arXiv:1711.02610v2 [math.CV] 14 Oct 2019 h lsia ae-inrTermasrsta o a for that asserts Theorem Paley-Wiener classical The Introduction 1 5/07A;NF rn o 1057 SCGatN.119015 No. 0123/2018/A3. Grant SAR: NSFC Macao 11701597; Fund, opment No. Grant NSFC 154/2017/A3; ute h o-agnilbudr ii NB)o ucini t in function a of (NTBL) limit boundary non-tangential the further h pe afpaei n nyi h ore rnfr of transform Fourier the if only and if plane half upper the relation ∗ hswr a upre yteSineadTcnlg Developmen Technology and Science the by supported was work This orsodn author Corresponding ore pcrmo Clifford of Spectrum Fourier f ˆ = sfrhrtenntneta onaylmto oefnto in function some of limit boundary non-tangential H the further is p in nteCiodagbavle ad spaces Hardy algebra-valued func Clifford of the characterization spectrum in Fourier tions the studies article This is rnfr,CiodAgba ojgt amncSystem Harmonic Conjugate Algebra, Clifford Transform, setting. Riesz Stein-Weiss the rather in words cases spaces Key vector-valued Hardy the algebra-valued only Clifford than whole the for systems epnigsnua nerlvrino h etrvle ae f cases cor- vector-valued the the proved latter of The 1 version Weiss. harmonic integral and conjugate singular Stein the of responding work with Cliffor the vector-valued identical in the are systems McIn- results, sense). functions our Alan of of distribution case Hardy context particular the relevant a the As in develop tosh. cases further some results These (for interpreted suitably 2 1 χ 1+ (1 + p ≤ ∞ ≤ ( f, ˆ R < p + n where i +1 | ξ ξ . | ) ) , ∞ , aey for Namely, 1 h ore rnfrainadteaoerlto are relation above the and transformation Fourier the ad pc,MngncFnto,FuirSpectrum, Fourier Function, Monogenic space, Hardy : e ag exogMai Weixiong Dang, Pei . χ ≤ + eas bantegnrlzdcnuaeharmonic conjugate generalized the obtain also We nmmr fAa McIntosh Alan of memory In stecaatrsi idctr ucino h e (0 set the of function (indicator) characteristic the is p ∞ ≤ 1 , f fadol if only and if ≤ ∈ Abstract L p p 1 ( R ∞ ≤ n ,Ciodalgebra-valued, Clifford ), H f ˆ a Qian Tao , ∗ p L = 2 pcson Spaces ( χ 4 h cec n ehooyDevel- Technology and Science The 94; R + )-function f, ˆ ud aa A:0006/2019/A1, SAR: Macao Fund, t f eoe by denoted , where H p ( R eHardy he + n χ f +1 + clrvle,i is it scalar-valued, , ) ( R , ξ 1 = ) f or ≤ + n ˆ f d - aifisthe satisfies , +1 H 2 , ∞ pc in space for ) , that takes the value 1 when its argument is in (0, ∞) and otherwise zero. This amounts to saying that a characteristic property of boundary limit functions of the Hardy space functions is that their Fourier transforms vanish at the negative spectra. This classical Fourier spectrum characterization of the Hardy H2 space functions has been studied and generalized by different authors (see, for instance, [5, 9]). Among recent studies [18] and [19] give a throughout treatment to the analogous results for Lp(R) for p ∈ [1, ∞]. Those papers prove that if f is a function in Lp(R), 1 ≤ p ≤∞, then f is the NTBL of some function in the Hardy Hp space in the upper half plane if and only if the Fourier transform fˆ satisfies the relation (f,ˆ ψ)=0 for ψ being any function in the Schwartz class whose support lays in the closure of left half of the real line. We note that for 2 < p ≤ ∞ this last statement is interpreted as that fˆ, as a distribution, is supported in [0, ∞). The Fourier spectrum characterization results have implications to the characterizations of the Hardy space functions, as well as to Hardy space decompositions of Lp functions, the latter being through the Fourier spectrum decomposition. We note that the Hardy spaces decomposition can also be extended to the Lp spaces for 0 2, there are two distinguished com- plex structures, of which one is several complex variables and the other is Clifford algebra. Both those complex structures in relation to their respective Hardy spaces are treated in [20] and [21]. The several complex variables setting corresponds to the Hardy spaces on tubes. The Clifford algebra setting corresponds to the conjugate harmonic systems. Fourier spectrum properties of Hardy spaces on tubes were first studied in [20] and [21] p with the restriction on p =2. Certain one-way results for H (TΩ) and a partial range of the n n space index p, where Ω is an irreducible symmetric cone and TΩ = {x + iy ∈ C ; x ∈ R ,y ∈ Ω ⊂ Rn}, were obtained in [8]. In [12] H¨ormander proved some results corresponding to the type of Paley-Wiener Theorem for bandlimited functions involving entire functions in several complex variables. Fourier spectrum characterizations of the Hardy spaces on tubes for all cases 1 ≤ p ≤∞ are thoroughly studied in [14]. The present paper gives Fourier spectrum characterizations for functions in the Clifford algebra-valued Hardy spaces for the whole range p ∈ [1, ∞]. As a particular case, the vector-valued case corresponding to the conjugate harmonic systems was previously and fundamentally studied in [21, 20], and further in [10, 16, 13]. The previous studies, besides the restriction to vector-values, were also restricted to the singular integral version, and the index range is restricted to 1 ≤ p < ∞. The main results of this study imply the Hilbert transformation eigenvalue characterizations of the Hardy spaces and the Hardy spaces decompositions of the Lp functions. The crucial notion with the Clifford algebra setting of the Euclidean spaces is the projec- tion functions χ± defined by

1 ξ χ (ξ)= 1 ± i ± 2 |ξ|   and the associated generalizations of the trigonometrical exponential function

2 ± 2πihx,ξi ∓2πx0|ξ| e (x, ξ)= e e χ±(ξ), where x = x0 + x. The purpose of this paper is to declare the Fourier multiplier form of the Clifford Cauchy integral representation formula for the Clifford algebra-valued Hardy Hp functions F in the n+1 upper-half space R+ for all p ∈ [1, ∞]:

