Holomorphic stability for Carleman pairs of function spaces
Abstract.
We introduce a notion of holomorphic stability for pairs of function spaces on a planar domain . In the case of the open unit disk equipped with a radial measure , by establishing Bourgain-Brezis type inequalities, we show that the pair
of weighted harmonic Bergman space and harmonic Hardy space is holomorphically stable if and only if is a -Carleson measure. With some extra efforts, we also obtain an analogous result for the upper half plane equipped with horizontal-translation invariant measures.
Key words and phrases:
holomorphic stability, weigthed Bergman spaces, Hardy spaces, Zen-type spaces, Bourgain-Brezis type inequalities1991 Mathematics Subject Classification:
Primary 30H10, 30H20; Secondary 32A10, 32A361. Introduction
1.1. Holomorphic stability for Carleman pairs
Let be a planar domain and let be the space of holomorphic functions on . In what follows, we shall always consider a pair of vector spaces both consisting of functions on .
Definition (Holomorphic stability).
The pair is called holomorphically stable if one of the following equivalent conditions is satisfied:
-
(i)
;
-
(ii)
-
(iii)
Here and .
The Venn diagram in Figure 1 illustrates a holomorphically stable pair .

For any and , we interpret as a holomorphic perturbation of . Then the holomorphic stability of a pair means that a holomorphic perturbation of any element in remains in provided that it is contained in after the perturbation.
Note that a holomorphically stable pair always satisfies the condition
Such pairs will be referred as Carleman pairs. This terminology comes from the classical Carleman embeddings of some classical holomorphic-function spaces on the open unit disk in the complex plane (see [Car21, Sai79, GK89, Vuk03]). For its generalization to complex domains of arbitrary dimension, see Hörmander [Hor67]. Note that if , then the pair is trivially holomorphically stable. Therefore, the notion of the holomorphic stability is of interests only for the pairs satisfying
Our research is inspired by Da Lio, Rivière and Wettstein’s very recent work [DLRW21] on Bourgain-Brezis type inequalities, where they essentially proved that the pair
of the harmonic Bergman space and the classical harmonic Hardy space is holomorphically stable (the precise definitions of and are given in §1.2). The notion of holomorphic stability leads to many natural questions. Generalization of our work in more general planar domains (including the non-simply connected ones) is more involved and will be given in the sequel to this paper.
1.2. Main results
The harmonic Hardy space is defined by
where denotes the space of all harmonic functions on . The Hardy space is then defined as .
Throughout the paper, all measures are assumed to be positive measures. Given a measure on , the associated weighted harmonic Bergman space and weighted Bergman space are defined as
both of which inherit the norm of . If is the Lebesgue measure on , then we use the simplified notation and .
A measure on is called boundary-accessable if its support is not relatively compact in . Note that if is boundary-inaccessable, then it is trivial to verify that the pair is holomorphically stable. Therefore, in what follows, we always assume that is boundary-accessable.
A measure on is called a -Carleson measure if there exists a constant such that
(1.1) |
Theorem 1.1.
Let be a radial boundary-accessable measure on . Then the pair
is holomorphically stable if and only if is a -Carleson measure.
A natural conjecture is
Conjecture.
For any boundary-accessable measure on , the pair is holomorphically stable if and only if is a -Carleson measure.
For a finite measure on which is not necessarily radial, a simple situation (which in general is rather different from the situation in Theorem 1.1, see Theorem 1.2 below for more details) for the holomorphic stability of is provided as follows. Consider the linear map defined on the space by
(1.2) |
Then the pair is holomorphically stable if both
(1.3) |
and
(1.4) |
are bounded linear operators.
It is not hard to see that the operator (1.3) is bounded if and only if
(1.5) |
Hence the boundedness of the operator (1.3) in general fails even for a radial -Carleson measure. On the other hand, the boundedness of (1.4) holds for all radial finite measure on . For general weights, a clear sufficient condition for the boundedness of the operator (1.4) is that the Bergman projection being bounded on (which then is equivalent to the condition that is a -weight à la Békollé-Bonami, see [BB78] for more details on Bergman projections). Consequently, for any -weight on satisfying (1.5), the pair is holomorphically stable.
