INTRODUCTION
When tasked with creating a mathematical model of a given scenario, one approach is to interpret the input data and output solutions as measurable functions, which belong to some function spaces, and the assignment of data to a solution as an operator between these spaces. We then obtain the abstract model
where
and
are function spaces containing the data and solutions, respectively, and
is a linear operator mapping the data to its solution. For example, in economics, the problem of increasing one's capital in the stock market can be modeled as the relation above, where
is the behavior of the market,
are possible decisions of a trader (purchases or sales), and
maps the current state of the market to the best possible decision. Numerous examples of such modeling in physics, biology, and other fields follow naturally.
One of the most important questions when working with such models is, what can we say about the quality of the solutions based on the quality of the input data. This is the motivation behind using different function spaces, as we group together functions with the same quality into one space. At this point, it is natural to ask whether we may find an optimal space for a given problem. There are two forms of such a task: either we are given data and search for the smallest group of possible solutions (= given space and finding the smallest space ), or we are given conditions on the solutions and search for the largest amount of data whose solutions satisfy such demands (= given space and finding the largest space ).
When solving optimality problems, it is important to remember that we also have to consider the properties of the chosen classes of function spaces in terms of accessibility and expressivity. For instance, Lebesgue spaces are a very well understood and accessible class of spaces (as they are described by a single parameter), though the spaces may be too sparse to provide accurate enough information, and are as such not expressive enough for certain needs. On the other side of the spectrum sit rearrangement-invariant spaces, where an optimal space virtually always exists, however, the optimal space is described implicitly, and is as such practically impossible to work with. A good middle ground is provided by Orlicz spaces, a class of function spaces described by Young functions. As such, these spaces provide both accessibility and expressivity.
Optimality problems are not a new discipline—the earlier results go back into the second half of the 20th century, but it was not until the turn of the millennium that they saw a boom in interest. As such, the field is now supported by a vast amount of literature, which includes the works [1, 3, 4, 6, 7, 9, 11, 12, 17, 21, 22].
We focus on the optimal form of Sobolev embeddings within rearrangement-invariant spaces, and within Orlicz spaces, on Maz'ya classes of Euclidean domains. Using the general theory we introduce, we prove the nonexistence of certain optimal Orlicz spaces in Orlicz–Sobolev embeddings, namely that there is no largest domain Orlicz space in the embedding
for appropriate values of the parameters involved.
The motivation for studying embeddings into this particular space stems from the fact that it is the optimal (smallest) rearrangement-invariant space which renders the embedding true for every in the very important case when the corresponding domain Orlicz space is the critical (limiting) Lebesgue space . This was first observed by [3] in connection with the special case , that is, , where is the dimension of the underlying ambient Euclidean space and is the order of differentiation. There are various reasons for establishing results involving such Lorentz-type refinements on the target side, perhaps the most notable one being the fact that when such embeddings are considered, then no loss of information occurs under their iterations (for more details, see [5], and references therein).
We shall now describe our principal results in detail. We first establish a general formula for the fundamental function of an operator-induced space based on the isoperimetric behavior of the underlying domain under certain mild assumptions. Next, we consider a specific situation concerning Maz'ya domains. We then establish a theorem which enables us to transfer the information of nonexistence of an optimal Orlicz domain partner space from a Marcinkiewicz space to any space on the same fundamental level. We finally apply this general scheme to a limiting Sobolev embedding. We thereby obtain a wide variety of results applicable to any Maz'ya domain having isoperimetric profile with whose particular case for Lipschitz domains was obtained in the earlier work [16].
The text is structured as follows. In Chapter 1, we present definitions and basic knowledge about the relevant function spaces and the isoperimetric function. In Chapter 2, we collect background results necessary for proofs of our main results, namely the principal alternative and reduction principle for Sobolev embeddings, and the forms of optimal r.i. domain and target spaces. Lastly, in Chapter 3, we present the main results of this work.
1 PRELIMINARIES
1.1 Function spaces
In this section, we recall some definitions and basic facts from the theory of various function spaces. For further details, the standard reference is [2].
Let , . In this work, denotes the -dimensional Lebesgue measure of for measurable. We use the convention that and .