2πihx,ξi −2πx0|ξ| F (x0 + x)= e e χ+(ξ)Fˆ(ξ)dξ. Rn Z The above formula is also the Laplace transform of functions on Rn in the Clifford algebra setting provided that the formula makes sense. We note that when 1 ≤ p ≤ 2, the above relation is valid in the Lebesgue integration sense, while when p > 2 it is valid in the distribution sense. We will prove that, for all p ∈ [1, ∞], a Clifford algebra-valued Lp(Rn)- p n+1 function F satisfies F ∈ H (R+ ) (the NTBL of some Clifford algebra-valued Hardy space function) if and only if Fˆ = χ+F,ˆ where the multiplication between χ+ and the distribution ˆ p n+1 F will be precisely defined in the rest part of the paper; and F ∈ H (R+ ) if and only if n HF = F, where H = − k=1 ekRk is the Hilbert transformation, and Rk,k =1, ..., n are the Riesz transformations. For the Clifford algebra-valued Hardy spaces in the lower-half of the n+1 P space R we have the counterpart results Fˆ = χ−Fˆ; and HF = −F. In [21] (see also Propositions 2.3-2.5) the characterization by the Riesz transformations of the NTBLs of the vector-valued Hardy spaces functions, or alternatively the conjugate harmonic systems, for 1 ≤ p < ∞, is proved. When p = 2, the corresponding Fourier 2 n+1 spectrum characterization of the vector-valued functions in H (R+ ) can be directly derived from Theorem 3.1 in [21, page 65] (see also Proposition 2.3) through the relation Fˆ = ˆ ∧ ∧ ˆ ξ1 ˆ ξn ˆ (f0, (R1(f0)) , ..., (Rn(f0)) )=(f0, −i |ξ| f0, ..., −i |ξ| f0). Our study systematically treats the Fourier multiplier aspect for the Clifford algebra-valued functions case, and for the whole index range 1 ≤ p ≤∞, involving distributional and BMO . The paper is organized as follows. In §2 some notations and terminologies are given. In p n+1 §3 we prove the main results. In §4, as an application of the Fourier spectrum of H (R+ ), n+1 we prove the analogous result in the Clifford algebra-valued Bergman spaces on R+ .

2 Preliminaries

2.1 Hp space in terms of conjugate harmonic systems

n+1 n+1 n Let R+ = {x = (x, x0) ∈ R ; x0 > 0, x ∈ R }, and uj, 0 ≤ j ≤ n, be functions defined n+1 on R+ . Suppose F =(u0,u1,...,un) satisfies the generalized Cauchy-Riemann systems in n+1 R+ , i.e.,

n ∂u j =0, ∂xj j=0 X (2.1) ∂u ∂u j = k , j =6 k, 0 ≤ j, k ≤ n. ∂xk ∂xj

3 Such (n + 1)-tuple F =(u0,u1,...,un) is called a conjugate harmonic system ([21, 20]). For p n+1 0

1 p p ||F ||p = sup |F (x, x0)| dx < ∞. (2.2) x > Rn 0 0 Z  ∞ n+1 For p = ∞, we say that F ∈ H (R+ ) if F satisfies (2.1), and

n+1 ||F ||∞ = sup{|F (x, x0)| : (x, x0) ∈ R+ } < ∞. (2.3)

||F ||p is a when 1 ≤ p ≤ ∞, and a norm when p = 2 (see c.f. [21, page p n+1 220]). For H (R+ ), the following results for harmonic functions are well-known. n+1 Proposition 2.1 ([21, page 78]) Let u(x, x0) be harmonic in R+ . One has: p n (a) If 1 < p ≤ ∞, u(x, x0) is the Poisson integral of an L (R ) function if and only if sup ku(x, x0)kp < ∞. x0>0 (b) u(x, x0) is the Poisson integral of a Borel if and only if sup ku(x, x0)k1 < ∞. x0>0

p n Proposition 2.2 ([3, Theorem 1.3, page 62]) For f ∈ L (R ), 1 ≤ p ≤ ∞, if u(x, x0) is the Poisson integral of f, then ∗ (a) u (x) = sup|y−x|<αx0 |u(y, x0)| ≤ AM(f)(x), where α> 0, M(f) is the Hardy-Littlewood maximal function of f. (b) For x ∈ Rn, one has

lim u(y, x0)= f(x). (y,x0)→(x,0),|y−x|<αx0

2 n Proposition 2.3 ([21, page 65]) Let f0 and f1,...,fn belong to L (R ), and let their re- spective Poisson integrals be u0(x, x0)= Px0 ∗ f0,u1(x, x0)= Px0 ∗ f1,...,un(x, x0)= Px0 ∗ fn. Then a necessary and sufficient condition that

fj = Rj(f0), j =1, . . . , n, is that (u0, ..., un) satisfies the generalized Cauchy-Riemann equations (2.1), where Rj is the j-th Riesz transformation, that is,

1 xj − yj fj(x)= Rj(f0)(x) = lim f0(y)dy (2.4) ǫ→0 σ |x − y|n+1 n Z|x−y|>ǫ n+1 π 2 with σn = n+1 . Γ( 2 )

p n+1 Proposition 2.4 ([21, page 220]) Suppose that F ∈ H (R+ ), 1

4 1 n+1 Proposition 2.5 ([21, page 221]) Let F ∈ H (R+ ). Then lim F (x, x0) = F (x) exists x0→0 almost everywhere, as well as in the L1(Rn) norm. Also

sup |F (x, x0)|dx ≤ A sup |F (x, x0)|dx = AkF k1. Rn Rn Z x0>0 x0>0 Z 1 n+1 1 n Moreover, the space H (R+ ) is naturally isomorphic with the space of L (R ) functions f0 1 n 1 which have the property that Rj(f0) ∈ L (R ), j =1, . . . , n. The H norm is then equivalent n 1 n 1 n with kf0k1 + kRj(f0)k1. Conversely, suppose f0 ∈ L (R ) and Rj(f0) ∈ L (R ), Rj(f0)= j=1 fj, j =1, . . . ,P n. Then

n n

sup |uj(x, x0)|dx ≤ A kfjkL1 , Rn x0>0 j=0 j=0 X Z X where uj = Px0 ∗ fj, j =0, ..., n.

