Duality of Hardy and BMO spaces associated with operators with heat kernel bounds

By Xuan Thinh Duong and Lixin Yan

Abstract

Let be the infinitesimal generator of an analytic semigroup on with suitable upper bounds on its heat kernels. Auscher, Duong, and McIntosh defined a Hardy space by means of an area integral function associated with the operator . By using a variant of the maximal function associated with the semigroup , a space of functions of BMO type was defined by Duong and Yan and it generalizes the classical BMO space. In this paper, we show that if has a bounded holomorphic functional calculus on , then the dual space of is where is the adjoint operator of . We then obtain a characterization of the space in terms of the Carleson measure. We also discuss the dimensions of the kernel spaces of BMO when is a second-order elliptic operator of divergence form and when is a Schrödinger operator, and study the inclusion between the classical BMO space and spaces associated with operators.

1. Introduction

The introduction and development of Hardy and BMO spaces on Euclidean spaces in the 1960s and 1970s played an important role in modern harmonic analysis and applications in partial differential equations. These spaces were studied extensively in Reference 32, Reference 22, Reference 18, Reference 19, Reference 31 and many others.

An function on is in the Hardy space if the area integral function of the Poisson integral satisfies

There are a number of equivalent characterizations of functions in the space, including the all-important atomic decomposition (see Reference 21, Reference 31).

A locally integrable function defined on is said to be in BMO, the space of functions of bounded mean oscillation, if

where the supremum is taken over all balls in , and stands for the mean of over , i.e.,

In Reference 19, Fefferman and Stein showed that the space BMO is the dual space of the Hardy space . They also obtained a characterization of the BMO space in terms of the Carleson measure, the - boundedness of convolution operators which satisfy the Hörmander condition, and an interpolation theorem between spaces and the BMO space. From the viewpoint of Calderón-Zygmund operator theory, and BMO spaces are natural substitutes for and spaces, respectively.

Recently, Auscher, McIntosh and the first-named author introduced a class of Hardy spaces associated with an operator by means of the area integral functions in (Equation 1.1) in which the Poisson semigroup was replaced by the semigroup (Reference 4). They then obtained an -molecular characterization for by using the theory of tent spaces developed by Coifman, Meyer and Stein (Reference 7, Reference 8 and Reference 4). See also Sections 3.2.1 and 4.1 below. In Reference 16, we introduced and developed a new function space BMO associated with an operator by using a maximal function introduced by Martell in Reference 25. Roughly speaking, if is the infinitesimal generator of an analytic semigroup on with kernel (which decays fast enough), we can view as an average version of (at the scale ) and use the quantity

to replace the mean value in the definition (Equation 1.2) of the classical BMO space, where is scaled to the radius of the ball We then say that a function (with suitable bounds on growth) is in if

See Section 3.2.2 below. We also studied and established a number of important features of the space such as the John-Nirenberg inequality and complex interpolation (Reference 16, Section 3). Note that the spaces and coincide with the classical Hardy and BMO spaces, respectively (Reference 16, Section 2).

The main purpose of this paper is to prove a generalization of Fefferman and Stein’s result on the duality of and BMO spaces. We will show that if has a bounded holomorphic functional calculus on and the kernel of the operator in (Equation 1.3) satisfies an upper bound of Poisson type, then the space is the dual space of the Hardy space in which denotes the adjoint operator of We also obtain a characterization of functions in in terms of the Carleson measure. See Theorems 3.1 and 3.2 below.

We note that a valid choice of in (Equation 1.3) is the Poisson integral , which is defined by

For this choice of , Theorems 3.1 and 3.2 of this article give the classical results of Theorem 2 and the equivalence (i) (iii) of Theorem 3 of Reference 19, respectively. See also Chapter IV of Reference 31.

Note that in our main result, Theorem 3.1, we assume only an upper bound on the kernel of in (Equation 1.3) and no regularities on the space variables or . Another feature of our result is that we do not assume the conservation property of the semigroup for . This allows our method to be applicable to a large class of operators .

The paper is organised as follows. In Section 2 we will give some preliminaries on holomorphic functional calculi of operators and on integral operators with kernels satisfying upper bounds of Poisson type. In Section 3 we introduce and describe the assumptions of the operator in this paper, and recall the definitions of and BMO spaces as in Reference 4 and Reference 16. We then state our main result, Theorem 3.1, which says that the dual space of is . In Section 4 we prove a number of important estimates for functions in and BMO spaces. We then prove Theorem 3.1 in Section 5 by combining the key estimates of Section 4 with certain estimates using the theory of tent spaces and Carleson measures. In Section 6, we study the dimensions of the kernel spaces of BMO when is a second-order elliptic operator of divergence form and when is a Schrödinger operator. We conclude this article with a study of inclusion between the classical BMO space and spaces associated with some differential operators, including a sufficient condition for the classical BMO and spaces to coincide.

Throughout this paper, the letter will denote (possibly different) constants that are independent of the essential variables.

2. Preliminaries

We first give some preliminary definitions of holomorphic functional calculi as introduced by McIntosh Reference 26.

Let . We define the closed sector in the complex plane by

and denote the interior of by .

We employ the following subspaces of the space of all holomorphic functions on :

where , and

Let . A closed operator in is said to be of type if , and for each there exists a constant such that

If is of type and , we define by

where is the contour parametrized clockwise around , and . Clearly, this integral is absolutely convergent in , and it is straightforward to show, using Cauchy’s theorem, that the definition is independent of the choice of If, in addition, is one-one and has dense range and if , then can be defined by

where . It can be shown that is a well-defined linear operator in . We say that has a bounded calculus on if there exists such that , and for ,

For a detailed study of operators which have holomorphic functional calculi, see Reference 6.

In this paper, we will work with a class of integral operators , which plays the role of generalized approximations to the identity. We assume that for each , the operator is defined by its kernel in the sense that

for every function which satisfies the growth condition (Equation 3.3) in Section 3.1 below.

We also assume that the kernel of satisfies a Poisson bound of order

in which is a positive, bounded, decreasing function satisfying

for some .

It is easy to check that there exists a constant such that satisfies

uniformly in , . See Section 2 of Reference 14.

