Abstract

Homogeneous Triebel-Lizorkin spaces with full range of parameters are introduced on stratified Lie groups in terms of Littlewood-Paley-type decomposition. It is shown that the scale of these spaces is independent of the choice of Littlewood-Paley-type decomposition and the sub-Laplacian used for the construction of the decomposition. Some basic properties of these spaces are given. As the main result of this paper, boundedness of a class of singular integral operators on these function spaces is obtained.

1. Introduction

In recent years there were several efforts of extending Besov and Triebel-Lizorkin spaces from Euclidean spaces to other domains and non-isotropic settings. In particular, Han et al. [1] developed a theory of these function spaces on spaces of homogenous type with the additional reverse doubling property. That setting is quite general and includes for example Lie groups of polynomial growth. However, the high level of generality imposes restrictions on the possible values of the parameters of the function spaces.

For the purpose of studying subelliptic regularity, Folland [2] introduced fractional Sobolev spaces and Lipschitz spaces on stratified Lie groups. Later, Folland and Stein [3] established the theory of Hardy spaces on general homogeneous groups. Besov spaces on stratified Lie groups were first introduced by Saka [4], by means of the heat semigroup associated to the sub-Laplacian. Recently, Führ and Mayeli [5] introduced homogeneous Besov spaces on stratified Lie groups in terms of Littlewood-Paley-type decomposition and established wavelet characterization of them. However, the integrability parameter and the summability parameter of the function spaces studied in both [4, 5] are restricted to be no less than . Moreover, systematic treatment of Triebel-Lizorkin spaces on stratified Lie groups can not be found in the literature, to our best knowledge.

The purpose of this paper is to introduce and study homogeneous Triebel-Lizorkin spaces with full range of parameters on stratified Lie groups. Motivated by [5], we define these function spaces via Littlewood-Paley-type decomposition. We find that a helpful way to treat the case that either the integrability parameter or the summability parameter is less than is to take the Peetre type maximal function into consideration. With the help of the almost orthogonality estimate on stratified Lie groups (see Lemma 2), we show that our definition of homogeneous Triebel-Lizorkin spaces is independent of the choice of the Littlewood-Paley-type decomposition and the sub-Laplacian used for the construction of the decomposition. Thus, these function spaces reflect of properties of the group, not of the sub-Laplacian used for the construction of the decomposition.

Singular integral theory is a powerful tool for the study of partial differential equations. The -boundedness of convolution operators with homogeneous distribution kernels on Lie groups endowed with suitable homogeneous structure was proved by Knapp and Stein [6] (for ) and Korányi and Vági [7] (for ). In Section 4 of this paper, we prove the boundedness on homogeneous Triebel-Lizorkin spaces of a class of convolution type singular integral operators on stratified Lie groups, which includes convolution operators with homogeneous distribution kernels.

This paper is organized as follows. After reviewing some basic notions concerning stratified Lie groups and their associated sub-Laplaicans in Section 2, in Section 3 we introduce homogeneous Triebel-Lizorkin spaces on stratified Lie groups, and give some basic properties of them. In Section 4 we show the -boundedness of a class of convolution singular integral operators. Throughout this paper the letter will denote a positive constant which is independent of the main variables involved but whose value may differ from line to line. The notation or for some variable quantities and means that for some constant ; stands for . We agree that the set of natural numbers contains .

2. Preliminaries

In this section we briefly review the basic notions concerning stratified Lie groups and their associated sub-Laplacians. For more details we refer the reader to the monograph by Folland and Stein [3]. A Lie group is called a stratified Lie group if it is connected and simply connected, and its Lie algebra may be decomposed as a direct sum , with for and . Such a group is clearly nilpotent, and thus it may be identified with (as a manifold) via the exponential map . Examples of stratified Lie groups include Euclidean spaces and the Heisenberg group .

The algebra is equipped with a family of dilations which are the algebra automorphisms defined by Under our identification of with , may also be viewed as a map . We generally write instead of , for . We shall denote by the homogeneous dimension of .

