Variational Principles and Sobolev-Type Estimates for ... - CiteSeerX

20 downloads 0 Views 323KB Size Report
Nira Dyn, Francis J. Narcowich, and Joseph D. Ward. Abstract .... Laplace-Beltrami operator, which in local coordinates has the form 8, p. 11] w := g ?1. 2 m. X.
Variational Principles and Sobolev-Type Estimates for Generalized Interpolation on a Riemannian Manifold by Nira Dyn, Francis J. Narcowich, and Joseph D. Ward

Abstract

The purpose of this paper is to study certain variational principles and Sobolev-type estimates for the approximation order resulting from using strictly positive de nite kernels to do generalized Hermite interpolation on a closed (i.e., no boundary), compact, connected, orientable, m-dimensional C 1 Riemannian manifold M, with C 1 -metric gij . The rate of approximation can be more fully analyzed with rates of approximation given in terms of Sobolev norms. Estimates on the rate of convergence for generalized Hermite and other distributional interpolants can be obtained in certain circumstances and nally the constants appearing in the approximation order inequalities are explicit. Our focus in this paper will be on approximation rates in the cases of the circle, other tori and the 2-sphere.

x0. Introduction

Over the last few years, there has been an increasing interest in approximation methods based on variational techniques in a reproducing kernel Hilbert space setting. Perhaps the rst paper on this topic was that of Golomb and Weinberger [10]. More recently, Duchon [5, 6] and Madych and Nelson [15, 16] have extensively investigated approximation properties of functions having nonnegative Fourier transforms or, more generally, conditionally positive de nite functions of order n on R s . Wu-Schaback [30] used kriging methods, which are based on variational techniques, in their investigation of the approximation properties for this class of functions. In this framework, the approximation rates apply to functions in a certain \induced" space. The object of this paper is to investigate variational principles and Sobolev-type estimates for the approximation order resulting from using strictly positive de nite kernels to do generalized Hermite interpolation on a closed (i.e., no boundary), compact, connected, orientable, m-dimensional C 1 Riemannian manifold M m , with C 1 -metric gij [1, 8, 11]. Generalized Hermite interpolation in R s was introduced in [20, 29], and on a manifold in [19]. We should also mention that Xu and Cheney [31] gave conditions for Schoenberg's spherical positive de nite functions [24] to be strictly positive, thereby providing a wide class of basis functions with which to do interpolation on Sm. It frequently happens that the induced spaces contain certain Sobolev spaces. The rate of approximation can therefore be analyzed in terms of Sobolev norms. Estimates on the rate of convergence for generalized Hermite and other distributional interpolants can be obtained in certain circumstances, and nally the constants appearing in the approximation order inequalities are explicit. Of special interest in this paper will be the approximation 1

rates obtained in the setting of the 2-sphere. Some work in this direction has already been done [7, 17, 22, 26, 27]. A crucial ingredient for our estimates is a recent sampling theorem for band-limited functions on the 2-sphere given in [4]. The outline for the paper is as follows: In Section 1, relevant notation and certain basic results about Sobolev spaces on Riemannian manifolds are given. In Section 2, material about complex-valued kernels are presented. Special attention is given to those kernels which are positive de nite and strictly positive de nite within our framework. Also the abstract interpolation problem is presented. In Section 3, a variational principle is derived which shows that the solution to the abstract interpolation problem satis es a certain minimal norm criterion which in turn gives rise to the \rate of approximation" estimates. In Section 4, the notion of conditionally positive de nite kernels on manifolds is introduced. In Section 5, rate of approximation estimates are derived in the case of the circle and other tori for both equally-spaced and scattered points. In Section 6, approximation rates for certain kernels on S2 are derived.

x1. Notation and Background

We will introduce notation and discuss pertinent matters related to manifolds. For the most part, we will use the notation for manifolds and distributions that is used in Friedlander [8, x2.8] and in [19]. In particular, in local coordinates the invariant volume element d induced by the Riemannian metric gij is

d(x) = g(x)1=2dx; where g(x) := det gij (x) > 0 and dx := dx1    dxm: As usual, we denote the components of the inverse of gij by gij . By C k (M m ) we will mean the collection of all k-times continuously di erentiable functions on M m ; the compactness of the closed manifold M m implies that C k (M m ) = C0k (M m ), where C0k (M m ) comprises all C k functions whose support is contained in a compact subset of M m . Much of what we do here is related to expansions involving eigenfunctions of the Laplace-Beltrami operator, which in local coordinates has the form [8, p. 11]

 1 @w  m m 1 XX @ ij ? w := g 2 i g 2 g @xj ; where g = det gij : @x i=1 j =1 It is possible to show that  is self-adjoint relative to L2(M m ; g) [1, p, 54] and that it is elliptic [28, p. 250]. The eigenfunctions mentioned above arise in connection with the eigenvalue problem F + F = 0; F 2 C 2 (M m ); concerning which we have the following well-known result: Proposition 1.1: The eigenvalues are discrete, nonnegative, and form an increasing sequence with +1 as the only accumulation point. The eigenspaces corresponding to a given eigenvalue are nite dimensional subspaces of C 1 (M m ). The eigenspaces are orthogonal in L2(M m ; g), and the direct sum of these spaces is all of L2(M m ; g). Proof: See [1, Chapter 3] and [28, Chapter 6]. 2

We remark that much of we do here could also be done with other eigenfunction expansions, provided one replaces the sobolev spaces by weighted eigenfunction expansions. Occasionally we need to label the eigenvalues. A common choice for the index j comes from labeling in increasing order the eigenvalues j , with appropriate repetitions for degenerate cases. The compactness of M m implies 1 = 0 is the lowest eigenvalue; it is easy to show that it has multiplicity 1, so j > 0 if j > 1. Other indexing schemes are, of course, possible. We will let Fj be the eigenfunction corresponding to the eigenvalue j . Of course, one can choose the Fj 's so that they form an orthonormal basis for L2 (M m ; g), and we do so here. We also want to point out that the theorem above implies that each Fj is actually a C 1 funtion. In [19], we discussed in detail the de nition of a distribution on a Riemannian manifold, and we refer the reader to that paper for the necessary background material. We shall denote the set of all distributions on the manifold M m by D0(M m ). To evaluate any distribution, we may use any partition of unity subordinated to a nite covering of M m to evaluate (u; ). Let P = f(  ;  )g be a nite set of coordinate charts that cover M m , and let X = f g be a partition of unity subordinated to this covering. We have (u; ) =

XZ 

 (  )

u  ?1 (x)( )  ?1(x)d(x):

This is the precise meaning of the more standard expression (u; ) =

Z

Mm

u(p)(p)d(p):

We will make heavy use of Sobolev spaces [9, Chapter I; 13, x1.7] in our analysis. In terms of the eigenvalue problem mentioned above, these spaces can be characterized in a relatively simple way [13, x1.7.3]. If s is an arbitrary real number, then

Hs (M m ) = fu 2 D0 (M m ) :

1 X j =1

sj jub(j )j2 < 1g; ub(j ) := (u; Fj );

where Fj being in C 1 (M m ) implies that (u; Fj ) is well de ned. The norm on this space will be taken to be 11=2 0 1 X @ kuks = (1 + j )sjub(j )j2A : j =1

Below, we collect some important, standard results about Sobolev spaces: Lemma 1.2: If s is an arbitrary real number, H?s = Hs  . If s > k + m2 , then Hs(M m )  C k (M m ). Proof: See Gilkey's book, [9, p. 35]. We need to consider tensor products of distributions. Let  be a C 1 function  : M m  M m ! C , and let u and v be in D 0 (M m ). Using [8, Theorem 2.3.2, p. 42] together 3

with the partitions of unity discussed above, one may de ne the tensor product of u and v as the distribution u v that is given via (u v; ) = =

Z

Mm

Z

Mm

u(p) v(q)



Z

Z

Mm Mm

v(q)(p; q)d(q) d(p)



u(p)(p; q)d(p) d(q):

It is important to note that the following result holds for the tensor products of distributions. Proposition 1.3: Let s and t be positive real numbers. If u 2 H?s (M m ) and v 2 H?t (M m ), then u v 2 H?s?t (M m  M m ). Moreover,

ku vk?s?t  kuk?s kvk?t: Proof: The Laplace-Beltrami operator for the product manifold M m  M m has eigenvalues j + k and corresponding eigenfunctions Fj Fk . Clearly, (u v; Fj Fk ) = (u; Fj )(v; Fk ). Consequently, 1 X 2 ku vk?s?t = (1 + j + k )?s?t j(u v; Fj Fk )j2 j;k=1 1 X = (1 + j + k )?s?t j(u; Fj )(v; Fk )j2 j;k=1 1 X (1 + k )t (1 +  )?sj(u; F )j2(1 +  )?t j(v; F )j2 (1 + j )s = j j k k s t j;k=1 (1 + j + k ) (1 + j + k ) 1 1 X X ? s 2   ( (1 + j ) j(u; Fj )j )( (1 + k )?t j(v; Fk )j2) = kuk2?skvk2?t j =1 k=1

Taking square roots yields the norm-inequality in the lemma. We will close our introductory remarks with a brief discussion of the mapping properties associated with a kernel  2 H2s \ C 0 (M m  M m ). Clearly, one can de ne the linear transformation  : C 0 (M m ) ! C 0 (M m ) via the expression  [f ](p) =

Z

Mm

(p; q)f (q)d(q):

Using the same kind of proof as in the last proposition, we can obtain the following result. Proposition 1.4: If 0  t  2s and if  2 H2s \ C 0 (M m  M m ), then the linear map  de ned above continuously maps H?t into H2s?t. Moreover, we have that

k  k  kk2s : 4

Proof: Omitted.

