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

Compact embeddings for fractional super and sub harmonic functions with radial symmetry

Jacopo Bellazzini Jacopo Bellazzini
Dipartimento di Matematica
Università di Pisa
Largo B. Pontecorvo 5, 56100 Pisa, Italy
 and  Vladimir Georgiev V. Georgiev
Dipartimento di Matematica Università di Pisa Largo B. Pontecorvo 5, 56100 Pisa, Italy
and
Faculty of Science and Engineering
Waseda University
3-4-1, Okubo, Shinjuku-ku, Tokyo 169-8555
Japan and Institute of Mathematics and Informatics, Bulgarian Academy of Sciences, Acad. Georgi Bonchev Str., Block 8, 1113 Sofia, Bulgaria
Abstract.

We prove compactness of the embeddings in Sobolev spaces for fractional super and sub harmonic functions with radial symmetry. The main tool is a pointwise decay for radially symmetric functions belonging to a function space defined by finite homogeneous Sobolev norm together with finite L2L^{2} norm of the Riesz potentials. As a byproduct we prove also existence of maximizers for the interpolation inequalities in Sobolev spaces for radially symmetric fractional super and sub harmonic functions.

Key words and phrases:
Interpolation inequalities, fractional Sobolev inequality, Riesz potential, radial symmetry, compact embeddings
2010 Mathematics Subject Classification:
Primary 46E35; Secondary 39B62
The authors thank N. Visciglia and L. Forcella for the reading of a preliminary version of the paper. J.B. and V.G. were partially supported by “Problemi stazionari e di evoluzione nelle equazioni di campo non-lineari dispersive” of GNAMPA 2020 and by the project PRIN 2020XB3EFL by the Italian Ministry of Universities and Research. V.G. was partially supported by by the Top Global University Project, Waseda University, by the University of Pisa, Project PRA 2018 49 and by Institute of Mathematics and Informatics, Bulgarian Academy of Sciences.

1. Introduction

The classical embedding in Sobolev spaces HS(d)H˙r(d)H^{S}(\mathbb{R}^{d})\subset\dot{H}^{r}(\mathbb{R}^{d}) for 0rS0\leq r\leq S follows from the interpolation inequality in homogeneous Sobolev spaces

(1.1) DrφLp(d)C(r,S,p,d)φL2(d)1θDSφL2(d)θ,,\|D^{r}\varphi\|_{L^{p}(\mathbb{R}^{d})}\leq C(r,S,p,d)\,\|\varphi\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\,\|D^{S}\varphi\|_{L^{2}(\mathbb{R}^{d})}^{\theta},\,,

where φHS(d)\varphi\in H^{S}(\mathbb{R}^{d}) and DsφD^{s}\varphi is defined by

(Dsφ^)(ξ)=|ξ|sφ^(ξ).(\widehat{D^{s}\varphi})(\xi)=|\xi|^{s}\widehat{\varphi}(\xi).

The inequality (1.1) holds, see [5], Corollary 1.5 in [11], [6] or Theorem 2.44 in [1] provided that

  • 1p=12+rθSd,\frac{1}{p}=\frac{1}{2}+\frac{r-\theta S}{d},

  • rSθ1\frac{r}{S}\leq\theta\leq 1,

  • 0<rS0<r\leq S, p>1p>1.

We notice that at the endpoint case p=2p=2, corresponding to θ=rS\theta=\frac{r}{S}, we have

(1.2) DrφL2(d)C(r,S,2,d)φL2(d)1rSDSφL2(d)rS,φHS(d),\|D^{r}\varphi\|_{L^{2}(\mathbb{R}^{d})}\leq C(r,S,2,d)\|\varphi\|_{L^{2}(\mathbb{R}^{d})}^{1-\frac{r}{S}}\,\|D^{S}\varphi\|_{L^{2}(\mathbb{R}^{d})}^{\frac{r}{S}},\qquad\forall\varphi\in H^{S}(\mathbb{R}^{d}),

and hence the embedding HSH˙rH^{S}\subset\dot{H}^{r} for 0rS0\leq r\leq S is just a consequence of (1.2). If we look at the endpoint cases θ=rS\theta=\frac{r}{S} and θ=1\theta=1 in (1.1) we obtain that the range of exponents pp without any symmetry and positivity assumption fulfills

(1.3) p[2,2dd2(Sr)]\displaystyle p\in[2,\frac{2d}{d-2(S-r)}] if Sr<d2,\displaystyle\text{if }S-r<\frac{d}{2},
p[2,)\displaystyle p\in[2,\infty) if Srd2.\displaystyle\text{if }S-r\geq\frac{d}{2}.

We remark that the lower endpoint does not depend on dimension dd.

Moreover, looking at (1.2), it is easy to prove that the best constant in (1.2) is C(r,S,2,d)=1C(r,S,2,d)=1. Indeed from Hölder’s inequality in frequency applied to l.h.s. of (1.2) we get C(r,S,2,d)1C(r,S,2,d)\leq 1 and calling An={ξd s.t. 11n<|ξ|<1+1n}A_{n}=\left\{\xi\in\mathbb{R}^{d}\text{ s.t. }1-\frac{1}{n}<|\xi|<1+\frac{1}{n}\right\} it suffices to consider a sequence φn\varphi_{n} such that φ^n(ξ)=𝟙An(ξ)\hat{\varphi}_{n}(\xi)=\mathbbm{1}_{A_{n}}(\xi) to prove that C(r,S,2,d)=1C(r,S,2,d)=1.

In the sequel we consider r,S,dr,S,d as fixed quantities and we aim to study the range of pp such that (1.1) holds in case we restrict to radially symmetric functions φ\varphi in HS(d)H^{S}(\mathbb{R}^{d}) such that DrφD^{r}\varphi is not only radially symmetric but also either positive or negative.

We introduce the notation for 0<r<s0<r<s

(1.4) H˙rads(d):={φH˙s(d),φ=φ(|x|)},\dot{H}^{s}_{rad}(\mathbb{R}^{d}):=\{\varphi\in\dot{H}^{s}(\mathbb{R}^{d}),\ \ \varphi=\varphi(|x|)\},
(1.5) Hrads(d):={φHs(d),φ=φ(|x|)},H^{s}_{rad}(\mathbb{R}^{d}):=\{\varphi\in H^{s}(\mathbb{R}^{d}),\ \ \varphi=\varphi(|x|)\},
(1.6) Hrad,+s,r(d):={φHrads(d),Drφ0},H^{s,r}_{rad,+}(\mathbb{R}^{d}):=\{\varphi\in H^{s}_{rad}(\mathbb{R}^{d}),\ \ \ D^{r}\varphi\geq 0\},
(1.7) Hrad,s,r(d):={φHrads(d),Drφ0}.H^{s,r}_{rad,-}(\mathbb{R}^{d}):=\{\varphi\in H^{s}_{rad}(\mathbb{R}^{d}),\ \ \ D^{r}\varphi\leq 0\}.

By the relation (Δφ^)(ξ)=4π2|ξ|2φ^(ξ)=4π2(D2φ^)(ξ)(\widehat{-\Delta\varphi})(\xi)=4\pi^{2}|\xi|^{2}\widehat{\varphi}(\xi)=4\pi^{2}(\widehat{D^{2}\varphi})(\xi) we shall emphasize that Hrad,+s,2(d)H^{s,2}_{rad,+}(\mathbb{R}^{d}) corresponds to the set of superharmonic radially symmetric functions belonging to Hs(d)H^{s}(\mathbb{R}^{d}) while Hrad,s,2(d)H^{s,2}_{rad,-}(\mathbb{R}^{d}) corresponds to the set of subharmonic radially symmetric functions belonging to Hs(d)H^{s}(\mathbb{R}^{d}). In the sequel we will call when r2r\neq 2 fractional superharmonic radially symmetric functions belonging to Hs(d)H^{s}(\mathbb{R}^{d}) the functions belonging to Hrad,+s,r(d)H^{s,r}_{rad,+}(\mathbb{R}^{d}) and fractional subharmonic radially symmetric functions belonging to Hs(d)H^{s}(\mathbb{R}^{d}) the functions belonging to Hrad,s,r(d)H^{s,r}_{rad,-}(\mathbb{R}^{d}).

The main questions we are interesting in are the following ones:

Question A: Can we find appropriate values of (r,S)(r,S) such that pp can be chosen below 22 in (1.1) for fractional superharmonic (resp. subharmonic) functions belonging to Hrad,+S,r(d)H^{S,r}_{rad,+}(\mathbb{R}^{d})?

Question B: If the answer of question A is positive, then can we expect a compact embedding of type

(1.8) Hrad,+S,r(d)H˙r(d)?H^{S,r}_{rad,+}(\mathbb{R}^{d})\subset\subset\dot{H}^{r}(\mathbb{R}^{d})?

In the sequel we will consider the case φHrad,+S,r(d)\varphi\in H^{S,r}_{rad,+}(\mathbb{R}^{d}) but all the results are still valid if we consider φHrad,S,r(d)\varphi\in H^{S,r}_{rad,-}(\mathbb{R}^{d}). The first result of the paper gives a positive answer to Question A.

Theorem 1.1.

Let d2d\geq 2 and 12<r<min(d2,S12)\frac{1}{2}<r<\min(\frac{d}{2},S-\frac{1}{2}), then

(1.9) DrφLp(d)Crad,+(r,S,p,d)φL2(d)1θDSφL2(d)θ,\displaystyle\|D^{r}\varphi\|_{L^{p}(\mathbb{R}^{d})}\leq C_{rad,+}(r,S,p,d)\,\|\varphi\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\,\|D^{S}\varphi\|_{L^{2}(\mathbb{R}^{d})}^{\theta},
φHrad,+S,r(d),\displaystyle\forall\varphi\ \in H^{S,r}_{rad,+}(\mathbb{R}^{d})\,,

with

(1.10) p(p0,2dd2(Sr)]\displaystyle p\in(p_{0},\frac{2d}{d-2(S-r)}] if Sr<d2,\displaystyle\text{if }S-r<\frac{d}{2},
(1.11) p(p0,)\displaystyle p\in(p_{0},\infty) if Srd2,\displaystyle\text{if }S-r\geq\frac{d}{2},

with θ\theta fixed by the scaling equation

1p=12+rθSd,\frac{1}{p}=\frac{1}{2}+\frac{r-\theta S}{d},

and p0<2p_{0}<2 is given by

p0=d2r+2(Sr)(d1)((Sr)12)(d2r)+2(Sr)(d1).p_{0}=\frac{d-2r+2(S-r)(d-1)}{-((S-r)-\frac{1}{2})(d-2r)+2(S-r)(d-1)}.
Remark 1.1.

Theorem 1.1 holds also for φHrad,S,r(d)\varphi\ \in H^{S,r}_{rad,-}(\mathbb{R}^{d}). The crucial condition is that DrφD^{r}\varphi does not change sign.

The constant Crad,+(r,S,p,d)C_{rad,+}(r,S,p,d) in (1.9) is defined as best constant in case of functions belonging to Hrad,+S,r(d)H^{S,r}_{rad,+}(\mathbb{R}^{d}).

