ISOTROPIC POSITIVE DEFINITE FUNCTIONS ON SPHERES GENERATED FROM THOSE IN EUCLIDEAN SPACES

A Dissertation by

Zhihui Nie

Master of Science, Tianjin University of Technology and Education, 2014

Bachelor of Science, Tianjin University of Technology and Education, 2011

Submitted to the Department of Mathematics, Statistics, and Physics and the faculty of the Graduate School of Wichita State University in partial fulfillment of the requirements for the degree of Doctor of Philosophy

May 2019 © Copyright 2019 by Zhihui Nie All Rights Reserved ISOTROPIC POSITIVE DEFINITE FUNCTIONS ON SPHERES GENERATED FROM THOSE IN EUCLIDEAN SPACES

The following faculty members have examined the final copy of this dissertation for form and content, and recommend that it be accepted in partial fulfillment of the requirement for the degree of Doctor of Philosophy with a major in Applied Mathematics.

Chunsheng Ma, Committee Chair

Tianshi Lu, Committee Member

Xiaomin Hu, Committee Member

Adam Jaeger, Committee Member

Hyuck M. Kwon, Committee Member

Accepted for the College of Liberal Arts and Sciences

Andrew Hippisley, Dean

Accepted for the Graduate School

Kerry Wilks, Interim Dean

iii DEDICATION

To my parents

iv ACKNOWLEDGEMENTS

This dissertation would not have been possible without the guidance and help of several individuals who in one way or another contributed and extended their valuable as- sistance in the preparation and completion of this study. My first debt of gratitude must go to my advisor, Professor Chunsheng Ma, who was truly an inspiration. Without his invalu- able guidance, this thesis would not have been possible. I would also like to to express my gratitude to Professor Daowei Ma, Professor Tianshi Lu, Professor Xiaomin Hu, Professor Adam Jaeger, Professor Hyuck M. Kwon and Professor Hua Liu for the contribution of their valuable time. In addition, I am grateful to acknowledge my friends in Wichita State Uni- versity, J. Chen, X. Wei, L. Liang, Y. Wang, N. Nguyen, S. Rahmati, S. Zhang and F. Yan, for their friendship and warmth. Last but not the least, huge thank you to my parents, for their full supports.

v ABSTRACT

For a continuous function g(x) on [0, π] with g(π) = 0, if it satisfies the inequality

Z π α+ 1 u 2 g(u)J 1 (xu)du ≥ 0, x ≥ 0, α− 2 0 it is shown in this thesis that

Z π (α) 2α g(u)Pn (cos u) sin udu ≥ 0, n ∈ N0, 0

(ν) where α is a nonnegative integer, and Jν(x) and Pn (x) denote the Bessel function and the ultraspherical polynomial, respectively. As a consequence, an isotropic and continuous positive definite function in the Euclidean space, if it is compactly supported, can be adopted as an isotropic positive definite function on a sphere.

vi TABLE OF CONTENTS

Chapter Page

1 INTRODUCTION ...... 1

2 BACKGROUND INFORMATION ...... 5

2.1 Definitions ...... 5 2.2 The Gamma Function ...... 7 2.3 Orthogonal Polynomial ...... 8 2.4 Bessel Functions ...... 9 2.5 Abel-Summability ...... 10

3 ISOTROPIC POSITIVE DEFINITE FUNCTIONS IN EUCLIDEAN SPACE . . 12

3.1 Integral Transforms ...... 12 3.2 Characterizations of Positive Definite Functions ...... 13 3.3 Schoenberg’s Characterization ...... 14 3.4 Properties of Positive Definite Functions ...... 15

4 ISOTROPIC POSITIVE DEFINITE FUNCTIONS ON SPHERES ...... 20

4.1 Some Classical Results Concerning Ultraspherical Polynomials ...... 20 4.2 Schoenberg’s Characterization ...... 23 4.3 Recent Results ...... 24

5 ISOTROPIC POSITIVE DEFINITE FUNCTIONS ON SPHERES GENERATED FROM THOSE IN EUCLIDEAN SPACES ...... 26

5.1 Main Results ...... 26 5.2 Some Lemmas and Their Proofs ...... 30 5.3 Proof of Theorems 5.1.1 and 5.1.2 ...... 38

6 PLANS FOR FUTURE WORK ...... 41

REFERENCES ...... 42

vii CHAPTER 1

INTRODUCTION

There has run on renewed interest in isotropic positive definite functions in the Euclidean

space Rd or on the sphere Sd, with applications in approximation theory [4], [8], [10], [28],

machine learning [20], probability, and spatial statistics [6], [15], [16], where Sd is the spherical

shell of radius 1 and center 0 in Rd+1, i.e., Sd = {x ∈ Rd+1, kxk = 1}, kxk is the Euclidean norm of x, and d is a positive integer.

An isotropic (radial, or spherically symmetric) function g(kxk): Rd → R, is said to be positive definite, if the inequality

n n X X aiajg(kxi − xjk) ≥ 0 i=1 j=1

d holds for every n ∈ N, any ai ∈ R and xi ∈ R (i = 1, . . . , n), where N is the set of positive integers.

When g(kxk) is continuous in Rd or g(·) is continuous on R, it is well-known [21] that

g(kxk) is positive definite in Rd if and only if g(x) admits a representation of the form Z ∞ − d−2 g(x) = (xu) 2 J d−2 (xu)dF (u), x ≥ 0, (1.0.1) 0 2 where F (x) is a bounded and nondecreasing function on [0, ∞), and Jν(x) denotes a Bessel function of order ν.

As an example [2],  λ λ  (a − x) , 0 ≤ x ≤ a, g(x) = (a − x)+ = (1.0.2)  0, x > a,

d  d  makes g(kxk) an isotropic positive definite function in R whenever λ ≥ 2 , where a is a positive constant and [x] denotes the greatest integer less than or equal to x.

For two points x and y on Sd, their spherical (great circle, angular, or geodesic) distance,

denotes by ϑ(x, y), is the distance between x and y on the largest circle on Sd that passes

1 through them. More precisely,

0 d ϑ(x, y) = arccos(x y), x, y ∈ S , (1.0.3) where x0y denotes the inner product between x and y. Since

kx − yk2 = kxk2 + kyk2 − 2x0y = 2 − 2x0y,

another expression for ϑ(x, y) in terms of the Euclidean distance kx − yk is

 1  ϑ(x, y) = arccos 1 − kx − yk2 , x, y ∈ d. 2 S

Evidently, 0 ≤ ϑ(x, y) ≤ π.

A function g(ϑ(x1, x2)), or g(ϑ) for simplicity, is called an isotropic positive definite

function on Sd if n n X X aiajg(ϑ(xi, xj)) ≥ 0 i=1 j=1 d holds for every n ∈ N, any ai ∈ R and xi ∈ S , i = 1, . . . , n. Thus, an isotropic positive function on a sphere depends on its arguments via the spher-

0 ical distance ϑ(x1, x2),or equivalently, via the inner product x1x2.

For a continuous function g(ϑ) on [0, π], it is positive definite on Sd if and only if it can be expressed [22] as

∞ d−1 X ( 2 ) g(ϑ) = bnPn (cos ϑ), ϑ ∈ [0, π], (1.0.4) n=0

∞ d−1 P ( 2 ) (ν) where {bn, n ∈ N0} is a sequence of nonnegative constants, bnPn (1) converges, Pn (x) n=0 denotes the Gagenbauer or ultraspherical polynomial [25] of degree n, and N0 is the set of nonnegative integers. In other words, g(ϑ) is positive definite on Sd if and only if

Z π d−1 ( 2 ) d−1 bn = g(ϑ)Pn (cos ϑ) sin ϑdϑ ≥ 0, n ∈ N0, (1.0.5) 0

∞ d−1 ∞ d−1 P ( 2 ) P ( 2 ) and bnPn (1) converges. For d ≥ 2, the convergence of bnPn (1) is equivalent n=0 n=0 ∞ P d−2 to that of n bn [6]. n=0

2 A question of interest is: if a function g(x) on [0, ∞) makes g(kxk) an isotropic positive

definite function in Rd, would it also make g(ϑ) an isotropic positive definite function on

Sd? Such a question may be answered on a case-by-case basis. A counterexample is g(x) = exp(−x2), x ≥ 0, it is known that g(kxk) is positive definite in all dimensions, but g(ϑ) is

not positive definite even on S1. On the other hand, a stimulating example is (1.0.2), for which it is shown in [4] that g(ϑ) is an isotropic positive definite function on Sd if d ≤ 7,

but, actually, the restriction d ≤ 7 can be removed [28]. More interestingly, for α, β ∈ N0, and max(α, β) > 0, Xu [28] proves that

Z a 1 1  2α  2β (α− 2 , β− 2 ) ϑ ϑ g(ϑ)Pn (cos ϑ) sin cos dϑ > 0, n ∈ N, 0 2 2

(α,β) for every a ∈ (0, π] and λ ≥ α + 1, where Pn (x) is the Jacobi polynomial [25] of degree n.

(α,β) When α = β, Pn (x) reduces to the ultraspherical polynomial, with

1  Γ α + Γ(n + 2α) (α− 1 ,α− 1 ) 1 P (α)(x) = 2 P 2 2 (x), n ∈ , α > − ; (1.0.6) n 1  n N0 2 Γ(2α)Γ n + α + 2 see, e.g., (4.7.1) of [25].

A simple but important feature of the function (1.0.2) is that it is of compact support

and takes nonnegative values. This motivates us to consider the following question: for a

compactly supported function g(x) on [0, ∞), if it makes g(kxk) an isotropic positive definite

function in Rd, will it also make g(ϑ) an isotropic positive definite function on Sd? The answer is yes, in a particualr case d = 1 or 3, but, as is conjectured in [15], it would be of interest to

see whether such a result holds for a general d, for which a difficulty arises when one deals

with the connection between the two bases, the Bessel functions for Rd and ultraspherical

polynomials for Sd. A useful bridge for the connection employed in [28] is Theorem 8.1 of [25], which is generalized in Theorem 5.1.3 below for us to answer the above question.

In this thesis we develop isotropic positive definite functions on odd dimensional spheres from those in Euclidean spaces. The rest of the thesis is organized as follows. We state the background information in chapter 2. The characterizations of isotropic positive definite functions on Rd were reviewed in chapter 3. We discuss the characterizations of positive

3 definite functions on Sd in chapter 4. We introduce and prove our main result in chapter 5. There is an open question conjectured in chapter 6 for further research.

