Abstract

We characterize those measures 𝜇 for which the Hardy-Orlicz (resp., weighted Bergman-Orlicz) space 𝐻Ψ1 (resp., 𝐴Ψ1𝛼) of the unit ball of 𝑁 embeds boundedly or compactly into the Orlicz space 𝐿Ψ2(𝔹𝑁,𝜇) (resp., 𝐿Ψ2(𝔹𝑁,𝜇)), when the defining functions Ψ1 and Ψ2 are growth functions such that 𝐿1𝐿Ψ𝑗 for 𝑗{1,2}, and such that Ψ2/Ψ1 is nondecreasing. We apply our result to the characterization of the boundedness and compactness of composition operators from 𝐻Ψ1 (resp., 𝐴Ψ1𝛼) into 𝐻Ψ2 (resp., 𝐴Ψ2𝛼).

1. Introduction and Preliminaries

1.1. Introduction

Let 𝔹𝑁={𝑧=(𝑧1,,𝑧𝑁)𝑁|𝑧|2=𝑁𝑖=1|𝑧𝑖|2<1} and 𝕊𝑁=𝜕𝔹𝑁, 𝑁1, denoting, respectively, the unit ball and the unit sphere of 𝑁. For 𝑁=1, we denote by 𝔻 the unit disc of the complex plane.

For a large class of spaces 𝑋 of holomorphic functions in the unit disc or the unit ball, characterizations of the boundedness and compactness of the canonical embedding 𝑋𝐿𝑝(𝜇) have been given and applied to different areas, for example, interpolation, multipliers, integral operators, composition operators, and so forth. These results are known as Carleson’s type theorems.

First, when 𝑋=𝐻𝑝(𝔻), Carleson [1] proved that 𝐻𝑝(𝔻)𝐿𝑝(𝜇) if and only if the finite positive Borel measure 𝜇 on 𝔻 (or 𝔻) is a so-called Carleson measure. This result was extended to the unit ball by Hörmander [2], whose proof was simplified by Power [3]. Duren [4] characterized those measures 𝜇 such that 𝐻𝑝(𝔻)𝐿𝑞(𝜇), with 0<𝑝𝑞<, in terms of (𝑞/𝑝)-Carleson measures. For the unweighted and weighted Bergman spaces 𝐴𝑝𝛼(𝔹𝑁), 𝑁1, similar results were obtained by Cima and Wogen [5], Luecking [6], Ueki [7]. We recall also that the compactness of 𝑋𝐿𝑝(𝜇) was also characterized in the previous cases, in terms of vanishing Carleson’s type measures. It is usual to assume that when the measure 𝜇 is defined on 𝔹𝑁, then 𝜇𝕊𝑁 is absolutely continuous with respect to 𝕊𝑁. This assumption will be done, without mentioning it any further.

Some observations may be done. First, it appears that the characterizations of both boundedness and compactness of 𝐻𝑝(𝔹𝑁)𝐿𝑞(𝜇) are not always satisfied and just depend on the ratio 𝑞/𝑝 for 0<𝑝𝑞<; in particular, when 𝑝=𝑞, they are independent of 𝑝. Now, since the restriction to 𝕊𝑁 of the finite positive Borel 𝜇 is assumed to be absolutely continuous with respect to the Lebesgue measure, then it is trivial that 𝐻𝐿(𝜇) always holds. On the contrary, the compactness of this inclusion implies strong condition on 𝜇. This suggests to think about this question when the space 𝑋 is intercalated between every 𝐻𝑝 and 𝐻.

This motivation is reinforced by a second observation: if the measure 𝜇𝜙 is the pull-back measure of the invariant-rotation measure on 𝕊𝑁 under a holomorphic map 𝜙𝔹N𝔹𝑁, 𝑁1, then 𝜇𝜙 is always a Carleson measure when 𝑁=1 (this is the Littlewood Subordination Principle, see [8]), but it is not systematic for 𝑁>1. For the compactness, there is still a big gap between 𝐻𝑝 and 𝐻 for any 𝑁1. This observation is directly connected to the study of composition operators 𝐶𝜙 on 𝐻𝑝, which are defined by 𝐶𝜙(𝑓)=𝑓𝜙 for 𝑓𝐻(𝔹𝑁), and which may be seen as the embedding operators 𝐻𝑝𝐿𝑝(𝜇𝜙). By the way, this leads the authors of [9, 10] to state Carleson theorems for Hardy-Orlicz and Bergman-Orlicz spaces (resp., denoted by 𝐻Ψ(𝔻) and 𝐴Ψ(𝔻)) in the unit disc, when the defining function Ψ is an Orlicz function. These spaces appear as good candidates for generalizing 𝐻𝑝 and 𝐴𝑝 spaces, 1𝑝<, and for covering the gap with 𝐻. This fact was still more pointed out in [11, 12], where the author gave Carleson theorems in the unit ball, under some mild conditions on the defining function Ψ. Indeed, he showed that, when 𝜇=𝜇𝜙 is the pull-back measure under a holomorphic self-map 𝜙 of 𝔹𝑁, then 𝐻Ψ always embeds into 𝐿Ψ(𝜇) whenever Ψ satisfied a fast growth condition (namely, the Δ2-Condition, that implies Ψ(𝑥)𝑒𝑥 for large value of 𝑥 and means that we are close to 𝐻). Similarly, it was shown in [13] that if 𝐻Ψ compactly embeds into 𝐿Ψ(𝜇) for every Ψ, then 𝐻𝐿(𝜇) compactly (the converse easily holds). We mention that there is no Ψ such that 𝐻Ψ𝐿Ψ(𝜇) if and only if 𝐻𝐿(𝜇) [14], and that such previous results are not true for arbitrary measure 𝜇. Yet, a link has been made between the involved Carleson conditions and the type of growth of the Orlicz function Ψ. This is also strengthened by the fact that if the Orlicz function Ψ is dominated by a power function (exactly Ψ satisfies the Δ2-Condition), then 𝐻Ψ𝐿Ψ(𝜇) boundedly (resp., compactly) if and only if 𝐻𝑝𝐿𝑝(𝜇), that is, if and only if 𝜇 is a classical Carleson measure (resp., a vanishing Carleson measure).

Let us note that similar results hold for Bergman-Orlicz spaces.

The purpose of the present paper is to deal with the same kind of question on the opposite side, that is, for Hardy-Orlicz and Bergman-Orlicz spaces which are larger than 𝐻1(𝔹𝑁). It seems that nothing has been done in this direction, except for explicit functions Ψ. In particular, the second author gave a necessary and sufficient condition for the inclusion 𝐻Ψ(𝔹𝑁)𝐿𝑞(𝜇) to be bounded, when Ψ(𝑡)=(𝑡/log(𝑒+𝑡))𝑠, 0<𝑠1, and 1𝑞< [15]. Moreover, [16] characterized Carleson measures for 𝐴Ψ𝛼(𝔻), where Ψ=log𝑝+, 1𝑝<, that is, when 𝐴Ψ𝛼(𝔻) is the area Nevanlinna space. These measures reveal to be those 𝜇 for which 𝐴𝑝𝛼𝐿𝑝(𝜇) holds, which are Bergman-Carleson measures.

A first difficulty when dealing with large Hardy-Orlicz or Bergman-Orlicz spaces is that we do not have normed spaces any more, and we need to exhibit the good properties of the function Ψ, in order to define spaces with which it is reasonable to work. For that, we were inspired by [17, 18] and the references therein. Then, we obtain more generally a complete characterization of those finite positive Borel measures 𝜇 such that 𝐻Ψ1(𝔹𝑁)𝐿Ψ2(𝔹𝑁,𝜇) (resp., 𝐴Ψ1𝛼(𝔹𝑁)𝐿Ψ2(𝔹𝑁,𝜇)) is bounded or compact, when Ψ1 and Ψ2 are two growth functions (i.e., for which 𝐻1𝐻Ψ𝑖𝐻𝑝 (resp., 𝐴1𝛼𝐴Ψ𝑖𝛼𝐴𝑝𝛼) for some 𝑝, 𝑖,𝑗{1,2}), and such that Ψ2  grows faster than Ψ1. It appears that if Ψ1=Ψ2, then such measures are exactly those which are Carleson (resp., Bergman-Carleson) measures for some 0<𝑝<1. For the Bergman-Orlicz case, these results let one think that, between the area Nevanlinna class [16] and 𝐴1𝛼, there is no difference regarding to Carleson theorems, whenever the defining functions Ψ share some natural properties. This is in contrast with what happens between 𝐴1𝛼 and 𝐻.

