1. Introduction
Historically, the analysis of tempered distributions as boundary values of analytic functions has found applications in mathematical physics, in the study of quantum field theory. An important reference in this study is Streater and Wightman [
1]. In field theory, the “vacuum expectation values” are tempered distributions that are boundary values in the tempered distribution topology of analytic functions, with the analytic functions being Fourier–Laplace transforms. In addition, a field theory can be recovered from its “vacuum expectation values” [
1] (Chapter 3). A similar field theory analysis is contained in the work by Simon [
2].
Of particular interest with respect to the contents of this paper is the work of Raina [
3] in mathematical physics. In [
3], Raina considered analytic functions in the upper half plane that satisfied a pointwise growth condition associated with the analytic functions that have tempered distributions as boundary value when 
. The important mathematical result in [
3] showed that if the tempered distributional boundary value was an element of 
, then the analytic function was in the Hardy space 
 of analytic functions in the upper half plane. A converse result was proved. Raina described the importance of the results of this type concerning tempered distributional boundary values and the Hardy spaces 
, which, in mathematical physics, are associated with “form factor bounds”, including the use of Hardy spaces in general in related topics in mathematical physics. Several associated references are given in [
3]. Importantly, the tempered distributions are used in the analysis of the mathematical physics in [
1,
2,
3].
The results in [
3] have led the author to consider the results of the type in [
3] for higher dimensions and for the analytic and 
 functions being both scalar-valued and vector-valued. We have also desired to obtain representations of the analytic functions involved in terms of Fourier–Laplace transforms, Cauchy integrals, and Poisson integrals. Further, we have desired to obtain new results concerning both the scalar-valued and vector-valued Hardy functions in higher dimensions, including the growth properties of these functions.
Given our desires expressed in the previous paragraph, we first considered the scalar-valued case in [
4] where we obtained the pointwise growth of scalar-valued 
 functions on tubes in 
. In [
4], we considered scalar-valued analytic functions on tubes in 
 that had a specified pointwise growth, leading to the existence of tempered distributions as boundary values, and showed that if these boundary values were a 
 function, 
, then the scalar-valued analytic function was in 
. Related results for other spaces of distributions were obtained in [
4].
Continuing to the vector-valued case and building upon the results of [
4], in [
5] we considered vector-valued analytic functions in tube domains in 
 that have pointwise growth, leading to the existence of vector-valued tempered distributions as boundary values, and proved that if the boundary value is a vector-valued 
 function then the analytic function must be in the Hardy space 
. We obtained integral representations of the analytic functions and obtained pointwise growth of vector-valued 
 functions in tubes, 
.
In [
6], we considered vector-valued analytic functions in tube domains without a defining pointwise growth so that any boundary value would be considered to be in the Schwartz vector-valued 
 space. We showed that if the analytic functions obtained a distributional boundary value in the vector-valued distribution 
 sense with the boundary value being a vector-valued function in 
, then the analytic function is in the vector-valued Hardy space. We obtained a Poisson integral representation of the analytic functions in this case.
The cases for 
 in the setting of [
5] as described above are missing from our analysis at this point. That is, we desire to consider vector-valued analytic functions in tube domains that have specified pointwise growth that leads to the existence of vector-valued tempered distributions as boundary values. We then desire to prove that if the boundary value is a vector-valued 
 function, then the analytic function is in the vector-valued 
, space. This additional analysis is desirable in order to obtain the appropriate extension of the important Raina results to all of 
 in our generalized setting. Thus, the analysis in this paper concerns the values of 
p in 
.
  2. Definitions and Notation
All notation and definitions needed in this paper are the same as described or referred to in [
5]. We mention and refer to several of the most frequently used definitions and notations here.
 will denote a Banach space,  will denote a Hilbert space,  will denote the norm of the specified Banach or Hilbert space, and  will denote the zero vector of the specified Banach or Hilbert space.  is a cone with a vertex at  in  if  implies  for all . The intersection of a cone C with the unit sphere  is the projection of C and is denoted . A cone  such that  is a compact subcone of C. The dual cone  of C is defined as  { for all }. An open convex cone that does not contain any entire straight line is called a regular cone. Let  be any of the  n-tuples whose entries are 0 or 1. The  n-rants  {} are examples of regular cones that will be useful in this paper.
The 
 functions, 
, with values in 
 and their norms 
, the Schwartz test spaces 
 and 
, and the spaces of tempered vector-valued distributions with values in 
 and 
, are all noted in ([
5], Section 2). The reference for the 
 functions is Dunford and Schwartz [
7]. The references for vector-valued distributions are Schwartz [
8,
9].
The Fourier transform on 
 and on 
 or 
 is given in [
5] (Section 2). The Fourier transform of 
 comes from [
8], and will be denoted 
, with the inverse Fourier transform being denoted 
. Similarly, all Fourier (inverse Fourier) transforms on scalar-valued or vector-valued functions will be denoted 
 or 
. Of particular importance in this paper are the Fourier and inverse Fourier transforms on the vector-valued 
 functions; the results that we need for these functions are discussed and proved in [
10] (Section 1.8). As stated in this reference and referenced in [
10] (Section 1.11), the Plancherel theory is not valid for vector-valued functions except when 
, a Hilbert space. That is, in order for the Fourier transform 
 to be an isomorphism of 
 onto itself with the Parseval identity 
 holding, it is necessary and sufficient that 
, a Hilbert space; this fact comes from Kwapień [
11]. The Plancherel theory is complete in the 
 setting in that the inverse Fourier transform is the inverse mapping of the Fourier transform with 
, with 
I being the identity mapping. As stated in [
10] (Section 1.8), the Plancherel theory stated there is valid for functions of several variables with values in Hilbert space. In the analysis of this paper, we need the Plancherel theory holding on 
, and thus where needed we take 
, a Hilbert space.
Associated with the Fourier transform on vector-valued functions with values in Banach space is the concept of Banach space of type 
 discussed in [
12] (Section 6). We note that every Banach space has Fourier type 1 and leave pursuit of this concept of Fourier type to the interested reader.
Let 
B be an open subset of 
. The Hardy space 
 consists of those analytic functions 
 on the tube 
 with values in a Banach space 
 such that
      