We recall that the Hardy-Littlewood maximal operator is defined by

where the sup is taken over all balls containing . It is well known that the Hardy-Littlewood maximal operator is bounded on for all . Because of the decay of the kernel in (Equation 2.2) and (Equation 2.3), one has

Proposition 2.1.

There exists a constant such that for any , we have

for all .

Proof.

This is a consequence of the conditions (Equation 2.2), (Equation 2.3) and the definition of See Reference 15, Proposition 2.4.

3. Duality between and spaces

In this section, we will give the framework and the main result of this paper.

3.1. Assumptions and notation

Let be a linear operator of type on with ; hence generates a holomorphic semigroup , . Assume the following two conditions.

Assumption (a).

The holomorphic semigroup , , is represented by the kernel which satisfies the upper bound

for , for and is defined on by (Equation 2.2).

Assumption (b).

The operator has a bounded -calculus on . That is, there exists such that , and for

for any .

We now give some consequences of assumptions (a) and (b) which will be useful in the sequel.

(i) If is a bounded analytic semigroup on whose kernel satisfies the estimate (Equation 2.2), then for all , the time derivatives of satisfy

for all and almost all . For each , the function might depend on but it always satisfies (Equation 2.3). See Lemma 2.5 of Reference 5.

(ii) has a bounded -calculus on if and only if for any non-zero function , satisfies the square function estimate and its reverse

for some , where . Note that different choices of and lead to equivalent quadratic norms of See Reference 26.

As noted in Reference 26, positive self-adjoint operators satisfy the quadratic estimate (Equation 3.2), as do normal operators with spectra in a sector, and maximal accretive operators. For definitions of these classes of operators, we refer the reader to Reference 36.

(iii) Under the assumptions (a) and (b), it was proved in Theorem 3.1 of Reference 15 and Theorem 6 of Reference 14 that the operator has a bounded holomorphic functional calculus on , ; that is, there exists such that , and for :

for any . For , the operator is of weak-type . In Reference 16, it was proved that for , the operator is bounded from into .

We now define the class of functions that the operators act upon. For any , a function is said to be a function of -type if satisfies

We denote by the collection of all functions of -type. If the norm of in is denoted by

It is easy to see that is a Banach space under the norm Note that we use instead of the space as in Reference 19 and Reference 16 since this gives the appropriate setting for the duality between and BMO. For any given operator , we let and define

Note that if is the Laplacian on , then . When , we have

For any and , we define

and

It follows from the estimate (Equation 3.1) that the operators and are well-defined. Moreover, the operator has the following properties:

(i) for any and almost all ,

(ii) the kernel of satisfies

where the function satisfies the condition (Equation 2.3). This property is the same as the estimate (Equation 3.1).

3.2. Hardy spaces and BMO spaces associated with operators

3.2.1. Hardy space

We assume that is an operator which satisfies the assumptions of Section 3.1. will denote the usual upper half-space in . The notation denotes the standard cone (of aperture ) with vertex . For any closed subset , will be the union of all cones with vertices in , i.e., If is an open subset of , then the “tent” over , denoted by , is given as .

Given a function , the area integral function associated with an operator is defined by

It follows from the assumption (b) of that the area integral function is bounded on (Reference 26). It then follows from the assumption (a) of that is bounded on , . See Theorem 6 of Reference 4. More specifically, there exist constants such that and

for all . See also Reference 35.

By duality, the operator also satisfies the estimate (Equation 3.8), where is the adjoint operator of .

The following definition was introduced in Reference 4. We say that belongs to a Hardy space associated with an operator , denoted by , if . We define its norm by

Note that if is the Laplacian on , then it follows from the area integral characterization of a Hardy space by using convolution that the classical space coincides with the spaces and and their norms are equivalent. See Reference 19 and Reference 31.

3.2.2. The function space

Following Reference 16, we say that is of bounded mean oscillation associated with an operator (abbreviated as ) if

where the sup is taken over all balls in , and is the radius of the ball . The class of functions of BMO, modulo , where

is a Banach space with the norm defined as in (Equation 3.9). We refer to Corollary 5.2 in Section 5 for completeness of the space BMO. See also Section 6.1 for a discussion of the kernel space .

We now give the following list of a number of important properties of the spaces BMO. For the proofs, we refer the reader to Sections 2 and 3 of Reference 16.

(i) If a function is in the classical space BMO, then it follows from the John-Nirenberg inequality that and . See Reference 22. Under the extra condition that satisfies the conservation property of the semigroup for every , it can be verified that BMO is a subspace of BMO. Moreover, the spaces BMO, and coincide and their norms are equivalent. See also Theorem 6.10 in Section 6.

(ii) If BMO, then for every and every , there exists a constant such that for almost all , we have

(iii) If BMO, then for any and any , there exists a constant which depends on such that

(iv) A variant of the John-Nirenberg inequality holds for functions in BMO. That is, there exist positive constants and such that for every ball and ,

This and (Equation 3.9) imply that for any BMO and , the norms

with different choices of are all equivalent.

3.3. Main theorems

We now state the main result of this paper.

Theorem 3.1.

Assume that the operator satisfies the assumptions (a) and (b) in Section . Denote by the adjoint operator of . Then, the dual space of the space is the BMO space, in the following sense.

(i) Suppose BMO. Then the linear functional given by

initially defined on the dense subspace , has a unique extension to .

(ii) Conversely, every continuous linear functional on the space can be realized as above; i.e., there exists BMO such that Equation 3.14 holds and

To state the next theorem, we recall that a measure defined on is said to be a Carleson measure if there is a positive constant such that for each ball on ,

where is the tent over . The smallest bound in (Equation 3.15) is defined to be the norm of and is denoted by .

The Carleson measure is closely related to the classical BMO space. We note that for every ,

is a Carleson measure on . See Reference 19 and Chapter 4 of Reference 21.

For the space BMO, we have the following characterization of BMO functions in terms of the Carleson measure.

Theorem 3.2.

Assume that the operator satisfies the assumptions (a) and (b) in Section . The following conditions are equivalent:

(i) is a function in BMO;

(ii) , and is a Carleson measure, with .

The proofs of Theorem 3.1 and the implication (ii) (i) of Theorem 3.2 will be given in Section 5. For the proof of the implication (i) (ii) of Theorem 3.2, we refer to Lemma 4.6 of Section 4.

