Abstract

We give characterizations for homogeneous and inhomogeneous Besov-Lizorkin-Triebel spaces (H. Triebel 1983, 1992, and 2006) in terms of continuous local means for the full range of parameters. In particular, we prove characterizations in terms of Lusin functions (tent spaces) and spaces involving the Peetre maximal function to apply the classical coorbit space theory according to Feichtinger and Gröchenig (H. G Feichtinger and K. Gröchenig 1988, 1989, and 1991). This results in atomic decompositions and wavelet bases for homogeneous spaces. In particular we give sufficient conditions for suitable wavelets in terms of moment, decay and smoothness conditions.

1. Introduction

This paper deals with Besov-Lizorkin-Triebel spaces ̇𝐵𝑠𝑝,𝑞(𝑑) and ̇𝐹𝑠𝑝,𝑞(𝑑) on the Euclidean space 𝑑 and their interpretation as coorbits. For this purpose we prove a number of characterizations for homogeneous and inhomogeneous spaces for the full range of parameters. Classically introduced in Triebel’s monograph [1, Section 2.3.1] by means of a dyadic decomposition of unity, we use more general building blocks and provide in addition continuous characterizations in terms of Lusin and maximal functions. Equivalent (quasi-) normings of this kind were first given by Triebel in [2]. His proofs use in an essential way the fact that the function under consideration belongs to the respective space. Therefore, the obtained equivalent (quasi-)norms could not yet be considered as a definition or characterization of the space. Later on, Triebel was able to solve this problem partly in his monograph [3, Sections 2.4.2, 2.5.1] by restricting to the Banach space case. Afterwards, Rychkov [4] completed the picture by simplifying a method due to Bui et al. [5, 6]. However, [4] contains some problematic arguments. One aim of the present paper is to provide a complete and self-contained reference for general characterizations of discrete and continuous type by avoiding these arguments. We use a variant of a method from Rychkov’s subsequent papers [7, 8], which is originally due to Strömberg and Torchinsky developed in their monograph [9, Chapter 5].

In a different language, the results can be interpreted in terms of the continuous wavelet transform (see Appendix A.1) belonging to a function space on the 𝑎𝑥+𝑏-group 𝒢. Spaces on 𝒢 considered here are mixed norm spaces like tent spaces [10] and Peetre-type spaces. The latter are indeed new and received their name from the fact that quantities related to the classical Peetre maximal function are involved. This leads to the main intention of the paper. We use the established characterizations for the homogeneous spaces in order to embed them in the abstract framework of coorbit space theory originally due to Feichtinger and Gröchenig [1115] in the 1980s. This connection was already observed by them in [11, 14, 15]. They worked with Triebel’s equivalent continuous normings from [2] and the results on tent spaces, which were introduced more or less at the same time by Coifman et al. [10] to interpret Lizorkin-Triebel spaces as coorbits. On the one hand the present paper gives a late justification, and on the other hand, we observe that Peetre-type spaces on 𝒢 are a much better choice for this issue. Their two-sided translation invariance is immediate and much more transparent as we will show in Section 4.1. Furthermore, generalizations in different directions are now possible. In a forthcoming paper, we will show how to apply a generalized coorbit space theory due to Fornasier and Rauhut [16] in order to recover generalized inhomogeneous spaces based on the characterizations given here. Moreover, the extension of the results to quasi-Banach spaces using a theory developed by Rauhut in [17, 18] is possible.

Once we have interpreted classical homogeneous Besov-Lizorkin-Triebel spaces as certain coorbits, we are able to benefit from the achievements of the abstract theory in [1115]. The main feature is a powerful discretization machinery which leads in an abstract universal way to atomic decompositions. We are now able to apply this method, which results in atomic decompositions and wavelet bases for homogeneous spaces. More precisely, sufficient conditions in terms of vanishing moments, decay, and smoothness properties of the respective wavelet function are given. Compact support of the used atoms does not play any role here. In particular, we specify the order of a suitable orthonormal spline wavelet system depending on the parameters of the respective space.

The paper is organized as follows. After giving some preliminaries, we start in Section 2 with the definition of classical Besov-Lizorkin-Triebel spaces and their characterization via continuous local means. In Section 3, we give a brief introduction to abstract coorbit space theory, which is applied in Section 4 on the 𝑎𝑥+𝑏-group 𝒢. We recover the homogeneous spaces from Section 2 as coorbits of certain spaces on 𝒢. Finally, several discretization results in terms of atomic decompositions and wavelet isomorphisms are established. The underlying decay result of the continuous wavelet transform and some basic facts about orthonormal wavelet bases are shifted to the appendix.

1.1. Notation

Let us first introduce some basic notation. The symbols ,,,0,and denote the real numbers, complex numbers, natural numbers, natural numbers including 0, and the integers. The dimension of the underlying Euclidean space for function spaces is denoted by 𝑑, its elements will be denoted by 𝑥,𝑦,𝑧,, and |𝑥| is used for the Euclidean norm. We will use |𝑘|1 for the 𝑑1-norm of a vector 𝑘. For a multi-index 𝛼 and 𝑥𝑑, we write𝑥𝛼=𝑥𝛼11𝑥𝛼𝑑𝑑(1.1)

and define the differential operators 𝐷𝛼 and Δ by𝐷𝛼=𝜕|𝛼|1𝜕𝑥𝛼11𝜕𝑥𝛼𝑑𝑑,Δ=𝑑𝑘=1𝜕2𝜕𝑥2𝑘.(1.2)

If 𝑋 is a (quasi-)Banach space and 𝑓𝑋, we use 𝑓𝑋 or simply 𝑓 for its (quasi-)norm. Operator norms of linear mappings 𝐴𝑋𝑌 are denoted by 𝐴𝑋𝑌 or simply by 𝐴. As usual, the letter 𝑐 denotes a constant, which may vary from line to line but is always independent of 𝑓, unless the opposite is explicitly stated. We also use the notation 𝑎𝑏 if there exists a constant 𝑐>0 (independent of the context-dependent relevant parameters) such that 𝑎𝑐𝑏. If 𝑎𝑏 and 𝑏𝑎, we will write 𝑎𝑏.

2. Function Spaces on 𝑑

2.1. Vector-Valued Lebesgue Spaces

The space 𝐿𝑝(𝑑), 0<𝑝, denotes the collection of complex-valued functions (equivalence classes) with finite (quasi-)norm𝑓𝐿𝑝𝑑=𝑑||||𝑓(𝑥)𝑝𝑑𝑥1/𝑝,(2.1) with the usual modification if 𝑝=. The Hilbert space 𝐿2(𝑑) plays a separate role, see for instance, Section 3. Having a sequence of complex-valued functions {𝑓𝑘}𝑘𝐼 on 𝑑, where 𝐼 is a countable index set, we put𝑓𝑘𝑞𝐿𝑝𝑑=𝑘𝐼𝑓𝑘𝐿𝑝𝑑𝑞1/𝑞,𝑓𝑘𝐿𝑝𝑞,𝑑=𝑘𝐼||𝑓𝑘||(𝑥)𝑞1/𝑞𝐿𝑝𝑑,(2.2) where we modify appropriately in the case 𝑞=.

2.2. Maximal Functions

For a locally integrable function 𝑓, we denote by 𝑀𝑓(𝑥) the Hardy-Littlewood maximal function defined by(𝑀𝑓)(𝑥)=sup𝑥𝑄1||𝑄||𝑄||||𝑓(𝑦)𝑑𝑦,𝑥𝑑,(2.3) where the supremum is taken over all cubes centered at 𝑥 with sides parallel to the coordinate axes. The following theorem is due to Fefferman and Stein [19].

Theorem 2.1. For 1<𝑝< and 1<𝑞, there exists a constant 𝑐>0, such that 𝑀𝑓𝑘𝐿𝑝𝑞,𝑑𝑓𝑐𝑘𝐿𝑝𝑞(2.4) holds for all sequences {𝑓𝑘}𝑘 of locally Lebesgue-integrable functions on 𝑑.

Let us recall the classical Peetre maximal operator, introduced in [20]. Given a sequence of function {Ψ𝑘}𝑘𝒮(𝑑), a tempered distribution 𝑓𝒮(𝑑) and a positive number 𝑎>0, we define the system of maximal functionsΨ𝑘𝑓𝑎(𝑥)=sup𝑦𝑑||Ψ𝑘||𝑓(𝑥+𝑦)1+2𝑘||𝑦||𝑎,𝑥𝑑,𝑘.(2.5)

Since (Ψ𝑘𝑓)(𝑦) makes sense pointwise (see the following paragraph), everything is well-defined. However, the value “’’ is also possible for (Ψ𝑘𝑓)𝑎(𝑥). This was the reason for the problematic arguments in [4] mentioned in the introduction. We will often use dilates Ψ𝑘(𝑥)=2𝑘𝑑Ψ(2𝑘𝑥) of a fixed function Ψ𝒮(𝑑), where Ψ0(𝑥) might be given by a separate function. Also continuous dilates are needed. Let the operator 𝒟𝐿𝑝𝑡, 𝑡>0, generate the 𝑝-normalized dilates of a function Ψ given by 𝒟𝐿𝑝𝑡Ψ=𝑡𝑑/𝑝Ψ(𝑡1). If 𝑝=1, we omit the super index and use additionally Ψ𝑡=𝒟𝑡Ψ=𝒟𝐿1𝑡Ψ. We define (Ψ𝑡𝑓)𝑎(𝑥) byΨ𝑡𝑓𝑎(𝑥)=sup𝑦𝑑||Ψ𝑡||𝑓(𝑥+𝑦)||𝑦||1+/𝑡𝑎,𝑥𝑑,𝑡>0.(2.6) We will refer to this construction later on. It turned out that this maximal function construction can be used to interpret classical smoothness spaces as coorbits of certain function spaces on the group.

2.3. Tempered Distributions and Fourier Transform

As usual, 𝒮(𝑑) is used for the locally convex space of rapidly decreasing infinitely differentiable functions on 𝑑, where the topology is generated by the family of seminorms𝜑𝑘,=sup𝑥𝑑,|𝛼|1||𝐷𝛼||𝜑(𝑥)(1+|𝑥|)𝑘,𝜑𝒮𝑑,𝑘,0.(2.7) The space 𝒮(𝑑) is called the set of all tempered distributions on 𝑑 and defined as the topological dual of 𝒮(𝑑). Indeed, a linear mapping 𝑓𝒮(𝑑)belongs to 𝒮(𝑑) if and only if there exist numbers 𝑘,0 and a constant 𝑐=𝑐𝑓 such that||||𝑓(𝜑)𝑐𝑓sup𝑥𝑑,|𝛼|1||𝐷𝛼||𝜑(𝑥)(1+|𝑥|)𝑘(2.8)for all 𝜑𝒮(𝑑). 𝒮(𝑑) is equipped with the weak-topology.

The convolution 𝜑𝜓 of two integrable functions 𝜑,𝜓 is defined via the integral(𝜑𝜓)(𝑥)=𝑑𝜑(𝑥𝑦)𝜓(𝑦)𝑑𝑦.(2.9) If 𝜑,𝜓𝒮(𝑑), then (2.9) still belongs to 𝒮(𝑑). The convolution can be generalized to 𝒮(𝑑)×𝒮(𝑑) via (𝜑𝑓)(𝑥)=𝑓(𝜑(𝑥)), makes sense pointwise, and is a 𝐶-function in 𝑑 of at most polynomial growth.

As usual, the Fourier transform defined on both 𝒮(𝑑) and 𝒮(𝑑) is given by (𝑓)(𝜑)=𝑓(𝜑), where 𝑓𝒮(𝑑),𝜑𝒮(𝑑), and𝜑(𝜉)=(2𝜋)𝑑/2𝑑𝑒𝑖𝑥𝜉𝜑(𝑥)𝑑𝑥.(2.10) The mapping is a bijection (in both cases) and its inverse is given by 1𝜑=𝜑().

In order to deal with homogeneous spaces, we need to define the subset 𝒮0(𝑑)𝒮(𝑑). Following [1, Chapter 5], we put𝒮0𝑑=𝜑𝒮𝑑𝐷𝛼(𝜑)(0)=0foreverymulti-index𝛼𝑑0.(2.11) The set 𝒮0(𝑑) denotes the topological dual of 𝒮0(𝑑). If 𝑓𝒮(𝑑), the restriction of 𝑓 to 𝒮0(𝑑) clearly belongs to 𝒮0(𝑑). Furthermore, if 𝑃(𝑥) is an arbitrary polynomial in 𝑑, we have (𝑓+𝑃())(𝜑)=𝑓(𝜑) for every 𝜑𝒮0(𝑑). Conversely, if 𝑓𝒮0(𝑑), then 𝑓 can be extended from 𝒮0(𝑑) to 𝒮(𝑑), that is, to an element of 𝒮(𝑑). However, this fact is not trivial and makes use of the Hahn-Banach theorem in locally convex topological vector spaces. We may identify 𝒮0(𝑑) with the factor space 𝒮(𝑑)/𝒫(𝑑), since two different extensions differ by a polynomial.

2.4. Besov-Lizorkin-Triebel Spaces

Let us first introduce the concept of a dyadic decomposition of unity, see also [1, Section 2.3.1].

Definition 2.2. (a) Let Φ(𝑑) be the collection of all systems {𝜑𝑗(𝑥)}𝑗0𝒮(𝑑) with the following properties:(i)𝜑𝑗(𝑥)=𝜑(2𝑗𝑥),𝑗, (ii)supp𝜑0{𝑥𝑑|𝑥|2},supp𝜑{𝑥𝑑1/2|𝑥|2},(iii)𝑗=0𝜑𝑗(𝑥)=1 for every 𝑥𝑑.
(b) Moreover, the system ̇Φ(𝑑) denotes the collection of all systems {𝜑𝑗(𝑥)}𝑗𝒮(𝑑) with the following properties: (i)𝜑𝑗(𝑥)=𝜑(2𝑗𝑥),𝑗, (ii)supp𝜑={𝑥𝑑1/2|𝑥|2},(iii)𝑗=𝜑𝑗=1 for every 𝑥𝑑{0}.

Remark 2.3. If we take 𝜑0𝒮(𝑑) satisfying 𝜑0(𝑥)=1:|𝑥|10:|𝑥|>2(2.12) and define 𝜑(𝑥)=𝜑0(𝑥)𝜑0(2𝑥), then the system {𝜑𝑗(𝑥)}𝑗0 belongs to Φ(𝑑) and the system {𝜑𝑗(𝑥)}𝑗 with 𝜑0=𝜑 belongs to ̇Φ(𝑑).

Now we are ready for the definition of the Besov and Lizorkin-Triebel spaces. See for instance [1, Section 2.3.1] for details and further properties.