For a radial boundary-accessable finite measure on , the space is complete and is a Banach space equipped with the norm:
Recall that a closed subspace of a Banach space is called complemented in if there exists a bounded linear projection from onto .
Theorem 1.2.
Let be a radial boundary-accessable -Carleson measure on . Then
is a closed subspace of . Moreover, the above subspace is complemented in if and only if satisfies the condition
(1.6) |
In general, the notion of holomorphic stability for the pair is not conformally invariant, where is assumed to have a nice boundary and is then defined as the set of all the Poisson convolutions of functions in (where is the arc-length measure on ). In particular, our following result for the upper half plane does not seem to be a direct consequence of the result on the unit disk and its proof requires more efforts.
Let denote the upper half plane. In this case, the suitable spaces for studying the holomorphic stability are the harmonic Zen-type spaces (which reduces to the ordinary harmonic Bergman space when the weight is the Lebesgue measure on ), see [Har09, JPP13].
A measure on is called boundary-accessable if its support is not contained in
for any . Given a horizontal translation-invariant boundary-accessable measure on , define the harmonic Zen-type space by
(1.7) |
where denotes the set of all harmonic functions on . It is easy to see that the above space is complete and thus is a Hilbert space.
The relation between and the ordinary weighted harmonic Bergman space
is given as follows. For any and any , define by
(1.8) |
Recall the Poisson kernel for at the point :
(1.9) |
Set
(1.10) |
One can easily check that, for any horizontal translation-invariant boundary-accessable measure on ,
(1.11) |
The harmonic Hardy space is defined by
and the Hardy space is defined as . A measure on is called a -Carleson measure if there exists a constant such that
(1.12) |
Theorem 1.3.
Let be a horizontal translation-invariant boundary-accessable measure on . Then the pair
is holomorphically stable if and only if is a -Carleson measure.
1.3. Sketch of the proof
Here we give a sketch of the proof of Theorem 1.1. The proof is based on a generalization of the following one-dimensional Bourgain-Brezis-type inequality due to Da Lio-Rivière-Wettstein: there exists a universal constant such that for any smooth function with ,
(1.13) |
where the Hilbert transform of is given by
and the -fractional Laplace transform is given by
The inequality (1.13) implies
(1.14) |
where, slightly by abusing the notation, is defined by
(1.15) |
In our situation, we are able to prove that, if is a radial boundary-accessable measure, then is a -Carleson measure on if and only if there exists a constant such that for any bounded harmonic function ,
(1.16) |
The inequality (1.16) applied to holomorphic functions immediately gives the result stated in Theorem 1.1.
The proof of the inequality (1.16) relies on a weighted version of Bourgain-Brezis-type inequality obtained in Theorem 1.4 below. More precisely, define a Fourier multiplier operator by
(1.17) |
where
Define also a Sobolev-type space corresponding to the radial weight by
(1.18) |
Remark.
For a general , the coefficients of a formal Fourier series may have non-polynomial growth and it may not represent a distribution in . However, if is the Lebesgue measure on , then is the Sobolev space .
Theorem 1.4.
Let be a radial boundary-accessable finite measure on . Then is a -Carleson measure if and only if there exists a universal constant such that for any smooth function ,
(1.19) |
2. Preliminaries on Carleson measures
Recall that throughout the paper, all measures are assumed to be positive. We shall use the famous geometric characterization of the -Carleson measures on defined in (1.1). For any interval , the Carleson box is defined by
where denotes the arc-length of . Let denote the Lebesgue measure of , then a measure on is a -Carleson measure if and only if (see [Car62] and [Dur69])
In particular, we have
Lemma 2.1.
Let be a radial measure on . Then is a -Carleson measure if and only if
(2.1) |
The -Carleson measures on the upper half plane is defined in (1.12) and its geometric characterization (see, e.g., [Ryd20, Thm. 2.1]) is given as follows: a positive Radon measure on is a -Carleson measure on if and only if
where is the Lebesgue measure of the Carleson box defined by
here denotes the Lebesgue measure of the interval . In particular, we have
Lemma 2.2.
Let be a horizontal translation-invariant measure on . Then is a -Carleson measure if and only if
(2.2) |
Remark.
While all -Carleson measures on are finite, a -Carleson measure on needs not be.