In order to emphasize the kernel generating  , we will adopt the convention that

 ? u :=  [u]: The symbol \?" is used to suggest convolution, although in using the symbol one should keep in mind that ? is not commutative. Any continuous kernel (p; q) de ned on M m has an eigenfunction expansion of the form X (p; q) = j;k Fk (p)Fj (q): j;k

Later on, we will work with the subclass of kernels for which j;k = ak j;k . Speci cally, these kernels have the form (1:1)

(p; q) :=

1 X

k=1

ak Fk (p)Fk (q); where

1 X

k=1

(1 + 2k )2sjak j2 < 1:

The condition on the ak 's is equivalent to assuming that  is in H2s(M m  M m ). Kernels of the form (1.1) are analogous to convolution kernels and are similar to those treated in x3 of [19]. Concerning such kernels, we have the following: Proposition 1.5: Let s be a positive real number. If (p; q ) has the form (1.1), then  is a selfadjoint kernel, (p; q) = (q; p). In addition, if u and v be in H?s (M m ), then (1:2)

(u v; ) =

1 X k=1

ak (u; Fk )(v; Fk ):

Proof: That  is selfadjoint is obvious from (1.1). By Proposition 1.3, u v is in H?2s(M m  M m ). Since  2 H2s(M m  M m ) has the expansion (1.1), which is obviously convergent in H2s(M m  M m ), we can replace  by (1.1) on the left in (1.2), and then inter-

change the sum with the continuous linear functional u v to obtain the series expansion on the left in (1.2).

x2. Strictly Positive De nite Kernels A complex-valued kernel  2 C (M m  M m ) is termed positive de nite on M m if (q; p) = (p; q) and if for every nite set of points C = fp1 ; : : : ; png in M m , the selfadjoint, N  N matrix with entries (pj ; pk ) is positive semi-de nite. The case in which 

is C 1 was studied in detail in [19]. Positive de nite kernels have arisen in many contexts [25]. The de nition that we use here is motivated by Schoenberg's [24], and is related to the ones studied by M. G. Krein [12], Yu. M. Berezanskii [2], and K. Maurin [14]. We will be especially interested in H2s positive de nite kernels, for which we have the following result, which is similar to one for the C 1 case [19, Theorem 2.1]: 5

Theorem 2.1: Let the kernel  be in H2s (M m  M m ) \ C 0 (M m  M m ), and let  satisfy (q; p) = (p; q). Then,  is positive de nite if and only if for every u 2 H?s (M m ),

(u u; )  0:

(2:1)

Proof: The proof is similar to that given in [19, Theorem 2.1], and we will omit it.

Later on, we will construct a number of examples of positive de nite kernels, but rst we need to de ne precisely the generalized Hermite interpolation problem that we wish to study. Let fuj gNj=1  H?s be a linearly independent set of distributions on M m , and let f 2 Hs . We may think of f as being \measured" and producing data in the form:

Z

Mm

uj (p)f (p)d(p) = dj ; for j = 1; : : :; N:

where the dj 's are complex numbers. The condition that the set fuj gNj=1 be linearly independent is equivalent to there being no redundancy in the data. m Problem 2.2: Given a linearly independent set fuj gN j =1  H?s (M ), complex numbers dj , j = 1; : : : ; N , and an H2s(M m  M m ) \ C 0 (M m  M m ) positive de nite kernel  on M m , nd u 2 Spanfu1 ; : : :; uN g such that

Z

uj (p)( ? u)(p)d(p) = dj ; for j = 1; : : :; N:

Mm

Equivalently, nd u 2 Spanfu1 ; : : :; uN g such that the function  ? u is a Hs (M m ) interpolant for the data generated by the function f above. The problem itself encompasses the problem of scattered data interpolation on a manifold; it also includes almost any problem involving data generated by derivatives, di erences, and even continuous distributions. We will say that such a problem is wellpoised if there exists a solution and the solution is unique. The next theorem characterizes the types of positive de nite kernels  on M m relative to which generalized Hermite problems are well-poised. Before we can state it, we need a de nition: Let  be a H2s(M m M m \C 0 (M m M m ) positive de nite kernel on M m . If the equation (u u; ) = 0 implies that the distribution u = 0, then we will say that  is strictly positive de nite (SPD) on M m . The results given below have proofs nearly identical to the corresponding results [19, Theorem 2.4 and Corollary 2.5]; they will be stated without proof. Theorem 2.3: Problem 2.2 is well-poised if  is strictly positive de nite on M m . Corollary 2.4: If  is a strictly positive de nite kernel on M m , and if the set of distributions fuj gNj=1 is linearly independent, then the interpolation matrix A with entries given by (2:2)

Aj;k :=

Z

Mm

uj (p)( ? uk )(p)d(p); where j; k = 1; : : :; N:

is self-adjoint and positive de nite. 6

Theorem 2.3 establishes that Problem 2.2 will be well-poised provided one has a strictly positive de nite kernel  de ned on M m , and Corollary 2.4 gives us that the corresponding interpolation matrices will be positive de nite. The next result gives us a large class of strictly positive de nite kernels. Corollary 2.5: Let s be a positive real number. If (p; q ) has the form (1.1), then  is strictly positive de nite if and only if ak > 0 for all k  1. Proof: By Proposition 1.5, we have that  (p; q ) = (q; p), and that, if u 2 H?s (M m ), (u u; ) = = (2:3)

=

1 X

k=1 1 X k=1 1 X k=1

ak (u; Fk )(u; Fk ) ak j(u; Fk )j2 ak jub(k)j2;

where ub(k) = (u; Fk )is the kth coecient in the eigenfunction expansion for u. Since ak  0, the expansion (2.3) clearly establishes that  is a positive de nite kernel. Moreover, choosing u = Fj in (2.3) gives (u u; ) = aj . Thus, if  is strictly positive de nite, we see that all the ak 's are positive. We want to show that the converse is also true. Assume that ak > 0 for all k. If (u u; ) = 0 for some u 2 H?s(M m ), then since ak > 0 we have that ub(k) = (u; Fk ) = 0 for all k, from which it follows that u = 0. Hence,  is strictly positive. Many manifolds of interest|tori of dimension two or greater, for example| are products of compact, connected Riemannian manifolds. Choosing the Riemannian metric on M m := M m1      M md to be the one induced by the Riemannian metrics on each of the M m 's, one gets a Laplace-Beltrami operator that is the sum of those coming from the M m 's. For such an operator, one easily sees that the eigenvalues and eigenfunctions have, respectively, the forms (2:4)  = 1j1 +    + d d and F (p1 ; : : :; pd ) = F 11 (p1)    F dd (pd ); where = ( 1 : : : ; d) is a multi-index. The eigenfunctions being products allows one to obtain strictly positive de nite kernels on M m1      M md from products of kernels that are strictly positive de nite on each M m and that have the form (1.1). Corollary 2.6: If the factors in the product (2:5) (p1 ; : : :; pd ; q1 ; : : :; qd ) := 1 (p1; q1 )    d (pd ; qd ): are strictly positive de nite kernels in H2s(M m  M m ) and have the form (1.1), then the kernel  de ned by (2.5) is in H2s(M m  M m ), has the form (1.1), and is strictly positive de nite on the product manifold M m1      M md . Moreover, (2:6)

kk2s 

Yd

=1

7

k k2s :

Proof: The proof that  has an expansion of the form (1.1) with positive coecients is identical to the one for [19, Corollary 3.10] in the C 1 case. If

 =

1 X k=1

a k Fk (p )Fk (q )

then coecient multiplying the term F (p1 ; : : :; pd )F (q1 ; : : :; qd ) is (2:7)

a =

Yd

=1

a :

Q From the inequality (1 + 21 1 +    2d d )2s  d =1 (1 + 2 )2s; it follows that kk22s = 

X

(1 + 21 1 +    2d d )2s(a )2

0 1 Yd X

@

1 (1 + 2 )2s(a )2A

=1 =1 Yd = k k22s

=1

Taking square roots in the last inequality establishes (2.6) and completes the proof.

x3. A Variational Principle Let  be a strictly positive de nite kernel in H2s(M m  M m ). We can use  to de ne an inner product on the space H?s (M m ). If u and v are in H?s (M m ), then v u is in H?2s(M m  M m ). Thus, the form (3:1)

[ u ; v] := (v u; )

is well de ned, and by Proposition 1.3 satis es the inequality (3:2)

j[ u ; v] j  kuk?s kvk?skk2s:

The selfadjointness and strict positivity of  imply that (3.1) is an inner product on H?s (M m ). The norm associated with this inner product is (3:3)

p

[]u[] := (u u; ):