The fact that p0<2p_{0}<2 in the above Theorem implies DrφLpD^{r}\varphi\in L^{p} with p(p0,2)p\in(p_{0},2) and this allows us to obtain also a positive answer to Question B.

Theorem 1.2.

Let d2d\geq 2 and 12<r0<min(d2,S12)\frac{1}{2}<r_{0}<\min(\frac{d}{2},S-\frac{1}{2}), then the embedding

Hrad,+S,r0(d)H˙radr(d)H^{S,r_{0}}_{rad,+}(\mathbb{R}^{d})\subset\subset\dot{H}^{r}_{rad}(\mathbb{R}^{d})

is compact for any 0<r<S.0<r<S.

Remark 1.2.

Theorem 1.2 holds also in Hrad,S,r0(d)H^{S,r_{0}}_{rad,-}(\mathbb{R}^{d}). Clearly the main difficult in Theorem 1.2 is to prove that the embedding Hrad,+S,r0(d)H˙radr0(d)H^{S,r_{0}}_{rad,+}(\mathbb{R}^{d})\subset\subset\dot{H}^{r_{0}}_{rad}(\mathbb{R}^{d}) is compact, the compactness for rr0r\neq r_{0} will follow by interpolation.

As a second byproduct we have also the following result concerning the existence of maximizers for the interpolation inequality (1.9) in case p=2p=2.

Theorem 1.3.

Let d2d\geq 2 and 12<r<min(d2,S12)\frac{1}{2}<r<\min(\frac{d}{2},S-\frac{1}{2}) then

DrφL2(d)Crad,+(r,S,2,d)φL2(d)1rSDSφL2(d)rS,\|D^{r}\varphi\|_{L^{2}(\mathbb{R}^{d})}\leq C_{rad,+}(r,S,2,d)\|\varphi\|_{L^{2}(\mathbb{R}^{d})}^{1-\frac{r}{S}}\,\|D^{S}\varphi\|_{L^{2}(\mathbb{R}^{d})}^{\frac{r}{S}},
φHrad,+S,r(d),\forall\varphi\ \in H^{S,r}_{rad,+}(\mathbb{R}^{d}),

and the best constant Crad,+(r,S,2,d)C_{rad,+}(r,S,2,d) is attained and Crad,+(r,S,2,d)<1.C_{rad,+}(r,S,2,d)<1.

The strategy to prove Theorem 1.1 and as a byproduct, the compactness result given in Theorem 1.2, it to rewrite (1.1) involving L2L^{2} norms of Riesz potentials when 0<r<d0<r<d. By defining u=Drφu=D^{r}\varphi we obtain

(1.12) uLp(d)C(α,s,p,d)1|x|αuL2(d)1θDsuL2(d)θ\|u\|_{L^{p}(\mathbb{R}^{d})}\leq C(\alpha,s,p,d)\,\|\frac{1}{|x|^{\alpha}}\star u\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\,\|D^{s}u\|_{L^{2}(\mathbb{R}^{d})}^{\theta}

where α=dr\alpha=d-r, s=Srs=S-r. With respect to the new variables α,s\alpha,s we get without any symmetry or positivity assumption

(1.13) p[2,2dd2s]\displaystyle p\in[2,\frac{2d}{d-2s}] if s<d2,\displaystyle\text{if }s<\frac{d}{2},
p[2,)\displaystyle p\in[2,\infty) if sd2.\displaystyle\text{if }s\geq\frac{d}{2}.

If one considers functions fulfilling Drφ=u0D^{r}\varphi=u\geq 0, inequality (1.12) is hence equivalent to the following inequality

(1.14) uLp(d)C(α,s,p,d)1|x|α|u|L2(d)1θDsuL2(d)θ\|u\|_{L^{p}(\mathbb{R}^{d})}\leq C(\alpha,s,p,d)\,\|\frac{1}{|x|^{\alpha}}\star|u|\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\,\|D^{s}u\|_{L^{2}(\mathbb{R}^{d})}^{\theta}

considering |u||u| instead of uu in the Riesz potential. The strategy is hence to prove that the radial symmetry increases the range of pp for which (1.14) holds and therefore as byproduct the range of pp for which (1.12) holds when Drφ=uD^{r}\varphi=u is positive and radially symmetric (resp. negative). In particular we will show that the lower endpoint is allowed to be below p=2p=2. A reasonable idea to prove that the lower endpoint exponent in (1.14) decreases with radial symmetry is to look at a suitable pointwise decay in the spirit of the Strauss lemma [17] (see also [15, 16] for Besov and Lizorkin-Triebel classes). In our context where two terms are present, the Sobolev norm and the Riesz potential involving |u||u|, we have been inspired by [13] where the case s=1s=1 in (1.14) has been studied (see also [4] and [3]). For our purposes the fact that ss is in general not integer makes however the strategy completetly different from the one in [13] and we need to estimate the decay of the high/low frequency part of the function to compute the decay. To this aim we compute the high frequency part using the explicit formula for the Fourier transform for radially symmetric function involving Bessel functions, in the spirit of [7], while we use a weighted L1L^{1} norm to compute the decay for the low frequency part. The importance of a pointwise decay for the low frequency part involving weighted LpL^{p} norms goes back to [8] and we need to adapt it to our case in order to involve the Riesz potential. Here is the step where positivity is crucial. Indeed if one is interested to show a scaling invariant weighted inequality as

(1.15) d|u(x)||x|γ𝑑xC1|x|α|u|L2(d)\int_{\mathbb{R}^{d}}\frac{|u(x)|}{|x|^{\gamma}}dx\leq C\|\frac{1}{|x|^{\alpha}}\star|u|\|_{L^{2}(\mathbb{R}^{d})}

a scaling argument forces the exponent γ\gamma to verify the relation γ=αd2\gamma=\alpha-\frac{d}{2}. Unfortunately (1.15) cannot hold in the whole Euclidean space following a general argument that goes back to [13] and [14]. However a scaling invariant inequality like (1.15) restricted on balls and on complementary of balls is enough for our purposes. Eventually, using all these tools, we are able to compute a pointwise decay that allows the lower endpoint for (1.14) to be below the threshold p=2p=2. Computed the pointwise decay we will follow the argument in [4] to estimate the lower endpoint for fractional superharmonic (resp. subharmonic) radially symmetric functions.

Concerning the compactness we prove that taking a bounded sequence φnHrad,+S,r\varphi_{n}\in H^{S,r}_{rad,+} then φnφ\varphi_{n}\to\varphi H˙r\dot{H}^{r} with r>0r>0. Our strategy is to prove the smallness of Dr(φnφ)L2(Bρ)\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2}(B_{\rho})} and Dr(φnφ)L2(Bρc)\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2}(B_{\rho}^{c})} for suitable choice of the ball Bρ.B_{\rho}. For the first term we use Rellich-Kondrachov argument combined with commutator estimates, while for the exterior domain we use the crucial fact that Dr(φnφ)D^{r}(\varphi_{n}-\varphi) is in Lp(|x|>ρ)L^{p}(|x|>\rho) for some p(1,2).p\in(1,2).

Looking at the case r=0r=0, by Rellich-Kondrachov we have φnφL2(Bρ)=o(1)\|\varphi_{n}-\varphi\|_{L^{2}(B_{\rho})}=o(1), however we can not obtain the smallness in the complementary BρcB_{\rho}^{c} of the ball so the requirement r>0r>0 seems to be optimal.

It is interesting to look at the lower endpoint exponent p0p_{0} given in Theorem 1.1 in case we consider radially symmetric superharmonic (or subharmonic), namely when r=2r=2. In this case the condition 12<r<min(d2,S12)\frac{1}{2}<r<\min(\frac{d}{2},S-\frac{1}{2}), imposes to consider the case d5d\geq 5 and S>52S>\frac{5}{2}. As an example we show on Figure 1 the graph of the function p0(S)p_{0}(S), that now is only a function of SS, in lowest dimensional case d=5d=5 that is a branch of hyperbola with asymptote p=limSp0(S)=8/7.p_{\infty}=\lim_{S\rightarrow\infty}p_{0}(S)=8/7. It is interesting how the regularity improves the lower endpoint p0(S)p_{0}(S).

As a final comment we notice that for d2d\geq 2 if D2φ0D^{2}\varphi\geq 0 then D34φ=D54(D2φ)0D^{\frac{3}{4}}\varphi=D^{-{\frac{5}{4}}}\left(D^{2}\ \varphi\right)\geq 0 then, taking r0=3/4r_{0}=3/4 and using the positivity of the Riesz kernel of D54,D^{-{\frac{5}{4}}}, we apply Theorem 1.2 and we get the following corollary.

Corollary 1.1.

Let φn\varphi_{n} be a sequence of radially symmetric superharmonic functions uniformly bounded in H2(d)H^{2}(\mathbb{R}^{d}), d2d\geq 2. Then for any 0<r<20<r<2, up to subsequence φnφ\varphi_{n}\to\varphi in H˙r(d)\dot{H}^{r}(\mathbb{R}^{d}).

Refer to caption
Figure 1. The graph of the function p0(S)=(16S30)/(14S27)p_{0}(S)=(16S-30)/(14S-27) in the case of superharmonic or subharmonic functions. Here r=2,d=5r=2,d=5 and S>5/2.S>5/2.

2. Interpolation inequalities for radial functions involving Riesz potentials.

Let d2d\geq 2, 0<α<d0<\alpha<d, 12<s,\frac{1}{2}<s, we define

X=Xs,α,d={uH˙rads(d),1|x|α|u|L2<+}.X=X_{s,\alpha,d}=\left\{u\in\dot{H}^{s}_{rad}(\mathbb{R}^{d}),\ \ \left\|\frac{1}{|x|^{\alpha}}\star|u|\right\|_{L^{2}}<+\infty\ \right\}.

The aim of this section is to prove the following

Theorem 2.1.

Let uXu\in X with d2d\geq 2, s>12s>\frac{1}{2}, d2<α<d12\frac{d}{2}<\alpha<d-\frac{1}{2}, then uLp(d)u\in L^{p}(\mathbb{R}^{d}) with

p(prad,2dd2s]\displaystyle p\in(p_{rad},\frac{2d}{d-2s}] if s<d2,\displaystyle\text{if }s<\frac{d}{2},
p(prad,)\displaystyle p\in(p_{rad},\infty) if sd2.\displaystyle\text{if }s\geq\frac{d}{2}.

where prad<2p_{rad}<2 with

prad=2(αd2)+2s(d1)(2s1)(αd2)+2s(d1).p_{rad}=\frac{2(\alpha-\frac{d}{2})+2s(d-1)}{-(2s-1)(\alpha-\frac{d}{2})+2s(d-1)}.