Remark.

Using Theorems 3.1 and 3.2, we can obtain more information about the Hardy spaces and the spaces. We will discuss the inclusion between the classical BMO space and the spaces associated with some differential operators. See Section 6.

4. Properties of and BMO spaces

In Reference 7, Reference 8, Coifman, Meyer and Stein introduced and studied a new family of function spaces, the so-called “tent spaces”. These spaces are useful for the study of a variety of problems in harmonic analysis. In particular, we note that the tent spaces give a natural and simple approach to the atomic decomposition of functions in the classical Hardy space by using the area integral functions and the connection with the theory of Carleson measure. In this paper, we will adopt the same approach of tent spaces.

4.1. Tent spaces and applications

For any function defined on we will denote

and

As in Reference 8, the “tent space” is defined as the space of functions such that when The resulting equivalence classes are then equipped with the norm . When the space is the class of functions for which and the norm . Thus, if and only if , i.e.,

Next, a function is called a -atom if

the function is supported in (for some ball ;

The following proposition on duality and atomic decomposition for functions in was proved in Reference 8.

Proposition 4.1.

(a) The following inequality holds, whenever and

(b) The pairing

realizes as equivalent to the Banach space dual of

(c) Every element can be written as where the are atoms, , and

Proof.

For the proof of Proposition 4.1, we refer to Theorem 1 of Reference 8. See also Theorem 1 of Reference 11 for a proof of (a).

Proposition 4.1 gives a quick proof of the atomic decomposition for the classical Hardy space . Let . For any , we denote by the Poisson integral and set . The atomic decomposition of in leads to the atomic decomposition of in by using the following identity on :

where for all the function is radial and in with , and for all . Note that instead of the condition , we may assume that for some . Then, the operator maps atoms to appropriate “molecules”. See Lemma 7 of Reference 7.

We now give a short discussion of the Hardy space . For more details, see Reference 4. First, we need a variant of formula (Equation 4.4), which is inspired from the -calculus for . We start from the identity:

which is valid for all in a sector with . As a consequence, one has

where the integral converges strongly in . See Reference 26. For any , we let . We then have the following identity for all :

Recall that in Reference 4, a function is called an -molecule if

where is a -atom supported in the tent of some ball , and satisfies the condition By using the identity (Equation 4.6) in place of (Equation 4.4), an -molecule decomposition of in the space is obtained in Theorem 7 of Reference 4 as follows.

Proposition 4.2.

Let . There exist -molecules and numbers for such that

The sequence satisfies . Conversely, the decomposition Equation 4.8 satisfies

Proof.

The proof of Proposition 4.2 follows from an argument using certain estimates on area integrals and tent spaces. For the details, we refer the reader to Theorem 7 of Reference 4.

4.2. Properties for and BMO spaces

Let be the set of all with compact support in Consider the operator of (Equation 4.6) initially defined on by

Note that for any compact set in

This and the estimate (Equation 3.2) imply that the integral (Equation 4.9) is well-defined, and for

Lemma 4.3.

The operator , initially defined on , extends to a bounded linear operator from

(a) to , if

(b) to

(c) to .

Proof.

The property (b) is contained in the second part of Proposition 4.2. The property (c) will be shown in Section 5.2 as it is a direct result of Theorem 3.1 and the duality of and spaces.

We now verify (a). By using (5.1) of Reference 8, we have

This, together with (Equation 4.9) and the estimate (Equation 3.8), yield

for any Hence, we obtain

As a consequence of Lemma 4.3, we have the following corollary.

Corollary 4.4.

The space is dense in .

Proof.

For any , by the definition of we have Define and let

for all This family of functions satisfies

(i)

(ii) as

By (a) and (b) of Lemma 4.3, the estimate (i) is straightforward since for each , . Moreover, by (b) of Lemma 4.3,

as This proves property (ii) and completes the proof of Corollary 4.4.

Remark.

From Corollary 4.4, it follows from a standard argument that for any , has an -molecular decomposition (Equation 4.8). See, for example, Chapter III of Reference 31.

We next prove the following -estimate for functions in the space , which will be useful in proving our Theorems 3.1 and 3.2 in Section 5.

Lemma 4.5.

For any -function supported on a ball with radius , there exists a positive constant such that

Proof.

Assume that is a ball of radius and centered at . One writes

Note that for any . Using Hölder’s inequality and the fact that the area integral function is bounded on , one obtains

We now estimate the term . First, we will show that there exists a constant such that for any ,

Let us verify (Equation 4.11). Let

and Since we obtain

It follows from the estimate (Equation 3.1) that the kernel of the operator satisfies

where is the positive constant in (Equation 2.3). Therefore,

We only consider the term since the estimate of the term is even simpler. For and , we set , where . For any and , we have

which implies ; hence . Obviously, for any and , we also have . Note that

It follows from elementary integration that

The estimate (Equation 4.11) then follows readily. Therefore,

Combining the estimates of the terms I and II, we obtain that The proof of Lemma 4.5 is complete.

We now follow Theorem 2.14 of Reference 16 to prove the implication (i) (ii) of Theorem 3.2. For the implication (ii) (i) of Theorem 3.2, we will present its proof in Section 5.3.

Lemma 4.6.

If BMO, then is a Carleson measure with .

Proof.

We will prove that there exists a positive constant such that for any ball on ,

Note that

Hence, (Equation 4.12) follows from the following estimates (Equation 4.13) and (Equation 4.14):

and

We will prove these two estimates by adapting the argument in pp. 85-86 of Reference 21. To prove (Equation 4.13), let us consider the square function given by

From (Equation 3.2), the function is bounded on . Let and . Using the properties (Equation 3.13) and (Equation 3.11), we obtain

On the other hand, for any and , one has . By (Equation 3.6) and the property (Equation 3.12),

Therefore,

This, together with (Equation 4.15), give the estimate (Equation 4.13).

Let us prove (Equation 4.14). Noting that for , it follows from the property (Equation 3.11) that for any

By (Equation 3.6), the kernel of the operator satisfies

Using the commutative property of the semigroup and the estimate (Equation 3.6), we then obtain

Hence,

which gives the estimate (Equation 4.14). Hence, the proof of the implication of Theorem 3.2 is complete.