It is noted that the Riesz transformations Rj, 1 ≤ j ≤ n, defined as (2.4) are from Lp(Rn) to Lp(Rn), 1

1 xj − yj −yj Rj(f0)(x) = lim − χ{|y|>1}(y) f0(y)dy, ǫ→0 σ |x − y|n+1 | − y|n+1 n Z|x−y|>ǫ   ∞ n which are well-defined for f0 ∈ L (R ) up to a constant, and χ{|y|>1} is the characteristic function for {y ∈ Rn; |y| > 1}. The revised Riesz transformations are bounded from L∞(Rn) to BMO(Rn) ( space).

2.2 Clifford algebra-valued Hardy Hp spaces

Let e1, ..., en be basic elements satisfying

ejek + ekej = −2δjk, j,k =1, ..., n,

n where δjk is the Kronecker delta function. Let R = {x = x1e1 + ··· + xnen; xj ∈ R, 1 ≤ j ≤ n n+1 n n} be identical with the usual Euclidean space R , and R+ = {x0 + x; x0 > 0, x ∈ R }. (n) (n) The real (complex) Clifford algebra R (C ) generated by e1, ..., en, is the associative (n) algebra generated by e1, ..., en over the real (complex) field R (C). The elements of R (n) (C ) are of the form x = T xT eT , where T = {1 ≤ j1 < j2 < ··· < jl ≤ n} runs over all ordered subsets of {1, ..., n}, xT ∈ R (C) with x∅ = x0, and eT = ej1 ej2 ··· ej with the P l identity element e∅ = e0 = 1. Sc x := x0 and NSc x := x − Sc x are respectively called the scalar part and the non-scalar part of x. In this paper, we denote the conjugate of x ∈ C(n) by x = T xT eT , where eT = ejl ··· ej2 ej1 with e0 = e0 and ej = −ej for j =6 0. The n 1 1 norm of x ∈ C( ) is defined as |x| := (Sc xx) 2 = ( |x |2) 2 . Generally, a Clifford algebra P T T is not a division algebra, unless n = 2, that corresponds to the algebra of quaternions. n+1 P x = x0 + x ∈ R is called a vector or a para-vector, and the conjugate of a vector x is

5 −1 x x = x0 − x. If x is a vector then x = |x|2 . For more information about Clifford algebra, we refer to [2]. A Clifford algebra-valued function F is left-monogenic (resp. right-monogenic) if

n n

DF =( ∂kek)F =0 resp. FD = F ( ∂kek)=0 , k k ! X=0 X=0 ∂ where ∂k = , 0 ≤ k ≤ n, and D is the Dirac operator. Note that D(DF ) = ∆F = 0 if F ∂xk is left-monogenic, which means that each component of F is harmonic. A function that is both left- and right-monogenic is called a monogenic function. Vector-valued left-monogenic functions are simultaneously right-monogenic functions, and vice-versa, and thus they are monogenic. The Fourier transform of a function in L1(Rn) is defined as

fˆ(ξ)= F(f)(ξ)= e−2πihx,ξif(x)dx, Rn Z n where ξ = ξ1e1 + ··· + ξnen ∈ R , and the inverse Fourier transform is formally defined as

g∨(x)= F −1(g)(x)= e2πihx,ξig(ξ)dξ. Rn Z We note that the Fourier transformation is linear and thus it, together with some of its properties, can be extended to Clifford algebra-valued functions. In particular, the holds for Clifford algebra-valued functions: For Clifford algebra-valued functions f,g ∈ L2(Rn) there holds

f(x)g(x)dx = fˆ(ξ)gˆ(ξ)dξ. Rn Rn Z Z An alternative form of the Plancherel Theorem is

f(x)g(x)dx = fˆ(ξ)g∨(ξ)dξ. Rn Rn Z Z Define, for x = x0 + x, e(x, ξ)= e+(x, ξ)+ e−(x, ξ) with ± 2πihx,ξi ∓2πx0|ξ| e (x, ξ)= e e χ±(ξ)

1 ξ (see e.g. [15]), where χ±(ξ)= 2 (1 ± i |ξ| ). χ± enjoy the following projection-like properties:

2 χ−χ+ = χ+χ− =0, χ± = χ±, χ+ + χ− =1. (2.5)

n+1 Definition 2.6 Let F (x) = T fT (x)eT , where x = x0 + x ∈ R+ . If F is left-monogenic on Rn+1 and satisfies + P p p ||F ||Hp = sup |F (x0 + x)| dx < ∞, 1 ≤ p< ∞, (2.6) Rn x0>0 Z 6 p n+1 n+1 then we say F (x) belongs to the Hardy space H (R+ ). If F is left-monogenic on R+ and satisfies ||F ||H∞ = sup |F (x)| < ∞, Rn+1 x∈ + ∞ n+1 then we say F (x) ∈ H (R+ ).