Moreover, we have the scaling invariant inequality for uXu\in X

uLp(d)C(α,s,p,d)1|x|α|u|L2(d)1θDsuL2(d)θ,\|u\|_{L^{p}(\mathbb{R}^{d})}\leq C(\alpha,s,p,d)\,\|\frac{1}{|x|^{\alpha}}\star|u|\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\,\|D^{s}u\|_{L^{2}(\mathbb{R}^{d})}^{\theta},

with p(prad,2dd2s]p\in(p_{rad},\frac{2d}{d-2s}] if s<d2s<\frac{d}{2} and p(prad,)p\in(p_{rad},\infty) if sd2s\geq\frac{d}{2}. Here θ\theta is fixed by the scaling invariance

dp=(1θ)((dα)+d2)+θ(s+d2).\frac{d}{p}=(1-\theta)((d-\alpha)+\frac{d}{2})+\theta(-s+\frac{d}{2}).

In order to show Theorem 2.1 we need to prove some preliminary results.

Proposition 2.1.

Let d1d\geq 1, q>1q>1, dq<α<d\frac{d}{q}<\alpha<d, δ>0\delta>0, then there exists C>0C>0 such that

(2.1) BR(0)c|u(x)||x|αdq+δ𝑑xCRδ1|x|α|u|Lq(d)\int_{B_{R}(0)^{c}}\frac{|u(x)|}{|x|^{\alpha-\frac{d}{q}+\delta}}dx\leq\frac{C}{R^{\delta}}||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{q}(\mathbb{R}^{d})}
(2.2) BR(0)|u(x)||x|αdqδ𝑑xCRδ1|x|α|u|Lq(d).\int_{B_{R}(0)}\frac{|u(x)|}{|x|^{\alpha-\frac{d}{q}-\delta}}dx\leq CR^{\delta}||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{q}(\mathbb{R}^{d})}.

The proposition for q=2q=2 has been proved in [13], we follow the same argument for q>1q>1. In order to prove Proposition 2.1 two crucial lemmas are necessary. The case q=2q=2 has been proved in [13] and we follow the same argument.

Lemma 2.1.

Let d1d\geq 1, q1q\geq 1, 0<α<d0<\alpha<d, then there exists C>0C>0 such that for any ada\in\mathbb{R}^{d}

0(Bρ(a)|u(y)|𝑑y)qρ(dα)q+d1𝑑ρC1|x|α|u|Lq(d)q.\int_{0}^{\infty}\left(\fint_{B_{\rho}(a)}|u(y)|dy\right)^{q}\rho^{(d-\alpha)q+d-1}d\rho\leq C||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{q}(\mathbb{R}^{d})}^{q}.
Proof.

Let us take x𝒜ρ=Bρ(a)Bρ2(a)x\in\mathcal{A_{\rho}}=B_{\rho}(a)\setminus B_{\frac{\rho}{2}}(a), then

1|x|α|u|(x)=d|u(y)||xy|α𝑑y\frac{1}{|x|^{\alpha}}\star|u|(x)=\int_{\mathbb{R}^{d}}\frac{|u(y)|}{|x-y|^{\alpha}}dy\geq
Bρ(a)|u(y)||xy|α𝑑yCρdαBρ(a)|u(y)|𝑑y.\geq\int_{B_{\rho}(a)}\frac{|u(y)|}{|x-y|^{\alpha}}dy\geq C\rho^{d-\alpha}\fint_{B_{\rho}(a)}|u(y)|dy.

Thus we obtain for x𝒜ρx\in\mathcal{A_{\rho}}

(1|x|α|u|(x))qCρ(dα)q(Bρ(a)|u(y)|𝑑y)q\left(\frac{1}{|x|^{\alpha}}\star|u|(x)\right)^{q}\geq C\rho^{(d-\alpha)q}\left(\fint_{B_{\rho}(a)}|u(y)|dy\right)^{q}

and hence

𝒜ρ(1|x|α|u|(x))q𝑑xCρ(dα)q+d(Bρ(a)|u(y)|𝑑y)q.\int_{\mathcal{A_{\rho}}}\left(\frac{1}{|x|^{\alpha}}\star|u|(x)\right)^{q}dx\geq C\rho^{(d-\alpha)q+d}\left(\fint_{B_{\rho}(a)}|u(y)|dy\right)^{q}.

By integration we conclude that

0ρ(dα)q+d1(Bρ(a)|u(y)|𝑑y)q𝑑ρ\int_{0}^{\infty}\rho^{(d-\alpha)q+d-1}\left(\fint_{B_{\rho}(a)}|u(y)|dy\right)^{q}d\rho\leq
C0(𝒜ρ(1|x|α|u|(x))q𝑑x)dρρ=C1|x|α|u|Lq(d)q.\leq C\int_{0}^{\infty}\left(\int_{\mathcal{A_{\rho}}}\left(\frac{1}{|x|^{\alpha}}\star|u|(x)\right)^{q}dx\right)\frac{d\rho}{\rho}=C||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{q}(\mathbb{R}^{d})}^{q}.

Let us call W(ρ)=ρw(s)𝑑sW(\rho)=\int_{\rho}^{\infty}w(s)ds where w:(0,)w:(0,\infty)\rightarrow\mathbb{R} is a measurable function such that

(2.3) 0|w(ρ)|qq1ραq+1dq1𝑑ρ<+.\int_{0}^{\infty}|w(\rho)|^{\frac{q}{q-1}}\rho^{\frac{\alpha q+1-d}{q-1}}d\rho<+\infty.
Lemma 2.2.

Let d1d\geq 1, q>1q>1, 0<α<d0<\alpha<d, then

|d|u(x)|W(|x|)𝑑x||\int_{\mathbb{R}^{d}}|u(x)|W(|x|)dx|\lesssim
(0|w(ρ)|qq1ραq+1dq1𝑑ρ)q1q(0(Bρ(a)|u(y)|𝑑y)qραq+d1𝑑ρ)1q,\left(\int_{0}^{\infty}|w(\rho)|^{\frac{q}{q-1}}\rho^{\frac{\alpha q+1-d}{q-1}}d\rho\right)^{\frac{q-1}{q}}\left(\int_{0}^{\infty}\left(\fint_{B_{\rho}(a)}|u(y)|dy\right)^{q}\rho^{\alpha q+d-1}d\rho\right)^{\frac{1}{q}},

and hence

(2.4) |d|u(x)|W(|x|)𝑑x|C1|x|α|u|Lq(d).|\int_{\mathbb{R}^{d}}|u(x)|W(|x|)dx|\leq C||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{q}(\mathbb{R}^{d})}.
Proof.

We have, thanks to Fubini Theorem,

d|u(x)|W(|x|)𝑑x=d|u(x)|(|x|w(ρ)𝑑ρ)𝑑x=\int_{\mathbb{R}^{d}}|u(x)|W(|x|)dx=\int_{\mathbb{R}^{d}}|u(x)|\left(\int_{|x|}^{\infty}w(\rho)d\rho\right)dx=
=C0w(ρ)ρd(Bρ(0)|u(y)|𝑑y)𝑑ρ=C\int_{0}^{\infty}w(\rho)\rho^{d}\left(\fint_{B_{\rho}(0)}|u(y)|dy\right)d\rho

such that by Hölder’s inequality we obtain

|d|u(x)|W(|x|)𝑑x|=C|0w(ρ)ρdβ(Bρ(0)|u(y)|𝑑y)ρβ𝑑ρ||\int_{\mathbb{R}^{d}}|u(x)|W(|x|)dx|=C|\int_{0}^{\infty}w(\rho)\rho^{d-\beta}\left(\fint_{B_{\rho}(0)}|u(y)|dy\right)\rho^{\beta}d\rho|\lesssim
(0|w(ρ)|qq1ραq+1dq1𝑑ρ)q1q(0(Bρ(0)|u(y)|𝑑y)qραq+d1𝑑ρ)1q,\left(\int_{0}^{\infty}|w(\rho)|^{\frac{q}{q-1}}\rho^{\frac{\alpha q+1-d}{q-1}}d\rho\right)^{\frac{q-1}{q}}\left(\int_{0}^{\infty}\left(\fint_{B_{\rho}(0)}|u(y)|dy\right)^{q}\rho^{\alpha q+d-1}d\rho\right)^{\frac{1}{q}},

choosing β\beta such that βq=(dα)q+d1.\beta q=(d-\alpha)q+d-1. Eq. (2.4) comes from Lemma 2.1. ∎

Proof of Proposition 2.1.

If we choose

w(ρ)={0,if 0<ρ<R;1ραdq+1+δ,if ρ>R.w(\rho)=\left\{\begin{array}[]{ll}0,&\hbox{if $0<\rho<R$;}\\ \frac{1}{\rho^{\alpha-\frac{d}{q}+1+\delta}},&\hbox{if $\rho>R$.}\end{array}\right.

thanks to Lemma 2.2 we get (2.1). In order to get (2.2) it is enough to choose

w(ρ)={0,if ρ>R;1ραdq+1δ,if 0<ρ<R.w(\rho)=\left\{\begin{array}[]{ll}0,&\hbox{if $\rho>R$;}\\ \frac{1}{\rho^{\alpha-\frac{d}{q}+1-\delta}},&\hbox{if $0<\rho<R$.}\end{array}\right.

Lemma 2.3.

Let d1d\geq 1, d2<α<d\frac{d}{2}<\alpha<d and DsuL2(d)=1|x|α|u|L2(d)=1||D^{s}u||_{L^{2}(\mathbb{R}^{d})}=||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{2}(\mathbb{R}^{d})}=1, then for any δ>0\delta>0 such that 0<δ<dα0<\delta<d-\alpha,

d|u(x)||x|αd2+δ𝑑xC(α,s,δ,d).\int_{\mathbb{R}^{d}}\frac{|u(x)|}{|x|^{\alpha-\frac{d}{2}+\delta}}dx\leq C(\alpha,s,\delta,d).
Proof.

Let 0<ϵ<d20<\epsilon<\frac{d}{2} be a number to be fixed later. We have

B(0,1)|u(x)||x|αd2+δ𝑑x=B(0,1)|u(x)||x|αd2+δϵ1|x|ϵ𝑑x\int_{B(0,1)}\frac{|u(x)|}{|x|^{\alpha-\frac{d}{2}+\delta}}dx=\int_{B(0,1)}\frac{|u(x)|}{|x|^{\alpha-\frac{d}{2}+\delta-\epsilon}}\frac{1}{|x|^{\epsilon}}dx\leq
cd,ϵ(B(0,1)|u(x)|2|x|2(αd2+δϵ))12,\leq c_{d,\epsilon}\left(\int_{B(0,1)}\frac{|u(x)|^{2}}{|x|^{2(\alpha-\frac{d}{2}+\delta-\epsilon)}}\right)^{\frac{1}{2}},

where cd,ϵ=(B(0,1)1|x|2ϵ𝑑x)12c_{d,\epsilon}=\left(\int_{B(0,1)}\frac{1}{|x|^{2\epsilon}}dx\right)^{\frac{1}{2}}. Now choose ϵ=αd2+δ\epsilon=\alpha-\frac{d}{2}+\delta. Notice that ϵ<d2\epsilon<\frac{d}{2} such that

B(0,1)|u(x)||x|αd2+δ𝑑xcd,ϵ(B(0,1)|u(x)|2𝑑x)12.\int_{B(0,1)}\frac{|u(x)|}{|x|^{\alpha-\frac{d}{2}+\delta}}dx\leq c_{d,\epsilon}\left(\int_{B(0,1)}|u(x)|^{2}dx\right)^{\frac{1}{2}}.

which implies

B(0,1)|u(x)||x|αd2+δ𝑑x1.\int_{B(0,1)}\frac{|u(x)|}{|x|^{\alpha-\frac{d}{2}+\delta}}dx\lesssim 1.