The paper is organized as follows. In the next subsection, we introduce the Hardy-Orlicz and Bergman-Orlicz spaces under further considerations, proving or generalizing some useful and classical results. The second section consists of the statements and the proofs of our Carleson theorems for these spaces. A third and last part is an immediate application of our result to composition operators.

Notation. Given two points 𝑧,𝑤𝑁, the euclidean inner product of 𝑧 and 𝑤 will be denoted by 𝑧,𝑤, that is, 𝑧,𝑤=𝑁𝑖=1𝑧𝑖𝑤𝑖; the notation || will stand for the associated norm, as well as for the modulus of a complex number.

𝜎 will stand for the invariant-rotation measure on the unit sphere. For 𝛼>1,𝑣𝛼 will be the measure on 𝔹𝑁 defined by 𝑑𝑣𝛼=𝑐𝛼(1|𝑧|)𝛼𝑑𝑣, where 𝑣 is the Lebesgue measure on 𝔹𝑁 and 𝑐𝛼 is the constant of normalization.

We will use the notations and for one-sided estimates up to an absolute constant, and the notation for two-sided estimates up to an absolute constant.

Without possible confusions, we will write 𝐻𝜓 (resp., 𝐴𝜓𝛼) instead of 𝐻𝜓(𝔹𝑁) (resp., 𝐴𝜓𝛼(𝔹𝑁)).

1.2. Preliminaries: Menagerie of Spaces

Let Ψ[0,)[0,) be a continuous nondecreasing function which vanishes and is continuous at 0. Given a probabilistic space (Ω,), we define the Orlicz class 𝐿Ψ(Ω,) as the set of all (equivalence classes of) measurable functions 𝑓 on Ω such that ΩΨ(|𝑓|/𝐶)𝑑< for some 0<𝐶<. We use to define the Morse-Transue space 𝑀Ψ(Ω,) by𝑀Ψ(Ω,)=𝑓Ωmeasurable;ΩΨ||𝑓||𝐶,<forany𝐶>0(1.1) and we also introduce the following set:Ψ(Ω,)=𝑓Ωmeasurable;ΩΨ||𝑓||<.(1.2) In general, these three sets are not vector spaces and do not coincide, but we trivially have𝑀Ψ(Ω,)Ψ(Ω,)𝐿Ψ(Ω,).(1.3) We also define the Luxembourg gauge on 𝐿Ψ(Ω,) by𝑓Ψ=inf𝜆>0,ΩΨ||𝑓||𝜆.1(1.4) This functional is homogeneous and is 0 if and only if 𝑓=0a.e., but it is not subadditive a priori.

We say that two functions Ψ1 and Ψ2 as above are equivalent if there exists some constant 𝑐 such that𝑐Ψ1(𝑐𝑥)Ψ2(𝑥)𝑐1Ψ1𝑐1𝑥,(1.5) for any 𝑥 large enough. Two equivalent functions define the same Orlicz class with equivalent Luxembourg functionals.

In order to define a good topology on 𝐿Ψ(Ω,) and to get properties convenient for our purpose, we will assume that Ψ satisfies the following definition.

Definition 1.1. Let 0<𝑝1. We say that Ψ[0,)[0,) is a growth function of order 𝑝 if it satisfies the two following conditions:(1)Ψ is of lower type 𝑝, that is, Ψ(𝑦𝑥)𝑦𝑝Ψ(𝑥) for any 0<𝑦1 and at least for 𝑥 large enough;(2)𝑥Ψ(𝑥)/𝑥 is nonincreasing, at least for every 𝑥 large enough.

We shall say that 𝜓 is a growth function if it is a growth function of order 𝑝 for some 0<𝑝1.

In particular, a growth function Ψ is equivalent to the function 𝑥𝑥0(Ψ(𝑠)/𝑠)𝑑𝑠 which is concave (see [17]). Now, for such a concave growth function of order 𝑝,𝐿1(Ω,)𝑀Ψ(Ω,)=Ψ(Ω,)=𝐿Ψ(Ω,)𝐿𝑝(Ω,),(1.6)𝑓ΨminΩΨ||𝑓||𝑑,ΩΨ||𝑓||𝑑𝑝,(1.7) while𝑓Ψ=ΩΨ||𝑓||𝑑max𝑓Ψ,𝑓𝑝Ψ.(1.8) Moreover, if we define 𝑑Ψ(𝑓,𝑔)=𝑓𝑔Ψ (resp., 𝑑Ψ(𝑓,𝑔)=𝑓𝑔Ψ), then 𝑑Ψ and 𝑑Ψ (that we will simply denote by Ψ and Ψ) are two equivalent metrics on 𝐿Ψ(Ω,) for which it is complete. Without loss of generality, we then assume that every growth function that we consider further is concave and diffeomorphic.

Remark 1.2. Note that every concave function which vanishes at 0 satisfies the Δ2-condition, that is, Ψ(2𝑥)𝐾Ψ(𝑥) for some 𝐾>1 and 𝑥 large enough. This condition is very classical when the function Ψ is an Orlicz function, that is, a nondecreasing continuous convex function (see [19, 20]). For large Orlicz class, this condition is natural in order to have vector space.

When we deal with spaces of holomorphic functions, it is very natural to require subharmonicity. Then we will assume that Ψ is such that Ψ(|𝑓|) is subharmonic when 𝑓 is holomorphic. We will refer to such a function as a subharmonic-preserving function.

Here are some examples of (concave) growth function that we may consider further.

Example 1.3. (1)  Ψ1(𝑥)=𝑥𝑝 for 0<𝑝1 and any 𝑥𝑥0.
(2)  Ψ2(𝑥)=𝑥𝑝log𝑞(𝐶+𝑥) (for 𝑥𝑥0) with 𝐶>0 large enough, 0<𝑝1 and 𝑞1.
(3)  Ψ3(𝑥)=Φ𝑝(𝑥) at least for any 𝑥 large, where Φ is an Orlicz function (see Remark 1.2) and 𝑝>0 is such that Φ𝑝(𝑥)/𝑥 is nonincreasing. Note that in this case, Ψ(|𝑓|) is subharmonic whenever 𝑓 is holomorphic, because we have the following.

Proposition 1.4. Let Ψ be as in (3) above. There exists a convex function 𝜓 such that Ψ(𝑥)=𝜓(𝑥𝑝) for any 𝑥>0 large enough.

Proof. Let 𝑝>0 and Φ be an Orlicz function such that Ψ(𝑥)=Φ𝑝(𝑥) for any 𝑥0. Since Ψ is (assumed to be) bijective, we can define the function 𝜓 by 𝜓(𝑥𝑝)=Φ𝑝(𝑥) for every 𝑥. Now, using that Φ is convex, 𝜓 is convex since 𝜓(𝑥)𝑥=Φ𝑥1/𝑝𝑥1/𝑝𝑝Φ𝑦1/𝑝𝑦1/𝑝𝑝=𝜓(𝑦)𝑦,(1.9) for any 0𝑥𝑦.

It follows that Ψ(|𝑓|)=𝜓(|𝑓|𝑝) is subharmonic (for 𝑓 holomorphic).

The following lemma gives an upper estimate of the Luxembourg norm of a function in 𝐿(Ω,).