4 CHAPTER 2

BACKGROUND INFORMATION

2.1 Definitions

In this expository section, we list the definitions of isotropic functions in Rd and on

Sd, and those of positive definite functions in Rd and on Sd, as well as their connections to positive definite matrices.

Definition 2.1.1. (i) A function G : Rd × Rd → R is said to be isotropic, radial or spherical

symmetric if there exists a function g : [0, ∞) → R such that

G(x, y) = g(kx − yk)

for all x, y ∈ Rd.

(ii) A function G : Rd → R is said to be isotropic, radial or spherical symmetric if there

exists a function g : [0, ∞) → R such that

G(x) = g(kxk)

for all x ∈ Rd.

Definition 2.1.2. A complex-valued function g(x, y) of two arbitrary points of Rd is called positive definite, if it enjoys the following properties: (i) Hermitian symmetry

g(x − y) = g(x − y).

d (ii) For any finite set of xj ∈ R and complex numbers zj (with conjugate complexz ¯j), 1 ≤ j ≤ n, we have n n X X g(xj − xk)zjz¯k ≥ 0. (2.1.1) j=1 k=1 If the quadratic form (2.1.1) is zero only for z ≡ 0 , then g is called strictly positive definite.

5 Here, we introduce a general definition for complex-valued positive definite functions.

The reason for this is that it allows us to use techniques such as Fourier transforms more naturally in Chapter 3. However, we will see that for even, real-valued functions it suffices

to investigate the quadratic form only for real numbers z ∈ R.

Definition 2.1.3. A d×d symmetric real matrix M is said to be positive definite if x0Mx ≥ 0 for all x ∈ Rd.

Strictly positive definite matrices are defined similarly, except that the above scalar x0Mx is positive for every x 6= 0.

Now, we define positive definite functions by using positive definite matrices.

Definition 2.1.4. A function G : Rd × Rd → R is said to be positive definite if for any

d N ∈ N and any finite set of points {x1, ··· , xN } in R , the N × N matrix

N G[XN ] = [G(xi, xj)]i,j=1

is positive definite. It is said to be strictly positive definite if the matrix G[XN ] is strictly positive definite.

Definition 2.1.4 is equivalent to definition 2.1.5 (i).

Definition 2.1.5. (i) A function G : Rd × Rd → R is said to be positive definite, if the inequality n n X X aiajG(xi, xj) ≥ 0 i=1 j=1 d holds for every n ∈ N, any ai ∈ R, and any xi ∈ R (i = 1, ··· , n).

(ii) An isotropic function g(x): Rd → R, is said to be positive definite, if the inequality

n n X X aiajg(kxi − xjk) ≥ 0 i=1 j=1

d holds for every n ∈ N, any ai ∈ R and xi ∈ R (i = 1, . . . , n).

6 Definition 2.1.6. A function G : Sd × Sd → R is said to be isotropic, if there exists a

function g : [0, π] → R such that G(x, y) = g(ϑ(x, y)) for all x, y ∈ Sd, where ϑ(x, y) is the spherical distance between x and y defined in (1.0.3).

Definition 2.1.7. (i) A function G : Sd × Sd → R is said to be positive definite on Sd, if the inequality n n X X aiajG(xi, xj) ≥ 0 i=1 j=1 d holds for every n ∈ N, any ai ∈ R and xi ∈ S (i = 1, ··· , n).

(ii) A function g(ϑ(x1, x2)), or g(ϑ) for simplicity, is called an isotropic positive definite

function on Sd if n n X X aiajg(ϑ(xi, xj)) ≥ 0 i=1 j=1 d holds for every n ∈ N, any ai ∈ R and xi ∈ S , i = 1, . . . , n.

2.2 The Gamma Function

The Euler integral of the second kind Z ∞ Γ(z) = xz−1e−xdx 0 defines the Gamma function Γ(z) for all complex numbers which have the positive real part.

In particular, Γ(1) = 1. Using integration by parts, we get the recurrence formula Z ∞ z−1 −x ∞ z−2 −x Γ(z) = −[x e ]0 + (z − 1) x e dx = (z − 1)Γ(z − 1). (2.2.1) 0 If we restrict the domain on the set of positive integers N, then we get

Γ(n) = (n − 1)!

Now, we list the following lemma [19] as a reference.

Lemma 2.2.1. If n → ∞, then Γ(n + k) ∼ nk, k ∈ , (2.2.2) Γ(n) R where the symbol ∼ means that the ratio of the two sides converges to 1 or asymptotically converges.

7 2.3 Orthogonal Polynomial

(α) Gegenbauer polynomials or ultraspherical polynomials, denoted by Pn (x), are orthog-

2 α− 1 onal polynomials on the interval [−1, 1] with respect to the weight function (1 − x ) 2 .

They are special cases of , occasionally called hypergeometric polynomi-

(α,β) als, denoted by Pn (x), which are a class of classical . They are orthogonal with respect to the weight (1 − x)α(1 − x)β on the interval [−1, 1].

(α) The Gegenbauer polynomials Pn (x) of degree n can be defined in terms of their gener- ating function [9]:

∞ 1 X = P (α)(x)tn, t ∈ (−1, 1). (2.3.1) (1 − 2xt + t2)α n n=0

(α) The function Pn (x) is even if n is even and odd if n is odd. For α = 0, we follow [22] and set

(0) Pn (cos ϑ) = cos nϑ, n ∈ N0, ϑ ∈ [0, π].

The Gegenbauer polynomials satisfy the [24]:

α P(0)(x) = 1,

α P(1)(x) = 2αx, 1 P α (x) = [2x(n + α − 1)P α (x) − (n + 2α − 2)P α (x)], n ∈ . (n) n n−1 n−2 N0

The Jacobi polynomials are defined as follows:

(α + 1) P (α,β)(x) = n F −n, 1 + α + β + n; α + 1; 1 (1 − x) , n n! 2 1 2

where ∞ n X (a)n(b)n z F (a, b; c; z) = 2 1 (c) n! n=0 n

is the which is defined for |z| < 1 by the power series and (α + 1)n denotes the Pochhammer’s symbol (for the rising factorial) ,i.e.,

(α + 1)n = (α + 1)(α + 2), ··· , (α + n + 1).

8 In this case, the series for the hypergeometric function is finite, and an explicit expression

(α,β) [25] of Pn (x) is

n m X Γ(α + n + 1)  n Γ(α + β + n + m + 1) x − 1 P (α,β)(x) = , n ∈ N . n n! Γ(α + β + n + 1) m Γ(α + m + 1) 2 0 m=0 For reference later on, we note the identities of Jacobi polynomials

n + α P (α,β)(1) = , n n

(α,β) P0 (x) = 1.

(α,β) If α = β, the Jacobi polynomial Pn (x) reduces to the Gegenbauer polynomial, with

1  (α− 1 ,α− 1 ) Γ(2α)Γ n + α + 1 P 2 2 (x) = 2 P (α)(x), n ∈ , α > − . n 1  n N0 2 Γ α + 2 Γ(n + 2α) 2.4 Bessel Functions

The Bessel function of the first kind of order α, denoted as Jα(z), is defined by its series expansion around z = 0, ∞ X (−1)ν(z/2)α+2ν J (z) = . α ν!Γ(ν + α + 1) ν where Γ(x) is the Gamma function, α is arbitrary real.

Bessel functions are solutions of Bessel’s differential equation

d2y dy x2 + x + x2 − α2 y = 0. dx2 dx

1 If α = − 2 , 2 1 2 J− 1 (z) = ( ) cos z. 2 πz 1 If α = 2 , 2 1 J 1 (z) = ( ) 2 sin z. 2 πz Bessel functions have the following asymptotic forms, when α is fixed and z → 0,

Jα(z) → 1,

z 1 ( ) 2 J (z) ∼ 2 . α Γ(α + 1)

9 The integral representation

z 1 1 ( ) 2 Z 1 2 2 (α− 2 ) izx Jα(z) = 1 1 (1 − x ) e dx Γ(α + 2 )Γ( 2 ) −1

1 holds for α > − 2 .

2.5 Abel-Summability

In order to use the Abel-summability in the proof of the convergence of the series

d−1 P∞ 2 n=0 bnPn (1) in Chapter 4 and Chapter 5, we state Abel-summability and Abel theo- rem here.

Abel-summability is a generalized convergence criterion for power series. It extends the usual definition of the sum of a series, and gives a way of summing up certain divergent P∞ series. Let us start with a series n=0 cn, convergent or not, and use that series to define P∞ n a power series f(x) = n=0 cnx . Note that for |x| < 1 the summability of f(x) is easier to achieve than the summability of the original series. Starting with this observation we P∞ say that the series n=0 cn is Abel-summable if the defining series f(x) is convergent for all |x| < 1, and if f(x) converges to some limit s as x → 1−. If this is so, we shall say that P∞ n=0 cn Abel converges to s. In other words, If

∞ X n lim cnx = s, x→1− n=0 then ∞ X cn = s. n=0 .

Of course it is important to ask whether an ordinary convergent series is also Abe- summable, and whether it converges to the same limit? This is true, and the result is known as Abel’s limit theorem, or simply as Abel’s theorem [3].

Theorem 2.5.1 (Abel’s Theorem). Let

∞ X n f(x) = cnx n=0

10 be a power series with real coefficients cn and radius of convergence 1. Suppose that the series ∞ X cn n=0 is convergent. Then f(x) is continuous from the left at x = 1, i.e.

∞ ∞ X n X lim cnx = cn. x→1− n=0 n=0

11 CHAPTER 3

ISOTROPIC POSITIVE DEFINITE FUNCTIONS IN EUCLIDEAN SPACE

This chapter reviews some properties of isotropic definite functions in Euclidean space

Rd. The main result is the first inequality in the abstract which is the key to prove the positive definiteness of an isotropic function on spheres Sd. 3.1 Integral Transforms

Before getting into the discussion of the characterizations of positive definite functions we summarize some formulas for various integral transforms to be used later.

We start with the conventions of the Fourier transform, for which a reference book is

Bochner and Chandrasekhara [5].

Definition 3.1.1. The Fourier transform of f ∈ L1(Rd) is given by Z iu0x d gˆ(u) = g(x)e dx, u ∈ R , Rd and its inverse Fourier transform is given by Z 1 −ix0u d gˇ(x) = d g(u)e du, x ∈ R (2π) Rd where x0u is the inner product between x and u.