Of course, the inner product space that we obtain this way is not a Hilbert space. We could complete it to be one, but the space that we would obtain would be too large to be 8

useful, at least in the H1 = C 1 -case. Finally, by means of the inner product de ned in (3.1), we can recast Problem 2.2. m Problem 2.20 : Given a linearly independent set fuj gN j =1  H?s (M ), complex numbers dj , j = 1; : : : ; N , and a positive de nite kernel  2 H2s(M m ) \ C 0 (M m ), nd u 2 U := Spanfu1 ; : : :; uN g such that [ u ; uj ] = dj ; for j = 1; : : :; N: Since we are assuming that  is strictly positive de nite, we knowPby Theorem 2.3 that N this problem is well poised. The unique solution to it is u = j=1 cj uj , where c = ( c1    cN )T 2 R N is the unique solution to Ac = d. Here, d = ( d1    dN )T . The interpolant one arrives at in this way is  ? u. The distribution u satis es the following minimization principle. Theorem 3.1: Let u 2 H?s (M m ) be the unique distribution in U for which ?u solves Problem 2.20. If v 2 H?s (M m ) is any other distribution for which ?v also interpolates the data in Problem 2.20|that is, [ v ; uj ] = dj ; for j = 1; : : :; N |then v ? u is orthogonal to Spanfu1 ; : : :; uN g with respect to the inner product (3.1). In addition, []v[]2 = []v ? u[]2 + []u[]2 : Finally, if v 6= u,

[]u[] < []v[]: Proof: Since both u and v interpolate the data, we have that [ v ? u ; uj ] = 0 for j = 1 : : :N ; thus, v ? u is orthogonal to U . The rest of the theorem follows from standard linear algebra. We remark that this is quite similar to the variational principle derived by Madych and Nelson [15, 16] in the radial basis function (RBF) case. (For a review of RBFs, see [21]). The last inequality in the theorem amounts to saying that among all v 2 H?s (M m ) for which ?v interpolates the data, the distribution u minimizes the norm (3.3). We now turn to the estimates that we mentioned earlier. We again suppose that u and v are as above, that the functions (3:4) f :=  ? v and f~ :=  ? u both interpolate the data, and that u 2 U is the unique solution to Problem 2.20. If we regard f as having generated the data in the rst place, we can ask how well the interpolant f~ approximates the original function f . Proposition 3.2: Let f and f~ be as in (3.4), and let w be an arbitrary distribution in H?s (M m ). If g and gw are de ned by (3:5) g := f ? f~ =  ? (v ? u) and gw := (w;  g) then (3:6)

jgw j  dist(v; U ) dist(w; U ); 9

where the distances are computed relative to the norm (3.3). Proof: As we noted earlier, v ? u is orthogonal to U = Spanfu1 ; : : :; uN g. Thus, if c1 ; : : : ; cN are arbitrary complex numbers. then

Z

gw := m w(p)g(p)d(p) M =[[v ? u ; w] (3:7)

=[[v ? u ; w ?

N X j =1

cj u j ]

Applying Schwarz's inequality to (3.7) yields (3:8)

jgw j  []v ? u[] []w ?

N X j =1

cj uj []:

Since v ? u is orthogonal to U , u is the orthogonal projection of v onto U . Hence, (3:9)

[]v ? u[] = dist(v; U ) :

Also, we have that (3:10)

dist(v; U ) = c1min ;:::;cN []w ?

N X j =1

cj uj []:

Coupling these facts with (3.8) yields (3.6). Remark 3.3: Formula (3.6) is a version of the hypercircle inequality [3, pg. 230]. The utility in (3.6) is that, because the right side separates into a product of a term depending only on g and a term involving only w, we can estimate the e ect of w independently of the original g. Another feature is that the distributions w and v appear symmetrically on the right side of (3.6), in factors that involve only distances to U . Thus our strategy for obtaining estimates on gw is to get upper bounds on the distance from a distribution to U , given that the norm we employ is (3.3). Remark 3.4: The distance dist(v; U ) = []v ? u[] depends on the interpolant  ? u. We can, however, bound dist(v; U ) independently of u. From Theorem 3.1, we see that (3:11)

dist(v; U ) = []v ? u[]  []v[];

which is independent of  ? u. Up to now in this section we have treated general SPD kernels. For the rest of it, the types of kernels that we will work with are ones having the form (1.1), with ak > 0 for 10

all k. By Corollary 2.5, these are strictly positive de nite kernels. Moreover, this class includes a wide variety of C k kernels, including the C 1 kernels treated in x3 of [19]. If we impose additonal restictions on the rate of decay of ak 's, then for these kernels we can obtain estimates on dist(w; U ) and []v[] in terms of more standard norms. We have the following result. Proposition 3.5: Let  and s be as in (1.1). If there is a t > s for which the ak 's satisfy a bound of the form (3:12) ak  c?1 (1 + k )?(s+t)=2 for all k and some constant c, then every f 2 Ht (M m ) can be written as f = ?v for some v in H?s (M m ). Moreover, (3:13)

kvk?s  ckf kt and []v[]  kk21s=2 ckf kt :

Proof: Use the eigenfunctions Fk to expand f , 1 X f (p) = fb(k)Fk (p): k=1

Let vb(k) = fb(k)=ak , and observe that for all k, (1 + k )?sjvb(k)j2  c2(1 + k )t jfb(k)j2 Summing both sides over all k  0 yields 1 X

From this, it follows that

k=1

(1 + k )?s jvb(k)j2  c2 kf k2t :

v(p) :=

1 X

vb(k)Fk (p)

k=1 m de nes a distribution in H?s (M ) that satis es

kvk?s  ckf kt . This inequality when

inequality []v[]  kk21s=2ckf kt .

coupled with (3:2) yields the Our next result sets up a framework that will enable us to discuss how well our interpolants approximate a given function. The idea is to estimate the distance dist(w; U ), given that the kernels used to generate the inner product on U are of the form (1.1). Proposition 3.6: Let M be a positive integer, let s > 0, and let  have the form (1.1) with aPk N> 0 for all k  1. If there are coecients c1 ; : : :; cN such that, for k = 1 : : :M , (w ? j=1 cj uj ; Fk ) = 0, and if there is a sequence bk > 0, k = M + 1; M + 2; : : :, for P which j(w ? Nj=1 cj uj ; Fk )j2  bk when k  M + 1, then (3:14)

dist(w; U )  []w ?

N X j =1

cj uj []  11

1 X k=M +1

ak bk

!1=2

:

Proof: From(3.3) and (2.3), we have that

(3:15)

N X

X X []w ? cj uj []2 = ak j(w ? cj uj ; Fj )j2 ; j =1 j =1 k=1 1

N

from which (3.14) follows immediately. Remark 3.7: It is not necessary to choose the ordering of the eigenfunctions to be that obtained from the labeling the eigenvalues in increasing order. In fact, the ordering is arbitrary, and may be chosen to be whatever is convenient. We can combine the results above to obtain a variety of estimates on how well the interpolant  ? u ts the original function  ? v. Perhaps the most useful case is when the function generating the data belongs to one of the Sobolev spaces. This is the situation that we address below. Theorem 3.8: Let  and s be as in (1.1), and let t > s satisfy the conditions in Proposition 3.5. If f is in Ht (M m ), if w is a distribution in H?s, and if f~ is the solution to Problem 2.20 for data generated by f , then

j(w;  f ? f~)j  kk12=s2ckf kt dist(w; U ): In addition, if w and the sequence fbk g1 k=M +1 satisfy the conditions of Proposition 3.6,

(3:16) then

(3:17)

j(w;  f ? f~)j  kk12=s2ckf kt

1 X k=M +1

ak bk

!1=2

:

Proof: From Proposition 3.5, we have that f =  ? v for some v in H?s (M m ). Since ~ f = ?u, we also have g = ? (v ? u), so (w;  f ? f~) = (w;  g) = gw . Thus, from Proposition 3.2, Remark 3.4, and (3.13), we obtain (3.16). Combining (3.16) with Proposition 3.6 then yields (3.17). There are a number of other estimates one can obtain using the results of this section. The ones given in Theorem 3.8 should be considered representative. We now turn to the case in which M m is the product manifold M m1     M md , and the kernels are of the form (2.5). Consider the space of distributions

(3:18)

U := U 1    U d ;

where each space U is a nite dimensional subset of H?s (M m ) and where s1 +    sd = s. Standard formulas involving tensor products apply to the inner product induced by the kernel (2.5). In particular, we have (3:19)

[ w1 w2    wd ; z 1 z 2    z d ] = 12

Yd

=1

[ w ; z ] ;

where w ; z are in H?s (M m ) and the inner products are of the form (3.1), with the kernel being  . Proposition 3.9: If w := w1    wd , where w 6= 0 is in H?s (M m ) for

= 1; : : :; d and if the space U is given by (3.18), then we have

v u Yd u t dist(w; U ) = []w[] 1 ? (1 ? dist(w ; U )2 =[]w []2 )

(3:20)

=1

and in addition (3:21)

dist(w; U )  []w[]

d dist(w ; U )2 !1=2 X

=1

[]w []2

Proof: Let u~ be the orthogonal projection of w onto U relative to the inner product [  ; ] , and set u~ = u~1    u~d . Note that, from (3.19) and the well-known properties

of orthogonal projections, we can easily show that [ w ? u~; u~] = 0 ; consequently u~ is the orthogonal projection of w onto the subspace U . This and the fact that u~ is the orthogonal projection of w onto U imply that dist(w; U )2 = []w[]2 ? []u~[]2 = []w[]2 = []w[]2

1? 1?

Yd

[]u~ []2 =[]w []2

=1 Yd 

=1

!

1 ? dist(w ; U )2=[]w []2

!

