Abstract

The boundedness of the various operators on -Morrey spaces is considered in the framework of the Littlewood-Paley decompositions. First, the Littlewood-Paley characterization of -Morrey-Campanato spaces is established. As an application, the boundedness of Riesz potential operators is revisted. Also, a characterization of -Lipschitz spaces is obtained: and, as an application, the boundedness of Riesz potential operators on -Lipschitz spaces is discussed.

1. Introduction

One of the main aims of the present paper is to investigate the structure of the -spaces by proving Theorem 1 below. To state the main result, we need the following setup. We write for and . We abbreviate to . For a set , we denote by its indicator function. The space denotes the set of all compactly supported -functions supported on . Now we formulate our main result in the simplest form.

Theorem 1. Let , and . Let be a function such that If a measurable function belongs to , that is, satisfies then and where denotes the inverse Fourier transform and the implicit constant in does not depend on .

When , then this is due to Mazzucato [1]. See [2] as well.

The space appearing in Theorem 1 is a special case of spaces. Theory of spaces stems from the Beurling work [3]. Beurling introduced the space together with its predual so-called the Beurling algebra [3]. Later, to extend Wiener’s ideas [4, 5] which describe the behavior of functions at infinity, Feichtinger [6] gave an equivalent norm on , which is a special case of norms to describe nonhomogeneous Herz spaces introduced in [7]. On the other hand, Lu and Yang [8, 9] introduced the central bounded mean oscillation space with the norm Recently, in [10], -Morrey-Campanato spaces have been introduced to unify central Morrey spaces, central bounded mean oscillation spaces, and usual Morrey-Campanato spaces. Using -spaces, we can study both local and global regularities of functions simultaneously. For example, when we consider , the underlying norm given by measures local regularities of functions, and the parameter plays the role of global regularity. Moreover, the proof of Theorem 13 reveals this aspect of our studying both local and global regularities of functions simultaneously.

Theorem 1 concerns the Littlewood-Paley theory. The Littlewood-Paley theory is one of the most powerful tools in harmonic analysis. Roughly speaking, this is a technique of transforming functions into good ones in order to measure the norms. Another side of Littlewood-Paley theory is that the functions are broken into good pieces of functions. We shall illustrate that the Littlewood-Paley theory is very useful by applying this to the fractional integral operator of order . The fractional integral operator , which is given by plays an important role not only in harmonic analysis but also in partial differential equations. It is well known that can be seen as the inverse operator of modulo a multiplicative constant, and hence has the smoothing effect. However, due to this smoothing effect, still seems to have a lot to be investigated.

To describe the related function spaces and formulate Theorem 1 in the full statement, we now fix some more notations. Since cubes play the central role, let us fix our notation of cubes. We denote by the set of all cubes of the form with and . Also, we denote by the set of dyadic cubes: where to define the right-hand side we used the Minkowski sum. Given , we denote by the center of a cube and by its volume. The notation stands for . Given a set , we define

With these notations in mind, let us recall the definition of the function space and related function spaces defined in [10]. Here we redefine in terms of dyadic cubes. However, a geometric observation shows that the definition of through dyadic cubes and that through cubes are equivalent.

Definition 2. Let , , and . Define as the set of all such that , where Define in analogy with . The norm is given by

As we remarked above, (3) and (10) are equivalent definitions, and the same can be said for weak type spaces. Namely, for any measurable function on . In view of (12), we identify the right-hand side and the left-hand side in these formulas. Note that from the definition of norms, we have for any measurable function on .

We need to pay attention to the word “local,” when we discuss . Burenkov and Guliyev, together with their successors, investigated local Morrey-type spaces in [11]. By “local” in [11], they meant that they defined the “local Morrey-type” norm by which indicates that appearing in the definition is restricted to all balls centered at the origin. However, by “local” in the present paper we meant that we measure the regularity of functions by (6). Nowadays there are many different definitions related to classical Morrey spaces, so we need to carefully distinguish the different definitions and the names given to the definitions. See [12] for related usage of the word “local” such as central mean oscillation, central , -central bounded mean oscillation, and -central Morrey spaces.

The goal of the present paper is to show that these function spaces fall under the scope of the Littlewood-Paley theory. As an application of this fact we show the boundedness property of and singular integral operators. The Littlewood-Paley theory is a powerful tool to investigate the boundedness property of . To consider the connection between and the Littlewood-Paley theory, we present definitions. Here and below we use the definition of the Fourier transform below for definiteness: Let be a function such that Following [13], we define It may be helpful to observe that for each .