A homogeneous norm on G is a continuous function from to smooth away from (the group identity), vanishing only at , and satisfying and for all and . Homogeneous norms on always exist and any two of them are equivalent. We assume is provided with a fixed homogeneous norm. It satisfies a triangle inequality: there exists a constant such that for all . If and we define the ball of radius about by . The Lebesgue measure on induces a bi-invariant Haar measure on . As done in [3], we fix the normalization of Haar measure by requiring that the measure of be . We shall denote the measure of any measurable by . Clearly we have . All integrals on are with respect to (the normalization of) Haar measure. Convolution is defined by

We consider as the Lie algebra of all left-invariant vector fields on , and fix a basis of , obtained as a union of bases of the . In particular, , with , is a basis of . We denote by the corresponding basis for right-invariant vector fields, that is, If is a multi-index we set and . Moreover, we set where the integers are given according to that . Then (resp., ) is a left-invariant (resp., right-invariant) differential operator, homogeneous of degree , with respect to the dilations , .

A complex-valued function on is called a polynomial on if is a polynomial on . Let be the basis for the linear forms on dual to the basis for , and set . From our definition of polynomials on , are generators of the algebra of polynomials on . Thus, every polynomial on can be written uniquely as where all but finitely many of the coefficients vanish, and . A polynomial of the type (6) is called of homogeneous degree , where , if holds for all multi-indices with . We let denote the space of all polynomials on , and let denote the space of polynomials on of homogeneous degree . Note that is invariant under left and right translations (see [3, Proposition 1.25]). A function is said to have vanishing moments of order , if with the absolute convergence of the integral.

The Schwartz class on is defined by that is, if and only if . As is indicated in [3, p. 35], is a Fréchet space and several different choices of families of norms induce the same topology on . In this paper, for our purpose we use the family of norms given by Here and in what follows, we use the notation convention for any function . The dual space of is the space of tempered distributions on . If and we shall denote the evaluation of on by .

Using the above conventions for the choice of the basis , and , the sub-Laplacian is defined by . When restricted to smooth functions with compact support, is essentially self-adjoint. Its closure has domain , where is taken in the sense of distributions. We denote this extension still by the symbol . By the spectral theorem, admits a spectral resolution where is the projection measure. If is a bounded Borel measurable function on , the operator is bounded on , and commutes with left translations. Thus, by the Schwartz kernel theorem, there exists a tempered distribution on such that Note that the point may be neglected in the spectral resolution, since the projection measure of is zero (see [8, p. 76]). Consequently we should regard as functions on rather than on .

Let denote the space of restrictions to of functions in . An important fact proved by Hulanicki [9] is as in the following lemma.

Lemma 1. If then the distribution kernel of is in .

Moreover, from the proof of [10, Corollary 1] we see that if is a function in which vanishes identically near the origin, then is a Schwartz function with all moments vanishing.

In the sequel, if not other specified, we will generally use Greak alphabets with hats to denote functions in , and use Greek alphabets without hats to denote the associated distribution kernels; for example, for we shall denote by the distribution kernel of the operator , where is a sub-Laplacian fixed in the context.

3. Homogeneous Triebel-Lizorkin Spaces on Stratified Lie Groups

For any function on and , we define the -normalized dilation of by Before we introduce the homogeneous Triebel-Lizorkin spaces on stratified Lie groups, we prove the following basic estimate, which is a generalization of [11, Lemma B.1] and will be frequently used throughout this paper.

Lemma 2. Let with . Suppose both have vanishing moments of order . Then there exists a constant such that for all and all , where .

Proof. Using dilations and the facts and (see (9)), we may assume . To proceed we follow the idea in the proof of [11, Lemma B.1]. Let , and . Let be the left Taylor polynomial of at of homogeneous degree (see [3, pp. 26-27]). Then using vanishing moments of
For , the stratified Taylor formula (cf. [3, Corollary 1.44]) yields that, with a suitable positive constant, since if . Hence we have where for the last inequality we used [3, Corollary 1.17] and that .
For , we have . On the other hand, . Thus, we have Also, we note that by [12, Proposition 20.3.14] the left Taylor polynomial is of the form where the integers are given according to that . From these remarks, it follows that where we used that and .
For we have , and, hence where for the last inequality we used [3, Corollary 1.17] and .
Combining the above estimates, we arrive at This is exactly what we need.