Lemma 1.5. Let (Ω,) be a probabilistic space and let Ψ be a growth function of order 𝑝. For any 𝑓𝐿(Ω,), one has 𝑓Ψ𝑓Ψ1𝑓/𝑓𝑝𝑝.(1.10)

Proof. It is quite identical to that of [10, Lemma 3.9], but we prefer to give the details. Without loss of generality, we may assume that 𝑓=1. For every 𝐶>0, one has the following, using that Ψ is of lower type 𝑝: ΩΨ||𝑓||𝐶𝑑Ω||𝑓||𝑝Ψ1𝐶𝑑𝑓𝑝𝑝Ψ1𝐶.(1.11) Then the last expression is less than or equal to 1 if and only if 𝐶1/Ψ1(1/𝑓𝑝𝑝).

1.2.1. Large Bergman-Orlicz Spaces

For Ψ a subharmonic-preserving growth function and 𝛼>1, the weighted Bergman-Orlicz space 𝐴Ψ𝛼 of the ball consists of those holomorphic functions on 𝔹𝑁 which belongs to the Orlicz space 𝐿Ψ(𝔹𝑁,𝑣𝛼). To avoid further confusion, we will denote by 𝛼,Ψ the corresponding Luxembourg (quasi)norm and by Ψ𝛼 the quantity 𝔹𝑁Ψ(|𝑓|)𝑑𝑣𝛼. 𝐴Ψ𝛼 is metric space for the distance 𝑑𝛼,Ψ or 𝑑Ψ𝛼 defined by, respectively, 𝑑𝛼,Ψ(𝑓,𝑔)=𝑓𝑔𝛼,Ψ and 𝑑Ψ𝛼(𝑓,𝑔)=𝑓𝑔Ψ𝛼. If Ψ(𝑡)=𝑡𝑝, then we recover the usual weighted Bergman space 𝐴𝑝𝛼. One checks that we have the followings inclusions:𝐴1𝛼𝐴Ψ𝛼𝐴𝑝𝛼,(1.12) whenever Ψ is a growth function of order 𝑝.

It seems important to us to mention that a linear operator 𝑇 from 𝐴Ψ1𝛼 to 𝑋 with 𝑋=𝐿Ψ2(𝔹𝑁,𝑣𝛼) or 𝑋=𝐴Ψ2𝛼, where Ψ1 and Ψ2 are two growth functions, is continuous if and only if it maps a bounded set into a bounded set, or equivalently if and only if there exists a constant 𝐶>0 such thatmax𝑇(𝑓)𝛼,Ψ2,𝑇(𝑓)Ψ2𝛼𝐶,(1.13) for any 𝑓𝐴Ψ1𝛼 such that min(𝑓𝛼,Ψ1,𝑓Ψ1𝛼)1. If 𝑋=, then 𝑇 is bounded if and only if |𝑇(𝑓)|𝐶 for 𝑓 as previously.

The next proposition says that the point evaluation functionals are continuous on 𝐴Ψ𝛼.

Proposition 1.6. Let 𝛼>1 and let Ψ be a subharmonic-preserving growth function. For any 𝑎𝔹𝑁 and any 𝑓𝐴Ψ𝛼, one has ||||𝑓(𝑎)Ψ121|𝑎|(𝑁+𝛼+1)𝑓𝛼,Ψ.(1.14)

The proof is the same as that of [11, Proposition 1.9] and so is omitted (still use the hypothesis that Ψ(|𝑓|) is subharmonic). We easily deduce from this and the completeness of 𝐿Ψ the following result.

Corollary 1.7. 𝐴Ψ𝛼, endowed with 𝑑𝛼,Ψ or 𝑑Ψ𝛼, is a complete metric space.

Let 𝑝>0 and 𝛼>1. For 𝑎𝔹𝑁, we introduce the following “test’’ function 𝑓𝑎,𝛼,𝑝 defined by𝑓𝑎,𝛼,𝑝=1|𝑎|2(1𝑧,𝑎)2(𝑁+𝛼+1)/𝑝.(1.15)|𝑓𝑎,𝛼,𝑝|𝑝 is nothing but the Berezin kernel, hence 𝑓𝑎,𝛼,𝑝=((1+|𝑎|)/(1|𝑎|))(𝑁+𝛼+1)/𝑝 while 𝑓𝑎,𝛼,𝑝𝑝=1. Then, as a consequence of Lemma 1.5, we have𝑓𝑎,𝛼,𝑝𝛼,Ψ21|𝑎|(𝑁+𝛼+1)/𝑝1Ψ1(2/1|𝑎|)(𝑁+𝛼+1),(1.16) whenever Ψ is a growth function of order 𝑝. These functions will be of interest to us later, when proving Carleson theorem for large Bergman-Orlicz spaces.

We now define a maximal operator which was introduced in [11] and that will be bounded on 𝐴Ψ𝛼. The definition needs to introduce the sets 𝑄(𝜁,), defined by𝑄(𝜁,)=𝑧𝕊𝑁,||||1𝑧,𝜁<,(1.17) and requires the construction of convenient sets based on the following lemma ([11, Lemma 2.1]; we also refer to the forthcoming Section 2.1).

Lemma 1.8. There exists an integer 𝑀>0 such that for any 0<𝑟<1, one can find a finite sequence {𝜉𝑘}𝑚𝑘=1 (𝑚 depending on 𝑟) in 𝕊𝑁 with the following properties.(1)𝕊𝑁=𝑘𝑄(𝜉𝑘,𝑟).(2)The sets 𝑄(𝜉𝑘,𝑟/4) are mutually disjoint.(3)Each point of 𝕊𝑁 belongs to at most 𝑀 of the sets 𝑄(𝜉𝑘,4𝑟).

From now on, 𝑀 denotes the constant involved in the previous lemma. Let 𝑛0 be an integer and let 𝐶𝑛 be the corona𝐶𝑛=𝑧𝔹𝑁1,12𝑛1|𝑧|<12𝑛+1.(1.18) For any 𝑛0, let (𝜉𝑛,𝑘)𝑘𝕊𝑁 be given by Lemma 1.8 putting 𝑟=1/2𝑛. For 𝑘0, we set𝑇0,𝑘=𝑧𝔹𝑁𝑧{0},𝜉|𝑧|𝑄0,𝑘,1{0}.(1.19) Then we define the sets 𝑇𝑛,𝑘, for 𝑛1 and 𝑘0, by𝑇𝑛,𝑘=𝑧𝔹𝑁𝑧{0},𝜉|𝑧|𝑄𝑛,𝑘,12𝑛.(1.20) We have both𝑛0𝐶𝑛=𝔹𝑁,𝑘0𝑇0,𝑘=𝔹𝑁,𝑘0𝑇𝑛,𝑘=𝔹𝑁{0},𝑛1.(1.21) For (𝑛,𝑘)2, we finally define the subset Δ(𝑛,𝑘) of 𝔹𝑁 by Δ(𝑛,𝑘)=𝐶𝑛𝑇𝑛,𝑘. These sets have good covering properties that we do not recall here (we refer to [11]). Anyway, we define the following maximal function Λ𝑓 for 𝑓𝐴Ψ𝛼(𝔹𝑁) byΛ𝑓=𝑛,𝑘0supΔ(𝑛,𝑘)||||𝜒𝑓(𝑧)Δ(𝑛,𝑘),(1.22) where 𝜒Δ(𝑛,𝑘) is the characteristic function of Δ(𝑛,𝑘). Now we may easily adapt the proof of [11, Proposition 2.2] (which only relies on the subharmonicity of Ψ(|𝑓|)) to get the following.

Proposition 1.9. Let Ψ be a subharmonic-preserving growth function and let 𝛼>1. Then the maximal operator Λ, which carries 𝑓 to Λ𝑓, is bounded from 𝐴Ψ𝛼 to 𝐿Ψ(𝔹𝑁,𝑣𝛼). More precisely, there exists 𝐵1 such that for every 𝑓𝐴Ψ𝛼, one has Λ𝑓𝐿Ψ(𝔹𝑁,𝑣𝛼)𝐵𝑓𝛼,Ψ.(1.23) In particular, a holomorphic function 𝑓 belongs to 𝐴Ψ𝛼 if and only if Λ𝑓 belongs to 𝐿Ψ(𝔹𝑁,𝑣𝛼).