:

Equation (3.20) then follows from taking square roots inQthe previous equation. To establish P d d 2 (3.21), use induction on d to show that the inequality =1 (1 ? x )  1 ? =1 x2 ; holds for all x 's that satisfy 0  x < 1. Apply this to the right side of (3.20) to get (3.21). The quality of an approximation method is measured by the error estimates, by its stability and by the amount of computation it requires. It is known for methods based on radial functions that better rates of approximation are coupled with worse stability constants [23]. Numerical results indicate a similar trade-o in the cases where one deals with the torus or the sphere, and it should be possible to quantify such a trade-o in the situation described here. We will discuss speci c examples in sections 5 and 6. We now turn to dealing with kernels that are analogous to RBFs of order 1 or higher.

x4. Conditionally Positive De nite Kernels on Manifolds

In practical applications one often wants interpolants that reproduce exactly some xed, nite dimensional space of functions. For example, in euclidean spaces one would 13

like to use interpolants that reproduce exactly the space of polynomials of degree n ? 1 or less. On the circle, the functions to be reproduced exactly would be all trigonometric polynomials of degree less than n, and on the sphere they would be all spherical harmonics, again with degree less than some xed number. Trigonometric polynomials are linear combinations of eik 's and spherical harmonics are linear combinations of Y`;m 's (see x6); the eik 's and the Yj;k 's are eigenfunctions of the Laplace-Beltrami operators for the circle and the sphere, respectively. It is thus natural to consider interpolation problems on M m in which one wants interpolants that reproduce exactly a space of functions with a basis comprising nitely many eigenfunctions of the Laplace-Beltrami operator. To formulate and solve such problems, we will introduce a class of conditionally positive de nite kernels. Let I be a nite set of indices in Z+, and consider the nite dimensional space de ned by SI := SpanfFi : i 2 Ig. In addition, de ne the space of distributions SI? := fu 2 H?s (M m ) : (u; F ) = 0; F 2 SI g . We will say that a selfadjoint kernel  in H2s(M m  M m ) \ C 0 (M m  M m ) is conditionally positive de nite (CPD) with respect to a nite set of indices I  Z+ if for every u 2 SI? we have (u u; )  0: If the inequality above is strict for u 6= 0, then we will say that  is conditionally strictly positive de nite (CSPD) on M m . Since the proofs of elementary properties of the CPD and CSPD kernels are similar to those of the corresponding properties for PD and SPD kernels, in the propositions below we will list them without proofs. Proposition 4.1: If  is a conditionally strictly positive de nite kernel on M m , and if the set of distributions fvj gNj=1  SI? is linearly independent, then the interpolation matrix A having entries Aj;k := (vj ;  ? vk ); where j; k = 1; : : :; N; is self-adjoint and positive de nite. We remark that, as in the case of SPD kernels, one has a whole class of CSPD kernels arising from self-adjoint kernels of the form (1.1). Proposition 4.2: Let  2 H2s (M m  M m ) \ C 0 (M m  M m ) be a selfadjoint kernel of the form (1.1).  is a CPD kernel if and only if ak  0 for all k 62 I . In addition,  is a CSPD kernel if and only if ak > 0 for all k 62 I . We will refer to such kernels at the end of the section. For the present, we do not need to require that  be of the form (1.1). Proposition 4.3: Let  be a conditionally strictly positive de nite kernel on M m . P P If u and v are in H?s(M m ), and if uI = u ? i2I (u; Fi)Fi and vI = v ? i2I (v; Fi)Fi , then the Hermitian form (4:1) [ u; v] I := (vI uI ; ); de nes a semi-inner product on H?s(M m ), and an inner product on SI? . We will denote the associated semi-norm (or norm, if we are on SI? ) by (4:2)

[]u[]I := [ u; v] I1=2: 14

We now want to consider the following problem, which is analogous to Problem 2.2: m Problem 4.4: Given a linearly independent set fuj gN j =1  H?s (M ), complex numbers fdj gNj=1 , and a CPD kernel  in H2s(M m ) \ C 0 (M m ), nd u 2 UI := Spanfuj : j = 1; : : :; N g \ SI? and F 2 SI , such that (uj ;  ? u + F ) = dj ; for j = 1; : : :; N ; that is, the function f :=  ? u + F interpolates the data. Concerning this problem, we have the result below. Theorem 4.5: Let  be CSPD in H2s (M m ) \ C 0 (M m ), and let the matrix C with entries Cj;i = (uj ; Fi), where j 2 f1; : : :; N g and i 2 I , be of full rank. If rank(C ) = N , then Problem 4.4 has one or more solutions, all of the form f = F 2 SI . If the rank(C ) = jIj < N , then Problem 4.4 has a unique solution; moreover, when the data set fdj gNj=1 is generated by a function in SI , then this solution reproduces the function. Proof: Problem 4.4 can be reformulatedPin terms of a system P of equations for the unknowns b 2 C N and e 2 C jIj for which u = Nj=1 bj uj and F = i2I ei Fi . The system has the form (4:3)

Bb + Ce = d ; C  b = 0 :

where B is the matrix with elements (4:4)

Bj;k := (uj ;  ? uk ); j; k = 1; : : :; N:

The condition that C  b = 0 comes from requiring that u be in SI? . In the case N  jIj, we have rank(C ) = N ; this implies both that the only vector b satisfying the second condition in (4.3) is b = 0 and that there is a vector e 2 C jIj satisfying Ce = d : If N < jIj, then C has a non-trivial null space, and e will not be unique. InPthat case, there are in nitely many solutions to Problem 4.4, all of the form f = F = i2I ei Fi . For rank(C ) = jIj < N , the matrices C  and C have nullities N ? jIj and 0, respectively. We want to show that the rank of the system in (4.3) is N + jIj, from which it follows that the (4.3) can be uniquely solved for b and e. To nd the nullity of this system, let d = 0 in (4.3), and observe that C  b = 0Pimplies that b Ce = 0; hence, bBb = 0. On the other hand, bBb = (u u; ), with u = j bj uj . Taken together, these yield (u u; ) = 0. Since  is a CSPD kernel, u = 0. The linear independence of the uj 's then also gives us that b = 0. From this and (4.3), we then get Ce = 0, and because the nullity of C is 0, e = 0. Thus the nullity of (4.3) is 0. The statement concerning the reproduction of functions in SI follows from the existence and uniqueness result just proved. The solution of the generalized Hermite problem satis es a variational principle relative to the seminorm (4.2). Since the proof is straightforward, we omit it. 15

Theorem 4.6: Let  be a CSPD kernel in H2s (M m )\C 0 (M m ), let rank(C ) = jIj < N ,

and let f =  ? u + F be the unique solution to Problem 4.4. If (4:5) f~ =  ? u~ + F~ ; u~ 2 SI? and F~ 2 SI ;

also satis es (uj ; f~) = dj , j = 1; : : :; N , then u~ ? u is orthogonal to the subspace UI , relative to the inner product in (4.1). Moreover []u~[]2I = []u~ ? u[]2I + []u[]2I ;

(4:6)

from which one concludes that the solution to Problem 4.4 minimizes the seminorm (4.2) among all other functions that are of the form (4.5) and satisfy the interpolation conditions. We want to obtain an error estimate for distributions in H?s(M m ) applied to the di erence f~ ? f . That is, if w 2 H?s (M m ), we want bounds on (w;  f~ ? f ) similar to those given in x3. To do this, we will prove the following lemma. Lemma 4.7: Adopt the notation and assumptions of Theorem 4.6. If w 2 H?s (M m ), P N then there exists a set f j gNj=1  C for which w ? j=1 j uj 2 SI? . Moreover, for any such set (4:7)

(w;  f~ ? f ) = (w ?

N X j =1

j uj ; f~ ? f ) = [ u~ ? u; w ?

N X j =1

j uj ] I :

Proof: Since f~ and f both satisfy the interpolation conditions in Problem 4.4, we have that (uj ; f~?P f ) = 0 for j = 1; : : :; N . Consequently, for every choice of j 's, P we see that N ~ ~ (w;  f ?f ) = (w ? j=1 j uj ; ?(~u?u)+F ?F ). If we can nd j 's such that w ? Nj=1 j uj P is in SI? , then, since F~ ? F is in SI , we would have (w ? Nj=1 j uj ; F~ ? F ) = 0, and so

(w;  f~ ? f ) = (w ?

N X j =1

j uj ;  ? (~u ? u)):

P

From this equatiion, the de nition of [  ;  ] I , and that both u~ ? u and w ? Nj=1 j uj are in SI? , we obtain (4.7). Thus, the whole result will follow if we can show the existence of j 's for which P w ? Nj=1 j uj is in SI? . Doing this is equavalent to solving the system of equations below for all i 2 I : N X (w; Fi) = j |(uj{z; Fi}) = [C  ]i : j =1 Cj;i Solving this is possible because rank(C ) = jIj. If we combine the results in Theorem 4.5, Theorem 4.6, and Lemma 4.7, we easily obtain the following variant of the hypercircle inequality [3, pg. 230]: 16

Corollary 4.8: With the notation and assumptions of Theorem 4.6, we have

X j(w;  f~ ? f )j  dist(~u; UI ) dist(w ? j uj ; UI ); N

j =1

where distances are computed relative to the norm (4.2). Remark 4.9: With the exception of results that make speci c use of kernels of the form (1.1), all the results in this section hold if we regard SI as a nite dimensional subspace of smooth functions and the set fFi : i 2 Ig is interpreted as an orthonormal basis for SI . Remark 4.10: With minor modi cations, most results invoving Problem 2.2 and based on Proposition 3.6 will also be true for Problem 4.4, provided the CSPD kernels employed have the form (1.1).