Let
be a bounded open set. We assume, without loss of generality, that
, and define
and
The
distribution function of a function
is defined as
and the
non-increasing rearrangement of a function
is defined as
The operation
is monotone in the sense that for
,
The
elementary maximal function of a function
is defined as
The operation
is subadditive in the sense that for
,
The
Hardy–Littlewood inequality is a classical property of function rearrangements, which asserts that, for
,
() A specialization of the inequality states that for every
and for every
measurable,
Next, we define the rearrangement-invariant norm. We say that a functional
is a
function norm, if for all
and
in
, and every
, the following properties hold:
- (P1)
a.e.,
,
;
- (P2) a.e. ;
- (P3) a.e. ;
- (P4) ;
- (P5) for some constant independent of .
If, in addition, the property
- (P6) if
holds, we call the functional
a
rearrangement-invariant function norm. For any such rearrangement-invariant function norm
, we define the functional
as
for
. The functional
is then also a rearrangement-invariant function norm, see [
2, Chapter 1, Theorem 2.2], and it is called the
associate function norm of
and, by [
2, Chapter 1, Theorem 2.7] it holds that
. We say that
is a
rearrangement-invariant function quasinorm, if it satisfies the conditions (P2), (P3), (P4), and (P6), and (Q1), a weaker version of (P1), where
- (Q1)
a.e.,
,
there exists ,
for all
and every
.
Given a rearrangement-invariant function norm
, we define the functional
as
and we call the set
a
rearrangement-invariant space. Furthermore, the space
is called the
representation space of
, and we define the
associate space of
as
Then, the
Hölder inequality
holds for every
and
.
For any rearrangement-invariant spaces
and
, the continuous embedding of
into
is denoted by
and means that there exists a constant
such that for any
, it holds that
and
. By [
2, Chapter 1, Proposition 2.10] it holds that
and by [
2, Chapter 1, Theorem 1.8], it holds that
Note that the functional may also be defined if its corresponding functional is only a rearrangement-invariant quasinorm. However, some of the properties listed here for the case, where is a rearrangement-invariant norm then do not necessarily hold.
Occasionally, when no confusion can arise, we will, for simplicity's sake, omit the underlying domain in the notation, more precisely, we will write in place of or , etc.
For given rearrangement-invariant spaces and , we denote the boundedness of an operator from to by .
Let
. For any
, the
dilation operator is defined as
Such an operator is bounded on any rearrangement-invariant space
, with its norm smaller than or equal to max{
}.
We introduce, for
, the operator
()
Given any
such that
by
Hardy's lemma the inequality
holds for every non-increasing function
. Consequently, the
Hardy–Littlewood–Pólya principle, which asserts that if
satisfy
then
holds for every rearrangement-invariant space
.
The
fundamental function of a rearrangement-invariant space
is defined as
where
is measurable and such that
. Thanks to the rearrangement invariance of
, the function
is well defined. By [
20, Proposition 1.1], the fundamental function is locally absolutely continuous on (0,1]. We define the
fundamental level as the collection of all rearrangement-invariant spaces, which share the same fundamental function.
We say that a function is quasiconcave, if if and only if , it is positive and non-decreasing on , and the function is non-decreasing. Recall that by [2, Chapter 2, Corollary 5.3] for any rearrangement-invariant space , its fundamental function is quasiconcave. Furthermore, by [2, Chapter 2, Proposition 5.10] it holds that for any quasiconcave function there exists a concave function such that for every , the inequality holds.
Let
be a quasiconcave function and let
be a concave function such that
for every
. We then define the functionals
and
By [
2, Chapter 2, Theorem 5.13], these functionals are rearrangement-invariant function norms, and as such, we define the corresponding rearrangement-invariant spaces
, and
. Both of these spaces have the same fundamental function equivalent to
. Furthermore, the space
is the smallest rearrangement-invariant space with the fundamental function
, while
is the largest rearrangement-invariant space with the fundamental function
.
Given a rearrangement-invariant space
, we define the corresponding
Lorentz space and
Marcinkiewicz space . Let us now recall the Lorentz–Marcinkiewicz sandwich
()
We recall the embeddings into Lebesgue spaces for any rearrangement-invariant space
Let
. The functionals
and
are respectively defined as
for
. Let us recall that if
,
and if one of the conditions
- (L1) ,
- (L2) ,
- (L3) ,
is met, then
is equivalent to a rearrangement-invariant function norm. Then, the corresponding rearrangement-invariant function space
is called a
Lorentz space.
