Radial Basis Functions by Buhmann
Radial Basis Functions by Buhmann
Radial Basis Functions by Buhmann
CONTENTS
1 Introduction 1
2 Convergence rates 5
3 Compact support 11
4 Iterative methods for implementation 16
5 Interpolation on spheres 25
6 Applications 28
References 34
1. Introduction
There is a multitude of ways to approximate a function of many variables:
multivariate polynomials, splines, tensor product methods, local methods
and global methods. All of these approaches have many advantages and
some disadvantages, but if the dimensionality of the problem (the number
of variables) is large, which is often the case in many applications from stat-
istics to neural networks, our choice of methods is greatly reduced, unless
2 M. D. Buhmann
suitable adjustments being made when x is not from the whole space, and
the coefficient vector λ = (λξ )ξ∈Ξ is an element of RΞ . In many instances,
particularly those that will interest us in Section 3, the interpolation re-
quirements
s |Ξ = f |Ξ (1.2)
for given data f |Ξ lead to a positive definite interpolation matrix A =
{φ(kξ − ζk)}ξ,ζ∈Ξ . In that case, we call the radial basis function ‘positive
definite’ as well. If it is, the linear system of equations that comes from (1.1)
and (1.2) and uses precisely that matrix A yields a unique coefficient vector
λ ∈ RΞ for the interpolant (1.1).
All radial basis functions of Section 3 have this property of positive defin-
Radial basis functions 3
2 2
iteness, as does for instance the Gaussian radial basis function φ(r) = e−c r
for√all positive parameters c and the inverse multiquadric function φ(r) =
1/ r2 + c2 .
However, in some instances such as the so-called thin-plate spline radial
basis function, the radial function φ is only conditionally positive definite
of some order k on Rn , say, a notion that we shall explain and use in the
subsequent section. Now, in this event, polynomials p(x) ∈ Pk−1
n (x) of degree
k − 1 in n unknowns are augmented to the right-hand side of (1.1) so as to
render the interpolation problem again uniquely solvable. Consequently we
have as approximant
X
s(x) = λξ φ(kx − ξk) + p(x), x ∈ Rn . (1.3)
ξ∈Ξ
Then the extra degrees of freedom are taken up by requiring that the coeffi-
cient vector λ ∈ RΞ is orthogonal to the polynomial space Pk−1n (Ξ), that is,
all polynomials of total degree less than k in n variables restricted to Ξ:
X
RΞ 3 λ ⊥ Pk−1
n (Ξ) ⇐⇒ λξ q(ξ) = 0, ∀ q ∈ Pk−1
n . (1.4)
ξ∈Ξ
2. Convergence rates
As usual in the study of methods for the approximation of functions, one of
the central themes in the analysis of radial basis functions is their conver-
gence behaviour when the centres become dense in a domain or in the whole
underlying Euclidean space. This is highly relevant because it shows how
6 M. D. Buhmann
well we can approximate smooth functions even in the practical case when
the centres become close together but do not cover a whole domain. Several
of the results were initiated by the work of Duchon and this explains the
title of the following subsection.
where we admit nonintegral k but always demand k > 12 n. The typical case
we always think of is the thin-plate spline in two dimensions, namely n = k =
2 and therefore φ(r) = r2 log r. The aforementioned odd powers – linear or
cubic, for instance – also belong to this class. Several further important cases
such as multiquadrics√or shifted logarithms are in fact derived from the above
by composition with r2 + c2 : namely, φ(r) = r altered in this fashion √ gives
multiquadrics and φ(r) = log r provides the shifted logarithm log r2 + c2
(although in this case 2k = n). The reason for this transformation is to gain
infinite smoothness when c is positive (recall that all of the above are not
smooth at the origin when composed with the Euclidean norm).
In this context, the approximants are usually taken from the ‘native
spaces’ X of distributions (Jones (1982), for instance, for generalized func-
tions or distributions) in n unknowns whose total kth degree partial derivat-
ives are square-integrable: we call the space that depends on the choice of φ
the space X := D−k L2 (Rn ). The Sobolev embedding theorem tells us that
this space consists of continuous functions as long as k > n2 . This is the first
and perhaps most important example of the general notion of native spaces,
namely semi-Hilbert spaces X that ‘belong’ to the radial function and are
defined by all distributions f : Rn → R that render a certain φ-dependent
seminorm kf kφ finite. In the present case, the seminorm is the homogeneous
Sobolev norm of order k. We will return to this concept soon in somewhat
more generality.
A beautiful convergence result that holds on a very general domain Ω
and does not yet require explicitly native space seminorms (although they
are used implicitly in the proof, as we shall see) is the following one of
Powell (1994). So long as all domains as general as those in the statement
of the theorem are admitted, this is the best possible bound achievable for
this general class of domains, that is, the constant C on its right-hand side
Radial basis functions 7
This notion is also used in the rest of this section, as is the notation k · k∞,Ω
for the Chebyshev norm restricted to Ω.