Let denote the space of Schwartz functions with all moments vanishing. We then consider as a subspace of , including the topology. It is shown in [5, Lemma 3.3] that is a closed subspace of , and the topology dual of can be canonically identified with the factor space .

We now have the following Calderón type reproducing formula.

Lemma 3. Suppose is a sub-Laplacian on , and is a function with compact support, vanishing identically near the origin, and satisfying Then for all , it holds that with convergence in . Duality entails that, for all , and the convergence is in .

Proof. First note that the -homogeneity of implies that the distribution kernel of coincides with . Let and be arbitrarily chosen. Then take such that . Since both and are Schwartz functions with all moments vanishing, it follows by Lemma 2 that where the constant is a suitable multiple of . This implies that converges uniformly in , for every and every . Consequently there exists such that converges in the topology of to , as . On the other hand, by (23) and the spectral theorem (cf. [13, Theorem VII.2]), holds in -norm. Therefore, , which completes the proof.

Let denote the class of all functions in satisfying

Definition 4. Let , and . Let be a sub-Laplacian on and . We define as the space of all such that with the usual modification for .

We then introduce the Peetre type maximal functions: Given , , a sub-Laplacian, and , we define

Lemma 5. Suppose is a sub-Laplacian and . Then for every there is a constant such that for all , all , and all ,

Proof. Because of (28) it is possible to find a function supported in such that for . Set , . Then for all . Consequently, for and , where for the last inequality we used [3, Corollary 1.17] and that . Dividing both sides of the above estimate by , and then taking the supremum over and , we obtain the desired estimate.

Lemma 6. Suppose is a sub-Laplacian, , and . Then there exists a constant such that for all , all , and all , where , and is the Hardy-Littlewood maximal operator on .

Proof. Let and . The stratified mean value theorem (cf. [3, Theorem 1.41]) gives that for every and every with , where is a suitable positive constant. Hence we have Putting , dividing both sides by , and using Lemma 5, we have where we have set and used that . Finally, taking sufficiently small (such that ), and taking the supremum over , we get the desired estimate.

Theorem 7. Suppose are any two sub-Laplacians on , and are any two functions in . Then, for , and , we have the (quasi-)norm equivalence

Proof. First we note that if then we have for , where denotes the distribution kernel of . Indeed, the direction “” of (38) is obvious, and the other direction follows by Lemma 6 and the Fefferman-Stein vector-valued maximal inequality on spaces of homogeneous type (see, e.g., [14]). Thus, to prove (37), it suffices to show that
To this end, let be a function in with support in such that for . For , by Lemma 3 we have with convergence in . Here is the distribution kernel of . Hence, since , we have the pointwise representation It follows that Since both and are Schwartz functions with all moments vanishing, we can use Lemma 2 to estimate that, with sufficiently large, Here, the constant is a suitable multiple of . On the other hand we observe that Putting these estimates into (42), multiplying both sides by , dividing both sides by and then taking the supremum over , we obtain In view of [15, Lemma 2], taking in the above inequality yields the direction “” of (39). By symmetricity, (39) holds, and the proof is complete.

Remark 8. From Theorem 7 we see that the space is actually independent of the choice of and . Thus, in what follows we don't specify the choice of and and write instead of . Henceforth we shall fix any sub-Laplacian . Moreover, for the sake of briefness, we will write instead of .

Proposition 9. For , and , one has the continuous inclusion maps .

Proof. Let and . Choose and . Since both and are Schwartz functions with all moments vanishing, it follows by Lemma 2 that for all and all , where the constant is a suitable multiple of . This together with [3, Corollary 1.17] give that which implies that continuously.
Now we show the other embedding. Let and take any . Let , and then let be a function with support in such that for . Then by Lemma 3 we have To proceed we claim that for all . Assuming the claim for a moment, it follows from (48) that It is easy to see that To estimate the sum in (50), we note that if we choose and then similarly to (46) we have where the constant is a suitable multiple of . From this it follows that Therefore, This implies that continuously.
We are left with showing the claim. Indeed, if is fixed then by Lemma 6 we have, for all , Taking in the above estimate and using Hardy-Littlewood maximal inequality, we have Since is arbitrary, the claim follows.

Since all the necessary tools are developed in the above arguments, the following proposition can be proved in the same manner as its Euclidean counterpart; see, for example, the proof of [16, Theorem 2.3.3].