This lemma, together with the estimate (Equation 3.12), give the following result. We leave the details of the proof to the reader.

Corollary 4.7.

Assume that For any BMO,

is a Carleson measure on with .

5. Proofs of Theorems 3.1 and 3.2

5.1. An identity related to Carleson measures

Suppose that is a function in such that is a Carleson measure and is an -molecule of . Let

We first establish the following identity, which will play an important role in the proof of Theorems 3.1 and 3.2.

Proposition 5.1.

For any functions defined as in Equation 5.1, we have the following identity with constant :

As a consequence, for any and , the above identity Equation 5.2 holds.

Proof.

For any -molecule of , we first observe that , where the mapping is given in (Equation 4.1). Since is a Carleson measure, then by (a) of Proposition 4.1 and the dominated convergence theorem, the following integral converges absolutely and satisfies

Next, by Fubini’s theorem, together with the commutative property of the semigroup , we have

Without loss of generality, we assume that where is a -atom supported in , and the ball is centered at and of radius . We have

where and .

We first consider the term I. From (a) of Lemma 4.3, the function . Since has a bounded -calculus on , we obtain

in , where is the constant such that . See Reference 26. Since , (Equation 3.3) ensures that . Hence

In order to estimate the term II, we need to show that for all , there exists a constant such that

Let us verify (Equation 5.4). Let

By (Equation 3.1), we have

Note that for we have Using the inequality

together with Hölder’s inequality and elementary integration, it can be verified that there exists a positive constant independent of such that for all ,

Estimate (Equation 5.4) then follows readily.

We now estimate the term II. For , it follows from (Equation 3.3) that the function . The estimate (Equation 5.4) implies that there exists a constant such that

This allows us to pass the limit inside the integral of II. Hence

Combining the estimates of I and II, we obtain the identity (Equation 5.2). The proof of Proposition 5.1 is complete.

5.2. Proof of Theorem 3.1

First, we prove (i) of Theorem 3.1. Note that for any and BMO, the assumptions of Proposition 5.1 are satisfied since we have

and by Lemma 4.6,

with .

Let be the constant in Proposition 5.1. Applying the identity (Equation 5.2), together with (a) of Proposition 4.1, we obtain

and thus Since is dense in , (i) of Theorem 3.1 follows from a standard density argument.

We now prove (ii) of Theorem 3.1. We define

By the definition of , we have that , where is the standard tent space. See Section 4.1. Note that by (b) of Lemma 4.3,

for every .

On the other hand, from (Equation 4.6) we have that for any ,

Therefore, for each continuous linear functional on , we obtain

for all . Furthermore, is a continuous linear functional on which satisfies

Applying the Hahn-Banach theorem, we can extend to a continuous linear functional on . Note that by (b) of Proposition 4.1, the dual of is equivalent to . By restricting attention to , we can conclude that if is a continuous linear functional on , then it follows from (Equation 5.5) that there exists a such that

where .

We now prove that BMO. For any ball , it follows from (Equation 5.6) and Lemma 4.5 that

This proves that BMOwith . Hence, the proof of (ii) of Theorem 3.1 is complete.

Proof of (c) of Lemma 4.3.

We now use Theorem 3.1 to prove property (c) of Lemma 4.3. As in Definition (Equation 4.9), we consider the operator associated with defined on by

In order to prove (c) of Lemma 4.3, it suffices to prove that is bounded from to . Note that for any and ,

Since Theorem 3.1 shows that the predual space of is the Hardy space , property (c) of Lemma 4.3 follows readily.

Corollary 5.2.

The spaces and BMO are Banach spaces.

Proof.

Note that is a normed linear space. It follows from Theorem 3.1 and a standard argument of functional analysis that is a Banach space. See, for example, page 111 of Reference 36. The same argument holds for the space . Hence, the proof of Corollary 5.2 is complete.

5.3. Proof of Theorem 3.2

In Lemma 4.6, we proved the implication (i) (ii) of Theorem 3.2. We now prove the implication (ii) (i). Suppose that such that is a Carleson measure. For any , using the identity (Equation 5.2) with in place of , we obtain

which gives and thus with . The proof of Theorem 3.2 is complete.

6. The and spaces associated with some differential operators

In this section, we conduct further study on the Hardy and BMO spaces associated with some differential operators such as the divergence form operators and the Schrödinger operators on (Section 6.1). We will also discuss the inclusion between the classical BMO space and the BMO spaces associated with operators (Section 6.2).

Note first that smooth functions with compact support do not necessarily belong to in general. The reason is that is a Banach space, with the norm vanishing on the kernel space of (Equation 3.10) defined by

hence if , then satisfies the cancellation condition

for all .

6.1. Kernel spaces of some differential operators

We first note that the classical BMO space is a Banach space modulo the constant functions. In this section, we will study the kernel spaces of spaces associated with second-order uniformly elliptic operators of divergence form and with Schrödinger operators with certain potentials.

6.1.1. Second-order elliptic operators of divergence form

Let be an matrix of bounded complex coefficients defined on which satisfies the ellipticity (or “accretivity”) condition

for and for some such that . We define the second-order divergence form operator

on , which we interpret in the weak sense via a sesquilinear form. See Reference 3.

Since is maximal accretive, it has a bounded -calculus on (Reference 1, Reference 3); i.e., satisfies assumption (b) of Section 3.1. Note that when has real entries, or when the dimension or in the case of complex entries, the operator generates an analytic semigroup on with a kernel satisfying a Gaussian upper bound; that is,

for and all . In this case, satisfies assumption (a) of Section 3.1. For dimensions 5 and higher, it is known that the Gausssian bounds (Equation 6.4) may fail. See Reference 2 and Chapter 1 of Reference 3.

Recall that is said to be -harmonic if it is a weak solution of the equation , i.e., for any ,

For any real number , one denotes

which is the space of all polynomial growth -harmonic functions of degree at most . See Reference 23 and Reference 24.