Theorem 1. Let φ be from the class (2.1) with k = n = 2 and Ω be
bounded and not contained in a straight line. Let s be the radial basis
function interpolant to f |Ξ satisfying (1.2) for Ξ ⊂ Ω where k keeps the
same meaning as in the introduction, that is, linear polynomials are added
to s and the appropriate side conditions (1.4) demanded. Then there is an
h- and Ξ-independent C such that
p
ks − f k∞,Ω ≤ Ch log(h−1 ), 0 < h < 1.
In fact we can be even more specific about the constant C in the above
error estimate. It is, of course, independent of h and Ξ, but its depend-
ence on f can be expressed by C = c̃kf kφ , where c̃ depends only on Ω and
kf kφ is the homogeneous Sobolev seminorm of order 2 of f , depending on
the aforementioned native space. The general approach to convergence es-
timates of this form is always to bound the error |s(x) − f (x)| by a fixed
constant multiple of
p
kf kφ Φ(α) (2.3)
that depends on the radial basis function (2.1) and the dimension n, where
α = (αξ ) ∈ RΞ are the coefficients of the representation
X
x= αξ ξ, x ∈ Ω,
ξ∈Ξ
Consequently, the main work lies in bounding this power functional from
above. If this is done judiciously we can obtain optimal bounds, because
the bound of the error function by a suitable constant multiple of (2.3) is
best possible (Wu and Schaback 1993).
Bejancu (1997) has generalized this result to arbitrary k and n, and his
theorem includes the above result (see also Matveev (1997)). There are no
further restrictions on Ω except, in general, its Pk−1
n -unisolvency which we
demand for the following theorem.
Theorem 2. Let φ be from the class (2.1). Let Ω be bounded and contain
a Pk−1
n -unisolvent subset. Let s be the radial basis function interpolant
(1.3) to f |Ξ for Ξ ⊂ Ω as in the introduction where k keeps the same
8 M. D. Buhmann
In this expression, the Lagrange function L that satisfies L(k) = δ0k for all
multi-integers k, and may be expanded as
X
L(x) = λi φ(kx − ik), x ∈ Rn ,
i∈Zn
Precise upper bounds on the approximation order that can be identified for
Ξ = (hZ)n may be stated in a very general context even for h-dependent ra-
dial basis functions φh that are from a sequence of radial functions {φh }h>0 .
Therefore we study approximants from spaces
· n
Sh (φh ) = span φh −j |j ∈Z .
h
A special case is, of course, φh ≡ φ for all h, perhaps taken from one of
our (2.1). For the statement of the next theorem we let Ω be the unit ball
and σ : Rn → R a smooth (C ∞ ) cut-off function that is supported in Ω
and satisfies σ | 1 Ω = 1. We recall that the notation fˇ stands for the inverse
2
Fourier transform
Z
1
fˇ(x) = eix·t f (t) dt,
(2π)n
Then the Lp -approximation order from Sh (φh ) cannot be more than 2k,
that is, the distance in the Lp -norm between the Lp -closure of Sh (φh ) and
the class of band-limited f whose Fourier transform is infinitely differentiable
cannot be o(h2k ) as h tends to zero.
We note that the class of all band-limited f whose Fourier transform is
infinitely differentiable is a class of very smooth local functions and if the
Lp -approximation order to such smooth functions cannot be more than h2k ,
it cannot be more than h2k to any general nontrivial larger set of less smooth
functions.
A typical example occurs when we again use radial basis functions of the
form (2.1) where φ̂(r) is a constant multiple of r−2k . Then condition (1) of
the above theorem is certainly true for φ̂h ≡ φ̂. Moreover, condition (2) is
true because the smoothness of the cut-off function, the smoothness of the
function φ̂(k · +2πj0 k) in a neighbourhood of the origin for nonzero j0 and
the fact that 2k > n imply by Lemma 2.7 of Buhmann and Micchelli (1992)
that condition (2) holds. Finally, we can take ρ̃ = φ̂−1 and get the required
result from condition (3), namely that O(h2k ) is best possible.
This is the obtainable (saturation) order and an inverse theorem of Scha-
back and Wendland (1998) tells us that all functions for which a better order
is obtainable must be trivial in the sense of polyharmonic functions. For its
statement we recall the standard notation ∆ for the Laplace operator.
Theorem 7. Let Ω be as in Theorem 2, Ξ ⊂ Ω a finite centre-set with
distance h as in (2.2) and φ as in (2.1). If, for any f ∈ C 2k (Ω), and all
compact Ω̃ ⊂ Ω,
ks − f k∞,Ω̃ = o(h2k ), 0 < h < 1,
then ∆k f = 0 on Ω.