Let
. Then,
, and
where, if
, equality of the spaces is attained if and only if
. Note that by equality of two rearrangement-invariant spaces
and
, we mean that
and
coincide in set-theoretical sense, and, moreover, that their norms are equivalent in the sense that there exists a constant
such that
for every
.
Let
. The functionals
and
are defined as
for
. If one of the conditions
- (Z1) , , ,
- (Z2) ,
- (Z3) ,
- (Z4) ,
is met, then
is equivalent to a rearrangement-invariant function norm. Then, the corresponding rearrangement-invariant function space
is called a
Lorentz–Zygmund space.
We say that
is a
Young function, if it is a convex non-constant left-continuous function such that
. Let us recall that any such function may be written in the integral form
where
is a non-decreasing, left-continuous function, which is not identically 0 or
.
The
Luxemburg function norm is defined as
By [
2, Chapter 4, Theorem 8.9], the Luxemburg function norm
is a rearrangement-invariant function norm. The
Orlicz space is defined as the rearrangement-invariant space associated with the Luxemburg function norm. Then, for some
and
,
; and for
,
.
Let
and
be Young functions. We say that
A dominates
B near infinity if there exist constants
and
such that
We say that
and
are equivalent near infinity if they dominate each other near infinity. Furthermore, it holds that
We denote certain Orlicz spaces without explicitly defining the corresponding Young functions. The Orlicz space associated with a Young function equivalent near infinity to
, where
and
, or
and
, is denoted by
, and the Orlicz space associated with a Young function equivalent near infinity to
, where
, is denoted by exp
.
In certain cases, the classes of Lorentz–Zygmund and Orlicz spaces overlap. If either
,
, or
,
, then
. Additionally, if
, then
()
For certain classes of function spaces, their fundamental functions are known. By [
18], it holds that
and
Furthermore, by [
19, Example 7.9.4 (iv), p. 260], it holds that
() Therefore, for every fundamental level of rearrangement-invariant spaces, there exists a unique Orlicz space with the same fundamental function. This
fundamental Orlicz space is denoted by
, where
is a rearrangement-invariant space.
Let , . The fact that there exists a positive constant such that , or for any is denoted by or , respectively. If both inequalities and hold, we write .
Let , . By for , we denote that there exist constants , such that for every .
We say that
satisfies the
condition, if it is non-decreasing and the inequality
holds for every
. Then, by [
19, Theorem 4.7.3] it holds that
1.2 Isoperimetric functions and Sobolev spaces
In this section, we shall define some basic notions and recall simple facts concerning the isoperimetric function and Sobolev spaces.
Let
be as in Section
1.1. Let
. We define the
perimeter , of a Lebesgue-measurable set
as
where
denotes the essential boundary of
(for details, see [
14]). We then define the
isoperimetric function of
as
Note that
is finite for
(for the detailed proof, see [
5, Chapter 4]). Also, by [
5, Proposition 4.1], there exists a constant
such that
Thus, the best possible behavior of the isoperimetric function at 0 is
. What this means is that, essentially,
cannot decay more slowly than
as
.
We will say that has a Lipschitz boundary, or simply is a Lipschitz domain, if at each point of the boundary of , the boundary is locally the graph of a Lipschitz function.
We shall call
a
John domain, if there exists
and
such that for every
there exists a rectifiable curve parameterized by arclength
, such that
,
and
An important link between John domains and the theory of isoperimetric functions is that if
is a John domain, then
.
Let
. We define the
Maz'ya class of Euclidean domains
as
We observe that all Lipchitz domains are John domains and the inclusion holds for .
Let
, let
be a rearrangement-invariant function space. The
th order Sobolev space is defined as
where
denotes the vector of partial derivatives of
of order
, and the
th order Sobolev space is defined as
Assume now that
() Then, by [
5, Proposition 4.5]
in set-theoretical sense with their norms equivalent. If we only consider a weaker form of (
1.6), namely that there exists a positive constant
such that
it then holds by [
5, Chapter 4, Corollary 4.3, Proposition 4.4] that
,
for every
, and furthermore for any
rearrangement-invariant space,
if and only if there exists a positive constant
such that
for all
, where
In the case where is an Orlicz space, we define an Orlicz–Sobolev space as .