In this thesis, we will focus on isotropic positive definite functions. Note that the Fourier transform of an isotropic function is again isotropic [26]. Indeed, the Fourier transform of an isotropic function is its own inverse.

d Theorem 3.1.2. Let G ∈ L1(R ) be continuous and isotropic, i.e, G(x) = g(kxk). Then its ˆ ˆ Fourier transform is also isotropic, i.e, G is also isotropic, i.e, G(x) = Fdg(kxk) with

Z ∞ d − d−2 d Fdg(r) = (2π) 2 x 2 g(u)u 2 J d−2 (xu)du, (3.1.1) 0 2

d−2 where r = kxk and J d−2 (x) is the classical Bessel function of the first kind of order . 2 2

12 Remark 1: The integral transform in (3.1.1) is also referred to as a Fourier-Bessel transform or Hankel transform.

Remark 2: The Hankel inversion theorem [23] ensures that the Fourier transform for an isotropic function is its own inverse, i.e., for an isotropic function g,

Fd[Fdg] = g. (3.1.2)

3.2 Characterizations of Positive Definite Functions

One of the most celebrated results on positive definite function in Euclidean space Rd is Bochner’s integral characterization, which is established by Bochner in 1932 for d = 1 and

1933 for d > 1.

The following theorem established the identity of positive definite functions in Rd with characteristic functions of probability measures in Rd.

Theorem 3.2.1 (Bochner’s Theorem). A complex-valued function g ∈ C(Rd) is positive

definite in Rd if and only if it is the Fourier transform of a finite nonnegative Borel measure

ν on Rd, i.e., Z ix0y d g(x) = e ν(dy), x ∈ R . Rd Remark: This is the Bochner’s integral characterization for the positive definite func-

tions in Rd. One of the proof is given by Wendland [27]. The proof for d = 1 can be found in the probability textbook [7].

From Bochner’s Theorem, an isotropic positive definite function g in Rd is the Fourier

transform of a finite nonnegative Borel measure ν on Rd. Then the inverse Fourier transform

of an isotropic positive definite function g is a finite nonnegative Borel measure ν on Rd, so it is nonnegative. From (3.1.2), the inverse Fourier transform for an isotropic function is its Fourier transform. Therefore, the inverse Fourier transform or Fourier transform of an isotropic positive definite function is nonnegative. The Fourier transform of an isotopic

13 positive definite function is given by (3.1.1). Now we obtain,

Z ∞ d − d−2 d (2π) 2 x 2 g(u)u 2 J d−2 (xu)du ≥ 0, x ≥ 0, 0 2

or, simplify, ∞ Z d g(u)u 2 J d−2 (xu)du ≥ 0, x ≥ 0. (3.2.1) 0 2 3.3 Schoenberg’s Characterization

A well-known result on isotropic positive definite functions in Euclidean space Rd is Schoenberg’s characterization [21] established in 1938. It is stated as the following theorem

[10].

Theorem 3.3.1. A continuous function g(kxk): Rd → R is isotropic positive definite if and only if g(x) is the Bessel transform of a finite nonnegative Borel measure ν on [0, ∞), i.e.,

Z ∞ g(x) = Ωd(xt)dν(t) 0

where ( cos x d = 1 Ωd(x) = d 2 d−2 Γ( )( ) 2 J d−2 (x) d ≥ 2 2 r 2 is an integral transcendental function and J d−2 (x) is the Bessel function of the first kind of 2 d−2 order 2 .

If a function is isotropic positive definite in Rd, then it is also isotropic positive definite

in Rd−1. Therefore, we are interested in those functions which are isotropic positive definite

in Rd for all d ∈ N. Those functions are identified by the following theorem.

Theorem 3.3.2. A continuous function g : [0, ∞) → R makes g(kxk) an isotropic positive

definite function in all Rd if and only if it is of the form

Z ∞ g(x) = e−x2t2 dν(t), 0

where ν is a finite nonnegative Borel measure on [0, ∞).

14 3.4 Properties of Positive Definite Functions

In order to derive the properties of positive definite functions easily, we have expanded

the range of positive definite functions on the set of complex numbers in Definition 2.1.2.

The reader should note that we call a function positive definite if the left side of (2.1.1) is

nonnegative and strictly positive definite if the left side of (2.1.1) is positive for z 6= 0 . This seems to be natural. However, there is an alternative terminology around in the literature: some authors call a function positive semi-definite if the left side of (2.1.1) is nonnegative and positive definite if if the left side of (2.1.1) is positive for z 6= 0. We do not follow this approach here, but the reader should always keep this in mind when looking at other texts.

From Definition 2.1.2, we can read off the elementary properties [27] of such a positive definite function.

Theorem 3.4.1. Suppose g is a complex-valued positive definite function in Rd. Then the following properties are satisfied.

(a) g(0) ≥ 0.

(b) g(x) = g(−x) for all x ∈ Rd.

(c) |g(x)| ≤ g(0) for all x ∈ Rd, i.e. g(x) is bounded (d) g(0) = 0 if and only if g ≡ 0.

(e) For d = 1, g is continuous at x = 0 implies that it is uniformly continuous on R.

(f) For d = 1, g is uniformly continuous in R, we have for every continuous complex- valued function ζ on R and every T > 0:

Z T Z T g(s − t)ζ(s)ζ(t)dsdt ≥ 0. (3.4.1) 0 0

Pm (g) If c1, ··· , cm are nonnegative, and g1, ··· , gm are positive definite, then G := l=1 clgl is also positive definite. If one of gl is strictly positive definite and the corresponding cl is positive then G also is strictly positive definite.

(h) For any a ∈ R, f(x) = g(x)eiax is also positive definite. Generally, the product of two positive definite functions is positive.

15 d (i) If {gn(x)} be a sequence of positive definite in R , gn(x) → g(x) as n → ∞, and g(x)

is continuous, then g(x) is also a positive definite function in Rd.

Proof. (a) Assume that n = 1 and z1 = 1. By (2.1.1), we obtain

g(0) ≥ 0.

(b) Assume that n = 2. By (2.1.1), we have

2 2 g(0)|z1| + g(x1 − x2)z1z2 + g(x2 − x1)z2z1 + g(0)|z2| ≥ 0 (3.4.2)

Taking x1 = 0, x2 = x, z1 = z2 = i in (3.4.2), we have

2g(0) + g(−x) + g(x) ≥ 0 which implies g(−x) + g(x) is real.

Taking x1 = 0, x2 = x, z1 = i, z2 = 1 in (3.4.2), we have

2g(0) + ig(−x) − ig(x) ≥ 0 which implies g(−x) − g(x) is pure imaginary. Hence, we have

g(x) = g(−x)

(c) Taking x1 = 0, x2 = x, z1 = g(x), z2 = −|g(x)| in (3.4.2), we have

g(0)|g(x)|2 + g(−x)g(x)(−|g(x)|) + g(x)(−|g(x)|)g(x) + g(0)|g(x)|2 ≥ 0

By (b), we have

g(0)|g(x)|2 + g(x)g(x)(−|g(x)|) + g(x)(−|g(x)|)g(x) + g(0)|g(x)|2

= 2g(0)|g(x)|2 + 2g(−x)g(x)(−|g(x)|)

= 2|g(x)|2(g(0) − |g(x)|) ≥ 0

Therefore,

g(0) ≥ |g(x)|.

16 (d) For the “ only if ” part , g(0) = 0 implies |g(x)| = 0 for all x ∈ Rd by (c). The converse is obvious.

n (e) From (2.1.1) and (b), we have that [g(xj − xk)]j,k=1 is a Hermitian positive definite

matrix. In particular, for n = 3, x1 = t, x2 = s and t3 = 0, the matrix   g(0) g(t − s) g(t)     g(t − s) g(0) g(s)   g(t) g(s) g(0) must have a nonnegative determinant, which gives

0 ≤ g(0)3 + g(s)g(t − s)g(t) + g(t)g(s)g(t − s) − g(0)(|g(t)|2 + |g(s)|2 + |g(t − s)|2)

= g(0)3 + 2Re[g(s)g(t − s)g(t)] − g(0)(|g(t) − g(s)|2 + 2Re[g(s)g(t)] + |g(t − s)|2)

= g(0)3 + 2Re[g(s)g(t)(g(t − s) − g(0)] − g(0)(|g(t) − g(s)|2 + |g(t − s)|2)

≤ g(0)3 + 2g(0)2|g(0) − g(t − s)| − g(0)(|g(t) − g(s)|2 + |g(t − s)|2) where in the last inequality we used (c). After simplifying and adding |g(t) − g(s)|2 to both

sides of the inequality we have

|g(t) − g(s)|2 ≤ g(0)2 − |g(t − s)|2 + 2g(0)|g(0) − g(t − s)|

= (g(0) − |g(t − s)|)(g(0) + |g(t − s)|) + 2g(0)|g(0) − g(t − s)|

≤ (g(0) − |g(t − s)|)(g(0) + g(0)) + 2g(0)|g(0) − g(t − s)|

≤ 4g(0)|g(0) − |g(t − s)|

Thus the “ modulus of continuity” at each t is bounded by twice the square root of the that at 0, and so the continuity at 0 implies uniform continuity everywhere on R. (f) The integrand of the double integral in (3.4.1) is continuous and the integral is the limit of the Riemann sums, hence it is nonnegative by (2.1.1), i.e.

Z T Z T g(s − t)ζ(s)ζ(t) ≥ 0 0 0

17 (g) Since G := clgl, we have

n n X X G(xj − xk)zjz¯k j=1 k=1 n n m X X X = ( clgl(xj − xk))zjz¯k j=1 k=1 l=1 m n n X X X = clgl(xj − xk)zjz¯k l=1 j=1 k=1 m n n X X X = cl( gl(xj − xk)zjz¯k) l=1 j=1 k=1 ≥ 0

where in the last line we used cl ≥ 0 and the positive definiteness of gl for 1 = 1, 2, ··· , m.

If one of gl is strictly positive definite and the corresponding cl is positive, without loss of generality, let l = p (0 ≤ p ≤ m), then we have

n n X X cp gp(xj − xk)zjz¯k) > 0, j=1 k=1 hence, m n n X X X cl( gl(xj − xk)zjz¯k) > 0. l=1 j=1 k=1 By Definition 2.1.2, G is strictly positive definite.

(h) Changing g to f in (2.1.1), we have

n n X X f(xj − xk)zjzk j=1 k=1 n n X X ia(xj −xk) = g(xj − xk)e zjzk j=1 k=1 m n X X iaxj −iaxj = g(xj − xk)e e zjzk l=1 j=1 m n X X iaxj iax = g(xj − xk)e e j zjzk l=1 j=1 m n X X iaxj iax = g(xj − xk)e zje j zk l=1 j=1 ≥ 0

18 where the inequality is because of the positive definiteness of g. By the definition, f(x) =

g(x)eiax is also positive definite. The proof for the general case can be found in [27].

d (i) Since {gn(x)} be a sequence of positive definite in R , gn(x) → g(x) as n → ∞, and g(x) is continuous, we have

n n X X g(xj − xk)zjzk j=1 k=1 n n X X = lim gi(xj − xk)zjzk i→∞ j=1 k=1 n n X X = lim gi(xj − xk)zjzk i→∞ j=1 k=1 ≥ 0.