3. Holomorphic stability: the disk case
This section is mainly devoted to proving Theorems 1.1 and 1.2. We shall use the following elementary observation: if is a radial boundary-accessable finite measure on , then
-
•
both and are closed in ;
-
•
for any , the evaluation map defined by
(3.1) is a continuous linear functional on ;
-
•
for any and any ,
(3.2)
Recall the definition (1.18) of the space : for any , we set
(3.3) |
By (3.2) and (3.3), for any , the Poisson transformation of an element is a smooth function given by
Any has a natural harmonic extension (denoted again by ) on :
(3.6) |
3.1. The derivation of Theorem 1.1 from Theorem 1.4
If is a holomorphically stable pair, then we have set-theoretical inclusion
It follows that is a finite measure, which, when combined with the assumption of the theorem, implies that is a radial boundary-accessable finite measure on . Therefore, is complete. Hence the embedding is continuous by the Closed Graph Theorem. In other words, is a -Carleson measure on .
Now assume that is a radial boundary-accessable -Carleson measure on . To prove the holomorphic stability of the pair , it suffices to show that there exists a constant such that
(3.7) |
Indeed, assuming (3.7) and let with , we obtain
and hence . It follows that and this gives the holomorphic stability of the pair .
It remains to prove (3.7). For any and any , write . Then, since is radial,
and
Therefore, it suffices to show that (3.7) holds for all belonging to the following class:
We now proceed to the derivation of the inequality (3.7) for any from the inequality (1.19) obtained in Theorem 1.4. Observe that, any has the form
where is defined as in (3.6). Then, by the radial assumption on ,
(3.8) |
Comparing (3.8) and (1.18), we obtain a natural isometric isomorphism that associates any to its harmonic extension in defined by (3.6). Similarly, there is a natural isometric isomorphism . Therefore, we get a natural identification of the Banach spaces
Hence, for any , by taking in (1.19), we obtain
with defined in (1.17), defined in (1.15). Since for all and , we get
Finally, notice that, for any ,
Thus we obtain the desired inequality (3.7) for all and complete the derivation of Theorem 1.1 from Theorem 1.4.
3.2. The proof of Theorem 1.4
A pair of sequences and is called -adapted if the following conditions are satisfied:
-
(i)
;
-
(ii)
for any ,
(3.9) -
(iii)
there exists a constant such that
(3.10)
Note that the condition (3.9) implies in particular that for all .
Proposition 3.1.
Suppose is a radial boundary-accessable -Carleson measure on . Let be a -adapted pair of sequences. Then there exists a constant such that
(3.11) |
where and are the Fourier multipliers defined by
The next criterion of radial -Carleson measures will be useful for us.
Lemma 3.2.
Let be a finite measure on . Then the inequality
(3.12) |
holds if and only if
Lemma 3.3.
Let be a radial boundary-accessable -Carleson measure on , then there exists a function such that
(3.13) |
Proof.
Proof of Proposition 3.1.
Let . Note that Clearly, if with , then and
It follows that . Therefore, from now on, we may assume that . Take any pairs of decompositions
(3.17) |
with and . Then for any , we have (the following Poisson convolutions will be used in the proof of the equality (3.29) below)
(3.20) |
That is,
(3.23) |
From (3.23), we have
(3.24) |
By Cauchy-Schwarz’s inequality,
(3.25) |
where we used the fact that if is a -adapted pair of sequences, then by (3.9) and (3.10), for any ,
(3.26) |
By (3.23),
(3.27) | II |
Then by Cauchy-Schwarz’s inequality,
Using similar inequality as (3.26), under the conditions (3.9) and (3.10), we have
(3.28) |
We now proceed to the estimate of term in the decomposition (3.27). Note that
For any , set
A priori, we only have , but for any . By (3.9),
By Lemma 3.3, . Then the Plancherel’s identity implies
(3.29) |
Hence
(3.30) |
By (3.24), (3.25), (3.27), (3.28) and (3.30), there is a constant , depending only on and the measure but not on , such that
Therefore, by a standard argument, there is a constant such that
where the last inequality is due to the contractive property of the Poission convolution on both and . Let approach to , then
Since the decompositions (3.17) are arbitrary, we obtain the desired inequality (3.11). ∎
Proof of Theorem 1.4.