Definition 2.4. Let {𝜑𝑗(𝑥)}𝑗=0Φ(𝑑) and Φ𝑗=1𝜑𝑗, 𝑗0. Let further <𝑠< and 0<𝑞. (i) If 0<𝑝, then𝐵𝑠𝑝,𝑞𝑑=𝑓𝒮𝑑𝑓𝐵𝑠𝑝,𝑞𝑑=𝑗=02𝑗𝑠𝑞Φ𝑗𝑓𝐿𝑝𝑑𝑞1/𝑞<.(2.13)(ii)If 0<𝑝<, then𝐹𝑠𝑝,𝑞𝑑=𝑓𝒮𝑑𝑓𝐹𝑠𝑝,𝑞𝑑=𝑗=02𝑗𝑠𝑞||Φ𝑗||𝑓(𝑥)𝑞1/𝑞𝐿𝑝𝑑.<(2.14) In case 𝑞=, we replace the sum by a supremum in both cases.

The homogeneous counterparts are defined as follows. For details, further properties and how to deal with ocurring technicalities we refer to [1, Chapter 5].

Definition 2.5. Let {𝜑𝑗(𝑥)}𝑗̇Φ(𝑑) and Φ𝑗=1𝜑𝑗. Let further <𝑠< and 0<𝑞. (i) If 0<𝑝, theṅ𝐵𝑠𝑝,𝑞𝑑=𝑓𝒮0𝑑̇𝐵𝑓𝑠𝑝,𝑞𝑑=𝑗=2𝑗𝑠𝑞Φ𝑗𝑓𝐿𝑝𝑑𝑞1/𝑞.<(2.15) (ii) If 0<𝑝<, theṅ𝐹𝑠𝑝,𝑞𝑑=𝑓𝒮0𝑑̇𝐹𝑓𝑠𝑝,𝑞𝑑=𝑗=2𝑗𝑠𝑞||Φ𝑗||𝑓(𝑥)𝑞1/𝑞𝐿𝑝𝑑.<(2.16) In case 𝑞=, we replace the sum by a supremum in both cases.

2.5. Inhomogeneous Spaces

Essential for the sequel are functions Φ0,Φ𝒮(𝑑) satisfying||Φ0||||||𝜀(𝑥)>0on{|𝑥|<2𝜀},(Φ)(𝑥)>0on2,<|𝑥|<2𝜀(2.17) for some 𝜀>0, and𝐷𝛼||(Φ)(0)=0𝛼||𝑅.(2.18) We will call the functions Φ0 and Φ kernels for local means. Recall that Φ𝑘=2𝑘𝑑Φ(2𝑘), 𝑘, and Ψ𝑡=𝒟𝑡Ψ.

Theorem 2.6. Let 𝑠, 0<𝑝<, 0<𝑞, 𝑎>𝑑/min{𝑝,𝑞} and 𝑅+1>𝑠. Let further Φ0,Φ𝒮(𝑑) be given by (2.17) and (2.18). Then the space 𝐹𝑠𝑝,𝑞(𝑑) can be characterized by 𝐹𝑠𝑝,𝑞𝑑=𝑓𝒮𝑑𝑓𝐹𝑠𝑝,𝑞𝑑𝑖<,𝑖=1,,5,(2.19) where 𝑓𝐹𝑠𝑝,𝑞1=Φ0𝑓𝐿𝑝𝑑+10𝑡𝑠𝑞||Φ𝑡||𝑓(𝑥)𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑑,(2.20)𝑓𝐹𝑠𝑝,𝑞2=Φ0𝑓𝐿𝑝𝑑+10𝑡𝑠𝑞sup𝑧𝑑||Φ𝑡||𝑓(𝑥+𝑧)(1+|𝑧|/𝑡)𝑎𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑑,(2.21)𝑓𝐹𝑠𝑝,𝑞3=Φ0𝑓𝐿𝑝𝑑+10𝑡𝑠𝑞|𝑧|<𝑡||Φ𝑡||𝑓(𝑥+𝑧)𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑,(2.22)𝑓𝐹𝑠𝑝,𝑞4=𝑘=02𝑠𝑘𝑞sup𝑧𝑑||Φ𝑘||𝑓(𝑥+𝑧)(1+2𝑘|𝑧|)𝑎𝑞1/𝑞𝐿𝑝𝑑,(2.23)𝑓𝐹𝑠𝑝,𝑞5=𝑘=02𝑠𝑘𝑞||Φ𝑘||𝑓(𝑥)𝑞1/𝑞𝐿𝑝𝑑,(2.24) with the usual modification in case 𝑞=. Furthermore, all quantities 𝑓𝐹𝑠𝑝,𝑞(𝑑)𝑖, 𝑖=1,,5, are equivalent (quasi-)norms in 𝐹𝑠𝑝,𝑞(𝑑).

For the inhomogeneous Besov spaces, we have the following characterizations.

Theorem 2.7. Let 𝑠, 0<𝑝,𝑞, 𝑎>𝑑/𝑝, and 𝑅+1>𝑠. Let further Φ0,Φ𝒮(𝑑) be given by (2.17) and (2.18). Then the space 𝐵𝑠𝑝,𝑞(𝑑) can be characterized by 𝐵𝑠𝑝,𝑞𝑑=𝑓𝒮𝑑𝑓𝐵𝑠𝑝,𝑞𝑑𝑖<,𝑖=1,,4,(2.25) where 𝑓𝐵𝑠𝑝,𝑞1=Φ0𝑓𝐿𝑝𝑑+10𝑡𝑠𝑞Φ𝑡𝑓(𝑥)𝐿𝑝𝑑𝑞𝑑𝑡𝑡1/𝑞,𝑓𝐵𝑠𝑝,𝑞2=Φ0𝑓𝐿𝑝𝑑+10𝑡𝑠𝑞sup𝑧𝑑||Φ𝑡||𝑓(𝑥+𝑧)(1+|𝑧|/𝑡)𝑎𝐿𝑝𝑑𝑞𝑑𝑡𝑡1/𝑞,𝑓𝐵𝑠𝑝,𝑞3=𝑘=02𝑠𝑘𝑞sup𝑧𝑑||Φ𝑘||𝑓(𝑥+𝑧)1+2𝑘|𝑧|𝑎𝐿𝑝𝑑𝑞1/𝑞,𝑓𝐵𝑠𝑝,𝑞4=𝑘=02𝑠𝑘𝑞Φ𝑘𝑓(𝑥)𝐿𝑝𝑑𝑞1/𝑞,(2.26) with the usual modification if 𝑞=. Furthermore, all quantities 𝑓|𝐵𝑠𝑝,𝑞(𝑑)𝑖, 𝑖=1,,4, are equivalent quasinorms in 𝐵𝑠𝑝,𝑞(𝑑).

2.6. Homogeneous Spaces

The homogeneous spaces can be characterized in a similar way. Here we do not have a separate function Φ0 anymore. We put Φ0=Φ.

Theorem 2.8. Let 𝑠, 0<𝑝<, 0<𝑞, 𝑎>𝑑/min{𝑝,𝑞}, and 𝑅+1>𝑠. Let further Φ𝒮(𝑑) be given by (2.17) and (2.18). Then the space ̇𝐹𝑠𝑝,𝑞(𝑑) can be characterized by ̇𝐹𝑠𝑝,𝑞𝑑=𝑓𝒮0𝑑̇𝐹𝑓𝑠𝑝,𝑞𝑑𝑖<,𝑖=1,,5,(2.27) where ̇𝐹𝑓𝑠𝑝,𝑞1=0𝑡𝑠𝑞||Φ𝑡||𝑓(𝑥)𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑑,̇𝐹𝑓𝑠𝑝,𝑞2=0𝑡𝑠𝑞sup𝑧𝑑||Φ𝑡||𝑓(𝑥+𝑧)(1+|𝑧|/𝑡)𝑎𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑑,̇𝐹𝑓𝑠𝑝,𝑞3=0𝑡𝑠𝑞|𝑧|<𝑡||Φ𝑡||𝑓(𝑥+𝑧)𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑,̇𝐹𝑓𝑠𝑝,𝑞4=𝑘=2𝑠𝑘𝑞sup𝑧𝑑||Φ𝑘||𝑓(𝑥+𝑧)(1+2𝑘|𝑧|)𝑎𝑞1/𝑞𝐿𝑝𝑑,̇𝐹𝑓𝑠𝑝,𝑞5=𝑘=2𝑠𝑘𝑞||Φ𝑘||𝑓(𝑥)𝑞1/𝑞𝐿𝑝𝑑(2.28) with the usual modification if 𝑞=. Furthermore, all quantities ̇𝐹𝑓|𝑠𝑝,𝑞(𝑑)𝑖, 𝑖=1,,5, are equivalent quasinorms in ̇𝐹𝑠𝑝,𝑞(𝑑).

For the Besov spaces, we obtain the following characterizations.

Theorem 2.9. Let 𝑠,0<𝑝,𝑞, 𝑎>𝑑/𝑝, and 𝑅+1>𝑠. Let further Φ𝒮(𝑑) be given by (2.17) and (2.18). Then the space ̇𝐵𝑠𝑝,𝑞(𝑑) can be characterized by ̇𝐵𝑠𝑝,𝑞𝑑=𝑓𝒮0𝑑̇𝐵𝑓𝑠𝑝,𝑞𝑑𝑖<,𝑖=1,,4,(2.29) where ̇𝐵𝑓𝑠𝑝,𝑞1=0𝑡𝑠𝑞Φ𝑡𝑓(𝑥)𝐿𝑝𝑑𝑞𝑑𝑡𝑡1/𝑞,̇𝐵𝑓𝑠𝑝,𝑞2=0𝑡𝑠𝑞sup𝑧𝑑||Φ𝑡||𝑓(𝑥+𝑧)(1+|𝑧|/𝑡)𝑎𝐿𝑝𝑑𝑞𝑑𝑡𝑡1/𝑞,̇𝐵𝑓𝑠𝑝,𝑞3=𝑘=2𝑠𝑘𝑞sup𝑧𝑑||Φ𝑘||𝑓(𝑥+𝑧)1+2𝑘|𝑧|𝑎𝐿𝑝𝑑𝑞1/𝑞,̇𝐵𝑓𝑠𝑝,𝑞4=𝑘=2𝑠𝑘𝑞Φ𝑘𝑓(𝑥)𝐿𝑝𝑑𝑞1/𝑞,(2.30) with the usual modification if 𝑞=. Furthermore, all quantities ̇𝐵𝑓𝑠𝑝,𝑞(𝑑)𝑖, 𝑖=1,,4, are equivalent quasinorms in ̇𝐵𝑠𝑝,𝑞(𝑑).

Remark 2.10. Observe, that the (quasi-)norms ̇𝐹𝑠𝑝,𝑞(𝑑)3 and 𝐹𝑠𝑝,𝑞(𝑑)3 are characterizations via Lusin functions, see [3, Section 2.4.5] and [1, Section 2.12.1] and the references given there. We will return to it later when defining tent spaces, see Definition 4.1 and (4.3).

2.7. Particular Kernels

For more details concerning particular choices for the kernels Φ0 and Φ, we refer mainly to Triebel [3, Section 3.3].

The most prominent nontrivial examples (besides the one given in Remark 2.3) of functions Φ0 and Φ satisfying (2.17) and (2.18) are the classical local means. The name comes from the compact support of Φ0,Φ, which is admitted in the following statement.

Corollary 2.11. Let 𝑝,𝑞,𝑠 as in Theorem 2.6. Let further 𝑘0,𝑘0𝒮(𝑑) such that 𝑘0(0),𝑘0(0)0(2.31) and define Φ0=𝑘0,Φ=Δ𝑁𝑘0,(2.32) where 𝑁such that 2𝑁>𝑠. Then (2.20), (2.21), (2.22), (2.23), and (2.24) characterize 𝐹𝑠𝑝,𝑞(𝑑).

Corollary 2.12. Let 𝑝,𝑞,𝑠 as in Theorem 2.6. Let further Φ0𝒮(𝑑) be a radial function such that Φ0 is non-increasing and atisfying Φ0(0)0,𝐷𝛼Φ0(0)=0(2.33) for 1|𝛼|1𝑅, where 𝑅+1>𝑠. Define Φ(𝑥)=Φ01(𝑥)2𝑑Φ0𝑥2.(2.34) Then (2.23), and (2.24) characterize 𝐹𝑠𝑝,𝑞(𝑑).

2.8. Proofs

We give the proof for Theorem 2.6 in full detail. The proof of Theorem 2.8 is more or less the same, even a bit simpler. We refer to the next paragraph for the necessary modifications. The proofs in the Besov scale are analogous, so we omit them completely. The strategy is a modification of Rychkov [4], where he proved the discrete case, that is, that (2.23) and (2.24) characterize 𝐹𝑠𝑝,𝑞(𝑑). However, Hansen [21, Remark 3.2.4] recently observed that the arguments used for proving (23) and (23′) in [4] are somehow problematic. The finiteness of the Peetre maximal function is assumed. But this is not true in general under the stated assumptions. Consider for instance in dimension 𝑑=1 the functionsΨ0(𝑡)=Ψ1(𝑡)=𝑒𝑡2(2.35) and, if 𝑎>0 is given, the tempered distribution 𝑓(𝑡)=|𝑡|𝑛 with 𝑎<𝑛. Then (Ψ𝑘𝑓)𝑎(𝑥) is infinite in every point 𝑥. The mentioned incorrect argument was inherited to some subsequent papers dealing with similar topics, for instance [2224]. Anyhow, the stated results hold true. An alternative strategy, in order to avoid the problematic Lemma  3 in [4], is given in Rychkov [7] as well as [8]. A variant of this method, which is originally due to Strömberg/Torchinsky [9, Chapter V], is also used in our proof below.

We start with a convolution-type inequality which will be often needed below. The following lemma is essentially [4, Lemma 2].

Lemma 2.13. Let 0<𝑝,𝑞 and, 𝛿>0. Let {𝑔𝑘}𝑘0 be a sequence of nonnegative measurable functions on 𝑑 and put 𝐺(𝑥)=𝑘2|𝑘|𝛿𝑔𝑘(𝑥),𝑥𝑑,.(2.36) Then there is some constant 𝐶=𝐶(𝑝,𝑞,𝛿), such that 𝐺𝑞𝐿𝑝𝑑𝑔𝐶𝑘𝑘𝑞𝐿𝑝𝑑,𝐺𝐿𝑝𝑞,𝑑𝑔𝐶𝑘𝑘𝐿𝑝𝑞,𝑑(2.37) hold true.

