslant\frac{1}{|x|^{n+2s+1}}\left(2\beta\,\Gamma\left(\frac{5}{2}\right)+\Lambda_{2}\Gamma(2)M_{2}^{n+2s}\right).

Step 3. Combining (B.17) and (B.20), we conclude that, for every |x|>1|x|>1,

(B.21) 𝒢(x)cn,s|x|n+2s+1.\mathcal{G}(x)\leqslant\frac{c_{n,s}}{|x|^{n+2s+1}}.

Finally, recalling (B.6), we have that

𝒦(r)=(2π)n2(n2)2r𝒦(r)n+22r𝒦(r)+(2s2)𝒢(r)+2r𝒦(r)2r0+tet(r,t,t)𝑑t.\mathcal{K}^{\prime}(r)=\frac{-(2\pi)^{\frac{n}{2}}(n-2)}{2r}\mathcal{K}(r)-\frac{n+2}{2r}\mathcal{K}(r)+(2s-2)\mathcal{G}(r)+\frac{2}{r}\mathcal{K}(r)-\frac{2}{r}\int_{0}^{+\infty}te^{-t}\mathcal{H}(r,t,t)\,dt.

Owing to the decay of 𝒦\mathcal{K}, by combining (B.8) with (B.21), we deduce that, for every |x|>1|x|>1,

|𝒦(x)||𝒦(r)|Crn+2s+1=C|x|n+2s+1|\nabla\mathcal{K}(x)|\leqslant|\mathcal{K}^{\prime}(r)|\leqslant\frac{C}{r^{n+2s+1}}=\frac{C}{|x|^{n+2s+1}}

for some constant CC depending only on nn and ss.

Using similar arguments, we infer that, for every |x|>1|x|>1,

|D2𝒦(x)|C|x|n+2s+2|D^{2}\mathcal{K}(x)|\leqslant\frac{C}{|x|^{n+2s+2}}

for some constant CC depending only on nn and ss. ∎

B.3. Proof of Theorem 3.2

We observe that, gathering together the results of Sections B.1 and B.2, we obtain the desired claims in (a)-(d) of Theorem 3.2. Thus, it suffices to prove property (e) of Theorem 3.2. To this end, we introduce the following two ancillary results:

Lemma B.4.

Let ϕ:n\phi:\mathbb{R}^{n}\to\mathbb{R} be in the Schwartz space and t>0t>0. Then, the function

n×n(x,ξ)et(|ξ|2s+|ξ|2)+2πixξϕ(x)\mathbb{R}^{n}\times\mathbb{R}^{n}\ni(x,\xi)\mapsto e^{-t(|\xi|^{2s}+|\xi|^{2})+2\pi ix\cdot\xi}\,\phi(x)

belongs to L1(n×n,)L^{1}(\mathbb{R}^{n}\times\mathbb{R}^{n},\,\mathbb{C}).

Proof.

We have that

n×n|et(|ξ|2s+|ξ|2)+2πixξϕ(x)|𝑑x𝑑ξ=n×net(|ξ|2s+|ξ|2)|ϕ(x)|𝑑x𝑑ξ=ϕL1(n)net(|ξ|2s+|ξ|2)𝑑ξ,\iint_{\mathbb{R}^{n}\times\mathbb{R}^{n}}\left|e^{-t(|\xi|^{2s}+|\xi|^{2})+2\pi ix\cdot\xi}\,\phi(x)\right|\,dx\,d\xi=\iint_{\mathbb{R}^{n}\times\mathbb{R}^{n}}e^{-t(|\xi|^{2s}+|\xi|^{2})}\,|\phi(x)|\,dx\,d\xi=\|\phi\|_{L^{1}(\mathbb{R}^{n})}\int_{\mathbb{R}^{n}}e^{-t(|\xi|^{2s}+|\xi|^{2})}\,d\xi,

which is finite. ∎

Lemma B.5.