2 n+1 When p =2, the inner product of H (R+ ) is defined as

hF,Gi = G(x)F (x)dx, Rn Z 2 n+1 where F (x) and G(x) are respectively the NTBL functions of F and G. The norm of H (R+ ) is, in fact, equal to

2 ||F ||H2 = Sc hF, F i.

p n+1 In particular, if F ∈ H (R+ ), 1 ≤ p ≤ ∞, is a vector-valued function, i.e., F (x) = n p n+1 f0(x)+ j=1 fj(x)ej, then F corresponds to an element in H (R+ ) ([13]). To see this, we let U(x0 + x)= u0(x0 + x) − u1(x0 + x)e1 −···− un(x0 + x)en. With e0 =1, we have P n

DU = ( ∂kek)[u0(x0 + x)e0 − u1(x0 + x)e1 −···− un(x0 + x)en] k X=0 ∂u0 ∂u1 ∂un ∂uj ∂uk = + + ··· + + ( − )ejek, (2.7) ∂x ∂x ∂xn ∂xk ∂xj 0 1 j

n

U(x)= u0(x) − Rj(u0)(x)ej. j=1 X p n+1 For general vector-valued F ∈ H (R+ ), 1 ≤ p< ∞, there holds n

F (x)= f0(x) − Rj(f0)(x)ej =(I + H)f0(x), j=1 X where the operator H, the Hilbert transformation, is defined by

n

H = − Rjej. j=1 X ξj Note that the Fourier multiplier of Rj is −i |ξ| (see e.g. [11]), i.e.,

ξ ∨ R (f )(x)= −i j fˆ (ξ) (x). j 0 |ξ| 0  

7 1 This implies that the Fourier multipliers of 2 (I ± H) are, respectively, χ±(ξ). The Cauchy integral formula in Clifford analysis may be regarded as the reproducing for- p n+1 mula. In particular, for F ∈ H (R+ ), the following reproducing formula plays an important role.

p n+1 Proposition 2.7 (see e.g. [10, 16]) For F ∈ H (R+ ), 1 ≤ p< ∞, we have

F (x)= E(x − y)F (y)dy, Rn Z 1 x where E(x)= n+1 is the Cauchy kernel, and F (y) is the NTBL function of F . 2σn |x| When 1 0 such that

||Cx0 (f)||Lp ≤ L||f||Lp , where

Cx0 (f)(x)= E(x − y)f(y)dy, x = x0 + x, x0 > 0, Rn Z and L is independent of x0 and f. As a consequence, as a function of x, the function Cx0 (f)(x) p n+1 belongs to H (R+ ). Moreover,

1 n lim Cx0 (f)(x)= (f(x)+ Hf(x)), a.e. x ∈ R . x0→0+ 2 There are parallel results for the lower-half of the space Rn+1. If we apply Proposition 2.8 to the NTBL of a Hardy space function F, then we conclude, from the last relation, H2 = I. Let Ψ(Rn) be the Clifford algebra-valued , whose elements are given by

ψ(ξ)= ψT (ξ)eT , T X n ± n n where ψT are in the Schwartz space S(R ). Denote by Ψ (R ) the subclasses of Ψ(R ) consisting of the Clifford algebra-valued Schwartz functions of, respectively, the forms

ψ(ξ)= ψ(ξ)χ±(ξ), where ψ(ξ) takes the zero value in some neighborhood of the origin. It is easy to show that the direct sum Ψ+(Rn) ⊕ Ψ−(Rn) is dense in Ψ(Rn). ± n ± n Denote by Ψscalar(R ) the subclasses of Ψ (R ), whose elements are of the form

ψ(ξ)= ψ0(ξ)χ±(ξ), where ψ0 is scalar-valued. e For more information on the Clifford algebra-valued distribution theory, we refer to, e.g., [2] and [17].e

8 3 Main Results

Our main results are as follows.

p n+1 Theorem 3.1 For F ∈ H (R+ ), 1 ≤ p ≤∞, we have

(Fˆ , ψ)=(F, ψˆ)= ψˆ(x)F (x)dx =0, Rn Z where ψ ∈ Ψ−(Rn).

Conversely, we have

p n − n Theorem 3.2 Let F ∈ L (R ), 1 ≤ p ≤ ∞. If (F,ˆ ψ)=0 for ψ ∈ Ψ (R ) (Fˆ = χ+Fˆ if p n+1 1 ≤ p ≤ 2), then F (x) is the NTBL function of some F (x) ∈ H (R+ ). In particular, we have

p n+1 p n ˆ ˆ Corollary 3.3 F ∈ H (R+ ), 1 ≤ p ≤ 2 if and only if F ∈ L (R ) and F (ξ)= χ+(ξ)F (ξ), and

F (x)= e+(x, ξ)Fˆ(ξ)dξ = E(x − y)F (y)dy. Rn Rn Z Z Remark 3.4 Theorem 3.1 and Theorem 3.2 extend Proposition 2.3 and Proposition 2.5, respectively, to the non-vector-valued Hardy spaces cases. Moreover, the sufficiency and the necessity of the result for the case p = ∞ are not proved in either the vector-valued or the Clifford algebra-valued Hardy spaces in literatures.

Proof of Theorem 3.1: We first prove the result for 1

p where q = p−1 , L is given in Proposition 2.8, and Cn is a constant depending on n. For any ǫ> 0, G is chosen as a smooth function with compact support such that

||F − G||Lp < ǫ.

9 2 n+1 Since Cx0 (G) ∈ H (R+ ), there follows

Cx0 (G)(x) = E(x0 + x − y)G(y)dy Rn Z x0 = E( + x − y)C x0 (G)(y)dy 2 Rn 2 Z x 2πihx,ξi −2π 0 |ξ| ∧ = e e 2 χ (ξ)(C x0 (G)) (ξ)dξ. + 2 Rn Z This shows that x0 ∧ −2π 2 |ξ| x ∧ (Cx0 (G)) (ξ)= e χ (ξ)(C 0 (G)) (ξ). + 2 Therefore,

π x0 ξ ˆ ∧ −2 2 | | x ∧ (Cx0 (G)(x), ψ(x)) = ((Cx0 (G)) (ξ), ψ(ξ))=(e χ (ξ)(C 0 (G)) (ξ), ψ(ξ))=0. + 2 Hence, through (3.8), for any ǫ> 0, we have

|(F, ψˆ)| ≤ CnLǫ||ψˆ||Lq +0, which shows (F, ψˆ)=0. Now we deal with the case p =1. We use the following Lemma (for the setting, see also e.g. [4]).

1 n+1 2 n Lemma 3.1 If F ∈ H (R+ ), then Fx0 (·)= F (x0 + ·) ∈ L (R ).