Proof of Theorem 2.6. The strategy of the proof is as follows. First, we prove the equivalence of the “continuous’’ characterizations (2.20) and (2.21). The next step is to build the bridge between the “continuous’’ (2.21) and the “discrete’’ characterization (2.23) and to change from the system (Φ0,Φ) to a system (Ψ0,Ψ). The equivalence of (2.23) and (2.24) goes parallel to (2.20) and (2.21). This was the original proof by Rychkov in [4]. So, up to this point, we have that (2.20), (2.21), (2.23), and (2.24) generate the same space for every chosen functions (Φ0,Φ) satisfying (2.17) and (2.18), namely, 𝐹𝑠𝑝,𝑞(𝑑). Indeed, Definition 2.4 can be seen as a special case of (2.24).Step 1. We are going to prove the following inequalities: 𝑓𝐹𝑠𝑝,𝑞2𝑓𝐹𝑠𝑝,𝑞1𝑓𝐹𝑠𝑝,𝑞2(2.38) for every 𝑓𝒮(𝑑).Substep 1.1. Put 𝜑0=Φ0 and 𝜑=(Φ)(2) if 1. Because of (2.17), it is possible to find functions 𝜓0,𝜓𝒮(𝑑) with supp𝜓0{𝜉𝑑|𝜉|2𝜀}, supp𝜓{𝜉𝑑𝜀/2|𝜉|2𝜀} and 𝜓(𝑥)=𝜓(2𝑥) such that 0𝜑(𝜉)𝜓(𝜉)=1.(2.39) We need a bit more. Fix a 1𝑡2. Clearly, we have also 0𝜑(𝑡𝜉)𝜓(𝑡𝜉)=1(2.40) for all 𝜉𝑑. With Ψ0=1𝜓0 and Ψ=1𝜓, we obtain then 𝑔=𝑚0Ψ𝑚𝑡Φ𝑚𝑡𝑔.(2.41) The 2 dilation gives then 𝑔=𝑚0Ψ𝑚𝑡2Φ𝑚𝑡2𝑔(2.42) for every 𝑔𝒮(𝑑), where 𝑔(𝜂)=𝑔(𝜂), 𝜂𝒮(𝑑). Obviously, we can rewrite (2.42) to obtain 𝑔=𝑚0Ψ𝑚𝑡2Φ𝑚𝑡2𝑔(2.43) for all 𝑔𝒮(𝑑). Let us now choose 𝑔=(Φ)𝑡𝑓. This gives the final version of the convolution identity Φ𝑡𝑓=𝑚0Φ𝑡Ψ𝑚𝑡2Φ𝑚𝑡2𝑓.(2.44) For 𝑚,0 we define Λ𝑚,2(𝑥)=𝑑Φ02𝑥Φ:𝑚=0,(𝑥):𝑚>0,𝑥𝑑.(2.45) Clearly, we have Φ𝑡Φ𝑚𝑡2=Λ𝑚,𝑡Φ𝑚+𝑡.(2.46) Plugging this into (2.44), we end up with the pointwise representation Φ𝑡(𝑓𝑦)=𝑚0Ψ𝑚2𝑡Λ𝑚,𝑡Φ𝑚+𝑡(=𝑓𝑦)𝑚0Ψ𝑚2𝑡Λ𝑚,𝑡Φ𝑚+𝑡=𝑓(𝑦)𝑚0𝑑Ψ𝑚2𝑡Λ𝑚,𝑡Φ(𝑦𝑧)𝑚+𝑡𝑓(𝑧)𝑑𝑧(2.47) for all 𝑦𝑑.Substep 1.2. Let us prove the following important inequality first. For every 𝑟>0 and every 𝑁0, we have ||Φ𝑡||𝑓(𝑥)𝑟𝑐𝑘02𝑘𝑁𝑟2(𝑘+)𝑑𝑑||Φ𝑘+𝑡||𝑓(𝑦)𝑟1+2||||𝑥𝑦𝑁𝑟𝑑𝑦,(2.48) where 𝑐 is independent of 𝑓𝒮(𝑑), 𝑥𝑑, and 0.
The representation (2.47) will be the starting point to prove (2.48). Namely, we have for 𝑦𝑑||Φ𝑡||𝑓(𝑦)𝑚0𝑑||Ψ𝑚𝑡2Λ𝑚,𝑡||||Φ(𝑦𝑧)𝑚+𝑡||𝑓(𝑧)𝑑𝑧𝑚0𝑆𝑚,,𝑡𝑑||Φ𝑚+𝑡||𝑓(𝑧)1+2||||𝑦𝑧𝑁𝑑𝑧,(2.49) where 𝑆𝑚,,𝑡=sup𝑥𝑑||Ψ𝑚2𝑡Λ𝑚,𝑡||(𝑥)1+2|𝑥|𝑁.(2.50) Elementary properties of the convolution yield (compare with (2.84)) 𝑆𝑚,,𝑡=2𝑑𝑡𝑑sup𝑥𝑑||||Ψ𝑚Λ𝑚,2𝑥2𝑡||||1+2|𝑥|𝑁=2𝑑𝑡𝑑sup𝑥𝑑||𝜓𝑚𝜂𝑚,||(𝑥)(1+|𝑡𝑥|)𝑁,(2.51) where 𝜂𝑚,Φ(𝑥)=Φ(𝑥):0,𝑚>0,0(𝑥):otherwise.(2.52) Lemma A.3 yields 𝑆𝑚,,𝑡𝑐𝑁2𝑑2𝑚𝑁,(2.53) which we put into (2.49) to obtain ||Φ𝑡||𝑓(𝑦)𝐶𝑁𝑚02𝑚𝑁𝑑2(𝑚+)𝑑||Φ𝑚+𝑡||𝑓(𝑧)1+2||||𝑦𝑧𝑁𝑑𝑧.(2.54) We prefer the strategy used by Rychkov in [7, Theorem 3.2] and [8, Lemma 2.9], which is a variant of the Strömberg/Torchinsky technique introduced in [9, Chapter V].
Let us continue by replacing by 𝑘+ in (2.54) and multiply on both sides with 2𝑘𝑁. Then we can estimate 2𝑘𝑁||Φ𝑘+𝑡||𝑓(𝑦)𝐶𝑁𝑚02𝑘𝑁2𝑚𝑁𝑑2(𝑚+𝑘+)𝑑||Φ𝑚+𝑘+𝑡||𝑓(𝑧)1+2𝑘+||||𝑦𝑧𝑁𝑑𝑧𝐶𝑁𝑚02(𝑚+𝑘)𝑁𝑑2(𝑚+𝑘+)𝑑||Φ𝑚+𝑘+𝑡||𝑓(𝑧)1+2||||𝑦𝑧𝑁𝑑𝑧=𝐶𝑁𝑚𝑘+02𝑚𝑁𝑑2(𝑚+)𝑑||Φ𝑚+𝑡||𝑓(𝑧)1+2||||𝑦𝑧𝑁𝑑𝑧,(2.55)𝐶𝑁𝑚02𝑚𝑁𝑑2(𝑚+)𝑑||Φ𝑚+𝑡||𝑓(𝑧)1+2||||𝑦𝑧𝑁𝑑𝑧.(2.56) Next, we apply the elementary inequalities 1+2||||𝑦𝑧1+2||||𝑥𝑦1+2,||Φ|𝑥𝑧|𝑚+𝑡||||Φ𝑓(𝑧)𝑚+𝑡||𝑓(𝑧)𝑟1+2|𝑥𝑧|𝑁(1𝑟)×sup𝑦𝑑||Φ𝑚+𝑡||𝑓(𝑦)1𝑟1+2||||𝑥𝑦𝑁(1𝑟),(2.57) where 0<𝑟1. We define the maximal function 𝑀,𝑁(𝑥,𝑡)=sup𝑘0sup𝑦𝑑2𝑘𝑁||Φ𝑘+𝑡||𝑓(𝑦)1+2||||𝑥𝑦𝑁,𝑥𝑑,(2.58) and estimate 𝑀,𝑁(𝑥,𝑡)𝐶𝑁𝑚02𝑚𝑁𝑑2(𝑚+)𝑑||Φ𝑚+𝑡||𝑓(𝑧)1+2|𝑥𝑧|𝑁𝑑𝑧(2.59)𝐶𝑁𝑚02𝑚𝑁𝑟2𝑚𝑁sup𝑦𝑑||Φ𝑚+𝑡||𝑓(𝑦)1+2||||𝑥𝑦𝑁1𝑟×𝑑2(𝑚+)𝑑||Φ𝑚+𝑡||𝑓(𝑧)𝑟1+2|𝑥𝑧|𝑁𝑟𝑑𝑧.(2.60) Observe that we can estimate the term ()1𝑟 in the right-hand side of (2.60) by 𝑀,𝑁(𝑥,𝑡)1𝑟. Hence, if 𝑀,𝑁(𝑥,𝑡)< we obtain from (2.60) 𝑀,𝑁(𝑥,𝑡)𝑟𝐶𝑁𝑚02𝑚𝑁𝑟𝑑2(𝑚+)𝑑||Φ𝑚+𝑡||𝑓(𝑧)𝑟1+2|𝑥𝑧|𝑁𝑟𝑑𝑧,(2.61) where 𝐶𝑁 is independent of 𝑥, 𝑓, , and 𝑡[1,2]. We claim that there exists 𝑁𝑓0 such that 𝑀,𝑁(𝑥,𝑡)< for all 𝑁𝑁𝑓. Indeed, we use that 𝑓𝒮(𝑑), that is, there is an 𝑀0 and 𝑐𝑓>0 such that ||Φ𝑘+𝑡||𝑓(𝑦)𝑐𝑓sup|𝛼|1𝑀sup𝑧𝑑||𝐷𝛼Φ𝑘+||||||(𝑧)1+𝑦𝑧𝑀.(2.62) Assuming 𝑁>𝑀, we estimate as follows: ||Φ𝑡||𝑓(𝑥)𝑀,𝑁(𝑥,𝑡)𝑐sup𝑘0sup𝑦𝑑2𝑘𝑁||Φ𝑘+𝑡||𝑓(𝑦)||||1+𝑥𝑦2𝑁/2𝑐sup𝑘0sup𝑦𝑑2𝑘𝑁2(𝑘+)(𝑀+𝑑)sup𝑧𝑑sup|𝛼|1𝑀||𝐷𝛼𝛾𝑘+(||||||𝑧)1+𝑦𝑧𝑀||||1+𝑥𝑦𝑁𝑐2(𝑀+𝑑)sup𝑘0sup𝑧𝑑sup|𝛼|1𝑀||𝐷𝛼𝛾𝑘+||(𝑧)(1+|𝑥𝑧|)𝑁,(2.63) where we again used the inequality (compare with (2.57)) ||||||||1+𝑦𝑧1+𝑥𝑦(1+|𝑥𝑧|)(2.64) and have set 𝛾Φ(𝑡)=0(𝑡)Φ(𝑡)=0,>0.(2.65) Hence 𝛾𝑘+ gives us only two different functions from 𝒮(𝑑). This implies the boundedness of 𝑀,𝑁(𝑥,𝑡) for 𝑥𝑑 if 𝑁>𝑀=𝑁𝑓. Therefore, (2.61) together with (2.63) yield (2.48) with 𝑐=𝐶𝑁, independent of 𝑥, 𝑓, and , for all 𝑁𝑁𝑓. But this is not yet what we want. Observe that the right-hand side of (2.48) decreases as 𝑁 increases. Therefore, we have (2.48) for all 𝑁0 but with 𝑐=𝑐(𝑓)=𝐶𝑁𝑓  depending on 𝑓. This is still not yet what we want. Now we argue as follows: starting with (2.48) where 𝑐=𝑐(𝑓) and 𝑁0 arbitrary, we apply the same arguments as used from (2.55) to (2.56), switch to the maximal function (2.58) with the help of (2.57), and finish with (2.61) instead of (2.59) but with a constant that depends on 𝑓. But this does not matter now. It is Important, that a finite right-hand side of (2.61) (which is the same as rhs(2.48)) implies 𝑀,𝑁(𝑥,𝑡)<.
We assume rhs (2.48) <. Otherwise, there is nothing to prove in (2.48). Returning to (2.60) and having in mind that now 𝑀,𝑁(𝑥,𝑡)<, we end up with (2.61) for all 𝑁 and 𝐶𝑁 independent of 𝑓. Finally, from (2.61) we obtain (2.48) and are done in case 0<𝑟1.
Of course, (2.48) also holds true for 𝑟>1 with a much simpler proof. In that case, we use (2.54) with 𝑁+1 instead of 𝑁 and apply Hölder’s inequality with respect to 1/𝑟+1/𝑟=1 first for integrals and then for sums.
Substep 1.3. The inequality (2.48) implies immediately a stronger version of itself. Using (2.57) again, we obtain for 𝑎𝑁 and Φ2𝑡𝑓𝑎(𝑥)𝑟𝑐𝑘02𝑘𝑁𝑟2(𝑘+)𝑑𝑑||Φ𝑘+𝑡||𝑓(𝑦)𝑟1+2||||𝑥𝑦𝑎𝑟𝑑𝑦.(2.66) We proved that the inequality (2.66) holds for all 𝑡[1,2] where 𝑐>0 is independent of 𝑡. If we choose 𝑟<min{𝑝,𝑞}, we can apply the norm 21||𝑞/𝑟𝑑𝑡𝑡𝑟/𝑞,(2.67) on both sides and use Minkowski’s inequality for integrals, which yields 21|||Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡𝑟/𝑞𝑐𝑘02𝑘𝑁𝑟2(𝑘+)𝑑𝑑21|Φ𝑘+𝑡𝑓(𝑦)|𝑞(𝑑𝑡/𝑡)𝑟/𝑞1+2||||𝑥𝑦𝑎𝑟𝑑𝑦.(2.68) If 𝑎𝑟>𝑑,then we have 𝑔2(𝑦)=𝑑1+2||𝑦||𝑎𝑟𝐿1𝑑,(2.69) and we observe 21|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡𝑟/𝑞𝑐𝑘02𝑘𝑁𝑟2𝑘𝑑2𝑠𝑟𝑔21|2𝑠Φ𝑘+𝑡𝑓()|𝑞𝑑𝑡𝑡𝑟/𝑞(𝑥).(2.70) Now we use the majorant property of the Hardy-Littlewood maximal operator (see Section 2.2 and [25, Chapter 2]) and continue estimating 21|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡𝑟/𝑞𝑐𝑘02𝑟𝑠2𝑘(𝑁𝑟+𝑑)𝑀21||Φ𝑘+𝑡||𝑓()𝑞𝑑𝑡𝑡𝑟/𝑞(𝑥).(2.71) An index shift on the right-hand side gives 21|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡𝑟/𝑞=𝑐𝑘+02𝑟𝑠2(𝑘)(𝑁𝑟+𝑑)𝑀21||Φ𝑘𝑡||𝑓()𝑞𝑑𝑡𝑡𝑟/𝑞(𝑥)=𝑐𝑘+02(𝑘)(𝑁𝑟𝑑+𝑟𝑠)2𝑘𝑟𝑠𝑀21||Φ𝑘𝑡||𝑓()𝑞𝑑𝑡𝑡𝑟/𝑞(𝑥).(2.72) Choose now 1/𝑎<𝑟<min{𝑝,𝑞}, 𝑁>max{0,𝑠}+𝑎 and put 𝑑𝛿=𝑁+𝑠𝑟>0.(2.73) We obtain for 21|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡𝑟/𝑞𝑐𝑘2𝛿𝑟|𝑘|2𝑘𝑟𝑠𝑀21||Φ𝑘𝑡||𝑓()𝑞𝑑𝑡𝑡𝑟/𝑞(𝑥).(2.74) Now we apply Lemma 2.13 in 𝐿𝑝/𝑟(𝑞/𝑟,𝑑), which yields 21|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡𝑟/𝑞𝐿𝑝/𝑟𝑞/𝑟𝑀𝑐21||2𝑘𝑠Φ𝑘𝑡||𝑓()𝑞𝑑𝑡𝑡𝑟/𝑞𝐿𝑝/𝑟𝑞/𝑟.(2.75) The Fefferman-Stein inequality (see Section 2.2 /Theorem 2.1, having in mind that 𝑝/𝑟,𝑞/𝑟>1) gives 21|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑞𝑟=𝑀21|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡𝑟/𝑞𝐿𝑝/𝑟𝑞/𝑟21|2𝑘𝑠Φ𝑘𝑡𝑓()|𝑞𝑑𝑡𝑡𝑟/𝑞𝑞/𝑟𝑞/𝑟=21||2𝑘𝑠Φ𝑘𝑡||𝑓()𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑞𝑟.(2.76) Hence, we obtain 10||𝜆𝑠𝑞Φ𝜆𝑓𝑎||(𝑥)𝑞𝑑𝜆𝜆1/𝑞𝐿𝑝𝑑=121|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝=21|||2𝑠Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑞10||𝜆𝑠𝑞Φ𝜆||𝑓(𝑥)𝑞𝑑𝜆𝜆1/𝑞𝐿𝑝𝑑.(2.77) This proves 𝑓𝐹𝑠𝑝,𝑞(𝑑)2𝑓𝐹𝑠𝑝,𝑞(𝑑)1. Since the reverse inequality is trivial, this finishes Step 1.Step 2. Let Ψ0,Ψ𝒮(𝑑) be functions satisfying (2.18). In fact, the condition (2.17) for Ψ0,Ψ is not necessary for what follows.Substep 2.1. We are going to prove 𝑓𝐹𝑠𝑝,𝑞𝑑Ψ4𝑓𝐹𝑠𝑝,𝑞𝑑Φ2(2.78) for all 𝑓𝒮(𝑑).We decompose 𝑓 similar as in Step 1. Exploiting the property (2.17) for the system Φ, we find 𝒮(𝑑)-functions 𝜆0,𝜆𝒮(𝑑)such that supp𝜆0{𝜉𝑑|𝜉|2𝜀} and supp𝜆{𝜉𝑑𝜀/2|𝜉|2𝜀} and 0𝜆(𝑡𝑥)𝜑(𝑡𝑥)=1(2.79) for 𝑥𝑑 and 𝑡[1,2] fix. Putting Λ0=𝜆0 and Λ=Λ, we obtain the decomposition 𝑔=0Λ𝑡Φ𝑡𝑔(2.80) for every 𝑔𝒮(𝑑). We put 𝑔=Ψ𝑓 and see Ψ𝑓=𝑘0ΨΛ𝑘𝑡Φ𝑘𝑡𝑓.(2.81) Now, we estimate as follows: ||ΨΛ𝑘𝑡Φ𝑘𝑡||𝑓(𝑦)𝑑||ΨΛ𝑘𝑡||||Φ(𝑧)𝑘𝑡||Φ𝑓(𝑦𝑧)𝑑𝑧2𝑘𝑡𝑓𝑎(𝑦)𝑑||ΦΛ𝑘𝑡||(𝑧)1+2𝑘|𝑧|𝑎Φ𝑑𝑧2𝑘𝑡𝑓𝑎(𝑦)𝐽,𝑘,(2.82) where 𝐽,𝑘=𝑑||ΨΛ𝑘𝑡(||𝑧)1+2𝑘|𝑧|𝑎𝑑𝑧.(2.83) We first observe that for 𝑥𝑑 and functions 𝜇,𝜂𝒮(𝑑), the following identity holds true for 𝑢,𝑣>0𝜇𝑢𝜂𝑣1(𝑧)=𝑢𝑑𝜇𝜂𝑣/𝑢𝑧𝑢=1𝑣𝑑𝜇𝑢/𝑣𝑧𝜂𝑣.(2.84) This yields in case 𝑘 (with a minor change if 𝑘=0 or =0) 𝐽,𝑘=𝑑||Ψ𝑘1/𝑡(||(Λ𝑧)1+|𝑡𝑧|)𝑎sup𝑧𝑑||Ψ𝑘1/𝑡||Λ(𝑧)(1+|𝑧|)𝑎+𝑑+12(𝑘)(𝑅+1),(2.85) where we used Lemma A.3 for the last estimate.
If 𝑘>𝑙, we change the roles of Ψ and Λ to obtain again with Lemma A.3𝐽,𝑘=𝑑||ΛΨ𝑘(||||2𝑧)1+𝑘𝑧||𝑎𝑑𝑥2(𝑘)𝑎sup𝑧𝑑||ΛΨ𝑘𝑡||(𝑧)(1+|𝑧|)𝑎+𝑑+12(𝑘)(𝐿+1𝑎),(2.86) where 𝐿 can be chosen arbitrary large since Λ satisfies (𝑀𝐿) for every 𝐿according to its construction. Let us further use the estimate Ψ𝑘𝑓𝑎Ψ(𝑦)𝑘𝑓𝑎(𝑥)1+2𝑘||||𝑥𝑦𝑎Ψ𝑘𝑓𝑎(𝑥)1+2||||𝑥𝑦𝑎max1,2(𝑘)𝑎.(2.87)
Consequently, sup𝑦𝑑2𝑠||ΨΛ𝑘𝑡Φ𝑘𝑡||𝑓(𝑦)1+2||||𝑥𝑦𝑎2𝑘𝑠Φ2𝑘𝑡𝑓𝑎(𝑥)2(𝑘)𝑠max1,2(𝑘)𝑎𝐽,𝑘2𝑘𝑠Φ2𝑘𝑡𝑓𝑎2(𝑥)(𝑘)(𝐿+1𝑎+𝑠)2𝑘>𝑙,(𝑘)(𝑅+1𝑠)𝑘.(2.88) Plugging this into (2.81) and choosing 𝐿𝑎+|𝑠| and 𝛿=min{1,𝑅+1𝑠}, we obtain the inequality 2𝑠Ψ𝑓𝑎(𝑥)𝑘02|𝑘|𝛿2𝑘𝑠Φ2𝑘𝑡𝑓𝑎(𝑥)(2.89) for all 𝑥𝑑 and all 𝑡[1,2]. Suppose first that 𝑞1. Then we take on both sides (21||𝑞𝑑𝑡/𝑡)1/𝑞, which gives 2𝑠Ψ𝑓𝑎(𝑥)𝑘02|𝑘|𝛿2𝑘𝑠21|||Φ2𝑘𝑡𝑓𝑎(|||𝑥)𝑞𝑑𝑡𝑡1/𝑞.(2.90) Applying Lemma 2.13 yields 2𝑠Ψ𝑓𝑎(𝑥)𝐿𝑝𝑞,𝑑𝑘=12𝑘𝑠𝑞21|||Φ2𝑘𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑑,(2.91) which gives the desired result.
In case 𝑞<1, we argue as follows. The quantity (21||𝑞𝑑𝑡/𝑡)1/𝑞 is not longer a norm, but a 𝑞-norm. This gives 2𝑠Ψ𝑓𝑎(𝑥)𝑞𝑘02|𝑘|𝛿𝑞2𝑘𝑠𝑞21|||Φ2𝑘𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡.(2.92) Notice that the right-hand side is nothing more than a convolution (𝛾𝛽) of the sequences 𝛾𝑘=2|𝑘|𝛿𝑞,𝛽𝑘=2𝑘𝑠𝑞21|||Φ2𝑘𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡.(2.93) Now we apply the 1-norm to both sides and get for all 𝑥𝑑2𝑠Ψ𝑓𝑎(𝑥)𝑞𝑞𝛾1𝛽1𝑘=12𝑘𝑠𝑞21|||Φ2𝑘𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡.(2.94) We take both sides to the power ()1/𝑞 and apply the 𝐿𝑝(𝑑)-norm. This gives (2.78).
Substep 2.2. With similar arguments and obvious modifications of Substep 2.1, we obtain for all 𝑓𝒮(𝑑)𝑓𝐹𝑠𝑝,𝑞𝑑Ψ2𝑓𝐹𝑠𝑝,𝑞𝑑Φ4.(2.95)Step 3. Choosing 𝑡=1 in Step 1 and omitting the integration over 𝑡, we see immediately 𝑓𝐹𝑠𝑝,𝑞𝑑5𝑓𝐹𝑠𝑝,𝑞𝑑4𝑓𝐹𝑠𝑝,𝑞𝑑5.(2.96)Step 4. What remains is to show that (2.22) is equivalent to the rest.Substep 4.1. Let us prove 𝑓𝐹𝑠𝑝,𝑞𝑑2𝑓𝐹𝑠𝑝,𝑞𝑑3.(2.97) We return to (2.66) in Substep 1.3. If |𝑧|<2(+𝑘)𝑡, formula (2.66) implies by shift in the integral the following: Φ2𝑡𝑓𝑎(𝑥)𝑟𝐶𝑁𝑘02𝑘𝑁𝑟2(𝑘+)𝑑𝑑||Φ𝑘+𝑡||𝑓(𝑦+𝑧)𝑟1+2||||𝑥𝑦𝑎𝑟𝑑𝑦.(2.98) Indeed, we have 1+2||||𝑥𝑦1+2||||𝑥(𝑦+𝑧)+|𝑧|1+2||||𝑥(𝑦+𝑧)+2𝑘1+2||||,𝑥(𝑦+𝑧)(2.99) where the last estimate follows from the fact that 𝑘0 in the sum. Instead of the integral (21||𝑞/𝑟𝑑𝑡/𝑡)𝑟/𝑞 (see Step 3), we take now on both sides of (2.98) the norm 21|𝑧|<𝑡||𝑞/𝑟𝑑𝑧𝑑𝑡𝑡𝑑+1𝑟/𝑞.(2.100) The integration over 𝑧 does not influence the left-hand side. Instead of (2.68), we obtain 21|||Φ2𝑡𝑓𝑎|||(𝑥)𝑞𝑑𝑡𝑡𝑟/𝑞𝑐𝑘02𝑘𝑁𝑠2(𝑘+)𝑑𝑑21|𝑧|<𝑡||Φ𝑘+𝑡||𝑓(𝑦)𝑞𝑑𝑧𝑑𝑡/𝑡𝑑+1𝑟/𝑞1+2||||𝑥𝑦𝑎𝑟𝑑𝑦.(2.101) We continue with analogous arguments as after (2.68) and end up with (2.97).Substep 4.2. We prove 𝑓𝐹𝑠𝑝,𝑞(𝑑)3𝑓𝐹𝑠𝑝,𝑞(𝑑)2. Indeed, it is easy to see, that we have for all 𝑡>01𝑡𝑑|𝑧|<𝑡||Φ𝑡||𝑓(𝑥+𝑧)𝑑𝑧sup|𝑧|<𝑡||Φ𝑡||𝑓(𝑥+𝑧)(1+1/𝑡|𝑧|)𝑎Φ𝑡𝑓𝑎(𝑥),(2.102) and we are done. The proof is complete.

