Compact embeddings for fractional super and sub harmonic functions with radial symmetry
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 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 embeddings2010 Mathematics Subject Classification:
Primary 46E35; Secondary 39B621. Introduction
The classical embedding in Sobolev spaces for follows from the interpolation inequality in homogeneous Sobolev spaces
(1.1) |
where and is defined by
The inequality (1.1) holds, see [5], Corollary 1.5 in [11], [6] or Theorem 2.44 in [1] provided that
-
•
-
•
,
-
•
, .
We notice that at the endpoint case , corresponding to , we have
(1.2) |
and hence the embedding for is just a consequence of (1.2). If we look at the endpoint cases and in (1.1) we obtain that the range of exponents without any symmetry and positivity assumption fulfills
(1.3) | |||||
We remark that the lower endpoint does not depend on dimension .
Moreover, looking at (1.2), it is easy to prove that the best constant in (1.2) is . Indeed from Hölder’s inequality in frequency applied to l.h.s. of (1.2) we get and calling it suffices to consider a sequence such that to prove that .
In the sequel we consider as fixed quantities and we aim to study the range of such that (1.1) holds in case we restrict to radially symmetric functions in such that is not only radially symmetric but also either positive or negative.
We introduce the notation for
(1.4) |
(1.5) |
(1.6) |
(1.7) |
By the relation we shall emphasize that corresponds to the set of superharmonic radially symmetric functions belonging to while corresponds to the set of subharmonic radially symmetric functions belonging to . In the sequel we will call when fractional superharmonic radially symmetric functions belonging to the functions belonging to and fractional subharmonic radially symmetric functions belonging to the functions belonging to .
The main questions we are interesting in are the following ones:
Question A: Can we find appropriate values of such that can be chosen below in (1.1) for fractional superharmonic (resp. subharmonic) functions belonging to ?
Question B: If the answer of question A is positive, then can we expect a compact embedding of type
(1.8) |
In the sequel we will consider the case but all the results are still valid if we consider . The first result of the paper gives a positive answer to Question A.
Theorem 1.1.
Let and , then
(1.9) | ||||
with
(1.10) | ||||
(1.11) |
with fixed by the scaling equation
and is given by
Remark 1.1.
Theorem 1.1 holds also for . The crucial condition is that does not change sign.
The constant in (1.9) is defined as best constant in case of functions belonging to .
The fact that in the above Theorem implies with and this allows us to obtain also a positive answer to Question B.
Theorem 1.2.
Let and , then the embedding
is compact for any
Remark 1.2.
As a second byproduct we have also the following result concerning the existence of maximizers for the interpolation inequality (1.9) in case .
Theorem 1.3.
Let and then
and the best constant is attained and
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 norms of Riesz potentials when . By defining we obtain
(1.12) |
where , . With respect to the new variables we get without any symmetry or positivity assumption
(1.13) | ||||||
If one considers functions fulfilling , inequality (1.12) is hence equivalent to the following inequality
(1.14) |
considering instead of in the Riesz potential. The strategy is hence to prove that the radial symmetry increases the range of for which (1.14) holds and therefore as byproduct the range of for which (1.12) holds when is positive and radially symmetric (resp. negative). In particular we will show that the lower endpoint is allowed to be below . 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 , we have been inspired by [13] where the case in (1.14) has been studied (see also [4] and [3]). For our purposes the fact that 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 norm to compute the decay for the low frequency part. The importance of a pointwise decay for the low frequency part involving weighted 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) |
a scaling argument forces the exponent to verify the relation . 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 . 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 then with . Our strategy is to prove the smallness of and for suitable choice of the ball For the first term we use Rellich-Kondrachov argument combined with commutator estimates, while for the exterior domain we use the crucial fact that is in for some
Looking at the case , by Rellich-Kondrachov we have , however we can not obtain the smallness in the complementary of the ball so the requirement seems to be optimal.
It is interesting to look at the lower endpoint exponent given in Theorem 1.1 in case we consider radially symmetric superharmonic (or subharmonic), namely when . In this case the condition , imposes to consider the case and . As an example we show on Figure 1 the graph of the function , that now is only a function of , in lowest dimensional case that is a branch of hyperbola with asymptote It is interesting how the regularity improves the lower endpoint .