For second-order uniformly elliptic operators with real measurable coefficients, De Giorgi-Nash-Moser theory asserts that any weak solution must be for some . A global version of this theory implies that there exists such that any -harmonic function satisfying the growth condition

as must be a constant function. This means that for all , the dimension of is . In Reference 23 and Reference 24, P. Li and J.P. Wang proved that for each real number , the space is of finite dimension. More specifically, there exists a constant depending only on , and in (Equation 6.2) such that the dimension of satisfies

For any fixed constant in (Equation 2.3), we let

Proposition 6.1.

Let be the divergence form operator as in Equation 6.3. Assume that the operator satisfies assumption (a) in Section for and some as in Equation 2.3. Then

(i) The results of Theorems and hold for the operator .

(ii) The following inclusion between the kernel space and the space holds:

(ii) ;

(ii) Conversely, we have that for any .

(iii) If the semigroup has a kernel satisfying the Gaussian upper bound Equation 6.4, then

(iv) In the case that has real coefficients, then for each , the kernel space has finite dimension.

In order to prove Proposition 6.1, we need the following Lemmas 6.2 and 6.3. For any two closed sets and of , we denote the distance between and by We first have

Lemma 6.2.

Let be the divergence form operator as in Equation 6.3 with ellipticity constants and as in Equation 6.2. For any two closed sets and of , the following off-diagonal estimate of Gaffney type holds:

where depends only on , and depends on

Proof.

For the proof, we refer to Lemma 2.1 of Reference 20. See also Lemma 2.1 of Reference 1.

Lemma 6.3.

Let be the divergence form operator as in Equation 6.3. Assume that the operator satisfies the assumption (a) in Section for and some as in Equation 2.3. Then for any ,

(i) for any , there exists a constant which depends on such that

for almost all .

(ii) For almost all ,

(iii) For any , .

Proof.

The proof of (i) is a simple consequence of direct integration using the decay of heat kernels (Equation 2.2), (Equation 2.3) and the triangle inequality. We omit the details.

We now prove (ii). We fix a ball of radius and set ourselves the task of showing that for almost every . Let be the ball with the same centre as and with radius . Let for and for ; and let . Then Note that under the conditions (Equation 2.2) and (Equation 2.3), satisfies the conservation property of the semigroup for all . See page 55 of Reference 3. By a standard argument using the heat kernel bounds, for example, Section 2, Chapter 3 of Reference 30 for the case of convolution operators, we have that for almost every . However for any and , we have , and then by the conditions (Equation 2.2) and (Equation 2.3),

as Hence, for almost all . Thus (ii) is proved.

For the proof of (iii), it suffices to prove that for any ball with its center at the origin and of radius , there exists a constant which depends on and such that

Let us prove (Equation 6.7). For any integer , we denote by the ball with center at the origin and of radius , except that the notation means the empty set . We define for any , and write Since , we have that for , Using Lemma 6.2, one has

This shows (Equation 6.7), and hence . The proof of Lemma 6.3 is complete.

Remark 6.4.

Property (iii) of Lemma 6.3 holds for any differential operator which satisfies the Gaffney estimate (Equation 6.6) and assumption (a) in Section 3.1 for . This will be used in the proof of Proposition 6.5 below.

Proof of Proposition 6.1.

For the proof of (i), it is straightforward that satisfies the assumptions (a) and (b) of Section 3.1; hence Theorems 3.1 and 3.2 hold.

We now prove (ii). If , then for any and . It follows from (i) of Lemma 6.3 that and Because of the growth of , we use a standard approximation argument through a sequence as follows. For any , we denote by a standard cut-off function which is inside the ball , zero outside , and let . Since , we have that for any ,

which proves that .

Next, we prove (ii). Since for , we have and . This gives that . Hence (ii) of Lemma 6.3 holds. Since , we have that for any ,

This gives a.e; hence a.e. This proves that , and (ii) is proved.

For (iii), that is a consequence of (ii). For (iv), it follows from (ii), Reference 23 and Reference 24 that for each , the kernel space has a finite dimension. The proof of Proposition 6.1 is complete.

6.1.2. Schrödinger operators

Let be a nonnegative function on . The Schrödinger operator with potential is defined by

The operator is a self-adjoint positive definite operator; hence it has a bounded -calculus on (Reference 26). From the Feynman-Kac formula, it is well known that the kernel of the semigroup satisfies the estimate

However, unless satisfies additional conditions, the heat kernel can be a discontinuous function of the space variables and the Hölder continuity estimates may fail to hold. See, for example, Reference 10.

As in Reference 29, a function is said to be a weak solution of in if for any ,

For any , one writes

and

Recall that a nonnegative locally integrable function on is said to belong to the reverse Hölder class with if there exists a constant such that the reverse Hölder inequality

holds for every ball in .

Note that if is a nonnegative polynomial, then for all , . If for some , then the fundamental solution decays faster than any power of . See page 517 of Reference 28. It follows from Corollary 2.8 of Reference 28 that in has a unique weak solution in . Hence for any ,

See also Proposition 2.3 of Reference 29.

Proposition 6.5.

Let be the Schrödinger operator as in Equation 6.8. Then,

(i) the results of Theorems and hold for the operator ;

(ii) for any , we have that .

As a consequence, if for some , then .

Proof.

For the proof of (i), it is straightforward that satisfies the assumptions (a) and (b) of Section 3.1; hence Theorems 3.1 and 3.2 hold.

We now prove (ii). Assume that . Let us prove that . First, for any two closed sets and of , we observe that satisfies the following off-diagonal estimate of Gaffney type:

The proof of this estimate for the Schrödinger operator is similar to that of the case when is a divergence form operator. See, for examples, Lemma 2.1 of Reference 20 and Lemma 2.1 of Reference 1. Then it follows from the Gaffney estimate and Remark 6.4 that .

Note that if , then for any . For any , we denote by a standard cut-off function which is inside the ball , zero outside , and let . Since , we have that for any ,

which proves that . The proof of Proposition 6.5 is complete.

6.2. Inclusion between the classical space and spaces associated with operators

An important application of the BMO space is the following interpolation result of operators.

Proposition 6.6.

Assume that is a sublinear operator which is bounded on for some and for any Then, is bounded on for all

Proof.

For the proof, we refer to Theorem 5.2 of Reference 16.

Because of this interpolation result, we would like to compare the classical BMO space with the spaces BMO associated with operators.

6.2.1. A necessary and sufficient condition for

The following proposition is essentially Proposition 3.1 of Reference 25.