Proof of Theorem 2.8. The proof of Theorem 2.8 works almost analogously to the previous one. It is even a bit simpler, since we do not have to deal with a separate function Φ0, which causes some technical difficulties. However, there are still some technical obstacles which have to be discussed. Although we are in the homogeneous world, we use the same decomposition as used in (2.43), even with the inhomogeneity Φ0. In the definition of Λ𝑚,(𝑥) in (2.45), we have to add Φ(𝑥) if =0 and 𝑚>0. The consequence is (2.47) for every , where Φ0=Φ. Hence, the inhomogeneity is buried in Λ𝑚,. This yields (2.66) for all , where 𝑘 still runs trough 0. We need this for the argument in Substep 4.1. In contrast to the previous decomposition, we use (2.80) now for , where Φ0=Φ, Λ0=Λ, and 𝑘. This works since we assume 𝑔𝒮0(𝑑). The rest is clear.

Proof of Corollaries 2.11 and 2.12. Corollary 2.11 is more or less clear. We know that Δ𝑁 gives (𝑑𝑘=1|𝜉𝑘|2)𝑁 as factor on the Fourier side. This gives (2.18) immediately, and together with (2.31) we have (2.17) for 𝜀>0 small enough.
In the case of Corollary 2.12, the situation is a bit delicate. Clearly, Condition (2.18) holds true. But the problem here is, that (2.17) may be violated for all 𝜀>0. However, we argue as follows. In Step 2 in the proof above we have seen, that we do not need (2.17) for the system (Ψ0,Ψ). Hence, we can estimate (2.20), (2.21), (2.22), (2.23), and (2.24) from above by a different characterization of 𝐹𝑠𝑝,𝑞(𝑑). For the remaining estimates we apply Theorem 2.6 with the system (Φ0,Φ), where Φ=Φ01(𝑥)2𝑘𝑑Φ0𝑥2𝑘(2.103) and 𝑘is chosen in such a way that (2.17) is satisfied. What remains is a simple consequence of the fact that Φ=Φ+Φ1++Φ(𝑘1)(2.104) and the triangle inequality. This type of argument is due to Triebel [3, Section 3.3.3].

3. Classical Coorbit Space Theory

In [1113, 15], a general theory of Banach spaces related to integrable group representations was developed. The ingredients are a locally compact group 𝒢 with identity 𝑒, a Hilbert space , and an irreducible, unitary, and continuous representation 𝜋 on , which is at least integrable. Then one can associate a Banach space Co𝑌 to any solid, translation-invariant Banach space 𝑌 of functions on the group 𝒢. The main achievement of this abstract theory is a powerful discretization machinery for Co𝑌, that is, a universal approach to atomic decompositions and Banach frames. It allows to transfer certain questions concerning Banach space or interpolation theory from the function space to the associated sequence space level, see [12, 13, 26]. In connection with smoothness spaces of Besov-Lizorkin-Triebel type, the philosophy of this approach is to measure smoothness of a function in decay properties of the continuous wavelet transform 𝑊𝑔𝑓, see the appendix for details. Indeed, homogeneous Besov and Lizorkin-Triebel-type spaces turn out to be coorbits of properly chosen spaces 𝑌 on the 𝑎𝑥+𝑏-group 𝐺.

The are many more examples according to this abstract theory. One main class of examples refers to the Heisenberg group and the short-time Fourier transform and leads to the well-known modulation spaces as coorbits of weighted 𝐿𝑝() spaces, see [11, Section 7.1] and also [27].

