You are on page 1of 87

These are classnotes for Math/CS 887, Spring 03 by Carl de Boor

corrections are welcome!


Approximation Theory overview
One considers special instances of the following
Problem. Given some element g in a metric space X (with metric d), nd a best approx-
imation (=:ba) m

to g from some given subset M of X, i.e., nd


m

M s.t. d(g, m

) = inf
mM
d(g, m) =: dist (g, M).
Abbreviation:
m

T
M
(g).
Basic questions
Existence: #T
M
(g) 1 ?
Uniqueness: #T
M
(g) 1 ? More generally, #T
M
(g) =?
Characterization: how would one recognize a ba (other than by comparing it with all
other candidates)? This is important for
Construction:
The metric
The metric is almost always a norm metric, i.e., d(x, y) := |x y|, and the set M
is usually a nite-dimensional linear subspace. But, as the following problem, of approxi-
mating a curve, shows, there are important practical instances in which linearity plays no
role, hence there is no suitable norm in which to pose the problem.
curve approximation problem X is the set of smooth closed curves, of nite
length, say, in R
2
; it is a metric space with the Hausdor metric
d(A, B) := maxdist (A, B), dist (B, A),
with
dist (A, B) := sup
aA
inf
bB
|a b|
2
.
M is the set of ellipses in R
2
, say (or some other class of simple curves).
23jan02 1 c _2003 Carl de Boor
Specic choices for normed X
(T is an interval or a suitable subset of R
d
)
L
2
(T), i.e., |x| := |x|
2
:=
_ _
T
[x(t)[
2
dt
_
1/2
Least-squares.
More generally: inner product spaces.
C(T), i.e., |x| := |x|

:= sup
tT
[x(t)[ uniform (or, Chebyshev).
L
1
(T), i.e., |x| := |x|
1
:=
_
T
[x(t)[ dt Least-mean.
L
p
(T, w), i.e., |x| := |x|
p,w
:=
_ _
T
w(t)[x(t)[
p
dt
_
1/p
weighted L
p
Specic choices for M
Usually, M is a nite-dimensional linear subspace, i.e., of the form
M = ran[f
1
, . . . , f
n
] := [f
1
, . . . , f
n
]a :=
n

j=1
f
j
a(j) : a F
n

with F either R or C.

k
:=
k
:= ran[()
j
: j = 0, . . . , k] (algebraic) polynomials of degree k. Here,
()
j
: t t
j
is the way Ill indicate the power function until someone comes along with
a better notation. More generally,

k
=
k
(R
d
) := ran[()

: [[ k]
with
()

: R
d
F : x x

:= x(1)
(1)
x(d)
(d)
, ZZ
d
+
,
and [[ := ||
1
=

i
(i). Also, for ZZ
d
+
,

:=

:= ran[()

: ].

1k,
:=
k
/

:= p/q : p
k
, q

rational functions of degrees k, .

k
:= ran[sin(), cos() : = 0, . . . , k] trigonometric polynomials of degree k. The
natural domain for trigonometric polynomials is the circle, i.e., the interval [0 . . 2]
or [ . . ] with the endpoints identied. Customary notation for this set:
TT = TT
1
.
If also complex scalars are admitted, we get the simpler description

k
= ran[e
ij
: [j[ k],
with i :=

1, and
e

: x exp(x)
the exponential with frequency , a denition which even makes sense for x F
d
with also in F
d
and
x :=

i
(i)x(i)
23jan02 2 c _2003 Carl de Boor
Exp

:= ran[e

: ] exponentials with frequencies . We already noted


that

k
(R) Exp
i{k,...,k}
. Also, for any R,
k
= lim
h0
Exp
h
whenever
# = k + 1.

k,
:= piecewise polynomials (=: pp) in C
()
, of degree k, with break sequence
= ( <
j
<
j+1
< ). (splines)
The rst and third are linear. The second is nonlinear. The last two are linear or
nonlinear depending on whether the frequencies (resp. the break sequence ) is xed or
variable.
Degree of approximation
considers the behavior of h dist (g, M
h
) as a function of the (discrete or continuous)
parameter h. E.g., k dist (g,
k
) as k , or h dist (g,

k,h
) as h 0. Usually,
one considers only
dist (K, M
h
) := sup
gK
dist (g, M
h
)
with K a class of functions sharing with the particular g of interest certain characteristics
(e.g., all functions whose 14th derivatives are no bigger than 7 in absolute value). Only the
behavior in the limit, as h 0 or h or whatever, is usually considered. If nothing is
said, then h 0.
Jackson type theorems: g K = dist (g, M
h
) = O(h