A very similar, also inverse, but more general theorem from the same
paper is the following one with which we close this section. In order to
state it we come back to the notion of native spaces and recall that for a
Radial basis functions 11
3. Compact support
This section deals with radial basis functions that are compactly supported,
quite in contrast to everything else we have seen before in this article. All of
the radial basis functions that we have considered so far have global support,
and in fact many of them, such as multiquadrics, do not even have isolated
zeros (thin-plate splines do, by contrast). Moreover, the radial basis func-
tions φ(r) are usually increasing with growing argument r → ∞, so that
square-integrability, for example, and especially absolute integrability are
immediately ruled out. In most cases, this poses no severe restrictions; we
can, in particular, interpolate with these functions and get good convergence
rates nonetheless, as we have seen in the previous section. There are, how-
ever, practical applications that demand local support of the basis functions
12 M. D. Buhmann
such as finite element applications (but note one of the approaches to partial
differential equations in the last section which works specifically with glob-
ally supported φ) or applications with very quickly growing data, or where
frequent evaluations with substantial amounts of centres are required, etc.
Therefore there is now a theory of radial basis functions of compact sup-
port where the entire class of φs gives rise to positive definite interpolation
matrices for distinct centres.
We note that, however, in the approximation theory community there
is an ongoing discussion about the usefulness of radial basis functions of
compact support because of their inferior convergence properties (unless we
wish to again forego the advantages of compact support – see our discussion
about scaling in the text below), and in comparison to standard, piecewise
polynomial finite elements. The latter, however, are harder to use in grid-
free environments and n > 3, because they require triangulations to be found
first.
As we shall see in this section, there exist at present essentially two ap-
proaches to constructing univariate, compactly supported φ : R+ → R such
that the interpolation problem is uniquely solvable with a positive definite
collocation matrix A = {φ(kξ − ζk)}ξ,ζ∈Ξ . There will be no restriction on
the geometry of the set Ξ of centres, but there are – in fact, there must be
– bounds on the maximal spatial dimension n which is admitted for each
radial function φ so that positive definiteness is retained. So they are still
useful in grid-free and high-dimensional applications.
In contrast to the radial basis functions of the previous section, the ap-
proximation orders we state, unfortunately, are not nearly as good as the
maximal ones available of the well-known globally supported radial func-
tions such as (2.1). This indeed puts a stricter limit to the usefulness of
compactly supported radial basis functions than we are used to for the glob-
ally supported ones. Be that as it may, it is nonetheless interesting to study
the question of when compactly supported radial functions give nonsingular
and convergent interpolation schemes.
upper bound on the spatial dimension n if we fix the degree of the piecewise
polynomial.
In order to derive a large class of compactly supported radial basis func-
tions starting from Askey’s result, Schaback and Wu (1995) introduced the
two so-called operators on radial functions
1
Df (x) := − f 0 (x), x > 0,
x
and Z ∞
If (x) := rf (r) dr, x > 0,
x
that are defined for suitably differentiable or asymptotically decaying f , re-
spectively, and are inverse to each other. Additive polynomial terms do
not arise on integration since we always restrict ourselves to compactly sup-
ported functions. Next, Wendland (1995) and Wu (1995) use the fact that
the said interpolation matrix A for the truncated power φ0 remains positive
definite if the basis function
φ(r) = φn,k (r) = I k φ0 (r), r ≥ 0, (3.1)
is used when ` = k+[n/2]+1. The way to establish that fact is by considering
the Fourier transform of the n-variate radially symmetric function φ(k · k),
which is also radially symmetric and computed by the univariate Hankel
transform
Z ∞
n/2 1−n/2
φ̂(r) = (2π) r sn/2 Jn/2−1 (rs)φ(s) ds, r = kxk ≥ 0, (3.2)
0
where Jn/2−1 is a Bessel function. This is the radial part of the radially
symmetric Fourier transform of φ(k · k). Hence we have to show that pos-
itivity of the Fourier transform, which is necessary and sufficient for the
positive definiteness of the matrix A by Bochner’s theorem, prevails when
the operator I above is applied, as long as the restrictions on dimension
are observed. This proof is carried out by studying the action of I on the
Hankel transform (3.2) and by direct computation and use of identities of
Bessel functions.
Starting from this, Wendland (1995, 1998), in particular, developed an
entire theory of the radial basis functions of compact support which are
piecewise polynomial and are positive definite. This theory encompasses
recursions for their coefficients when they are expanded in linear combin-
ations of powers and truncated powers of lower order, convergence results,
and minimality of polynomial degree for given dimension and smoothness.
Two results below serve as examples for the whole theory. The first states
the minimality of degree for given smoothness and dimension and the second
is a convergence result.
14 M. D. Buhmann
Theorem 9. The radial function (3.1) gives rise to a positive definite in-
terpolation matrix A with radial basis function φ = φn,k , and with distinct
centres Ξ in Rn . Further, among all such functions for dimension n and
smoothness C 2k , it is of minimal polynomial degree.
