Home Optimal Functional Inequalities for Fractional Operators on the Sphere and Applications
Article Open Access

Optimal Functional Inequalities for Fractional Operators on the Sphere and Applications

  • Jean Dolbeault and An Zhang EMAIL logo
Published/Copyright: October 18, 2016

Abstract

This paper is devoted to the family of optimal functional inequalities on the n-dimensional sphere 𝕊n, namely

F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 𝖢 q , s 𝕊 n F s F 𝑑 μ for all F H s / 2 ( 𝕊 n ) ,

where s denotes a fractional Laplace operator of order s(0,n), q[1,2)(2,q], q=2nn-s is a critical exponent, and dμ is the uniform probability measure on 𝕊n. These inequalities are established with optimal constants using spectral properties of fractional operators. Their consequences for fractional heat flows are considered. If q>2, these inequalities interpolate between fractional Sobolev and subcritical fractional logarithmic Sobolev inequalities, which correspond to the limit case as q2. For q<2, the inequalities interpolate between fractional logarithmic Sobolev and fractional Poincaré inequalities. In the subcritical range q<q, the method also provides us with remainder terms which can be considered as an improved version of the optimal inequalities. The case s(-n,0) is also considered. Finally, weighted inequalities involving the fractional Laplacian are obtained in the Euclidean space, by using the stereographic projection.

1 Introduction and Main Results

Let us consider the unit sphere 𝕊n with n1 and assume that the measure dμ is the uniform probability measure, which is also the measure induced on 𝕊n by Lebesgue’s measure on n+1, up to a normalization constant. With λ(0,n), p=2n2n-λ(1,2) or equivalently λ=2np where 1p+1p=1, according to [38], the sharp Hardy–Littlewood–Sobolev inequality on 𝕊n reads

(1.1) 𝕊 n × 𝕊 n F ( ζ ) | ζ - η | - λ F ( η ) 𝑑 μ ( ζ ) 𝑑 μ ( η ) Γ ( n ) Γ ( n - λ 2 ) 2 λ Γ ( n 2 ) Γ ( n p ) F L p ( 𝕊 n ) 2 .

For the convenience of the reader, the definitions of all parameters, their ranges and their relations have been collected in Appendix C.

By the Funk–Hecke formula, the left-hand side of the inequality can be written as

(1.2) 𝕊 n × 𝕊 n F ( ζ ) | ζ - η | - λ F ( η ) 𝑑 μ ( ζ ) 𝑑 μ ( η ) = Γ ( n ) Γ ( n - λ 2 ) 2 λ Γ ( n 2 ) Γ ( n p ) k = 0 Γ ( n p ) Γ ( n p + k ) Γ ( n p ) Γ ( n p + k ) 𝕊 n | F ( k ) | 2 𝑑 μ ,

where F=k=0F(k) is a decomposition on spherical harmonics, so that F(k) is a spherical harmonic function of degree k. See [33, Section 4] for details on the computations and, e.g., [42] for further related results. With the above representation, inequality (1.1) is equivalent to

(1.3) k = 0 Γ ( n p ) Γ ( n p + k ) Γ ( n p ) Γ ( n p + k ) 𝕊 n | F ( k ) | 2 𝑑 μ F L p ( 𝕊 n ) 2 .

By duality, with q=q(s) defined by

(1.4) q = 2 n n - s

or equivalently s=n(1-2q), we obtain the fractional Sobolev inequality on 𝕊n:

(1.5) F L q ( 𝕊 n ) 2 𝕊 n F 𝒦 s F 𝑑 μ for all F H s / 2 ( 𝕊 n )

for any s(0,n), where

(1.6) 𝕊 n F 𝒦 s F 𝑑 μ := k = 0 γ k ( n q ) 𝕊 n | F ( k ) | 2 𝑑 μ

and

γ k ( x ) := Γ ( x ) Γ ( n - x + k ) Γ ( n - x ) Γ ( x + k ) .

With s(0,n) the exponent q is in the range (2,). Inequalities (1.1) and (1.5) are related by q=p so that

p = 2 n n + s and λ = n - s .

We shall refer to q=q(s) given by (1.4) as the critical case and our purpose is to study the whole range of the subcritical interpolation inequalities

(1.7) F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 𝖢 q , s 𝕊 n F s F 𝑑 μ for all F H s / 2 ( 𝕊 n )

for any q[1,2)(2,q], where

s := 1 κ n , s ( 𝒦 s - Id ) with κ n , s := Γ ( n q ) Γ ( n - n q ) = Γ ( n - s 2 ) Γ ( n + s 2 ) .

If q=q, inequalities (1.5) and (1.7) are identical, the optimal constant in (1.7) is 𝖢q,s=κn,sq-2, and we recall that (1.5) is equivalent to the fractional Sobolev inequality on the Euclidean space (see the proof of Theorem 1.6 in Section 3 for details). The usual conformal fractional Laplacian is defined by

𝒜 s := 1 κ n , s 𝒦 s = s + 1 κ n , s Id .

For brevity, we shall say that s is the fractional Laplacian of order s, or simply the fractional Laplacian.

We observe that γ0(nq)-1=0 and γ1(nq)-1=q-2. A straightforward computation gives

𝕊 n F s F 𝑑 μ := k = 1 δ k ( n q ) 𝕊 n | F ( k ) | 2 𝑑 μ ,

where the spectrum of s is given by

δ k ( x ) := Γ ( n - x + k ) Γ ( x + k ) - Γ ( n - x ) Γ ( x ) .

The case corresponding to s=2 and n3, where 1κn,2=14n(n-2), 2=-Δ, 𝒜2=-Δ+14n(n-2), and Δ stands for the Laplace–Beltrami operator on 𝕊n, has been considered by W. Beckner; in [5, p. 233, (35)] he observed that

δ k ( n q ) δ k ( n q ) = k ( k + n - 1 )

if q(2,q(2)], where q=q(2)=2nn-2 and (k(k+n-1))k is the sequence of the eigenvalues of -Δ according to, e.g., [7]. This establishes the interpolation inequality

(1.8) F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 n F L 2 ( 𝕊 n ) 2 for all F H 1 ( 𝕊 n ) ,

where 𝖢q,2=1n is the optimal constant; see [5, (35), Theorem 4] for details. An earlier proof of the inequality with optimal constant can be found in [8, Corollary 6.2], with a proof based on rigidity results for elliptic partial differential equations. Our main result generalizes the interpolation inequalities (1.8) to the case of the fractional operators s, and relies on W. Beckner’s approach. In particular, as in [5], we characterize the optimal constant 𝖢q,s in (1.7) using a spectral gap property.

After dividing both sides of (1.8) by (q-2) we obtain an inequality which, for s=2, also makes sense for any q[1,2). When q=1, this is actually a variant of the Poincaré inequality (or, to be precise, the Poincaré inequality written for |F|), and the range q>1 has been studied using the carré du champ method, also known as the Γ2 calculus, by D. Bakry and M. Emery in [3]. Actually their method covers the range corresponding to 1q< if n=1 and

1 q 2 # := 2 n 2 + 1 ( n - 1 ) 2 if n 2 .

In the special case q=2, the left-hand side of (1.8) has to be replaced by the entropy

𝕊 n F 2 log ( F 2 F L 2 ( 𝕊 n ) 2 ) 𝑑 μ .

Still under the condition that s=2, the whole range 1q< when n=2, and 1q2nn-2 if n3 can be covered using nonlinear flows as shown in [21, 24, 25].

All these considerations motivate our first result, which generalizes known results for 2=-Δ to the case of the fractional Laplacian s.