As a final comment we notice that for if then then, taking and using the positivity of the Riesz kernel of we apply Theorem 1.2 and we get the following corollary.
Corollary 1.1.
Let be a sequence of radially symmetric superharmonic functions uniformly bounded in , . Then for any , up to subsequence in .

2. Interpolation inequalities for radial functions involving Riesz potentials.
Let , , we define
The aim of this section is to prove the following
Theorem 2.1.
Let with , , , then with
where with
Moreover, we have the scaling invariant inequality for
with if and if . Here is fixed by the scaling invariance
In order to show Theorem 2.1 we need to prove some preliminary results.
Proposition 2.1.
Let , , , , then there exists such that
(2.1) |
(2.2) |
The proposition for has been proved in [13], we follow the same argument for . In order to prove Proposition 2.1 two crucial lemmas are necessary. The case has been proved in [13] and we follow the same argument.
Lemma 2.1.
Let , , , then there exists such that for any
Proof.
Let us take , then
Thus we obtain for
and hence
By integration we conclude that
∎
Let us call where is a measurable function such that
(2.3) |
Lemma 2.2.
Let , , , then
and hence
(2.4) |
Proof.
Proof of Proposition 2.1.
∎
Lemma 2.3.
Let , and , then for any such that ,
Proof.
Let be a number to be fixed later. We have
where . Now choose . Notice that such that
The next Proposition concerning pointwise decay for radial functions in 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 be a radial function in with , , and
(2.5) |
Then for any satisfying
(2.6) |
we have
Remark 2.1.
It is easy to see that the above Proposition is equivalent to the following statement.
Let be a radial function in with , , and
(2.7) |
then for any such that ,
with
(2.8) |
Proof.
For any we can take a function such that is a radial nonnegative function with support in and for and then we make the decomposition of into low and high frequency part as follows
where . For the high frequency part we will use Fourier representation for radial functions in (identifying the function with its profile)
(2.9) |
where is the Bessel function of order Applying the results in [7] and [8], we find
(2.10) |
Indeed, using the uniform bound
we get
and this gives (2.10).
For low frequency term , since , we can take any so that there exists such that
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 then for any radially symmetric function we have
(2.11) |
Then we estimate as follows,
To this end we plan to apply Lemma 2.4 assuming . To check the assumption of the Lemma we use the inequalities
for Applying the Lemma 2.4 we deduce
Therefore, collecting our estimates and using the condition (2.5), we find
We use Lemma 2.3 and we get
This completes the proof.
∎
With all these preliminary results we are now ready to prove Theorem 2.1.
Proof.
Let with , then by Proposition 2.2
with
We aim to show that
, where is the lower endpoint for (1.12). Therefore it sufficies to show that
provided that and (indeed for all by interpolation).
We have, thanks to Proposition 2.2 and Lemma 2.3,
(2.12) |
provided that . This condition is equivalent, is defined in (2.8) and letting , to
An elementary computation shows that provided that .
Now consider an arbitrary and let us call where the parameters are chosen such that . By scaling we have
and hence we obtain the relations
By the previous estimates we have
which implies
where
It is easy to see that is fixed by the scaling invariance
∎
3. Proof of Theorem 1.1
Our goal is to represent in the form with and apply Theorem 2.1. Therefore, we choose (modulo constant)
Then the estimate of Theorem 2.1 gives
By the assumption
(3.1) |
for almost every then we deduce
and we obtain (1.9). The lower endpoint is hence nothing but of Theorem 2.1 substituting with and with . The condition is equivalent to the conditions , of Theorem 2.1. All these estimates remain valid if we consider , i.e if . Indeed if
4. Proof of Theorem 1.2
We prove that under the assumption of Theorem 1.2, the embedding
is compact. As a byproduct the embedding
(4.1) |
is compact for any . The embedding (4.1) follows noticing that if converges weakly to some in then converges weakly to the same in . Now if we prove that (taking a subsequence)
(4.2) |
as then by the following interpolation inequalities
if and
if , we get (4.1).