By Definition 2.1.2, g(x) is also positive definite in Rd.

19 CHAPTER 4

ISOTROPIC POSITIVE DEFINITE FUNCTIONS ON SPHERES

In this chapter, we proceed to review Schogenberg’s characterizations and constructions

of isotropic positive definite functions on Sd in terms of Gegenbauer polynomials, for which we require some classical results concerning Gegenbauer polynomials.

4.1 Some Classical Results Concerning Ultraspherical Polynomials

We introduce a representation of a real continuous function by Gegenbauer polynomials

[22].

Lemma 4.1.1. A real continuous function G(x) defined on Sd can be approximated by ∞ X n + α Z G(x) ∼ G(y)P (α)(cos θ)dω . (4.1.1) αω n y n=0 d Sd

d+1 d−1 2π 2 d where α = 2 , ωd = d+1 is the surface area of S and θ is the spherical distance between Γ( 2 ) x and y.

d Proof. Suppose that G(x) is a real continuous function defined on S . Let xr be a point on the radius ox such that oxr = r, 0 ≤ r < 1. Define G(xr) as a Poisson integral

1 Z 1 − r2 G(xr) = d+1 G(y)dωy. (4.1.2) ω 2 2 d Sd (1 − 2r cos θ + r ) By Proposition 2.3.2 in [14], we have that

lim G(xr) = G(x). (4.1.3) r→1

Let t = r and x = cos θ in (2.3.1), we obtain

∞ 1 X = rnP (α)(cos θ). (4.1.4) (1 − 2r cos θ + r2)α n n=0 Differentiating both sides of (4.1.5) with respect to r, we get

∞ −2 cos θ + 2r X = nrn−1P (α)(cos θ). (4.1.5) (1 − 2r cos θ + r2)α+1 n n=0

20 r Multiplying (4.1.5) by α , we have ∞ −2r cos θ + 2r2 1 X = nrnP (α)(cos θ). (4.1.6) α(1 − 2r cos θ + r2)α+1 α n n=0 Adding (4.1.4) and (4.1.6), we now have ∞ 1 − r2 X n = ( + 1)rnP (α)(cos θ). (4.1.7) (1 − 2r cos θ + r2)α+1 α n n=0 Multiplying (4.1.7) by G(y) dωy and integrating both sides over d, we have ωd S Z 2 Z ∞ 1 1 − r X n + α n (α) d+1 G(y)dωy = r G(y)Pn (cos θ)dωy. (4.1.8) ω 2 2 αω d Sd (1 − 2r cos θ + r ) Sd n=0 d As r → 1 in (4.1.8), we develop an approximation of G(x) by ∞ X n + α Z G(x) ∼ G(y)P (α)(cos θ)dω . (4.1.9) αω n y n=0 d Sd

The relation (4.1.3) implies that the approximation (4.1.9) is Abel-summable at any

point x on Sd to the sum G(x).

d+1 We may convert the Cartesian coordinates of x = (x, x1, ··· , xd) ∈ R to spherical coordinates with:

x = cos ϑ,

x1 = sin ϑ cos ϑ1,

x2 = sin ϑ sin ϑ1 cos ϑ2, . .

xd−2 = sin ϑ sin ϑ1 ··· cos ϑd−2,

xd−1 = sin ϑ sin ϑ1 ··· sin ϑd−2 cos φ,

xd = sin ϑ sin ϑ1 ··· sin ϑd−2 sin φ, where the angles ϑ, ϑ1, ··· , ϑd−2 ∈ [0, π] , and φ ∈ [0, 2π).

d−1 Let G(x) = g(ϑ) and α = 2 . The integral in (4.1.9) can be written Z π Z π Z 2π 0 (α) d−1 0 d−2 0 0 0 0 0 An = ··· g(ϑ )Pn (cos θ) sin ϑ sin ϑ1 ··· sin ϑd−2dϑ ··· dϑd−2dφ . (4.1.10) 0 0 0

21 0 0 0 0 0 0 0 where y = (x , x1, ··· , xd) in terms of Euclidean coordinate and y = (ϑ , ϑ1, ··· , ϑd−2, φ ) in

0 (α) d−1 0 terms of spherical coordinates. Since the integrand g(ϑ )Pn (cos θ) sin ϑ only involves in

0 0 0 0 ϑ , we may integrate first with the variable ϑ1, ··· , ϑd−2, φ as follows: Z π 0 d−1 0 0 An = g(ϑ )Cn sin ϑ dϑ . (4.1.11) 0 where

Z π Z π Z 2π (α) d−2 0 0 0 0 0 Cn = ··· Pn (cos θ) sin ϑ1 ··· sin ϑd−2dϑ ··· dϑd−2dφ . (4.1.12) 0 0 0

This integral can be reduced as follows. For two points p, q both lie in the equator Sd−1 π defined by ϑ = 2 with the polar coordinate: π π p = ( , ϑ , ··· , ϑ , φ), q = ( , ϑ0 , ··· , ϑ0 , φ0). 2 1 d−2 2 1 d−2

Then we get

0 0 0 0 0 0 cos θ = x y = xx + x1x1 + ··· + xdxd = cos ϑ cos ϑ + sin ϑ sin ϑ cos ψ, (4.1.13) where ψ is the spherical distance between p and q.

(α) d d Now, Cn is an integration of Pn (cos θ) over S . If we take in S a different polar coordinates (ψ, ψ1, ··· , ψd−3, ϕ) of pole p, we obtain

Z π Z π Z 2π (α) d−2 d−3 Cn = ··· Pn (cos θ) sin ψ sin ψ1 ··· sin ψd−3dψdψ1 ··· dψd−3dϕ 0 0 0 Since cos θ depends only on ψ by (4.1.13), we have

Z π (α) d−2 Cn = ωd−2 Pn (cos θ) sin ψdψ 0 which can be explicitly computed by

Z π 1 (α) 2α−1 Γ(α)Γ( 2 )Γ(n + 1)Γ(2α) α α 0 Pn (cos θ) sin ψdψ = 1 Pn (cos ϑ)Pn (cos ϑ ) 0 Γ(α + 2 )Γ(n + 2α)

Substituting Cn into (4.1.11), the expression (4.1.9) reduces to

∞ X (n + α)Γ(α)Γ(n + 1)Γ(2α) Z π g(ϑ) ∼ P α(cos ϑ) P α(cos ϑ0)g(ϑ0) sind−1 ϑ0dϑ0 (4.1.14) Γ(α + 1 )Γ( 1 )Γ(n + 2α) n n n=o 2 2 0

22 We note that this expansion is also Abel-summable.

We need the following two lemmas due to Schoenberg to prove the Schoenberg’s charac-

terization.

Lemma 4.1.2. If g(x) is a real continuous function on [−1, 1] such that g(θ) is positive

definite on Sd, then we have Z g(θ)dωy ≥ 0, (4.1.15) Sd where θ is the spherical distance between x and y.

Lemma 4.1.3. The Gegenbauer polynomials

d−1 2 Pn (cos ϑ), n ∈ N0, ϑ ∈ [0, π]

are all positive definite on Sd.

4.2 Schoenberg’s Characterization

Theorem 4.2.1 (Schoenberg’s Characterization). A real continuous function g(x) on [0, π]

is positive definite on Sd if and only if g(ϑ) can be expressed as the series in (4.1.14) with d−1 2 the series converges for ϑ = 0 and the coefficients of Pn (cos ϑ) are nonnegative, i.e.

∞ d−1 d−1 π (n + )Γ( )Γ(n + 1)Γ(d − 1) Z d−1 d−1 X 2 2 2 d−1 2 Pn (cos ϑ)g(ϑ) sin ϑdϑPn (1) Γ( d )Γ( 1 )Γ(n + d − 1) n=0 2 2 0 converges, and

d−1 d−1 π (n + )Γ( )Γ(n + 1)Γ(d − 1) Z d−1 2 2 2 d−1 d 1 Pn (cos ϑ)g(ϑ) sin ϑdϑ ≥ 0, Γ( 2 )Γ( 2 )Γ(n + d − 1) 0 which is equivalent to

Z π d−1 ( 2 ) d−1 bn = g(ϑ)Pn (cos ϑ) sin ϑdϑ, n ∈ N0, (4.2.1) 0

are nonnegative. In other words, g(ϑ) is positive definite on Sd if and only if ∞ X α g(ϑ) = bnPn (cos ϑ) (4.2.2) n=0 with bn ≥ 0 and the series converges for ϑ = 0.

23 α Proof. We can rewrite the integral coefficient of Pn (cos ϑ) in (4.1.14) as Z π 1 Z D = P α(cos ϑ0)g(ϑ0) sind−1 ϑ0dϑ0 = P α(cos θ0)g(θ0)dω n n ω n y 0 d−1 Sd where θ0 is the spherical distance between a = (1, 0, ··· , 0) and y. Note that g(θ0) is positive

d α 0 d definite in S , and Pn (cos θ ) is also positive definite on S , so is their product. By lemma

(4.1.2), Dn is nonnegative, so are all coefficients of (4.1.14). Thus, (4.1.14) may be rewritten as ∞ X α g(ϑ) ∼ bnPn (cos ϑ), (4.2.3) n=0 where bn is the same as in (4.2.1). On the other hand, we know this series is Abel-summable for all ϑ, hence, in particular, for ϑ = 0. Thus k k ∞ X α X (α) X n (α) bn |Pn (cos ϑ)| ≤ bnPn (1) ≤ lim bnr Pn (1) = g(1). r→1− n=0 n=0 n=0 This shows that the series (4.2.3) is absolutely and uniformly convergent for all ϑ, hence convergent to its Abel-sum which is g(ϑ).

The other side of prove is by the property i in theorem 3.4.1. Note that g(ϑ) is continuous because the series (4.2.2) must be convergent uniformly. As g(ϑ) is the continuous limit of a sequence of positive definite functions, it is also positive definite.