Theorem 1.1

Let n1, s(0,n], q[1,2)(2,q], with q given by (1.4) if s<n, and q[1,2)(2,) if s=n. Inequality (1.7) holds with sharp constant

𝖢 q , s = n - s 2 s Γ ( n - s 2 ) Γ ( n + s 2 ) .

With our previous notations, this amounts to

𝖢 q , s = κ n , s q - 2 = n - s 2 s κ n , s .

Remarkably, 𝖢q,s is independent of q. Equality in (1.7) is achieved by constant functions. The issue of the optimality of 𝖢q,s is henceforth somewhat subtle. If we define the functional

(1.9) 𝒬 [ F ] := ( q - 2 ) 𝕊 n F s F 𝑑 μ F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2

on the subset s/2 of the functions in Hs/2(𝕊n) which are not almost everywhere constant, then 𝖢q,s can be characterized by

𝖢 q , s - 1 = inf F s / 2 𝒬 [ F ] .

This minimization problem will be discussed in Section 4.

Our key estimate is a simple convexity observation that is stated in Lemma 2.2. The optimality in (1.7) is obtained by performing a linearization, which corresponds to an asymptotic regime as we shall see in Section 2.1. Technically, this is the reason why we are able to identify the optimal constant. The asymptotic regime can be investigated using a flow. Indeed, a first consequence of Theorem 1.1 is that we may apply entropy methods to the generalized fractional heat flow

(1.10) u t - q ( u 1 - 1 q ( - Δ ) - 1 s u 1 q ) = 0 .

Notice that (1.10) is a 1-homogeneous equation, but that it is nonlinear when q1 and s2. Let us define a generalized entropy by

q [ u ] := 1 q - 2 [ ( 𝕊 n u 𝑑 μ ) 2 q - 𝕊 n u 2 q 𝑑 μ ] .

It is straightforward to check that for any positive solution to (1.10) which is smooth enough and has sufficient decay properties as |x|+, we have

d d t q [ u ( t , ) ] = - 2 𝕊 n u 1 q ( - Δ ) - 1 s u 1 q d μ = - 2 𝕊 n u 1 q s u 1 q d μ ,

so that by applying (1.7) to F=u1q, we obtain the exponential decay of q[u(t,)].

Corollary 1.2

Let n1, s(0,n], q[1,2)(2,q] if s<n, with q given by (1.4), and q[1,2)(2,) if s=n. If u is a positive function in C1(R+;L(Sn)) such that u1qC1(R+;Hs/2(Sn)) and if u solves (1.10) on Sn with initial datum u0>0, then

q [ u ( t , ) ] q [ u 0 ] e - 2 𝖢 q , s - 1 t for all t 0 .

The exponential rate is determined by the asymptotic regime as t+. The value of the optimal constant𝖢q,s is indeed determined by the spectral gap of the linearized problem around non-zero constant functions. From the expression of (1.10), which is not even a linear equation whenever s2, we observe that the interplay of optimal fractional inequalities and fractional diffusion flows is not straightforward, while for s=2 the generalized entropy q enters in the framework of the so-called φ-entropies and is well understood in terms of gradient flows; see for instance [2, 13, 28]. When s=2, it is also known from [3] that heat flows can be used in the framework of the carré du champ method to establish the inequalities at least for exponents in the range q2# if n2, and that the whole subcritical range of exponents can be covered using nonlinear diffusions as in [21, 24, 25] (and also the critical exponent if n3). Even better, rigidity results, that is, uniqueness of positive solutions (which are therefore constant functions), follow by this technique. So far there is no analogue in the case of fractional operators, except for one example found in [12] when n=1.

When s=2, the carré du champ method provides us with an integral remainder term and, as a consequence, with an improved version of (1.7). As we shall see, our proof of Theorem 1.1 establishes another improved inequality by construction; see Corollary 2.3. This also suggests another direction, which is more connected with the duality that relates (1.1) and (1.5). Let us describe the main idea. The operator 𝒦s is positive definite and we can henceforth consider 𝒦s1/2 and 𝒦s-1. Moreover, using (1.2) and (1.6), we know that

𝕊 n × 𝕊 n G ( ζ ) | ζ - η | - λ G ( η ) 𝑑 μ ( ζ ) 𝑑 μ ( η ) = Γ ( n ) Γ ( s 2 ) 2 λ Γ ( n 2 ) Γ ( n + s 2 ) 𝕊 n G 𝒦 s - 1 G 𝑑 μ .

Expanding the square

𝕊 n | 𝒦 s 1 / 2 F - 𝒦 s - 1 / 2 G | 2 𝑑 μ

with G=Fq-1 so that FG=Fq=Gp where q and p are Hölder conjugates, we get a comparison of the difference of the two terms which show up in (1.1) and (1.5) and, as a result, an improved fractional Sobolev inequality on 𝕊n. The reader interested in the details of the proof is invited to consult [27] for a similar result.

Proposition 1.3

Let n1 and s(0,n). Consider q given by (1.4), p=q=2nn+s and λ=n-s. For any FHs/2(Sn) if G=Fq-1, then

G L p ( 𝕊 n ) 2 - 2 λ Γ ( n 2 ) Γ ( n + s 2 ) Γ ( n ) Γ ( s 2 ) 𝕊 n × 𝕊 n G ( ζ ) | ζ - η | - λ G ( η ) 𝑑 μ ( ζ ) 𝑑 μ ( η ) F L q ( 𝕊 n ) 2 ( q - 2 ) ( 𝕊 n F 𝒦 s F 𝑑 μ - F L q ( 𝕊 n ) 2 ) .

Still in the critical case q=q, by using the fractional Yamabe flow and taking inspiration from [23, 27, 37, 36, 40], it is possible to give improvements of the above inequality and in particular improve on the constant which relates the left- and the right-hand sides of the inequality in Proposition 1.3. We will not go further in this direction because of the delicate regularity properties of the fractional Yamabe flow and because so far the method does not allow to characterize the best constant in the improvement. Let us mention that, in the critical case q=q, further estimates of Bianchi–Egnell type have also been obtained in [15, 40] for fractional operators. In this paper, we shall rather focus on the subcritical range. It is however clear that there is still space for further improvements, or alternative proofs of (1.5) which rely neither on rearrangements as in [38] nor on inversion symmetry as in [31, 32, 33], for the simple reason that our method fails to provide us with a proof of the Bianchi–Egnell estimates in the critical case.

For completeness let us quote a few other related results. Symmetrization techniques and the method of competing symmetries are both very useful to identify the optimal functions; the interested reader is invited to refer to [39] and [10], respectively, when s=2. In this paper, we shall use notations inspired by [5], but at this point it is worth mentioning that in [5] the emphasis is put on logarithmic Hardy–Littlewood–Sobolev inequalities and their dual counterparts, which are n-dimensional versions of the Moser–Trudinger–Onofri inequalities. Some of these results were obtained simultaneously in [11] with some additional insight on optimal functions gained from rearrangements and from the method of competing symmetries. Concerning observations on duality, we refer to the introduction of [11], which clearly refers the earlier contributions of various authors in this area. For more recent considerations on n-dimensional Moser–Trudinger–Onofri inequalities see, e.g., [19].

Section 2 is devoted to the proof of Theorem 1.1. As already said, we shall take advantage of the subcritical range to obtain remainder terms and improved inequalities. Improvements in the subcritical range have been obtained in the case of non-fractional interpolation inequalities in the context of fast diffusion equations in [29, 30]. In this paper we shall simply take into account the terms which appear by difference in the proof of Theorem 1.1; see Corollary 2.3 in Section 2.3. Although this approach does not provide us with an alternative proof of the optimality of the constant 𝖢q,s in (1.7), variational methods will be applied in Section 4 in order to explain a posteriori why the value of the optimal value of 𝖢q,s is determined by the spectral gap of a linearized problem. Some useful information on the spectrum of s is detailed in Appendix A.