x5. The Torus In this section and section 6 below, we will apply Proposition 3.6, Theorem 3.8, and other results from previous sections to several types of interpolation problems. For the circle and higher dimensional tori, we will obtain rates for scattered data point-interpolation and for a restricted class of Hermite problems that involve interpolating a xed linear combination of function values and derivatives at scattered sites. In the next section we will deal with a point-evaluation problem for the 2-sphere. To apply Proposition 3.6 in any of these cases requires two things. First, one must show that the coecients fcj gNj=1 exist. Second, one needs to bound the sequence fbk g so that the estimates in (3.14) and (3.17) are useful. Doing these two things is in essence our goal here and in the next section. In all cases the rates that we obtain re ect the smoothness of both the underlying data-generating function and the functions in the approximating subspace. Let Tm be the m-torus. Consider a kernel of the form (5:1)

(; 0 ) = m =1P ( ? 0 );

where  and 0 are standard periodic coordinates. Suppose that for some s  0 each P 2 H2s(T1 ) is continuous and has the Fourier series expansion (5:2)

1 X a k eik ; where a k > 0: P () = p1 2 k=?1

Noting that the eigenfunctions for Tm are just the functions Fk1 ;:::;km (1 ; : : :; m ) = (2)?m=2ei(k1 1 +km m ) and using Corollary 2.6, one sees that the kernel (; 0 ) is strictly positive de nite, continuous, has the form (1.1), and also belongs to H2s(Tm ). C 1 kernels of this type were described in [19, x5]. 17

The simplest case is thepcircle, T1 . The kernel in (5.1) is then a single function, P ( ? 0 ), and Fk () = eik = 2. We want to apply the estimates from section 3 to the Hermite problem 2.20 in which the distributions uj are (5:3)

uj () :=

 X

n (n) ( ? j ); j = 0 : : :N ? 1;

n=0

where the j 's are distinct angles in the interval [0; 2). To get rates of approximation, we will have to put further restrictions on these angles; we will do this later. Also, we have labeled the distributions in the natural way, j = 0 : : :N ? 1 rather than j = 1 : : : N . In (5.3),  is some nite integer, and the coecients n are in C . In addition, we require (5:4)

 < s ? 1=4; 0 6= 0;  6= 0; and

 X

n=0

n in kn 6= 0 for all k 2 Z:

For later use, we remark that these conditions imply that 2 =2 (5:5) 0 < c?U 1  j P(1 + k )in kn j  CU < 1 for all k 2 Z: n=0 n To solve the Hermite problem using the distributions uj , we need P to be at least 2 times continuously di erentiable. From Lemma 1.2, we can achieve this if we assume s >  +1=4; this is the reason for the condition  < s ? 1=4. As in section 3, we will take U to be the span of the uj 's. We remark that the basis functions that we use to solve our Hermite problem are Pj () :=P ? uj () = (5:6)

=

Z 2 0

 X n=0

P ( ? 0 )uj (0 )d0

n P (n) ( ? j )

Our basic tool will be Proposition 3.6. Using it requires the following two lemmas. Lemma 5.1: Let the uj 's be as in (5.3), with the j 's being distinct angles in the interval [0; 2), let the cj 's be arbitraryPcomplex numbers, and suppose that (5.4) holds. If w 2 H?s (T1 ) and if Wk := (w; Fk ) ? jN=0?1 cj (uj ; Fk ); then for all k 2 Z we have (5:7)

! )

(

 X n in kn c~k ; Wk = p1 (w; e?ik) ? 2 n=0

where (5:8)

c~k =

NX ?1 j =0

cj e?ikj

18

Moreover, there exist unique scalars c1 ; : : :; cN such that Wk = 0 for all integers k 2 JN := [?[N=2]; [(N ? 1)=2] ] \ Z . Proof: We have the following chain of equations:

p

Wk =(w; e?ik= 2) ?

p =(w; e?ik= 2) ?

NX ?1 j =0 NX ?1

p

cj (uj ; e?ik = 2) cj

  (n) X

p 

n  ( ? j ); e?ik = 2

j =0 n=0 8 19 ! 0NX ? 1  = < X 1 ? ik ? ik n n j A @ cj e n i k = p :(w; e ) ? ; 2 n=0 j =0

Combining this and (5.8) yields (5.7). For k 2 JN , we wish to solve the system of equations NX ?1

(5:9)

j =0

?ik cj e?ikj = P(w; e in)kn : n=0 n

The determinant of the coecient matrix for this system is a Vandermonde determinant. Since the j 's are distinct angles in [0; 2), the exponentials e?ij correspond to distinct points on the unit circle. Inspecting the usual formula for a Vandermonde determinant shows that it is not zero; the system thus has a unique solution. With this choice of cj 's, we have ?ik (5:10) c~k = P(w; e in)kn ; for all k 2 JN : n=0 n Inserting this expression in (5.7) then yields Wk = 0 for k 2 JN . Lemma 5.2: With the notation and assumptions of Lemma 5.1, if k 2= JN , we have (5:11)

jWk

j  (1 + k2 )s=2





cU jc~k j : kwk?s + p 2(1 + ([N=2])2)(s? )=2

In addition, if the cj 's are taken to be the unique coecients for which Wk = 0, k 2 JN , then (5:12)

p

sup jc~k j  2CU (1 + [N=2]2)(s? )=2kwk?s:

k2JN

Proof: The distribution w is in H?s (T1 ); its Fourier coecients thus satisfy a bound of

the form (5:13)

p j(w; e?ik )j  2kwk?s (1 + k2 )s=2; k 2 Z ; 19

If we combine (5.13) with (5.7) and (5.5), we arrive at the estimate (5:14)

jWk

j  (1 + k2)s=2





kwk?s + p cU jc~2k j(s? )=2) : 2(1 + k )

Note that s >  and that if k 2= JN , then jkj  [N=2]. Consequently, (1 + k2)(s? )=2  (1 + ([N=2])2)(s? )=2. Using this in (5.14) yields (5.11). Next, take the cj 's to be those for which Wk = 0 holds when k 2 JN . The c~k 's then satisfy (5.10). If we combine (5.10), (5.5), and (5.13), we get the following inequality,

p jc~k j  2CU kwk?s(1 + k2)(s? )=2; for all k 2 JN : Again, since s >  and jkj  [N=2] if k 2 JN , we see that (1 + k2)(s? )=2  (1 + (5:15)

([N=2])2)(s? )=2. Employing this in conjunction with (5.15) then yields (5.12). Thus far, we have made no real restrictions on the j 's used in the uj 's. At this point, we will assume that these angles are distributed quasi-uniformly in the following sense. Speci cally, let (5:16) j = 2N (j + "j ); where the j 's are real numbers that satisfy (5:17)

sup jj j = L; 0  L < 1=4: j

Using a discrete analogue of a theorem of Kadec [32, p. 43], we are able to estimate the distance from w to U , provided the norm (3.3) is used to compute the distances. Theorem 5.3: Let the uj 's be as in (5.3), with the j 's being given by (5.16), and P 1 let JN be as in Lemma 5.1. If P 2 H2s(T ), and if k2Z(1 + k2 )sak converges, then, for any w 2 H?s (T1 ), we have (5:18)

0 11=2 X dist(w; U )  kwk?s (L; N ) @ (1 + k2 )sak A ; k=2JN

where the ak 's are the Fourier coecients of P , and  1; if L = 0; (5:19) (L; N ) := 1 + cU CU pN=2 csc(  ? L) if 0 < L < 1=4. 4 Here, cU and CU are given in (5.5). Proof: By Lemma 5.1, there are unique cj 's for which Wk = 0 when k 2 JN . With this choice of cj 's, when k is in JN , c~k is given by (5.10) and satis es (5.12). There are now two cases. If L = 0 in (5.17), then the j 's are equally spaced angles, and c~k = c^k , 20

the discrete Fourier transform of the cj 's. Since the c^k 's are periodic in k with period N , we see that supk2Z jc~k j = supk2JN jc~k j. The second case is the one in which 0 < L < 1=4. From Theorem A.2, we have p jc~k j  N=2 csc( 4 ? L) sup jc~j j; k 2= JN (0 < L < 1=4): j 2JN Combining the estimates for jc~k j from the two cases and using (5.19), we get

jc~k j  (L; N ) sup jc~j j; k 2= JN : j 2JN

Now use (5.12) to replace supj2JN jc~j j in the inequality above, and then insert the result in (5.11) to get (5:20)

jWk j  (1 + k2 )s=2kwk?s (L; N ); k 2= JN (0  L < 1=4):

Clearly, Proposition 3.6 and Remark 3.7 now apply, provided that in (3.14) we take bk to be the square of the right side of (5.20). Doing so then yields (5.18). Suppose that u 2 U is the unique solution to Problem 2.20 , given that the data is generated by a function f :=  ? v, where v 2 H?s(T1 ). Here,  is of the form (5.1), P is as in (5.2), and the uj 's are as in (5.3). In addition, the interpolant generated is f~ := ?u. We obtain the following: Corollary 5.4: With the notation used above and the assumptions from Theorem 5.3, if w 2 H?s (T1 ), then (5:21)

0 1 X j(w; f ? f~)j  kvk?skwk?s (L; N )2 @ (1 + k2 )sak A k=2JN

Proof: Combine Proposition 3.2 and Theorem 5.3 to get (5.21).