1.2.2. Large Hardy-Orlicz Spaces

Let Ψ still be a (concave) subharmonic-preserving growth function. With the notations of Section 1.2, let (Ω,)=(𝕊𝑁,𝜎). The Hardy-Orlicz space 𝐻Ψ of the ball consists of all holomorphic functions 𝑓 on 𝔹𝑁 such that𝑓𝐻Ψ=sup0<𝑟<1𝑓𝑟Ψ<,(1.24) where 𝑓𝑟(𝑧)=𝑓(𝑟𝑧), and where Ψ is the Luxembourg norm on the Orlicz space 𝐿Ψ(𝕊𝑁,𝜎). Note that we can replace sup0<𝑟<1𝑓𝑟Ψ< by lim𝑟1𝑓𝑟Ψ< thanks to the subharmonicity of Ψ(|𝑓|). Because Ψ is supposed to be a growth function, we have the following inclusion:𝐻1𝐻Ψ𝐻𝑝(1.25) for a growth function Ψ of order 𝑝. In particular, every 𝑓𝐻Ψ admits a boundary radial limit, denoted by 𝑓, 𝜎-almost everywhere on 𝕊𝑁. Let us also note that if Ψ is a growth function, then 𝑓𝐻Ψ if and only if𝑓Ψ𝐻Ψ=sup0<𝑟<1𝑓𝑟Ψ=lim𝑟1𝑓𝑟Ψ<(Inequalities(1.7)and(1.8)).(1.26)

Without possible confusion, we will write Ψ instead of 𝐻Ψ (resp., Ψ instead Ψ𝐻Ψ). As for Bergman-Orlicz spaces, a linear operator 𝑇 from 𝐻Ψ1 to some 𝐿Ψ2(𝕊𝑁,𝜎) or 𝐻Ψ2 where Ψ1 and Ψ2 are two growth functions is continuous (bounded) if and only if there exists a constant 𝐶>0 such thatmax𝑇(𝑓)Ψ2,𝑇(𝑓)Ψ2𝐶(1.27) for any 𝑓𝐻Ψ1 such that min(𝑓Ψ1,𝑓Ψ1)1.

In addition, it is clear that, for any 𝛼>1,𝐴Ψ𝛼𝐻Ψ and 𝑓𝛼,Ψ𝑓Ψ for any 𝑓𝐻Ψ. Therefore, letting 𝛼 tend to −1 in Proposition 1.6, we get the following.

Proposition 1.10. Let Ψ be a subharmonic-preserving growth function. For any 𝑎𝔹𝑁 and any 𝑓𝐻Ψ, one has ||||𝑓(𝑎)Ψ111|𝑎|𝑁𝑓Ψ.(1.28)

As a corollary, we have the following.

Corollary 1.11. Let Ψ be a subharmonic-preserving growth function. 𝐻Ψ is a complete metric space (with the equivalent distances induced by Ψ and Ψ, as usual).

For 𝑎𝔹𝑁 and 𝑝>0, we introduce the “test’’ function 𝑓𝑎,𝑝 defined for any 𝑧𝔹𝑁 by𝑓𝑎,𝑝(𝑧)=1|𝑎|1𝑧,𝑎2𝑁/𝑝.(1.29) It is easily seen that 𝑓𝑎,𝑝𝐻 with 𝑓𝑎,𝑝=1 and that |𝑓𝑎,𝑝(𝑧)|1 for any 𝑧𝕊𝑁. Moreover, let us observe that||𝑓𝑎,𝑝||(𝑧)𝑝=1|𝑎|1+|𝑎|𝑁𝑃𝑎|𝑎|𝑧,,|𝑎|(1.30) so that 𝑓𝑎,𝑝𝑝=((1|𝑎|)/(1+|𝑎|))𝑁/𝑝. Therefore, if Ψ is a growth function of order 𝑝, we have, by Lemma 1.5,𝑓𝑎,𝑝Ψ1Ψ1𝑓1/𝑎,𝑝𝑝1Ψ1(1/(1|𝑎|))𝑁/𝑝.(1.31)

It is very convenient to see 𝐻Ψ as a closed subspace of 𝐿Ψ(𝕊𝑁). When Ψ is an Orlicz function, this is possible thanks to the representation of any function in 𝐻1 by the Poisson integral of its boundary values. This does not work any more in 𝐻𝑝 with 0<𝑝<1, even in this case. However, using a radial maximal function, we can still see 𝐻𝑝 as a subspace of 𝐿𝑝. We are going to extend this to 𝐻Ψ for Ψ a growth function which preserves the subharmonicity. To this purpose, we recall the definition of the nonisotropic distance on 𝔹𝑁: for (𝑧,𝑤)𝔹𝑁,𝑑(𝑧,𝑤)=||||.1𝑧,𝑤(1.32) It is well known that 𝑑 is a distance on 𝕊𝑁 and a pseudodistance on 𝔹𝑁 [21, Paragraph 5.1]. It permits to define the Korányi approach region Γ(𝜁) for 𝜁𝕊𝑁:Γ(𝜁)=𝑧𝔹𝑁,𝑑(𝑧,𝜁)2<1|𝑧|2.(1.33) Then the maximal function 𝑁𝑓 of 𝑓, associated to Korányi approach region, is given by𝑁𝑓(𝜁)=sup𝑧Γ(𝜁)||||𝑓(𝑧)(1.34) for any 𝜁𝕊𝑁. [17, Theorem 1.3] will be very useful.

Theorem 1.12. Let Ψ be a growth function. Then, for any 𝑓𝐻Ψ, 𝕊𝑁Ψ||𝑁𝑓||𝑑𝜎𝑓Ψ.(1.35) In particular, a holomorphic function 𝑓 belongs to 𝐻Ψ if and only if 𝑁𝑓 belongs to 𝐿Ψ(𝕊𝑁,𝜎).

From this theorem, we deduce the following one.

Theorem 1.13. Let Ψ be a subharmonic-preserving growth function. Then for every 𝑓𝐻Ψ, one has(1)lim𝑟1𝕊𝑁Ψ(|𝑓𝑓𝑟|)𝑑𝜎=0;(2)𝑓Ψ=𝑓Ψ𝐻Ψ;(3)𝐻Ψ is separable. More precisely, the polynomials are dense in 𝐻Ψ.

Proof. Let 𝑀rad𝑓(𝜁)=sup0<𝑟<1|𝑓(𝑟𝜁)| for 𝜁𝕊𝑁. Obviously, 𝑀rad𝑓𝑁𝑓, hence 𝕊𝑁Ψ(𝑀rad𝑓)𝑑𝜎𝑓Ψ (Theorem 1.12). Since Ψ is concave and vanishes at 0, we have Ψ(|𝑓𝑓𝑟|)Ψ(|𝑓|)+Ψ(|𝑓𝑟|)2Ψ(𝑀rad𝑓). Now Ψ(|𝑓𝑓𝑟|(𝜁)) tends to 0 as 𝑟 goes to 1 for 𝜎-almost every 𝜁𝕊𝑁. By the dominated convergence theorem, (1) follows.
Then 𝑓Ψ=lim𝑟1𝑓𝑟Ψ and (2) comes from to the subharmonicity of Ψ(|𝑓|).
We proved in (1) that 𝑓𝑟 tends to 𝑓 in 𝐻Ψ for Ψ (hence for Ψ also). We approach every 𝑓𝑟 uniformly on 𝔹𝑁 by its Taylor series to get the third assertion.

2. Carleson Embedding Theorems

2.1. Statements of the Results

For 𝜁𝔹𝑁 and ]0,1], we define the nonisotropic “ball’’ of 𝔹𝑁 by𝑆(𝜁,)=𝑧𝔹𝑁,𝑑(𝜁,𝑧)2<(2.1) and its analogue in 𝔹𝑁 by𝒮(𝜁,)=𝑧𝔹𝑁,𝑑(𝜁,𝑧)2<.(2.2) Let us also denote by𝑄(𝜁,)=𝒮(𝜁,)𝕊𝑁,(2.3) the “true’’ balls in 𝕊𝑁. We have 𝜎(𝑄(𝜁,))𝑁 and 𝑣𝛼(𝑆(𝜁,))𝑁+𝛼+1 [22].