)
Bernstein type (or, inverse) theorems: dist (g, M
h
) = O(h

) = g K.
Saturation theorems: dist (g, M
h
) = o(h

) = g K
0
(for some appropriate
and with K
0
some very small set). E.g., dist (g,
0
1,
) = o([[
2
) = g
1
.
Typically, K := f X : |f|

1 for some stronger norm ||

. This leads to
consideration of the K-functional
K
f
: t inf
q
(|f q| +t|q|

),
which plays a major role in the precise description of h dist (K, M
h
) for such K.
Related question: Is M a good choice for approximating g, given that we know that
g K? Typical criterion involves the dimension of M, i.e., the degrees of freedom to be
used. If dimM = n, then one compares dist (K, M) with
d
n
(K) := inf
dimY n
dist (K, Y ),
the n-width of K (in the sense of Kolmogorov). While it is not easy to nd optimal
subspaces, i.e., Y with d
dimY
(Y ) = dist (K, Y ), one can often nd a ladder (M
n
) for
which dist (K, M
n
) d
n
(K) and dimM
n
n.
Here,
A(t) = O(B(t)) := limsup
t
[A(t)/B(t)[ < ;
A(t) = o(B(t)) := lim
t
[A(t)/B(t)[ = 0;
A(t) B(t) := A(t) = O(B(t)) and B(t) = O(A(t)),
23jan02 3 c _2003 Carl de Boor
with the limiting value of t usually clear from the context. E.g., the order of A(n) or A
n
will always be considered as the natural number, n, goes to , while the order of A(h) or
A
h
will always be considered as the real positive number, h, goes to zero.
Good approximation
In practice, best approximation is rarely used. Instead, one looks for cheap, but
good, approximation schemes. E.g., if A is a linear map into M, then |(1 A)
|K
| :=
sup
fK
|f Af| may well be close to dist (K, M).
Special case: a near-best A is one for which, for some const and for all f,
|f Af| const dist (f, M).
Any such A is necessarily a projector (onto M). Conversely, if A is a bounded linear
projector onto M, then, for any f, and any m M, |f Af| = |(1 A)(f m)|
|1 A||f m|, therefore
(1) |f Af| |1 A| dist (f, M).
course
The intent is to give a quick reading of these basics of AT, illustrated with the help
of splines, thereby giving also a quick introduction to (univariate) splines.
Weierstra, Korovkin, Lebesgue, Bernstein
Start o the course the way Lorentz starts o his book (the nicest book in classical
AT) and the way Tikhomirov starts o his survey of AT, namely with
(2) Weierstra (1885). For any (nite) interval I = [a . . b],
I
is dense in C(I).
Since both and | |

are invariant under translation


f f( +t)
and dilation
f f(/),
it is sucient to consider just one nontrivial interval, e.g., the special case I = [0 . . 1].
I will give three proofs (at least). The rst proof is Bernsteins, but done with Ko-
rovkins theorem. The second is Lebesgues, done with broken lines. The third is Stones.
23jan02 4 c _2003 Carl de Boor
(3) Korovkin (1957). F = R, T compact, (U
n
) in L(C(T)), U
n
positive, U
n
pw
1 on
some nite set F , 0. If there exists (a
f
: f F) in C(T) so that
p(t, s) :=

fF
f(t)a
f
(s) 0 with equality i t = s,
then U
n
pw
1.
Explanations:
(i) With U
n
and U being maps from the same domain and into the same target,
U
n
pw
U indicates that U
n
converges pointwise to U, i.e., x X lim
n
U
n
x =
Ux (in the topology of the common target of the U
n
and U). This is much weaker than
the more elusive uniform convergence, denoted U
n
u
U, which presupposes that the
common target is normed and means that
lim
n
sup
xX
|U
n
x Ux| = 0.
(ii) It is assumed here that C(T) is the set of real-valued continuous maps on T. For
those, there is a natural (partial) order, namely
f g : t T f(t) g(t).
U : C(T) C(T) is called positive (or, more precisely, nonnegative) if
0 f = 0 Uf.
Observation: Assume that U is positive and linear. Then f g = Uf Ug.
Further, with [f[ : T R : t [f(t)[,
[f[ f [f[ = U([f[) Uf U([f[),
hence,
[U(f)[ U([f[).
Proof of Korovkin (3): By assumption, U
n
pw
1 on F, hence (U
n
being
linear), also on
ran[F] :=

fF
fc(f) : c R
F
=: span F.
The latter is nite-dimensional, therefore U
n
1 uniformly on bounded subsets of ran[F].
The rest is a very nice trick, in which the arbitrary g C(T) to be approximated from
is locally related to some element of ran[F], as follows. From the assumption, ran[F]
contains (strictly) positive functions, e.g., the function t p(t, s) +p(t, s