Let Φ:n\Phi:\mathbb{R}^{n}\to\mathbb{C} be in the Schwartz space. Then, the function

n×(0,+)(ξ,t)et(1+|ξ|2s+|ξ|2)Φ(ξ)\mathbb{R}^{n}\times(0,+\infty)\ni(\xi,t)\mapsto e^{-t(1+|\xi|^{2s}+|\xi|^{2})}\,\Phi(\xi)

belongs to L1(n×(0,+),)L^{1}(\mathbb{R}^{n}\times(0,+\infty),\,\mathbb{C}) and

(B.22) n×(0,+)et(1+|ξ|2s+|ξ|2)Φ(ξ)𝑑ξ𝑑t=nΦ(ξ)1+|ξ|2s+|ξ|2𝑑ξ.\iint_{\mathbb{R}^{n}\times(0,+\infty)}e^{-t(1+|\xi|^{2s}+|\xi|^{2})}\,\Phi(\xi)\,d\xi\,dt=\int_{\mathbb{R}^{n}}\frac{\Phi(\xi)}{1+|\xi|^{2s}+|\xi|^{2}}\,d\xi.
Proof.

We have that

0+et(1+|ξ|2s+|ξ|2)𝑑t=11+|ξ|2s+|ξ|2.\int_{0}^{+\infty}e^{-t(1+|\xi|^{2s}+|\xi|^{2})}\,dt=\frac{1}{1+|\xi|^{2s}+|\xi|^{2}}.

Therefore, by Fubini-Tonelli’s Theorem,

n×(0,+)|et(1+|ξ|2s+|ξ|2)Φ(ξ)|𝑑ξ𝑑t=n×(0,+)et(1+|ξ|2s+|ξ|2)|Φ(ξ)|𝑑ξ𝑑t=n|Φ(ξ)|1+|ξ|2s+|ξ|2𝑑ξ,\iint_{\mathbb{R}^{n}\times(0,+\infty)}\left|e^{-t(1+|\xi|^{2s}+|\xi|^{2})}\,\Phi(\xi)\right|\,d\xi\,dt=\iint_{\mathbb{R}^{n}\times(0,+\infty)}e^{-t(1+|\xi|^{2s}+|\xi|^{2})}\,|\Phi(\xi)|\,d\xi\,dt=\int_{\mathbb{R}^{n}}\frac{|\Phi(\xi)|}{1+|\xi|^{2s}+|\xi|^{2}}\,d\xi,

which is finite. ∎

With this preparatory work, we can now check that the Bessel kernel is the fundamental solution of the mixed order operator Δ+(Δ)s-\Delta+(-\Delta)^{s}, as clarified by the following result:

Lemma B.6.

The Fourier transform of 𝒦{\mathcal{K}} equals 11+|ξ|2s+|ξ|2\frac{1}{1+|\xi|^{2s}+|\xi|^{2}} in the sense of distribution.

More explicitly, for every ϕ:n\phi:\mathbb{R}^{n}\to\mathbb{R} in the Schwartz space, we have that

(B.23) n𝒦(x)ϕ(x)𝑑x=nϕ^(ξ)1+|ξ|2s+|ξ|2𝑑ξ.\int_{\mathbb{R}^{n}}{\mathcal{K}}(x)\,\phi(x)\,dx=\int_{\mathbb{R}^{n}}\frac{\hat{\phi}(\xi)}{1+|\xi|^{2s}+|\xi|^{2}}\,d\xi.
Proof.

By (1.4) and Fubini-Tonelli’s Theorem (whose validity is a consequence here of Lemma B.2), we have that

n𝒦(x)ϕ(x)𝑑x=n(0+et(x,t)𝑑t)ϕ(x)𝑑x=0+(net(x,t)ϕ(x)𝑑x)𝑑t.\int_{\mathbb{R}^{n}}{\mathcal{K}}(x)\,\phi(x)\,dx=\int_{\mathbb{R}^{n}}\left(\int_{0}^{+\infty}e^{-t}\,\mathcal{H}(x,t)\,dt\right)\phi(x)\,dx=\int_{0}^{+\infty}\left(\int_{\mathbb{R}^{n}}e^{-t}\,\mathcal{H}(x,t)\,\phi(x)\,dx\right)\,dt.