Proposition 6.7.

Suppose is an operator which generates a semigroup with the heat kernel bounds Equation 2.2 and Equation 2.3. A necessary and sufficient condition for the classical space with

is that for every , almost everywhere, that is, for almost all .

Proof.

Assume that for every , almost everywhere. By Proposition 3.1 of Reference 25, we have that and the estimate Equation 6.12 holds. See also Proposition 2.5 of Reference 16. We now show that the condition a.e. is necessary for . Indeed, let us consider . Then, (Equation 6.12) implies that , and thus for every , almost everywhere.

We now give an example of

Proposition 6.8.

There exists an operator which satisfies the assumptions (a) and (b) of Section such that

Proof.

We recall that denotes the upper-half space of i.e.,

Similarly, denotes the lower-half space in

By (resp. ) we denote the Neumann Laplacian on (resp. on ). See page 57 of Reference 33. The Neumann Laplacians are self-adjoint and positive definite operators. Using the spectral theory one can define the semigroup (resp. ) generated by the operator (resp. ). For any defined on , we set

where and are restrictions of the function to and , respectively. Let be the uniquely determined unbounded operator acting on such that

for all such that and .

Then, generates the conservative semigroup for every , which satisfies the assumptions (a) and (b) of Section 3.1. Moreover, it can be proved that this operator generates the spaces and such that and . For the details, we refer the reader to Reference 12.

6.2.2. A sufficient condition for BMO spaces to coincide with the classical BMO space

Assume that is a linear operator of type on with ; hence generates an analytic semigroup . We assume that for each , the kernel of is Hölder continuous in both variables , and there exist positive constants , and such that for all , and ,

whenever ; and

We have the following lemma.

Lemma 6.9.

Assume that satisfies Equation 6.13 and Equation 6.14. Then the kernel of the operator also satisfies Equation 6.13 and Equation 6.14 in which the constants and are replaced by some constants and , respectively. Moreover, for any there exist constants , and such that for all with ,

and

whenever .

Proof.

The proof of Lemma 6.9 is standard. We give a brief argument of this proof for completeness and the convenience of the reader.

Assume that the statement on is proved. Then, using the Cauchy formula applied to the holomorphic function , we obtain the desired estimates for the kernel of . See, for example, Lemma 2.5 of Reference 5.

It remains to prove the statement on . An argument of Davies, as adapted in Proposition 3.3 of Reference 15, enables one to obtain (Equation 6.16). See also Lemma 2.4 of Reference 5.

We now prove (Equation 6.17). We only consider the part since the proof of is similar. It can be verified that it is equivalent to the following: there exist constants and such that for all and ,

Let us prove (Equation 6.18). By Lemma 17 of Chapter 1 of Reference 3, this inequality is equivalent to the boundedness of from to the homogeneous space with the right-hand side of (Equation 6.18) being its operator norm. For , we denote by the operator norm of from into . We deduce from (Equation 6.13) and Lemma 17 of Chapter 1 of Reference 3 that , and Hence, by interpolation,

On the other hand, it follows from (Equation 6.14) that

and

Hence, by interpolation,

One writes where , and . Then using the semigroup property , we have

which gives (Equation 6.18). This gives the desired estimate of in (Equation 6.17). Hence, Lemma 6.9 is proved.

Using Lemma 6.9, we have the following equivalence between the classical BMO space and BMO spaces associated with differential operators.

Theorem 6.10.

Assume that satisfies the assumptions Equation 6.13, Equation 6.14 and Equation 6.15. Then, the BMO space (modulo constant functions) and the BMO space (modulo ) coincide, and their norms are equivalent.

Proof.

We remark that for satisfying (Equation 6.13), (Equation 6.14) and (Equation 6.15), our proof below shows that has a bounded holomorphic functional calculus on because the area integral functions and are bounded on where is the adjoint operator of . Hence, Theorem 3.1 holds for the operators and .

This follows from Proposition 6.7 and the assumption (Equation 6.15) that We now prove . From Theorem 3.1 and a duality argument, this reduces to proving that with . Using the atomic decomposition of , it suffices to prove that for any atom , we have where is a positive constant independent of . See Reference 31. Denote by the kernel of the operator By (Equation 6.15) we have . It follows from Lemma 6.9 that there exist constants , and such that

and whenever ,

From Theorem 3 of Reference 27, the area integral function is bounded on ; hence . It follows from a standard harmonic analysis argument that we have See, for example, Proposition 1.2, Chapter 14 of Reference 34.

This proves that ; hence . The proof of Theorem 6.10 is complete.

Remarks.

(i) As noted in Section 6.1.1, the assumptions (Equation 6.13), (Equation 6.14) and (Equation 6.15) are satisfied for the divergence form operator in (Equation 6.3) when has real coefficients or when the dimension or in the case of complex coefficients. See Chapter 1 of Reference 3 and Reference 2.

(ii) The Laplacian on satisfies the assumptions of Theorem 6.10; hence the spaces and coincide with the classical BMO space and Theorem 6.10 generalizes the results of Theorems 2.14 and 2.15 of Reference 16.

6.2.3. An example of

In Reference 13, a space of BMO type associated with a Schrödinger operator was introduced as follows. Let on , where

is a nonnegative nonzero polynomial on , Such a function in (Equation 6.19) belongs to the reverse Hölder class for all , . See the condition (Equation 6.10) in Section 6.1.2.

Denote by The space associated with was defined by

It is obvious that . It was observed in Reference 13 that is a proper subspace of the classical BMO space (for example, ). In Reference 13, they also proved that

where the Hardy space is defined by means of a maximal function associated with the semigroup , i.e.,

See Reference 17. Note that by Theorem 3 of Reference 37,

Theorem 3.1, together with (Equation 6.20) and (Equation 6.21), give the following proposition.

Proposition 6.11.

Assume that , where is a nonnegative nonzero polynomial Equation 6.19. Then, the spaces and coincide and their norms are equivalent.

As a consequence, we have . That is, is a proper subspace of the classical BMO space.

Acknowledgments

The authors would like to thank the referee for helpful comments and suggestions. The authors thank A. McIntosh for helpful suggestions and P. Auscher as some of these ideas originate with him in Reference 4. The second-named author thanks C. Kenig and Z. Shen for useful discussions and references and D.G. Deng for all support, encouragement and guidance given over the years.