It remains to prove Lemma 3.2. We shall apply a result due to Garnett about the boundary behavior of Poisson integrals on the upper half plane .
Lemma 3.4 (Garnett, see, e.g., [RU88, pp. 210]).
Let be a measure on with Then the following two assertions are equivalent:
-
•
-
•
Proof of Lemma 3.2.
Note that
For any , consider the point and recall the Poisson kernel at the point given in (1.9), then
Consider the Möbius transformation defined by Then is an automorphism of the upper half plane and
Note that when ranges over , the image ranges over . Therefore,
By change-of-variable ,
Then
where . Clearly, . Therefore, by Lemma 3.4, the inequality (3.12) holds if and only if
(3.31) |
By the definition of , it is easy to see that , where is the open interval
It follows that, (3.31) holds if and only if , which in turn is equivalent to
By Lemma 2.1, the above inequality holds if and only if is a -Carleson measure on . This completes the whole proof. ∎
3.3. Proof of Theorem 1.2
Fix a radial boundary-accessable -Carleson measure By Theorem 1.1,
(3.32) |
By (3.7), there exists a constant , such that for all ,
(3.33) |
That is, the identity map
is an isomorphic isomorphism. Thus is a closed subspace of .
Now suppose that the extra condition (1.6) is satisfied. We are going to show that the closed subspace is complemented in . Indeed, under the condition (1.6), for any , the following holomorphic function
belongs to and
(3.34) |
Recall the definition (1.2) of . Clearly, since is radial, defines an orthgonal projection from onto . Then
(3.35) |
Note also that if , that is,
then it is easy to see that
Hence, by (3.34),
(3.36) |
By (3.35) and (3.36) and the definition of the norm on ,
It follows that defines a bounded linear projection from onto
Hence is complemented in .
Finally, assume that the condition (1.6) is not satisfied. Then
Let us show that is not complemented in . Otherwise, there exists a bounded linear projection operator
onto the closed subspace . That is,
-
•
,
-
•
for all ,
-
•
for all .
Since is radial, for any , the rotation map defined by preserves both the norms of functions in and the norms of functions in . Therefore, preserves the norms of functions in :
Consequently, the operator-norm of the composition operator
is bounded by that of :
It can be easily checked that is also a projection operator from onto .
Define a bounded linear operator via the Bochner integral (see, e.g., [Yos95, Section V.5])
Then
(3.37) |
and
(3.38) |
Since the evalutation map defined in (3.1) is a continuous linear functional on for any ,
(3.39) |
Note that for any and any , thus
On the other hand, for any integer ,
(3.40) |
Indeed, for any ,
Thus . By (3.32),
we can write
with
(3.41) |
Thus, for all ,
where the last series converges absolutely by the inequalities (3.2), (3.41) and
Therefore, by (3.39), for all ,
This is the desired equality (3.40).
However, if we take the harmonic extension of the Féjer kernel on :
then
(3.42) |
Since is a projection onto and satisfies (3.40), we have
and, by the radial assumption on ,
Then, by (3.33),
This contradicts to the following inequality (which is a consequence of (3.37) and (3.42))
Hence we complete the whole proof of the theorem.
4. Holomorphic stability: the upper half plane case
In this section, we will prove Theorem 1.3. For any Radon measure on satisfying the condition (2.2), define
Recall the following definition of the Fourier transform for :
Recall the definition (1.7) of . For any and , recall the definition of the function defined in (1.8). Note that the Fourier transform of the Poisson kernel given in (1.9) has the following form (see [Kat04, Chapter VI, p. 140]):
Then, by (1.10) and (1.11), we have for all and hence
(4.1) |
Definition.
By (4.2) and the Plancherel’s identity, the norm of any has the form:
(4.3) |
Remark.
Definition.
Suppose that is a boundary-accessable -Carleson measure on . Let be the Hilbert space defined by the norm completion as follows:
For any , set
Immediately from the definition of , we see that the map
extends to a unitary map from to .
Similar to the disk case, for a given measure on , a pair of two functions on is called -adapted if the following conditions are satisfied:
-
(i)
for any ,
(4.4) -
(ii)
there exists a constant such that
(4.5)
Given any Radon measure on satisfying (2.2), Garnett’s result stated in Lemma 3.4 implies that the following function belongs to :
(4.6) |
Proposition 4.1.