Proposition 10. For , and , is a quasi-Banach space.

Let us introduce a class of functions. We say that , if there exists whose support is compact and which vanishes identically near the origin, and , such that . Clearly .

Lemma 11. Let and . Then is dense in . In particular, is dense in .

Proof. Take any and any . In the appendix we show that admits smooth atomic decomposition. By the smooth atomic decomposition, we see that is dense in , for and . Thus, we can find such that . On the other hand, the argument in Step 5 of the proof of [16, Theorem 2.3.3] shows that there exists a sufficiently large such that . Now we put Then , and we have This proves the claimed statement.

We next consider lifting property of . For , the power is naturally given by

Remark 12. By [17, Theorem 13.24], we have for all . As a consequence, is dense in , for all and .
We now have the lifting property of .

Theorem 13. Let and , .(i)The operator , initially defined on , extends to a continuous operator from to .(ii)Let denote the continuous extension of . Then is an isomorphism, and is an equivalent quasi-norm of .

Proof. (i) Let . Set and . Clearly is also in , and . By [17, Theorem 13.24], we have , and moreover Hence, for every , we have .
Now let . By the above remarks, we have . It follows that Since is dense in (see Remark 12), the mapping extends to a continuous operator from to . We denote this extension by .
(ii) Let us first show that the mapping is injective. Indeed, assume such that is the zero element of . By Remark 12 we can find a sequence in which converges in to . Then applying (i) to yields that converges in to the zero element. Since and , applying (i) to the operator we see that converges in to the zero element. Therefore, is the zero element in . This proves that is injective.
Next we show that is surjective. Indeed, given , we let be a sequence in which converges in to . Then from (i) we see that converges in . Denote this limit by . We claim that . Indeed, since , it follows from (i) that converges to in . Hence in . This proves that is surjective.
The above arguments also show that both and are identity operators on . Furthermore, by an easy density argument we see that (61) holds for all , provided that in (61) is replaced by . Thus, is an isomorphism, and is an equivalent quasi-norm of .

Our next goal is to show the Lusin and Littlewood-Paley function characterizations of . If , , , , and is a function on , we define The following proposition shows that the spaces are characterized by Lusin and Littlewood-Paley functions.

Proposition 14. Let , , , and . Let and . Put , for and . Then one has

Proof. Step  1. Show that if then Indeed, the proof of the first inequality in (64) is essentially the same as that of [18, Theorem 2.3]. To see the second inequality in (64), one only needs to examine the proofs of Theorems 1, 2 and 4 in [19, Chapter 4] and observe that, although the function considered in [19] is defined on the half space , the arguments there can also be adapted to functions which are defined on .
Step  2. We prove that . First note that, by an argument similar to the proofs of [19, Theorems 1 and 2], it follows that for every fixed . Hence, in view of (38), it is enough to show that But this is a consequence of the following elementary estimate: The proof of Proposition 14 is thus complete.

Corollary 15. Let be a stratified Lie group. Then with equivalent (quasi-)norms, for . Here are Hardy spaces on .

Proof. In [3, Chapter 7], Folland and Stein proved the characterization of Hardy spaces by continuous version Lusin function, for . Note that the arguments in [3, Chapter 7] are still valid if we replace the continuous version Lusin function by discrete version one defined above; see also [20] for a treatment of discrete version Lusin funcion. This fact together with Proposition 14 yield the identification of with for .

4. Convolution Singular Integral Operators on

In this section we study boundedness of convolution singular integral operators on homogeneous Triebel-Lizorkin spaces on stratified Lie groups. Motivated by [21, Section 5.3 in Chapter XIII], we introduce a class of singular convolution kernels as follows.

Definition 16. Let be a positive integer. A kernel of order is a distribution with the following properties:
(i) coincides with a function away from the group identity and enjoys the regularity condition:
(ii) satisfies the cancellation condition: For all normalized bump function and all , we have where , and is a constant independent of and . Here, by a normalized bump function we mean a function supported in and satisfying for some fixed positive integer .
The convolution operator with kernel of order is called a singular integral operator of order .