Mathematical Fragments

Equation (1.1)
Equation (1.2)
Equation (1.3)
Equation (2.2)
Equation (2.3)
Equation (3.1)
Equation (3.2)
Equation (3.3)
Equation (3.6)
Equation (3.8)
Equation (3.9)
Equation (3.10)
Equation (3.11)
Equation (3.12)
Equation (3.13)
Theorem 3.1.

Assume that the operator satisfies the assumptions (a) and (b) in Section . Denote by the adjoint operator of . Then, the dual space of the space is the BMO space, in the following sense.

(i) Suppose BMO. Then the linear functional given by

initially defined on the dense subspace , has a unique extension to .

(ii) Conversely, every continuous linear functional on the space can be realized as above; i.e., there exists BMO such that 3.14 holds and

Equation (3.15)
Equation (4.1)
Equation (4.4)
Equation (4.6)
Proposition 4.2.

Let . There exist -molecules and numbers for such that

The sequence satisfies . Conversely, the decomposition 4.8 satisfies

Equation (4.9)
Lemma 4.3.

The operator , initially defined on , extends to a bounded linear operator from

(a) to , if

(b) to

(c) to .

Equation (4.11)
Equation (4.12)
Equation (4.13)
Equation (4.14)
Equation (4.15)
Equation (5.1)
Proposition 5.1.

For any functions defined as in Equation 5.1, we have the following identity with constant :

As a consequence, for any and , the above identity 5.2 holds.

Equation (5.4)
Equation (5.5)
Equation (5.6)
Equation (6.2)
Equation (6.3)
Equation (6.4)
Proposition 6.1.

Let be the divergence form operator as in Equation 6.3. Assume that the operator satisfies assumption (a) in Section for and some as in Equation 2.3. Then

(i) The results of Theorems and hold for the operator .

(ii) The following inclusion between the kernel space and the space holds:

(ii) ;

(ii) Conversely, we have that for any .

(iii) If the semigroup has a kernel satisfying the Gaussian upper bound Equation 6.4, then

(iv) In the case that has real coefficients, then for each , the kernel space has finite dimension.

Lemma 6.2.

Let be the divergence form operator as in Equation 6.3 with ellipticity constants and as in Equation 6.2. For any two closed sets and of , the following off-diagonal estimate of Gaffney type holds:

where depends only on , and depends on

Equation (6.7)
Equation (6.8)
Equation (6.10)
Proposition 6.7.

Suppose is an operator which generates a semigroup with the heat kernel bounds Equation 2.2 and Equation 2.3. A necessary and sufficient condition for the classical space with

is that for every , almost everywhere, that is, for almost all .

Equations (6.13), (6.14)
Equation (6.15)
Lemma 6.9.

Assume that satisfies Equation 6.13 and Equation 6.14. Then the kernel of the operator also satisfies Equation 6.13 and Equation 6.14 in which the constants and are replaced by some constants and , respectively. Moreover, for any there exist constants , and such that for all with ,

and

whenever .

Equation (6.18)
Equation (6.19)
Equation (6.20)
Equation (6.21)