Suppose that is a boundary-accessable -Carleson measure on and let be a -adapted pair of functions defined on . Then for any ,
(4.7) |
where are the Fourier multipliers associated to given by
and the constant can be taken to be
(4.8) |
Remark.
If either or does not belong to , then the right hand side of (4.7) is understood as .
4.1. The derivation of Theorem 1.3 from Proposition 4.1
Let be a boundary-accessable Radon measure on . If the pair is holomorphically stable, then and by the Closed Graph Theorem, this embedding is continuous: there exists such that
(4.9) |
Recall the definition (1.10) of the space . Since ,
(4.10) |
The inequality (4.9) and the equality (4.10) together imply that the measure is a -Carleson measure.
Suppose now that is a boundary-accessable -Carleson measure on . Assume that
(4.11) |
Then, the goal is to show that . It suffices to show
(4.12) |
To avoid technical issues, we first consider the truncated measures of . That is, for any , define
Define in a similar way as in (4.6). Then for any . Therefore, the desired inequality (4.12) follows from
(4.13) |
Now we are going to apply Proposition 4.1. For any , define a -adapted pair of functions by
In particular, by (2.2),
(4.14) |
For any , define and by
And are defined similarly.
Claim I.
For any , the function belongs to the classical analytic Hardy space and hence
(4.15) |
Remark.
The assertion (4.15) does not follow from the fact that is the restriction onto the real line of a holomorphic function defined on a neighborhood of the closed upper-half plane. For instance, the following function
belongs to and is the restriction of an entire function on the complex plane. However, .
Since , there exists with
(4.16) |
Thus By (1.11), the assumption implies that . Consequently
(4.17) |
Again by (4.16) and (1.11), for any ,
Therefore, for any ,
The above inequality combined with implies that . This completes the proof of Claim I.
Since , for any , the function belongs to and hence . Then by the natural unitary map between and ,
Note that the equality (4.17) and the inequality (4.14) together imply that the function belongs to . Then there exists a unique function with
(4.18) |
Hence, by (4.15), . It follows that,
That is, Therefore, since , we may apply (4.7) and get
where the last inequality is due to the simple observation: for any ,
Finally, by Plancherel’s identity, the equalities (4.18) and (4.3),
Thus,
The inequality (4.13) now follows immediately since
4.2. The proof of Proposition 4.1
Lemma 4.2.
Let be a Radon measure on satisfying (2.2). Then there is a function such that the following equality
(4.19) |
holds for all satisfying
(4.20) |
Remark.
The equality (4.19) means that the Fourier transform, in a certain distributional sense, of the function , is given by . If is the Lebesgue measure on , then
In general, the Fourier transform of can only be understood in a certain distributional sense and the condition (4.20) in Lemma 4.2 can not be removed.
The proof of Lemma 4.2 is postponed to the end of this section.
Proof of Proposition 4.1.
Take . Suppose that we have decompositions
(4.21) |
with and . That is,
(4.22) |
For any fixed , applying the Poisson convolution to both sides of (4.21), we have
By Plancherel’s identity and (4.22),
(4.23) |
Cauchy-Schwarz’s inequality and the conditions (4.4), (4.5) together imply
(4.24) |
And, by (4.22) and , the integral can be decomposed as
(4.25) |
The integral can be easily controlled. Indeed, again by Cauchy-Schwarz’s inequality and (4.4), (4.5),
(4.26) |
It remains to estimate the integral . Since , for any , one can define
(4.27) |
In particular,
(4.28) |
Claim A.
Indeed, since . Therefore, (4.29) follows from the definition (4.27) and the simple observation that . By (4.22),
The assumptions combined with the conditions (4.4), (4.5) on the pair imply that both functions and belong to . Since and hence , we obtain, by using (4.4) again, that
By Claim A, the function satisfies all the required conditions of Lemma 4.2. Hence, by (4.28), (4.4) and (4.19),
It follows that
(4.31) |
Combining (4.23), (4.24), (4.25), (4.26) and (4.31), we get
Therefore, by a standard argument, there exists a constant depending only on the constants and such that
The constant in the above inequality can be taken to be
Since the decompositions (4.21) are arbitrary, we get
Finally, by taking the limit and using
we obtain the desired inequality (4.7) and complete the whole proof of the proposition. ∎
Proof of Lemma 4.2.