We now wish to obtain a result for the case in which f is in Ht(T1 ), where t > s. P To do this, we will assume that (3.12) P holds. If we also assume that the series k2Z(1 + 2 s k ) ak converges, then we must have k2Z(1 + k2 )(s?t)=2 be convergent, which implies the additional constraint that t has to be larger than s + 1. Corollary 5.5: Let t > s +1 and let f 2 Ht (T1 ) have f~ as the interpolant described P above, and suppose that k2Z(1+ k2 )sak converges. With the assumptions from Theorem 5.3, if (3.12) holds for the ak 's and if w 2 H?s(T1 ), then (5:22)

0 1 X j(w; f ? f~)j  ckf kt kwk?s(L; N )2 @ (1 + k2 )sak A k=2JN

where c is as in (3.12). 21

Proof: Since (3.12) holds and since t > s + 1 > s, Proposition 3.5 applies. Thus, we see that f =  ? v, with kvk?s  ckf kt. Using this in (5.21) yields (5.22). Remark 5.6: In Corollary 5.5, the right side of (5.22) is the tail end of a convergent series, so one would like to have the ak 's decay as quickly as possible. On the other hand, (3.12) limits the rate at which the coecients can decay. Now, the rate at which the ak 's decay roughly translates into the smoothness class to which P belongs. Let s0  s. With a little work one can show that if P 2 Hs+s0 (T1 ), then s  s0 < t ? 1=2. We also remark that it should be true that if w 2 H?r , r < s, then one should get better estimates. Unfortunately, the Hilbert space techniques used here do not provide a suitable means for establishing this. Finally, we note the meshsize hN = max0jN ?1 (j+1 ? j ), where j is given by (5.16) and N := 0 + 2, satis es =N  hN  3=N . Consequently, the estimates we derive can be related to classical estimates employing the meshsize. We now turn to higher dimensional tori, Tm , with m  2. We will use a kernel of the form (5.1), and our space of distributions will be (5:23) U := U 1    U m ; where each space U is a nite dimensional subset of H?s (T1 ) and where s1 +    sm = s. In addition, we will assume that each U is spanned by N distributions of the form (5.3), with angles involved being not only distinct but also of the form (5.16), with N replaced by N , and the corresponding j 's there satisfying (5.17) with L replaced by L . P Corollary 5.7: If k2Z(1 + k2 )s a k converges for = 1; : : :; m, and if w := w1    wm , where w 6= 0 is in H?s (T1 ), then (5:24) ! 0X 11=2 m  (L ; N )2  X m Y

(1 + k2)s a k A ; dist(w; U )  k k2s kw k?s @ 2 k  k

2s

=1

=1 k=2IN

where the a k 's are the Fourier coecients of P , the set IN is as in Lemma 5.1, and the factor (  ;  ) is given in (5.19). Proof: By theorem 5.3, we have that

0 1 X (1 + k2)s a k A : dist(w ; U )2  kw k2?s (L ; N )2 @

From equation (3.19), we also see that []w[]2 =

k=2JN

m Y =1

[]w []2 :

Multiply both sides of the previous inequality by []w[]2 =[]w []2 and use the last equation to obtain 0 1m 2

2 X []w[] dist(w ; U )  kw k2 (L ; N )2 @ 2 )s a A Y []w []2 : (1 + k ?s k []w []2 6= k=2JN 22

Recall that

[]w []2  k k22s kw k2?s : Combining this with the previous inequality yields []w[]2dist(w ; U )2 []w []2



(L ; N )2 k k22s

0 10 m 1 X Y @ (1 + k2)s a k A @ k k22s kw k2?s A k=2JN

=1

The inequality (5.24) follows from using the inequality above in conjunction with (3.20) of Proposition 3.9. We can now use this result to obtain rate estimates for the case of Tm . Theorem 5.8: With the notation and assumptions of Corollary 5.7, if u 2 H?s (Tm ) is the unique distribution in U for which  ? u solves Problem 2.20, and if v 2 H?s (Tm ) is any other distribution for which  ? v also interpolates the data in Problem 2.20, then (5:25)

(5:26)

j(w;  f ? f~)j  Ks kvk?s

m Y

=1

kw k?s

!

0m 11=2   2 X X  @ (kL ;kN2 ) (1 + k2 )s a k A

2s

=1 m Y where Ks := kk2s k k2s

=1

k=2IN

Proof: By Remark 3.4 and the inequality []v []  kk2s kv k?s , we see that dist(v; U )  kk2skvk?s . The result then follows on combining this with Proposition 3.2 and Corollary

5.7.

If each of the P 's in the product (5.1) has coecients that satisfy a bound of the form (5:27)

a k  c? 1 (1 + k2 )?(s +t )=2

Q a satis es then ak1 ;:::;km = m

=1 k

m X ? 1 ak1 ;:::;km  c (1 + k 2 )?(s+t)=2 = c?1 (1 + k1 ;:::;km )?(s+t)=2

=1

Q c , s = Pm s , and t = Pm t . This coupled with Theorem 3.8 and where c = m

=1

=1

=1 Theorem 5.8 yields the following result. 23

Corollary 5.9: If the a k 's satisfy (5.27), and if f =  ? v is in Ht (Tm ), then we

may replace (5.25) by (5:28)

j(w;  f ? f~)j  Ks ckf kt

m Y

=1

kw k?s

!

11=2 0m   2 X X (1 + k2 )s a k A :  @ (kL ;kN2 )

=1

2s

k=2IN

We close this section by remarking that the higher dimensional tori estimates we obtained in Theorem 5.8 and Corollary 5.9 are not optimal. We conjecture that estimates similar to ones for the circle can be found.

x6. The 2-Sphere

Our next example deals with the 2-sphere, S2. The interpolation problem that we will deal with in this section will be one in which the the distributions in Problem 2.20 are point evaluations at points pj;k that we will describe below; that is,

uj;k = pj;k and U := Spanfuj;k g The results we obtain in this section depend on recent results of Driscoll and Healy [4], and we will adopt the convention for spherical coordinates that is used by them and that is customary in physics: the angle  2 [0; ] is measured o the positive z-axis and the angle  2 [0; 2) is measured o the x-axis in the x-y plane. In addition, take  to be a xed positive integer, then let 8 j ; j = 0; : : :; 2 ? 1; > j = 2 > < (6:2) ; k = 0; : : :; 2 ? 1; k = k >  > : (j ; k ) = coordinates of pj;k :

(6:1)

The eigenfunctions of the Laplace-Beltrami operator for S2 are the Y`;m's, where ` = 0; 1; 2; : : : and where, for ` xed, m = ?`; : : :; `; these functions are described in detail in [4(x2),18, 19], and in spherical coordinates are related to associated Legendre functions. We will not need their explicit form here, although we will need some of their properties. The Y`;m's form an orthonormal basis for L2(S2) with respect to the standard measure on the sphere, d, which has the form d = sin()dd in the coordinates above. As in [4], we will denote the expansion coecient for a distribution h relative to Y`;m by bh(`; m); that is, set Z bh(`; m) = hY `;md: S2

We will say that h is band-limited if the orthogonal expansion for h contains only nitely many terms. Driscoll and Healy have established the following \sampling theorem." 24

Theorem 6.1: [4, Theorem 3] If f (p) is a band-limited function on S2 for which fb(`; m) = 0 for `  , then there exist coecients () j that are independent of f such that p 2X?1 2X?1 2 () fb(`; m) = 2 j f (pj;k )Y `;m (pj;k ) j =0 k=0

for ` <  and jmj  `. We are now ready to proceed with the \rate of approximation" estimates, which we will obtain by applying Proposition 3.6. To apply this proposition, we will need the lemmas below. Lemma 6.2: Let w 2 H?s (S2) and let the expansion coecients for w be wb(`; m) = (w; Y `;m ). If w is the band-limited function de ned by (6:3)

w(p) :=

and if

` X

`=0 m=?`

wb(`; m)Y`;m(p);

p

2 w (p )  ; 0  j; k  2 ? 1 cj;k := 2  j;k j

(6:4) then (6:5)

X ?1

wb(`; m) ?

2 X?1 2X?1

j =0 k=0

cj;k (uj;k ; Y `;m) = 0; 0  ` < ; jmj  `

Proof: By Theorem 6.1,

p

2 X?1 2X?1 2  w (pj;k ) j Y`;m (pj;k ); 0  ` < ; jmj  `: wb(`; m) = 2 j =0 k=0

Since Y`;m(pj;k ) = (uj;k ; Y`;m), choosing the cj;k 's to be the coecients given in (6.4) results in (6.5). Next, we need to estimate various quantities related to ones appearing in the lemma above. For all ` and m, de ne (6:6)

p

?1 2 X?1 2 2X W`;m := wb(`; m) ? 2 w (pj;k ) j (uj;k Y `;m ) ; j =0 k=0

where the j 's are the coecients from Theorem 6.1. By the previous lemma, we see that W`;m = 0 for all 0  ` < ; jmj  `. The results below provide a bound on this quantity in the case where `  . 25

Lemma 6.3: If w (p) is given by (6.3) and if   1, then for all p 1+s

jw (p)j  p kwk?s: 4

(6:7)

In addition, for all `  , we have (6:8)

p2 2X?1 2X?1 r 2` + 1 2?1  s +1 kwk X j  j: w ( p ) ( u   Y ) ?s j 2 j=0 k=0  j;k j j;k `;m 8 j =0

Proof: From Schwarz's inequality applied to the right-hand side of (6.3), we see that

(6:9)

jw (p)j2 

X ?1

` X

`=0 m=?`

jwb(`; m)j

! X ?1 X ` 2

`=0 m=?`

!