In the present paper, the following function space of Littlewood-Paley type will play a key role. Here and below we denote for a cube . Observe that

Definition 3. Let . Given , set The function space denotes the set of all tempered distributions for which the quantity is finite.

Note that Definition 3 is closely related to the space defined by Yang and Yuan and the function space is investigated in [1421].

In Theorem 1, we did not mention what happens if , and the right-hand side of (4) is finite. Here including this problem, we reformulate and reinforce Theorem 1.

Theorem 4. Let . Then the spaces and are isomorphic. More precisely, we have the following.(a) in the sense of continuous embedding. (b) in the sense of continuous embedding. (c)Let . If , then . Conversely, if , then there exists a polynomial such that , and in this case the norms and are equivalent. (d)Different choices of satisfying (17) will yield the equivalent norms in the definition of .

Instead of Theorem 1 itself, we shall prove Theorem 4 in the present paper.

Let and . The Morrey space , which can be realized as , admits two different Littlewood-Paley characterizations. In [1], Mazzucato established Meanwhile, combining this equivalence with what we proved in [15], we can say that Thus, (23) is closer to Definition 3 than (22). Also, (23) seems to have stemmed from the famous technique due to Uchiyama [22]. We take advantage of equivalence of Theorem 4 in the proof of Theorem 6.

To establish Theorem 4, we will need an auxiliary vector-valued estimate of the Hardy-Littlewood maximal operator . Define the Hardy-Littlewood maximal operator by The following proposition is proven in our earlier paper [12]. This is an extension of [23] to .

Proposition 5. Let , and . Assume in addition that . Then we have for some independent of , where we modify (25) obviously when .

Chiarenza and Frasca obtained the boundedness of Hardy-Littlewood maximal operators on global Morrey spaces in [24]. In [11] Burenkov and Guliyev considered local Morrey-type spaces, where they showed that maximal operators are bounded [2527].

With Theorem 4 and Proposition 5 in mind, we investigate the boundedness property of again as we announced in the beginning. More precisely, we shall provide an alternative proof of the following theorems, which were proven earlier in [10, 12].

Theorem 6 (see [10, 12]). Suppose that the parameters , and satisfy Assume in addition that Then is a bounded operator from to .

Theorem 7 (see [10, 12]). Suppose that the parameters , and satisfy (26) and with . Then is a bounded linear operator from to .

We can also consider Campanato spaces and Lipschitz spaces in this framework. First, let be the set of all polynomials having a degree at most . For a cube , a locally integrable function over and a nonnegative integer , there exists a unique polynomial such that, for all , Denote this unique polynomial by . It follows immediately from the definition that if . We can characterize spaces and spaces (cf. [28]).

Definition 8. Let and . Then, for , define that where we use the obvious modification when . The spaces and are the sets modulo of all for which the quantities and are finite, respectively.

Note that, in particular, with norm coincidence. For the definition of and , we refer to [12].

Definition 9. For , , and with , let be the set of all such that , where where we use the obvious modification when .

Observe that the quotient space is a Banach space equipped with the norm .

Proposition 10. Let , , . Assume that and that . Then with equivalent norms.

Now we define a function space by way of difference. For , an integer and a function , we define inductively.

Definition 11. For , and with , let be the set of all continuous functions such that , where
Then is a Banach space equipped with the norm and also the quotient space is a Banach space equipped with the norm .

Definition 12. For and let be the set of all such that , where
In the present paper, we aim to show the following equivalence as well. Here, for , we let the largest integer such that .

Theorem 13. Let and . Assume that the integer satisfies . (1)If is a continuous function such that the quantity is finite, then and the inequality holds. (2)If , then can be represented by a continuous function and the inequality holds.

Remark 14. It is absolutely necessary to assume that is a continuous function in Theorem 13, when . We remark that there exists a discontinuous function such that for all . See [29, Proposition A1] for such an example constructed algebraically.

Now we explain notations and we describe its organization of the present paper. We use the following notations for the inequalities. First, we use standard notation for inequalities. For example, in the present paper a chain of inequalities of the form appears in (110) below. The inequality (38) means that there exist such that If the implicit constants in or do depend on some important parameters , then we write or . We shall prove Theorem 4 in Section 3. We prove Theorem 6 in Section 4. Theorem 7 is covered in Section 5. We prove Theorem 13 in Section 6. Finally in Section 7 we present another application of Theorem 4 by showing that the Fourier multiplier is bounded on .

2. Preliminaries

In the present paper, we frequently use the following fundamental inequalities.