Proof Since F is left-monogenic, by the mean-value theorem, we have 1 F (x0 + x)= F (y0 + y)dy, x V x0 B ( 0 ) 2 Z x 2

x0 x0 x0 where Bx( ) is the ball centered at x with radius , V x0 is the volume of Bx( ). Then we 2 2 2 2 have 1 |F (x0 + x)| ≤ |F (y0 + y)|dy x V x0 B ( 0 ) 2 Z x 2 3x0 C 2 ≤ n+1 |F (y0 + y)|dydy0 x x0 Rn 0 Z 2 Z C ≤ n sup |F (y0 + y)|dy, x Rn 0 y0>0 Z ∞ n where C is a constant. This implies that Fx0 (x) ∈ L (R ). Then, we have

2 |F (x0 + x)| dx = |F (x0 + x)||F (x0 + x)|dx Rn Rn Z Z ≤ sup |F (x0 + x)| |F (x0 + x)|dx < ∞. Rn Rn x∈ Z

10 The proof of the lemma is complete. 1 n+1 Now we continue to prove Theorem 3.1. Let F ∈ H (R+ ), x0 > 0. Then we have ˆ ˆ ˆ (F, ψ)=(F − Fx0 , ψ)+(Fx0 , ψ).

For any ǫ> 0, for a suitably chosen x0 > 0, we have ˆ ˆ ˆ |(F − Fx0 , ψ)| ≤ CnkF − Fx0 k1kψk∞ ≤ Cnǫkψk∞. On the other hand, as in the proof for the case 1 0. That shows (F, ψˆ)=0. For p = ∞, we will mainly adopt the technique used in [19]. In Lemma 3.2, we will show ∞ n+1 that for F ∈ H (R+ ), F (x)= HF (x)+ c, where c is a constant. It is easily shown that (c, ψˆ)=0, ψ ∈ Ψ−(Rn). Hence, in the following we only need to consider 1 1 F (x)= (I + H)F (x)= (I + H)f (x)e . 2 2 T T T X − n n We first consider that ψ ∈ Ψscalar(R ). We have ψ(ξ)= ψ0(ξ)χ−(ξ)= ψ0(ξ)+ j=1 ψj(ξ)ej, where ψ0 is a scalar-valued Schwartz function. In particular, we note that e P 1 ξ −2πihx,ξi e ψˆ(x)= ψ0(ξ) (1 − i )e dξ Rn 2 |ξ| Z ξ e 1 2πihx,ξi = ψ0(−ξ) (1 + i )e dξ, Rn 2 |ξ| Z ˆ e 1 n+1 which means that ψ(x) is the NTBL function of functions in H (R+ ). In the following we first accept, and use this property to prove that (F, ψˆ) = 0 for ψ ∈ Ψ−(Rn). Since ˆ 1 n+1 n ˆ ψ ∈ H (R+ ) and (I + H)fT (x) ∈ BMO(R ), ((I + H)fT (x), ψ) is well-defined. As shown in [7, page 146], we have n

ψˆ(x)Rj(fT )(x)dx = ekψˆk(x)Rj(fT )(x)dx Rn Rn Z k=0 Z Xn

= − ekRj(ψˆk)(x)fT (x)dx Rn k X=0 Z = − Rj(ψˆ)fT (x)dx. Rn Z

11 Hence we have n n

ψˆ(x)(fT (x) − Rj(fT )(x)ej)dx = (ψˆ(I + Rjej))(x)fT (x)dx. Rn Rn j=1 j=1 Z X Z X

ˆ n One can easily show that ψ(I + j=1 Rjej)=0. In fact, since the Fourier multiplier of Rj is −i ξj , we have |ξ| P

n n ξ e 1 ξ ξ (ψˆ(I + R e ))∨(ξ)= ψ(ξ)(1 + i j j )= ψ (ξ) (1 − i )(1 + i )=0. j j |ξ| 0 2 |ξ| |ξ| j=1 j=1 X X e ˆ n ˆ − n Thus ψ(I + j=1 Rjej)=0, and hence ((I + H)fT (x), ψ) = 0 for ψ ∈ Ψscalar(R ). Conse- F, ψˆ ψ − n . ψ − n , quently, ( P)=0 for ∈ Ψscalar(R ) For ∈ Ψ (R ) we have

∧ (F, ψˆ)= eS(F, (ψSχ−) ). S X Then, for each S, we have ∧ (F, (ψSχ−) )=0 by applying the above argument. We have (F, ψˆ)=0 for ψ ∈ Ψ−(Rn). ✷

Proof of Theorem 3.2: We first consider the case p =1. Define

+ Φ(x0 + x)= e (x, ξ)Fˆ(ξ)dξ Rn Z and

Gx0 (x)= G(x0 + x)= Px0 (x − y)F (y)dy. Rn Z We then have

+ ˆ Φ(x0 + x)= e (x, ξ)F (ξ)dξ = Px0 (x − y)F (y)dy = G(x0 + x). Rn Rn Z Z 1 n We also note that Gx0 ∈ L (R ) since

1 1 1 ||Gx0 ||L ≤ Cn||F ||L ||Px0 ||L < ∞.

By interchanging the derivatives with the integral, we can show that Φ is left-monogenic + 1 n+1 since e (x, ξ) is monogenic. Thus Φ(x) ∈ H (R+ ). Next we consider the case for Lp(Rn), 1

0=(F,ˆ ψ)=((F +)∧ +(F −)∧, ψ)=((F −)∧, ψ), for all ψ ∈ Ψ−(Rn), (3.9)

12 + 1 − 1 + ∧ where F = 2 (I + H)F and F = 2 (I − H)F, and ((F ) , ψ) = 0 follows from the proof of Theorem 3.1 (Note that this argument only works for 1 < p ≤ ∞, not including p = 1). Consequently, we can have, for ϕ being a scalar-valued Schwartz function taking zero in some neighborhood of the origin,