Our next result is devoted to the singular case of inequality (1.7) corresponding to the limit as q=2. We establish a family of sharp fractional logarithmic Sobolev inequalities in the subcritical range.

Corollary 1.4

Let s(0,n]. Then we have the sharp logarithmic Sobolev inequality

(1.11) 𝕊 n | F | 2 log ( | F | F 2 ) 𝑑 μ 𝖢 2 , s 𝕊 n F s F 𝑑 μ for all F H s / 2 ( 𝕊 n ) .

Equality is achieved only by constant functions, and C2,s=n-s2sκn,s is optimal.

This result completes the picture of Theorem 1.1 and shows that, under appropriate precautions, the case q=2 can be put in a common picture with the cases corresponding to q2. Taking the limit as s0+, we recover Beckner’s fractional logarithmic Sobolev inequality as stated in [4, 6]. In that case, q=2 is critical from the point of view of the fractional operator. The proof of Corollary 1.4 and further considerations on the s=0 limit will be given in Section 2.4.

Definition (1.6) of 𝒦s also applies to the range s(-n,0) and the reader is invited to check that

𝒦 s - 1 = 𝒦 - s for all s ( 0 , n )

is defined by the sequence of eigenvalues γk(np) where p=2nn+s is the Hölder conjugate of q(s) given by (1.4). It is then straightforward to check that the sharp Hardy–Littlewood–Sobolev inequality on 𝕊n (see (1.3)) can be written as

(1.12) F L p ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 p - 2 κ n , - s 2 - p 𝕊 n F - s F 𝑑 μ for all F L 2 ( 𝕊 n ) ,

where

p = 2 n n + s ( 1 , 2 ) , - s := 1 κ n , - s ( Id - 𝒦 - s ) , κ n , - s = Γ ( n + s 2 ) Γ ( n - s 2 ) .

Notice that κn,-s=1κn,s. A first consequence is that we can rewrite the result of Proposition 1.3 as

G L p ( 𝕊 n ) 2 - 𝕊 n G 𝒦 - s G 𝑑 μ F L q ( 𝕊 n ) 2 ( q - 2 ) ( 𝕊 n F 𝒦 s F 𝑑 μ - F L q ( 𝕊 n ) 2 )

for any FHs/2(𝕊n) and G=Fq-1, where n1, s(0,n), q is given by (1.4) and p=q. A second consequence of the above observations is the extension of Theorem 1.1 to the range (-n,0).

Theorem 1.5

Let n1, s(-n,0) and q[1,2nn-s). Inequality (1.7) holds with Ls:=κn,-s(Id-Ks) and sharp constant

𝖢 q , s = n - s 2 | s | Γ ( n - s 2 ) Γ ( n + s 2 ) .

The results of Theorems 1.1 and 1.5 are summarized in Figure 1.

Figure 1 
					The optimal constant 𝖢q,s${\mathsf{C}_{q,s}}$ in (1.7) is independent of q and determined for any given s by the critical case q=q⋆⁢(s)${q=q_{\star}(s)}$ which corresponds to the Hardy–Littlewood–Sobolev inequality (1.1) if s∈(-n,0)${s\in(-n,0)}$ and to the Sobolev inequality (1.5) if s∈(0,n)${s\in(0,n)}$. The case s=0${s=0}$ is covered by Corollary 2.4, while q=2${q=2}$ corresponds to the fractional logarithmic Sobolev inequality (2.3) if s=0${s=0}$ and the subcritical fractional logarithmic Sobolev inequality by Corollary 1.4 if s∈(0,n]${s\in(0,n]}$.
Figure 1

The optimal constant 𝖢q,s in (1.7) is independent of q and determined for any given s by the critical case q=q(s) which corresponds to the Hardy–Littlewood–Sobolev inequality (1.1) if s(-n,0) and to the Sobolev inequality (1.5) if s(0,n). The case s=0 is covered by Corollary 2.4, while q=2 corresponds to the fractional logarithmic Sobolev inequality (2.3) if s=0 and the subcritical fractional logarithmic Sobolev inequality by Corollary 1.4 if s(0,n].

To conclude with the outline of this paper, Section 3 is devoted to the stereographic projection and consequences for functional inequalities on the Euclidean space. By stereographic projection, (1.5) becomes

f L q ( n ) 2 𝖲 n , s f H ˙ s / 2 ( n ) 2 for all f H ˙ s / 2 ( n ) ,

where

f H ˙ s / 2 ( n ) 2 := n f ( - Δ ) s 2 f 𝑑 x

and the optimal constant is such that

𝖲 n , s = κ n , s | 𝕊 n | 2 q - 1 .

The fact that (1.5) is equivalent to the fractional Sobolev inequality on the Euclidean space is specific to the critical exponent q=q(s). In the subcritical range weights appear. Let us introduce the weighted norm

f L q , β ( n ) q := n | f | q ( 1 + | x | 2 ) - β 2 𝑑 x .

The next result is inspired by a non-fractional computation done in [26] and relies on the stereographic projection.

Theorem 1.6

Let n1, s(0,n), q(2,q) with q given by (1.4), and β=2n(1-qq). Then we have the weighted inequality

(1.13) f L q , β ( n ) 2 𝖺 f H ˙ s / 2 ( n ) 2 + 𝖻 f L 2 , 2 s ( n ) 2 for all f C 0 ( n ) ,

where

𝖺 = q - 2 q - 2 κ n , s 2 n ( 2 q - 2 q ) | 𝕊 n | 2 q - 1 𝑎𝑛𝑑 𝖻 = q - q q - 2 2 n ( 1 - 2 q ) | 𝕊 n | 2 q - 1 .

Moreover, if q<q, equality holds in (1.13) if and only if f is proportional to fs,(x):=(1+|x|2)-(n-s)/2.

This result is one of the few examples of optimal functional inequalities involving fractional operators on n. It touches the area of fractional Hardy–Sobolev inequalities and weighted fractional Sobolev inequalities, for which we refer to [34, 14] and [16], respectively, and the references therein. The wider family of Caffarelli–Kohn–Nirenberg type inequalities raises additional difficulties, for instance related with symmetry and symmetry breaking issues, which are so far essentially untouched in the framework of fractional operators, up to few exceptions like [14].

Inequality (1.13) holds not only for the space C0(n) of all smooth functions with compact support but also for the much larger space of functions obtained by completion of C0(n) with respect to the norm defined by

f 2 := f H ˙ s / 2 ( n ) 2 + f L 2 , 2 s ( n ) 2 .

2 Subcritical Interpolation Inequalities

In this section, our purpose is to prove Theorem 1.1.

2.1 A Poincaré Inequality

We start by recalling some basic facts:

  1. If q and q are Hölder conjugates, then nq=n-x with x=nq.

  2. γ0(x)=1 for any x>0.

  3. γk(n2)=1 and δk(n2)=0 for any k.

  4. γ1(x)=n-xx, γ1(nq)=q-1 and δ1(nq)=q-2κn,s. As a consequence, we know that the first positive eigenvalues of 𝒦s and s are

    λ 1 ( 𝒦 s ) = γ 1 ( n q ) = q - 1 and λ 1 ( s ) = δ 1 ( n q ) = q - 2 κ n , s = 2 s ( n - s ) κ n , s .

A straightforward consequence is the following sharp Poincaré inequality.

Lemma 2.1

For any FHs/2(Sn) we have

F - F ( 0 ) L 2 ( 𝕊 n ) 2 𝖢 1 , s 𝕊 n F s F 𝑑 μ ,

where

F ( 0 ) = 𝕊 n F 𝑑 μ ,

and

𝖢 1 , s = κ n , s q - 2

is the optimal constant. Any function F=F(0)+F(1), with F(1) such that LsF(1)=λ1(Ls)F(1), realizes the equality case.