Theorem 10. Let φ be defined by (3.1), let f be a function in the Sobolev
space H k+(n+1)/2 (Rn ), and let k be at least 1 when n = 1 or n = 2. Then,
for a compact domain Ω with centres Ξ ⊂ Ω, the interpolant (1.1) satisfies
ks − f k∞,Ω = O hk+1/2 , h → 0,
Here, pξ ∈ P12 and λ·,ξ ⊥ P12 (Ξ) for each ξ as usual. Clearly, the computation
of such full Lagrange functions would be just as expensive as solving the
full usual linear interpolation system of equations. Therefore the idea is to
replace (4.2) by local Lagrange conditions which require for each ξ only that
the identity holds for some q = 30 points, say, ζ that are nearby ξ. Hence
we take Lagrange functions that are still of the form (4.3) but with at most
q nonzero coefficients. We end up with an approach to our interpolation
18 M. D. Buhmann
We shall see in the proof of the next theorem why λξξ is positive and we
may therefore divide by λξξ . The correction (4.5) is added for all ξ ∈ Ξ \ Σ
for each sweep of the algorithm. The final stage of each sweep consists of
Radial basis functions 19
the correction
s(x) −→ s(x) + σ(x),
where σ is the full standard thin-plate spline solution of the interpolation
problem with centres Σ computed with a direct method
σ(ξ) = f (ξ) − s(ξ), ξ ∈ Σ.
Here also, s denotes the current approximation to s∗ after all intermediate
steps. This finishes the sweep of the algorithm, and, if we started with a
trivial approximation s = s0 = 0 to s∗ , the sweep replaces sm by sm+1 that
goes into the next sweep. The stopping criterion can be, for instance, that
we terminate the algorithm if all the residuals are sufficiently small. The
main work is clearly the computation of the residuals, the coefficients of the
local Lagrange made available before the start. For that, a method such as
the one described in the next subsection about multipole methods can be
used.
It is quite a novelty that there is a convergence proof of this algorithm,
although it had been known for some time that the method performs ex-
ceptionally well in practical computations even if Ξ contains as many as
50000 points. It turns out that it is not unusual to have an increase of ac-
curacy of one digit per sweep of the algorithm, that is, each sweep reduces
max |s(ξ) − f (ξ)| by a factor of ten, which indicates very fast convergence.
We provide the essentials of the proof of this theorem (Faul and Powell
1999) – which is in fact not very long – because it is highly instructive about
the working of the algorithm and involves many concepts that are typical
for the use and analysis of radial basis functions, such as native spaces and
their seminorms and semi-inner products.
Theorem 13. Let {sj }∞ ∗
j=0 be a sequence of approximations to s gener-
ated by s0 = 0 and the above algorithm. Let the radial function be the
thin-plate spline function and n = 2. Then limj→∞ sj = s∗ .
Proof. We show first that the native space norm ks∗ − sj kφ is monotonic-
ally decreasing with increasing j. Let (·, ·) be the semi-inner product that
corresponds to the native space norm, namely
Z
1 1
(u, v) = n
û(t)v̂(t) dt. (4.6)
(2π) φ̂(ktk)
Here, the requirements kukφ < ∞, kvkφ < ∞ are sufficient for the above
integral to exist, by the Cauchy–Schwarz inequality. In the thin-plate spline
case and in two dimensions, this is by Parseval–Plancherel a multiple of
Z 2
∂ u(x, y) ∂ 2 v(x, y) ∂ 2 u(x, y) ∂ 2 v(x, y) ∂ 2 u(x, y) ∂ 2 v(x, y)
+ 2 + dx dy
∂x2 ∂x2 ∂x∂y ∂x∂y ∂y 2 ∂y 2
and has as kernel the space of linear polynomials.
20 M. D. Buhmann
= λ ξi ξi ,
by (4.4) and because pζ is annihilated by the side conditions on the coeffi-
cients λζξi . This also shows, incidentally, that the important inequality
λ ξi ξi > 0 (4.8)
holds because kLξi k2φ is not zero. Otherwise Lξi would be in the kernel of
our semi-inner product and not able to satisfy the cardinality conditions.
Moreover, by the same token, we get from the reproducing kernel properties
X
(s∗ − sj−1,i−1 , Lξi ) = λζξi (s∗ − sj−1,i−1 , φ(k · −ζk))
ζ∈Lξi
X
= λζξi (s∗ (ζ) − sj−1,i−1 (ζ))
ζ∈Lξi
X
= λζξi (f (ζ) − sj−1,i−1 (ζ)).
ζ∈Lξi
when this happens, we need to prove that the following inner product van-
ishes !
X
∗
s − sj−1,m − σ, λ̂τ φ(k · −τ k) + p̂
τ ∈Σ
!
X
= s∗ − sj−1,m − σ, λ̂τ φ(k · −τ k)
τ ∈Σ
= 0.