      where 
 and the constant 
 is independent of 
; the usual modification is made for the case 
.
Let 
C be an open convex cone in 
 will denote the set of all infinitely differentiable complex valued functions on 
. We define the function 
, as in [
5] (Section 2).
We define and state known results concerning the Cauchy and Poisson kernel functions corresponding to tubes 
. Let 
C be a regular cone in 
 and 
 be the corresponding dual cone of 
C. The Cauchy kernel corresponding to 
 is
      
      where 
 is the dual cone of 
C as noted. The Poisson kernel corresponding to 
 is
      
Referring to [
13] (Chapters 1 and 4) for details, we know for 
 that 
; and 
, where ∗ is Beurling 
 or Roumieu 
. These ultradifferentiable functions are contained in the Schwartz space 
. We also use the results [
4] (Lemmas 3.1 and 3.2). Because of the combined properties of the Cauchy and Poisson kernels from [
13,
14], we know that the Cauchy and Poisson integrals
	  
 are well defined for 
, and 
, respectively, where 
 is a Banach space.
We use [
5] (Lemma 3.4) several times in this paper. For convenience to the reader, we state this result here to conclude this section. Throughout 
 denotes the closed ball of radius 
 centered at 
.
Theorem 1. Let f be analytic in  with values in a Banach space , where C is a regular cone in , and have the Poisson integral representationfor . We have . For  in the weak-star topology of  as ; for , in  as ; for for all compact subcones  being a constant depending on  and not on , whilewhere  is a constant depending on ; and for for all compact subcones  and all  being a constant depending on  and on , but not on , whilewhere  is a constant depending on .    3. Tempered Distributional Boundary Values
Let C be an open convex cone in  and . We denote the set of analytic functions on  with values in a Banach space  by . As above,  denotes the closed ball about  of radius .
In [
5] (Theorem 4.1), we have stated the following result which we need here.
Theorem 2. Let C be an open convex cone. Let . For every compact subcone  and every , letwhere  is a constant depending on  and on  is a nonnegative integer, k is an integer greater than 1, and neither R nor k depend on  or r. There exists a positive integer m and a unique element  such that  In Theorem 2, and in the remainder of this paper, by , we mean that , for every compact subcone  of C.
In [
5] (Theorem 4.4), we proved for 
C, a regular cone, and the boundary value 
U in Theorem 2 being a function 
, that the analytic function 
 in Theorem 2 is, in fact, in 
. In [
5], we were not able to obtain this result for the cases 
. We now have a proof for the cases 
, and we obtain the result [
5] (Theorem 4.4) for the cases 
 here.
To obtain [
5] (Theorem 4.4) for 
, we follow some of the structure of [
5] by first proving our result for the case that the cone 
C is a n-rant cone 
 or is contained in a n-rant cone and then using this case to obtain the general result for the cone 
C being any regular cone. Because 
, here the details of our proof in the case 
 in Theorem 3 below are different in many instances than those of [
5] (Theorems 4.2 and 4.3). The values of the functions and distributions in the remainder of this section will be in Hilbert space 
 because of the need for the Fourier transform properties on 
, as described in 
Section 2 above.