4.3 Recent Results

In [15], it has been proved a compactly supported continuous function is positive definite

on Sd, if the function is positive definite in Rd, where d = 1, 3. In [28], Xu showed that the integral of Jacobi polynomials defined as

Z a 1 1  2α  2β (α,β),δ δ (α− 2 ,β− 2 ) ϑ ϑ Fn = (a − ϑ) Pn (cos ϑ) sin cos dϑ 0 2 2 is nonnegative. It can be used to prove the truncated power function  δ δ  (a − ϑ) , 0 ≤ ϑ ≤ a, g(x) = (a − ϑ)+ = (4.3.1)  0, ϑ > a,

24 d−1 d 1 is positive definite on S for δ > 2 + 2 .

25 CHAPTER 5

ISOTROPIC POSITIVE DEFINITE FUNCTIONS ON SPHERES GENERATED FROM THOSE IN EUCLIDEAN SPACES

5.1 Main Results

One of the main results in this thesis is the following theorem, which provides an approach

for generating isotropic positive definite functions on the sphere from those on the Euclidean

space.

Theorem 5.1.1. Suppose that g(x) is a continuous function on [0, ∞), and g(x) = 0, x ≥ π.

For an odd integer d, if g(kxk) is a positive definite function in Rd, then g(ϑ), ϑ ∈ [0, π], is

a positive definite function on Sd.

d Note: The variables x in 5.1.1 can be any point in R , but we should have kx1 − x2k ∈

[0, π] in order to keep g(kxk) is positive definite in Rd. For the function g(x) = cos x, which

is isotropic positive definite in R and on S1, it is not satisfies the conditions of 5.1.1, since g(π) = −1.

It is not clear whether a similar result holds for an even integer d. Nevertheless, the next

corollary is a consequence of Theorem 5.1.1, since an isotropic positive definite function in

Rd is also an isotropic positive definite function in Rd−1 (d ≥ 2), and d − 1 is odd for an even d.

Corollary 1. Let g(x) be as in Theorem 5.1.1. For an even integer d, if g(kxk) is a positive

definite function in Rd, then g(ϑ), ϑ ∈ [0, π], is a positive definite function on Sd−1. Although the requirement g(x) = 0, x ≥ π, is crucial in Theorem 5.1.1, it is always

possible to change the scale for a compactly supported function. This result is in the following corollary. Corollary 2. Let g(x) be a continuous function on [0, ∞), and g(x) = 0, x ≥ l. For an odd

d l  integer d, if g(kxk) is a positive definite function in R , then g π ϑ , ϑ ∈ [0, π], is a positive definite function on Sd.

26 Since g(x) is continuous with compact support on [0, ∞), the positive definiteness of g(kxk) in Rd implies that

π Z d 2 u J d −1(xu)g(u)du ≥ 0, x ≥ 0, 0 2

Hence, Theorem 5.1.1 can be rephrased as the two inequality implication statement as fol- lows, in term of the Bessel function and the ultraspherical polynomial.

Theorem 5.1.2. For a nonnegative integer α and a continuous function g(x) on [0, π] with

g(π) = 0, if Z π α+ 1 u 2 J 1 (xu)g(u)du ≥ 0, x ≥ 0, (5.1.1) α− 2 0 then Z π (α) 2α bn = g(ϑ)Pn (cos ϑ) sin ϑdϑ ≥ 0, n ∈ N0, (5.1.2) 0 ∞ P (α) and bnPn (1) converges. n=0 Moreover, if inequality (5.1.1) is strict for every x ≥ 0, then so is inequality (5.1.2) for

every n ∈ N0.

The proof of Theorems 5.1.1 and 5.1.2 occupies Section 4, for which we need some lemmas in Section 3 and Theorem 5.1.3 that builds a useful bridge between the Jacobi polynomial and the Bessel function, as an extension of Theorem 8.1.1 of [25].

The base for an isotropic positive definite function in Rd is the Bessel function, as is released by the integral representation (1.0.1), and the base for that on Sd is the ultraspherical polynomial, by (1.0.4). The latter is a special case of the Jacobi polynomial. To prove Theorem 5.1.1 or 5.1.2 we need a bridge to connect these two types of special functions.

According to Theorem 8.1.1 of [25], the Bessel function is related to the Jacobi polynomial by 1 1 − α− 1 1 (α− ,β− )  z  z  ( 2 ) −(α− 2 ) 2 2 lim m Pm cos = Jα− 1 (z), z ∈ C, (5.1.3) m→∞ m 2 2

27 where the limit holds uniformly in every bounded region of the complex z-plane C. This relation is empolyed in [28] to deduce the positivity of the integrals of the Jacobi polynomial from that of the Bessel function. A generalized version of (5.1.3) is given in the next theorem,

where γm is assumed to be a rational number for ensuring that the degree, mγm + γ0, of a Jacobi polynomial is a positive integer.

1 1 Theorem 5.1.3. For α > − 2 , β > − 2 , and γ0 ∈ N0, if {γm, m ∈ N} is a sequence of

bounded positive rational numbers such that mγm ∈ and lim γm = γ > 0, then N m→∞

1 1 1 1 −(α− ) − α− (α− 2 ,β− 2 )  z  z  2 ( 2 ) 1 lim m Pmγm+γ0 cos = Jα− (zγ), z ∈ C. (5.1.4) m→∞ m 2 2

1 1 (α− 2 ,β− 2 ) Proof. For mγm +γ0 ∈ N, an explicit expression of Pmγm+γ0 (x) is given by formula (4.21.2) of [25], which reads,

1 1 (α− 2 ,β− 2 ) Pmγm+γ0 (x) mγm+γ0 1   ν X 1 Γ(mγm + γ0 + α + β + ν)Γ mγm + γ0 + α + x − 1 = 2 . ν!Γ ν + α + 1  Γ(mγ + γ + α + β)Γ(mγ + γ − ν + 1) 2 ν=0 2 m 0 m 0

z after substituting x by cos m ,

(α− 1 ,β− 1 ) z P 2 2 (cos ) mγm+γ0 m mγm+γ0 1   z ν X 1 Γ(mγm + γ0 + α + β + ν)Γ mγm + γ0 + α + cos − 1 = 2 m ν!Γ ν + α + 1  Γ(mγ + γ + α + β)Γ(mγ + γ − ν + 1) 2 ν=0 2 m 0 m 0 mγm+γ0 1  X 1 Γ(mγm + γ0 + α + β + ν)Γ mγm + γ0 + α +  z ν = 2 − sin2 ν!Γ ν + α + 1  Γ(mγ + γ + α + β)Γ(mγ + γ − ν + 1) 2m ν=0 2 m 0 m 0

Γ(n+k) k Keeping ν and z fixed, and letting m → ∞, by the asymptotic formula (2.2.2), Γ(n) ∼ n z z and sin 2m ∼ 2m , we obtain the following asymptotic expression for the (ν + 1)st term in the

28 above series,

1 Γ(mγ + γ + α + β + ν)Γ mγ + γ + α + 1   z ν m 0 m 0 2 − sin2 1  Γ(mγ + γ + α + β)Γ(mγ + γ − ν + 1) 2m ν!Γ ν + α + 2 m 0 m 0 1 Γ(mγ + γ + α + β + ν)Γ mγ + γ + α + 1   z 2ν ∼ 0 0 2 1  Γ(mγ + γ + α + β)Γ(mγ + γ − ν + 1) 2m ν!Γ ν + α + 2 0 0 ν ν+α− 1 (mγ) (mγ) 2  z 2ν ∼ 1  2m ν!Γ ν + α + 2 α− 1 (−1)ν(mγ) 2 zγ 2ν ∼ 1 , ν!Γ(ν + α + 2 ) 2 and

1 (α− 1 ,β− 1 )  z  −(α− 2 ) 2 2 lim m Pmγ +γ cos m→∞ m 0 m mγm+γ0 −(α− 1 ) 1  X m 2 Γ(mγm + γ0 + α + β + ν)Γ mγm + γ0 + α +  z ν = lim 2 − sin2 m→∞ ν!Γ ν + α + 1  Γ(mγ + γ + α + β)Γ(mγ + γ − ν + 1) 2m ν=0 2 m 0 m 0 ∞ ν α− 1 X (−1) γ 2 zγ 2ν = ν!Γ(ν + α + 1 ) 2 ν=0 2 − α− 1 z  ( 2 ) = Jα− 1 (zγ), 2 2

where passing to the limit under the summation sign is valid, due to the existence of a dominant for the total sum which is readily derived. Indeed if m is large enough, then we

have

1 1  −(α− 2 ) 2ν m Γ(mγ + γ0 + α + β + ν)Γ mγ + γ0 + α + 2  z  ν!(mγ + γ − ν)! 1  2m 0 Γ(mγ + γ0 + α + β)Γ ν + α + 2 (2mγ + 2γ + α + β)νΓ mγ + γ + α + 1  z2ν ≤ 0 0 2 1 1 α− 2  2ν ν!(mγ + γ0 − ν)!m Γ ν + α + 2 (2m) (mγ + γ )ν(2mγ + 2γ + α + β)νΓ mγ + γ + α + 1  z2ν ≤ 0 0 0 2 1 1 α− 2  2ν ν!(mγ + γ0)!m Γ ν + α + 2 (2m) = O(1).

29 5.2 Some Lemmas and Their Proofs

In order to prove Theorems 5.1.1 and 5.1.2 in the next section, we need some lemmas

involving the Jacobi polynomial and recurrent formulas related to bn in (5.1.2), besides Theorem 5.1.3.

For a nonnegative integer β, Lemma 5.2.1 expresses the n + β + 1 degree polynomial α− 1 ,β+ 1 α− 1 ,− 1 β+1 ( 2 2 ) ( 2 2 ) (1 + x) Pn (x) as a nonnegative combination of polynomials Pn+j (x), j = 0, 1, . . . , β + 1.

1 Lemma 5.2.1. For α > − 2 and β ∈ N0,

β+1 1 1 1 1 β+1 (α− 2 ,β+ 2 ) X α,β (α− 2 ,− 2 ) (1 + x) Pn (x) = aj,n Pn+j (x), x ∈ R, n ∈ N0, (5.2.1) j=0 where

β+1 3 α,β 2 (α + 2n + 2j)(n + j)!Γ(α + 2n + j)Γ(β + n + 2 )Γ(β + 2) aj,n = 1 . (5.2.2) n!j!Γ(α + β + 2n + j + 2)Γ(β − j + 2)Γ(n + j + 2 ) Proof. We will give two proofs of the Lemma. The first one is by mathematical induction, in the second one we use some identities from the references.

Proof 1. The proof of (5.2.1) is by induction on β.