Fix a Radon measure on satisfying (2.2). By Garnett’s result stated in Lemma 3.4, one can define a function by (4.6).
Now we show that satisfies the equality (4.19). For any , set
(4.32) |
Claim B.
For any , we have and the Fourier transform of is given by the Bochner integral for -vector valued function:
(4.33) |
In particular, can be identified with a -function by the formula
(4.34) |
Indeed, for any , recall that the conjugate Poisson kernel (see, e.g., [Gra14, formula (4.1.16)]) of is given by
Clearly, for all and the map is continuous from to , hence it is uniformly continuous from to . Consequently, using the definition (4.32) of and the fact that is a finite measure with support contained in , we obtain that and the following equality in the sense of the Bochner integral for -vector valued functions:
Then the equality (4.33) follows immediately since (see, e.g., [Gra14, formula (4.1.33)])
Claim C.
For any ,
(4.35) |
Indeed, for any , set . Then and by the assumption (2.2), there exists a constant such that
(4.36) |
By integration by parts for the absolutely continuous function ,
(4.37) |
In particular, if is the Lebesgue measure on , then the equality (4.37) becomes
(4.38) |
Comparing (4.37) and (4.38) and using (4.36), we obtain
Therefore, by dominated convergence theorem, for any ,
Claim D.
References
- [BB78] David Bekollé and Aline Bonami. Inégalités à poids pour le noyau de Bergman. C. R. Acad. Sci. Paris Sér. A-B, 286(18):A775–A778, 1978.
- [Car21] Torsten Carleman. Zur Theorie der linearen Integralgleichungen. Math. Z., 9(3-4):196–217, 1921.
- [Car62] Lennart Carleson. Interpolations by bounded analytic functions and the corona problem. Ann. of Math. (2), 76:547–559, 1962.
- [DLRW21] Francesca Da Lio, Tristan Rivière, and Jerome Wettstein. Bergman-Bourgain-Brezis-type inequality. J. Funct. Anal., 281(9):Paper No. 109201, 33, 2021.
- [Dur69] Peter L. Duren. Extension of a theorem of Carleson. Bull. Amer. Math. Soc., 75:143–146, 1969.
- [GK89] Theodore W. Gamelin and Dmitry S. Khavinson. The isoperimetric inequality and rational approximation. Amer. Math. Monthly, 96(1):18–30, 1989.
- [Gra14] Loukas Grafakos. Classical Fourier analysis, volume 249 of Graduate Texts in Mathematics. Springer, New York, third edition, 2014.
- [Har09] Zen Harper. Boundedness of convolution operators and input-output maps between weighted spaces. Complex Anal. Oper. Theory, 3(1):113–146, 2009.
- [Hor67] Lars Hormander. estimates for (pluri-) subharmonic functions. Math. Scand., 20:65–78, 1967.
- [JPP13] Birgit Jacob, Jonathan R. Partington, and Sandra Pott. On Laplace-Carleson embedding theorems. J. Funct. Anal., 264(3):783–814, 2013.
- [Kat04] Yitzhak Katznelson. An introduction to harmonic analysis. Cambridge Mathematical Library. Cambridge University Press, Cambridge, third edition, 2004.
- [RU88] Wade Ramey and David Ullrich. On the behavior of harmonic functions near a boundary point. Trans. Amer. Math. Soc., 305(1):207–220, 1988.
- [Ryd20] Eskil Rydhe. On Laplace-Carleson embeddings, and -mapping properties of the Fourier transform. Ark. Mat., 58(2):437–457, 2020.
- [Sai79] Saburou Saitoh. The Bergman norm and the Szego norm. Trans. Amer. Math. Soc., 249(2):261–279, 1979.
- [Vuk03] Dragan Vukotić. The isoperimetric inequality and a theorem of Hardy and Littlewood. Amer. Math. Monthly, 110(6):532–536, 2003.
- [Yos95] Kōsaku Yosida. Functional analysis. Classics in Mathematics. Springer-Verlag, Berlin, 1995. Reprint of the sixth (1980) edition.