Let 𝜇 be a positive Borel measure on 𝔹𝑁 whose restriction to 𝕊𝑁 is absolutely continuous with respect to 𝜎 and let 0<𝑝𝑞<. By definition, 𝜇 is a (𝑞/𝑝)-Carleson measure if 𝜇(𝒮(𝜁,))𝐶𝑁𝑞/𝑝, while it is a vanishing (𝑞/𝑝)-Carleson measure if 𝜇(𝒮(𝜁,))=𝑜(𝑁𝑞/𝑝) when goes to 0. A variant of the well-known Carleson theorem for Hardy spaces [1, 3] ensures that the embedding 𝐻𝑝𝐿𝑞(𝔹𝑁,𝜇) is bounded (resp., compact) if and only if 𝜇 is a (𝑞/𝑝)-Carleson measure (resp., a vanishing (𝑞/𝑝)-Carleson measure).

Similarly, we define the (𝛼,(𝑞/𝑝))-Bergman Carleson measures (resp., vanishing (𝛼,(𝑞/𝑝))-Bergman Carleson measures) for weighted Bergman spaces by 𝜇(𝑆(𝜁,))𝐶(𝑁+𝛼+1)𝑞/𝑝 (resp., 𝜇(𝑆(𝜁,))=𝑜((𝑁+𝛼+1)𝑞/𝑝)). When 𝑝=𝑞, we just speak about 𝛼-Bergman Carleson measures (resp., vanishing 𝛼-Bergman Carleson measures). Ueki [7] showed that 𝐴𝑝𝛼𝐿𝑞(𝔹𝑁,𝜇) is bounded (resp., compact) if and only if 𝜇 is a (𝛼,(𝑞/𝑝))-Bergman Carleson measure (resp., a vanishing (𝛼,(𝑞/𝑝))-Carleson measure).

In the context of Hardy-Orlicz spaces (resp., weighted Bergman-Orlicz spaces) smaller than 𝐻1 (resp., 𝐴1𝛼) (i.e., when the defining function Ψ is an Orlicz function), much general results were obtained in [9, 10] in the unit disc, and in [11, 12] in the unit ball.

For Hardy-Orlicz (resp., weighted Bergman-Orlicz) spaces larger than 𝐻1 (resp., 𝐴1𝛼), we state that the characterizations of the boundedness and compactness of 𝐻Ψ1𝐿Ψ2(𝔹𝑁,𝜇) (resp., 𝐴Ψ1𝛼𝐿Ψ2(𝔹𝑁,𝜇)), where Ψ1 and Ψ2 are two growth functions such that 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing at least for large values of 𝑥 (or equivalently 𝑥Ψ2Ψ11(𝑥)/𝑥 nondecreasing, since Ψ1 is increasing and does vanish except in 0), only depend on the growth of Ψ2Ψ11 at infinity.

Note that if Ψ𝑖(𝑥)=𝑥𝑝𝑖, Ψ𝑖=Φ𝑝𝑖 (with Φ an Orlicz function), or Ψ𝑖(𝑥)=𝑥𝑝𝑖log𝑞𝑖(𝐶+𝑥) with 𝑝2𝑝1, then 𝑥Ψ2Ψ11(𝑥)/𝑥 is nondecreasing.

Theorem 2.1. Let Ψ1 and Ψ2 be two subharmonic-preserving growth functions such that 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing. Let 𝜇 be a finite positive Borel measure on 𝔹𝑁 (whose restriction to 𝕊𝑁 is absolutely continuous with respect to 𝜎). Then,(1)𝐻Ψ1 embeds into 𝐿Ψ2(𝔹𝑁,𝜇) if and only if there exists some 0(0,1) such that, for any (0,0), 1𝜇(𝒮(𝜁,))Ψ2Ψ111/𝑁(2.4) uniformly in 𝜁𝕊𝑁,(2)the embedding 𝐻Ψ1𝐿Ψ2(𝔹𝑁,𝜇) is compact if and only if 𝜇(𝒮(𝜁,))=𝑜01Ψ2Ψ111/𝑁(2.5) uniformly in 𝜁𝕊𝑁.A measure 𝜇 which satisfies (2.4) (resp., (2.5)) will be called a (Ψ1,Ψ2)-Carleson measure (resp., a vanishing (Ψ1,Ψ2)-Carleson measure).

For big Bergman-Orlicz spaces, we have the following.

Theorem 2.2. Let Ψ1 and Ψ2 be two subharmonic-preserving growth functions such that 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing, and let 𝛼>1. Let also 𝜇 be a finite positive Borel measure on 𝔹𝑁. Then,(1)𝐴Ψ1𝛼 embeds into 𝐿Ψ2(𝔹𝑁,𝜇) if and only if there exists 0(0,1) such that, for any (0,0), 1𝜇(𝑆(𝜁,))Ψ2Ψ111/𝑁+𝛼+1,(2.6) uniformly in 𝜁𝕊𝑁,(2)the embedding 𝐴Ψ1𝛼𝐿Ψ2(𝔹𝑁,𝜇) is compact if and only if 𝜇(𝑆(𝜁,))𝑜01Ψ2Ψ111/𝑁+𝛼+1,(2.7) uniformly in 𝜁𝕊𝑁.A measure 𝜇 which satisfies (2.6) (resp., (2.7)) will be called a (𝛼,Ψ1,Ψ2)-Bergman-Carleson measure (resp., a vanishing (𝛼,Ψ1,Ψ2)-Bergman-Carleson measure).

Remark 2.3. By the closed graph theorem, the above embeddings are bounded as soon as they exist.

We immediately deduce from the previous theorems the following corollaries.

Corollary 2.4. Let 𝛼>1 and let 0<𝑝𝑞<1 and Ψ1=Φ𝑝, Ψ2=Φ𝑞 as in (3) of Example 1.3. Let 𝜇 be a finite positive Borel measure on 𝔹𝑁 (whose restriction to 𝕊𝑁 is absolutely continuous with respect to 𝜎) (resp., on 𝔹𝑁). Then,(1)𝐻Φ𝑝 (resp., 𝐴Φ𝑝𝛼) embeds into 𝐿Φ𝑞(𝔹𝑁,𝜇) (resp., 𝐿Φ𝑞(𝔹𝑁,𝜇)) if and only if 𝜇 is a (𝑞/𝑝)-Carleson measure (resp., a (𝛼,(𝑞/𝑝))-Bergman-Carleson measure),(2)the embedding 𝐻Φ𝑝𝐿Φ𝑞(𝔹𝑁,𝜇) (resp., 𝐴Φ𝑝𝛼𝐿Φ𝑞(𝔹𝑁,𝜇)) is compact if and only if 𝜇 is a vanishing (𝑞/𝑝)-Carleson measure (resp., a vanishing (𝛼,(𝑞/𝑝)) Bergman-Carleson measure).

If Ψ1=Ψ2 (or equivalently if Ψ1 and Ψ2 are equivalent), we have the following.

Corollary 2.5. Let 𝛼>1, and let Ψ be subharmonic-preserving growth function. Let 𝜇 be a finite positive Borel measure on 𝔹𝑁 (whose restriction to 𝕊𝑁 is absolutely continuous with respect to 𝜎) (resp., on 𝔹𝑁). Then,(1)𝐻Ψ (resp., 𝐴Ψ𝛼) embeds into 𝐿Ψ(𝔹𝑁,𝜇) (resp., 𝐿Ψ(𝔹𝑁,𝜇)) if and only if 𝜇 is a Carleson measure (resp., a 𝛼-Bergman-Carleson measure),(2)the embedding 𝐻Ψ𝐿Ψ(𝔹𝑁,𝜇) (resp., 𝐴Ψ𝛼𝐿Ψ(𝔹𝑁,𝜇)) is compact if and only if 𝜇 is a vanishing Carleson measure (resp., a vanishing 𝛼-Bergman-Carleson measure).