For β = 0, (5.2.1) appears as identity (22.7.16) of [1],

1 1 (α− 2 ,β+ 2 ) (2n + α + β + 1)(1 + x)Pn (x) 1 1 1 1 (5.2.3) (α− 2 ,β− 2 ) (α− 2 ,β− 2 ) = (2n + 2β + 1)Pn (x) + 2(n + 1)Pn+1 (x). or,

1 1 (α− 2 ,β+ 2 ) (1 + x)Pn (x) 2n + 2β + 1 (α− 1 ,β− 1 ) 2(n + 1) (α− 1 ,β− 1 ) = P 2 2 (x) + P 2 2 (x). 2n + α + β + 1 n 2n + α + β + 1 n+1

Taking β = 0, we have

1 1 (α− 2 , 2 ) (1 + x)Pn (x) 2n + 1 (α− 1 ,− 1 ) 2(n + 1) (α− 1 ,− 1 ) = P 2 2 (x) + P 2 2 (x), 2n + α + 1 n 2n + α + 1 n+1

30 where 2n + 1 aα,0 = 0,n 2n + α + 1 and 2(n + 1) aα,0 = . 1,n 2n + α + 1 Assume that (5.2.1) is true for β = λ − 1, i.e,

λ 1 1 1 1 λ (α− 2 ,λ− 2 ) X α,λ−1 (α− 2 ,− 2 ) (1 + x) Pn (x) = aj,n Pn+j (x), j=0 substituting n by n + 1,

λ 1 1 1 1 λ (α− 2 ,λ− 2 ) X α,λ−1 (α− 2 ,− 2 ) (1 + x) Pn+1 (x) = aj,n+1 Pn+1+j (x). j=0 Then for β = λ, by the identity (5.2.3)

1 1 λ+1 (α− 2 ,λ+ 2 ) (1 + x) Pn (x)

1 1 λ (α− 2 ,λ+ 2 ) = (1 + x)(1 + x) Pn (x) 2n + 2λ + 1 (α− 1 ,λ− 1 ) 2(n + 1) (α− 1 ,λ− 1 ) = (1 + x)λP 2 2 (x) + (1 + x)λP 2 2 (x) 2n + α + λ + 1 n 2n + α + λ + 1 n+1 λ λ 2n + 2λ + 1 X (α− 1 ,− 1 ) 2(n + 1) X (α− 1 ,− 1 ) = aα,λ−1P 2 2 (x) + aα,λ−1P 2 2 (x) 2n + α + λ + 1 j,n n+j 2n + α + λ + 1 j,n+1 n+1+j j=0 j=0 2n + 2λ + 1 (α− 1 ,− 1 ) 2(n + 1) (α− 1 ,− 1 ) = aα,λ−1P 2 2 (x) + aα,λ−1P 2 2 (x) 2n + α + λ + 1 0,n n 2n + α + λ + 1 λ,n+1 n+1+λ λ λ−1 2n + 2λ + 1 X (α− 1 ,− 1 ) 2(n + 1) X (α− 1 ,− 1 ) + aα,λ−1P 2 2 (x) + aα,λ−1P 2 2 (x) 2n + α + λ + 1 j,n n+j 2n + α + λ + 1 j,n+1 n+1+j j=1 j=0 2n + 2λ + 1 (α− 1 ,− 1 ) 2(n + 1) (α− 1 ,− 1 ) = aα,λ−1P 2 2 (x) + aα,λ−1P 2 2 (x) 2n + α + λ + 1 0,n n 2n + α + λ + 1 λ,n+1 n+1+λ λ λ 2n + 2λ + 1 X (α− 1 ,− 1 ) 2(n + 1) X (α− 1 ,− 1 ) + aα,λ−1P 2 2 (x) + aα,λ−1 P 2 2 (x) 2n + α + λ + 1 j,n n+j 2n + α + λ + 1 j−1,n+1 n+j j=1 j=1 2n + 2λ + 1 (α− 1 ,− 1 ) 2(n + 1) (α− 1 ,− 1 ) = aα,λ−1P 2 2 (x) + aα,λ−1P 2 2 (x) 2n + α + λ + 1 0,n n 2n + α + λ + 1 λ,n+1 n+1+λ λ   1 1 X 2n + 2λ + 1 2(n + 1) (α− ,− ) + aα,λ−1 + aα,λ−1 P 2 2 (x) 2n + α + λ + 1 j,n 2n + α + λ + 1 j−1,n+1 n+j j=1 λ+1 1 1 X α,λ (α− 2 ,λ− 2 ) = aj,n Pn+j (x) j=0

31 where

2n + 2λ + 1 aα,λ = aα,λ−1, 0,n 2n + α + λ + 1 0,n

2(n + 1) aα,λ = aα,λ−1, λ+1,n 2n + α + λ + 1 λ,n+1

and

2n + 2λ + 1 2(n + 1) aα,λ = aα,λ−1 + aα,λ−1 j,n 2n + α + λ + 1 j,n 2n + α + λ + 1 j−1,n+1

for j = 1, 2, ··· , λ.

Proof 2. Successively using this identity (5.2.3), the n + β + 1 degree polynomial (1 + α− 1 ,β+ 1 α− 1 ,− 1 β ( 2 2 ) ( 2 2 ) x) Pn (x) can be written as a linear combination of polynomials Pn+j (x), j = α,β 0, 1, . . . , β + 1. To verify (5.2.2) for aj,n in (5.2.1), we multiply both sides of (5.2.1) by α− 1 ,− 1 α− 1 − 1 ( 2 2 ) (1 − x) 2 (1 + x) 2 Pn+j (x) and then integrate from -1 and 1,

2αΓ(α + n + j + 1 )Γ(n + j + 1 ) 2 2 aα,β (n + j)!(α + 2n + 2j)Γ(α + n + j) j,n β+1 1 1 1 1 1 Z 1 1 X α,β (α− 2 ,− 2 ) (α− 2 ,− 2 ) α− 2 − 2 = ak,n (1 − x) (1 + x) Pn+j (x)Pn+k (x)dx k=0 −1 1 1 1 1 1 Z 1 1 (α− 2 ,− 2 ) (α− 2 ,β+ 2 ) α− 2 β+ 2 = (1 − x) (1 + x) Pn+j (x)Pn (x)dx −1 2α+β+1Γ(α + n + j + 1 )Γ(α + 2n + j)Γ(β + n + 3 )Γ(β + 2) = 2 2 , n!j!Γ(α + n + j)Γ(α + β + 2n + j + 2)Γ(β − j + 2)

where the first equality is obtained from identity (7.391.1) of [13], and the last equality from

identity (7.391.9) of [13].

1 Suppose that g(ϑ) is a continuous function on [0, ∞) and g(ϑ) = 0, ϑ ≥ π. For α, β > − 2 , define

Z π 1 1    2α  2β (α,β) (α− 2 ,β− 2 ) ϑ ϑ ϑ bn,k = g(ϑ)Pn cos k sin k+1 cos k+1 dϑ, n ∈ N0, k ∈ N0, 0 2 2 2

32 (α,β) (α,β) and, for simplicity, write bn for bn,0 when k = 0. Thus,

Z π 1 1  2α  2β (α,β) (α,β) (α− 2 ,β− 2 ) ϑ ϑ bn = bn,0 = g(ϑ)Pn (cos ϑ) sin cos dϑ, 0 2 2

and, for bn in (5.1.2),

2α 1  2 Γ α + 2 Γ(n + 2α) (α,α) bn = 1  bn , n ∈ N0, Γ(2α)Γ n + α + 2 which is a direct result of (1.0.6).

1 Lemma 5.2.2. For α, β > − 2 ,

2n + 2β + 1 n + 1 b(α,β+1) = b(α,β) + b(α,β), n ∈ . (5.2.4) n 2(2n + α + β + 1) n 2n + α + β + 1 n+1 N0

1 If α > − 2 and β ∈ N0, then

β+1 (α,β+1) −(β+1) X α,β (α,0) bn,k = 2 aj,n bn+j,k, n ∈ N0, k ∈ N0, (5.2.5) j=0

and, in particular, for α ∈ N,

α (α,α) −α X α,α−1 (α,0) bn,k = 2 aj,n bn+j,k, n ∈ N0, k ∈ N0, (5.2.6) j=0 and α (α,α) −α X α,α−1 (α,0) bn = 2 aj,n bn+j , (5.2.7) j=0 α,β where aj,n ’s are positive constants defined in (5.2.2).

Proof. Substituting x by cos ϑ in (5.2.3) yields

1 1 (α− 2 ,β+ 2 ) (2n + α + β + 1) (1 + cos ϑ) Pn (cos ϑ)

1 1 1 1 (α− 2 ,β− 2 ) (α− 2 ,β− 2 ) = (2n + 2β + 1)Pn (cos ϑ) + 2(n + 1)Pn+1 (cos ϑ) ,

or   1 1 ϑ (α− ,β+ ) 2(2n + α + β + 1) cos2 P 2 2 (cos ϑ) 2 n

33 1 1 1 1 (α− 2 ,β− 2 ) (α− 2 ,β− 2 ) = (2n + 2β + 1)Pn (cos ϑ) + 2(n + 1)Pn+1 (cos ϑ) .

ϑ 2α ϑ 2β Multiplying both sides of the last equation by g(ϑ) sin 2 cos 2 and integrating from 0 to π, we obtain (5.2.4).