On the other hand by Proposition 2.1, when d2<α<d\frac{d}{2}<\alpha<d

B(0,1)c|u(x)||x|αd2+δ𝑑xC1|x|α|u|L2(d)\int_{B(0,1)^{c}}\frac{|u(x)|}{|x|^{\alpha-\frac{d}{2}+\delta}}dx\leq C||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{2}(\mathbb{R}^{d})}

and hence we obtain the claim.∎

The next Proposition concerning pointwise decay for radial functions in XX follows the strategy of Theorem 3.1 in [8]. We will decompose the function in high/low frequency part, estimating the high frequency part involving the Sobolev norm while we control the low frequency part involving the Riesz norm.

Proposition 2.2.

Let d2,d\geq 2, uu be a radial function in XX with s>12s>\frac{1}{2}, d2<α<d\frac{d}{2}<\alpha<d, and

(2.5) DsuL2(d)=1|x|α|u|L2(d)=1.||D^{s}u||_{L^{2}(\mathbb{R}^{d})}=||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{2}(\mathbb{R}^{d})}=1.

Then for any σ\sigma satisfying

(2.6) 2s(d21)+(d2)2s+1<σ<2s(d1)(2s1)(αd2)2s+1\frac{2s\left(\frac{d}{2}-1\right)+\left(\frac{d}{2}\right)}{2s+1}<\sigma<\frac{2s(d-1)-(2s-1)\left(\alpha-\frac{d}{2}\right)}{2s+1}

we have

|u(x)|C(α,s,σ,d)|x|σ.|u(x)|\leq C(\alpha,s,\sigma,d)|x|^{-\sigma}.
Remark 2.1.

It is easy to see that the above Proposition is equivalent to the following statement.

Let uu be a radial function in XX with s>12s>\frac{1}{2}, d2<α<d\frac{d}{2}<\alpha<d, and

(2.7) DsuL2(d)=1|x|α|u|L2(d)=1||D^{s}u||_{L^{2}(\mathbb{R}^{d})}=||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{2}(\mathbb{R}^{d})}=1

then for any δ>0\delta>0 such that 0<δ<dα0<\delta<d-\alpha,

|u(x)|C(α,s,δ,d)|x|σ|u(x)|\leq C(\alpha,s,\delta,d)|x|^{-\sigma}

with

(2.8) σ=(2s1)(αd2+δ)+2s(d1)2s+1.\sigma=\frac{-(2s-1)(\alpha-\frac{d}{2}+\delta)+2s(d-1)}{2s+1}.
Proof.

For any R>1R>1 we can take a function ψR(x)=Rdψ(x/R)\psi_{R}(x)=R^{-d}\psi(x/R) such that ψ^(ξ)\widehat{\psi}(\xi) is a radial nonnegative function with support in |ξ|2|\xi|\leq 2 and ψ^(ξ)=1\widehat{\psi}(\xi)=1 for |ξ|1|\xi|\leq 1 and then we make the decomposition of uu into low and high frequency part as follows

u(x)=ψRu(x)+h(x)u(x)=\psi_{R}\star u(x)+h(x)

where h^(ξ)=(1ψ^(R|ξ|))u^(ξ)\hat{h}(\xi)=(1-\hat{\psi}(R|\xi|))\hat{u}(\xi). For the high frequency part we will use Fourier representation for radial functions in d\mathbb{R}^{d} (identifying the function with its profile)

(2.9) |h(x)|=(2π)d2|x|d220Jd22(|x|ρ)(1ψ(Rρ))u^(ρ)ρd2𝑑ρ|h(x)|=(2\pi)^{\frac{d}{2}}|x|^{-\frac{d-2}{2}}\int_{0}^{\infty}J_{\frac{d-2}{2}}(|x|\rho)(1-\psi(R\rho))\hat{u}(\rho)\rho^{\frac{d}{2}}d\rho

where Jd22J_{\frac{d-2}{2}} is the Bessel function of order d22.\frac{d-2}{2}. Applying the results in [7] and [8], we find

(2.10) |h(x)|cRs12|x|12(d1)uH˙s(d),s>12.|h(x)|\leq cR^{s-\frac{1}{2}}|x|^{-\frac{1}{2}(d-1)}||u||_{\dot{H}^{s}(\mathbb{R}^{d})},\ s>\frac{1}{2}.

Indeed, using the uniform bound

|Jd22(ρ)|(1+ρ)1/2,|J_{\frac{d-2}{2}}(\rho)|\lesssim(1+\rho)^{-1/2},

we get

|h(x)||x|d220|(Jd22)(|x|ρ)(1ψ(Rρ))u^(ρ)|ρd2𝑑ρ|h(x)|\lesssim|x|^{-\frac{d-2}{2}}\int_{0}^{\infty}|(J_{\frac{d-2}{2}})(|x|\rho)||(1-\psi(R\rho))||\hat{u}(\rho)|\rho^{\frac{d}{2}}d\rho\lesssim
|x|d22(1/R|Jd22(|x|ρ)|2dρρ2s1)1/2(0|u^(ρ)|2ρ2s+d1𝑑ρ)1/2|x|^{-\frac{d-2}{2}}\left(\int_{1/R}^{\infty}|J_{\frac{d-2}{2}}(|x|\rho)|^{2}\frac{d\rho}{\rho^{2s-1}}\right)^{1/2}\left(\int_{0}^{\infty}|\hat{u}(\rho)|^{2}\rho^{2s+d-1}d\rho\right)^{1/2}\lesssim
|x|d22Rs1(1(1+|x|ρ/R)1dρρ2s1)1/2uH˙s(d)|x|^{-\frac{d-2}{2}}R^{s-1}\left(\int_{1}^{\infty}(1+|x|\rho/R)^{-1}\frac{d\rho}{\rho^{2s-1}}\right)^{1/2}\|u\|_{\dot{H}^{s}(\mathbb{R}^{d})}\lesssim
Rs1/2|x|d12uH˙s(d)\lesssim R^{s-1/2}|x|^{-\frac{d-1}{2}}\|u\|_{\dot{H}^{s}(\mathbb{R}^{d})}

and this gives (2.10).

For low frequency term ψRu(x)\psi_{R}\star u(x), since ψ𝒮(d)\psi\in\mathcal{S}\left(\mathbb{R}^{d}\right), we can take any γ>1\gamma>1 so that there exists C>0C>0 such that

|ψ(x)|C(1+|x|2)γ/2.|\psi(x)|\leq C\left(1+|x|^{2}\right)^{-\gamma/2}.

We shall need the following estimate that can be found also in [12] and [8]. For sake of completeness we give an alternative proof of the Lemma in the Appendix.

Lemma 2.4.

If b(d+1,0),γ>d1,b\in(-d+1,0),\gamma>d-1, then for any radially symmetric function f(|y|)f(|y|) we have

(2.11) |df(|y|)dy(1+|xy|2)γ/2|1|x|d1+b|y|bfL1(d).\left|\int_{\mathbb{R}^{d}}\frac{f(|y|)dy}{(1+|x-y|^{2})^{\gamma/2}}\right|\lesssim\frac{1}{|x|^{d-1+b}}\left\||y|^{b}f\right\|_{L^{1}(\mathbb{R}^{d})}.

Then we estimate ψRu(x)\psi_{R}\star u(x) as follows,

|ψRu(x)|\displaystyle|\psi_{R}\star u(x)| |ψR(x)||u(x)|Cd1Rd|u(y)|(1+|xyR|)γ/2𝑑y\displaystyle\leq\left|\psi_{R}(x)\right|*|u(x)|\leq C\int_{\mathbb{R}^{d}}\frac{1}{R^{d}}\frac{|u(y)|}{\left(1+\left|\frac{x-y}{R}\right|\right)^{\gamma/2}}dy
Cd|u(Rz)|(1+|xRz|2dz)γ/2𝑑z(y=Rz).\displaystyle\leq C\int_{\mathbb{R}^{d}}\frac{|u(Rz)|}{\left(1+\left|\frac{x}{R}-z\right|^{2}dz\right)^{\gamma/2}}dz\quad(y=Rz).

To this end we plan to apply Lemma 2.4 assuming b=(αd/2+δ)b=-(\alpha-d/2+\delta). To check the assumption of the Lemma we use the inequalities

αd2+δ<d2d1\alpha-\frac{d}{2}+\delta<\frac{d}{2}\leq d-1

for d2.d\geq 2. Applying the Lemma 2.4 we deduce

|ψRu(x)|\displaystyle|\psi_{R}\star u(x)|
C|xR|(d1+b)d|u(Rz)||z|b𝑑z\displaystyle\leq C\left|\frac{x}{R}\right|^{-(d-1+b)}\int_{\mathbb{R}^{d}}|u(Rz)||z|^{b}dz
CR(d1+b)|x|(d1+b)d|u(y)||yR|bdyRd\displaystyle\leq CR^{(d-1+b)}|x|^{-(d-1+b)}\int_{\mathbb{R}^{d}}|u(y)|\left|\frac{y}{R}\right|^{b}\frac{dy}{R^{d}}
CR1|x|(d1+b)|y|buL1(d).\displaystyle\leq CR^{-1}|x|^{-(d-1+b)}\||y|^{b}u\|_{L^{1}(\mathbb{R}^{d})}.

Therefore, collecting our estimates and using the condition (2.5), we find

|u(x)|\displaystyle|u(x)| C[|x|(d1)/2Rs1/2+|x|(d1+b)R1|y|buL1(d)].\displaystyle\leq C\left[|x|^{-(d-1)/2}R^{s-1/2}+|x|^{-(d-1+b)}R^{-1}\||y|^{b}u\|_{L^{1}(\mathbb{R}^{d})}\right].

We use Lemma 2.3 and we get

|u(x)|\displaystyle|u(x)| C[|x|(d1)/2Rs1/2+|x|(d1+b)R1].\displaystyle\leq C\left[|x|^{-(d-1)/2}R^{s-1/2}+|x|^{-(d-1+b)}R^{-1}\right].

Minimizing in RR or equivalently choosing R>0R>0 so that

|x|(d1)/2Rs1/2=|x|(d1+b)R1,|x|^{-(d-1)/2}R^{s-1/2}=|x|^{-(d-1+b)}R^{-1},

i.e.

Rs+1/2=|x|b(d1)/2,R^{s+1/2}=|x|^{-b-(d-1)/2},

we find

|u(x)|C(d,s,α,δ)|x|σ.|u(x)|\leq C(d,s,\alpha,\delta)|x|^{-\sigma}.

where σ\sigma is defined in (2.8).

This completes the proof.

With all these preliminary results we are now ready to prove Theorem 2.1.

Proof.