Proof.

The proof is elementary. With the usual notations, we may write

𝕊 n F s F 𝑑 μ = 𝕊 n ( F - F ( 0 ) ) s ( F - F ( 0 ) ) 𝑑 μ = k = 1 δ k ( n q ) 𝕊 n | F ( k ) | 2 𝑑 μ
δ 1 ( n q ) F - F ( 0 ) L 2 ( 𝕊 n ) 2 = λ 1 ( s ) F - F ( 0 ) L 2 ( 𝕊 n ) 2 ,

because δk(nq) is increasing with respect to k. ∎

The sharp Poincaré constant 𝖢1,s is a lower bound for 𝖢q,s, for any q(1,q] if s<n, or any q>1 if s=n. Indeed, if q2, by testing inequality (1.7) with F=1+εG1, where G1 is an eigenfunction of s associated with the eigenvalue λ1(s), it is easy to see that

ε 2 G 1 L 2 ( 𝕊 n ) 2 F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 𝖢 q , s 𝕊 n F s F 𝑑 μ = 𝖢 q , s ε 2 𝕊 n G 1 s G 1 𝑑 μ

as ε0, which means that

G 1 L 2 ( 𝕊 n ) 2 = λ 1 ( s ) 𝖢 q , s G 1 L 2 ( 𝕊 n ) 2 ,

by keeping only the leading order term in ε. Altogether, this proves that

(2.1) 𝖢 q , s 1 λ 1 ( s ) = κ n , s q - 2 .

A similar computation, with (1.7) replaced by (1.11) and F=1+εG1, shows that

𝕊 n | F | 2 log ( | F | F 2 ) 𝑑 μ 𝖢 2 , s ε 2 𝕊 n G 1 s G 1 𝑑 μ

as ε0, so that (2.1) also holds if q=2. Hence, under the assumptions of Theorem 1.1, inequality (2.1) holds for any q1. In order to establish Theorem 1.1 and Corollary 1.4, we now have to prove that (2.1) is actually an equality.

2.2 Some Spectral Estimates

Let us start with some observations on the function γk in (1.6). Expanding its expression, we get that

γ k ( x ) = ( n + k - 1 - x ) ( n + k - 2 - x ) ( n - x ) ( k - 1 + x ) ( k - 2 + x ) x

for any k1. Taking the logarithmic derivative, we find that

(2.2) α k ( x ) := - γ k ( x ) γ k ( x ) = j = 0 k - 1 β j ( x ) with β j ( x ) := 1 n + j - x + 1 j + x ,

and observe that αk is positive. As a consequence, γk<0 on [0,n] and, from the expression of γk, we read that γk(n)=0. Since γk(n2)=1, we know that γk(nq)>1 if and only if q>2. Using the fact that

γ k ′′ ( x ) γ k ( x ) = ( α k ( x ) ) 2 - α k ( x ) = ( γ k ( x ) γ k ( x ) ) 2 + j = 0 k - 1 ( 2 j + n ) ( n - 2 x ) ( n + j - x ) 2 ( j + x ) 2 ,

we have γk′′(x)0, which establishes the convexity of γk on [0,n2]. Moreover, we know that

γ k ( n 2 ) = - α k ( n 2 ) = - j = 0 k - 1 4 n + 2 j .

See Figure 2. Taking these observations into account, we can state the following result.

Lemma 2.2

Assume that n1. With the above notations, the function

q γ k ( n q ) - 1 q - 2

is strictly monotone increasing on (1,) for any k2.

Figure 2 
						The functions x↦γk⁢(x)${x\mapsto\gamma_{k}(x)}$ and q↦γk⁢(nq)${q\mapsto\gamma_{k}(\frac{n}{q})}$ are both convex, and such that γk⁢(n2)=1${\gamma_{k}(\frac{n}{2})=1}$.
Figure 2

The functions xγk(x) and qγk(nq) are both convex, and such that γk(n2)=1.

Proof.

Let us prove that qγk(nq) is strictly convex with respect to q for any k2. Written in terms of x=nq, it is sufficient to prove that

x γ k ′′ + 2 γ k > 0 for all x ( 0 , n ) ,

which can also be rewritten as

α k 2 - α k - 2 x α k > 0 .

Let us prove this inequality. Using the estimates

α k 2 = ( j = 0 k - 1 β j ) 2 2 β 0 j = 1 k - 1 β j + j = 0 k - 1 β j 2 , β 0 2 - β 0 - 2 x β 0 = 0 ,

and

2 β 0 β j + β j 2 - β j - 2 x β j = 2 ( n + j ) ( n + 2 j ) ( n - x ) ( n + j - x ) ( j + x ) 2

for any j1, we actually find that

α k 2 - α k - 2 x α k j = 1 k - 1 2 ( n + j ) ( n + 2 j ) ( n - x ) ( j + n - x ) ( j + x ) 2 for all k 2 ,

which concludes the proof. Note that as a byproduct, we also proved the strict convexity of γk for the whole range x(0,n). See Figure 2 for a summarization of properties of the spectral functions. ∎

Proof of Theorem 1.1.

We deduce from (1.5) that

F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 k = 1 γ k ( n q ) - 1 q - 2 𝕊 n | F ( k ) | 2 𝑑 μ

because γ0(x)=1. It follows from Lemma 2.2 that

F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 k = 1 γ k ( n q ) - 1 q - 2 𝕊 n | F ( k ) | 2 𝑑 μ
= κ n , s q - 2 k = 1 δ k ( n q ) 𝕊 n | F ( k ) | 2 𝑑 μ = κ n , s q - 2 𝕊 n F s F 𝑑 μ .

This proves that 𝖢q,sκn,sq-2. The reverse inequality has already been shown in (2.1).∎

Proof of Theorem 1.5.

With s(-n,0), it turns out that q defined by (1.4) is in the range (1,2) and plays the role of p in (1.12). According to Lemma 2.2, the inequality holds with the same constant for any q(1,q), and this constant is optimal because of (2.1).∎

2.3 An Improved Inequality with a Remainder Term

What we have shown in Section 2.2 is actually that the fractional Sobolev inequality (1.5) is equivalent to the following improved subcritical inequality.

Corollary 2.3

Assume that n1, q[1,2)(2,q) if s(0,n), and q[1,2)(2,) if s=n. For any FHs/2(Sn) we have

F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 + 𝕊 n F q , s F 𝑑 μ κ n , s q - 2 𝕊 n F s F 𝑑 μ ,

where Rq,s is a positive semi-definite operator whose kernel is generated by the spherical harmonics corresponding to k=0 and k=1.

Proof.

We observe that

𝕊 n F q , s F 𝑑 μ := k = 2 ϵ k 𝕊 n | F ( k ) | 2 𝑑 μ ,

where

ϵ k := γ k ( n q ) - 1 q - 2 - γ k ( n q ) - 1 q - 2

is positive for any k2 according to Lemma 2.2.∎

Equality in (1.7) is realized only when F optimizes the critical fractional Sobolev inequality and, if q<q, when F(k)=0 for any k2, which is impossible unless F is an optimal function for the Poincaré inequality of Lemma 2.1. This observation will be further exploited in Section 4.

2.4 Fractional Logarithmic Sobolev Inequalities

Proof of Corollary 1.4.