Hence, by (1.3),

n𝒦(x)ϕ(x)𝑑x=0+(n(net(1+|ξ|2s+|ξ|2)+2πixξϕ(x)𝑑ξ)𝑑x)𝑑t.\int_{\mathbb{R}^{n}}{\mathcal{K}}(x)\,\phi(x)\,dx=\int_{0}^{+\infty}\left(\int_{\mathbb{R}^{n}}\left(\int_{\mathbb{R}^{n}}e^{-t(1+|\xi|^{2s}+|\xi|^{2})+2\pi ix\cdot\xi}\,\phi(x)\,d\xi\right)\,dx\right)\,dt.

This and Fubini-Tonelli’s Theorem (whose validity is a consequence here of Lemma B.4) yield that

n𝒦(x)ϕ(x)𝑑x=0+(n(net(1+|ξ|2s+|ξ|2)+2πixξϕ(x)𝑑x)𝑑ξ)𝑑t=0+(net(1+|ξ|2s+|ξ|2)ϕ^(ξ)𝑑ξ)𝑑t.\begin{split}\int_{\mathbb{R}^{n}}{\mathcal{K}}(x)\,\phi(x)\,dx&=\int_{0}^{+\infty}\left(\int_{\mathbb{R}^{n}}\left(\int_{\mathbb{R}^{n}}e^{-t(1+|\xi|^{2s}+|\xi|^{2})+2\pi ix\cdot\xi}\,\phi(x)\,dx\right)\,d\xi\right)\,dt\\ &=\int_{0}^{+\infty}\left(\int_{\mathbb{R}^{n}}e^{-t(1+|\xi|^{2s}+|\xi|^{2})}\,\hat{\phi}(\xi)\,d\xi\right)\,dt.\end{split}

From this and (B.22) (utilized here with Φ:=ϕ^\Phi:=\hat{\phi}) we arrive at (B.23), as desired. ∎

Proof of (e) of Theorem 3.2.

Let q[1,+)q\in[1,+\infty) and fLq(n)f\in L^{q}(\mathbb{R}^{n}). Since 𝒮(n)\mathcal{S}(\mathbb{R}^{n}) is dense in Lq(n)L^{q}(\mathbb{R}^{n}), there exists a sequence fm𝒮(n)f_{m}\in\mathcal{S}(\mathbb{R}^{n}) such that

(B.24) fmfin Lq(n).f_{m}\rightarrow f\quad\text{in }L^{q}(\mathbb{R}^{n}).

From Lemma B.6, it follows that the function um:=𝒦fmu_{m}:=\mathcal{K}\ast f_{m} is a solution of

Δum+(Δ)sum+um=fm in n.-\Delta u_{m}+(-\Delta)^{s}u_{m}+u_{m}=f_{m}\quad\text{ in }\mathbb{R}^{n}.

Namely, for every φCc(n)\varphi\in C_{c}^{\infty}(\mathbb{R}^{n}),

(B.25) nΔφum+n(Δ)sφum+numφ=nfmφ.\int_{\mathbb{R}^{n}}-\Delta\varphi\,u_{m}+\int_{\mathbb{R}^{n}}(-\Delta)^{s}\varphi\,u_{m}+\int_{\mathbb{R}^{n}}u_{m}\,\varphi=\int_{\mathbb{R}^{n}}f_{m}\,\varphi.

Moreover, owing to (B.24) and taking into account the fact that 𝒦L1(n)\mathcal{K}\in L^{1}(\mathbb{R}^{n}), we infer that umu:=𝒦fu_{m}\rightarrow u:=\mathcal{K}\ast f in Lq(n).L^{q}(\mathbb{R}^{n}).

By passing to the limit in (B.25), we conclude that the function u=𝒦fu=\mathcal{K}\ast f is a solution of

Δu+(Δ)su+u=f in n.-\Delta u+(-\Delta)^{s}u+u=f\quad\text{ in }\mathbb{R}^{n}.