) for any s ,= s

in case #T > 1 and [f[ for any f F0 otherwise. Let p

be one such. For s T, set


g =:
g(s)
p

(s)
p

+h(, s).
23jan02 5 c _2003 Carl de Boor
Then
(4) (U
n
g)(s) =
g(s)
p

(s)
(U
n
p

)(s) + (U
n
h(, s))(s),
and
U
n
p

.
Since T is compact, |1/p

< , hence
g(s)
p

(s)
(U
n
p

)(s)
g(s)
p

(s)
p

(s) = g(s)
uniformly in s. Korovkins result therefore follows from the following claim.
Claim. U
n
(h(, s))(s) 0 uniformly in s.
Proof: By the positivity of U
n
, [U
n
h(, s)[ U
n
([h(, s)[). Take any > 0. Then
[h[ + a bound for [h[ on the set

:= (t, s) : [h(t, s)[ . This set is closed


since h is continuous, hence compact; it also does not contain the zero-set of p, i.e., the
set (t, t) : t T, since h vanishes there. Therefore,
:= inf p(

) > 0,
hence, [h[ (|h|

/)p on

. So,
[h[ + (|h|

/)p.
Consequently, for any s,
[U
n
(h(, s))[ U
n
([h(, s)[) M + (|h|

/)U
n
(p(, s)),
with
M := sup
n
|U
n
()
0
|

nite since, by the strict positivity of p

, there is some positive b so that ()


0
bp

, therefore
[U
n
()
0
[ U
n
()
0
bU
n
(p

) bp

.
Now, p(, s) : s T is a bounded subset of the nite-dimensional linear space ran[F],
hence, on it, U
n
converges to 1 uniformly. Since p(s, s) = 0, this means that, for large n,
U
n
(p(, s))(s) is close to zero uniformly in s. Therefore, for all n n

,
[U
n
(h(, s))(s)[ (M + 1).
Since was arbitrary, this proves the Claim.
27jan03 6 c _2003 Carl de Boor
From this, (2)Weierstra for I = [0 . . 1] follows, with the choices F = ()
0
, ()
1
, ()
2

hence ran[F] =
2
, p : (t, s) (t s)
2
, and U
n
= B
n
, n > 0, the Bernstein operator
(introduced by Bernstein in 1912 for a dierent proof of Weierstra)
B
n
: f
n

j=0

j,nj
f(
j
n
),
with

r,s
: t
_
r +s
r
_
t
r
(1 t)
s
.
Indeed, B
n
is linear, and is positive as a map on C(I). Further,
DB
n
f =
n

j=1
_
n
j
_
j()
j1
(1 )
nj
f(
j
n
)
n1

j=0
_
n
j
_
(n j)()
j
(1 )
nj1
f(
j
n
),
therefore (and this is of independent interest)
(5) DB
n
f = n
n1

j=0
_
n 1
j
_
()
j
(1 )
n1j
f(
j
n
),
with
f := f( +
1
n
) f,
using the facts that
_
n
j
_
j = n
_
n1
j1
_
and
_
n
j
_
(nj) = n
_
n1
j
_
. Now note that (
k
)
k1
.
Therefore,
k+1
:=
k
vanishes identically on
k
, hence
B
n
(
k
)
k
(for k n; its trivial for k > n). Since also
B
n
f(t) = f(t), t = 0, 1,
it follows that B
n
= 1 on
1
, therefore B
n
()
j
()
j
for j = 0, 1 trivially, and we are
done once we show that B
n
f f for some f
2