ϑ To derive (5.2.5), we substitute x by cos( 2k ) in (5.2.1),

β+1 1 1   X (α− ,− ) ϑ aα,βP 2 2 cos j,n n+j 2k j=0

 β+1 1 1   ϑ (α− ,β+ ) ϑ = 1 + cos P 2 2 cos 2k n 2k

 2(β+1) 1 1   ϑ (α− ,β+ ) ϑ = 2β+1 cos P 2 2 cos , 2k+1 n 2k

and obtain

Z π  2(β+1) 1 1    2α (α,β+1) ϑ (α− 2 ,β+ 2 ) ϑ ϑ bn,k = g(ϑ) cos k+1 Pn cos k sin k+1 dϑ 0 2 2 2 β+1 π 1 1    2α X Z (α− ,− ) ϑ ϑ = 2−(β+1) aα,β g(ϑ)P 2 2 cos sin dϑ j,n n+j 2k 2k+1 j=0 0 β+1 −(β+1) X α,β (α,0) = 2 aj,n bn+j,k. j=0

1 Lemma 5.2.3. For α > − 2 and n, k ∈ N0, Γ 2n + α + 1  Γ(n + 1) b(α,α) = 2 b(α,0), (5.2.8) 2n,k+1 2α 1  n,k 2 Γ n + α + 2 Γ(2n + 1) and, in particular, Γ 2n + α + 1  Γ(n + 1) b(α,α) = 2 b(α,0). (5.2.9) 2n,1 2α 1  n 2 Γ n + α + 2 Γ(2n + 1) Proof. By identity (4.1.5) of [25],

1  α− 1 ,α− 1 Γ 2n + α + Γ(n + 1) α− 1 ,− 1 ( 2 2 ) 2 ( 2 2 ) 2 P2n (x) = 1  Pn (2x − 1), Γ n + α + 2 Γ(2n + 1)

34 ϑ which reads, after substituting x by cos 2k+1 ,

1 1   1  1 1   (α− ,α− ) ϑ Γ 2n + α + Γ(n + 1) (α− ,− ) ϑ P 2 2 cos = 2 P 2 2 cos . 2n 2k+1 1  n 2k Γ n + α + 2 Γ(2n + 1) From the last equality we obtain

Z π 1 1    2α (α,α) (α− 2 ,α− 2 ) ϑ ϑ ϑ b2n,k+1 = g(ϑ)P2n cos k+1 sin k+2 cos k+2 dϑ 0 2 2 2 Z π 1 1    2α −2α (α− 2 ,α− 2 ) ϑ ϑ = 2 g(ϑ)P2n cos k+1 sin k+1 dϑ 0 2 2 1  π 2α Γ 2n + α + Γ(n + 1) Z α− 1 ,− 1     2 ( 2 2 ) ϑ ϑ = g(ϑ)Pn cos sin dϑ 2α 1  2k 2k+1 2 Γ n + α + 2 Γ(2n + 1) 0 Γ 2n + α + 1  Γ(n + 1) = 2 b(α,0). 2α 1  n,k 2 Γ n + α + 2 Γ(2n + 1)

Lemma 5.2.4. For α ∈ N, k ∈ N, and n ∈ N0, α α (α,α) (k−1)α P P α,α−1 α,α−1 bn = 2 ... a ··· a k k−1 j0,n jk,2 (n+j0)+2 j1+...+2jk−1 j0=0 jk=0 (5.2.10) (α,0) ×c (j0, . . . , jk, n) b k k−1 , 2 (n+j0)+2 j1+...+jk,k where

c (j0, j1, . . . , jk, n) k−1 l l−1 1  l+1 l  Y Γ 2 (n + j0) + 2 j1 + ... + jl + α + Γ 2 (n + j0) + 2 j1 + ... + 2jl + 1 = 2 Γ 2l+1(n + j ) + 2lj + ... + 2j + α + 1  Γ (2l(n + j ) + 2l−1j + ... + j + 1) l=0 0 1 l 2 0 1 l α,α−1 is a nonnegative constant and a k k−1 is defined as in (5.2.2) which is also jk,2 (n+j0)+2 j1+...+2jk−1 nonegative.

Proof. The proof of (5.2.10) is by induction on k. For k = 1, (5.2.10) follows from identities

(5.2.7), (5.2.9) and (5.2.6)

(α,α) bn α X (α,0) = 2−α aα,α−1b j0,n n+j0 j0=0 α 2α 1  X 2 Γ n + j0 + α + Γ(2(n + j0) + 1) (α,α) = 2−α aα,α−1 2 b j0,n 1  2(n+j0),1 Γ 2(n + j0) + α + Γ(n + j0 + 1) j0=0 2

35 α 2α 1  α X 2 Γ n + j0 + α + Γ(2(n + j0) + 1) X (α,0) = 2−α aα,α−1 2 2−α aα,α−1 b j0,n 1  j1,2(n+j0) 2(n+j0)+j1,1 Γ 2(n + j0) + α + Γ(n + j0 + 1) j0=0 2 j1=0 α α 1  X X Γ n + j0 + α + Γ(2(n + j0) + 1) (α,0) = aα,α−1 aα,α−1 2 b . j0,n j1,2(n+j0) 1  2(n+j0)+j1,1 Γ 2(n + j0) + α + Γ(n + j0 + 1) j0=0 j1=0 2 Suppose that (5.2.10) is true for k = i, i.e.,

α α (α,α) (i−1)α P P α,α−1 α,α−1 b = 2 ... a ··· a i i−1 n j0,n ji,2 (n+j0)+2 j1+...+2ji−1 j0=0 ji=0 (α,0) ×c (j0, . . . , ji, n) b i i−1 , 2 (n+j0)+2 j1+...+ji,i then for k = i + 1, by (5.2.8) and (5.2.6)

(α,α) bn α α (i−1)α X X α,α−1 α,α−1 = 2 ... a ··· a i i−1 j0,n ji,2 (n+j0)+2 j1+...+2ji−1 j0=0 ji=0 (α,0) ×c (j0, . . . , ji, n) b i i−1 2 (n+j0)+2 j1+...+ji,i α α (i−1)α X X α,α−1 α,α−1 = 2 ... a ··· a i i−1 c (j0, . . . , ji, n) j0,n ji,2 (n+j0)+2 j1+...+2ji−1 j0=0 ji=0 22αΓ 2i(n + j ) + 2i−1j + ... + j + α + 1  Γ(2(2i(n + j ) + 2i−1j + ... + j ) + 1) × 0 1 i 2 0 1 i i i−1 1  i i−1 Γ 2(2 (n + j0) + 2 j1 + ... + ji) + α + 2 Γ(2 (n + j0) + 2 j1 + ... + ji + 1) (α,α) ×b i i−1 2(2 (n+j0)+2 j1+...+ji),i+1 α α (i−1)α X X α,α−1 α,α−1 = 2 ... a ··· a i i−1 c (j0, . . . , ji, n) j0,n ji,2 (n+j0)+2 j1+...+2ji−1 j0=0 ji=0 22αΓ 2i(n + j ) + 2i−1j + ... + j + α + 1  Γ(2(2i(n + j ) + 2i−1j + ... + j ) + 1) × 0 1 i 2 0 1 i i i−1 1  i i−1 Γ 2(2 (n + j0) + 2 j1 + ... + ji) + α + 2 Γ(2 (n + j0) + 2 j1 + ... + ji + 1) α −α X α,α−1 (α,0) ×2 a i i−1 b i i−1 ji+1,2(2 (n+j0)+2 j1+...+ji) 2(2 (n+j0)+2 j1+...+ji)+ji+1,i+1 ji+1=0 α α α iα X X X α,α−1 α,α−1 α,α−1 = 2 ... a ··· a i i−1 a i i−1 j0,n ji,2 (n+j0)+2 j1+...+2ji−1 ji+1,2(2 (n+j0)+2 j1+...+ji) j0=0 ji=0 ji+1=0

×c (j0, . . . , ji, n) Γ 2i(n + j ) + 2i−1j + ... + j + α + 1  Γ2i+1(n + j ) + 2ij + ... + 2j + 1) × 0 1 i 2 0 1 i i+1 i 1  i i−1 Γ 2 (n + j0) + 2 j1 + ... + 2ji + α + 2 Γ(2 (n + j0) + 2 j1 + ... + ji + 1) (α,0) ×b i+1 i 2 (n+j0)+2 j1+...+2ji+ji+1,i+1

36 α α iα X X α,α−1 α,α−1 = 2 ... a ··· a i i−1 c (j0, . . . , ji+1, n) j0,n ji+1,2(2 (n+j0)+2 j1+...+ji) j0=0 ji+1=0 (α,0) ×b i+1 i , 2 (n+j0)+2 j1+...+2ji+ji+1,i+1

where

c (j0, . . . , ji+1, n) i l l−1 1  l+1 l  Y Γ 2 (n + j0) + 2 j1 + ... + jl + α + Γ 2 (n + j0) + 2 j1 + ... + 2jl + 1 = 2 . Γ 2l+1(n + j ) + 2lj + ... + 2j + α + 1  Γ (2l(n + j ) + 2l−1j + ... + j + 1) l=0 0 1 l 2 0 1 l

By the principle of mathematical induction, (5.2.10) holds for any k ∈ N.

A common and important feature of recurrent formulas (5.2.6)-(5.2.10) is that all coeffi-

cients are nonnegative. More importantly, k at the right-hand side of (5.2.10) is an arbitrary positive integer, making which large enough and combining with the following lemma help us to prove Theorems 5.1.1 and 5.1.2 in the next section.

1 Lemma 5.2.5. For α > − 2 and γ0 ∈ N0, if {γk, k ∈ N} is a sequence of bound positive k rational numbers such that 2 γk ∈ N and lim γk = γ > 0, then k→∞ Z π k α+ 1 (α,0) 1 1 ( 2 ) −α− 2 α+ 2 lim 2 b k = 2 g(ϑ)ϑ J 1 (γϑ)dϑ. (5.2.11) 2 γk+γ0,k α− 2 k→∞ 0

1 (α,0) k(α+ 2 ) Proof. Rewriting 2 b k as 2 γk+γ0,k

1 (α,0) k(α+ 2 ) 2 b k 2 γk+γ0,k !2α Z π  2α 1 1   ϑ ϑ 1 (α− ,− ) ϑ sin k+1 −k(α− 2 ) 2 2 2 = g(ϑ) 2 P k cos dϑ, 2 2 γk+γ0 2k ϑ 0 2k+1 we obtain (5.2.11) from Theorem 5.1.3, by letting k → ∞,

1 (α,0) k(α+ 2 ) lim 2 b k k→∞ 2 γk+γ0,k

37 !2α Z π  2α 1 1   ϑ ϑ 1 (α− ,− ) ϑ sin k+1 −k(α− 2 ) 2 2 2 = g(ϑ) lim 2 P k cos lim dϑ 2 k→∞ 2 γk+γ0 2k k→∞ ϑ 0 2k+1 π 2α −(α− 1 ) Z ϑ ϑ 2 = g(ϑ) J 1 (ϑγ)dϑ α− 2 0 2 2 Z π −(α+ 1 ) α+ 1 = 2 2 g(ϑ)ϑ 2 J 1 (ϑγ)dϑ, α− 2 0 where the exchange between the limit and integral at the first equality is ensured by the dominated convergence theorem.