− ∧ − ∧ − ∧ ((F ) ,ϕ)=((F ) , ϕχ+)+((F ) , ϕχ−)=0,

− ∧ − ∧ where ((F ) , ϕχ+) = 0 follows from the proof of Theorem 3.1, and ((F ) , ϕχ−) = 0 is given by (3.9). This implies that (F −)∧ is either zero or a distribution with support at the origin. For the latter case, (F −)∧ has to be a finite linear combination of the partial derivatives of the (see e.g. [12, 17]), which contradicts to F ∈ Lp(Rn), 1

∞ n+1 Lemma 3.2 For F ∈ H (R+ ), we have

F (x)= HF (x)+ c, (3.10) where c is a constant.

n n Proof: Denote by S0(R ), where φ ∈ S0(R ) is a scalar-valued Schwartz function such that φˆ takes the zero value in some neighborhood of the origin. To show (3.10), it suffices to show that

((I − H)F (x),φ(x)) = 0 (3.11)

n holds for all φ ∈ S0(R ). In fact, we have

((I − H)F (x),φ(x)) = (((I − H)F )∨(ξ), φˆ(ξ)), which means that ((I − H)F )∨ is either zero or a distribution with support at the origin. Thus (I − H)F (x)= c, and hence F (x)= HF (x)+ c, where c is a constant. ∨ φˆ(ξ) Let F = T fT eT , and ψ(x)= 2π|ξ| (x). Moreover, we have P   −2π|ξ|e−2πx0|ξ|ψˆ(ξ)= −e−2πx0|ξ|φˆ(ξ), and hence,

(∂0Px0 ∗ ψ)(x)= −Px0 ∗ φ(x).

13 n Note that for φ ∈ S0(R ), ψ is still a Schwartz function. For each T, we have

n

((I + ejRj)fT (x0 + x), −φ(x)) j=1 X n ˆ ∨ = ((I + ejRj)fT (x0 + x), (−2π|ξ|ψ(ξ)) (x)) j=1 X n ∨ =(fT (x0 + x), (−2π|ξ|ψˆ(ξ)) (x)(I − Rjej)) j=1 n X ˆ ξjej ∨ =(fT (x0 + x), [(−2π|ξ|ψ(ξ))(1 + i )] (x)) (3.12) |ξ| j=1 X n ∨ =(fT (x0 + x), −φ(x))+(fT (x0 + x), [ −2πiξjψˆ(ξ)ej] (x)) j=1 Xn ˆ ∨ =(fT (x), −Px0 ∗ φ(x))+(fT (x0 + x), [ −2πiξjψ(ξ)ej] (x)) j=1 X n

=(fT (x), (∂0Px0 ∗ ψ)(x))+(fT (x0 + x), − ∂jψ(x)ej). j=1 X In the second equality we have used the result proved in [7, page 146], while the other equalities follow from the properties of the Fourier transform and the Fourier multiplier of the Riesz transformations. Since ψ is a Schwartz function, for any ǫ> 0, we can find ψ,a C∞- function with compact support, such that ||ψ − ψ||L1 < ǫ and ||∂jψ − ∂jψ||L1 < ǫ, 1 ≤ j ≤ n. We also note that e

e 2 2 e 1 ∂ x0 1 −nx0 + |x| ∂0Px0 (x)= n+1 = n+1 . σ ∂x 2 2 2 σ 2 2 2 +1 n 0 (x0 + |x| ) n (x0 + |x| ) Thus,

|(fT (x), (∂0Px0 ∗ (ψ − ψ))(x))| ′ 2 C x0 ≤ e n+1 |ψ(y) − ψ(y)|dydx σn Rn Rn x2 x y 2 2 +1 Z Z ( 0 + | − | ) C′′ |x − y|2 e + n+1 |ψ(y) − ψ(y)|dydx σn Rn Rn x2 x y 2 2 +1 Z Z ( 0 + | − | ) ′′ ′ C e 1 1 ≤ C ||Px0 ∗ (|ψ − ψ|)||L + ||Px0 ∗ (|ψ − ψ|)||L x0 ′′ ′ C ≤ (C + )||ψ −eψ||L1 , e x0 ≤ Cǫ. e

14 where C′,C′′ and C are constants. Hence, n

(3.12) = ((∂0Px0 ∗ fT )(x), ψ(x))+( ∂jfT (x0 + x)ej, ψ(x)) j=1 X e n e +(fT (x), ∂0Px0 ∗ (ψ − ψ)(x))+(fT (x0 + x), −(∂jψ(x) − ∂jψ(x))ej) j=1 X n e e = ((∂0 + ∂jej)fT (x0 + x), ψ(x)) j=1 X e n +(fT (x), ∂0Px0 ∗ (ψ − ψ)(x))+(fT (x0 + x), −(∂jψ(x) − ∂jψ(x))ej) j=1 X =(DfT (x0 + x), ψ(x)) e e n

+(fT (x), ∂0Px0 ∗ (eψ − ψ)(x))+(fT (x0 + x), −(∂jψ(x) − ∂jψ(x))ej). j=1 X e e Then, we have n

((I + ejRj)F (x0 + x), −φ(x)) j=1 X =(DF (x0 + x), ψ(x)) n

+(F (x), ∂0Px0 ∗ (eψ − ψ)(x))+(F (x0 + x), −(∂jψ(x) − ∂jψ(x))ej). j=1 X Since F is left-monogenic, we havee e n

((I + ejRj)F (x0 + x), −φ(x)) j=1 X n

=(F (x), ∂0Px0 ∗ (ψ − ψ)(x))+(F (x0 + x), −(∂j ψ(x) − ∂jψ(x))ej). j=1 X Thus, for any ǫ> 0, we have e e n

|((I + ejRj)F (x0 + x), −φ(x))| j=1 X n

≤|(F (x), ∂0Px0 ∗ (ψ − ψ)(x))| + |(F (x0 + x), −(∂jψ(x) − ∂jψ(x))ej)| j=1 X n e e ′′′ ≤ C (||ψ − ψ||L1 + ||∂jψ − ∂jψ)||L1 ) j=1 X ≤ (n + 1)C′′′eǫ, e

15 where C′′′ is a constant. Hence,

n

((I + ejRj)F (x0 + x), −φ(x))=0. j=1 X Finally, we have

n n

((I + ejRj)F (x),φ(x)) = lim ((I + ejRj)F (x0 + x),φ(x))=0 x0→0 j=1 j=1 X X n for all φ ∈ S0(R ). The proof is completed.