We give an outline of the proof of Theorem 3 for the benefit of the reader. Given the assumed function  in Theorem 3, we will divide it by a structured analytic function , and put .  is represented as the Fourier transform involving a function , which has support in .  is shown to have boundary value  in  as , and then is shown to equal the Cauchy integral and the Poisson integral of a function involving the boundary value  of . After establishing some important limit analysis, we proceed to prove that  equals the Poisson integral of the boundary value , which will then yield the conclusions of Theorem 3.
Theorem 3. Let C be an open convex cone which is contained in or is any of the  n-rants . Let  be a Hilbert space. Let  and satisfy . Let the unique boundary value U of Theorem 2 be . We have , and  Proof.  As noted above, the proof has a structure similar to that of [
5] (Theorems 4.2 and 4.3), but many details are different. We refer to [
5] (Theorems 4.2 and 4.3) where appropriate. Put 
, where
        
 satisfies (1). (By Theorem 2, there is a unique 
 such that (2) holds for 
 in 
, a fact that we use later in this proof.) By the same analysis as in the proof of [
5] (Theorem 4.2), we obtain [5, (15)] here; that is,
        
        for all compact subcones 
 and all 
 where 
 is a constant. Put
        
Using (4), the same proof as in the proof of [
5] (Theorem 4.2) yields that 
 is a continuous function of 
 for 
 and 
, is independent of 
, and has support in 
, the dual cone of 
C.    □
 For any compact subcone 
, any 
, and any 
 Equation (
4) yields
      
      from which 
 for 
 and for all 
, by ([
5], Lemma 3.1). From Equation (
5), 
, with the transform holding in both the 
 and 
 cases, and in 
:
From the properties of 
, the Fourier transform in (7) is in both the 
 and 
 cases, and (7) becomes
      
Both 
 and 
 are elements of 
, and 
. Thus, 
, in 
 now. Let 
 and 
. We have
      
      as 
. As noted above, by Theorem 2, there is a unique 
 such that 
 in 
 as 
; hence, 
 in 
 and 
.
Since 
, we have 
. By hypothesis, 
 has boundary value 
 in 
 as 
, and 
 in 
 as 
. Thus, 
 in 
. For 
      and 
. For 
, we have 
, since 
 for all 
. We put 
; thus 
. Since supp(
, then supp
 almost everywhere. Recalling the function 
 defined in 
Section 2, we have 
. For 
      with 
. From [
4] (Lemma 2.1), 
 for all 
, for 
 where 
 is the characteristic function of 
, and the integral on the right of 
 is convergent, since 
 and supp
 almost everywhere. From 
, and the fact that supp
, we have for 
We proceed to construct a Poisson integral representation for 
 in addition to the Cauchy integral representation in (11). Let 
w be an arbitrary point of 
. Using [
4] (Lemma 3.2), we have for 
 that 
 is analytic in 
 and satisfies the growth (1) of 
. Further,
      
      in 
 with 
, since both 
 and 
 are bounded for 
. The same proof leading to (11) applied to 
, yields
      
For 
, we choose 
. Then, (12) combined with (11) becomes
      
We now present some limited analyses, which we need to analyze the function, the Poisson integral of 
, that we will show represents 
 and from which the conclusion of the proof of this theorem will follow. Since 
, both 
 and 
 are in 
. We have
      
      with the right side being independent of 
. Further,
      
By the Lebesgue dominated convergence theorem
      
      which proves 
 in 
, as 
.
We now define and analyze the function which we desire to be the Poisson integral representation of 
, as noted in the preceding paragraph; this function is
      