2.2. Proofs of Theorems 2.1 and 2.2

For the compactness parts, we will use a criterion given in [12, Proposition 2.11] and [11, Proposition 2.8]. Its proof is easy to adapt as soon as we have checked that the convergence in 𝐻Ψ (resp., 𝐴Ψ𝛼), for Ψ a subharmonic-preserving growth function, implies the convergence on every compact subset of 𝔹𝑁, but it stems from Proposition 1.10 (resp., Proposition 1.6).

Proposition 2.6. Let 𝛼>1, let Ψ1 and Ψ2 be two subharmonic-preserving growth functions and let 𝜇 (resp., 𝜇) be a finite positive Borel measure on 𝔹𝑁 (resp., 𝔹𝑁) whose restriction to 𝕊𝑁 is absolutely continuous with respect to 𝜎. One assume that 𝑗𝜇𝐻Ψ1𝐿Ψ2(𝔹𝑁,𝜇) (resp., 𝑗𝜇𝛼𝐴Ψ1𝛼𝐿Ψ2(𝔹𝑁,𝜇)) is well defined (hence bounded).(1)The two following assertions are equivalent:(a)the canonical embedding 𝑗𝜇 (resp., 𝑗𝜇𝛼) is compact;(b)every sequence in the unit ball of 𝐻Ψ1 (resp., 𝐴Ψ1𝛼), which is convergent to 0 uniformly on every compact subset of 𝔹𝑁, is convergent to 0 in 𝐿Ψ2(𝜇) (resp., 𝐿Ψ2(𝜇)).(2)If lim𝑟1𝐼𝑟=0 (resp., lim𝑟1𝐼𝑟,𝛼=0), where 𝐼𝑟(𝑓)=𝑓𝜒𝔹𝑁𝑟𝔹𝑁 (resp., 𝐼𝑟(𝑓)=𝑓𝜒𝔹𝑁𝑟𝔹𝑁), then the canonical embedding 𝑗𝜇 (resp., 𝑗𝜇𝛼) is compact.

2.2.1. Proof of Theorem 2.1

We assume that the hypothesis of Theorem 2.1 is fulfilled. The proof will be based on two lemmas, whose proofs follow that of Theorem 2.4 and Lemma 2.6 of [12]. These results are refinement of Carleson theorem and are the key to deal with different Hardy-Orlicz spaces which are not classical Hardy spaces. We need to introduce the function 𝐾𝜇, associated to 𝜇 by𝐾𝜇()=sup0<𝑡<sup𝜁𝕊𝑁𝜇(𝒮(𝜁,𝑡))𝑡𝑁,for(0,1),(2.8) so that 𝜇 is a Carleson measure if and only if 𝐾𝜇() is finite for some (0,1).

Theorem 2.7 (see [12, Theorem 2.4]). There exist two constants 𝐶>0 and 𝐶>1 such that, for every 𝑓 continuous on 𝔹𝑁 which admits boundary values almost everywhere on 𝕊𝑁, and every finite positive Borel measure 𝜇 on 𝔹𝑁, one has 𝜇𝑧𝔹𝑁||||,|𝑧|>1,𝑓(𝑧)>𝑡𝐶𝐾𝜇𝑁(𝐶)𝜎𝑓>𝑡,(2.9) for every (0,1/𝐶) and every 𝑡>0.

The next lemma is a the technical key to obtain Carleson theorems in Hardy-Orlicz context. Its proof closely follows that of [12, Lemma 2.6], but we prefer to give the details.

Lemma 2.8. Let 𝜇 be a finite positive Borel measure on 𝔹𝑁 and let Ψ1 and Ψ2 be two subharmonic-preserving growth functions. Let 𝐶1 be the constant appearing in Theorem 2.7. Assume that there exist 𝐴>0,𝜂>0 and 𝐴(0,1/𝐶) such that 𝐾𝜇()𝜂1/𝑁Ψ2𝐴Ψ111/𝑁,(2.10) for every (0,𝐴). Then, for every 𝑓𝐻Ψ1(𝔹𝑁) such that 𝑓Ψ11 and every Borel subset 𝐸 of 𝔹𝑁, 𝐸Ψ2||𝑓||𝑑𝜇𝜇(𝐸)+𝜂𝕊𝑁Ψ1𝑁𝑓𝑑𝜎,(2.11) where involves a constant which is independent of 𝑓, 𝜂, and 𝐸.

Proof. Let 𝑓𝐻Ψ1(𝔹𝑁) with 𝑓Ψ11. With the notations of the statement of the lemma and using Proposition 1.10, the proof of [12, Lemma 2.6] directly yields. 𝐸Ψ2𝐴2𝑁+1(𝐶+1)𝐶𝑁1||𝑓||𝑑𝜇𝑥𝐴0Ψ2+𝜂𝐶(𝑠)𝜇(𝐸)𝑑𝑠𝐶𝑁𝑥𝐴Ψ2Ψ(𝑠)1(𝐶+1)𝐶𝑁1𝑆/𝐴Ψ2𝑁(((𝐶+1)/𝐶)𝑆)×𝜎𝑓>2𝑁+1(𝐶+1)𝐶𝑁1𝐴𝑠𝑑𝑠,(2.12) for some constant 𝑥𝐴 which depends only on 𝐴. Now, since Ψ1 and Ψ2 are two concave growth functions, we have Ψ𝑖(𝑠)Ψ𝑖(𝐾𝑠)Ψ𝑖(𝑠),(2.13) where involve constants which only depend on 𝐾 and Ψ𝑖. Therefore, (2.12) becomes 𝐸Ψ2||𝑓||𝑑𝜇𝜇(𝐸)+𝜂𝑥𝐴Ψ2Ψ(𝑠)1(𝑠)Ψ2𝜎𝑁(𝑠)𝑓>2𝑁+1(𝐶+1)𝐶𝑁1𝐴𝑠𝑑𝑠.(2.14) Since 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing and since Ψ1 and Ψ2 and their derivatives do not vanish except in 0, we have, for any 𝑠>0, Ψ1(𝑠)Ψ2Ψ(𝑠)1(𝑠)Ψ2(𝑠).(2.15) It follows that 𝐸Ψ2||𝑓||𝑑𝜇𝜇(𝐸)+𝜂𝑥𝐴Ψ1𝑁(𝑠)𝜎𝑓>2𝑁+1(𝐶+1)𝐶𝑁1𝐴𝑠𝑑𝑠𝜇(𝐸)+𝜂0+Ψ1𝐴𝑁𝑓2𝑁+1(𝐶+1)𝐶𝑁1>𝑠𝑑𝑠=𝜇(𝐸)+𝜂𝕊𝑁𝜓1𝐴𝑁𝑓2𝑁+1(𝐶+1)𝐶𝑁1𝑑𝜎𝜇(𝐸)+𝜂𝕊𝑁𝜓1𝑁𝑓𝑑𝜎.(2.16)

We now can prove Theorem 2.1.