In the above, the coefficients λ̂τ are real and p̂ ∈ Pk−1 n . This orthogonality
relation is established using the same facts as those we required for showing
kLξi k2φ = λξi ξi and therefore we do not repeat the arguments.
In summary, we now know that ksj,i − s∗ kφ tends to a limit for all fixed
i, and j increasing, because it is bounded below by zero and monotonically
decreasing, as it is a subsequence of the monotonically decreasing doubly
indexed sequence {ksj,i − s∗ kφ }∞ j=0,0≤i≤m . Moreover, by the definition of our
semi-inner product
(s∗ − sj−1,i−1 , Lξi )2
ks∗ − sj−1,i−1 − θξi Lξi k2φ = ksj−1,i−1 − s∗ k2φ − .
kLξi k2φ
Therefore, for all centres
(s∗ − sj−1,i−1 , Lξi ) → 0, j → ∞, ξi ∈ Ξ \ Σ, (4.9)
because ksj,i − s∗ kφ converges. In particular, it follows that
(s∗ − sj , Lξ1 ) → 0, j → ∞. (4.10)
Moreover, sj,1 = sj + (s∗ − sj , Lξ1 )Lξ1 /kLξ1 k2φ , so that we also get
∗ (s∗ − sj , Lξ1 )Lξ1
s − sj − , Lξ2 → 0, j → ∞.
kLξ1 k2φ
Therefore
(s∗ − sj , Lξ2 ) → 0, j → ∞,
and indeed
(s∗ − sj , Lξ ) → 0, j → ∞, ξ ∈ Ξ \ Σ. (4.11)
Finally, we observe
X
(s∗ − sj , Lξi ) = λζξi (s∗ (ζ) − sj (ζ)) → 0, j → ∞. (4.12)
ζ∈Lξi
Recalling that s∗ −sj restricted to Σ vanishes anyway, and recalling that λξi ξi
is positive, we now go backwards and start from i = m with an induction
argument, whereupon (4.8), (4.9) and (4.12) imply sj (ζ) → s∗ (ζ) as j → ∞
22 M. D. Buhmann
children the intersections of the four quarter squares of the unit square with
Ξ. They are in turn divided into four grandchildren each in the same fashion.
We stop this process at a predetermined level that in part determines the
accuracy of the calculation in the end. Each member of this family is called
a panel. We point out that there are implementations where the divisions
are only by two and not by four (Powell 1993). There is no principal reason
for making this choice that forms a binary tree instead of a quad-tree and
both approaches have been tried successfully.
Now we must define the far field and the near field for each x where s(x) is
computed. Given x, all centres ξ that are in the near field give rise to explicit
evaluations of φ(kx − ξk). The near field consists simply of contributions
from all points which are not ‘far’, according to the following definition: we
say x is far away from a panel T and therefore from all centres in that panel,
if there is at least one more panel between x and T , and if this panel is on
the same level of parenthood as T itself.
Next, we have to decide how to group the far points. All panels Q are
in the evaluation list of a panel T if Q is either at the same or a coarser
(higher) level than T and every point in T is far away from Q, and if, finally,
T contains a point that is not far away from the parent of Q. Thus the far
field of an x whose closest centre is in a panel T is the sum of all
X
sQ (x) = λξ φ(kx − ξk) (4.13)
ξ∈Q
such that Q is in the evaluation list of T . For each (4.13), a common Laurent
series is computed. Since we do not know the value of x that is to be inserted
into (4.13), we compute the coefficients of the Laurent series and insert x
later on.
In a set-up process, we compute these expansions of the radial function for
large argument, that is, their coefficients, and store them; when x is provided
at the evaluation stage, we combine those expansions for all centres from
each evaluation list, and in an additional, final step we can simplify further
by approximating the whole far field by one Taylor series.
A typical Laurent series expansion to thin-plate spline terms, which we
still use as a paradigm for algorithms with more general classes of radial basis
functions, is as follows. To this end, we work in two dimensions n = 2 and
identify two-dimensional real space with one-dimensional complex space.
Lemma 1. Let z and t be complex numbers, and let
φt (z) := kt − zk2 log kt − zk.
Then, for all kzk > ktk, and denoting the real part by Re ,
( ∞ !)