3.1. Function Spaces on 𝒢

Integration on 𝒢 will always be with respect to the left Haar measure 𝑑𝜇(𝑥). The Haar module on 𝒢 is denoted by Δ. We define further 𝐿𝑥𝐹(𝑦)=𝐹(𝑥1𝑦) and 𝑅𝑥𝐹(𝑦)=𝐹(𝑦𝑥), 𝑥,𝑦𝒢, the left and right translation operators. A Banach function space 𝑌 on the group 𝒢 is supposed to have the following properties:(i)𝑌 is continuously embedded in 𝐿1loc(𝒢), (ii)𝑌 is invariant under left and right translation 𝐿𝑦 and 𝑅𝑦, which represent continuous operators on 𝑌,(iii) 𝑌 is solid, that is, 𝐻𝑌 and |𝐹(𝑥)||𝐻(𝑥)| a.e. imply 𝐹𝑌 and 𝐹𝑌𝐻|𝑌.

The continuous weight 𝑤 is called submultiplicative if 𝑤(𝑥𝑦)𝑤(𝑥)𝑤(𝑦) for all 𝑥,𝑦𝒢. Further, another weight 𝑚 is called 𝑤-moderate if 𝑚(𝑥𝑦𝑧)𝑤(𝑥)𝑚(𝑦)𝑤(𝑧), 𝑥,𝑦,𝑧𝒢. The space 𝐿𝑤𝑝(𝒢) of functions 𝐹 on the group 𝒢 is defined via the norm𝐹𝐿𝑤𝑝=(𝒢)𝒢||||𝐹(𝑥)𝑤(𝑥)𝑝𝑑𝜇(𝑥)1/𝑝,(3.1) where 1𝑝 (modification if 𝑝=). If 𝑤1, then we simply write 𝐿𝑝(𝒢). These spaces provide left and right translation invariance, which is easy to show. Later, in Section 4.1, we are going to introduce certain mixed norm spaces where the translation invariance is not longer automatic.

3.2. Sequence Spaces

Definition 3.1. Let 𝑋={𝑥𝑖}𝑖𝐼 be some discrete set of points in 𝒢 and 𝑉 a relatively compact neighborhood of 𝑒𝒢. (i) 𝑋 is called 𝑉-dense if 𝒢=𝑖𝐼𝑥𝑖𝑉.(ii) 𝑋 is called relatively separated if for all compact sets 𝐾𝒢,there exists a constant 𝐶𝐾 such that sup𝑗𝐼𝑖𝐼𝑥𝑖𝐾𝑥𝑗𝐾𝐶𝐾.(3.2) (iii) 𝑋 is called 𝑉-well-spread (or simply well-spread) if it is both relatively separated and 𝑉-dense for some 𝑉.

Definition 3.2. For a family 𝑋={𝑥𝑖}𝑖𝐼 which is V-well-spread with respect to a relatively compact neighborhood 𝑉 of 𝑒𝒢 we define the sequence space 𝑌𝑏 and 𝑌 associated to 𝑌 as 𝑌𝑏=𝜆𝑖𝑖𝐼𝜆𝑖𝑖𝐼𝑌𝑏=𝑖𝐼||𝜆𝑖||||𝑥𝑖𝑉||1𝜒𝑥𝑖𝑉,𝑌𝑌<=𝜆𝑖𝑖𝐼𝜆𝑖𝑖𝐼𝑌=𝑖𝐼||𝜆𝑖||𝜒𝑥𝑖𝑉.𝑌<(3.3)

Remark 3.3. For a well-spread family 𝑋, the spaces 𝑌𝑏 and 𝑌 do not depend on the choice of 𝑉, that is, different sets 𝑉 define equivalent norms on 𝑌𝑏 and 𝑌, respectively. For more facts on these sequence spaces, confer [12].

3.3. Coorbit Spaces

The starting point is a Hilbert space and an integrable, irreducible, unitary and continuous representation 𝜋 of 𝒢 on . Then the general voice transform 𝑉𝑔𝑓 is a function on the group 𝒢 given by𝑉𝑔𝑓(𝑥)=𝜋(𝑥)𝑔,𝑓,(3.4) where the brackets denote the inner product in .

The continuous wavelet transform 𝑊𝑔𝑓 (appendix) is a voice transform for the situation =𝐿2(𝑑) and the 𝑎𝑥+𝑏-group.

Definition 3.4. For a weight 𝑤()1 on 𝒢, we define the space 𝐴𝑤 of admissible vectors by 𝐴𝑤=𝑔𝑉𝑔𝑔𝐿𝑤1(𝒢).(3.5) If 𝐴𝑤{0} and 𝑔𝐴𝑤, we define further 1𝑤𝑑=𝑓𝑓1𝑤=𝑉𝑔𝑓𝐿𝑤1(𝒢)<.(3.6) Finally, we denote by (1𝑤) the canonical antidual of 1𝑤, that is, the space of conjugate linear functionals on 1𝑤.

We see immediately that 𝐴𝑤1𝑤. The voice transform (3.4) can now be extended to 𝑤×(1𝑤) by the usual dual pairing. The space 1𝑤 can be considered as the space of test functions, whereas the space (1𝑤) serves as reservoir, that is, distributions.

Let now 𝑌 be a space on 𝒢 such that (i)–(iii) in Section 3.1 hold true. We define further𝑤𝑌𝐿(𝑥)=max𝑥,𝐿𝑥1,𝑅𝑥𝑥,Δ1𝑅𝑥1,𝑥𝒢,(3.7) where the operator norms are considered from 𝑌 to 𝑌.

Definition 3.5. Let 𝑌 be a space on 𝒢 satisfying (i)–(iii) and 𝑤𝑌(𝑥) be given by (3.7). Let further 𝑔𝐴𝑤. We define the space Co𝑌, called coorbit space of Y, through C𝑜𝑌=𝑓1𝑤𝑉𝑔𝑉𝑓𝑌with𝑓Co𝑌=𝑔𝑓𝑌.(3.8)
The following result is of crucial importance. See [14] for details.

Lemma 3.6 (Correspondence principle). Let 𝐺=𝑉𝑔𝑔𝐿𝑤1(𝒢) for a fixed analyzing vector 𝑔𝐴𝑤 with proper normalization. Then a function 𝐹𝑌 is of the form 𝑉𝑔𝑓 for some 𝑓C𝑜𝑌 if and only if 𝐹(𝑥)=(𝐹𝐺)(𝑥)=𝒢𝐹(𝑦)𝐿𝑦𝐺(𝑥)𝑑𝜇(𝑦),𝑥𝒢.(3.9) In other words, Co𝑌 is isometrically isomorphic to the closed subspace 𝑌𝐺 of 𝑌.

The following basic properties are proved for instance, in [28, Theorem 4.5.13].

Theorem 3.7. (i) The space C𝑜𝑌 is independent of the analyzing vector 𝑔𝐴𝑤.
(ii) The definition of the space Co𝑌 is independent of the reservoir in the following sense. Assume that 𝑆1𝑤 is a nontrivial locally convex vector space, which is invariant under 𝜋. Assume further that there exists a nonzero vector 𝑔𝑆𝐴𝑤 for which the reproducing formula 𝑉𝑔𝑓=𝑉𝑔𝑔𝑉𝑔𝑓(3.10) holds for all 𝑓𝑆. Then we have Co𝑌=𝑓1𝑤𝑉𝑔=𝑓𝑌𝑓𝑆𝑉𝑔𝑓𝑌.(3.11)

3.4. Discretizations

This section collects briefly the basic facts concerning atomic (frame) decompositions in coorbit spaces. We are interested in atoms of type {𝜋(𝑥𝑖)𝑔}𝑖𝐼, where {𝑥𝑖}𝑖𝐼𝒢 represents a discrete subset, whereas 𝑔 denotes a fixed admissible analyzing vector.

Definition 3.8. A family {𝑔𝑖}𝑖𝐼 in a Banach space 𝐵 is called an atomic decomposition for B if there exists a family of bounded linear functionals {𝜆𝑖}𝑖𝐼𝐵 (not necessary unique) and a Banach sequence space 𝐵=𝐵(𝐼) such that (a){𝜆𝑖(𝑓)}𝑖𝐼𝐵 for all 𝑓𝐵 and there exists a constant 𝐶1>0 with𝜆𝑖(𝑓)𝑖𝐼𝐵𝐶1𝑓𝐵.(3.12)(b)For all 𝑓𝐵, we have 𝑓=𝑖𝐼𝜆𝑖(𝑓)𝑔𝑖(3.13) in some suitable topology. (c) If {𝜆𝑖}𝑖𝐼𝐵, then 𝑖𝐼𝜆𝑖𝑔𝑖𝐵 and there exists a constant 𝐶2>0 such that𝑖𝐼𝜆𝑖𝑔𝑖𝐵𝐶2𝜆𝑖𝑖𝐼𝐵.(3.14)

Definition 3.9. A family {𝑖}𝑖𝐼𝐵 is called a Banach frame for 𝐵 if there exists a Banach sequence space 𝐵𝑏=𝐵𝑏(𝐼) and a linear bounded reconstruction operator Θ𝐵𝑏𝐵 such that(a){𝑖(𝑓)}𝑖𝐼𝐵𝑏 for all 𝑓𝐵 and there exist constants 𝐶1,𝐶2 such that𝐶1𝑓𝐵𝑖(𝑓)𝑖𝐼𝐵𝑏𝐶2𝑓𝐵,(3.15)(b)Θ({𝑖(𝑓)}𝑖𝐼)=𝑓.

Remark 3.10. This setting differs slightly from the understanding of Triebel in [3, 29].

The following abstract result for the atomic decomposition in Co𝑌 is due to Feichtinger and Gröchenig (see [12, Theorem 6.1]).

Theorem 3.11. Let 𝑌 be a function space on the group 𝒢 which satisfies the hypotheses (i)–(iii). Let further 𝑤𝑌𝐿(𝑥)=max𝑥,𝐿𝑥1,𝑅𝑥𝑥,Δ1𝑅𝑥1,(3.16) where the operator norms are taken from 𝑌 to𝑌, and 𝑔𝐴𝑤 such that 𝒢sup𝑦𝑥𝑉||||𝜋(𝑦)𝑔,𝑔𝑤(𝑥,𝑡)𝑑𝜇(𝑥)<.(3.17) Then there exists a neighborhood 𝑈 of 𝑒𝒢 and constants 𝐶0,𝐶1>1 such that for every 𝑈-well-spread discrete set 𝑋={𝑥𝑖}𝑖𝐼𝒢, the following is true.(i) (Analysis) Every 𝑓Co𝑌 has a representation 𝑓=𝑖𝐼𝜆𝑖𝜋𝑥𝑖𝑔(3.18) with coefficients {𝜆𝑖}𝑖𝐼 depending linearly on 𝑓 and satisfying the estimate 𝜆𝑖𝑖𝐼𝑌𝐶0𝑓𝑌.(3.19)(ii) (Synthesis) Conversely, for any sequence {𝜆𝑖}𝑖𝐼𝑌, the element 𝑓=𝑖𝐼𝜆𝑖𝜋(𝑥𝑖)𝑔 is in Co𝑌 and one has 𝑓Co𝑌𝐶1𝜆𝑖𝑖𝐼𝑌.(3.20) In both cases, convergence takes place in the norm of Co𝑌 if the finite sequences are norm dense in 𝑌, and in the weak-sense of (1𝑤) otherwise.

Remark 3.12. According to Definition 3.8, the family {𝜋(𝑥𝑖)𝑔}𝑖𝐼 represents an atomic decomposition for Co𝑌.

Theorem 3.13. Under the same conditions as in Theorem 3.11, the system {𝜋(𝑥𝑖)𝑔}𝑖𝐼 represents a Banach frame for Co𝑌, that is, 𝜋𝑥𝑓Co𝑌𝑖𝑔,𝑓𝑌𝑏.(3.21)

The following powerful result goes back to Gröchenig [14] and was generalized by Rauhut [17].

Theorem 3.14. Suppose that the functions 𝑔𝑟,𝛾𝑟, 𝑟=1,,𝑛, satisfy (3.17). Let 𝑋={𝑥𝑖}𝑖𝐼 be a well-spread set such that 𝑓=𝑛𝑟=1𝑖𝐼𝜋𝑥𝑖𝛾𝑟𝜋𝑥,𝑓𝑖𝑔𝑟(3.22) for all 𝑓𝐻. Then expansion (3.22) extends to all 𝑓Co𝑌. Moreover, 𝑓(1𝑤) belongs to Co𝑌 if and only if {𝜋(𝑥𝑖)𝛾𝑟,𝑓}𝑖𝐼 belongs to 𝑌 for each 𝑟=1,,𝑛. The convergence is considered in Co𝑌 if the finite sequences are dense in 𝑌. In general, we have weak-convergence.

Proof. The proof of this result relies on the fact that there exists an atomic decomposition {𝜋(𝑦𝑖)𝑔}𝑖𝐼 by Theorem 3.11 with a certain 𝑔 satisfying (3.17) and a corresponding sequence of points 𝑍={𝑦𝑖}𝑖𝐼. This has to be combined with Theorem 3.13 and Theorem 3.11/(ii) and we are done. See [14] for the details.

4. Coorbit Spaces on the 𝑎𝑥+𝑏-Group

Let 𝒢=𝑑+ the 𝑑-dimensional 𝑎𝑥+𝑏-group. Its multiplication is given by(𝑥,𝑡)(𝑦,𝑠)=(𝑥+𝑡𝑦,𝑠𝑡).(4.1)

The left Haar measure 𝜇 on 𝒢 is given by 𝑑𝑡/𝑡𝑑+1, the Haar module is Δ(𝑥,𝑡)=𝑡𝑑. Giving a function 𝐹 on 𝒢, the left and right translation 𝐿𝑦=𝐿(𝑦,𝑟) and 𝑅𝑦=𝑅(𝑦,𝑟) are given by𝐿(𝑦,𝑟)𝐹(𝑥,𝑡)=𝐹(𝑦,𝑟)1(𝑥,𝑡)=𝐹𝑥𝑦𝑟,𝑡𝑟,𝑅(𝑦,𝑟)𝐹(𝑥,𝑡)=𝐹((𝑥,𝑡)(𝑦,𝑟))=𝐹(𝑥+𝑡𝑦,𝑟𝑡).(4.2)

4.1. Peetre-Type Spaces on 𝒢

The present paragraph is devoted to the definition of certain mixed norm spaces on the group. Such spaces were considered in various papers, see [10, 11, 14, 15]. Especially tent spaces have some nice applications in harmonic analysis. In particular, they were used by many authors in order to apply coorbit space theory to Lizorkin-Triebel spaces.