Proof of Theorem 2.1. (1) We assume that 𝜇 is a (Ψ1,Ψ2)-Carleson measure, and we intend to show that 𝑗𝜇𝐻Ψ1𝐿Ψ2(𝔹𝑁,𝜇) is well defined (hence bounded). We observe that (2.10) is satisfied for 𝐴=𝜂=1. Indeed we have, for any (0,0), 𝜇(𝒮(𝜁,))𝑁1/𝑁Ψ2Ψ111/𝑁,(2.17) uniformly in 𝜁𝕊𝑁. Now, 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing (at least for large 𝑥) so that we can find 1(0,min(1/𝐶,0)) such that, for any (0,1), 𝐾𝜇()1/𝑁Ψ2Ψ111/𝑁.(2.18) Therefore, we may and will apply Lemma 2.8 with 𝐸=𝔹𝑁 to get 𝔹𝑁Ψ2||𝑓||𝑑𝜇𝜇(𝐸)+𝕊NΨ1𝑁𝑓𝑑𝜎.(2.19) We finish the proof of (1) using Theorem 1.12 and Inequality (1.7): 𝔹𝑁Ψ2||𝑓||𝑑𝜇𝜇(𝐸)+𝑓Ψ1.(2.20)
For the converse, let 𝑔𝑎,𝑝=Φ1((1/(1|𝑎|))𝑁/𝑝)𝑓𝑎,𝑝, where 𝑓𝑎,𝑝 is the test function introduced at the end of Section 1.2.2, with 𝑝 such that Ψ1 is a growth function of order 𝑝. According to (1.31), 𝑔𝑎,𝑝 lies in the unit ball of 𝐻Ψ1. Let 𝐶 be the (finite) norm of 𝑗𝜇. Using a classical computation which gives that |1𝑧,𝑎|2(1|𝑎|) for any 𝑧𝒮(𝑎/|𝑎|,1|𝑎|), we get |𝑓𝑎,𝑝(𝑧)|1/4𝑁/𝑝 for such 𝑧, hence 1𝔹𝑁Ψ2||𝑔𝑎,𝑝||𝐶𝒮𝑎𝑑𝜇𝜇Ψ|𝑎|,1|𝑎|214𝑁/𝑝𝐶Ψ1111|𝑎|𝑁/𝑝(2.21) for any 𝑎𝔹𝑁. We may assume that 𝐶1 and since Ψ2 is concave and vanishes at 0, it follows that 𝜇𝒮𝑎|𝑎|,1|𝑎|4𝑁/𝑝𝐶1Ψ2Ψ111/(1|𝑎|)𝑁.(2.22)
(2) We turn to the compactness part. We first prove the sufficient part. Let us assume that 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing and that 𝜇 is a vanishing (Ψ1,Ψ2)-Carleson measure. In particular, 𝜇 is a (Ψ1,Ψ2)-Carleson measure and we may apply Proposition 2.6. Then it is sufficient to prove that for every 𝜀>0, there exists 𝑟 close enough to 1, such that 𝐼𝑟<𝜀 (where 𝐼𝑟𝐻Ψ1𝐿Ψ2(𝔹𝑁𝑟𝔹𝑁,𝜇)). We fix 𝜀>0 and let 𝑓 be in the unit ball of 𝐻Ψ2. As in the proof of the sufficient part of (1), since 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing, 𝜇 is a vanishing (Ψ1,Ψ2)-Carleson measure implying that 𝐾𝜇()𝜀1/𝑁Ψ2Ψ111/𝑁,(2.23) for small enough. Then, we apply Lemma 2.8 with 𝜂=𝜀 and 𝐸=𝔹𝑁𝑟𝔹𝑁 to get the existence of 𝑟 close enough to 1 (independent of 𝑓), such that 𝐼𝑟(𝑓)Ψ2𝜇𝔹𝑁𝑟𝔹𝑁+𝜀𝕊𝑁Ψ1𝑁𝑓𝑑𝜎𝜇𝔹𝑁𝑟𝔹𝑁+𝜀𝑓Ψ1,(2.24) for any 𝜀>0. Now, we may argue as in [12, Lemma 2.13] to prove that, under the assumption that 𝜇 is a (Ψ1,Ψ2)-Carleson measure, 𝜇(𝕊𝑁)=0. Hence limsup𝑟1𝐼𝑟(𝑓)Ψ2𝜀𝑓Ψ1,(2.25) for any 𝜀>0, which gives (2).
For the converse, we assume that 𝑗𝜇 is compact but that 𝜇 is not a vanishing (Ψ1,Ψ2)-Carleson measure, that is there exist 𝜀0>0, a sequence (𝑛)𝑛 decreasing to 0 and a sequence (𝜁𝑛)𝑛𝕊𝑁 such that 𝜇𝒮𝜁𝑛,𝑛𝜀01Ψ2Ψ111/𝑁𝑛.(2.26) Now, let 𝑔𝑎𝑛,𝑝=Φ1((1/(1|𝑎𝑛|))𝑁/𝑝)𝑓𝑎𝑛,𝑝 be as in the proof of the necessity in the boundedness part with 𝑎𝑛=(1𝑛)𝜁𝑛 and 𝑝 such that Ψ1 is growth function of order 𝑝. Since (𝑔𝑎𝑛,𝑝)𝑛 is bounded in 𝐻Ψ1 and converges uniformly on every compact subset of 𝔹𝑁, then we must have 𝑗𝜇(𝑔𝑎𝑛,𝑝)Ψ2𝑛0, because of Proposition 2.6. But, for 𝑧𝒮(𝜁𝑛,𝑛), we saw that ||𝑔𝑎𝑛,𝑝||1(𝑧)4𝑁/𝑝Ψ111𝑁𝑛.(2.27) Since Ψ2 is a growth function of order, let say 𝑞, we have, for any 𝑡(0,1), Ψ2(𝑦/𝑡)Ψ2(𝑦)/𝑡𝑞. Hence, we get 𝔹𝑁Ψ24𝑁𝑞/𝑝||𝑔𝑎𝑛,𝑝||𝜀01/𝑞𝑑𝜇Ψ21𝜀01/𝑞Ψ111/𝑁𝑛𝜇𝒮𝜁𝑛,𝑛Ψ2Ψ111/𝑁𝑛𝜀0𝜀01Ψ2Ψ111/𝑁𝑛=1.(2.28) Therefore, 𝑗𝜇(𝑔𝑎𝑛,𝑝)Ψ2𝜀01/𝑞/4𝑁𝑞/𝑝, which is a contradiction.

2.2.2. Proof of Theorem 2.2

We introduce the following function:𝐾𝜇,𝛼()=sup0<𝑡<sup𝜁𝕊𝑁𝜇(𝑆(𝜁,𝑡))𝑡𝑁+𝛼+1.(2.29) As for Hardy-Orlicz spaces, we need a refinement of Carleson theorem which allows to deal with weighted Bergman-Orlicz spaces associated to different growth functions. It is [11, Theorem 2.3].

Theorem 2.9. There exists a constant 𝐶>0 such that, for every 𝑓 continuous on 𝔹𝑁 and every positive finite Borel measure 𝜇 on 𝔹𝑁, one has 𝜇𝑧𝔹𝑁||||,|𝑧|>1,𝑓(𝑧)>𝑡𝐶𝐾𝜇,𝛼(2)𝑣𝛼Λ𝑓,>𝑡(2.30) for every (0,1/2) and every 𝑡>0.

Our Carleson theorem for large weighted Bergman-Orlicz spaces is a consequence of the next technical lemma. Its proof is an easy adaptation and combination of the proofs of Lemma 2.8 and [11, Lemma 2.4], so we omit it.

Lemma 2.10. Let 𝜇 be a finite positive Borel measure on 𝔹𝑁 and let Ψ1 and Ψ2 be two subharmonic-preserving growth functions. Assume that there exist 𝐴>0, 𝜂>0, and 𝐴(0,1/2) such that 𝐾𝜇,𝛼()𝜂1/𝑁+1+𝛼𝜓2𝐴Ψ111/𝑁+1+𝛼,(2.31) for every (0,𝐴). Then, for every 𝑓𝐴Ψ1𝛼(𝔹𝑁) with 𝑓𝐴Ψ1𝛼1 and every Borel subset 𝐸 of 𝔹𝑁, one has 𝐸Ψ2||𝑓||𝑑𝜇𝜇(𝐸)+𝜂𝔹𝑁Ψ1Λ𝑓𝑑𝑣𝛼,(2.32) where involves constants which are independent of 𝑓, 𝜂, and 𝐸.

We now prove Theorem 2.2.