According to Theorem 1.1, we know by (1.7) that

F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 n - s 2 s κ n , s 𝕊 n F s F 𝑑 μ

for any function FHs/2(𝕊n) and any q[1,2)(2,q) with q=q(s) given by (1.4) (and the convention that q= if s=n). Taking the limit as q2 for a given s(0,n), we obtain that (1.11) holds with 𝖢2,sn-s2sκn,s. The reverse inequality has already been shown in (2.1) written with q=2.∎

Let us comment on the results of Corollary 1.4, in preparation for Section 4. Instead of fixing s and letting q2 as in the proof of Corollary 1.4, we can consider the case q=q(s) and let s0, or equivalently rewrite (1.5) as

F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 k = 0 γ k ( n q ) - 1 q - 2 𝕊 n | F ( k ) | 2 𝑑 μ

and take the limit as q2. By an endpoint differentiation argument, we recover the conformally invariant fractional logarithmic Sobolev inequality

(2.3) 𝕊 n F 2 log ( | F | F L 2 ( 𝕊 n ) ) 𝑑 μ n 2 𝕊 n F 𝒦 0 F 𝑑 μ

as in [4, 6], where the differential operator 𝒦0 is the endpoint derivative of 𝒦s at s=0. The equality 𝒦0=0 holds because κn,0=1 and 𝒦0=Id. More specifically, the right-hand side of (2.3) can be written using the identities

𝕊 n F 𝒦 0 F 𝑑 μ = 𝕊 n F 0 F 𝑑 μ = 1 2 k = 0 α k ( n 2 ) 𝕊 n | F ( k ) | 2 𝑑 μ

with

α k ( n 2 ) = - γ k ( n 2 ) = j = 0 k - 1 4 n + 2 j .

Inequality (2.3) is sharp, and equality holds if and only if F is obtained by applying any conformal transformation on 𝕊n to constant functions. Finally, let us notice that (2.3) can be recovered as an endpoint of (1.11) by letting s0. The critical case is then achieved as a limit of the subcritical inequalities (1.11). The optimal constant can be identified, but the set of optimal functions in the limit is larger than in the subcritical regime, because of the conformal invariance.

Even more interesting is the fact that the fractional logarithmic Sobolev inequality is critical for s=0 and q=2 but subcritical inequalities corresponding to q[1,2) still make sense.

Corollary 2.4

Assume that n1 and q[1,2). For any FL2(Sn) such that SnFK0F𝑑μ is finite, we have

F L q ( 𝕊 n ) 2 - F L 2 ( 𝕊 n ) 2 q - 2 n 2 𝕊 n F 𝒦 0 F 𝑑 μ .

As for Corollary 1.4, the proof relies on Lemma 2.2. Details are left to the reader.

3 Stereographic Projection and Weighted Fractional Interpolation Inequalities on the Euclidean Space

This section is devoted to the proof of Theorem 1.6. Various results concerning the extension of the Caffarelli–Kohn–Nirenberg inequalities introduced in [9] (see also [20, Theorem 1] in our context) are scattered throughout the literature, and one can consult for instance [18, Theorem 1.8] for a quite general result in this direction. However, very little is known so far on optimal constants or even estimates of such constants, except for some limit cases like fractional Sobolev or fractional Hardy–Sobolev inequalities (see, e.g., [44]). What we prove here is that the interpolation inequalities on the sphere provide inequalities on the Euclidean space with weights based on (1+|x|2) with optimal constants.

Proof of Theorem 1.6.

Let us consider the stereographic projection 𝒮, whose inverse is defined by

𝒮 - 1 : n 𝕊 n , x ζ = ( 2 x 1 + | x | 2 , 1 - | x | 2 1 + | x | 2 )

with Jacobian determinant |J|=2n(1+|x|2)-n. Given s(0,n) and q(2,q), and using the conformal Laplacian, we can rewrite inequality (1.7) as

F L q ( 𝕊 n ) 2 - q - q q - 2 F L 2 ( 𝕊 n ) 2 q - 2 q - 2 κ n , s 𝕊 n F 𝒜 s F 𝑑 μ ,

where 𝒜s and the fractional Laplacian on n are related by

| J | 1 - 1 q ( 𝒜 s F ) 𝒮 - 1 = ( - Δ ) s 2 ( | J | 1 q F 𝒮 - 1 ) .

Then the interpolation inequality (1.7) on the sphere is equivalent to the fractional interpolation inequality on the Euclidean space

| 𝕊 n | 1 - 2 q ( n | f | q | J | 1 - q q 𝑑 x ) 2 q - q - q q - 2 n f 2 | J | 1 - 2 q 𝑑 x q - 2 q - 2 κ n , s n f ( - Δ ) s 2 f 𝑑 x

by using the change of variables Ff=|J|1/qF𝒮-1. The equality case is now achieved only by f=|J|1/q for any q(2,q), up to a multiplication by a constant, and the inequality is equivalent to (1.13).∎

4 Concluding Remarks

A striking feature of inequality (1.7) is that the optimal constant 𝖢q,s is determined by a linear eigenvalue problem, although the problem is definitely nonlinear. This deserves some comments. Let q[1,2)(2,q) if s<n and q[1,2)(2,) if s=n. With 𝒬 defined by (1.9) on s/2, the subset of the functions in Hs/2(𝕊n) which are not almost everywhere constant, we investigate the relation

𝖢 q , s inf F s / 2 𝒬 [ F ] = 1 .

Notice that both numerator and denominator of 𝒬[F] converge to 0 if F approaches a constant, so that 𝒬 becomes undetermined in the limit. As we shall see next, this happens for a minimizing sequence and explains why a linearized problem appears in the limit.

By compactness of the Sobolev embedding Hs/2(𝕊n)Lq(𝕊n) (see [1, 18] for fundamental properties of fractional Sobolev spaces, [22, Sections 6 and 7] and [41] for application to variational problems), any minimizing sequence (Fn)n for 𝒬 is relatively compact if we assume that FnLq(𝕊n)=1 for any n. This normalization can be imposed without loss of generality because of the homogeneity of 𝒬. Hence (Fn)n converges to a limit FHs/2(𝕊n). Assume that F is not a constant. Then the denominator in 𝒬[F] is positive and by semicontinuity we know that

𝕊 n F s F 𝑑 μ lim n + 𝕊 n F n s F n 𝑑 μ .

On the other hand, by compactness, up to the extraction of a subsequence, we have that

F L 2 ( 𝕊 n ) 2 = lim n + F n L 2 ( 𝕊 n ) 2 and F L q ( 𝕊 n ) 2 = lim n + F n L q ( 𝕊 n ) 2 = 1 .

Hence F is optimal and solves the Euler–Lagrange equations

( q - 2 ) 𝖢 q , s s F + F = F q - 1 .

Using Corollary 2.3, we also get that F lies in the kernel of q,s, that is, the space generated by the spherical harmonics corresponding to k=0 and k=1. From the Euler–Lagrange equations, we read that F has to be a constant. Because of the normalization FLq(𝕊n)=1, we obtain that F=1 a.e., a contradiction.

Hence (Fn)n converges to 1 in Hs/2(𝕊n). With εn=1-FnHs/2(𝕊n) and vn:=Fn-1εn, we can write that

F n = 1 + ε n v n with v n H s / 2 ( 𝕊 n ) = 1    for all n ,

and

lim n + ε n = 0 .

On the other hand, (Fn)n being a minimizing sequence, it turns out that

𝖢 q , s - 1 = lim n + 𝒬 [ F n ] = lim n + ε n 2 ( q - 2 ) 𝕊 n v n s v n 𝑑 μ 1 + ε n v n L q ( 𝕊 n ) 2 - 1 + ε n v n L 2 ( 𝕊 n ) 2 .

If q>2, an elementary computation shows that

(4.1) 1 + ε n v n L q ( 𝕊 n ) 2 - 1 + ε n v n L 2 ( 𝕊 n ) 2 = ( q - 2 ) ε n 2 v n - v ¯ n L 2 ( 𝕊 n ) 2 ( 1 + o ( 1 ) )

as n+, where v¯n:=𝕊nvn𝑑μ, so that

𝖢 q , s - 1 = lim n + 𝒬 [ F n ] = lim n + 𝕊 n v n s v n 𝑑 μ v n - v ¯ n L 2 ( 𝕊 n ) 2 .