Remark 17. Using [3, Proposition 1.29], it is easy to verify that (67) is equivalent to the following condition:

Examples of such kernels include the class of distributions which are homogeneous of degree (see Folland and Stein [3, p. 11] for definition) and agree with functions away from . Indeed, assume is such a distribution, then it is easy to verify that satisfies the regularity condition (i) in Definition 16; moreover, from [3, Proposition 6.13] we see that is a principle value distribution such that for all . Hence, for every normalized bump function , by the homogeneity of we have Using stratified mean value theorem (cf. [3, Theorem 1.41]) and (67)–(69), it is easy to verify that the last integral converges absolutely and is bounded by a constant independent of and . Hence satisfies the condition (ii) in Definition 16.

Now we state the main result of this section.

Theorem 18. Let , , and let be a positive integer such that . Suppose is a singular integral operator of order . Then extends to a bounded operator on .

If and , we define as the tempered distribution given by (). For the proof of Theorem 18, we will need the following lemma, in which is the positive constant as in [3, Corollary 1.44].

Lemma 19. Let be a positive integer. Suppose is a kernel of order , and is a smooth function supported in and having vanishing moments of order . Then, there exists a constant such that for all , all with , and all
Moreover, have vanishing moments of the same order as .

Proof. Recall that the convolution of with is defined by , where is the function given by , and as before for any function . From [3, p. 38] we see that are functions, . We claim that for every with , the function is a normalized bump function multiplied with a constant independent of . Indeed, using the quasi-triangle inequality satisfied by the homogeneous norm it is easy to verify that the function is supported in ; moreover, since and since where are polynomials of homogeneous degree (see [3, Proposition 1.29]), we have Here is a constant depending on but not on . Hence the claim is true. The above argument also shows that, for every with and for every , is a normalized bump function multiplied with a constant independent of . Thus, by the condition (ii) in Definition 16, there exits a constant such that for all , all with , and all with From this and [3, Proposition 1.29], we also get that, for all , all with , and all with
Let now . Let . Let with , and denote by the right Taylor polynomial of at of homogeneous degree (see [3, pp. 26-27]). Then by the right-invariant version of [3, Corollary 1.44], we have Observe that satisfies (67) with the bound independent of . Also note that for and we have . Thus, for all with and all with , by using (73) and (67) (with replaced by ) we have Here, for the second inequality we also used the observation that when and , we have . Inserting (78) into (77) we obtain Notice that for , and , we have . Thus, by using the vanishing moments of and (79), we have Combining (76) and (80), we see that, for all , all with , and all From this and (73), we also get that, for all , all with , and all Since , (81) along with (82) yield (72).
It is straightforward to verify that have vanishing moments of the same order as . The proof of Lemma 19 is therefore complete.

The proof of Theorem 18 also relies on the existence of smooth functions with compact support and having arbitrarily high order vanishing moments.

Lemma 20. Given any nonnegative integer and any positive number , there exists a function with the following properties:(i) for , with some (large) positive integer;(ii) is a Schwartz function on having vanishing moments of order ;(iii).

Proof. From the appendix of [22] we see that there exists such that and has compact support. Now let us define , . Here and is a nonnegative integer. Then . Hence, if we take sufficiently large, then (ii) and (iii) follow immediately. Moreover, since , it is easy to see that (i) is also satisfied, provided that is sufficiently large.

We are now ready to prove Theorem 18.

Proof of Theorem 18. Choose a function which satisfies conditions (i)–(iii) in Lemma 20 with and . The condition (i) guarantees the existence of a function with the following properties: Note that . For , by Lemma 3 we have with convergence in . Let , and let be the convolution kernel of the operaotor . Then we have the representation which holds pointwise and also in the sense of . Since (by Lemma 19) satisfies the decay condition (72) (with the bound independent of ) and has vanishing moments of the same order as , from the proof of Lemma 2 we see that This together with (85) gives that where for the last inequality we used that . By the hypothesis we can choose such that and . From [3, Corollary 1.17] we see that the last integral converges absolutely. Consequently, we obtain Hence, it follows by [15, Lemma 2] that This together with (38) imply that for all . Since is dense in , extends to an bounded operator on . This completes the proof of Theorem 18.