Proof of Theorem 2.2. Let Ψ1 and Ψ2 be two subharmonic-preserving growth functions such that 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing.
(1) We first assume that 𝜇 is a (𝛼,Ψ1,Ψ2)-Bergman Carleson measure, which implies, as in the beginning of the proof of (1) of Theorem 2.1, that condition (2.31) with 𝐴=𝜂=1 is fulfilled. Applying (2.31) to 𝑓, with 𝐸=𝔹𝑁, gives 𝔹𝑁Ψ2||𝑓||𝑑𝜇𝜇(𝐸)+𝔹𝑁Ψ1Λ𝑓𝑑𝑣𝛼.(2.33) We conclude the proof using Proposition 1.9 together with (1.7).
To prove the converse, we argue as in the proof of the corresponding part of Theorem 2.1, using the test function 𝑔𝑎,𝛼,𝑝=Φ1((1/(1|𝑎|))(𝑁+𝛼+1)/𝑝)𝑓𝑎,𝛼,𝑝, where 𝑓𝑎,𝛼,𝑝 has been introduced in Section 1.2.1, with 𝑝 such that Ψ1 is a growth function of order 𝑝.
(2) The proof of the compactness part of Theorem 2.2 is still similar to that for Hardy-Orlicz spaces: we apply Lemma 2.10 to 𝑓𝐴Ψ1𝛼, with 𝜂=𝜀 and 𝐸=𝔹𝑁𝑟𝔹𝑁 to show that 𝑓.𝜒𝔹𝑁𝑟𝔹𝑁Ψ2 tends to 0, as 𝑟 tends to 1. Proposition 2.6 then shows that 𝐴Ψ1𝛼𝐿Ψ2(𝜇) is compact.
For the converse, we procede as for Theorem 2.1 using the test function introduced in the boundedness part above.

3. Applications to Composition Operators

Theorems 2.1 and 2.2 will be used to characterize both boundedness and compactness of composition operators on 𝐻Ψ and 𝐴Ψ𝛼. As usual, for 𝜙𝔹𝑁𝔹𝑁 holomorphic, 𝐶𝜙 will be seen as an embedding operator. To do so, we define the pullback measures 𝜇𝜙 and 𝜇𝜙,𝛼 of, respectively, 𝜎 and 𝑣𝛼 under 𝜙:𝜇𝜙𝜙(𝐸)=𝜎1(𝐸)𝕊𝑁,(3.1) where 𝜙 is the radial limit almost everywhere of 𝜙 and 𝐸 is any Borel subset of 𝔹𝑁. And𝜇𝜙,𝛼(𝐸)=𝑣𝛼𝜙1(𝐸)(3.2) for any Borel subset 𝐸 of 𝔹𝑁.

By a classical formula for pull-back measures,𝑗𝜇𝜙(𝑓)Ψ=𝐶𝜙(𝑓)Ψ𝑗resp.,𝜇𝜙,𝛼(𝑓)𝛼,Ψ=𝐶𝜙(𝑓)𝛼,Ψ(3.3) for any growth function Ψ and every polynomial 𝑓 (resp., for any 𝑓𝐴Ψ𝛼). Then we extend this equality to the whole space 𝐻Ψ by density of polynomials in 𝐻Ψ (Theorem 1.13) and using Cauchy’s formula.

Remark 3.1. Observe that it is not true that 𝑗𝜇𝜙(𝑓)=𝐶𝜙(𝑓) for any 𝑓𝐻Ψ but it is for any polynomial 𝑓. Such an equality holds for every function in 𝐴Ψ𝛼.

Moreover, we have the following criterion for compactness of composition operators.

Proposition 3.2. Let 𝛼>1 and let Ψ1 and Ψ2 be as in Proposition 2.6. Let also 𝜙𝔹𝑁𝔹𝑁 be holomorphic. 𝐶𝜙 is compact from 𝐻Ψ1 into 𝐻Ψ2 (resp., 𝐴Ψ1𝛼 into 𝐴Ψ2𝛼) if and only if, for every bounded sequence (𝑓𝑛)𝑛𝐻Ψ1 (resp., (𝑓𝑛)𝑛𝐴Ψ1𝛼) which converges to 0 on every compact of 𝔹𝑁,𝐶𝜙(𝑓)Ψ2 (resp., 𝐶𝜙(𝑓)Ψ2,𝛼) tends to 0.

The proof of this proposition is quite similar to that of Proposition 2.6.

Therefore, Theorems 2.1 and 2.2 together with Propositions 2.6 and 3.2 give the following characterizations of the boundedness and compactness of 𝐶𝜙 on 𝐻Ψ and 𝐴Ψ𝛼.

Theorem 3.3. Let Ψ1 and Ψ2 be two subharmonic-preserving growth functions such that 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing. Let also 𝜙𝔹𝑁𝔹𝑁 be holomorphic. Then,(1)𝐶𝜙 is bounded from 𝐻Ψ1 into 𝐻Ψ2 if and only if 𝜇𝜙 is a (Ψ1,Ψ2)-Carleson measure,(2)𝐶𝜙 is compact from 𝐻Ψ1 into 𝐻Ψ2 if and only if 𝜇𝜙 is a vanishing (Ψ1,Ψ2)-Carleson measure.

For Bergman-Orlicz spaces, we have the following.

Theorem 3.4. Let 𝛼>1 and let Ψ1 and Ψ2 be two subharmonic-preserving growth functions such that 𝑥Ψ2(𝑥)/Ψ1(𝑥) is nondecreasing. Let also 𝜙𝔹𝑁𝔹𝑁 be holomorphic. Then,(1)𝐶𝜙 is bounded from 𝐴Ψ1𝛼 into 𝐴Ψ2𝛼 if and only if 𝜇𝜙,𝛼 is a (𝛼,Ψ1,Ψ2)-Bergman Carleson measure,(2)𝐶𝜙 is compact from 𝐴Ψ1𝛼 into 𝐴Ψ2𝛼 if and only if 𝜇𝜙,𝛼 is a vanishing (𝛼,Ψ1,Ψ2)-Bergman Carleson measure.

We immediately deduce the following corollary, which reminds a similar result when the function Ψ is an Orlicz function satisfying the Δ2-Condition (see [11, 12]).

Corollary 3.5. Let 𝛼>1 and let Ψ be a subharmonic preserving growth function. Let also 𝜙𝔹𝑁𝔹𝑁 be holomorphic. Then,(1)C𝜙 is bounded (resp., compact) on 𝐻Ψ if and only if it is bounded (resp., compact) on one (or equivalently every) 𝐻𝑝, if and only if 𝜇𝜙 is a Carleson measure (resp., a vanishing Carleson measure),(2)𝐶𝜙 is bounded (resp., compact) on 𝐴Ψ𝛼 if and only if it is bounded (resp., compact) on one (or equivalently every) 𝐴𝑝𝛼, if and only if 𝜇𝜙 is a Bergman-Carleson measure (resp., a vanishing Bergman-Carleson measure).

The following corollary shows that the behavior of composition operators between certain classes of different Hardy-Orlicz or Bergman-Orlicz spaces is still the same as that in the classical cases (see [23]).

Corollary 3.6. Let 𝛼>1, let 0<𝑝𝑞< and Ψ1=Φ𝑝, and Ψ2=Φ𝑞 as in (3) of Example 1.3. Let also 𝜙𝔹𝑁𝔹𝑁 be holomorphic. Then,(1)𝐶𝜙 is bounded (resp., compact) from 𝐻Ψ1 to 𝐻Ψ2 if and only if it is bounded (resp., compact) from 𝐻𝑝 to 𝐻𝑞, if and only if 𝜇𝜙 is a 𝑞/𝑝-Carleson measure (resp., a vanishing 𝑞/𝑝-Carleson measure),(2)𝐶𝜙 is bounded (resp., compact) from 𝐴Ψ1𝛼 to 𝐴Ψ2𝛼 if and only if it is bounded (resp., compact) from 𝐴𝑝𝛼 to 𝐴𝑞𝛼, if and only if 𝜇𝜙 is a 𝑞/𝑝-Bergman-Carleson measure (resp., a vanishing 𝑞/𝑝-Bergman-Carleson measure).

Acknowledgments

The authors would like to thank Ueki for providing his paper [7]. The second author acknowledges support from the Irish Research Council for Science, Engineering and Technology.