Lemma 15 (see [30, page 466]). Let , and satisfy Suppose that such that for some . Assume in addition that is a measurable function such that and that for some . Then we have

To formulate the next lemma, we recall the definition of with . A measurable function , which takes values in almost everywhere, is said to be an weight or belongs to the class , if For all , it is easy to see that and that .

Lemma 16 (see [31]). Let and satisfy (17). Then, for , we have for all .

We also need a piece of information on dilation of the space .

Lemma 17. Let , and . Let and . Then

It is just a matter of handling the left-hand side carefully. But, for the sake of convenience, we supply the proof.

Proof. From the definition of , we deduce proving the lemma.

In the course of the proof of Theorem 6, we need another piece of information on the space . Let , and . Define as the set of all such that , where

The next lemma concerns the norm of the translation operator.

Lemma 18. Let , and . Then

Proof. Let and be fixed and consider Then we have and . Note that since . This implies that If we consider the supremum over and such that , then we obtain the desired result.

Lemma 19. Let . Let , and . Then for all .

Proof. We have from Lemma 18. Hence, by the triangle inequality, we obtain that Meanwhile, since , for , we have, on , Assuming that , we have Putting (57)–(59) together, we obtain the desired result.

Lemma 20. Suppose that the parameters , and satisfy Let . Then

Proof. Note that for . Recall that has a scaling invariance, as we have verified in Lemma 17. So, to prove (61), we can assume that . In this case, (61) is Lemma 19 itself.

We need the following sequence of functions.

Lemma 21. There exists a sequence of functions such that, for any fixed constant , provided .

In the present paper the sequence above is called a Rademacher sequence.

3. Littlewood-Paley Characterization of

3.1. Proof of Proposition 10

It follows from the definition of the norm that for any constant function . Note that this implies that is a subset of .

Let . Then, since , we can use the Hölder inequality and we have for . Assuming that , the sequence is convergent by the Cauchy test. Thus, we can consider the mapping Let us check that the range of is in .

Let and be fixed. Let be the largest integer such that . Then observe that . We decompose We consider the -norm over and multiply .

As for the first term, directly from the definition of , we obtain that Also, a geometric observation shows that the second term can be estimated similarly. Since , we obtain that By using (64) and , we can handle the third term: In view of the way in which we chose , we obtain that In summary, It remains to estimate the fourth term. We employ the following estimate: Since , (72) is summable and we obtain that where the implicit constant in (73) is independent of .

It follows from (67), (68), (71), and (73) that sends to boundedly; for all .

Meanwhile, it follows from the norm that for all . In view of (63), is a surjection:

Finally observe that, for , if and only if is a constant function, that is, . Namely, Thus, from (75), (76) and (77), we conclude (33).

3.2. Proof of Theorem 4 Part (a)

For and , let us define the function space of uniformly functions by where, for , we write , and the norm is given by Then from the definition of the norms (10) and (78), the following chain of continuous inclusion holds. For , Thus, (a) follows.

Remark 22. The space is a special case of amalgams investigated in [32].

3.3. Proof of Theorem 4 Part (b)

Let be a fixed function satisfying (17). It follows from (17) that there exist and such that Let and fix a dyadic cube such that . We are going to show that converges in the topology of , and that converges in the topology of .

The presice meaning of (82) and (83) is that as follows: (82) means as and (83) means that as . Once (82) and (83) are proven, we will have converges in the topology of and that is a polynomial. Hence it follows that . Remark also that the convergence in of the sum defining is a generality. So let us prove (82) and (83).

Let us begin with proving (82). To this end we take a dyadic cube containing . Since we are assuming (81), we deduce that Since , we have from (19) and (20). By the Hölder inequality and the fact that , we have for all . By decomposing the last integral dyadically, we obtain that Observe from Definition 3 that for all . Thus, we obtain Hence, from (92) and the fact that , we have which proves the embedding into . To prove that the sum defining converges in the topology of , we first fix a cube containing and observe from (91) and the fact that . Meanwhile, again by (92), we have for all cubes containing . As a result, we obtain that from the fact that . It follows from (94) and (96) that the sum defining converges in .

Now let us prove (83). First, choose . By virtue of Lemma 16 and the fact that has a bound independent of [33], we obtain that Here for the last inequality, we employed a geometric observation of the support of . Therefore, of the last inequality depends only on .

Since , we have Recall that we are assuming that . So (98) is summable and we have showing (83).

With (82) and (83), the proof of (b) is now complete.