We say that
is the optimal (largest) rearrangement-invariant domain space in the embedding
() if
is a rearrangement-invariant space, embedding (
1.7) holds, and if (
1.7) holds with
replaced by a rearrangement-invariant space
, then the embedding
holds. Similarly, we say that
is the optimal (smallest) r.i. target space in embedding (
1.7), if
is a rearrangement-invariant space, embedding (
1.7) holds, and if embedding (
1.7) holds with
replaced by a rearrangement-invariant space
, then the embedding
holds.
2 BACKGROUND RESULTS
Let
,
, let
be an open set, and let
be a rearrangement-invariant space. For the sake of brevity, we shall refer to rearrangement-invariant spaces as r.i. spaces from this point onward. In this work, we consider Sobolev spaces
together with their norm defined as
for
. Furthermore, we consider Sobolev embeddings of the form
() where
is an r.i. space. We restrict ourselves to such sets
which fulfill the property
and classes of such sets. Furthermore, it is known that given such an r.i. space
, the optimal r.i. domain space always exists and can be explicitly described. Namely, the combination of a reduction principle from [
5] with the characterization of the optimal domain for a Copson integral operator from [
15, Proposition 3.3], such optimal r.i. space
obeys
() where the supremum is taken over all
such that
.
We first examine possible approaches to the reduction of Sobolev embeddings to significantly simpler one-dimensional inequalities for Maz'ya domains. Such problems have already been examined and solved, and as such, for our purposes, it suffices to use the reduction principle stated and proven in [5, Theorem 6.4], which follows. For the proof, see the original paper.
Theorem 2.1. (Reduction principle for Maz'ya domains)Let , , and . Let and be r.i. function norms. Let be such that
() for any
nonnegative. Then, the Sobolev embedding (
2.1) holds for every
. Conversely, if the Sobolev embedding (
2.1) holds for every
, then the inequality (
2.3) holds.
Note that we have omitted the case . While our results may be, after some modification, applied to such a case, it remains rather technical and is beyond the scope of this work.
As a consequence of Theorem 2.1, we can identify the optimal r.i. target space associated with a given domain space in the Sobolev embedding (2.1) for any . This is also a known result, for the proof see [5, Theorem 6.5].
Theorem 2.2. (optimal r.i. target for Maz'ya domains)Let and be as in Theorem 2.1. Define the functional as
where
denotes the associate function norm to
. Then, the functional
is an r.i. function norm, whose associate norm
satisfies
() for every
, and for some constant
depending on
,
,
, and
() for every
and every
. Furthermore, the function norm
is optimal in (
2.4) and (
2.5) among all r.i. norms, as
ranges in
.
We shall now use Theorem 2.2 to show the optimal target r.i. space for certain critical r.i. spaces.
Example.Let . Then, for any and for any , given the critical spaces , we obtain the embedding
where
is the smallest (= optimal) space with this property for any
. In the case
, we obtain the embedding
where
is the optimal target space. In the case
, we obtain the embedding
where
is the optimal target space. In the case
, we obtain the embedding
() where
is the optimal target space, and which recovers, as its special case for
, the result of [
3].
The question remains whether we can say anything about the optimality of r.i. domains, given a target r.i. space. This problem has also been extensively studied, and as such, for our purposes, it suffices to modify a known result by [13, Theorem 3.3].
Theorem 2.3. (optimal r.i. domain space)Let , , . Let be an r.i. space such that
() Define the functional
as
where the supremum is extended over all
such that
, and the set
Then, the embedding
holds for every
. Furthermore,
is the largest r.i. space with this property.
The proof of Theorem 2.3 is a simple modification of the proof given in [13, Theorem 3.3] and therefore is omitted. Note that, by [15, Proposition 3.3] the theorem may be strengthened by omitting condition (2.7).
We shall now discuss the fundamental Orlicz spaces of r.i. spaces. The following theorem was established and proved by [16].
Theorem 2.4. (The principal alternative)
- (i) Let be an r.i. space and its fundamental Orlicz space. Then, either and is the smallest Orlicz space containing , or and no smallest Orlicz space containing exists.
- (ii) Let be an r.i. space and its fundamental Orlicz space. Then, either and is the largest Orlicz space contained in space , or and no largest Orlicz space contained in exists.
Next, we specify the principal alternative to Sobolev embeddings. The following theorem is an adjustment to [16, Theorem 4.1], its proof is a simple modification of the proof in the original paper and therefore is omitted.
Theorem 2.5. (principal alternative for Sobolev embeddings)Let , , , and let , be r.i. norms.