Let uXu\in X with DsuL2(d)=1|x|α|u|L2(d)=1||D^{s}u||_{L^{2}(\mathbb{R}^{d})}=||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{2}(\mathbb{R}^{d})}=1, then by Proposition 2.2

|u(x)|C(d,s,α,δ)|x|σ|u(x)|\leq C(d,s,\alpha,\delta)|x|^{-\sigma}

with

σ=(2s1)(αd2+δ)+2s(d1)2s+1.\sigma=\frac{-(2s-1)(\alpha-\frac{d}{2}+\delta)+2s(d-1)}{2s+1}.

We aim to show that prad<2p_{rad}<2, where p=2p=2 is the lower endpoint for (1.12). Therefore it sufficies to show that |x|>1|u|p𝑑x<+\int_{|x|>1}|u|^{p}dx<+\infty provided that uXu\in X and prad<pp_{rad}<p (indeed |x|1|u|p𝑑x<+\int_{|x|\leq 1}|u|^{p}dx<+\infty for all 0<p<20<p<2 by interpolation).
We have, thanks to Proposition 2.2 and Lemma 2.3,

(2.12) |x|>1|u||u|p1𝑑x|x|>1|u||x|σ(p1)𝑑x1\int_{|x|>1}|u||u|^{p-1}dx\lesssim\int_{|x|>1}\frac{|u|}{|x|^{\sigma(p-1)}}dx\lesssim 1

provided that σ(p1)>αd2\sigma(p-1)>\alpha-\frac{d}{2}. This condition is equivalent, σ\sigma is defined in (2.8) and letting δ0\delta\rightarrow 0, to

p>σ+αd2σ=2(αd2)+2s(d1)(2s1)(αd2)+2s(d1):=prad.p>\frac{\sigma+\alpha-\frac{d}{2}}{\sigma}=\frac{2(\alpha-\frac{d}{2})+2s(d-1)}{-(2s-1)(\alpha-\frac{d}{2})+2s(d-1)}:=p_{rad}.

An elementary computation shows that prad<2p_{rad}<2 provided that d2<α<d12\frac{d}{2}<\alpha<{d-\frac{1}{2}}.

Now consider an arbitrary vXv\in X and let us call u=λv(μx)u=\lambda v(\mu x) where the parameters λ,μ>0\lambda,\mu>0 are chosen such that DsuL2(d)=1|x|α|u|L2(d)=1||D^{s}u||_{L^{2}(\mathbb{R}^{d})}=||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{2}(\mathbb{R}^{d})}=1. By scaling we have

1=DsuL2(d)=λμsd2DsvL2(d)1=||D^{s}u||_{L^{2}(\mathbb{R}^{d})}=\lambda\mu^{s-\frac{d}{2}}||D^{s}v||_{L^{2}(\mathbb{R}^{d})}
1=1|x|α|u|L2(d)=λμα32d1|x|α|v|L2(d)1=||\frac{1}{|x|^{\alpha}}\star|u|||_{L^{2}(\mathbb{R}^{d})}=\lambda\mu^{\alpha-\frac{3}{2}d}||\frac{1}{|x|^{\alpha}}\star|v|||_{L^{2}(\mathbb{R}^{d})}

and hence we obtain the relations

μ=(DsvL2(d)1|x|α|v|L2(d))1αsd,λ=1|x|α|v|L2(d)sd2αdsDsvL2(d)α32dαds.\mu=\left(\frac{||D^{s}v||_{L^{2}(\mathbb{R}^{d})}}{||\frac{1}{|x|^{\alpha}}\star|v|||_{L^{2}(\mathbb{R}^{d})}}\right)^{\frac{1}{\alpha-s-d}},\ \ \lambda=\frac{||\frac{1}{|x|^{\alpha}}\star|v|||_{L^{2}(\mathbb{R}^{d})}^{\frac{s-\frac{d}{2}}{\alpha-d-s}}}{||D^{s}v||_{L^{2}(\mathbb{R}^{d})}^{\frac{\alpha-\frac{3}{2}d}{\alpha-d-s}}}.

By the previous estimates we have

uLp(d)=λμdpvLp(d)1||u||_{L^{p}(\mathbb{R}^{d})}=\lambda\mu^{-\frac{d}{p}}||v||_{L^{p}(\mathbb{R}^{d})}\lesssim 1

which implies

vLp(d)λ1μdp=DsvL2(d)θ1|x|α|v|L2(d)1θ,||v||_{L^{p}(\mathbb{R}^{d})}\lesssim\lambda^{-1}\mu^{\frac{d}{p}}=||D^{s}v||_{L^{2}(\mathbb{R}^{d})}^{\theta}\|\frac{1}{|x|^{\alpha}}\star|v|\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta},

where

θ=d22αp+3dp2ds2p(d+sα), 1θ=(2sd)(d+p)2p(d+sα).\theta=\frac{d^{2}-2\alpha p+3dp-2ds}{2p(d+s-\alpha)},\ \ 1-\theta=\frac{(2s-d)(d+p)}{2p(d+s-\alpha)}.

It is easy to see that θ\theta is fixed by the scaling invariance

dp=(1θ)((dα)+d2)+θ(s+d2).\frac{d}{p}=(1-\theta)((d-\alpha)+\frac{d}{2})+\theta(-s+\frac{d}{2}).

3. Proof of Theorem 1.1

Our goal is to represent φ\varphi in the form φ=1|x|αu=cDru,\varphi=\frac{1}{|x|^{\alpha}}\star u=cD^{-r}u, with α=dr,\alpha=d-r, c=πd/2Γ((dα)/2)Γ(α/2)c=\frac{\pi^{d/2}\Gamma((d-\alpha)/2)}{\Gamma(\alpha/2)} and apply Theorem 2.1. Therefore, we choose (modulo constant) u=Drφ.u=D^{r}\varphi.

Then the estimate of Theorem 2.1 gives

DrφLp(d)=uLp(d)1|x|α|u|L2(d)1θDsuL2(d)θ=\|D^{r}\varphi\|_{L^{p}(\mathbb{R}^{d})}=\|u\|_{L^{p}(\mathbb{R}^{d})}\lesssim\,\|\frac{1}{|x|^{\alpha}}\star|u|\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\,\|D^{s}u\|_{L^{2}(\mathbb{R}^{d})}^{\theta}=
=Dr|Drφ|L2(d)1θDSφ|L2(d)θ.=\left\|D^{-r}|D^{r}\varphi|\right\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\|D^{S}\varphi|\|_{L^{2}(\mathbb{R}^{d})}^{\theta}.

By the assumption

(3.1) Drφ(x)0D^{r}\varphi(x)\geq 0

for almost every xd,x\in\mathbb{R}^{d}, then we deduce

Dr|Drφ|L2(d)1θDSφ|L2(d)θ=DrDrφL2(d)1θDSφ|L2(d)θ\left\|D^{-r}|D^{r}\varphi|\right\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\|D^{S}\varphi|\|_{L^{2}(\mathbb{R}^{d})}^{\theta}=\left\|D^{-r}D^{r}\varphi\right\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\|D^{S}\varphi|\|_{L^{2}(\mathbb{R}^{d})}^{\theta}

and we obtain (1.9). The lower endpoint p0p_{0} is hence nothing but pradp_{rad} of Theorem 2.1 substituting α\alpha with drd-r and ss with SrS-r. The condition 12<r<min(d2,S12)\frac{1}{2}<r<\min(\frac{d}{2},S-\frac{1}{2}) is equivalent to the conditions d2<α<d12\frac{d}{2}<\alpha<d-\frac{1}{2}, s>12s>\frac{1}{2} of Theorem 2.1. All these estimates remain valid if we consider Drφ(x)0D^{r}\varphi(x)\leq 0, i.e if φHrad,s,r(d)\varphi\in H^{s,r}_{rad,-}(\mathbb{R}^{d}). Indeed if φHrad,s,r(d)\varphi\in H^{s,r}_{rad,-}(\mathbb{R}^{d})

Dr|Drφ|L2(d)1θDSφ|L2(d)θ=\left\|D^{-r}|D^{r}\varphi|\right\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\|D^{S}\varphi|\|_{L^{2}(\mathbb{R}^{d})}^{\theta}=
=DrDrφL2(d)1θDSφ|L2(d)θ=φL2(d)1θDSφ|L2(d)θ.=\left\|-D^{-r}D^{r}\varphi\right\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\|D^{S}\varphi|\|_{L^{2}(\mathbb{R}^{d})}^{\theta}=\left\|\varphi\right\|_{L^{2}(\mathbb{R}^{d})}^{1-\theta}\|D^{S}\varphi|\|_{L^{2}(\mathbb{R}^{d})}^{\theta}.

4. Proof of Theorem 1.2

We prove that under the assumption of Theorem 1.2, the embedding

Hrad,+S,r0(d)H˙radr0(d)H^{S,r_{0}}_{rad,+}(\mathbb{R}^{d})\subset\subset\dot{H}^{r_{0}}_{rad}(\mathbb{R}^{d})

is compact. As a byproduct the embedding

(4.1) Hrad,+S,r0(d)H˙radr(d)H^{S,r_{0}}_{rad,+}(\mathbb{R}^{d})\subset\subset\dot{H}^{r}_{rad}(\mathbb{R}^{d})

is compact for any 0<r<S0<r<S. The embedding (4.1) follows noticing that if φn\varphi_{n} converges weakly to some φ\varphi in HradS(d)H^{S}_{rad}(\mathbb{R}^{d}) then φn\varphi_{n} converges weakly to the same φ\varphi in Hradr0(d)H^{r_{0}}_{rad}(\mathbb{R}^{d}). Now if we prove that (taking a subsequence)

(4.2) Dr0(φnφ)L2=o(1)\|D^{r_{0}}(\varphi_{n}-\varphi)\|_{L^{2}}=o(1)

as n,n\to\infty, then by the following interpolation inequalities

Dr(φnφ)L2Dr0(φnφ)L21rr0Sr0DS(φnφ)L2rr0Sr0=o(1)\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2}}\lesssim\|D^{r_{0}}(\varphi_{n}-\varphi)\|_{L^{2}}^{1-\frac{r-r_{0}}{S-r_{0}}}\,\|D^{S}(\varphi_{n}-\varphi)\|_{L^{2}}^{\frac{r-r_{0}}{S-r_{0}}}=o(1)

if 0<r0<r<S0<r_{0}<r<S and

Dr(φnφ)L2(φnφ)L21rr0Dr0(φnφ)L2rr0=o(1)\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2}}\lesssim\|(\varphi_{n}-\varphi)\|_{L^{2}}^{1-\frac{r}{r_{0}}}\,\|D^{r_{0}}(\varphi_{n}-\varphi)\|_{L^{2}}^{\frac{r}{r_{0}}}=o(1)

if 0<r<r00<r<r_{0}, we get (4.1).