Here we use a different approach and define a new scale of function spaces on the group 𝒢. We call them Peetre-type spaces since the Peetre maximal function is involved in the definition. It turned out that they are very easy to handle in connection with translation invariance. Compared with the tent space approach, they are the more natural choice for considering Lizorkin-Triebel spaces as coorbit spaces. Additionally they seem to be the suitable choice for inhomogeneous spaces and more general situations like weighted spaces, which will be studied in a subsequent contribution to the subject.

Definition 4.1. Let 𝑠,0<𝑝,𝑞, and 𝑎>0. We define the spaces ̇𝐿𝑠𝑝,𝑞(𝒢), ̇𝑇𝑠𝑝,𝑞(𝒢), and ̇𝑃𝑠,𝑎𝑝,𝑞(𝒢) on the group 𝒢 via the finiteness of the following (quasi-)norms: ̇𝐿𝐹𝑠𝑝,𝑞=(𝒢)0𝑡𝑠𝑞𝐹(,𝑡)𝐿𝑝𝑑𝑞𝑑𝑡𝑡𝑑+11/𝑞,̇𝑇𝐹𝑠𝑝,𝑞=(𝒢)0𝑡𝑠𝑞𝐵(0,𝑡)||||𝐹(𝑥+𝑧,𝑡)𝑞𝑑𝑧𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑,̇𝑃𝐹𝑠,𝑎𝑝,𝑞=(𝒢)0𝑡𝑠𝑞sup𝑦𝑑||𝐹||(𝑥+𝑦,𝑡)||𝑦||(1+/𝑡)𝑎𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑,(4.3) using the usual modification in case 𝑞=.
Proposition 4.2. The spaces ̇𝐿𝑠𝑝,𝑞̇𝑇(𝒢),𝑠𝑝,𝑞(𝒢) and ̇𝑃𝑠𝑝,𝑞(𝒢), are left and right translation invariant. Precisely, we have 𝐿(𝑧,𝑟)̇𝐿𝑠𝑝,𝑞̇𝐿(𝒢)𝑠𝑝,𝑞(𝒢)=𝑟𝑑(1/𝑝1/𝑞)𝑠,𝑅(𝑧,𝑟)̇𝐿𝑠𝑝,𝑞̇𝐿(𝒢)𝑠𝑝,𝑞(𝒢)=𝑟𝑠+𝑑/𝑞,𝐿(𝑧,𝑟)̇𝑇𝑠𝑝,𝑞(̇𝑇𝒢)𝑠𝑝,𝑞(𝒢)=𝑟𝑑/𝑞𝑠,𝑅(𝑧,𝑟)̇𝑇𝑠𝑝,𝑞̇𝑇(𝒢)𝑠𝑝,𝑞(𝒢)𝐶𝑟𝑑/𝑞+𝑠max1,𝑟𝑏(1+|𝑧|)𝑏,(4.4) where 𝑏>0 is a constant depending on 𝑑, 𝑝, and 𝑞, and 𝐿(𝑧,𝑟)̇𝑃𝑠,𝑎𝑝,𝑞̇𝑃(𝒢)𝑠,𝑎𝑝,𝑞(𝒢)=𝑟𝑑(1/𝑝1/𝑞)𝑠,𝑅(𝑧,𝑟)̇𝑃𝑠,𝑎𝑝,𝑞̇𝑃(𝒢)𝑠,𝑎𝑝,𝑞(𝒢)𝑟𝑠+𝑑/𝑞max{1,𝑟𝑎}(1+|𝑧|)𝑎.(4.5)Proof. Step 1. The left and right translation invariance of ̇𝐿𝑠𝑝,𝑞(𝒢) and ̇𝑇𝑠𝑝,𝑞(𝒢) was shown in [28, Lemma. 4.7.10].Step 2. Let us consider ̇𝑃𝑠,𝑎𝑝,𝑞(𝒢). Clearly, we have for ̇𝑃𝐹𝑠,𝑎𝑝,𝑞(𝒢)𝐿(𝑧,𝑟)̇𝑃𝐹𝑠𝑝,𝑞(=𝒢)0𝑡𝑠𝑞sup𝑦𝑑||||𝐹((𝑥+𝑦𝑧)/𝑟,𝑡/𝑟)||𝑦||(1+/𝑡)𝑎𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑=𝑟𝑑/𝑝0𝑡𝑠𝑞sup𝑦𝑑||||𝐹(𝑥+𝑦,𝑡/𝑟)||𝑦||(1+𝑟/𝑡)𝑎𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑=𝑟𝑑(1/𝑝1/𝑞)𝑠0𝑡𝑠𝑞sup𝑦𝑑||||𝐹(𝑥+𝑦,𝑡)||𝑦||(1+/𝑡)𝑎𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝(𝑑).(4.6) Hence, we obtain 𝐿(𝑧,𝑟)̇𝑃𝑠,𝑎𝑝,𝑞̇𝑃(𝒢)𝑠𝑝,𝑞(𝒢)=𝑟𝑑(1/𝑝1/𝑞)𝑠.(4.7) The right translation invariance is obtained by 𝑅(𝑧,𝑟)̇𝑃𝐹𝑠,𝑎𝑝,𝑞(=𝒢)0𝑡𝑠𝑞sup𝑦𝑑||||𝐹(𝑥+𝑡𝑧+𝑦,𝑡𝑟)||𝑦||1+/𝑡𝑎𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑=0𝑡𝑠𝑞sup𝑦𝑑||||𝐹(𝑥+𝑦,𝑡𝑟)(1+|𝑦𝑡𝑧|/𝑡)𝑎𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑=𝑟𝑠+𝑑/𝑞0𝑡𝑠𝑞sup𝑦𝑑||||𝐹(𝑥+𝑦,𝑡)||||1+𝑦𝑡𝑧𝑟/𝑡𝑎𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑.(4.8) Observe that sup𝑦𝑑||||𝐹(𝑥+𝑦,𝑡)||||1+𝑦𝑡𝑧𝑟/𝑡𝑎=sup𝑦𝑑||||𝐹(𝑥+𝑦,𝑡)||𝑦||1+/𝑡𝑎||𝑦||1+/𝑡𝑎||||1+𝑦𝑡𝑧𝑟/𝑡𝑎,||𝑦||1+/𝑡𝑎||||1+𝑦𝑡𝑧𝑟/𝑡𝑎||||1+𝑦𝑡𝑧/𝑡+|𝑧|𝑎||||1+𝑦𝑡𝑧𝑟/𝑡𝑎=(1+|𝑦𝑡𝑧|/𝑡)𝑎)(1+|𝑧|)𝑎||||1+𝑦𝑡𝑧𝑟/𝑡𝑎.(4.9) This yields sup𝑦𝑑||||𝐹(𝑥+𝑦,𝑡)||||1+𝑦𝑡𝑧𝑟/𝑡𝑎max{1,𝑟𝑎}(1+|𝑧|)𝑎sup𝑦𝑑||||𝐹(𝑥+𝑦,𝑡)||𝑦||1+/𝑡𝑎(4.10) and consequently 𝑅(𝑧,𝑟)̇𝑃𝑠,𝑎𝑝,𝑞̇𝑃(𝒢)𝑠,𝑎𝑝,𝑞(𝒢)𝑟𝑠+𝑑/𝑞max{1,𝑟𝑎}(1+|𝑧|)𝑎.(4.11)

Remark 4.3. Note that we did neither use the translation invariance of the Lebesgue measure nor any change of variable in order to prove the right translation invariance of ̇𝑃𝑠𝑝,𝑞(𝒢). This gives room for generalizations, that is, replacing the space 𝐿𝑝(𝑑) by some weighted Lebesgue space 𝐿𝑝(𝑑,𝜔) in the definition.

4.2. New Old Coorbit Spaces

We start with =𝐿2(𝑑) and the representation𝜋(𝑥,𝑡)=𝑇𝑥𝒟𝐿2𝑡,(4.12) which is unitary continuous on . This representation is not irreducible. However, if we restrict to radial functions 𝑔𝐿2(𝑑), then span{𝜋(𝑥,𝑡)𝑔(𝑥,𝑡)𝒢} is dense in 𝐿2(𝑑). Another possibility to overcome this obstacle is to extend the group by 𝑆𝑂(𝑑), which is somehow equivalent. See [11, 12] for details.

The voice transform in this special situation is represented by the so-called continuous wavelet transform 𝑊𝑔𝑓, see Appendix A.1 in the appendix. Recall the abstract definition of the space 1𝑤 and 𝒜𝑤 from Definition 3.4. The following lemma is proved for instance in [28, Lemma 4.7.11] and is also a consequence of our Lemma A.3 on the decay of the continuous wavelet transform. It states under which conditions on the weight 𝑤 the space 1𝑤 is nontrivial.

Lemma 4.4. If the weight function 𝑤(𝑥,𝑡)1 satisfies the condition 𝑤(𝑥,𝑡)(1+|𝑥|)𝑟𝑡𝑠+𝑡𝑠(4.13) for some 𝑟,𝑠,𝑠0, then 𝒮0𝑑1𝑤.(4.14)

This is a kind of minimal condition which we need in order to define coorbit spaces in a reasonable way. Instead of (𝐻1𝑤), one may use 𝒮0(𝑑) as reservoir and a radial 𝑔𝒮0(𝑑) as analyzing vector. Considering (3.7), we have to restrict to such function spaces 𝑌 on 𝒢 satisfying (i),(ii),(iii) in Section 3.1 where additionally (recall the definition of 𝑤𝑌 in (3.16))

(iv)𝑤𝑌(𝑥,𝑡)(1+|𝑥|)𝑟𝑡𝑠+𝑡𝑠(4.15)

for some 𝑟,𝑠,𝑠0. The following theorem shows how the spaces of Besov-Lizorkin-Triebel type from Section 2 can be recovered as coorbit spaces.

Theorem 4.5. (i) For 1𝑝,𝑞, and 𝑠,we have ̇𝐵𝑠𝑝,𝑞(𝑑̇𝐿)=Co𝑠+𝑑/2𝑑/𝑞𝑝,𝑞(𝒢),
(ii) for 1𝑝<, 1𝑞 and 𝑠we have ̇𝐹𝑠𝑝,𝑞𝑑̇𝑇=Co𝑠+𝑑/2𝑝,𝑞(𝒢)(4.16)
(iii) and if additionally 𝑎>𝑑/min{𝑝,𝑞}, we obtain ̇𝐹𝑠𝑝,𝑞𝑑̇𝑃=Co𝑠+𝑑/2𝑑/𝑞,𝑎𝑝,𝑞(𝒢).(4.17)

Proof. Theorem 4.5 is a direct consequence of Proposition 4.2, Theorems 2.8, 2.9, and the abstract result in Theorem 3.7.

Remark 4.6. (a) The assertions (i) and (ii) are not new. They appear for instance in [11, 14, 15]. They rely on the characterizations given by Triebel in [2] and [3, Sections 2.4, 2.5], see in particular, [3, Section 2.4.5] for the variant in terms of tent spaces, which were invented in [10]. From the deep result in [10, Proposition 4], it follows that ̇𝑇𝑠𝑝,𝑞(𝒢) are translation invariant Banach function spaces on 𝒢, which makes them feasible for coorbit space theory.
(b) Assertion (iii) is indeed new and makes the rather complicated tent spaces ̇𝑇𝑠𝑝,𝑞(𝒢) obsolete for this issue. We showed that 𝑌=𝑃𝑠,𝑎𝑝,𝑞(𝒢) is a much better choice since the right translation invariance is immediate and gives more transparent estimates for its norm. Once we are interested in reasonable conditions for atomic decompositions, this is getting important, see Section 4.5.

4.3. Sequence Spaces

In the following, we consider a compact neighborhood of the identity element in 𝒢 given by 𝒰=[0,𝛼]𝑑×[𝛽1,1], where 𝛼>0 and 1<𝛽. Furthermore, we consider the discrete set of points𝑥𝑗,𝑘=𝛼𝑘𝛽𝑗,𝛽𝑗𝑗,𝑘𝑑.(4.18) Then the family {𝑥𝑗,𝑘𝒰}𝑗,𝑘𝑑 defines a partition of 𝒢. Indeed,𝑥𝑗,𝑘𝒰=𝑄𝑗,𝑘×𝛽(𝑗+1),𝛽𝑗,(4.19)

where𝑄𝑗,𝑘=𝛼𝑘1𝛽𝑗𝑘,𝛼1𝛽+1𝑗××𝛼𝑘𝑑𝛽𝑗𝑘,𝛼𝑑𝛽+1𝑗.(4.20)

Note that in this case the spaces 𝑌 and 𝑌𝑏 coincide. We will further use the notation𝜒𝑗,𝑘(𝑥)=1:𝑥𝑄𝑗,𝑘0:otherwise.(4.21)

Definition 4.7. Let 𝑌 be a function space on 𝒢 satisfying Section 3/(i), (ii), (iii). We put 𝑌𝜆(𝛼,𝛽)=𝑗,𝑘𝑗,𝑘𝜆𝑗,𝑘𝑗,𝑘𝑌(𝛼,𝛽)<,(4.22) where 𝜆𝑗,𝑘𝑗,𝑘𝑌=(𝛼,𝛽)𝑗,𝑘||𝜆𝑗,𝑘||𝜒𝑗,𝑘(𝑥)𝜒[𝛽𝑗,𝛽𝑗+1](𝑡)𝑌.(4.23)

Theorem 4.8. Let 1𝑝,𝑞, 𝑠,and 𝑎>𝑑/min{𝑝,𝑞}. Then 𝜆𝑗,𝑘𝑗,𝑘̇𝑃𝑠,𝑎𝑝,𝑞(𝛼,𝛽)𝑘𝑑𝛽(𝑠+𝑑/𝑞)𝑞||𝜆,𝑘||𝑞𝜒,𝑘(𝑥)1/𝑞𝐿𝑝𝑑,𝜆𝑗,𝑘𝑗,𝑘̇𝐿𝑠𝑝,𝑞(𝛼,𝛽)𝛽(𝑠+𝑑/𝑞𝑑/𝑝)𝑞𝑘𝑑||𝜆,𝑘||𝑝𝑞/𝑝1/𝑞.(4.24)