As a consequence of this, by Lemmata A.2, B.1, B.2 and B.3, we obtain (e) of Theorem 3.2. ∎

References

  • [17] C. J. Amick and J. F. Toland. Uniqueness and related analytic properties for the Benjamin-Ono equation—a nonlinear Neumann problem in the plane. Acta Math., 167(1-2):107–126, 1991. ISSN 0001-5962. URL https://doi.org/10.1007/BF02392447.
  • [18] N. H. Bingham, C. M. Goldie, and J. L. Teugels. Regular variation. Number 27. Cambridge university press, 1989.
  • [19] R. M. Blumenthal and R. K. Getoor. Some theorems on stable processes. Trans. Amer. Math. Soc., 95:263–273, 1960. ISSN 0002-9947. URL https://doi.org/10.2307/1993291.
  • [20] S. Bochner and K. Chandrasekharan. Fourier Transforms. Annals of Mathematics Studies, No. 19. Princeton University Press, Princeton, NJ; Oxford University Press, London, 1949. ix+219 pp.
  • [21] K. Bogdan, T. Grzywny, and M. Ryznar. Density and tails of unimodal convolution semigroups. J. Funct. Anal., 266(6):3543–3571, 2014. ISSN 0022-1236.
  • [22] H. Brezis. Functional analysis, Sobolev spaces and partial differential equations. Universitext. Springer, New York, 2011. ISBN 978-0-387-70913-0. xiv+599 pp.
  • [23] V. Coti Zelati and P. H. Rabinowitz. Homoclinic type solutions for a semilinear elliptic PDE on 𝐑n{\bf R}^{n}. Comm. Pure Appl. Math., 45(10):1217–1269, 1992. ISSN 0010-3640. URL https://doi.org/10.1002/cpa.3160451002.
  • [24] W. Cygan, T. Grzywny, and B. Trojan. Asymptotic behavior of densities of unimodal convolution semigroups. Trans. Amer. Math. Soc., 369(8):5623–5644, 2017. ISSN 0002-9947. URL https://doi.org/10.1090/tran/6830.
  • [25] J. Dávila, M. del Pino, S. Dipierro, and E. Valdinoci. Concentration phenomena for the nonlocal Schrödinger equation with Dirichlet datum. Anal. PDE, 8(5):1165–1235, 2015. ISSN 2157-5045. URL https://doi.org/10.2140/apde.2015.8.1165.
  • [26] J. Dávila, M. del Pino, and J. Wei. Concentrating standing waves for the fractional nonlinear Schrödinger equation. J. Differential Equations, 256(2):858–892, 2014. ISSN 0022-0396. URL https://doi.org/10.1016/j.jde.2013.10.006.
  • [27] S. Dipierro and E. Valdinoci. Elliptic Partial Differential Equations from an Elementary Viewpoint. A Fresh Glance at the Classical Theory. World Scientific, 2024.
  • [28] A. Erdélyi, W. Magnus, F. Oberhettinger, and F. G. Tricomi. Higher transcendental functions. Vols. I, II. McGraw-Hill Book Co., Inc., New York-Toronto-London, 1953. xxvi+302, xvii+396 pp. Based, in part, on notes left by Harry Bateman.
  • [29] M. M. Fall and E. Valdinoci. Uniqueness and nondegeneracy of positive solutions of (Δ)su+u=up(-\Delta)^{s}u+u=u^{p} in N\mathbb{R}^{N} when ss is close to 1. Comm. Math. Phys., 329(1):383–404, 2014. ISSN 0010-3616. URL https://doi.org/10.1007/s00220-014-1919-y.
  • [30] P. Felmer, A. Quaas, and J. Tan. Positive solutions of the nonlinear Schrödinger equation with the fractional Laplacian. Proc. Roy. Soc. Edinburgh Sect. A, 142(6):1237–1262, 2012. ISSN 0308-2105. URL https://doi.org/10.1017/S0308210511000746.
  • [31] R. Feynman and A. Hibbs. Quantum mechanics and path integrals. McGraw-Hill, New York, 1965.
  • [32] A. Floer and A. Weinstein. Nonspreading wave packets for the cubic Schrödinger equation with a bounded potential. J. Funct. Anal., 69(3):397–408, 1986. ISSN 0022-1236. URL https://doi.org/10.1016/0022-1236(86)90096-0.