To prove (4.2) we recall that (φn)n(\varphi_{n})_{n\in\mathbb{N}} is a bounded sequence in Hrad,+S,r0(d)H^{S,r_{0}}_{rad,+}(\mathbb{R}^{d}) and we can assume that φn\varphi_{n} converges weakly to some φ\varphi in HS(d)H^{S}(\mathbb{R}^{d}). To simplify the notation we will use rr instead of r0r_{0} in the proof of (4.2). We choose a bump function θC0(d)\theta\in C_{0}^{\infty}(\mathbb{R}^{d}), such that θ=1\theta=1 on B1B_{1} and θ=0\theta=0 in dB2\mathbb{R}^{d}\setminus B_{2} and for any ρ>1\rho>1 we define θρ(x)=θ(x/ρ).\theta_{\rho}(x)=\theta(x/\rho). Clearly the multiplication by θρ𝒮(d)\theta_{\rho}\in\mathcal{S}(\mathbb{R}^{d}) is a continuous mapping HS(d)HS(d)H^{S}(\mathbb{R}^{d})\rightarrow H^{S}(\mathbb{R}^{d}). Now setting vn=θρφnv_{n}=\theta_{\rho}\varphi_{n} and v=θρφv=\theta_{\rho}\varphi we aim to show that

(4.3) limnDr(vnv)L2(d)2=limnDr(θρ(φnφ))L2(d)2=0.\lim_{n\to\infty}\|D^{r}(v_{n}-v)\|_{L^{2}(\mathbb{R}^{d})}^{2}=\lim_{n\to\infty}\|D^{r}(\theta_{\rho}(\varphi_{n}-\varphi))\|_{L^{2}(\mathbb{R}^{d})}^{2}=0.

for any r[0,S).r\in[0,S).

Indeed, by Plancharel’s identity we have

Dr(vnv)L2(d)2=|ξ|R|ξ|2r|v^n(ξ)v^(ξ)|2𝑑ξ=I+|ξ|>R|ξ|2r|v^n(ξ)v^(ξ)|2𝑑ξ=II.\|D^{r}(v_{n}-v)\|_{L^{2}(\mathbb{R}^{d})}^{2}=\underbrace{\int_{|\xi|\leq R}|\xi|^{2r}|\widehat{v}_{n}(\xi)-\widehat{v}(\xi)|^{2}d\xi}_{=I}+\underbrace{\int_{|\xi|>R}|\xi|^{2r}|\widehat{v}_{n}(\xi)-\widehat{v}(\xi)|^{2}d\xi}_{=II}.

Clearly

II1R2(Sr)|ξ|>R|ξ|2S|v^n(ξ)v^(ξ)|2𝑑ξII\leq\frac{1}{R^{2(S-r)}}\int_{|\xi|>R}|\xi|^{2S}|\widehat{v}_{n}(\xi)-\widehat{v}(\xi)|^{2}d\xi

and then we can choose R>0R>0 such that IIϵ2II\leq\frac{\epsilon}{2}.
Since e2πixξLx2(B2ρ)e^{-2\pi ix\cdot\xi}\in L^{2}_{x}(B_{2\rho}), by weak convergence in L2(B2ρ)L^{2}(B_{2\rho}) we have v^n(ξ)v^(ξ)\widehat{v}_{n}(\xi)\rightarrow\widehat{v}(\xi) almost everywhere. Notice that v^nLvnL1(B2ρ)μ(B2ρ)12vnL2(B2ρ)μ(B2ρ)12vnHS(d)||\widehat{v}_{n}||_{L^{\infty}}\leq||v_{n}||_{L^{1}(B_{2\rho})}\leq\mu(B_{2\rho})^{\frac{1}{2}}||v_{n}||_{L^{2}(B_{2\rho})}\leq\mu(B_{2\rho})^{\frac{1}{2}}||v_{n}||_{H^{S}(\mathbb{R}^{d})} and hence |v^n(ξ)v^(ξ)|2|\widehat{v}_{n}(\xi)-\widehat{v}(\xi)|^{2} is estimated by a uniform constant so that by Lebesgue’s dominated convergence theorem

I=|ξ|R|ξ|2r|v^n(ξ)v^(ξ)|2𝑑ξ<ϵ2,I=\int_{|\xi|\leq R}|\xi|^{2r}|\widehat{v}_{n}(\xi)-\widehat{v}(\xi)|^{2}d\xi<\frac{\epsilon}{2},

for nn sufficiently large. This proves (4.3).

Our next step is to show that for a given ε>0\varepsilon>0 one can find ρ0=ρ0(ε)\rho_{0}=\rho_{0}(\varepsilon) sufficiently large and n0(ε)n_{0}(\varepsilon) sufficiently large so that

(4.4) θρDr(φnφ)L2(d)2ε2\|\theta_{\rho}D^{r}(\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}^{2}\leq\frac{\varepsilon}{2}

for nn0,ρρ0n\geq n_{0},\rho\geq\rho_{0} and any r[0,S).r\in[0,S).

We consider first the case 0r2,r<S.0\leq r\leq 2,r<S. The cases r=0r=0 and r=2r=2 are trivial, for this we assume 0<r<min(2,S).0<r<\min(2,S). We shall use the following statement (see Corollary 1.1 in [10]).

Proposition 4.1.

Let p,p1,p2p,p_{1},p_{2} satisfy 1<p,p1,p2<1<p,p_{1},p_{2}<\infty and 1/p=1/p1+1/p21/p=1/p_{1}+1/p_{2}. Let r,r1,r2r,r_{1},r_{2} satisfy 0r1,r210\leq r_{1},r_{2}\leq 1, and r=r1+r2r=r_{1}+r_{2}. Then the following bilinear estimate

Dr(fg)fDrg=[Dr,f]ggDrfLpCDr1fLp1Dr2gLp2\|\underbrace{D^{r}(fg)-fD^{r}g}_{=[D^{r},f]g}-gD^{r}f\|_{L^{p}}\leq C\|D^{r_{1}}f\|_{L^{p_{1}}}\|D^{r_{2}}g\|_{L^{p_{2}}}

holds for all f,g𝒮f,g\in\mathcal{S}.

By a density argument the statement holds for f,gHS(d).f,g\in H^{S}(\mathbb{R}^{d}). We choose f=θρ,f=\theta_{\rho}, g=φnφg=\varphi_{n}-\varphi and r1=r2=r/2r_{1}=r_{2}=r/2 and therefore we aim to use (4.3) and prove that

(4.5) [θρ,Dr](φnφ)L2(d)\displaystyle\|[\theta_{\rho},D^{r}](\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}\leq
O(ρr)φnφL2(d)+O(ρr/4)φnφHr(d).\displaystyle\leq O(\rho^{-r})\|\varphi_{n}-\varphi\|_{L^{2}(\mathbb{R}^{d})}+O(\rho^{-r/4})\|\varphi_{n}-\varphi\|_{H^{r}(\mathbb{R}^{d})}.

Indeed from the Proposition 4.1 we have

[θρ,Dr](φnφ)L2(d)DrθρL(d)φnφL2(d)=O(ρr)+\|[\theta_{\rho},D^{r}](\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}\lesssim\underbrace{\|D^{r}\theta_{\rho}\|_{L^{\infty}(\mathbb{R}^{d})}\|\varphi_{n}-\varphi\|_{L^{2}(\mathbb{R}^{d})}}_{=O(\rho^{-r})}+
+Dr/2θρLp1(d)Dr/2(φnφ))Lp2(d).+\|D^{r/2}\theta_{\rho}\|_{L^{p_{1}}(\mathbb{R}^{d})}\|D^{r/2}(\varphi_{n}-\varphi))\|_{L^{p_{2}}(\mathbb{R}^{d})}.

It is easy to check the estimate

Dr/2θρLp1(d)=O(ρr/4),\|D^{r/2}\theta_{\rho}\|_{L^{p_{1}}(\mathbb{R}^{d})}=O(\rho^{-r/4}),

as ρ,\rho\to\infty, and this is obviously fulfilled if dp1<r4\frac{d}{p_{1}}<\frac{r}{4}. To control Dr/2(φnφ))Lp2(d)\|D^{r/2}(\varphi_{n}-\varphi))\|_{L^{p_{2}}(\mathbb{R}^{d})} we use Sobolev inequality

Dr/2(φnφ))Lp2(d)φnφHr(d)\|D^{r/2}(\varphi_{n}-\varphi))\|_{L^{p_{2}}(\mathbb{R}^{d})}\lesssim\|\varphi_{n}-\varphi\|_{H^{r}(\mathbb{R}^{d})}

so we need

1p2>12rr/2d.\frac{1}{p_{2}}>\frac{1}{2}-\frac{r-r/2}{d}.

Summing up we have the following restrictions for 1/p1,1/p21/p_{1},1/p_{2}

(4.6) 1p1+1p2=12\displaystyle\frac{1}{p_{1}}+\frac{1}{p_{2}}=\frac{1}{2}
1p1<r4d,1p2>12rr/2d.\displaystyle\frac{1}{p_{1}}<\frac{r}{4d},\ \frac{1}{p_{2}}>\frac{1}{2}-\frac{r-r/2}{d}.

Choosing p2=2+κ,p1=2(2+κ)/κp_{2}=2+\kappa,p_{1}=2(2+\kappa)/\kappa with κ>0\kappa>0 sufficiently small we see that (4.6) is nonempty. Now notice that

(4.7) θρDr(φnφ)L2(d)Dr(θρ(φnφ)L2(d)+[θρ,Dr](φnφ)L2(d)\|\theta_{\rho}D^{r}(\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}\leq\|D^{r}(\theta_{\rho}(\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}+\|[\theta_{\rho},D^{r}](\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}

and we conclude that (4.4) is true for 0r<min(2,S)0\leq r<\min(2,S) thanks to (4.3) and (4.5).

Now we consider the case 2r<S.2\leq r<S. We have Dr=Dr1(Δ),D^{r}=D^{r_{1}}(-\Delta)^{\ell}, where 1\ell\geq 1 is integer and 0<r1<2.0<r_{1}<2. Then the commutator relation

[A,BC]=[A,B]C+B[A,C][A,BC]=[A,B]C+B[A,C]

implies

[θρ,Dr]=[θρ,Dr1](Δ)+Dr1[θρ,(Δ)].[\theta_{\rho},D^{r}]=[\theta_{\rho},D^{r_{1}}](-\Delta)^{\ell}+D^{r_{1}}[\theta_{\rho},(-\Delta)^{\ell}].

In fact, we have the relation

θρDr(φnφ)=[θρ,Dr1]((Δ)(φnφ))+Dr1[θρ,(Δ)](φnφ)\theta_{\rho}D^{r}(\varphi_{n}-\varphi)=[\theta_{\rho},D^{r_{1}}]((-\Delta)^{\ell}(\varphi_{n}-\varphi))+D^{r_{1}}[\theta_{\rho},(-\Delta)^{\ell}](\varphi_{n}-\varphi)

and we use (4.5) so that

[θρ,Dr1](Δ)(φnφ)L2(d)\|[\theta_{\rho},D^{r_{1}}](-\Delta)^{\ell}(\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}\leq
O(ρr1)(Δ)(φnφ)L2(d)+\leq O(\rho^{-r_{1}})\|(-\Delta)^{\ell}(\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}+
+O(ρr1/4)Dr1+2(φnφ)L2(d)=o(1)+O(\rho^{-r_{1}/4})\|D^{r_{1}+2\ell}(\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}=o(1)

for ρ.\rho\to\infty.