Proof. We prove the first statement. The proof for the second one is even simpler. Let 𝐹(𝑥,𝑡)=𝑗,𝑘||𝜆𝑗,𝑘||𝜒𝑗,𝑘(𝑥)𝜒[𝛽(𝑗+1),𝛽𝑗](𝑡).(4.25) Discretizing the integral over 𝑡 by 𝑡𝛽, we obtain 𝐹𝑌=0𝑡𝑠𝑞sup𝑤||||𝐹(𝑥+𝑤,𝑡)(1+|𝑤|/𝑡)𝑎𝑞𝑑𝑡𝑡𝑑+11/𝑞𝐿𝑝𝑑𝛽(𝑠+𝑑/𝑞)𝑞𝛽𝛽(+1)sup𝑤||||𝐹(𝑥+𝑤,𝑡)1+𝛽|𝑤|𝑎𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑑.(4.26) With 𝑡[𝛽(+1),𝛽], we observe 𝐹(𝑥,𝑡)=𝑘||𝜆,𝑘||𝜒,𝑘(𝑥)(4.27) and estimate 𝐹𝑌𝛽(𝑠+𝑑/𝑞)𝑞sup𝑤𝑑11+𝛽|𝑤|𝑎𝑘||𝜆,𝑘||𝜒,𝑘(𝑥+𝑤)𝑞1/𝑞𝐿𝑝𝑑.(4.28) In order to include also the situation min{𝑝,𝑞}1, we use the following trick. Obviously, we can rewrite and estimate (4.28) with 0<𝑟<1 in the following way: 𝐹𝑌𝑘𝛽(𝑠+𝑑/𝑞)𝑟||𝜆,𝑘||𝑟sup𝑤𝑑𝜒,𝑘(𝑥+𝑤)1+𝛽|𝑤|𝑎𝑟𝑞/𝑟1/𝑞𝐿𝑝𝑑.(4.29)
The next observation is the useful estimate sup𝑤||𝜒,𝑘||(𝑥+𝑤)1+𝛽|𝑤|𝑎𝑟11+𝛽||𝑥𝑘𝛽||𝑎𝑟𝜒,𝑘𝛽()𝑑1+𝛽||𝑎𝑟(𝑥).(4.30) Indeed, the first estimate is obvious. Let us establish the second one 𝜒,𝑘1()1+𝛽||𝑎𝑟(𝑥)=|𝑦𝑖𝑘𝑖𝛽|𝛽𝑖=1,,𝑑11+𝛽||||𝑥𝑦𝑎𝑟𝑑𝑦||𝑦||𝑐𝛽11+𝛽||𝑥𝑘𝛽||𝑦𝑎𝑟𝑑𝑦||𝑦||𝑐𝛽11+𝛽||𝑥𝑘𝛽||+𝛽||𝑦||𝑎𝑟𝑑𝑦𝛽𝑑10𝑢𝑑11+𝛽||𝑥𝑘𝛽||+𝑢𝑎𝑟𝛽𝑑𝑢𝑑1+𝛽||𝑥𝑘𝛽||𝑎𝑟.(4.31) Note, that the functions 𝑔𝛽(𝑥)=𝑑1+𝛽||𝑎𝑟(4.32) belong to 𝐿1(𝑑) with uniformly bounded norm where we need 𝑎𝑟>𝑑. Putting (4.31) and (4.30) into (4.29), we obtain 𝐹𝑌𝑟𝑔𝑘𝑑𝛽(𝑠+𝑑/𝑞)𝑟||𝜆,𝑘||𝑟𝜒,𝑘(𝑥)𝑞/𝑟𝑟/𝑞𝐿𝑝/𝑟𝑑.(4.33) Now we are in a position to use the majorant property of the Hardy/Littlewood maximal operator (see Section 2.2 and [25, Chapter 2]), which states that a convolution of a function 𝑓 with an 𝐿1(𝑑)-function (having norm one) can be estimated from above by the Hardy/Littlewood maximal function of 𝑓. We choose 𝑟<𝑚𝑖𝑛{𝑝,𝑞} and apply Theorem 2.1 for the 𝐿𝑝/𝑟(𝑞/𝑟) situation. This gives 𝐹𝑌𝑟𝑘𝑑𝛽(𝑠+𝑑/𝑞)𝑟||𝜆,𝑘||𝑟𝜒,𝑘(𝑥)𝑞/𝑟𝑟/𝑞𝐿𝑝/𝑟𝑑=𝑘𝑑𝛽(𝑠+𝑑/𝑞)𝑞||𝜆,𝑘||𝑞𝜒,𝑘(𝑥)1/𝑞𝐿𝑝𝑑𝑟(4.34) and finishes the upper estimate. Both conditions, 𝑎𝑟>𝑑 and 𝑟<min{𝑝,𝑞}, are compatible if 𝑎>𝑑/min{𝑝,𝑞} is assumed at the beginning.
For the estimate from below, we go back to (4.26) and observe sup𝑤||||𝐹(𝑥+𝑤,𝑡)1+𝛽|𝑤|𝑎||||𝐹(𝑥,𝑡),(4.35) which results in 𝐹𝑌𝛽(𝑠+𝑑/𝑞)𝑞𝛽𝛽(+1)||||𝐹(𝑥,𝑡)𝑞𝑑𝑡𝑡1/𝑞𝐿𝑝𝑑.(4.36) A further use of (4.27) gives finally 𝐹𝑌𝑘𝑑𝛽(𝑠+𝑑/𝑞)𝑞||𝜆,𝑘||𝑞𝜒,𝑘(𝑥)1/𝑞𝐿𝑝𝑑.(4.37) The proof is complete.

4.4. Atomic Decompositions

The following theorem is a direct consequence of the abstract results in Theorems 3.11, and 3.13.

Theorem 4.9. Let 1𝑝,𝑞, 𝑎>𝑑/min{𝑝,𝑞}, and 𝑠. Let further 𝑔𝒮0(𝑑) be a radial function. Then there exist numbers 𝛼0>0 and 𝛽0>1 such that for all 0<𝛼𝛼0 and 1<𝛽𝛽0 the family 𝑔𝑗,𝑘𝑗,𝑘𝑑=𝑇𝛼𝑘𝛽𝑗𝒟𝐿2𝛽𝑗𝑔(4.38) has the following properties.(i){𝑔𝑗,𝑘}𝑗,𝑘𝑑 forms a Banach frame for ̇𝐿Co𝑠𝑝,𝑞(𝒢) and ̇𝑃Co𝑠,𝑎𝑝,𝑞(𝒢), that is, we have a dual frame {𝑒𝑗,𝑘}𝑗,𝑘𝑑𝒮0(𝑑)with 𝑓=𝑗,𝑘𝑑𝑔𝑗,𝑘,𝑓𝑒𝑗,𝑘 and the norm equivalences ̇𝐿𝑓Co𝑠𝑝,𝑞𝑔(𝒢)𝑗,𝑘̇𝐿,𝑓𝑠𝑝,𝑞(𝒢)̇𝐿(𝛼,𝛽),𝑓Co𝑠𝑝,𝑞(𝒢)(4.39) as well as ̇𝑃𝑓Co𝑠,𝑎𝑝,𝑞𝑔(𝒢)𝑗,𝑘̇𝑃,𝑓𝑠,𝑎𝑝,𝑞(𝒢)̇𝑃(𝛼,𝛽),𝑓Co𝑠,𝑎𝑝,𝑞(𝒢).(4.40)(ii){𝑔𝑗,𝑘}𝑗,𝑘𝑑 is an atomic decomposition, that is, for ̇𝑃𝑓Co𝑠,𝑎𝑝,𝑞(𝒢), we have a (not necessary unique) decomposition 𝑗,𝑘𝑑𝜆𝑗,𝑘(𝑓)𝑔𝑗,𝑘 such that 𝜆𝑗,𝑘(𝑓)𝑗,𝑘̇𝑃𝑠,𝑎𝑝,𝑞(𝒢)̇𝑃(𝛼,𝛽)𝑓Co𝑠,𝑎𝑝,𝑞(𝒢).(4.41) Conversely, if {𝜆𝑗,𝑘}𝑗,𝑘𝑑̇𝑃(𝑠,𝑎𝑝,𝑞(𝒢))(𝛼,𝛽), then 𝑓=𝑗,𝑘𝑑𝜆𝑗,𝑘𝑔𝑗,𝑘 converges and belongs to ̇𝑃Co𝑠,𝑎𝑝,𝑞(𝒢) and moreover, ̇𝑃𝑓Co𝑠,𝑎𝑝,𝑞𝜆(𝒢)𝑗,𝑘𝑗,𝑘̇𝑃𝑠,𝑎𝑝,𝑞(𝒢)(𝛼,𝛽)(4.42) (analogously for ̇𝐿Co𝑠𝑝,𝑞(𝒢)). Convergence is considered in the strong topology if the finite sequences are dense in (̇𝑃𝑠,𝑎𝑝,𝑞(𝒢))(𝛼,𝛽) and in the weak-topology otherwise.

Remark 4.10. (i) Since the analyzing function or atom 𝑔 can be chosen arbitrarily, we allow more flexibility here than in the results given in Frazier/Jawerth [30] and Triebel [3, 29].
(ii) Instead of regular families of sampling points (𝛼𝛽𝑗𝑘,𝛽𝑗) rather irregular families of points in 𝒢 are allowed as long as they are distributed sufficiently dense, see Theorem 3.11.

4.5. Wavelet Frames

In the sequel, we consider wavelet bases on 𝑑 in the sense of Lemma A.5 from Appendix A.2 in the appendices. We have given an orthonormal scaling function Ψ0 and the associated wavelet Ψ1 on and consider the tensor products Ψ𝑐, 𝑐𝐸. Our aim is to specify, that is, give sufficient conditions to Ψ0, Ψ1, such that (A.19) represents an unconditional basis in 𝐵𝑠𝑝,𝑞(𝑑)and 𝐹𝑠𝑝,𝑞(𝑑). We will apply the abstract Theorem 3.14.

In order to do so we need to have (3.17) for all functions Ψ𝑐. We will state certain smoothness, decay, and moment conditions to Ψ1 and Ψ0, see Definition A.1 in the appendix, to ensure this. Let us fix the neighborhood 𝑉=[1,1]𝑑×(1/2,1]𝒢 of 𝑒𝒢.

Proposition 4.11. Let 𝐿, 𝐾>0, and Ψ0 be an orthogonal scaling function with associated wavelet Ψ1 on . The function Ψ0 is supposed to satisfy (𝐷) and (𝑆𝐾) and Ψ1 is supposed to satisfy (𝐷), (𝑆𝐾), and (𝑀𝐿1). For 𝑟1,𝑟2,the weight 𝑤(𝑥,𝑡) is given by 𝑤(𝑥,𝑡)=(1+|𝑥|)𝑣(𝑡𝑟2+𝑡𝑟1),(𝑥,𝑡)𝒢.(4.43) If now 𝑟1𝑑<min{𝐿,𝐾}2,𝑟2𝑑<min{𝐿,𝐾}+2𝑣,(4.44) then we have 𝑑0sup(𝑦,𝑠)(𝑥,𝑡)𝑉||𝜋(𝑦,𝑠)Ψ𝑐,Ψ𝑐||𝑤(𝑥,𝑡)𝑑𝑡𝑡𝑑+1𝑑𝑥<.(4.45)

Proof. With Lemma A.3, we obtain for 𝑊Ψ1Ψ1 the following estimates: ||𝑊Ψ1Ψ1(||𝑡𝑠,𝑡)min{𝐿,𝐾}+1/2(1+𝑡)2min{𝐿,𝐾}+11(1+|𝑠|/(1+𝑡))𝑁.(4.46) And in addition ||𝑊Ψ𝑖Ψ𝑖||𝑡(𝑠,𝑡)1/21(𝑡+1)(1+|𝑠|/(1+𝑡))𝑁,𝑖=1,2.(4.47) Hence, for any 𝑐𝐸, the tensor product structure gives (assume without restriction that 𝑐𝑑=1) ||𝑊Ψ𝑐Ψ𝑐||𝑡(𝑥,𝑡)min{𝐿,𝐾}(1+𝑡)2min{𝐿,𝐾}𝑡𝑑/2(1+𝑡)𝑑𝑑𝑖=11||𝑥1+𝑖||/(1+𝑡)𝑁.(4.48) The expression sup(𝑦,𝑠)(𝑥,𝑡)𝑉|𝑊Ψ𝑐Ψ𝑐(𝑦,𝑠)| can be estimated similar sup(𝑦,𝑠)(𝑥,𝑡)𝑉||𝑊Ψ𝑐Ψ𝑐||(𝑦,𝑠)=sup||𝑦𝑖𝑥𝑖||𝑡𝑡/2𝑠𝑡||𝑊Ψ𝑐Ψ𝑐||𝑡(𝑦,𝑠)min{𝐿,𝐾}(1+𝑡)2min{𝐿,𝐾}𝑡𝑑/2(1+𝑡)𝑑𝑑𝑖=11||𝑥1+𝑖||/(1+𝑡)𝑁𝑡min{𝐿,𝐾}(1+𝑡)2min{𝐿,𝐾}𝑡𝑑/2(1+𝑡)𝑑1(1+|𝑥|/(1+𝑡))𝑁.(4.49) Fubini’s theorem and a change of variable yields 𝑑0sup(𝑦,𝑠)(𝑥,𝑡)𝑉||𝜋(𝑦,𝑠)Ψ𝑐,Ψ𝑐||𝑤(𝑥,𝑡)𝑑𝑡𝑡𝑑+1𝑑𝑥0𝑡min{𝐿,𝐾}(1+𝑡)2min{𝐿,𝐾}𝑡𝑑/2(𝑡𝑟2+𝑡𝑟1)𝑑𝑡𝑡𝑑+1.(4.50) Finally it is easy to see that the latter is finite if the conditions in (4.44) are valid. This proves Proposition 4.11.

Theorem 4.12. Let 𝐿, 𝐾>0, and Ψ0 be an orthogonal scaling function with associated wavelet Ψ1 on . The function Ψ0 is supposed to satisfy (𝐷) and (𝑆𝐾), and Ψ1 is supposed to satisfy (𝐷), (𝑆𝐾) and (𝑀𝐿1). (a) If 1𝑝,𝑞 and 𝑑min{𝐿,𝐾}+𝑝1<𝑠<min{𝐿,𝐾}𝑑1𝑝,(4.51) then (A.19) is a Banach frame for ̇𝐵𝑠𝑝,𝑞(𝑑) in the sense of (3.22).(b) If 1𝑝,𝑞< and 1min{𝐿,𝐾}+2𝑑max𝑝,1𝑞1<𝑠<min{𝐿,𝐾}𝑑max𝑝,1𝑞1,1𝑝,(4.52) then (A.19) is a Banach frame for ̇𝐹𝑠𝑝,𝑞(𝑑)in the sense of (3.22).