1
, e.g., for f :=
1,1
: t t(1 t).
This f vanishes at 0 and 1 and is quadratic, hence B
n
f must have the same properties,
and therefore must equal
n
f for some
n
. It follows that DB
n
f(0) =
n
Df(0) =
n
,
while, by (5), DB
n
f(0) = n(f(1/n) f(0)) = nf(1/n) = 1
1
n
. Therefore
n
=
n1
n
, i.e.,
B
n
f = f f/n f as n .
29jan03 7 c _2003 Carl de Boor
(6) d-dim. Weierstra. The restriction
T
of the polynomials in d arguments to any
compact subset T of R
d
is dense in C(T).
Proof: It is sucient to prove the theorem for the special case T = I := [0 . .
1]
d
since T is compact, hence contained in some axi-parallel box that, after translation
and dilation, we may assume to be I, and, by Tietzes extension theorem, C(T) can be
isometrically imbedded into C(I).
Remember: The Tietze extension of f C(T) to an f
I
C(I) is given by the rule
f
I
: x
_
f(x), x T;
+ inf
tT
(f(t) ) dist (x, t)/ dist (x, T), x , T,
with := inf f(T) 1, say, i.e., so that f ()
0
is strictly positive. A proof is usually
given on the way to proving Urisohns Lemma.
Now choose U
n
as the tensor product B
n
B
n
of d copies of the Bernstein
operator. This means that
U
n
f :=

0j(n,...,n)

j
f(j/n),
with j Z
d
and

j
: x
j(1),nj(1)
(x(1))
j(d),nj(d)
(x(d)).
Evidently, U
n
is linear and positive. Moreover,
(f
i
) (C([0 . . 1]))
d
U
n

d
i=1
f
i
=
i
B
n
f
i
: x

i
(B
n
f
i
)(x(i)).
From the univariate argument, U
n
pw
1 on
2,...,2
since U
n
()

=
d
i=1
B
n
()
(i)
, and a
suitable p is
p : (x, y)

i
(x(i) y(i))
2
.
Korovkin also supplies a proof that the trigonometric polynomials are dense in C(TT),
the space of continuous functions on the circle. In this case, the role of p is played by
p : (t, s) 1 cos(t s) = (e
0
(e
i
+e
i
)/2)(t s),
i.e., ran[F]

1
, and the maps U
n
are the Fejer operators, i.e.,

n
: f
1

_
TT
F
n
( t)f(t) dt
with
F
n
() :=
1
2(n + 1)
_
sin((n + 1)/2)
sin(/2)
_
2
.
29jan03 8 c _2003 Carl de Boor
The Fejer operator associates with f the average of its truncated Fourier series of orders
0, 1, . . . , n, i.e.,

n
f =
1
n + 1
n

j=0
s
j
f
with
s
j
f :=

||j
e
i
_
TT
e
it
f(t) dt/(2).
In particular,
n
is a positive operator (which s
n
is not), and

n
(e
ik
) =
(n [k[ + 1)
+
n + 1
e
ik
,
hence
n
(e
ik
) converges to e
ik
for any k, in particular for [k[ 1. On the other hand,
recall that |s
j
| ln j as j , hence s
j
fails to converge to 1 on C(TT).
Lebesgues proof (1898) of (2)Weierstra Let I := [a. .b]. For any f C(I) and
any nite sequence := (a =
1
<
2
< <
+1
= b), the broken-line interpolant to f is,
by denition, the unique element P

f in
0
1,
that agrees with f at . For
j
t
j+1
,
f(t) P

f(t) = (f(t) f(
j
))

j+1
t

j+1

j
+ (f(t) f(
j+1
))
t
j

j+1

j
,
hence
|f P

f|


f
([[),
with
f
the (uniform) modulus of continuity of f and
[[ := max
j
[
j+1

j
[.
Conclusion: The collection

0
1
(I) :=
_
in I

0
1,
of continuous broken lines on I is dense in C(I).
Since each
0
1,
is contained in
1
+ ran[ [
j
[ : j = 2, . . . , ], the following Claim
therefore nishes the proof.
(7) Claim. For nite [a . . b] and any s, dist

([ s[, )[a . . b] = 0.
Proof: Only the case s (a. .b) needs proof. For such s, we may choose
1

0
that carries [a . . b] into [1 . . 1] in such a way that (s) = 0. Further, if p is close
to [ [ on [1 . . 1], then p is a polynomial that, on [a . . b], is that close to c[ s[
for some nonzero c that depends only on . Hence it is sucient to prove the Claim for
[a . . b] = [1 . . 1] and s = 0. Since [t[ = (t
2
)
1/2
, i.e.,
[ [ = ()
1/2
()
2
,
and ()
2
([1 . . 1]) = [0 . . 1], and ()
2
, the following Claim nishes the proof.
23jan03 9 c _2003 Carl de Boor
(8) Claim. dist