Details on the Taylor expansion used in (4.1) can be found in Appendix B. When q[1,2), we can estimate the denominator by restricting the integrals to {x𝕊n:εn|vn|<12} and Taylor expand t(1+t)q on (12,32).

Notice that by Fn being a function in s/2, we know that vn-v¯nL2(𝕊n)>0 for any n, so that the above limit makes sense. With the notations of Section 2.1, we know that

𝖢 q , s - 1 inf v s / 2 𝕊 n v s v 𝑑 μ v - v ¯ L 2 ( 𝕊 n ) 2 λ 1 ( s ) = 2 s κ n , s n - s

according to the Poincaré inequality of Lemma 2.1, which proves that we actually have equality in (2.1) and determines 𝖢q,s.

Additionally, we may notice that (vn)n has to be a minimizing sequence for the Poincaré inequality, which means that up to a normalization and after the extraction of a subsequence, vn-v¯n converges to a spherical harmonic function associated with the component corresponding to k=1. This explains why we obtain that 𝖢q,sλ1(s)=1.

The above considerations have been limited to the subcritical range q<q if s<n and q<+ if s=n. However, the critical case of the Sobolev inequality can be obtained by passing to the limit as qq (and even the Onofri-type inequalities when s=n) so that the optimal constants are also given by an eigenvalue in the critical case. However, due to the conformal invariance, the constant function F1 is not the only optimal function. At this point it should be noted that the above considerations heavily rely on Corollary 2.3 and, as a consequence, cannot be used to give a variational proof of Theorem 1.1.

Although the subcritical interpolation inequalities of this paper appear weaker than inequalities corresponding to a critical exponent, we are able to identify the equality cases and the optimal constants. We are also able to keep track of a remainder term which characterizes the functions realizing the optimality of the constant or, to be precise, the limit of any minimizing sequence and its first order correction. This first order correction, or equivalently the asymptotic value of the quotient 𝒬, determines the optimal constant and explains the role played by the eigenvalues in a problem which is definitely nonlinear.


Communicated by Laurent Veron


Award Identifier / Grant number: ANR-10-LABX-0098

Award Identifier / Grant number: ANR-12-BS01-0019

Award Identifier / Grant number: ANR-13-BS01-0004

Award Identifier / Grant number: 291214

Funding statement: This work is supported by a public grant overseen by the French National Research Agency (ANR) as part of the “Investissements d’Avenir” program (the second author, reference: ANR-10-LABX-0098, LabEx SMP) and by the projects STAB (both authors) and Kibord (the first author) of the French National Research Agency (ANR). The second author thanks the ERC Advanced Grant BLOWDISOL (Blow-up, dispersion and solitons; PI: Frank Merle) # 291214 for support.

A The Spectrum of the Fractional Laplacian

The standard approach for computing γk in (1.6) relies on the Funk–Hecke formula as it is detailed in [33, Section 4]. In this appendix, for completeness, we provide a simple direct proof of the expression of γk. For this purpose, we compute the eigenvalues λk=λk((-Δ)s/2) of the fractional Laplacian on n, that is,

( - Δ ) s 2 f k = λ k ( 1 + | x | 2 ) s f k in n

for any k. We shall then deduce the eigenvalues of s. This determines the optimal constant in (1.5) and (1.7) without using Lieb’s duality and without relying on the symmetry of the optimal case in (1.1) as in [38].

Proposition A.1

Given s(0,n), the spectrum of the fractional Laplacian is

λ k ( ( - Δ ) s 2 ) = 2 s Γ ( k + n q ) Γ ( k + n q ) = 2 s λ k ( 𝒜 s ) = 2 s Γ ( n q ) Γ ( n q ) λ k ( 𝒦 s ) .

Proof.

Using the stereographic projection and a decomposition in spherical harmonics, we can reduce the problem of computing the spectrum to the computation of the spectrum associated with the eigenfunctions

f k μ ( x ) = C k ( α ) ( z ) ( 1 + | x | 2 ) - μ with z = 1 - | x | 2 1 + | x | 2 ,

where μ=λ2=n-s2, α=n-12 and Ck(α) denotes the Gegenbauer polynomials. Let

f ^ ( ξ ) = ( f ) ( ξ ) := n f ( x ) e - 2 π i ξ x 𝑑 x

be the Fourier transform of a function f. Since the functions are radial, by the Hankel transform n2-1 we get that

f k μ ^ ( ξ ) = 2 π | ξ | n 2 - 1 0 f k μ ( r ) J n 2 - 1 ( 2 π r | ξ | ) r n 2 𝑑 r

(cf. [35, Appendix B.5, p. 578]), where Jν is the Bessel function of the first kind.

The Fourier transform of f0μ=(1+|x|2)-μ has been calculated, e.g., by E. Lieb in [38, (3.9)–(3.14)] in terms of the modified Bessel functions of the second kind Kν as

f 0 μ ^ ( ξ ) = π n 2 2 1 + 2 n - μ Γ ( μ ) ( 2 π | ξ | ) μ - n 2 K μ - n 2 ( 2 π | ξ | ) .

This is a special case of the modified Weber–Schafheitlin integral formula in [43, Section 13.45]. Using the expansion of Gegenbauer polynomials, we get

f k μ ^ ( ξ ) = 2 π Γ ( n - 1 2 ) | ξ | n 2 - 1 j = 0 [ k 2 ] l = 0 k - 2 j [ ( - 1 ) j + k - l 2 k + l - 2 j Γ ( n - 1 2 + k - j ) j ! k ! ( k - 2 j - l ) ! 0 ( 1 + r 2 ) - ( μ + l ) J n 2 - 1 ( 2 π r | ξ | ) r n 2 𝑑 r ]
= 1 Γ ( n - 1 2 ) j = 0 [ k 2 ] l = 0 k - 2 j ( - 1 ) j + k - l 2 k + l - 2 j Γ ( n - 1 2 + k - j ) j ! l ! ( k - 2 j - l ) ! f 0 μ + l ^ ( ξ )
= 2 1 + 2 n - μ π n 2 ( 2 π | ξ | ) μ - n 2 Γ ( n - 1 2 ) Γ ( μ + k ) I n , k μ ( | ξ | ) ,

where

I n , k μ ( | ξ | ) := l = 0 k c n , k , l Γ ( μ + k ) Γ ( μ + l ) ( 2 π | ξ | ) l K μ - n 2 + l ( 2 π | ξ | )

and

c n , k , l := 1 l ! j = 0 [ k - l 2 ] ( - 1 ) j + k - l 2 k - 2 j Γ ( n - 1 2 + k - j ) j ! ( k - 2 j - l ) ! .

From the recurrence relation

x ( K ν - 1 - K ν + 1 ) = - 2 ν K ν ,

we deduce the identity

l = 0 k c n , k , l x l ( Γ ( ν + n 2 + k ) Γ ( ν + n 2 + l ) K ν + l ( x ) - Γ ( - ν + n 2 + k ) Γ ( - ν + n 2 + l ) K ν - l ( x ) ) = 0 for all k 0

and observe that

I n , k μ 1 = I n , k μ 2 for all k

if μ1=λ2 and μ2=λ2+s, so that μ1+μ2=n and μ1-μ2=-s. It remains to observe that

( 2 π | ξ | ) s f k λ / 2 ^ = λ k ( f k λ / 2 ( 1 + | x | 2 ) - s ) with λ k = 2 s Γ ( k + n q ) Γ ( k + n q ) .