Y`;m (p)Y`;m(p) :

The rst term on the right above satis es this chain of inequalities: X ?1

` X

`=0 m=?`

jwb(`; m)j2 =

X ?1

` X

`=0 m=?`

j2 (1 + `(` + 1))s (1 j+wb(``;(`m+)1)) s

X ?1 X ` jwb(`; m)j2 s  (1 + ( ? 1)) s `=0 m=?` (1 + `(` + 1))  2s kwk2?s

(6:10)

The second term on the right of (6.9) can be evaluated exactly via the Addition Theorem for spherical harmonics [18, p. 10]; the result is this: X ?1

` X

X p  p); Y`;m(p)Y`;m(p) = 2`4+ 1 P` (|{z} ?1

`=0 m=?`

`=0

1

where P` is the standard LegendrePpolynomial of degree `, which satis es the normalization ?1 (2` + 1) = 2 , we see that condition that P` (1) = 1. Since `=0 (6:11)

X ?1

` X

`=0 m=?`

2 Y`;m (p)Y`;m(p) = 4

Taking sqare roots after combining (6.9), (6.10), and (6.11) yields (6.7). To establish (6.8), rst note that from (6.7) we have

p 2?1 2?1 2 X X X?1 2X?1  s 2 w (pj;k ) j (uj;k ; Y `;m)   kpwk?s j j jj(uj;k Y `;m )j (6:12) 2 8 j =0 k=0 j =0 k=0 26

As we noted in the previous proof, (uj;k ; Y `;m) = Y `;m (pj;k ). Also, by [18, Lemma 8, p. 14], we have r jY`;m(p)j  2`4+ 1 : Combining these two facts and using them in conjunction with (6.12), we nd that

p 2?1 2?1 2 X X w (p ) (u ; Y )  skwkp?sp2` + 1 2X?1 2X?1 j j j 2 j=0 k=0  j;k j j;k `;m 32 j =0 k=0 r 2 X?1  2`8+ 1 s+1 kwk?s

P

j =0

j j j :

We next wish to estimate the sum j2=0?1 j j j. This we can do when the integer  is a power of 2. Lemma 6.4: If  is a power of 2, then 2 X?1

(6:13)

j =0

j j j  log2 (16):

Proof: From [4, p. 216], if  is a power of two, then

p

 





?1 1 2 2 sin j X j (2` + 1) ; j = 0; : : :; 2 ? 1: sin j 2 2 `=0 2` + 1 2 From this it follows that, for j = 0; : : :; 2 ? 1,

 =

p X 2 ?1 1

j j  2 j

2 `=0 2` + 1 Z ?1 dx ! 2   1+ 2x + 1 0 p   2 1   1 + 2 ln(2 ? 1) 1 log (16):  2 2

p

(16) and so Consequently, j j j  log22 2 X?1

j =0

(16) (2) = log (16): j j j  log22 2 27

We can now bound the W`;m 's de ned in (6.6). The estimate that we obtain will allow us to use Proposition 3.6 to get rates of approximation. Lemma 6.5: Let  be a power of 2. If `  , then (6:14)





jW`;mj  1 + 2p1   23 log2 (16) (1 + `(` + 1))s=2kwk?s

Proof: Note that, since w is in H?s , its Fourier expansion coecients satisfy

jwb(`; m)j  (1 + `(` + 1))s=2kwk?s This provides a bound for the rst term in (6.6). A bound for the second term in (6.6) may be obtained by combining (6.8) and (6.13). This yields the following bound on W`;m.

r

jW`;mj  (1 + `(` + 1))s=2kwk?s + 2`8+ 1 s+1 kwk?s log2 (16): Rearranging terms on the right above, we have (6:15)

jW`;mj  (1 + `(` + 1))s=2kwrk?s(1 + As(`; ))

where As(`; ) = (1 + `(` + 1))?s=2 2`8+ 1 s+1 log2 (16):

With a little algebra, one sees that

  s? 21 3 1 1 As(`; ) = 2p (1 + 4 (` + 2 )?2)? 2s  32 log2 (16): 1 `+ 2

Since `  , it follows that

As(`; )  2p1   23 log2 (16): Using this bound in (6.15) results in (6.14). We are now ready to apply Proposition 3.6. In what follows, we assume that the SPD kernel  2 H2s(S2  S2) has the form (6:16)

(p; q) :=

1 X ` X `=0 m=?`

a`;mY`;m(p)Y `;m (q);

where the coecients satisfy a`;m > 0 for all ` and m. We will now prove the estimate on the distance from a distribution w to the subspace U . 28

Theorem 6.6: Let w be and  be as in (6.16). If  is a P in HP?`s, U (1be+as`(in` +(6.1), s a`;m is convergent, then power of 2, and if the series 1 1)) `=0 m=?`





!1=2

1 X ` X 1 3 (1 + `(` + 1))sa`;m (6:17) dist(w; U )  1 + p  2 log2 (16) kwk?s `= m=?` Proof: Combine Lemmas 6.2 and 6.5 with Proposition 3.6. We can now use this result to obtain rate estimates for the case of the 2-sphere, S2. Corollary 6.7: With the notation and assumptions of Theorem 6.6, if u 2 H?s (S2) is the unique distribution in U for which  ? u solves Problem 2.20, and if v 2 H?s(S2) is any other distribution for which  ? v also interpolates the data in Problem 2.20, then for all w 2 H?s (6:18) ! 2  1 3 1 X ` X (1 + `(` + 1))sa`;m j(w;  f ? f~)j  1 + 2p  2 log2(16) kvk?skwk?s `= m=?` Proof: By Proposition 3.2,

j(w;  f ? f~)j  dist(v; U )dist(w; U ): Both w and v are in H?s . Thus, (6.18) follows on applying the distance estimates in Theorem 6.6. We now wish to obtain a result for the case in which f is in f 2 Ht (S2), where tP> s.PTo do this, we will assume that (3.12) holds. If we alsoPassume that the series 1 1 ` s `=0 m=?` (1 + `(` + 1)) a`;m converges, then we must have `=0 (2` + 1)(1 + `(` + 1))?(t?s)=2 be convergent, which implies the additional constraint that t has to be larger than s + 2. Corollary 6.8: Let t > s +2, and adopt the notation of Corollary 6.7. If f 2 Ht (S2) and if (3.12) holds, then (6:19) !  1 3 2 X 1 X ` j(w; f ? f~)j  ckf ktkwk?s 1 + 2p  2 log2 (16) (1 + `(` + 1))sa`;m `= m=?` where c is as in (3.12). Proof: Since (3.12) holds and since t > s + 2 > s, Proposition 3.5 applies. Thus, we see that f =  ? v, with kvk?s  ckf kt. Using this in (6.18) yields (6.19). Remark 6.9: As in the the case of the circle (see Remark 5.6), one would like to have the a`;m's decay as quickly as possible, but again (3.12) limits the rate at which these coecients can decay. This roughly translates into limiting the smoothness class to which  belongs. If s0  s, then one can show that if  2 Hs+s0 (S2  S2), then s  s0 < t ? 1. 29

Appendix

Consider angles of the form (A:1) j = 2N (j + "j ); i = 0; : : :; N ? 1; j"j j  L < 1=4: Recall that in (5.8), we de ned (A:2)

c~k :=

NX ?1 j =0

cj e?ikj

for all k. Let JN := [?[N=2]; [(N ? 1)=2] ] \ Z . From the proof of Lemma 5.1, we can solve the equations above to get the cj 's in terms of the c~k 's, with k 2 JN . We wish to get estimates on the 2-norm of the column matrix c := ( c0    cN ?1 )T in terms of the 2-norm of c~ := ( c~?[N=2]    c~[(N ?1)=2] )T . We remark that, if c^ := ( c^?[N=2]    c^[(N ?1)=2] )T , where (A:3)

c^k =

NX ?1 j =0

cj e?

2ikj

N

is the discrete Fourier transform of the cj 's, then c~ = c^. The technique that we use in this section is a modi cation of the one employed by Young to prove a version of a famous theorem of Kadec [32, x1.10]. We begin with the observation that for k 2 JN , (A:4)

2i"j k

2ijk

e?ikj = e? N (1 ? (1 ? e? N )) 2i"j k 2ijk 2ijk = e? N ? e? N (1 ? e? N )

From this equation and (A.3), we can write c~k in the form (A:5)

c~k = c^k ?

For k 2 JN , set in the expansion [32, p. 43]

NX ?1 j =0

cj e?

2ijk

N

(1 ? e?

2i"j k

N

):

t = ? 2Nk and  = "j

1 ? eit = (1 ? sin  ) +

1 2(?1)m  sin( ) X

2 ?  2 ) cos(mt)   ( m m=1 1 (?1)m 2 cos( ) X sin((m ? 12 )t); +i 1 2 2 m=1  ((m ? 2 ) ?  )

30

and insert the result in (A.5): NX ?1





2ijk c~k = c^k ? cj e? N 1 ? sin(""j ) j j =0 1 NX ?1 m X "j  ) ? 2(?1) cos(2km=N ) cj e ?2Nijk "jmsin( 2 ? "2 j m=1 j =0 ?1 ?2ijk " cos("  ) 1 2i(?1)m+1 sin ?2k ?m ? 1  =  NX X 2 cj e N ? j 1 2j 2 ? N m ? 2 ? "j m=1 j =0

De ne the quantities (A:6)

This puts c~k into the form

c~k = c^k ?