Proof. Let us prove (a). First of all, we apply Theorem 4.5/(i). Afterwards, we use Proposition 4.2 in order to estimate the weight 𝑤𝑌(𝑦,𝑡) for ̇𝐿𝑌=𝑠+𝑑/2𝑑/𝑞𝑝,𝑞(𝒢). We obtain 𝑤𝑌𝑡(𝑥,𝑡)=max𝑑(1/𝑝1/2)𝑠,𝑡𝑠𝑑(1/𝑝1/2),𝑡𝑠+𝑑/2,𝑡𝑠+𝑑/2𝑡𝑟1𝑡0<𝑡<1,𝑟2𝑡1.(4.53) Let us distinguish the cases, 𝑠0 and 𝑠<0. In the first case we can put 𝑟1=max{𝑠𝑑(1/𝑝1/2),𝑠+𝑑(1/𝑝1/2),𝑠𝑑/2}, 𝑟2=max{𝑠+𝑑/2,𝑠+𝑑(1/𝑝1/2)}, and 𝑣=0. Now we apply first Proposition 4.11. This gives the condition 0𝑠<min{𝐿,𝐾}𝑑(11/𝑝).(4.54) In the second case we put 𝑟1=max{𝑠𝑑(1/𝑝1/2),𝑠+𝑑(1/𝑝1/2),𝑠𝑑/2}, 𝑟2=max{𝑠+𝑑/2,𝑠+𝑑(1/𝑝1/2)} and 𝑣=0. With Proposition 4.11, we obtain the condition min{𝐿,𝐾}+𝑑/𝑝<𝑠<0.(4.55) Finally (4.54), (4.55), and Theorem 3.14 yield (a).
Step 2. We prove (b). We apply Theorem 4.5/(iii) and afterwards Proposition 4.2 and obtain for ̇𝑃𝑌=𝑠+𝑑/2𝑑/𝑞,𝑎𝑝,𝑞(𝒢)𝑤𝑌𝑡(𝑥,𝑡)=max𝑑(1/𝑝1/2)𝑠,𝑡𝑠𝑑(1/𝑝1/2),𝑡𝑠+𝑑/2max{1,𝑡𝑎}(1+|𝑥|𝑎),𝑡𝑠+𝑑/2max{𝑡𝑎,𝑡𝑎}(1+|𝑥|)𝑎(1+|𝑥|)𝑎𝑡𝑟1𝑡:0<𝑡<1,𝑟2:𝑡1.(4.56) First, we consider the case 𝑠0. We can put 𝑟1=max{𝑠+𝑎𝑑/2,𝑠+𝑑/2𝑑/𝑝}, 𝑟2=max{𝑠+𝑑/2,𝑠+𝑑/2+𝑎}, and 𝑣=𝑎. Proposition 4.11 gives the condition 10𝑠<min{𝐿,𝐾}max𝑎,𝑑1𝑝(4.57) which can be rewritten to 10𝑠<min{𝐿,𝐾}𝑑max𝑝,1𝑞1,1𝑝(4.58) since 𝑎 can be chosen arbitrarily greater than 𝑑max{1/𝑝,1/𝑞}. This gives the upper bound in (b). Now we consider 𝑠<0. We put 𝑟1=max{𝑠+𝑑/𝑝𝑑/2,𝑠+𝑑/2𝑑/𝑝}, 𝑟2=𝑠+𝑑/2+𝑎. This yields min{𝐿,𝐾}+2𝑎<𝑠<0(4.59) and can be rewritten to min{𝐿,𝐾}+2𝑑max{1/𝑝,1/𝑞}<𝑠<0.(4.60) This yields the lower bound in (b) and we are done.

The following corollary is a direct consequence of Theorem 4.12 and the facts in Appendix A.2.

Corollary 4.13. Let 𝑚>0 and (Ψ0,Ψ1)=(𝜑𝑚,𝜓𝑚) the spline wavelet system of order 𝑚. Then (a) If 1𝑝,𝑞 and 𝑑𝑚+1+𝑝1<𝑠<𝑚1𝑑1𝑝,
then (A.19) is a Banach frame for ̇𝐵𝑠𝑝,𝑞(𝑑) in the sense of (3.22);
(b) If 1𝑝,𝑞 and 1𝑚+1+2𝑑max𝑝,1𝑞1<𝑠<𝑚1𝑑max𝑝,1𝑞1,1𝑝,(4.61)
then (A.19) is a Banach frame for ̇𝐹𝑠𝑝,𝑞(𝑑) in the sense of (3.22).

Remark 4.14. The (optimal) smoothness conditions in [31] are slightly weaker than (a) in case 𝑑=1. However, compared to the approach of Triebel [29, 32], we admit some more degree of freedom. The wavelet or atom does not have to be compactly supported. Additionally, in case 𝑑=1, we do not need that 𝜓𝐶𝑢() where 𝑢>𝑠. Indeed, the conditions in (a) and (b) are slightly weaker.

Remark 4.15. More examples can be obtained by using compactly supported Daubechies’ wavelets of order 𝑁 or Meyer’s wavelets. Based on the underlying abstract result in Theorem 3.14 even biorthogonal wavelet systems which provide sufficiently high smoothness and vanishing moments are suitable for this issue.

Appendix

A. Wavelets

A.1. The Continuous Wavelet Transform

The vector 𝑔 is said to be the analyzing vector for a function 𝑓𝐿2(𝑑).The continuous wavelet transform 𝑊𝑔𝑓 is then defined through𝑊𝑔𝑇𝑓(𝑥,𝑡)=𝑥𝒟𝐿2𝑡𝑔,𝑓,𝑥𝑑,𝑡>0,(A.1) where the bracket , denotes the inner product in 𝐿2(𝑑). We call 𝑔 an admissible wavelet if𝑐𝑔=𝑑||||𝑔(𝜉)2||𝜉||𝑑𝑑𝜉<.(A.2)

If this is the case, then the family {𝑇𝑥𝒟𝐿2𝑡𝑔}𝑡>0,𝑥𝑑 represents a tight continuous frame in 𝐿2(). In particular, this means that the above family is dense in 𝐿2(𝑑). For a proof we refer to Theorem  1.5.1 in [28].

Let us now specify the conditions (𝑀𝐿), (𝐷), and (𝑆𝐾), which we intend to impose on functions Φ,Ψ𝐿2(𝑑) in order to obtain a good decay of the continuous wavelet transform |𝑊ΨΦ(𝑥,𝑡)|.

Definition A.1. Let 𝐿+10, 𝐾>0, and fix the conditions (𝐷), (𝑀𝐿), and (𝑆𝐾) for a function Ψ𝐿2(𝑑). (𝐷) For every 𝑁,there exists a constant 𝑐𝑁 such that ||||𝑐Ψ(𝑥)𝑁(1+|𝑥|)𝑁.(A.3) (𝑀𝐿) We have vanishing moments 𝐷𝛼Ψ(0)=0(A.4) for all |𝛼|1𝐿.(𝑆𝐾) The function||𝜉||1+𝐾||𝐷𝛼||Ψ(𝜉)(A.5)
belongs to 𝐿1(𝑑)for every multi-index 𝛼𝑑0.

Remark A.2. If a function 𝑔𝐿2(𝑑) satisfies (𝑆𝐾) for some𝐾>0 then by well-known properties of the Fourier transform we have 𝑔𝐶𝐾(𝑑).

The following lemma provides a useful decay result for the continuous wavelet transform under certain smoothness, decay, and moment conditions, see also [4, 30, 33] for similar results in a different language. It represents a continuation of [4, Lemma  1] where one deals with 𝒮(𝑑)-functions.

Lemma A.3. Let 𝐿0, 𝐾>0, and Φ,Ψ,Φ0𝐿2(𝑑). (i) Let Φ satisfy (𝐷), (𝑀𝐿1), and let Φ0 satisfy (𝐷), (𝑆𝐾). Then for every 𝑁there exists a constant 𝐶𝑁 such that the estimate ||𝑊ΦΦ0(||𝑥,𝑡)𝐶𝑁𝑡min{𝐿,𝐾}+𝑑/2(1+|𝑥|)𝑁(A.6) holds true for 𝑥𝑑and 0<𝑡<1.(ii) Let Φ,Ψ satisfy (𝐷), (𝑀𝐿1) and (𝑆𝐾). For every 𝑁there exists a constant 𝐶𝑁 such that the estimate ||𝑊ΦΨ||(𝑥,𝑡)𝐶𝑁𝑡min{𝐿,𝐾}+𝑑/2(1+𝑡)2min{𝐿,𝐾}+𝑑1+|𝑥|1+𝑡𝑁(A.7) holds true for 𝑥𝑑and0<𝑡<.

Proof. Step 1. Let us prove (i). We follow the proof of Lemma  1 in [4]. This reference deals with 𝒮(𝑑)-functions, which makes the situation much more easy. We argue as follows. Clearly, ||𝑊ΦΦ0||(𝑥,𝑡)=𝑡𝑑/2||𝒟𝑡ΦΦ0||(𝑥).(A.8) Fix 0<𝑡<1. Obviously, the convolution (𝒟𝑡Φ)Φ0 satisfies (𝐷). By well-known properties of the Fourier transform, the derivative 𝐷𝛼((𝒟𝑡Φ)Φ0))(𝜉) exists for every multi-index 𝛼𝑑0. For fixed 𝛼, we estimate by using Leibniz’ formula ||𝐷𝛼𝒟𝑡ΦΦ0||=||𝐷)(𝜉)𝛼Φ(𝑡𝜉)Φ0||(𝜉)𝑐𝛼𝛽𝛼𝑡|𝛽|1|||𝐷𝛽Φ(𝑡𝜉)𝐷𝛼𝛽Φ0|||(𝜉)𝑐𝛼𝑡𝐿||𝜉||1+𝐿𝛽𝛼|||𝐷𝛼𝛽Φ0|||.(𝜉)(A.9) In the last step we used property (𝑀𝐿1). Assuming 𝐾𝐿 and exploiting (𝑆𝐾), we obtain that the left-hand side of (A.9) belongs to 𝐿1(𝑑) and 𝐷𝛼𝒟𝑡ΦΦ0)(𝜉)𝐿1𝑑𝑐𝛼𝑡𝐿.(A.10) We proceed as follows: max|𝛼|1𝑁+1𝐷𝛼𝒟𝑡ΦΦ0(𝜉)𝐿1𝑑max|𝛼|1𝑁+11𝐷𝛼𝒟𝑡ΦΦ0𝐿𝑑𝑐𝑁(1+|𝑥|)𝑁𝒟𝑡ΦΦ0(𝑥)𝐿𝑑.(A.11) This estimate together with (A.8) and (A.10) yields (A.6).
Let us finally assume 𝐾<𝐿 and return to (A.9). Clearly, the resulting inequality keeps valid if we replace the exponent 𝐿 by 𝐿0 with 𝐿𝐿. It is even possible to extend (A.9) to every 0𝐿<𝐿 by the following argument. Let 𝐿. We have on the one hand LHS(A.9)𝑐𝛼𝑡𝐿||𝜉||1+𝐿𝐺(𝜉)(A.12) and on the other hand LHS(A.9)𝑐𝛼𝑡𝐿+1||𝜉||1+𝐿+1𝐺(𝜉),(A.13) where 𝐺(𝜉)=𝛽𝛼|𝐷𝛼𝛽Φ0(𝜉)|. Choosing 0<𝜃<1 such that 𝐿=(1𝜃)𝐿+𝜃(𝐿+1), we obtain by a kind of interpolation argument LHS(A.9)=LHS(A.9)1𝜃LHS(A.9)𝜃𝑐𝛼𝑡𝐿||𝜉||1+𝐿𝐺(𝜉).(A.14) In particular, we obtain instead of (A.9) ||𝐷𝛼𝒟𝑡ΦΦ0||(𝜉)𝑐𝛼𝑡𝐾||𝜉||1+𝐾𝛽𝛼|||𝐷𝛼𝛽Φ0|||(𝜉),𝜉𝑑.(A.15) We exploit property (𝑆𝐾) for Φ0 and proceed analogously as above. This proves (A.6).
Step 2. The estimate in (ii) is an immediate consequence of (A.6) and the fact 𝑊ΦΨ𝑊(𝑥,𝑡)=ΨΦ𝑥𝑡,1𝑡.(A.16) This completes the proof.

Corollary A.4. Let Φ,Ψ belong to the Schwartz space 𝒮0(𝑑). By Lemma A.3/(ii) for every 𝐿,𝑁there is a constant 𝐶𝐿,𝑁>0 such that ||𝑊ΦΨ||(𝑥,𝑡)𝐶𝐿,𝑁𝑡𝐿+𝑑/2(1+𝑡)2𝐿+𝑑1+|𝑥|1+𝑡𝑁,𝑥𝑑,𝑡>0.(A.17) Additionally, we obtain for Φ𝒮0(𝑑) and Φ0𝒮(𝑑) such that ||𝑊ΦΦ0||(𝑥,𝑡)𝐶𝐿,𝑁𝑡𝐿+𝑑/2(1+|𝑥|)𝑁,𝑥𝑑,0<𝑡<1.(A.18)

A.2. Orthonormal Wavelet Bases

The following Lemma is proved in Wojtaszczyk [34, Section 5.1].

Lemma A.5. Suppose we have a multiresolution analysis in 𝐿2() with scaling functions Ψ0(𝑡) and associated wavelets Ψ1(𝑡). Let 𝐸={0,1}𝑑{(0,,0)}. For 𝑐=(𝑐1,,𝑐𝑑)𝐸, let Ψ𝑐=𝑑𝑗=1Ψ𝑐𝑗. Then the system 2𝑗𝑑/2Ψ𝑐2𝑗𝑥𝑘𝑐𝐸,𝑗,𝑘𝑑(A.19) is an orthonormal basis in 𝐿2(𝑑).

Spline Wavelets
As a main example, we will consider the spline wavelet system. The normalized cardinal B-spline of order 𝑚+1 is given by 𝒩𝑚+1(𝑥)=𝒩𝑚𝒳(𝑥),𝑥,𝑚,(A.20) beginning with 𝒩1=𝒳, the characteristic function of the interval (0,1). By 𝜑𝑚1(𝑥)=2𝜋1𝒩𝑚(𝜉)𝑘=||𝒩𝑚||(𝜉+2𝜋𝑘)21/2(𝑥),𝑥,(A.21) we obtain an orthonormal scaling function, which is again a spline of order 𝑚. Finally, by 𝜓𝑚(𝑥)=𝑘=𝜑𝑚𝑡2,𝜑𝑚((𝑡𝑘)1)𝑘𝜑𝑚(2𝑥+𝑘+1)(A.22) the generator of an orthonormal wavelet system is defined. For 𝑚=1, it is easily checked that 𝜓1(𝑥1) is the Haar wavelet. In general, these functions 𝜓𝑚 have the following properties:(i)𝜓𝑚 restricted to intervals [𝑘/2,(𝑘+1)/2], 𝑘, is a polynomial of degree at most 𝑚1.(ii)𝜓𝑚𝐶𝑚2() if 𝑚2.(iii)𝜓𝑚(𝑚2) is uniformly Lipschitz continuous on if 𝑚2.(iv)The function 𝜓𝑚 satisfies a moment condition of order 𝑚1, that is,𝑥𝜓𝑚(𝑥)𝑑𝑥=0,=0,1,,𝑚1.(A.23)
In particular, 𝜓𝑚 satisfies (𝐷), (𝑀𝐿) for 0<𝐿𝑚 and (𝑆𝐾) for 𝐾<𝑚1.

Acknowledgments

The author would like to thank Holger Rauhut, Martin Schäfer, Benjamin Scharf, and Hans Triebel for valuable discussions, a critical reading of preliminary versions of this paper and for several hints how to improve it.