X 1 t k
φt (z) = Re (z̄ − t̄)(z − t) log z − ,
k z
k=1
24 M. D. Buhmann
5. Interpolation on spheres
Because of the many applications that suit radial basis functions in geodesy,
there is already a host of papers that specialize radial basis function approx-
imation and interpolation to spheres. Freeden and co-workers (1981, 1986,
1995) have made a very large number of contributions to this aspect of ap-
proximation theory of radial basis functions. There are excellent and long
review papers available from the work of this group (see the cited references)
and we will therefore be relatively brief in this section. Of course, we no
longer use the conventional Euclidean norm in connection with a univariate
radial function when we approximate on the (n − 1) sphere S n−1 within
Rn but apply so-called geodesic distances. Therefore the standard notions
of positive definite functions and conditional positive definiteness no longer
apply, and one has to study new concepts of (conditionally) positive defin-
ite functions on the (n − 1) sphere. This started with Schoenberg (1942),
who characterized positive definite functions on spheres as those ones whose
expansions in series of Gegenbauer polynomials always have nonnegative
coefficients. Xu and Cheney (1992) studied strict positive definiteness on
spheres and gave necessary and sufficient conditions. This was further gen-
eralized by Ron and Sun in 1996
Recent papers by Jetter, Stöckler and Ward (1999) and Levesley, Light,
Ragozin and Sun (1999) use native spaces, (semi-)inner products and repro-
ducing kernels (cf. Saitoh (1988) for a treatise on the theory of reproducing
kernels) to derive approximation orders in a very similar fashion to the work
(`) `
summarized in Section 2. They all apply the spherical harmonics {Yk }dk=1
that form an orthonormal basis of the d` -dimensional space of polynomials
26 M. D. Buhmann
The native space X given in (5.2) and functions (5.1) will give rise to a
reproducing kernel that is positive definite, but if we enlarge the space by
starting the first sum in (5.2) only at ` = κ and thereby weakening condi-
tions, we can also get conditionally positive definite (reproducing) kernels
(Levesley et al. 1999) for the ensuing spaces Xκ . Then the native space will
be a semi-Hilbert space, that is, the inner product that we shall describe
shortly has a nullspace
( )
d
κ−1 X̀
X
(`)
K = f f = fˆ`k Y k .
`=0 k=1
Now, a standard choice for the positive weights for defining the space X
which are often independent of k is a`k = (1 + λ` )−s . This gives rise to the
Sobolev space H s (S n−1 ). Here the λ` = `(` + n − 2) are eigenvalues of the
Laplace–Beltrami operator.
The inner product that the native space is equipped with can be described
by
d
∞ X̀
X 1 ˆ
hf, gi = f`k ĝ`k ,
a`k
`=0 k=1
where the coefficients are defined through (5.1); they are still assumed to
be positive. The reproducing kernel that results from this Hilbert space X
with the above inner product and that corresponds to the function of our
previous radial basis functions in the native space is, when x and y are on
Radial basis functions 27
the sphere,
d
∞ X̀
X (`) (`)
φ(x, y) = a`k Yk (x)Yk (y), x, y ∈ S n−1 . (5.3)
`=0 k=1
This can be simplified by the famous addition theorem (Stein and Weiss
1971) to φ(x, y) = φ(xT y), where
∞
1 X
φ(t) = d` a`k P` (t), (5.4)
ωn−1
`=0
ωn−1 being the measure of the unit sphere, if the coefficients are constant
with respect to k. Here, the d` are as above and P` is a Gegenbauer poly-
nomial (Abramowitz and Stegun 1972) normalized by P` (1) = 1. Therefore,
we now use (5.3) or (5.4) for interpolation on the sphere, in the same place
and with the same centres Ξ as before, but they are from the sphere them-
selves of course. Convergence estimates are available from all three sources
mentioned above that vary in approaching the convergence question. Using
the mesh norm
h = sup inf arccos(xT ξ),
x∈S n−1 ξ∈Ξ
Jetter et al. prove the following theorem. The notation |Ξ| is for the cardin-
ality of the set Ξ as before.
Theorem 14. Let X and Ξ be as above with the given mesh norm h. Let
κ be a positive integer such that h ≤ 1/(2κ). Then, for any f ∈ X, there is
a unique interpolant s in
span {φ(ξ, ·) | ξ ∈ Ξ}
that interpolates f on Ξ and satisfies the error estimate
X∞
2 5(|Ξ| + 1) 2
ks − f k∞ ≤ kf kφ d` max a`k . (5.5)
ωn−1 k=1,...,d`
`=κ+1
Corollary 1. Let the assumptions of the previous theorem hold and sup-
pose further that |Ξ| + 1 ≤ C1 κn−1 and
C2 1
≤h≤ .
1+κ 2κ
Then the said interpolant s provides
(α−n)/2 !
h
ks − f k∞ = O
C2
or
exp(−αC2 /2h)
ks − f k∞ =O ,
h(n−1)/2
28 M. D. Buhmann
respectively, if d` ×maxdk=1
`
a`k is bounded by a constant multiple of (1+`)−α
for an α > n or by a constant multiple of exp(−α(1 + `)) for a positive α,
respectively.
An error estimate of Levesley et al., which includes conditionally positive
definite kernels for κ = 1 or κ = 2, is as follows, for n = 2 (see also Freeden
and Hermann (1986) for a similar, albeit slightly weaker result).