Remark 23. We did not use the structure of in the proof. Therefore, we can deduce the following variant.

Proposition 24. Suppose that we are given a collection such that and that Then converges in the topology of .

3.4. Proof of Theorem 4 Part (c)

Keeping what we proved in Parts (a) and (b) in mind, we shall now describe the isomorphism between and . If , then we claim that and that The proof of (101) is comparably easy.

Let us start to prove (101). Let and be fixed. Also, fix a Rademacher sequence . Define Then we have by the property of the Rademacher sequence. Let us set Then we have since with constant independent of .

By the Calderón-Zygmund theory for , we have As for , we invoke the following pointwise estimate obtained from (105): We write Notice that since . Assuming that , we have Consequently, from (106) and (110), we obtain (101).

The heart of the matter is how to construct the inverse mapping from to . Let . Then we have established in (b) that exists in the topology of in view of (92) and (99). We claim that Here and below in proving (112), we assume that by replacing with .

Let us prove (112). Suppose that we are given a dyadic cube and such that . We estimate First, we expand by using (111). If we denote for , then we have where the convergence, as we have established in Part (b), takes places in .

We set for .

We decompose (83) by using and . by using (115).

We estimate . Let be an auxiliary parameter again, which is taken so that By using the maximal operator , we have By Lemma 16, we have We write We decompose again the right-hand side of (120) dyadically. For each , we have , since . Thus, Consequently, from (118), (122), and (123), we obtain that Thus, the estimate for is now valid.

As for we choose first. Since , we have By the Hölder inequality, we have We have, for and , Consequently, from (126) and (127), we have We decompose the integral (128) dyadically: Notice that , since . Consequently, assuming that , we have We are going to add (130) over . To this end, observe that If we change variables, then we have Now that and , we obtain As a result, from (133) and the fact that , the estimate (130) is summable over , and we obtain that Consequently, (83) is proved, and we conclude the proof of Theorem 4(c).

3.5. Proof of Theorem 4 Part (d)

An important corollary of Proposition 5 is that we can relax the condition (17) on , which is stronger than (d) in Theorem 4.

Corollary 25. Let be such that Define Then the norm equivalence holds.

Proof. Choose integers and so that Define Then by virtue of (138). Observe also that in view of the size of supports. Hence, we have Since the numbers appearing in (138) are finite, we have from (141) and Proposition 5. The proof of is analogous; just swap the role of and . With (142) and (143), we see that the norms in (137) are equivalent.

4. - Boundedness of , Proof of Theorem 6

To prove Theorem 6, we need the lemmas. Note that Lemma 26 can be seen as the Plancherel-Polya-Nikolskij inequality for .

Lemma 26. Suppose that the parameters , and satisfy Let be chosen so that it satisfies (17). Then we have for all and .

Proof. The right-hand inequality is a consequence of Lemma 20.
Let be fixed. We choose sufficiently large. Also, take a compactly supported function so that equals on as we did in (138). Then we have We decompose dyadically. Then we have Thus, the proof is now complete.

In the course of the proof, we obtained the following chain of inequalities.

Corollary 27. Suppose that the parameters , and satisfy Let be chosen so that it satisfies (17). Then one has for , and such that .

The following estimate is somehow well known. Here we remark that the following form was recorded in [34]. Here we change variables to transform the result into the one needed in the present paper.

Lemma 28. Let . Then

Now we prove Theorem 6. Given , we shall define Note that, by using the Fourier transform, can be defined. Once we show that satisfies the condition in Proposition 24, then by Proposition 24   can be defined. Since this essentially amounts to showing by assuming that we can define , we omit the detail.

Since , we have by virtue of Theorem 4. Likewise since , we have Let be a fixed dyadic cube. Define . Then we have (135) and for and . Thus, we have since . If we use the embedding , then we have By Corollary 27 and , we have Now we invoke Lemma 28 with Since , we have . Thus, from Lemma 28, we obtain that We consider the -norm of the both sides. Then by the relation , we deduce In view of the definition of , we have Putting (156), (157) and (162) together, we obtain Consequently, from (153), Corollary 25 and the definition of the norm , we obtain By combining (164) with (154), we conclude that is a bounded operator from to .

Remark 29. Suppose that the parameters , and satisfy and (27). If we reexamine (157), we see that

5. - Boundedness of , Proof of Theorem 7

The definition of can be justified in the same was as Theorem 6. So we assume that is already defined.