(()
1/2
, )[0 . . 1] = 0.
Proof: A standard proof uses the Taylor series expansion
(1 )
1/2
= 1

n>0
()
n
n

k=1
[3/2 k[/k
which converges uniformly on any compact subset of (1 . . 1) (since the power coecients
are all bounded by 1). I prefer the following proof, from Dieudonnes Analysis.
Dene the sequence (u
n
: n = 0, 1, 2 ) by the iteration
u
n+1
:= u
n
+ (()
1
(u
n
)
2
)/2, n = 0, 1, 2, . . .
with
u
0
:= 0.
Claim: for all n, u
n
and
u
n
()
1/2
, on [0 . . 1].
Indeed, assuming this already to be true for n (as it is for n = 0), we observe that then
also u
n+1
and compute
()
1/2
u
n+1
= (()
1/2
u
n
)(1 (()
1/2
+u
n
)/2) (()
1/2
u
n
)(1 ()
1/2
) 0
on [0 . . 1].
It follows that ()
1
u
2
n
0 on [0 . . 1], hence u
n+1
= u
n
+(()
1
(u
n
)
2
)/2 u
n
. Thus,
(u
n
) is monotone increasing, yet bounded on [0 . . 1], therefore pointwise convergent there,
and its limit is necessarily a xed point of the iteration used to dene it, hence its limit
is ()
1/2
. But since this limit function is continuous and [0 . . 1] is compact, u
n
()
1/2
uniformly (by Dinis Theorem).
Stone-Weierstra
For an arbitrary set T, the collection F
T
of all scalar-valued maps (whether real or
complex) is not only closed under (pointwise) addition and multiplication by a scalar, but
also closed under (pointwise) multiplication of two elements, i.e., for f, g F
T
, also
fg : T F : t f(t)g(t)
is in F
T
, and F
T
is a ring wrto these two operations. In fact, it is a ring with identity
since it contains the multiplicative identity, i.e., the function 1 : t 1. Since also
f(g) = (f)g = (fg), F, f, g F
T
,
29jan03 10 c _2003 Carl de Boor
F
T
is an algebra with identity.
The Stone(-Weierstra) theorem employs the following notion: A F
T
is said to
separate points if, for any two distinct points s, t T, there exists a A with a(s) ,= a(t).
If A is a linear subspace and also contains the identity, then this is equivalent to the
statement that the linear map
A F
2
: a (a(s), a(t))
is onto (since its range contains the vector (1, 1) as well as some vector (, ) with ,= ),
hence the map has right inverses. In particular, for any f F
T
and any s, t T, there
exists a
f,s,t
A that agrees with f at s and t. It is this conclusion that is needed.
(9) Stone(-Weierstra) (1937). Let T be compact metric. The only closed subalgebra
A, of real C(T), that separates points and contains 1 is C(T) itself.
Proof: The range a(T) of any a A is a bounded subset of R (since T is compact),
hence, by Claim 7, [ [ can be approximated, uniformly on a(T), by polynomials p, and
A, being an algebra, contains p a : t p(a(t)) for any polynomial. This implies that A,
being closed, is closed under formation of absolute values; i.e.,
a A = [a[ A.
Since
maxf, g : T R : t maxf(t), g(t)
can also be written
maxf, g = ((f +g) +[f g[)/2,
A is also closed under formation of the maximum of nitely many functions. Since
minf, g = maxf, g,
A is also closed under formation of minimum.
Take f C(T), > 0. Let t T. For each s T, there is a
s,t
A agreeing with
f at s and t, hence, there is some neighborhood U
s
of s on which f < a
s,t
. T being
compact, there exists a nite set S for which T =
sS
U
s
. Hence, the function
a
t
:= max
sS
a
s,t
is in A and satises
f < a
t
on T, a
t
(t) = f(t).
The latter implies that, on some neighborhood V
t
of t, f + > a
t
. Hence, with S a nite
set for which T =
tS
V
t
, the function
a := min
tS
a
t
is in A and satises
f + > a
as well as f < a. Consequently, |f a|