Let 
 be an arbitrary but fixed point of 
. Choose the closed neighborhood 
 of [
5] (Lemma 3.3), and note that [
5] (Lemma 3.3) holds for all 
. Let the constant 
 in (16) below be the constant obtained in [
5] (Lemma 3.3). Using the Hölder inequality if 
 and the boundedness of 
 from the proof of [
5] (Lemma 3.3) ([
4] (Lemma 3.4)) if 
 and using (13) and (15), we have
      
      for 
. Using (14) and (16) for 
, we have
      
      uniformly in 
. Since 
 is analytic in 
, 
, we have that 
 is analytic at 
; hence 
 is analytic in 
 since 
 is an arbitrary point in 
. Applying Theorem 1, we have 
.
Let 
. Using Hölder’s inequality, if 
 and the boundedness of 
 if 
, we have
      
By Theorem 1,  in , as ; hence  in  as .
Now, consider 
, which is analytic in 
. For 
, we have the pointwise bound on 
 for 
 in any compact subcone 
 contained in Theorem 1. (see also ([
5], (6)).) Thus, combining the bounds (1) on 
 and the pointwise bound just noted on 
 for 
 in any compact subcone 
, we have the inequality
      
      on 
 for the cases 
 for any compact subcone 
 and any 
 where 
 is a constant depending on 
 and on 
. If 
, by combining inequalities ([
4], (10) and (11)) in the proof of Theorem 1 given in ([
5], Lemma 3.4), we have
      
      and hence
      
      where 
 depends only on 
 and not on 
C and 
 is the surface area of the unit sphere in 
. Combining this inequality on 
 with inequality (1) on 
, we again have that 
, also satisfies (17) for 
. In addition, we know from the boundary values of 
 and 
 that
      
      in 
 for 
.
Using (18), we now proceed to complete the proof by proving 
. Put 
, which is analytic in 
. 
 satisfies (17) and (18) for each 
. Consider 
, where 
 is defined at the beginning of this proof for 
. As in obtaining (4) for 
, we have for 
 and 
      where 
 is a constant. Now putting as in (5)
      
      and proceeding with the proof from (5) to (8), we have that 
 is continuous, is independent of 
; has support in 
; satisfies a growth as in (6); satisfies 
, with the transform holding in both the 
 and 
 cases; and with 
 for all 
; satisfies 
; and satisfies
      
For 
 and 
      as 
; and
      
      as 
. Combining this fact with (18) yields 
, since 
; hence, 
 in 
. Put
      
Since 
, and 
 is given in (15), we conclude
      
      and 
, since we have previously obtained 
, from Theorem 1. The proof of Theorem 3 is complete.
With Theorem 3 proved for 
, we now obtain this result for 
C being an arbitrary regular cone in 
; this is our desired result, which extends [
5] (Theorem 
) to the cases 
. The proof of the following theorem for the cases 
 is obtained using Theorem 3 by exactly the same proof that [
5] (Theorem 
) was proved using [
5] (Theorems 
 and 
); we ask the interested reader to follow the suggested proof if desired.
Theorem 4. Let C be a regular cone in . Let  be a Hilbert space. Let  and satisfy . Let the unique boundary value U of Theorem 2 be . We have , and  The Poisson integral representation of the function 
 in Theorem 4 follows from the fact that the unique 
 boundary value 
, is obtained independently of how 
, and follows from the structure of the tubes 
, in the referenced proof of [
5] (Theorem 4.4). 
 equals the Poisson integral of 
 in each of these tubes by Theorem 3 and hence in all of 
.
In summary concerning the proofs here of Theorems 3 and 4 for 
 and the proofs of the corresponding results in [
5] for 
, we note the following. In certain places in the analysis, the products or quotients involving the boundary value 
h and other terms must be analyzed carefully in order for the analysis to proceed. In both restrictions on 
p, many times we need the product or quotient to be Fourier transformable in 
 or 
 or both. The properties of such products or quotients can be different depending on whether 
 or 
; hence, the analysis must be suitably adjusted to proceed with the proof. Further, to obtain appropriate boundedness properties in the proofs the method to proceed depends on whether 
 or 
 for the case 
, and depends on whether 
 or 
 for the case 
. These and other technical difficulties must be overcome for the proofs to proceed, and the difficulties depend on the two cases, 
 or 
. Additionally, here we have stated the Poisson integral representation of 
 as a conclusion in Theorem 4, but should have done so in [
5] (Theorem 4.4) as well, where this conclusion is obtained by the same argument used in the paragraph below Theorem 4 above.