The term

Dr1[θρ,(Δ)](φnφ)D^{r_{1}}[\theta_{\rho},(-\Delta)^{\ell}](\varphi_{n}-\varphi)

can be evaluated pointwise via the classical Leibnitz rule and then via the fractional Leibnitz rule as follows

Dr1[θρ,(Δ)](φnφ)L2(d)\|D^{r_{1}}[\theta_{\rho},(-\Delta)^{\ell}](\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}\lesssim
1|α|,|α|+|β|=2Dr1(xαθρ)xβ(φnφ)L2(d)O(ρ1)φnφHr(d).\lesssim\sum_{1\leq|\alpha|,|\alpha|+|\beta|=2\ell}\|D^{r_{1}}(\partial_{x}^{\alpha}\theta_{\rho})\partial_{x}^{\beta}(\varphi_{n}-\varphi)\|_{L^{2}(\mathbb{R}^{d})}\lesssim O(\rho^{-1})\|\varphi_{n}-\varphi\|_{H^{r}(\mathbb{R}^{d})}.

Summing up, we conclude that (4.4) holds in case r[0,S).r\in[0,S).

To conclude that the embedding is compact it remains to show that also Dr(φnφ)L2(Bρc)2ϵ\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2}(B_{\rho}^{c})}^{2}\leq\epsilon. To this purpose we first use the pointwise decay in terms of homogeneous Sobolev norm, see [7]. Given rr there exists 0<δ<d120<\delta<\frac{d-1}{2} with r+12+δ<Sr+\frac{1}{2}+\delta<S such that

(4.8) |Dr(φnφ)(x)|C|x|γφnφH˙r+12+δ(d)C|x|γ|D^{r}(\varphi_{n}-\varphi)(x)|\leq\frac{C}{|x|^{\gamma}}||\varphi_{n}-\varphi||_{\dot{H}^{r+\frac{1}{2}+\delta}(\mathbb{R}^{d})}\lesssim\frac{C}{|x|^{\gamma}}

with γ=d12δ\gamma=\frac{d-1}{2}-\delta. Secondly we use that p0<2p_{0}<2, i.e. that p=2p=2 is non endpoint. By Theorem 1.1 there exists δ0>0\delta_{0}>0 sufficiently small such that DrφnD^{r}\varphi_{n} is uniformly bounded in L2δ0(d)L^{2-\delta_{0}}(\mathbb{R}^{d}) and the same holds hence for DrφD^{r}\varphi and Dr(φnφ)D^{r}(\varphi_{n}-\varphi). As a consequence we have

Dr(φnφ)L2(Bρc)2=Bρc|Dr(φnφ)|δ0|Dr(φnφ)|2δ0𝑑x\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2}(B_{\rho}^{c})}^{2}=\int_{B_{\rho}^{c}}|D^{r}(\varphi_{n}-\varphi)|^{\delta_{0}}|D^{r}(\varphi_{n}-\varphi)|^{2-\delta_{0}}dx\leq
Cργδ0Dr(φnφ)L2δ0(Bρc)2δ\leq\frac{C}{\rho^{\gamma}}^{\delta_{0}}\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2-\delta_{0}}(B_{\rho}^{c})}^{2-\delta}

with

Dr(φnφ)L2δ0(Bρc)Dr(φnφ)L2δ0(d)=O(1).\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2-\delta_{0}}(B_{\rho}^{c})}\leq\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2-\delta_{0}}(\mathbb{R}^{d})}=O(1).

This proves that Dr(φnφ)L2(Bρ)2ϵ\|D^{r}(\varphi_{n}-\varphi)\|_{L^{2}(B_{\rho})}^{2}\lesssim\epsilon and hence that the embedding is compact.

5. Proof of Theorem 1.3

For easier reference we state the following.

Lemma 5.1 (pqr Lemma [9]).

Let 1p<q<r1\leq p<q<r\leq\infty and let α,β,γ>0\alpha,\beta,\gamma>0. Then there are constants η,c>0\eta,c>0 such that for any measurable function fLp(X)Lr(X)f\in L^{p}(X)\cap L^{r}(X), XX a measure space, with

fLppα,fLqqβ,fLrrγ,\|f\|_{L^{p}}^{p}\leq\alpha,\quad\|f\|_{L^{q}}^{q}\geq\beta,\quad\|f\|_{L^{r}}^{r}\leq\gamma,\quad

one has (with |||\cdot| denoting the underlying measure on XX)

(5.1) |{xX:|f(x)|>η}|c.\left|\{x\in X:\ |f(x)|>\eta\}\right|\geq c\,.
Lemma 5.2 (Compactness up to translations in H˙s\dot{H}^{s} [2]).

Let s>0s>0, 1<p<1<p<\infty and unH˙s(d)Lp(d)u_{n}\in\dot{H}^{s}(\mathbb{R}^{d})\cap L^{p}(\mathbb{R}^{d}) be a sequence with

(5.2) supn(unH˙s(d)+unLp(d))<\sup_{n}\left(\|u_{n}\|_{\dot{H}^{s}(\mathbb{R}^{d})}+\|u_{n}\|_{L^{p}(\mathbb{R}^{d})}\right)<\infty

and, for some η>0\eta>0, (with |||\cdot| denoting Lebesgue measure)

(5.3) infn|{|un|>η}|>0.\inf_{n}\left|\{|u_{n}|>\eta\}\right|>0\,.

Then there is a sequence (xn)d(x_{n})\subset\mathbb{R}^{d} such that a subsequence of un(+xn)u_{n}(\cdot+x_{n}) has a weak limit u0u\not\equiv 0 in H˙s(d)Lp(d)\dot{H}^{s}(\mathbb{R}^{d})\cap L^{p}(\mathbb{R}^{d}).

The strategy to prove Theorem 1.3 follows the one developed in [2]. First we aim to show that the maximum of

W(φ)=DrφL2(d)φL2(d)1rSDSφL2(d)rSφHrad,+S,r(d),W(\varphi)=\frac{\|D^{r}\varphi\|_{L^{2}(\mathbb{R}^{d})}}{\|\varphi\|_{L^{2}(\mathbb{R}^{d})}^{1-\frac{r}{S}}\,\|D^{S}\varphi\|_{L^{2}(\mathbb{R}^{d})}^{\frac{r}{S}}}\ \ \ \varphi\in H^{S,r}_{rad,+}(\mathbb{R}^{d}),

is achieved in Hrad,+S,r(d)H^{S,r}_{rad,+}(\mathbb{R}^{d}) . Let us consider a maximizing sequence φn\varphi_{n}. Since WW is invariant under homogeneity φ(x)λφ(x)\varphi(x)\mapsto\lambda\varphi(x) and scaling φφ(λx)\varphi\mapsto\varphi(\lambda x) for any λ>0\lambda>0, we can choose a maximizing sequence φn\varphi_{n} such that

(5.4) DrφnL2(d)=Crad,+(r,S,2,d)+o(1)\|D^{r}\varphi_{n}\|_{L^{2}(\mathbb{R}^{d})}=C_{rad,+}(r,S,2,d)+o(1)

and

(5.5) φnL2(d)=DSφnL2(d)=1.\|\varphi_{n}\|_{L^{2}(\mathbb{R}^{d})}=\|D^{S}\varphi_{n}\|_{L^{2}(\mathbb{R}^{d})}=1\,.

The key observation is that, since we are looking at a non-endpoint case (i.e. p0<2p_{0}<2), there exists ϵ>0\epsilon>0 such that from inequality (1.9) we infer that

(5.6) supnmax{DrφnL2ϵ(d),DrφnL2+ϵ(d)}<.\sup_{n}\max\left\{\|D^{r}\varphi_{n}\|_{L^{2-\epsilon}(\mathbb{R}^{d})},\|D^{r}\varphi_{n}\|_{L^{2+\epsilon}(\mathbb{R}^{d})}\right\}<\infty\,.

The pqrpqr-lemma (Lemma 5.1) now implies that

(5.7) infn|{|Drφn|>η}|>0.\inf_{n}\left|\{|D^{r}\varphi_{n}|>\eta\}\right|>0.

Next, we apply the compactness modulo translations lemma (Lemma 5.2) to the sequence (Drφn)(D^{r}\varphi_{n}). This sequence is bounded in H˙Sr\dot{H}^{S-r} by (5.5), and (5.2) and (5.3) are satisfied by (5.4) and (5.7). Thus possibly after passing to a subsequence, we have Drφnψ0D^{r}\varphi_{n}\rightharpoonup\psi\not\equiv 0 in HSr(d)H^{S-r}(\mathbb{R}^{d}). By the fact the embedding is compact we deduce that φn(x)ψ0\varphi_{n}(x)\rightarrow\psi\not\equiv 0 in H˙r(d)\dot{H}^{r}(\mathbb{R}^{d}) and hence ψ\psi is a maximizer for WW.
Now we conclude showing that Crad,+(r,S,2,d)<1C_{rad,+}(r,S,2,d)<1.
Indeed if the best constant is Crad,+(r,S,2,d)=1C_{rad,+}(r,S,2,d)=1, the maximizer ψ\psi achieves the equality in Hölder’s inequality, which means

(5.8) d|ξ|2r|ψ^|2𝑑ξ=d|ψ^|22rS|ξ|2r|ψ^|2rS𝑑ξ=\displaystyle\int_{\mathbb{R}^{d}}|\xi|^{2r}|\widehat{\psi}|^{2}d\xi=\int_{\mathbb{R}^{d}}|\widehat{\psi}|^{2-\frac{2r}{S}}|\xi|^{2r}|\widehat{\psi}|^{\frac{2r}{S}}d\xi=
=(d|ψ^|2𝑑ξ)1rS(d|ξ|2S|ψ^|2𝑑ξ)rS,\displaystyle=\left(\int_{\mathbb{R}^{d}}|\widehat{\psi}|^{2}d\xi\right)^{1-\frac{r}{S}}\left(\int_{\mathbb{R}^{d}}|\xi|^{2S}|\widehat{\psi}|^{2}d\xi\right)^{\frac{r}{S}},

where we used as conjugated exponents SSr\frac{S}{S-r} and Sr\frac{S}{r}. Now we recall that if fLp(d)f\in L^{p}(\mathbb{R}^{d}) and gLq(d)g\in L^{q}(\mathbb{R}^{d}) with pp and qq conjugated exponents achieve the equality in Hölder’s inequality then |f|p|f|^{p} and |g|q|g|^{q} shall be linearly dependent, i.e. for a suitable μ,|f|p=μ|g|q\mu,|f|^{p}=\mu|g|^{q} almost everywhere. Therefore, calling f=|ψ^|22rSf=|\widehat{\psi}|^{2-\frac{2r}{S}} and g=|ξ|2r|ψ^|2rSg=|\xi|^{2r}|\widehat{\psi}|^{\frac{2r}{S}}, the maximizer ψ\psi should satisfy |ψ^|2=μ|ξ|2S|ψ^|2|\widehat{\psi}|^{2}=\mu|\xi|^{2S}|\widehat{\psi}|^{2} for a suitable μ\mu which drives to the contradiction ψ^=0.\widehat{\psi}=0.