∞ n+1 Remark 3.5 If F ∈ H (R+ ) is vector-valued, the above result becomes

n

F (x)= f0(x) − Rj(f0)(x)ej + c, j=1 X which is a special case of the celebrated characterization of BMO proved by Fefferman and Stein in [7]. p n+1 We note that the proof of the above lemma holds for H (R+ ), 1 ≤ p ≤∞. This means p n+1 that we have given an alternative proof of the Cauchy integral formula for F ∈ H (R+ ), 1 ≤ p< ∞, i.e.,

F (x0 + x)= E(x − y)F (y)dy, Rn Z ∞ n+1 and for F ∈ H (R+ ),

F (x0 + x)= E(x − y)F (y)dy + c, Rn Z where E(x − y) is revised corresponding to the revision of Rj, 1 ≤ j ≤ n. When p = 2, we can prove F (x) = HF (x) in the normal sense. The proof of the above lemma is motivated by the case p =2. In fact, a direct computation of DF =0 yields

n lj (∂0fT (x0 + x)+ (−1) ∂jfTj (x0 + x))eT =0, for all T, (3.13) j=1 X

lj where Tj satisfies (−1) ejeTj = eT . We note that

lj = N(j ∩ Tj)+ P (j, Tj), where N(A) = #A denotes the number of elements in some set A, and

P (j, Tj) = #{k; j >k,k ∈ Tj}.

16 Taking the Fourier transform on x, we have

n ˆ lj ˆ (∂0fT (x0 + ξ)+ (−1) (2πiξj)fTj (x0 + ξ))eT =0. j=1 X Using the fact that

ˆ ∧ −2πx0|ξ| ˆ ˆ lim ∂0fT (x0 + ξ) = lim ∂0(Px0 ∗ fT ) (ξ) = lim ∂0(e fT (ξ)) = −2π|ξ|fT (ξ), x0→0 x0→0 x0→0 we have n ξj (fˆ (ξ)+ (−1)lj (−i )fˆ (ξ))e =0, T |ξ| Tj T j=1 X and consequently,

n lj (fT (x)+ (−1) Rj(fTj )(x))eT =0. j=1 X This means that F (x)= HF (x). We also note that the system (3.13) is indeed a generalization of the conjugate harmonic system (2.1). When F is vector-valued, the system (3.13) is reduced to (2.1) (see the simple p n+1 n argument given in §2). Consequently, for H (R+ ) ∋ F = f0 + j=1 fjej, F = HF is then reduced to fj = −Rj(f0), 1 ≤ j ≤ n, which is the classical result for the conjugate harmonic system (see Propositions 2.3−2.5). P

4 Analogue in

In this section we prove a representation formula for the functions in the Clifford algebra- valued Bergman space in the upper half-space as an application of our results in the previous sections. For the analogous result in the complex analysis setting, we refer to [6] and [1].

n+1 n+1 Definition 4.1 Let F (x) = T fT (x)eT , where x ∈ R+ . If F is left-monogenic on R+ , and satisfies P ∞ p p ||F ||Ap = |F (x0 + x)| dxdx0 < ∞, 1 ≤ p< ∞, Rn Z0 Z p n+1 then we say that F belongs to the Bergman space A (R+ ). p n+1 As an application of the Fourier spectrum characterization of H (R+ ), 1 ≤ p< ∞, we have

p n+1 q n p Theorem 4.2 For F ∈ A (R+ ), 1 ≤ p ≤ 2, there exists a function G ∈ L (R ), q = p−1 , such that

F (x)= e+(x, ξ)G(ξ)dξ. Rn Z

17 Moreover, for 1

1 q q |χ+(ξ)G(ξ)| q dξ ≤ ||F ||Ap < ∞, Rn πp ξ p Z (2 | |) ! and for p =1,

|χ+(ξ)G(ξ)| sup ≤ ||F ||A1 < ∞. ξ∈Rn 2π|ξ|

p n+1 Theorem 4.3 For F ∈ A (R+ ), 2 0, there holds

(F (x0 + ·), ψˆ)=0 for ψ ∈ Ψ−(Rn).

Proof of Theorem 4.2 Since F is left-monogenic, |F |p is subharmonic for 1 ≤ p ≤ 2. We thus have 1 |F (x)|p ≤ |F (y + y)|pdy, V 0 δ ZBx(δ) x0 where Bx(δ) is the ball centered at x with radius 0 <δ

C x0+δ |F (x)|p ≤ |F (y + y)|pdydy δn+1 0 0 Zx0−δ Z|x−y|<δ C ∞ ≤ |F (y + y)|pdydy , δn+1 0 0 Z0 Z|x−y|<δ and then, C ∞ F x x pdx F y y pdydy dx | ( 0 + )| ≤ n+1 | ( 0 + )| 0 Rn δ Rn Z Z Z0 Z|x−y|<δ C ∞ χ x dx F y y pdydy = n+1 By(δ)( ) | ( 0 + )| 0 δ Rn Rn Z0 Z Z C′δn ∞ F y y pdydy = n+1 | ( 0 + )| 0 δ 0 Rn ′ Z∞ Z C p = |F (y0 + y)| dydy0 δ Rn Z0 Z where the second equality used Fubini’s theorem, and C and C′ are constants. The above p n+1 p n+1 inequality implies that Fy0 (x) = F (y0 + x) ∈ H (R+ ) for F ∈ A (R+ ) and y0 > 0. Consequently, we have

+ ˆ Fy0 (x)= e (x, ξ)Fy0 (ξ)dξ Rn Z

18 and

Fy0 (x)= Px0 (x − y)Fy0 (y)dy. Rn Z Thus

ˆ −2πx0|ξ| ˆ Fy0+x0 (ξ)= e Fy0 (ξ) and hence

2π(y0+x0)|ξ| ˆ 2πy0|ξ| ˆ e Fy0+x0 (ξ)= e Fy0 (ξ).