Theorem 15. Let X, Ξ, h and κ be as above, let s be the minimal norm
interpolants to f ∈ Xκ on Ξ. When φ is twice continuously differentiable
on [1 − ε, ε] for some ε ∈ (0, 1), then
ks − f k∞ ≤ Ch2 kf kφ ,
so that, in particular, a polynomial p ∈ Pκ−1
3 (S 2 ) is added to the s used in
Theorem 14.
6. Applications
6.1. Numerical solution of partial differential equations
Given that radial basis functions are known to be useful to approximate
multivariate functions efficiently, it is suitable to apply them to approxim-
ate solutions of partial differential equations numerically. Three approaches
have been tried and tested in this direction, namely collocation techniques,
variational formulations and boundary element methods, all in order to solve
elliptic partial differential equations with boundary values given. There are
various reasons why radial basis functions are useful for these three ap-
proaches. The first of these is useful because we know much about existence
and accuracy of radial basis function interpolants, especially when the data
are scattered, which is useful for non-grid collocation. The second resembles
typical finite element applications, where usually radial basis functions of
compact support are used to mimic the standard finite element approach
with multivariate piecewise polynomials. Finally, boundary element meth-
ods are suitable in several cases when radial basis functions are known to
be fundamental solutions (Green’s functions) of elliptic partial differential
operators, most notably powers of the Laplace operator. An example is the
thin-plate spline radial basis function and the bi-Laplacian operator. After
all, in boundary element methods, explicit solutions of the associated homo-
geneous problem are required in advance, for which it is immensely helpful
to have Green’s functions to work with.
Naturally, an important decision is the choice of radial basis function,
especially whether globally or locally supported ones should be used. In
a Galerkin approach, locally supported elements are almost always em-
ployed. Further, the use of radial basis functions becomes particularly in-
teresting when nonlinear partial differential equations are solved or non-grid
Radial basis functions 29
Λξ uΞ = f (ξ), ξ ∈ Ξ1 ,
Λζ uΞ = q(ζ), ζ ∈ Ξ2 . (6.3)
30 M. D. Buhmann
with Z
a(u, v) = (∇u)T (∇v) + uv.
Ω
If φ is a radial basis function of compact support such that its Fourier
transform satisfies the decay estimate |φ̂(r)| = O(r−2k ), then Franke and
Schaback (1998) establish the convergence estimate
ku − uΞ kH 1 (Ω) ≤ Chσ−1 kukH σ (Ω) ,
where h is given by (2.2) and k ≥ σ > n/2 + 1.
We now outline the third method, that is, a boundary element (BEM)
method, following Pollandt (1997). The dual reciprocity method uses the
second Green’s formula and a fundamental solution φ(k · k) of the Laplace
operator ∆, in order to reformulate a boundary value problem as a boundary
integral problem over a space of one dimension lower. This will then lead
through discretization to a linear system with a full matrix for collocation by
radial basis functions, in the way familiar from other applications of radial
basis function interpolation. The radial basis function that occurs in that
context is this fundamental solution, and, naturally, it is highly relevant in
this case that the Laplace operator is rotationally invariant and has radial
functions as Green’s functions. We give a concrete example. Namely, for a
nonlinear problem on a domain Ω ⊂ Rn with Dirichlet boundary conditions
such as the following one with a nonlinear right-hand side
∆u(x) = u2 (x), x ∈ Ω ⊂ Rn , (6.4)
u|∂Ω = q, (6.5)
the goal is to approximate the solution u of the elliptic partial differential
equation on the domain by g plus a boundary term r̃ that satisfies ∆r̃ ≡ 0
on the domain. To this end, one gets after an application of Green’s formula
the equation on the boundary that u(x) is the same as
Z Z
∂ ∂
u(y)2 φ(kx − yk) dy − φ(kx − yk) u(y) − u(y) φ(kx − yk) dΓy
Ω ∂Ω ∂ny ∂ny
(6.6)
for x ∈ Ω, where ∂n∂ y is the normal derivative with respect to y on Γ = ∂Ω.
The radially symmetric φ is still the fundamental solution of the Laplace
operator used in the formulation of the differential equation above. Further,
one gets after two applications of Green’s formula the equation
Z
1 ∂
(u(x) − g(x)) + φ(kx − yk) × (u(y) − g(y)) −
2 ∂Ω ∂n y
∂
(q(y) − g(y)) φ(kx − yk) dΓy = 0, x ∈ ∂Ω.
∂ny
We will later use this equation to approximate the boundary part of the
32 M. D. Buhmann
solution, that is, the part satisfying the boundary conditions. Now we as-
sume that there are real coefficients λξ such that the infinite expansion
(which will be truncated later on)
X
u2 (y) = λξ φe (ky − ξk), y ∈ Ω,
ξ
and
X
g(y) = e (ky − ξk),
λξ Φ y ∈ Ω.