References

Reference [1]
P. Auscher, S. Hofmann, M. Lacey, A. McIntosh and Ph. Tchamitchian, The solution of the Kato square root problem for second order elliptic operators on , Annals of Math., 156 (2002), 633-654. MR 1933726 (2004c:47096c)
Reference [2]
P. Auscher and P. Tchamitchian, Calcul fonctionnel précisé pour des opérateurs elliptiques complexes en dimension un (et applications à certaines équations elliptiques complexes en dimension deux), Ann. Institut Fourier (Grenoble), 45 (1995), 721-778. MR 1340951 (96f:35036)
Reference [3]
P. Auscher and P. Tchamitchian, Square root problem for divergence operators and related topics, Asterisque, 249, Soc. Math. France, 1998. MR 1651262 (2000c:47092)
Reference [4]
P. Auscher, X.T. Duong and A. McIntosh, Boundedness of Banach space valued singular integral operators and Hardy spaces, preprint, 2004.
Reference [5]
T. Coulhon and X.T. Duong, Maximal regularity and kernel bounds: Observations on a theorem by Hieber and Prss, Adv. Differential Equations, 5 (2000), 343-368. MR 1734546 (2001d:34087)
Reference [6]
M. Cowling, I. Doust, A. McIntosh and A. Yagi, Banach space operators with a bounded functional calculus, J. Austral. Math. Soc. Ser. A, 60 (1996), 51-89. MR 1364554 (97d:47023)
Reference [7]
R.R. Coifman, Y. Meyer and E.M. Stein, Un nouvel éspace adapté à l’étude des opérateurs définis par des intégrales singulières, in “Proc. Conf. Harmonic Analysis, Cortona”, Lecture Notes in Math., Vol. 992, pp. 1-15, Springer-Verlag, Berlin/New York, 1983. MR 0729344 (85j:42032)
Reference [8]
R.R. Coifman, Y. Meyer and E.M. Stein, Some new functions and their applications to harmonic analysis, J. Funct. Analysis, 62(1985), 304-335. MR 0791851 (86i:46029)
[9]
R. Coifman and G. Weiss, Extensions of Hardy spaces and their use in analysis, Bull. Amer. Math. Soc., 83 (1977), 569–645. MR 0447954 (56:6264)
Reference [10]
E.B. Davies, Heat kernels and spectral theory, Cambridge Univ. Press, 1989. MR 0990239 (90e:35123)
Reference [11]
D.G. Deng, On a generalized Carleson inequality, Studia Math., 78 (1984), 245-251. MR 0782661 (86h:42033)
Reference [12]
D.G. Deng, X.T. Duong, A. Sikora and L.X. Yan, Comparison between the classical BMO and the BMO spaces associated with operators and applications, preprint, (2005).
Reference [13]
J. Dziubański, G. Garrigós, T. Martínez, J. Torrea and J. Zienkiewicz, BMO spaces related to Schrödinger operators with potentials satisfying a reverse Hölder inequality, Math. Z., 249 (2005), 329-356. MR 2115447.
Reference [14]
X.T. Duong and A. McIntosh, Singular integral operators with non-smooth kernels on irregular domains, Rev. Mat. Iberoamericana, 15 (1999), 233-265. MR 1715407 (2001e:42017a)
Reference [15]
X.T. Duong and D.W. Robinson, Semigroup kernels, Poisson bounds, and holomorphic functional calculus, J. Funct. Anal., 142 (1996), 89–128. MR 1419418 (97j:47056)
Reference [16]
X.T. Duong and L.X. Yan, New function spaces of BMO type, the John-Nirenberg inequality, interpolation and applications, to appear, Comm. Pure Appl. Math., (2005).
Reference [17]
J. Dziubański and J. Zienkiewicz, Hardy spaces associated with some Schrödinger operator, Studia Math., 126 (1997), 149-160. MR 1472695 (98k:42029)
Reference [18]
C. Fefferman, Characterizations of bounded mean oscillation, Bull. Amer. Math. Soc., 77 (1971), 587-588. MR 0280994 (43:6713)
Reference [19]
C. Fefferman and E.M. Stein, spaces of several variables, Acta Math., 129 (1972), 137–195. MR 0447953 (56:6263)
Reference [20]
S. Hofmann, J.M. Martell, bounds for Riesz transforms and square roots associated to second order elliptic operators, Pub. Mat., 47 (2003), 497-515. MR 2006497 (2004i:35067)
Reference [21]
J.L. Journé, Calderón-Zygmund operators, pseudo-differential operators and the Cauchy integral of Calderón. Lecture Notes in Mathematics, 994. Springer, Berlin-New York, 1983. MR 0706075 (85i:42021)
Reference [22]
F. John and L. Nirenberg, On functions of bounded mean oscillation, Comm. Pure Appl. Math., 14(1961), 415–426. MR 0131498 (24:A1348)
Reference [23]
P. Li, Harmonic sections of polynomial growth, Math. Research Letters, 4 (1997), 35-44. MR 1432808 (98i:53054)
Reference [24]
P. Li and J.P. Wang, Counting dimensions of -harmonic functions, Ann. of Math., 152 (2000), 645-658. MR 1804533 (2002c:58032)
Reference [25]
J.M. Martell, Sharp maximal functions associated with approximations of the identity in spaces of homogeneous type and applications, Studia Math., 161(2004), 113-145. MR 2033231 (2005b:42016)
Reference [26]
A. McIntosh, Operators which have an -calculus, Miniconference on operator theory and partial differential equations, (1986) Proc. Centre Math. Analysis, ANU, Canberra 14 (1986), 210-231. MR 0912940 (88k:47019)
Reference [27]
S. Semmes, Square function estimates and the theorem, Proc. Amer. Math. Soc., 110 (1990), 721-726. MR 1028049 (91h:42018)
Reference [28]
Z. Shen, estimates for Schrödinger operators with certain potentials, Ann. Institut Fourier (Grenoble), 45 (1995), 513-546. MR 1343560 (96h:35037)
Reference [29]
Z. Shen, On fundamental solutions of generalized Schrödinger operators, J. Funct. Anal., 167 (1999), 521-564. MR 1716207 (2000j:35055)
Reference [30]
E.M. Stein, Singular integral and differentiability properties of functions, Princeton Univ. Press, 30, (1970). MR 0290095 (44:7280)
Reference [31]
E.M. Stein, Harmonic analysis: Real variable methods, orthogonality and oscillatory integrals, Princeton Univ. Press, Princeton, NJ, (1993). MR 1232192 (95c:42002)
Reference [32]
E.M. Stein and G. Weiss, On the theory of harmonic functions of several variables I, The theory of spaces, Acta Math., 103 (1960), 25-62. MR 0121579 (22:12315)
Reference [33]
W.A. Strauss, Partial differential equations: An introduction. John Wiley & Sons, Inc., New York, 1992. MR 1159712 (92m:35001)
Reference [34]
A. Torchinsky, Real-variable methods in harmonic analysis, Pure and Applied Math., Vol 123, Academic Press, (1986). MR 0869816 (88e:42001)
Reference [35]
L.X. Yan, Littlewood-Paley functions associated to second order elliptic operators, Math. Z., 246 (2004), 655-666. MR 2045834 (2005a:42015)
Reference [36]
K. Yosida, Functional Analysis (fifth edition), Springer-Verlag, Berlin, 1978. MR 0500055 (58:17765)
Reference [37]
Y.P. Zhu, Area functions on Hardy spaces associated to Schrödinger operators, Acta Math. Sci. Ser. B Engl. Ed., 23 (2003), 521–530. MR 2032556 (2004k:42029)

Article Information

MSC 2000
Primary: 42B30 (-spaces), 42B35 (Function spaces arising in harmonic analysis), 47F05 (Partial differential operators)
Keywords
  • Hardy space
  • BMO
  • semigroup
  • holomorphic functional calculi
  • tent space
  • Carleson measure
  • second-order elliptic operator
  • Schrödinger operator.
Author Information
Xuan Thinh Duong
Department of Mathematics, Macquarie University, NSW 2109, Australia
duong@ics.mq.edu.au
MathSciNet
Lixin Yan
Department of Mathematics, Macquarie University, NSW 2109, Australia and Department of Mathematics, Zhongshan University, Guangzhou, 510275, People’s Republic of China
lixin@ics.mq.edu.au, mcsylx@zsu.edu.cn
MathSciNet
Additional Notes

Both authors are supported by a grant from the Australia Research Council. The second author is also supported by NNSF of China (Grant No. 10371134) and the Foundation of Advanced Research Center, Zhongshan University.

Journal Information
Journal of the American Mathematical Society, Volume 18, Issue 4, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2005 American Mathematical Society; reverts to public domain 28 years from publication
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0894-0347-05-00496-0
  • MathSciNet Review: 2163867
  • Show rawAMSref \bib{2163867}{article}{ author={Duong, Xuan}, author={Yan, Lixin}, title={Duality of Hardy and BMO spaces associated with operators with heat kernel bounds}, journal={J. Amer. Math. Soc.}, volume={18}, number={4}, date={2005-10}, pages={943-973}, issn={0894-0347}, review={2163867}, doi={10.1090/S0894-0347-05-00496-0}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.