NX ?1 j =0

8 > Aj = 1 ? sin(""j ) > j > > < m "j sin("j ) B j = m2 ? " 2 > j > > "j ) > : Cjm = (m"?j cos( 1=2)2 ? "2 j

cj Aj e?

2ijk

N

1 m e? ijk N A ? c B j j  m=1 j =0 0 ?1 1 1 2(?1)m+1 i sin ? 2k (m ? 1 ) NX X ijk N 2 @ cj Cjm e? N A ? 0 ?1 1 2(?1)m cos ? 2km  NX X N @

2



m=1

2

j =0

By the convolution theorem, we have c~k = c^k ? N1 (^c  A^)k 1 2(?1)m  2km  (^c  B^ m ) X k ? cos  N N

? Setting

m=1 1 2(?1)m+1 i X

m=1



 2k

 (^c  C^m) 1 k: sin N (m ? 2 ) N





(^c  B^ m)k B^km = (?1)m cos 2km N N  (^c  C^ m)  2k 1 k; ^Ckm = (?1)m+1i sin ( m ? ) N 2 N 31

we may rewrite the previous equation as (A:7) c~ = c^ ? T c^ = (I ? T )^c where the linear transformation T : C N ! C N is de ned via 1 2 X (A:8) T c^ = N1 (^c  A^) + (B^m + C^m ):  m=1

We wish to estimate kT k2;2. By the triangle inequality, we have 1 2 X 1 ^ kT c^k  N kc^  Ak +  (kB^m k + kC^m k): m=1 p Recall that kX^ k2 = N kX k, so ^ k c^ N A k = kF [   cj Aj   ]k2 p : = N k    cj Aj   k2 p  N kck2kAk1 = kc^k2kAk1 Similarly, ^m kB^m k  k c^ NB k  kc^k2 kB mk1 ^m kC^ m k  k c^ NC k  kc^k2kC m k1 We thus have the following estimate on T c^: 1 2 X k T c ^ k 2 m k1 + kC m k1 ) : (A:9)  k A k + ( k B 1 kc^k2 m=1  From the de nitions of Aj ; Bjm; Cjm in (A.6), we have 8 > kAk1  1 ? sin(LL) > > < m L) k B k1  Lmsin( (A:10) 2 ? L2 > > L) > : kC mk1  (mL?cos( 1 )2 ? L 2 2 Combining (A.9) and (A.10), we get 1 2L sin(L) kT c^k2  1 ? sin(L) + X 2 2 kc^k2 L m=1  (m ? L ) 1 X 2L cos(L) : + 2 2 m=1  ((m ? 1=2) ? L ) 32

The series on the right are known expansions:

8 > > < 1 X > > :

1 X m=1

2L

(m2 ? L2)

1 ? cot(L); = L

2L 2 ? L2 ) = tan(L ):  (( m ? 1 = 2) m=1

Using these above gets us (A:11)

kT c^k2  1 ? cos(L) + sin(L) = 1 ? p2 sin(  ? L) kc^k2 4

We state this result as a lemma: Lemma A.1: If L < 1=4, then

p kT k2;2  1 ? 2 sin( 4 ? L) < 1: The point is that we can now invert the linear transformation in (A.7) and estimate the norm of I ? T . When L < 1=4, we have c^ = (I ? T )?1c~. Since kT k < 1, the standard Neumann expansion gives

(A:12)

kc^k2  1 ?kkc~Tk2k 2;2 p ( 2=2) csc( 4 ? L)kc~k2 ;

which leads to the following result. Theorem A.2: If L < 1=4 and c~ = ( c~?[N=2]

(A:13)

   c~[(N ?1)=2] )T , then for all k 2= JN

p p jc~k j  2=2 csc( 4 ? L)kc~k2  N=2 csc( 4 ? L) sup jc~k j: k2JN

p

Proof: From (A.2), Schwarz's inequality, and kX^ k2 = N kX k, we have that for all k 2= JN p jc~k j  N kck2 = kc^k2:

Combining this inequality with (A.12) then yields the p left inequality in (A.13). The right inequality follows from the observation that kc~k2  N kc~k1 . 33

Acknowledgments

The work of F. J. Narcowich was supported by Air Force Oce of Scienti c Research Grant F49620-95-1-0194, and that of J. D. Ward was supported by National Science Foundation Grant DMS-9505460 and by Air Force Oce of Scienti c Research Grant F4962095-1-0194. Nira Dyn School of Math. Sciences Tel-Aviv University Tel-Aviv 69978 ISRAEL

Addresses

Francis J. Narcowich Dept. of Mathematics Texas A&M University College Station, TX 77843 USA

References

Joseph D. Ward Dept. of Mathematics Texas A&M University College Station, TX 77843 USA

[1] P. H. Berard, Spectral Geometry: Direct and Inverse Problems, Springer-Verlag, Berlin, 1980. [2] Yu. M. Berezanskii, \A generalization of a multidimensional theorem of Bochner," (English) Soviet Math. 2 (1961) 143-147. [3] Philip J. Davis, Interpolation and Approximation, Dover Publications, New York, 1975. [4] J. R. Driscoll and D. M. Healy, \Computing Fourier Transforms and Convolutions for the 2-Sphere," Advances in Applied Math. 15 (1994), 202-250. [5] J. Duchon, \Interpolation des Fonctions de Deux Variables Suivant le Principe de la Flexion des Plaques Minces," Rairo Analyse Numerique 10 (1976), 5-12. [6] J. Duchon, \Splines minimizing rotation invariant semi-norms in Sobolev spaces", pp. 85-100 in Constructive Theory of Functions of Several Variables, Oberwolfach 1976, W. Schempp and K. Zeller, eds., Springer-Verlag, Berlin, 1977 [7] W. Freeden, \On Spherical Spline Interpolation and Approximation," Math. Meth. in the Appl. Sci. 3 (1981), 551-575. [8] F. G. Friedlander, The wave equation on a curved space-time, Cambridge University Press, Cambridge, 1975. [9] P. B. Gilkey, The Index Theorem and the Heat Equation, Publish or Perish, Inc., Boston, MA, 1974. [10] M. Golomb and H. F. Weinberg, \Optimal approximation and error Bounds," pp. 117190 in on Numerical Approximation, ed. by R. E. Langer, Madison, 1959. [11] N. J. Hicks, Notes on Di erential Geometry, D. van Nostrand Co., Princeton, NJ, 1965. [12] M. G. Krein, \On a general method for decomposing positive-de nite kernels into elementary products," (in Russian) Dokl. Acad. Nauk SSSR 53 (1946), 3-6. [13] J. L. Lions and E. Magenes, Non-Homogeneous Boundary Value Problems and Applications, Vol. I, Springer-Velag, New york, 1972. 34

[14] K. Maurin, Methods of Hilbert Spaces , Hafner Publishing Company, New york, 1967. [15] W.R. Madych and S.A. Nelson, \Multivariate interpolation and conditionally positive de nite functions", Approx. Theory and its Applications 4 (1988), 77-79. [16] W.R. Madych and S.A. Nelson, \Multivariate interpolation and conditionally positive de nite functions II", Math. Comp. 54 (1990), 211-230. [17] V. A. Menegatto, \Strictly Positive De nite Kernels on the Hilbert Sphere," Applicable Analysis, to appear. [18] C. Muller, Spherical Harmonics, Springer-Verlag, Berlin, 1966. [19] F. J. Narcowich, \Generalized Hermite Interpolation and Positive De nite Kernels on a Riemannian Manifold," J. Math. Anal. Applic. 190 (1995), 165-193. [20] F. J. Narcowich and J. D. Ward, \Generalized Hermite interpolation via matrix-valued conditionally positive de nite kernels," Math. Comp., 63 (1994), 661-688. [21] M. J. D. Powell, \The theory of radial basis approximation in 1990," in Wavelets, Subdivision and Radial Functions, W. Light, (ed.), Oxford University Press, 1990. [22] A. Ron and X. Sun, \Strictly Positive De nite Functions on Spheres," Math. Comp., 65 (1996), 1513-1530. [23] R. Schaback, \Error Estimates and Condition Numbers for Radial Basis Function Interpolation," Advances in Computational Mathematics 3 (1995), 251-264. [24] I. J. Schoenberg, \Positive De nite Functions on Spheres," Duke Math. J. 9 (1942), 96-108. [25] J. Stewart, \Positive de nite functions and generalizations, an historical survey," Rocky Mountain J. Math., 6 (1976), 409-434. [26] G. Wahba, \Spline Interpolation and Smoothing on the Sphere," SIAM J. Sci. Stat. Comput. 2(1981), 5-16. [27] G. Wahba, \Surface Fitting with Scattered Noisy Data on Euclidean d-Space and on the Sphere," Rocky Mountain J. Math. 14 (1984), 281-299. [28] F. W. Warner, Foundations of Di erentiable Manifolds and Lie Groups, Scott, Foresman, and Co., Glennview, IL, 1971. [29] Z. Wu, \Hermite-Birkho interpolation of scattered data by radial basis functions," Approx. Th. Applics. 24, (1992), 201-215. [30] Z. Wu and R. Schaback, \Local error estimates for radial basis function interpolation of scattered data," IMA journal of Numerical Analysis 13, (1993), 13-27. [31] Y. Xu and E. W. Cheney, \Strictly Positive De nite Functions on Spheres," Proc. Amer. Math. Soc. 116 (1992), 977-981. [32] R. M. Young, An Introduction to Nonharmonic Fourier Series, Academic Press, New York, 1980.

35

Suggest Documents