ξ∈Ξ
We require that the equation in the display after (6.6) holds for finitely many
points x = ζj ∈ ∂Ω, j = 1, 2, . . . , t, only. Then we solve for the coefficients
λξ by requiring that (6.7) holds for y = ξ, for all ξ ∈ Ξ. This fixes the λξ by
interpolation (collocation in the language of differential equations), whereas
the equation after (6.6) determines the normal derivative ∂n∂ y u(y) on ∂Ω,
where we are replacing ∂n∂ y u(y) by another approximant, a polynomial
spline τ (y), for instance. Thus the spline is found by requiring the above
equation for all x = ζj ∈ ∂Ω, j = 1, 2, . . . , t, and choosing a suitable t.
Finally, an approximation u e(x) to u(x) is determined on Ω by the identity
Z
∂
u
e(x) := g(x) + (q(y) − g(y)) φ(kx − yk) dΓy −
∂Ω ∂n y
Z
∂g(y)
φ(kx − yk) τ (y) − dΓy , x ∈ Ω, (6.8)
∂Ω ∂ny
where the boundary term r̃ corresponds to the second term on the right-hand
side of the display.
Now, all expressions on the right-hand side are known. This is an outline
of the approach but we have skipped several important details. Nonethe-
less, one can clearly see how radial basis functions appear in this algorithm;
Radial basis functions 33
indeed, it is most natural to use them here, since many of them are funda-
mental solutions of Laplace operators or their iterates in certain dimensions.
In the above example and n = 2, φ(r) = (2π)−1 log r, φ(r)e = r2 log r (thin-
e
plate splines) and Φ(r) 1 4 1 4
= 16 r log r − 32 r are the correct choices. As
the collocation matrices that appear in the boundary integral equation we
are left with in the method are dense and as interpolation is not absolutely
necessary in this approach, it can be substituted by quasi-interpolation.
Quasi-interpolation has asymptotically essentially the same approximation
behaviour (i.e., convergence speed) as interpolation but does not require
any data-dependent linear systems to be solved; see also Buhmann (1990a).
Convergence theorems for the method are available in the paper by Pollandt.
ively for this (e.g., Eckhorn (1999), Hochreiter and Schmidhuber (1999),
Anderson, Das and Keller (1998)).
The approximation to so-called learning situations by neural networks
usually leads to very high-dimensional interpolation problems with scattered
data. Girosi (1992) mentions radial basis functions as a very suitable ap-
proach to this, partly because of their availability in arbitrary dimensions,
and their smoothness. A typical application is in fire detectors. An advanced
type of fire detector has to consider several measured parameters, such as
colour, spectrum, intensity, movement of an observed object from which it
must decide whether, for instance, it is looking at a fire in the room or not,
because the apparent fire is reflected sunlight. There is a learning procedure
before the implementation of the device, where several prescribed situations
(these are the data) are tested and the values zero (no fire) and one (fire)
are interpolated, so that the device can ‘learn’ to interpolate between these
standard situations for general cases later when it is used in real life. Radial
basis function methods have been tried very successfully for this application
because they are excellent tools for high-dimensional problems that will
undoubtedly find many more applications in real life, such as polynomial
splines have done in at least the last 30 years and still do now.
REFERENCES
M. Abramowitz and I. Stegun, eds (1972), Handbook of Mathematical Functions,
National Bureau of Standards, Washington.
R. W. Anderson, S. Das and E. L. Keller (1998), ‘Estimation of spatiotemporal
neural activity using radial basis function networks’, J. Comput. Neuroscience
5, 421–441.
N. Arad, N. Dyn and D. Reisfeld (1994), ‘Image warping by radial basis functions:
applications to facial expressions’, Graphical Models and Image Processing
56, 161–172.
R. Askey (1973), ‘Radial characteristic functions’, MRC Report 1262, University
of Wisconsin, Madison.
I. Barrodale and C. A. Zala (1997), ‘MJDP–BCS industrial liaison: applications
to defence science’, in Approximation and Optimization: Tributes to M. J. D.
Powell (M. D. Buhmann and A. Iserles, eds), Cambridge University Press,
Cambridge, pp. 31–46.
I. Barrodale and C. A. Zala (1999), ‘Warping aerial photographs to orthomaps
using thin plate splines’, Adv. Comput. Math. 11, 211–227.
B. J. C. Baxter (1991), ‘Conditionally positive functions and p-norm distance
matrices’, Constr. Approx. 7, 427–440.
R. K. Beatson and W. A. Light (1992), ‘Quasi-interpolation in the absence of
polynomial reproduction’, in Numerical Methods of Approximation Theory
(D. Braess and L. L. Schumaker, eds), Birkhäuser, Basel, pp. 21–39.
Radial basis functions 35
Z. Wu and R. Schaback (1993), ‘Local error estimates for radial basis function
interpolation of scattered data’, IMA J. Numer. Anal. 13, 13–27.
Y. Xu and E. W. Cheney (1992), ‘Strictly positive definite functions on spheres’,
Proc. Amer. Math. Soc. 116, 977–981.