< . As is arbitrary and A is closed, f A


follows.
29jan03 11 c _2003 Carl de Boor
From this, even the multivariate Weierstra theorem, (6), follows since is an algebra
with identity that separates points.
Another consequence is the density of the even trigonometric polynomials, i.e., of
ran[cos(j) : j = 0, 1, 2, . . .] in C([0 . . ] (the function cos alone is enough to separate
points). Note that this algebra fails to be dense in C([0 . . ]) for any > .
The restriction to real-valued functions is essential here. Not only does the proof of
Stone make explicit use of it (it relies on the total ordering of the reals), but the polynomials
fail to be dense in C(z C : [z[ 1) of all complex-valued continuous functions of one
complex argument on the unit disk (even though they continue to form there an algebra
with identity that separates points), since their closure consists of functions analytic in the
interior of the disk.
On the other hand, Stones theorem does have the following (weaker) immediate
consequence for complex-valued functions which relies on nothing more than the fact that
C
C
(T) = C
R
(T) + iC
R
(T).
(10) complex Stone. Let T be compact metric. The only closed subalgebra A of complex
C(T) with identity which separates points and is closed under conjugation is C(T) itself.
Here, being closed under conjugation means that A contains, with f, also its conjugate,
f : T C : t f(t).
In particular,
D
is dense in C
C
(D) for D = x R
2
: |x|
2
1.
Existence
M X is called an existence set in case g X T
M
(g) ,= . Such a set is
necessarily closed, hence M is assumed to be closed from now on.
Having M closed does not ensure existence, even when M is a linear subspace and X
is a Banach space. A standard example is the kernel M := ker of any continuous linear
functional which does not take on its norm, i.e., for which there is no nontrivial
g X with [g[ = |||g|. For,
[g[ = || dist (g, ker ), g X, X

.
Hence, if |g m| = dist (g, ker ) with m ker , then
|| dist (g, ker ) = [g[ = [(g m)[ |||g m| = || dist (g, ker ),
therefore, necessarily, [(g m)[ = |||g m|. Hence, if fails to take on its norm, then
this can only happen if g = m, i.e., if g M (in which case T
M
(g) = g trivially).
Specically, take
X =
1
:= a R
N
: |a|
1
:=

j
[a(j)[ < .
Take
: a

j
(1 1/j)a(j).
29jan03 12 c _2003 Carl de Boor
Then
[a[

j
(1 1/j)[a(j)[

j
[a(j)[ = |a|
1
,
and the last inequality is sharp since, for any j, i
j
= 1 1/j
j
1 = |i
j
|
1
. However,
[a[ = |a|
1
implies equality throughout and thats not possible unless, for each j, [a(j)[ =
0, i.e., a = 0.
In eect, in the example, the sequence (i
j
: j N) is a maximizing sequence for
|| = sup
xX
[x[/|x|,
but this sequence fails to have limit points, hence the sup is not taken on.
Put positively, existence of bas from M is usually proved by establishing that, in some
weak enough topology, closed and bounded subsets of M are compact while the topology
is still strong enough to have x |x| at least lower semicontinuous (i.e., x
n
x =
liminf
n
|x
n
| |x|).
(11) Proposition. Any nite-dimensional linear subspace M of any nls X is an existence
set.
Proof: Since T
M
(g) necessarily lies in the intersection
M
g
:= M B