Let and a cube be fixed so that . Set and for some small . More precisely, we choose so that By interpolation described in [35] we have We calculate the -functional. By the definition of the -functional [35, Section 3.1] and by Remark 29, we have Note that . Hence we have

Let us define By Lemma 20 and Remark 29, we have From (172) and (173), we deduce that If we insert this estimate and (174), then we have The proof is therefore complete.

6. Structure of , Proof of Theorem 13

We prove Theorem 13. Let us recall (13), which we use now. To prove Theorem 13, we need the following lemma.

Lemma 30. For all multi-indices , , , we have

Proof. To prove this, we fix and so that Also, we take so that on . We set for as usual. Then we have Now that , we have This proves the lemma.

Proposition 31 (see [36]). Let . Suppose that we are given a sequence of such that is convergent for each with . Then there exists a sequence of polynomials such that is convergent in .

The next proposition reveals the fundamental structure of .

Proposition 32. Assume that the parameters and satisfy Let .   (1)There exists a sequence of polynomials of degree at most such that converges in . (2) If we denote by the limit (182), then (3) The limit is determined modulo a difference of polynomials of order no matter how we choose . (4) The distribution is a polynomial. (5) The distribution is a continuous function.

Proof. (1) Let and . Choose so that . Then we have for some cube with . By Lemma 30, we have Consequently, then we have The estimate (185) shows that converges in the topology of , provided that . So we are in the position of resorting to Proposition 31. (2) This is clear since . (3) Suppose that is a sequence of polynomials of degree such that converges in the topology of . Then we claim for all with length . Indeed, Consequently, and coincide modulo a polynomial of order . (4) Just observe that has frequency support in the origin. (5) In view of (4), we have only to show that itself is a continuous function. As we can see from (185), if , then the limit defines a function. If are multi-indices such that and that , then Consequently, there exists a function of polynomial growth such that .Observe also that when , Consequently, since , by virtue of the Weierstrass test, we see the series converges uniformly to a continuous function . Thus, is a desired function.

 Here and below we assume that itself satisfies and that is a continuous function.

Lemma 33. Let satisfy (192). Then we have in the topology of .

Remark that the left-hand side equals .

Proof. By (192), we have The proof is therefore completed.

Proof of Theorem 13. Choose such that . Let us define Then equals a nonzero constant near a neighborhood of the origin and has compact support. Let us set . From the above observation, can be used for the definition of .
Suppose that is a continuous function such that the quantity is finite. Let be a cube such that and fix . Then we have and hence Observe that and hence, for , From this estimate and the fact that and , we obtain that The estimate (200) shows that .
Conversely, suppose that . We need to prove that , that is, we need to find a constant such that for any , any cube and any pair of points satisfying and . Then, from , we have by Lemma 33. Then by the mean value theorem, we have and, by using the triangle inequality applied directly to the summation defining , we have Consequently, we have It follows directly from the definition of the norm that and from Lemma 30 and (181) that Consequently, by Lemma 28, we have Now that and , we deduce that This concludes the proof of Theorem 13.

Going through an argument similar to Theorem 6, we obtain the following.

Theorem 34. Let , and . Then the fractional integral operator is bounded from to .

As a corollary of Theorems 13 and 34, we have the following conclusion.

Corollary 35. Let , and . Assume in addition that the integer satisfies . Then the fractional integral operator is bounded from to .

7. Application of Theorem 4 to the Fourier Multipliers

As a corollary of Theorem 4, the boundedness of singular integral operators follows.

Corollary 36. Let be such that for all multi-indices . Then the mapping extends naturally to a -bounded operator.

Proof. To check this, by virtue of Theorem 4, we need only to verify for all and , where satisfies (17). As usual we take and so that it satisfies (138). Write Then it is easy to verify that and that . Consequently, we have Recall that we defined . Hence, a direct calculation yields If we insert (214) to (215), then we obtain that If we combine Proposition 5 with (216), then we have Thus, (211) is established.

Acknowledgments

The first author is partially supported by Grant-in-Aid for Scientific Research (C), no. 21540199 and no. 24540194, Japan Society for the Promotion of Science. The second author was partially supported by Grant for 2011 Overseas Researcher of Nihon University, Japan. The third author is partially supported by Grant-in-Aid for Scientific Research (C), no. 20540167 and no. 24540159, Japan Society for the Promotion of Science. The fourth author is supported financially by Grant-in-Aid for Young Scientists (B), no. 21740104 and no. 24740085, Japan Society for the Promotion of Science. The authors are thankful to the anonymous referee for his/her pointing out some improvements about the presentation of mathematics.