6. Appendix.

The statement of Lemma 2.4 can be found in [12]. Somehow, due to the fact that in the original paper the proof of Lemma 2.4 is not easy readable, being a part of a more general statement, we give an alternative short proof.

Proof of Lemma 2.4.

We divide the integration domain in two subdomains:

Ω={|x|<|y|/2}{|x|>2|y|}\Omega=\{|x|<|y|/2\}\cup\{|x|>2|y|\}

and its complementary set Ωc.\Omega^{c}. In Ω\Omega we use

|xy|max(|x|,|y|)2|x-y|\geq\frac{\max(|x|,|y|)}{2}

and via

(1+|xy|2)(d1)/2(1+(max(|x|,|y|))2)(d1)/2(1+|x-y|^{2})^{(d-1)/2}\gtrsim(1+(\max(|x|,|y|))^{2})^{(d-1)/2}\geq
max(|x|,|y|))(d1)|x|(d1+b)|y|b\geq\max(|x|,|y|))^{(d-1)}\gtrsim|x|^{(d-1+b)}|y|^{-b}

with d1+b>0,b>0d-1+b>0,-b>0 we deduce

1(1+|xy|2)γ/2=1(1+|xy|2)(d1)/21(1+|xy|2)(γd+1)/2\frac{1}{(1+|x-y|^{2})^{\gamma/2}}=\frac{1}{(1+|x-y|^{2})^{(d-1)/2}}\frac{1}{(1+|x-y|^{2})^{(\gamma-d+1)/2}}
1|x|d1+b|y|b1(1+|xy|2)(γd+1)/21|x|d1+b|y|b.\lesssim\frac{1}{|x|^{d-1+b}}|y|^{b}\frac{1}{(1+|x-y|^{2})^{(\gamma-d+1)/2}}\leq\frac{1}{|x|^{d-1+b}}|y|^{b}.

These estimates imply

(6.1) |Ωf(y)dy(1+|xy|2)γ/2|1|x|d1+b|y|bfL1(d).\left|\int_{\Omega}\frac{f(y)dy}{(1+|x-y|^{2})^{\gamma/2}}\right|\lesssim\frac{1}{|x|^{d-1+b}}\left\||y|^{b}f\right\|_{L^{1}(\mathbb{R}^{d})}.

For the complementary domain Ωc\Omega^{c} we use spherical coordinates x=rθ,y=ρω,x=r\theta,y=\rho\omega, where r=|x|,ρ=|y|.r=|x|,\rho=|y|. We have to estimate

Ωcf(y)dy(1+|xy|2)γ/2=r/22rK(r,ρ)f(ρ)ρd1𝑑ρ,\int_{\Omega^{c}}\frac{f(y)dy}{(1+|x-y|^{2})^{\gamma/2}}=\int_{r/2}^{2r}K(r,\rho)f(\rho)\rho^{d-1}d\rho,

where

(6.2) K(r,ρ)=Kθ,γ(r,ρ)=𝕊d1(1+|rθρω|2)γ/2𝑑ω.K(r,\rho)=K_{\theta,\gamma}(r,\rho)=\int_{\mathbb{S}^{d-1}}(1+|r\theta-\rho\omega|^{2})^{-\gamma/2}d\omega.

To get the desired estimate

(6.3) |Ωcf(y)dy(1+|xy|2)γ/2|1|x|d1+b|y|bfL1(d)\left|\int_{\Omega^{c}}\frac{f(y)dy}{(1+|x-y|^{2})^{\gamma/2}}\right|\lesssim\frac{1}{|x|^{d-1+b}}\left\||y|^{b}f\right\|_{L^{1}(\mathbb{R}^{d})}

it is sufficient to check the pointwise estimate

(6.4) K(r,ρ)r(d1+b)ρbr(d1)for r/2ρ2r.K(r,\rho)\lesssim r^{-(d-1+b)}\rho^{b}\sim r^{-(d-1)}\ \ \mbox{for $r/2\leq\rho\leq 2r$.}

To deduce this pointwise estimate of the kernel KK we note first that KK does not depend on θ\theta so we can take θ=ed=(0,,0,1)\theta=e_{d}=(0,\cdots,0,1) and ω=(ωsinφ,cosφ),\omega=(\omega^{\prime}\sin\varphi,\cos\varphi), ω𝕊d2\omega^{\prime}\in\mathbb{S}^{d-2} and get

K(r,ρ)=c0πsind2φdφ(1+r2+ρ22rρcosφ)γ/2.K(r,\rho)=c\int_{0}^{\pi}\frac{\sin^{d-2}\varphi d\varphi}{(1+r^{2}+\rho^{2}-2r\rho\cos\varphi)^{\gamma/2}}.

Using the relation

(1+r2+ρ22rρcosφ)=1+(rρ)2+rρsin2(φ/2),(1+r^{2}+\rho^{2}-2r\rho\cos\varphi)=1+(r-\rho)^{2}+r\rho\sin^{2}(\varphi/2),

we can use the

(1+r2+ρ22rρcosφ)rρr2(1+r^{2}+\rho^{2}-2r\rho\cos\varphi)\gtrsim r\rho\sim r^{2}

when ρr\rho\sim r and φ\varphi is not close to 0,0, say φ(π/4,π).\varphi\in(\pi/4,\pi). Then we get

π/4πsind2φdφ(1+r2+ρ22rρcosφ)γ/2π/4πrγ𝑑φrγrd+1.\int_{\pi/4}^{\pi}\frac{\sin^{d-2}\varphi d\varphi}{(1+r^{2}+\rho^{2}-2r\rho\cos\varphi)^{\gamma/2}}\lesssim\int_{\pi/4}^{\pi}r^{-\gamma}d\varphi\lesssim r^{-\gamma}\leq r^{-d+1}.

For φ\varphi close to 0, say φπ/4\varphi\leq\pi/4 we use

sind2φ(1+r2+ρ22rρcosφ)γ/2φd2(1+rρφ2)γ/2.\frac{\sin^{d-2}\varphi}{(1+r^{2}+\rho^{2}-2r\rho\cos\varphi)^{\gamma/2}}\lesssim\frac{\varphi^{d-2}}{(1+r\rho\varphi^{2})^{\gamma/2}}.

In this way, making the change of variable rφ=ηr\varphi=\eta we get

0π/4φd2dφ(1+rρφ2)γ/20φd2dφ(1+r2φ2)γ/2\int_{0}^{\pi/4}\frac{\varphi^{d-2}d\varphi}{(1+r\rho\varphi^{2})^{\gamma/2}}\lesssim\int_{0}^{\infty}\frac{\varphi^{d-2}d\varphi}{(1+r^{2}\varphi^{2})^{\gamma/2}}\leq
rd+10ηd2dη(1+η2)γ/2rd+1\leq r^{-d+1}\int_{0}^{\infty}\frac{\eta^{d-2}d\eta}{(1+\eta^{2})^{\gamma/2}}\lesssim r^{-d+1}

in view of ρr\rho\sim r and γ>d1.\gamma>d-1. Taking together the above estimates of the integrals over (0,π/4)(0,\pi/4) and (π/4,π)(\pi/4,\pi), we arrive at (6.4).

This completes the proof of the Lemma.

References

  • [1] H. Bahouri, J.-Y. Chemin, R. Danchin, Fourier Analysis and Nonlinear Partial Differential Equations, Springer, 2011.
  • [2] J.Bellazzini, R.L. Frank, N. Visciglia, Maximizers for Gagliardo-Nirenberg inequalities and related non-local problems, Math. Annalen (2014), no. 3-4, 653 – 673.
  • [3] J. Bellazzini, M. Ghimenti, C. Mercuri, V. Moroz, J. Van Schaftingen, Sharp Gagliardo-Nirenberg inequalities in fractional Coulomb-Sobolev spaces, Trans. Amer. Math. Soc. 370, 11, 8285 – 8310 (2018)
  • [4] J. Bellazzini, M. Ghimenti, and T. Ozawa, Sharp lower bounds for Coulomb energy, Math. Res. Lett. 23 (2016), no. 3, 621–632
  • [5] H. Brezis, P. Mironescu, Gagliardo-Nirenberg, composition and products in fractional Sobolev spaces. Dedicated to the memory of Tosio Kato. J. Evol. Equ. 1 (2001), no. 4, 387 – 404
  • [6] H. Brezis and P. Mironescu, Where Sobolev interacts with Gagliardo-Nirenberg, J. Funct. Anal. 277 (2019), no. 8, 2839 – 2864.
  • [7] Y. Cho, T. Ozawa, Sobolev inequality with symmetry, Commun. Contemp. Math. 11 (2009), no. 3, 355–365
  • [8] P.L. De Nápoli Symmetry breaking for an elliptic equation involving the Fractional Laplacian, Differential Integral Equations 31 (1/2) 75 - 94, January/February (2018).
  • [9] J. Fröhlich, E. H. Lieb, M. Loss, Stability of Coulomb systems with magnetic fields. I. The one-electron atom., Comm. Math. Phys. 104 (1986), no. 2, 251–270.
  • [10] K. Fujiwara, V.Georgiev, T. Ozawa, Higher order fractional Leibniz rule. J. Fourier Anal. Appl. 24 (2018), no. 3, 650 – 665.
  • [11] H. Hajaiej,L. Molinet, T. Ozawa, and B. Wang, Necessary and sufficient conditions for the fractional Gagliardo-Nirenberg inequalities and applications to Navier-Stokes and generalized boson equations, Harmonic analysis and nonlinear partial differential equations, 159 – 175, RIMS Kôkyûroku Bessatsu, B26, Res. Inst. Math. Sci. (RIMS), Kyoto, 2011
  • [12] P. D’Ancona, R. Luca`\grave{a}, Stein-Weiss and Caffarelli-Kohn-Nirenberg inequalities with angular integrability, J. Math. Anal. Appl., 388(2): 1061–1079, (2012)
  • [13] C. Mercuri, V. Moroz, J. Van Schaftingen Groundstates and radial solutions to nonlinear Schrödinger-Poisson-Slater equations at the critical frequency, Calc. Var. 55, 146 (2016)
  • [14] D. Ruiz, On the Schrödinger-Poisson-Slater system: behavior of minimizers, radial and nonradial cases, Arch. Ration. Mech. Anal. 198 (2010), no. 1, 349–368
  • [15] W. Sickel, L. Skrzypczak, Radial subspaces of Besov and Lizorkin-Triebel classes: extended Strauss lemma and compactness of embeddings, J. Fourier Anal. Appl. 6 (2000), no. 6, 639 – 662.
  • [16] W. Sickel, L. Skrzypczak, On the Interplay of Regularity and Decay in Case of Radial Functions II. Homogeneous Spaces, J Fourier Anal Appl (2012) 18:548 – 582 DOI 10.1007/s00041-011-9205-2
  • [17] W. A. Strauss. Existence of Solitary Waves in Higher Dimensions. Comm. Math. Phys. 55 (1977), 149–162.