- (i) If is the largest among all r.i. spaces rendering embedding (2.1) true for every , then either and is the largest Orlicz space such that
holds for every , or no such largest Orlicz space exists.
- (ii) If is the smallest among all r.i. spaces rendering embedding (2.1) true for every , then either and is the smallest Orlicz space such that
holds for every , or no such smallest Orlicz space exists.
3 MAIN RESULTS
First, we shall present a general result concerning the fundamental function of an operator-induced space.
Theorem 3.1. (Fundamental function of an operator-induced space)Let be a non-decreasing measurable function such that the function belongs to , where is an r.i. space such that
() and
() Let
be defined by
where the supremum is extended over all
such that
. Then,
is an r.i. space, and one has
Proof.The fact that is an r.i. space follows from Theorem 2.3. First, we will examine the lower bound of . Fix . By the boundedness of the dilation operator on , we get
showing that
Since the function
is non-increasing, it follows that
Thus, we have obtained the lower bound of
, as
() Let us focus on the upper bound of
. Note that since
is locally absolutely continuous on (0,1], it is differentiable a.e. and for any measurable
, one has
Therefore, by the fundamental embedding (
1.3), (
3.1) and Fubini's theorem, we obtain that for any
, it holds that
Hence, by the definition of
and by the Hardy–Littlewood inequality (
1.1), it holds for any
, that
These inequalities give us an upper bound of
, as
() Thus, by combining the lower bound (
3.3) and the upper bound (
3.4) of
for
, we obtain the estimates
Finally, applying inequality (
3.2) to the estimates above, we get the desired result
The particular case where is a Lipschitz domain was treated in [16, Theorem 4.2].
We shall now use the result obtained in Theorem 3.1 to extract important information about fundamental functions of optimal domain r.i. spaces in Sobolev embeddings corresponding to given target spaces sharing the same fundamental level.
Corollary 3.2.Let be with isoperimetric profile . Suppose that , are r.i. spaces over on the same fundamental level, that is, , and (3.1) and (3.2) hold for both and in place of . Let be the optimal r.i. domain spaces in the embedding
for
. Then,
,
are also on the same fundamental level, that is,
.
Proof.The proof follows immediately from the formula of the norm in the optimal r.i. domain space (2.2) and by Theorem 3.1.
The presuppositions of Theorem 3.1, namely (3.2), are not particularly flexible, as there possibly exist classes of isoperimetric functions for which the presupposition is needlessly strong, or even entirely unnecessary. We aim to prove that for Maz'ya classes of domains, Theorem 3.1 may be applied to their isoperimetric function directly, omitting the need for presupposition (3.2) entirely. To prove this, we must find a work-around, so we do not need to consider every domain in .
In order to obtain the necessary condition for embedding (2.1) for some in the reduction principle (Theorem 2.1), we need that the embedding holds for every . We will however point out that it suffices to focus on the worst domain with such property, denoted . To describe this domain, we use [5, Proposition 10.1 (i)], which follows. In the statement, denotes the Lebesgue measure of the unit ball in .
Proposition 3.3.Let , , . Define as
Let
be the Euclidean domain in
given by
Then
, and
()
The best case scenario happens when
and we are dealing with John domains. For example, if
and
, then the function
takes the form
and, as a result,
In other words,
takes the form of a triangle with vertices (denoted by
)
,
, and [2,0]. We may easily verify that
. The domains start to get worse as we choose
. It is on these domains
that we consider the following corollary.
Corollary 3.4.Let , , . Define as in Proposition 3.3. Then, for , Theorem 3.1 and Corollary 3.2 are applicable without the restriction (3.2).
Proof.Let . Returning to the proof of Theorem 3.1, in the case of , the lower bound (3.3) contains a computable integral. Thus, it holds that for some constant
Therefore,
We observe that, owing to our choice of
and (
3.5), this computation cannot produce a smaller lower bound for any
. Moreover, it follows from (
3.4) that
As such, to prove the corollary, it is sufficient to prove that for any
() Define
We claim that
() Indeed,
Therefore,
is non-decreasing. It also follows from equality (
3.7) that for any
and any
, one has
for
and therefore
Hence,
and by definition of
, we obtain the desired result (
3.6).
Combining all the results of this chapter established so far yields the following theorem, which is an extension of [16, Theorem 4.9].