2πy0|ξ| ˆ Therefore, if we let G(ξ)= e Fy0 (ξ), which is independent of y0, then we have

+ −2πy0|ξ| Fy0 (x)= e (x, ξ)e G(ξ)dξ Rn Z + = e (y0 + x, ξ)G(ξ)dξ. Rn Z Consequently,

F (x)= e+(x, ξ)G(ξ)dξ. Rn Z Moreover, by Hausdorff-Young’s inequality, we have, for 1

1 1 q p q −2πx0q|ξ| p |χ+(ξ)G(ξ)| e dξ ≤ |F (x0 + x)| dx , Rn Rn Z  Z  and for p =1,

−2πx0|ξ| sup |χ+(ξ)G(ξ)|e ≤ |F (x0 + x)|dx. Rn Rn ξ∈ Z For 1

p q p q ∞ p ∞ q p −2πx0p|ξ| q −2πx0q|ξ| |χ+(ξ)G(ξ)| e dx0 dξ ≤ |χ+(ξ)G(ξ)| e dξ dx0 Rn Rn Z Z0  ! Z0 Z  ∞ p ≤ |F (x0 + x)| dxdx0, Rn Z0 Z and thus

1 1 q q ∞ p |χ+(ξ)G(ξ)| p q ≤ |F (x0 + x)| dxdx0 < ∞. Rn p Rn Z (2πp|ξ|) ! Z0 Z 

19 For p =1,

∞ |χ+(ξ)G(ξ)| sup ≤ |F (x0 + x)|dxdx0 < ∞. Rn 2π|ξ| Rn ξ∈ Z0 Z ✷

Proof of Theorem 4.3 p n+1 p n+1 As in the proof of Theorem 4.2, we can show that Fy0 (x) ∈ H (R+ ) for F ∈ A (R+ ), 2 < p< ∞ and y0 > 0. Then, by Theorem 3.1 we complete the proof. ✷

References

[1] D. B´ekoll´e and A. Bonami, Hausdorff-Young Inequalities for Functions in Bergman Spaces on Tube Domains, Proceedings of the Edinburgh Mathematical Society, 41 (1998), pages: 553–566.

[2] F. Brackx, R. Delanghe and F. Sommen, Clifford Analysis, Research Notes in Mahtem- atics 76, Pitman Advanced Publishing Company, Boston London, Melbourne, 1982.

[3] D.-G. Deng and Y.-S. Han, Theory of Hp spaces (in Chinese), Peking University Press, Peking, 1992.

[4] G.-T. Deng and T. Qian, Rational Approximation of functions in Hardy spaces, Complex Analysis and , 10 (2016), pages: 903–320.

[5] P. Duren, Theory of Hp Spaces, Pure and Applied Mathematics, Volume 38, Academic Press, New York and Lodon, 1970.

[6] T. G. Genchev, Paley-Wiener Type Theorems for Functions in Bergman Spaces over Tube Domains, Journal of and Applications, 118 (1986), pages: 496–501.

[7] C. Fefferman and E. M. Stein, Hp spaces of several variables, Acta Math. 129 (1972), pages: 137–193.

[8] G. Garrig´os, Generalized Hardy Spaces on Tube Domains over cones, Colloq. Math. 90 (2001), pages: 213–251.

[9] J. B. Garnett, Bounded analytic functions, Academic Press, 1987.

[10] J. Gilbert and M. Murray, Clifford Algebras and Dirac Operators in , Cambridge University Press, 1991.

[11] L. Grafakos, Classical Fourier Analysis (Second Edition), Graduate Texts in Mahtem- atics 249, Springer, 2008.

20 [12] L. H¨ormander, The Analysis of Linear Partial Differential Operator, I. Distribution theorey and Fourier analysis, 2nd Ed., Springer-Berlag, Berlin, 2003.

[13] K.-I. Kou and T. Qian, The Paley-Wiener Theorem in Rn with the Clifford Analysis Setting, Journal of Functional Analysis, 189 (2002), pages: 227–241.

[14] H.-C. Li, G.-T. Deng and T. Qian, Fourier Spectrum Characterizations of Hp Spaces on Tubes over Cones for 1 ≤ p ≤ ∞, Complex Analysis and Operator Theory, 12 (2018), pages: 1193–1218.

[15] C. Li, A. McIntosh and T. Qian, Clifford Algebras, Fourier Transform and Singular Convolution Operators on Lipschitz Surfaces, Revista Mathematica Iberoamericana, 10 (1994), pages: 665–721.

[16] M. Mitrea, Clifford Wavelets, Singular Integrals and Hardy Spaces, Lecture Note in Mathematics, Springer-Verlag, Berlin Heidelberg, 1994.

[17] D. Mitrea, Distributions, Partial Differential Equations, and Harmonic Analysis, Uni- versitext, Springer, New York, 2013.

[18] T. Qian, Characterization of boundary values of functions in Hardy spaces with applica- tions in signal analysis, J. Integral Equations Appl. 17 (2005), pages: 159–198.

[19] T. Qian, Y. S. Xu, D. Y. Yan, L. X. Yan, B. Yu, Fourier spectrum characterization of Hardy spaces and applications, Proceedings of the American Mathematical Society, 137 (2009), pages: 971–980.

[20] E. M. Stein, G. Weiss, Introduction to fourier analysis on Euclidean spaces, Princeton University Press, 1971.

[21] E. M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, Princeton, New Jersey, 1970.

[22] A. Uchiyama, A Constructive Proof of the Fefferman-Stein Decomposition of BMO(Rn), Acta Math. 148 (1982), pages: 215–241.

Faculty of Information Technology, Macau University of Science and Technology, Macao, China E-mail address: [email protected] School of Mathematics (Zhuhai), Sun Yat-Sen University (Zhuhai), China E-mail address: [email protected] Macau Institute of Systems Engineering, Macau University of Science and Technology, Macao, China E-mail address: [email protected]

21