B A Taylor Formula with Integral Remainder Term

Let us define the function r: such that

| 1 + t | q = 1 + q t + 1 2 q ( q - 1 ) t 2 + r ( t ) for all t .

Lemma B.1

Let q(2,). With the above notations, there exists a constant C>0 such that

| r ( t ) | { C | t | 3 if | t | 1 , C | t | q if | t | 1 .

This result is elementary but crucial for the expansion of FLq(𝕊n)2-FL2(𝕊n)2 around F=1. This is why we give a proof with some details, although we claim absolutely no originality for that. Similar computations have been repeatedly used in a related context, e.g., in [15, 17, 40].

Proof.

Using the Taylor formula with integral remainder term

f ( t ) = f ( 0 ) + f ( 0 ) t + 1 2 f ′′ ( 0 ) t 2 + 1 2 0 t ( t - s ) 2 f ′′′ ( s ) 𝑑 s

applied to f(t)=(1+t)q with q>2, we obtain that

| 1 + t | q = 1 + q t + 1 2 q ( q - 1 ) t 2 + r ( t ) ,

where the remainder term is given by

r ( t ) = 1 2 q ( q - 1 ) ( q - 2 ) t q 0 1 ( 1 - σ ) 2 | 1 t + σ | q - 4 ( 1 t + σ ) 𝑑 σ .

Hence the remainder term can be bounded as follows:

  1. If t1, using σ<1t+σ<1+σ, we get that

    0 < r ( t ) < c q t q

    with

    c q = 1 2 q ( q - 1 ) ( q - 2 ) 0 1 ( 1 - σ ) 2 max { σ q - 3 , ( 1 + σ ) q - 3 } 𝑑 σ .

  2. If 0<t<1, using 1t<1t+σ<2t, we get that

    0 < r ( t ) < 1 6 q ( q - 1 ) ( q - 2 ) max { 1 , 2 q - 3 } t 3 .

  3. If -1<t<0, using 1t<1t+σ<1t+1<0, we get that

    - 1 6 q ( q - 1 ) ( q - 2 ) | t | 3 < r ( t ) < 0 .

  4. If t-1, using σ-1<1t+σ<σ, we get that

    - 1 2 ( q - 1 ) ( q - 2 ) t q < r ( t ) < t q .

C Notations and Ranges

For the convenience of the reader, this appendix collects various notations which are used throughout this paper and summarizes the ranges covered by the parameters.

The identity

λ = 2 n p where 1 p + 1 p = 1

means that

p = 2 n 2 n - λ .

With

λ = n - s ,

we have

p = 2 n n + s and p = q = 2 n n - s .

The limiting values of the parameters are summarized in Table 1.

Table 1

Correspondence of the limiting values of the parameters.

s 0 2 n
λ n n - 2 0
p 2 2 n n + 2 1
p = q 2 2 n n - 2 +

The coefficients γk and δk defined by

γ k ( x ) = Γ ( x ) Γ ( n - x + k ) Γ ( n - x ) Γ ( x + k ) and δ k ( x ) = 1 κ n , s ( γ k ( x ) - 1 ) = Γ ( n - x + k ) Γ ( x + k ) - Γ ( n - x ) Γ ( x )

are such that

δ k ( n q ) = 1 κ n , s ( γ k ( n q ) - 1 ) where κ n , s = Γ ( n q ) Γ ( n - n q ) = Γ ( n - s 2 ) Γ ( n + s 2 ) .

We recall that

γ 0 ( n q ) - 1 = 0 , γ 1 ( n q ) - 1 = q - 2 , δ k ( n q ) = k ( k + n - 1 ) , 1 κ n , 2 = 1 4 n ( n - 2 ) .

According to (2.2), we have that

α k ( x ) = - γ k ( x ) γ k ( x ) = j = 0 k - 1 β j ( x ) with β j ( x ) = 1 n + j - x + 1 j + x

for any k1. With these notations, the eigenvalues of 𝒦s, s and 𝒦0=0 are respectively given by

γ k ( n q ( s ) ) = γ k ( n - s 2 ) , κ n , s - 1 ( γ k ( n - s 2 ) - 1 ) , 1 2 α k ( n 2 )

with

α k ( n 2 ) = - γ k ( n 2 ) = 4 j = 0 k - 1 ( n + 2 j ) - 1 .

Finally, we recall that 𝒦s, the fractional Laplacian s and the conformal fractional Laplacian 𝒜s satisfy the relations

κ n , s 𝒜 s = 𝒦 s = κ n , s s + Id .

Acknowledgements

The authors thank Maria J. Esteban for fruitful discussions and suggestions. They are also grateful to Van Hoang Nguyen for pointing out an error in a former version.

References

[1] Adams R. A., Sobolev Spaces, Pure Appl. Math. 65, Academic Press, New York, 1975. Search in Google Scholar

[2] Arnold A., Markowich P., Toscani G. and Unterreiter A., On convex Sobolev inequalities and the rate of convergence to equilibrium for Fokker–Planck type equations, Comm. Partial Differential Equations 26 (2001), no. 1–2, 43–100. 10.1081/PDE-100002246Search in Google Scholar

[3] Bakry D. and Émery M., Diffusions hypercontractives, Séminaire de Probabilités XIX (Strasbourg 1983/84), Lecture Notes in Math. 1123, Springer, Berlin (1985), 177–206. 10.1007/BFb0075847Search in Google Scholar

[4] Beckner W., Sobolev inequalities, the Poisson semigroup, and analysis on the sphere 𝕊n, Proc. Natl. Acad. Sci. USA 89 (1992), no. 11, 4816–4819. 10.1073/pnas.89.11.4816Search in Google Scholar

[5] Beckner W., Sharp Sobolev inequalities on the sphere and the Moser–Trudinger inequality, Ann. of Math. (2) 138 (1993), no. 1, 213–242. 10.2307/2946638Search in Google Scholar

[6] Beckner W., Logarithmic Sobolev inequalities and the existence of singular integrals, Forum Math. 9 (1997), no. 3, 303–323. 10.1515/form.1997.9.303Search in Google Scholar

[7] Berger M., Gauduchon P. and Mazet E., Le Spectre d’une Variété Riemannienne, Lecture Notes in Math. 194, Springer, Berlin, 1971. 10.1007/BFb0064643Search in Google Scholar

[8] Bidaut-Véron M.-F. and Véron L., Nonlinear elliptic equations on compact Riemannian manifolds and asymptotics of Emden equations, Invent. Math. 106 (1991), no. 3, 489–539. 10.1007/BF01243922Search in Google Scholar

[9] Caffarelli L., Kohn R. and Nirenberg L., First order interpolation inequalities with weights, Compos. Math. 53 (1984), no. 3, 259–275. Search in Google Scholar

[10] Carlen E. A. and Loss M., Extremals of functionals with competing symmetries, J. Funct. Anal. 88 (1990), no. 2, 437–456. 10.1016/0022-1236(90)90114-ZSearch in Google Scholar

[11] Carlen E. and Loss M., Competing symmetries, the logarithmic HLS inequality and Onofri’s inequality on 𝕊n, Geom. Funct. Anal. 2 (1992), no. 1, 90–104. 10.1007/BF01895706Search in Google Scholar

[12] Carrillo J. A., Huang Y., Santos M. C. and Vázquez J. L., Exponential convergence towards stationary states for the 1D porous medium equation with fractional pressure, J. Differential Equations 258 (2015), no. 3, 736–763. 10.1016/j.jde.2014.10.003Search in Google Scholar

[13] Chafaï D., Entropies, convexity, and functional inequalities: On Φ-entropies and Φ-Sobolev inequalities, J. Math. Kyoto Univ. 44 (2004), no. 2, 325–363. 10.1215/kjm/1250283556Search in Google Scholar