  • [33] R. L. Frank and E. Lenzmann. Uniqueness of non-linear ground states for fractional Laplacians in \mathbb{R}. Acta Math., 210(2):261–318, 2013. ISSN 0001-5962. URL https://doi.org/10.1007/s11511-013-0095-9.
  • [34] R. L. Frank, E. Lenzmann, and L. Silvestre. Uniqueness of radial solutions for the fractional Laplacian. Comm. Pure Appl. Math., 69(9):1671–1726, 2016. ISSN 0010-3640. URL https://doi.org/10.1002/cpa.21591.
  • [35] D. Gilbarg and N. S. Trudinger. Elliptic partial differential equations of second order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. ISBN 3-540-41160-7. xiv+517 pp. Reprint of the 1998 edition.
  • [36] M. K. Kwong. Uniqueness of positive solutions of Δuu+up=0\Delta u-u+u^{p}=0 in 𝐑n{\bf R}^{n}. Arch. Rational Mech. Anal., 105(3):243–266, 1989. ISSN 0003-9527. URL https://doi.org/10.1007/BF00251502.
  • [37] N. Laskin. Fractional quantum mechanics and Lévy path integrals. Phys. Lett. A, 268(4-6):298–305, 2000. ISSN 0375-9601. URL https://doi.org/10.1016/S0375-9601(00)00201-2.
  • [38] N. Laskin. Fractional Schrödinger equation. Phys. Rev. E (3), 66(5):056108, 7, 2002. ISSN 1539-3755. URL https://doi.org/10.1103/PhysRevE.66.056108.
  • [39] N. Laskin. Principles of fractional quantum mechanics. In Fractional dynamics, pages 393–427. World Sci. Publ., Hackensack, NJ, 2012.
  • [40] J. Mawhin and M. Willem. Critical point theory and Hamiltonian systems, volume 74 of Applied Mathematical Sciences. Springer-Verlag, New York, 1989. ISBN 0-387-96908-X. xiv+277 pp. URL https://doi.org/10.1007/978-1-4757-2061-7.
  • [41] K. McLeod. Uniqueness of positive radial solutions of Δu+f(u)=0\Delta u+f(u)=0 in 𝐑n{\bf R}^{n}. II. Trans. Amer. Math. Soc., 339(2):495–505, 1993. ISSN 0002-9947. URL https://doi.org/10.2307/2154282.
  • [42] K. McLeod and J. Serrin. Uniqueness of positive radial solutions of Δu+f(u)=0\Delta u+f(u)=0 in 𝐑n{\bf R}^{n}. Arch. Rational Mech. Anal., 99(2):115–145, 1987. ISSN 0003-9527. URL https://doi.org/10.1007/BF00275874.
  • [43] P. H. Rabinowitz. On a class of nonlinear Schrödinger equations. Z. Angew. Math. Phys., 43(2):270–291, 1992. ISSN 0044-2275. URL https://doi.org/10.1007/BF00946631.
  • [44] M. Rao, R. Song, and Z. Vondraček. Green function estimates and Harnack inequality for subordinate Brownian motions. Potential Anal., 25(1):1–27, 2006. ISSN 0926-2601. URL https://doi.org/10.1007/s11118-005-9003-z.
  • [45] R. Song and Z. Vondraček. Parabolic Harnack inequality for the mixture of Brownian motion and stable process. Tohoku Math. J. (2), 59(1):1–19, 2007. ISSN 0040-8735. URL http://projecteuclid.org/euclid.tmj/1176734744.
  • [46] E. M. Stein. Singular integrals and differentiability properties of functions. Princeton Mathematical Series, No. 30. Princeton University Press, Princeton, NJ, 1970. xiv+290 pp.
  • [47] X. Su, E. Valdinoci, Y. Wei, and J. Zhang. On some regularity properties of mixed local and nonlocal elliptic equations. 2023. Preprint.
  • [48] X. Wang. On concentration of positive bound states of nonlinear Schrödinger equations. Comm. Math. Phys., 153(2):229–244, 1993. ISSN 0010-3616. URL http://projecteuclid.org/euclid.cmp/1104252679.