To prove (4.2) we recall that is a bounded sequence in and we can assume that converges weakly to some in . To simplify the notation we will use instead of in the proof of (4.2). We choose a bump function , such that on and in and for any we define Clearly the multiplication by is a continuous mapping . Now setting and we aim to show that
(4.3) |
for any
Indeed, by Plancharel’s identity we have
Clearly
and then we can choose such that .
Since , by weak convergence in we have almost everywhere. Notice that and hence is estimated by a uniform constant so that by Lebesgue’s dominated convergence theorem
for sufficiently large. This proves (4.3).
Our next step is to show that for a given one can find sufficiently large and sufficiently large so that
(4.4) |
for and any
We consider first the case The cases and are trivial, for this we assume We shall use the following statement (see Corollary 1.1 in [10]).
Proposition 4.1.
Let satisfy and . Let satisfy , and . Then the following bilinear estimate
holds for all .
By a density argument the statement holds for We choose and and therefore we aim to use (4.3) and prove that
(4.5) | ||||
Indeed from the Proposition 4.1 we have
It is easy to check the estimate
as and this is obviously fulfilled if . To control we use Sobolev inequality
so we need
Summing up we have the following restrictions for
(4.6) | ||||
Choosing with sufficiently small we see that (4.6) is nonempty. Now notice that
(4.7) |
Now we consider the case We have where is integer and Then the commutator relation
implies
In fact, we have the relation
and we use (4.5) so that
for
The term
can be evaluated pointwise via the classical Leibnitz rule and then via the fractional Leibnitz rule as follows
Summing up, we conclude that (4.4) holds in case
To conclude that the embedding is compact it remains to show that also . To this purpose we first use the pointwise decay in terms of homogeneous Sobolev norm, see [7]. Given there exists with such that
(4.8) |
with . Secondly we use that , i.e. that is non endpoint. By Theorem 1.1 there exists sufficiently small such that is uniformly bounded in and the same holds hence for and . As a consequence we have
with
This proves that 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 and let . Then there are constants such that for any measurable function , a measure space, with
one has (with denoting the underlying measure on )
(5.1) |
Lemma 5.2 (Compactness up to translations in [2]).
Let , and be a sequence with
(5.2) |
and, for some , (with denoting Lebesgue measure)
(5.3) |
Then there is a sequence such that a subsequence of has a weak limit in .
The strategy to prove Theorem 1.3 follows the one developed in [2]. First we aim to show that the maximum of
is achieved in . Let us consider a maximizing sequence . Since is invariant under homogeneity and scaling for any , we can choose a maximizing sequence such that
(5.4) |
and
(5.5) |
The key observation is that, since we are looking at a non-endpoint case (i.e. ), there exists such that from inequality (1.9) we infer that
(5.6) |
The -lemma (Lemma 5.1) now implies that
(5.7) |
Next, we apply the compactness modulo translations lemma (Lemma 5.2) to the sequence . This sequence is bounded in 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 in . By the fact the embedding is compact we deduce that in and hence is a maximizer for .
Now we conclude showing that .
Indeed if the best constant is , the maximizer achieves the equality in Hölder’s inequality, which means
(5.8) | ||||
where we used as conjugated exponents and . Now we recall that if and with and conjugated exponents achieve the equality in Hölder’s inequality then and shall be linearly dependent, i.e. for a suitable almost everywhere. Therefore, calling and , the maximizer should satisfy for a suitable which drives to the contradiction
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:
and its complementary set In we use
and via
with we deduce
These estimates imply
(6.1) |
For the complementary domain we use spherical coordinates where We have to estimate
where
(6.2) |
To get the desired estimate
(6.3) |
it is sufficient to check the pointwise estimate
(6.4) |
To deduce this pointwise estimate of the kernel we note first that does not depend on so we can take and and get
Using the relation
we can use the
when and is not close to say Then we get
For close to , say we use
In this way, making the change of variable we get
in view of and Taking together the above estimates of the integrals over and , 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. Luc, 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.