[14] Chen L., Liu Z. and Lu G., Symmetry and regularity of solutions to the weighted Hardy–Sobolev type system, Adv. Nonlinear Stud. 16 (2016), no. 1, 1–13. 10.1515/ans-2015-5005Search in Google Scholar

[15] Chen S., Frank R. L. and Weth T., Remainder terms in the fractional Sobolev inequality, Indiana Univ. Math. J. 62 (2013), no. 4, 1381–1397. 10.1512/iumj.2013.62.5065Search in Google Scholar

[16] Chen X. and Yang J., Weighted fractional Sobolev inequality in N, Adv. Nonlinear Stud. 16 (2016), no. 3, 623–641. 10.1515/ans-2015-5002Search in Google Scholar

[17] Christ M., A sharpened Hausdorff–Young inequality, preprint 2014, https://arxiv.org/abs/1406.1210. Search in Google Scholar

[18] D’Ancona P. and Lucà R., Stein–Weiss and Caffarelli–Kohn–Nirenberg inequalities with angular integrability, J. Math. Anal. Appl. 388 (2012), no. 2, 1061–1079. 10.1016/j.jmaa.2011.10.051Search in Google Scholar

[19] del Pino M. and Dolbeault J., The Euclidean Onofri inequality in higher dimensions, Int. Math. Res. Not. IMRN 2013 (2012), no. 15, 3600–3611. 10.1093/imrn/rns119Search in Google Scholar

[20] del Pino M., Dolbeault J., Filippas S. and Tertikas A., A logarithmic Hardy inequality, J. Funct. Anal. 259 (2010), no. 8, 2045–2072. 10.1016/j.jfa.2010.06.005Search in Google Scholar

[21] Demange J., Improved Gagliardo–Nirenberg–Sobolev inequalities on manifolds with positive curvature, J. Funct. Anal. 254 (2008), no. 3, 593–611. 10.1016/j.jfa.2007.01.017Search in Google Scholar

[22] Di Nezza E., Palatucci G. and Valdinoci E., Hitchhiker’s guide to the fractional Sobolev spaces, Bull. Sci. Math. 136 (2012), no. 5, 521–573. 10.1016/j.bulsci.2011.12.004Search in Google Scholar

[23] Dolbeault J., Sobolev and Hardy–Littlewood–Sobolev inequalities: Duality and fast diffusion, Math. Res. Lett. 18 (2011), no. 6, 1037–1050. 10.4310/MRL.2011.v18.n6.a1Search in Google Scholar

[24] Dolbeault J., Esteban M. J. and Loss M., Nonlinear flows and rigidity results on compact manifolds, J. Funct. Anal. 267 (2014), no. 5, 1338–1363. 10.1016/j.jfa.2014.05.021Search in Google Scholar

[25] Dolbeault J., Esteban M. J. and Loss M., Interpolation inequalities on the sphere: Linear vs. nonlinear flows, Ann. Fac. Sci. Toulouse Math. (6) (2016), https://hal.archives-ouvertes.fr/hal-01206975. 10.5802/afst.1536Search in Google Scholar

[26] Dolbeault J., Esteban M. J. and Tarantello G., Optimal Sobolev inequalities on N and 𝕊n: A direct approach using the stereographic projection, unpublished. Search in Google Scholar

[27] Dolbeault J. and Jankowiak G., Sobolev and Hardy–Littlewood–Sobolev inequalities, J. Differential Equations 257 (2014), no. 6, 1689–1720. 10.1016/j.jde.2014.04.021Search in Google Scholar

[28] Dolbeault J., Nazaret B. and Savaré G., A new class of transport distances between measures, Calc. Var. Partial Differential Equations 34 (2009), no. 2, 193–231. 10.1007/s00526-008-0182-5Search in Google Scholar

[29] Dolbeault J. and Toscani G., Improved interpolation inequalities, relative entropy and fast diffusion equations, Ann. Inst. Henri Poincaré Anal. Non Linéaire 30 (2013), no. 5, 917–934. 10.1016/j.anihpc.2012.12.004Search in Google Scholar

[30] Dolbeault J. and Toscani G., Stability results for logarithmic Sobolev and Gagliardo–Nirenberg inequalities, Int. Math. Res. Not. IMRN 2016 (2016), no. 2, 473–498.10.1093/imrn/rnv131Search in Google Scholar

[31] Frank R. L. and Lieb E. H., Inversion positivity and the sharp Hardy–Littlewood–Sobolev inequality, Calc. Var. Partial Differential Equations 39 (2010), no. 1–2, 85–99. 10.1007/s00526-009-0302-xSearch in Google Scholar

[32] Frank R. L. and Lieb E. H., Spherical reflection positivity and the Hardy–Littlewood–Sobolev inequality, Concentration, Functional Inequalities and Isoperimetry, Contemp. Math. 545, American Mathematical Society, Providence (2011), 89–102. 10.1090/conm/545/10767Search in Google Scholar

[33] Frank R. L. and Lieb E. H., A new, rearrangement-free proof of the sharp Hardy–Littlewood–Sobolev inequality, Spectral Theory, Function Spaces and Inequalities, Oper. Theory Adv. Appl. 219, Birkhäuser, Basel (2012), 55–67. 10.1007/978-3-0348-0263-5_4Search in Google Scholar

[34] Ghoussoub N. and Shakerian S., Borderline variational problems involving fractional Laplacians and critical singularities, Adv. Nonlinear Stud. 15 (2015), no. 3, 527–555. 10.1515/ans-2015-0302Search in Google Scholar

[35] Grafakos L., Classical Fourier Analysis, 3rd ed., Grad. Texts in Math. 249, Springer, New York, 2014. 10.1007/978-1-4939-1194-3Search in Google Scholar

[36] Jankowiak G. and Hoang Nguyen V., Fractional Sobolev and Hardy–Littlewood–Sobolev inequalities, preprint 2014, http://arxiv.org/abs/1404.1028. Search in Google Scholar

[37] Jin T. and Xiong J., A fractional Yamabe flow and some applications, J. Reine Angew. Math. 696 (2014), 187–223. 10.1515/crelle-2012-0110Search in Google Scholar

[38] Lieb E. H., Sharp constants in the Hardy–Littlewood–Sobolev and related inequalities, Ann. of Math. (2) 118 (1983), no. 2, 349–374. 10.1007/978-3-642-55925-9_43Search in Google Scholar

[39] Lieb E. H. and Loss M., Analysis, 2nd ed., Grad. Stud. Math. 14, American Mathematical Society, Providence, 2001. Search in Google Scholar

[40] Liu H. and Zhang A., Remainder terms for several inequalities on some groups of Heisenberg-type, Sci. China Math. 58 (2015), no. 12, 2565–2580. 10.1007/s11425-015-5070-9Search in Google Scholar

[41] Molica Bisci G., Radulescu V. D. and Servadei R., Variational Methods for Nonlocal Fractional Problems, Encyclopedia Math. Appl. 162, Cambridge University Press, Cambridge, 2016. 10.1017/CBO9781316282397Search in Google Scholar

[42] Müller C., Spherical Harmonics, Lecture Notes in Math. 17, Springer, Berlin, 1966. 10.1007/BFb0094775Search in Google Scholar

[43] Watson G. N., A Treatise on the Theory of Bessel Functions, Cambridge University Press, Cambridge, 1995. Search in Google Scholar

[44] Yang J., Fractional Sobolev–Hardy inequality in N, Nonlinear Anal. 119 (2015), 179–185. 10.1016/j.na.2014.09.009Search in Google Scholar

Received: 2016-08-30
Accepted: 2016-09-21
Published Online: 2016-10-18
Published in Print: 2016-11-01

© 2016 by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2016-0121/html
Scroll to top button