Corollary 21. Let , , and let be a nonnegative integer. Then

Proof. Note that by the Poincaré-Birkhoff-Witt theorem (cf. [23, I.2.7]), the operators form a basis of the algebra of the left-invariant differential operators on . By this fact and the stratification of , it suffices to show that To this end, we first note that when restricted to Schwartz functions, are convolution operators with distribution kernels homogeneous of degree and coincide with smooth functions in . This follows from the fact that the operator is a convolution operator whose distribution kernel is homogeneous of degree and coincides with a smooth function in (see [2, Proposition 3.17]). Hence, by Theorem 18, extend to bounded operators on . From this fact and the lifting property (Theorem 7), we deduce that Hence . To see the converse, we need to use [2, Lemma 4.12], which asserts that there exists tempered distributions homogeneous of degree and coinciding with smooth functions in such that for all . By this result and Theorem 7, we have, at least for (), where is the convolution kernel of the operator . As is indicated in [2, p. 190], are distributions homogeneous of degree and coincide with smooth functions away from . Thus it follows by Theorem 18 that Inserting this into (93), we obtain for all . Since is dense in , the latter inequality also holds for all . This completes the proof.

Appendix

Smooth Atomic Decomposition of

In this appendix we show that homogeneous Triebel-Lizorkin spaces on stratified Lie groups admit smooth atomic decomposition. We follow the proof of [24, Theorem 6.6.3] with necessary modifications.

Equipped with Haar measure and the quasi-distance defined by the homogeneous norm, the group is a space of homogeneous type in the sense of Coifman and Weiss [25]. On such type of spaces, Christ [26] constructed a dyadic grid analogous to that of the Euclidean space as follows.

Lemma A.1. Let be a stratified Lie group. There exists a collection of open subsets of , where is some (possibly finite) index set, and constants and such that(i) for each fixed and if ;(ii)for any with , either or ;(iii)for each and , there exists a unique such that ;(iv), where ;(v)each contains some ball , where .

The set can be thought of as a dyadic cube with diameter roughly and centered at . We denote by the family of all dyadic cubes on . For , we set , so that . For any dyadic cube , we denote by the “center” of and by the unique integer such that .

Without loss of generality, in what follows we assume . Otherwise we need to replace in Definition 4 by , and also make some other necessary changes; see [27, pp. 96–98] for more details.

Definition A.2. Let be a dyadic cube and let be a nonnegative integer. A smooth function on is called a smooth L-atom for if it satisfies(i) is supported in ;(ii) for all ;(iii) for all multi-indices with .

Definition A.3. Let and . The sequence space consists of all sequences such that the function is in . For such a sequence we set

The smooth atomic decomposition of homogeneous Triebel-Lizorkin spaces on stratified Lie groups can be stated as follows.

Theorem A.4. Let , , and let be a nonnegative integer satisfying . Then there is a constant such that for every sequence of smooth -atoms and every sequence of complex scalars one has Conversely, there is a constant such that given any distribution and any , there exists a sequence of smooth -atoms such that where the sum converges in and moreover

Proof. Let for some , and let be a smooth -atom for . We set . Then is supported in ; moreover (since for supp ), where can be chosen to be arbitrarily large. Set , that is, . Then the above inequality can be rewritten as Using this estimate, (73), and that , we also deduce that Thus, it follows from Lemma 2 that where , and can be taken to be arbitrarily large. Consequently,
Note that if and then we have Here we used the fact that for and one has , which can be easily verified by using Lemma 19 (iv) and the quasi-triangle inequality satisfied by the homogeneous norm. The above estimate and (A.10), along with the argument in [24, pp. 80-81], yield (A.3).
Now we show the converse statement of the theorem. By Lemma 20, there exists such that for for some (large) positive integer , and that is supported in and has vanishing moments of order . Then it is possible to find a function with the properties that , for , and for . Note that . Given , it follows from Lemma 3 that with the convergence in . Let us decompose as where we have set, for , It is straightforward to verify that is supported in , and that has vanishing moments of the same order as . Moreover, for we have for all . Hence the function a smooth -atom for . Choose any . We note that We thus obtain, using (38), This proves (A.5).

Acknowledgments

The author would like to thank Professor Hitoshi Arai for his patient guidance and constant support. He is also grateful to Professor Yoshihiro Sawano for his valuable comments.