Lemma 5.2.6. For α ∈ N0, k ∈ N0, and n ∈ N0,

1 α− 1 Z π 1 (α,0) 2 2 (n + α − )! 1 Jα− (xϑ) 2α(k+1) 2 α+ 2 2 lim 2 b = lim ϑ 1 g(ϑ)dϑ. (5.2.12) n,k + α− k→∞ n! x→0 0 x 2

Proof. It can be proved by the following two limiting forms, namely,

2α π ϑ ! Z 1 1 sin 2α(k+1) (α,0) 2α (α− 2 ,− 2 ) ϑ 2k+1 lim 2 bn,k = g(ϑ)ϑ lim Pn (cos ) dϑ k→∞ k→∞ 2k ϑ 0 2k+1 1 1 Z π (α− 2 ,− 2 ) 2α = Pn (1) g(ϑ)ϑ dϑ, 0

(α− 1 ,− 1 ) 1 2 2 n+α− 2  where Pn (1) = n by (4.1.1) of [25], and

π π Z 1 Jα− 1 (xϑ) Z 1 Jα− 1 (xϑ) α+ 2 2 α+ 2 2 lim ϑ 1 g(ϑ)dϑ = ϑ lim 1 g(ϑ)dϑ + α− + α− x→0 0 x 2 0 x→0 x 2 1 −α Z π 2 2 2α = 1 g(ϑ)ϑ dϑ, Γ(α + 2 ) 0

zα which is a direct consequence of (10.7.3) of [19], Jα(z) ∼ 2αΓ(α+1) by fixing α and letting z → 0.

5.3 Proof of Theorems 5.1.1 and 5.1.2

Since g(x) is continuous with compact support on [0, ∞) and g(kxk) is positive definite

in Rd, inequality (5.1.1) holds for every x ≥ 0, i.e.,

Z π α+ 1 u 2 J 1 (xu)g(u)du ≥ 0, x ≥ 0. α− 2 0

38 What needs a proof is inequality (5.1.2),i.e.,

Z π (α) 2α bn = g(ϑ)Pn (cos ϑ) sin ϑdϑ ≥ 0, n ∈ N0, 0

(α,α) d−1 or bn ≥ 0 for every n ∈ N0, where α = 2 is a nonnegative integer. While the case d = 1 or 3 is covered in [15], it suffices to consider d ≥ 5 or α ∈ N. (α,α) To prove bn ≥ 0 for each n ∈ N, from identity (5.2.10) we need to show

(α,0) b k k k−1 ≥ 0, ji = 0, 1, 2, . . . , α, i = 1, 2, . . . , k. 2 n+2 j0+2 j1+...+jk,k

Since k in (5.2.10) is an arbitrary positive integer, it can be chosen large enough. Noticing

that

k k k k−1 k 2 n ≤ 2 n + 2 j0 + 2 j1 + ... + jk ≤ 2 (n + 2α) − α,

k k k−1 we can write 2 n + 2 j0 + 2 j1 + ... + jk as

k k k−1 k 2 n + 2 j0 + 2 j1 + ... + jk = 2 γk + γ0,

where γ0 ∈ N0 and γk is a positive rational number bounded between n and n + 2α, so that

(α,0) (α,0) b k k k−1 = b k . 2 n+2 j0+2 j1+...+jk,k 2 γk+γ0,k

First, suppose that inequality (5.1.1) is strict, then the right-hand side of (5.2.11) is also

positive, i.e., Z π α+ 1 g(ϑ)ϑ 2 J 1 (γϑ)dϑ > 0. α− 2 0 This is due to (5.1.1) is strict and γ > 0. Then its left-hand side has to be positive.

Consequently, when k is large enough,

(α,0) (α,0) b k = b k k k−1 ≥ 0, 2 γk+γ0,k 2 n+2 j0+2 j1+...+jk,k

(α,α) and it implies bn ≥ 0 by (5.2.10). Moreover, since the left-hand side of (5.2.11) is positive,

(α,0) (α,α) there exists at least one number k such that b k k k−1 is positive, so that bn 2 n+2 j0+2 j1+...+jk,k must be positive.

39 k Now we consider the case n = 0, there are two possibilities for 2 γk + γ0. If it is

(α,0) unbounded, then, by Lemma 5.2.5, b k ≥ 0 for large k. If it is bounded, then it follows 2 γk+γ0,k (α,0) (α,α) from Lemma 5.2.6 that b k ≥ 0 when k is large enough. By (5.2.10), b ≥ 0. 2 γk+γ0,k 0 In a general case, (5.1.1) implies Z π α+ 1 ϑ 2 J 1 (γϑ)g(ϑ)dϑ ≥ 0, γ ≥ 0. α− 2 0 For an arbitrary  > 0, we define a function

α+1 g(ϑ) = g(ϑ) + (π − ϑ)+ , ϑ ≥ 0.

By Theorem 7 of [11], Z π α+ 1 α+1 ϑ 2 J 1 (γϑ)(π − ϑ) dϑ > 0, γ ≥ 0, α− 2 + 0 so that Z π α+ 1 ϑ 2 J 1 (γϑ)g(ϑ)dϑ > 0, γ ≥ 0. α− 2 0 It implies

Z π 1 1  2α  2α (α− 2 ,α− 2 ) ϑ ϑ g(ϑ)Pn (cos ϑ) sin cos ≥ 0, n ∈ N. 0 2 2 (α,α) Let  → 0, we obtain bn ≥ 0 for n ∈ N0. ∞ d−1 P ( 2 ) What remains is to verify that bnPn (1) converges. As is shown in [22], the con- n=0 ∞ d−1 P ( 2 ) tinuous function g(ϑ) possesses the expansion bnPn (cos ϑ), which is Abel-summable n=0 for every ϑ ∈ [0, π], and, in particular, for ϑ = 0. Thus,

k k ∞ d−1 d−1 d−1 X ( 2 ) X ( 2 ) X n ( 2 ) bn Pn (cos ϑ) ≤ bnPn (1) ≤ lim bnr Pn (1) = g(0). r→1− n=0 n=0 n=0 Let k d−1 X ( 2 ) Sk = bn Pn (cos ϑ) . n=0

Then sequence {Sk} is bounded and monotone, so it is convergent. Hence the sequence

k d−1 X ( 2 ) Sk = bnPn (cos ϑ). n=0 is absolutely and uniformly convergent for all ϑ ∈ [0, π], and converges to its Abel-sum g(ϑ).

40 CHAPTER 6

PLANS FOR FUTURE WORK

In this thesis, we have reviewed the classical characterizations of isotropic positive def- inite functions on Euclidean spaces and spheres, and we have applied them to develop a construction of isotropic positive definite functions on spheres. However, this construction only works when the dimension of the sphere is odd. Our next work is to develop isotropic positive definite functions on all spheres. It is stated as the following conjecture.

Conjecture. Suppose that g(x) is a continuous function on [0, ∞), and g(x) = 0, x ≥ π.

For a positive integer d, if g(kxk) is a positive definite function in Rd, then g(ϑ), ϑ ∈ [0, π], is a positive definite function on Sd.

41 REFERENCES

42 LIST OF REFERENCES

[1] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions, Dover Publi- cations, New York, 1965.

[2] R. Askey, Radial characteristic functions , Tech. Report No.1262, Math. Research Cen- ter, University of Wisconsin-Madison, 1973.

[3] H. F. Baker, Abel’s theorem and the allied theory, University Press, Cambridge, 1897

[4] R. Beatson, W. zu Castell and Y. Xu, A P´olya criterion for (strict) positive-definiteness on the sphere, IMA J. Numer. Anal. 34 (2014), 550-568.

[5] S. Bochner and K. Chandrasekharan, Fourier Transforms, Princeton University Press, Princeton, 1949.

[6] D. Cheng and Y. Xiao, Excursion probability of Gaussian random fields on sphere, Bernoulli. 22 (2016), 1113-1130.

[7] K.L. Chung, A Course in Probability Theory, Academic Press, 2001

[8] F. Dai and Y. Xu, Approximation Theory and Harmonic Analysis on Spheres and Balls, Springer, New York, 2013.

[9] Stein, Elias; Weiss, Guido , Introduction to Fourier Analysis on Euclidean Spaces, N.J.: Princeton University Press, Princeton, 1971.

[10] G. E. Fasshauer, Meshfree Approximation Methods with Matlab, World Scientific, New Jersey, 2007.

[11] G. Gasper, Positive sums of the classical orthogonal polynomials, SIAM J. Math. Anal. 8 (1977), 423-447.

[12] T. Gneiting, Strictly and non-strictly positive definite functions on spheres, Bernoulli, 2013.

[13] I. S. Gradshteyn and I. M. Ryzhik, Tables of Integrals, Series, and Products, 7th ed. Academic Press, Amsterdam, 2007.

[14] M. Jevti , D.Vukoti, M.Arsenovi, The Poisson Integral. In: Taylor Coefficients and Coefficient Multipliers of Hardy and Bergman-Type Spaces, RSME Springer Series, vol 2. Springer, Cham, 2016

[15] C. Ma, Stochastic representations of Isotropic vector random fields on spheres, Stoch. Anal. Appl. 34 (2016), 389-403.

43 LIST OF REFERENCES (continued)

[16] C. Ma, Isotropic covariance matrix polynomials on spheres, Stoch. Anal. Appl. 34 (2016), 679-706.

[17] C. Ma, Time varying isotropic vector random fields on spheres, J. Theor. Probab. 30 (2017), 1763-1785.

[18] Z. Nie, C. Ma, Isotropic positvie definite functions on spheres generated from those in Euclidean spaces, Proc. Amer. Math. Soc. 147 (2019), 3047-3056.

[19] F. W. J. Olver, D. W. Lozier, R. F. Boisvert, and C. W. Clark, NIST Handbook of Mathematical Functions, Cambridge University Press, Cambridge, 2010.

[20] C. E. Rasmussen and C. K. I. Williams, Gaussian Processes for Machine Learning, Adaptive Computation and Machine Learning, The MIT Press, Massachusetts, 2006.

[21] I. J. Schoenberg, Metric spaces and completely monotone functions, Ann. Math. 39 (1938), 811-841.

[22] I. J. Schoenberg, Positive definite functions on spheres, Duke Math. J. 9 (1942), 96-108.

[23] I. N. Sneddon, The Use of Integral Transforms, New York: McGraw-Hill, 1972.

[24] P.K. Suetin, Ultraspherical polynomials, in Hazewinkel, Michiel, Encyclopedia of Math- ematics, Springer Science+Business Media B.V. / Kluwer Academic Publishers, 2001.

[25] G. Szeg¨o, Orthogonal Polynomials, AMS Coll. Publ., Vol. 23, Providence, RI, 1975.

[26] Faris, G. William, Radial functions and the Fourier transform, Department of Mathe- matics, University of Arizona, 2008.

[27] H. Wendland, Scattered Data Approximation, Cambridge University Press (Cambridge), 2005.

[28] Y. Xu, Positive definite functions on the unit sphere and integrals of Jacobi polynomials, Proc. Amer. Math. Soc. 146 (2018), 2039-2048.

44