Theorem 3.5. (Nonexistence of an optimal Orlicz space)Let , let , and let be an r.i. space. Assume there is no largest Orlicz space such that
() for every
in
. Then, there is no largest Orlicz space
such that
() for every
in
.
Proof.We know that and share the same fundamental function. We denote by and the largest domain r.i. spaces in embeddings
and
respectively. Then, by Corollary
3.2 combined with Corollary
3.4, we obtain that
. Owing to the assumption, there is no largest Orlicz space
in the Sobolev embedding (
3.8). Therefore, by Theorem
2.5, it follows that
. Since it holds that
, it is also true that
. Consequently,
. But, since
, we have
. Altogether,
, whence, using Theorem
2.5 once again, there is no largest Orlicz space
in the Sobolev embedding (
3.9).
We shall now apply Theorem
3.5 to show that there is no optimal Orlicz domain space
in the embedding
thereby solving the open problem mentioned in the introduction. The value
corresponds to the case in which the target space is optimal in the embedding of the limiting Sobolev space, see (
2.6).
First, we shall prove two lemmas, which we will later use in the proof. The first lemma concerns the r.i. norm of the characteristic function.
Lemma 3.6.Let , and be an r.i. norm. Then
Proof.The inequality
follows from the property (P2) of
. Conversely, we observe that
since, by the triangle inequality and by the rearrangement-invariance of
,
Altogether, one has
with the constants of equivalence depending only on
.
The second lemma concerns the boundedness of the operator introduced in (1.2).
Lemma 3.7.Let . Then
()
Proof.This is a particular case of [10, Theorem 3.2], applied to , and , for . It is readily verified that the necessary and sufficient condition for (3.10), namely
holds.
We are now in position to formulate and prove our main result.
Theorem 3.8.Let , , , such that , and let . Then, there is no largest Orlicz space such that
() for every
in
.
Proof.By [18, Lemma 3.7], for every , the spaces and share the same fundamental function , where
and it holds that
. Moreover, by the aforementioned lemma, it holds that
for any r.i. space
with the fundamental function
, and in fact,
.
We aim to prove that there is no largest Orlicz space such that the embedding (3.11) holds for every in . Therefore, by Theorem 3.5, it suffices to prove that there is no largest Orlicz space such that the embedding
() holds for every
in
. By Theorem
2.5 i, it is therefore enough to prove that if
is the optimal domain r.i. space in
for every
in
, then
.
First, we find the optimal domain space . We use Theorem 2.3 and obtain the formula
By Lemma
3.7, applied to
, the operator
is bounded on
. Furthermore, by [
18, Theorem 6.11] and (
1.4), it holds that
. Hence, it follows from [
8, Theorem 4.7] that
As such, we will with no loss of generality suppose that
Next, we describe . For that purpose, we need to find so we can determine its Young function . Let . Then, as = and by Lemma 3.6, we get
We therefore know that
Hence, by (
1.5) and simple computation, it holds that
Finally, we need to prove that . We know that is optimal in the embedding , and consequently, it suffices to prove that . Then, by Theorem 2.1, it is enough to prove that the inequality
does not hold. Let
, where
is anywhere in the interval
. Owing to the assumption, the interval is non-empty. First, we prove that
, that is,
. Since
satisfies the
condition, it holds that
. Furthermore,
Therefore, our choice of
guarantees that
. Moreover, we know that
thus it is sufficient to prove that
() We compute the integral and obtain
so the equality holds if
. Hence, once again, our choice of
yields (
3.13). Therefore, we have found a function
such that
Consequently, by Theorem
2.1 . By Theorem
2.5, there is no largest Orlicz space which would render the embedding (
3.12) true for every
in
. Finally, by Theorem
3.5, there is no largest Orlicz space which would render the embedding (
3.11) true for every
in
.
Remark 3.9.In this work, we have focused primarily on Maz'ya classes of Euclidean domains. We are aware that there are plenty of open questions worth pursuing which we leave open. Pivotal examples are the case in Maz'ya domains, Gaussian–Sobolev embeddings, embeddings on domains endowed with Frostman–Ahlfors measures, or embeddings on probability spaces. The reason we do not consider these cases is that the corresponding integral operators get too complicated (e.g., they take the form, at least for higher-order embeddings, of kernel-type operators). Hence, such considerations reach beyond the scope of this text. We plan to study them in our following work.