2 dist (g,M)
(g)
of M with the closed ball around g with radius twice the distance of g from M, we have
M
g
not empty and
T
M
(g) = T
M
g
(g).
M is closed (since it is nite-dimensional), therefore M
g
is a closed and bounded subset of
a nite-dimensional ls, hence compact. In particular, the continuous function m |gm|
takes on its inmum on M
g
.
The same proof supports the claim that any closed set M which is nite-dimensional
in the sense that its ane hull is nite-dimensional, is an existence set. More than that, if
M is merely nite-dimensional in the sense that it is the image of some closed subset of F
n
under a continuous map with continuous inverse (as a map to M), then M is an existence
set.
However, here is another example. Let X = L
p
[0 . . 1], for some p [1 . . ], and
M =
k,l
[0 . . 1]
the collection of all pp functions of degree k on [0. .1] with l pieces. For k = 0, this would
be the space of splines of order 1 with l 1 free knots. This collection includes also
all pp functions of degree k with fewer than l breakpoints since the latter are obtained
from the former description when some adjacent polynomial pieces happen to come from
the same polynomial. This, however, indicates the technical diculty to be overcome here:
M appears to be nite-dimensional in the sense that each of its elements is specied by
29jan03 13 c _2003 Carl de Boor
nitely many parameters or degrees of freedom, namely the l 1 (interior) breakpoints
and the (k + 1)l polynomial coecients. However, this parametrization is badly awed
since any element with fewer than l 1 active breakpoints can be parametrized this way
in innitely many dierent ways, as one can think of the inactive breakpoints as being
located anywhere. In eect, the natural parametrization describes this M as the image of
: 0 =
1
< <
l+1
= 1 R
(k+1)l
,
with the map failing to be 1-1 on the boundary of this domain, yet no better parametriza-
tion is at hand.
Let (s
n
: n N) be a minimizing sequence from M =
k,l
for g, i.e., lim
n
|g
s
n
| = dist (g, M). Without loss, we assume that each s
n
consists of exactly l pieces and,
AGTASMAT (:= After Going To A Subsequence May Assume That) the corresponding
sequence (
(n)
: n N) of breakpoint sequences
(n)
:= (0 =
(n)
1
< <
(n)
l+1
= 1)
converges to some (l + 1)-vector .
Assume rst that is strictly increasing. Let (p
j,n
: j = 1, . . . , l) be the sequence of
polynomial pieces which make up s
n
, with p
j,n
the piece corresponding to the interval
I
j,n
:= (
(n)
j
. .
(n)
j+1
).
Since s
n
is a minimizing sequence, it is, in particular, bounded, and this implies that,
for each j, the sequence (|p
j,n
|(I
j,n
) : n N) is bounded. Here and below, and for any
domain-dependent norm || such as the L
p
-norms on some domain T, and for any subset
U of that domain, |g|(U) is the same norm, but for the domain U. Since the endpoints
of I
j,n
converge to the endpoints of I
j
:= [
j
. .
j+1
], it follows that, for some slightly
smaller, but still nontrivial, interval

I
j
, the sequence (|p
j,n
|(

I
j
) : n large ) is bounded.
Since p |p|(

I
j
) is a norm on
k
, and
k
is nite-dimensional, it follows AGTASMAT
that p
j,n
converges to some p
j

k
, and since there are only nitely many j involved, we
can make this assumption for all j. The argument is nished by verifying that the pp s
with break sequence and polynomial pieces (p
j
: j = 1, . . . , l) is, indeed, the norm limit
of (s
n
). This implies that
dist (g, M) |g s| = lim
n
|g s
n
| = dist (g, M),
hence s T
M
(g).
This argument is not only a little bit shaky (since the verication mentioned two
sentences ago was not explicitly carried out), but runs into trouble in case there is coales-
cence, i.e., in case is not strictly increasing. This implies that, for some j, I
j
= lim
n
I
j,n
has no interior, hence the boundedness of the sequence (|p
j,n
|(I
j,n
) : n N) does not
force convergence of some subsequence of (p
j,n
: n). At the same time, the fact that I
j
is
trivial would suggest that, somehow, we dont care about this polynomial sequence. On
the other hand, we cannot simply ignore it. Or can we?
Here is a soft-analysis approach around this. A sequence (x
n
) in a nls X is said to
nearly converge to x X if
(12) y X liminf
n
|x
n
y| |x y|.
29jan03 14 c _2003 Carl de Boor
Further, a subset Y of X is nearly compact in Z if every sequence in Y has a nearly
converging subsequence, with near-limit in Z.
This is certainly a useful notion here. For:
(13) Proposition. If bounded subsets of M X are nearly compact in M, then M is
an existence set.
Proof: Any minimizing sequence (m
n
) in M for g X is necessarily bounded,
hence, AGTASMAT, there exists m M so that
dist (g, M) |g m| liminf
n
|g m
n
| = dist (g, M),
hence m T
M
(g).
(It would have suced here to work with the even weaker notion of demanding only
that
y X limsup
n
|x
n
y| |x y|.
However, such a notion of convergence isnt even preserved by going to subsequences.)
Here is a ready source for nearly convergent sequences.
(14) Proposition. Assume that is a collection of seminorms on the nls X, and that
sup

(x) = |x|, x X.
If (x
n
) -converges to x, i.e.,
lim
n
(x x
n
) = 0,
then (x
n
) nearly converges to x.
Proof: Since seminorms provide a translation-invariant characterization of con-
vergence, it is sucient to prove (12) for just y = 0. For this, since (x
n
) |x
n
|, we have
(x) = lim
n
(x
n
) liminf
n
|x
n
|, hence
|x| = sup

(x) liminf
n
|x
n
|.
Returning to the example, choose =

: > 0, with

: x |x|([0 . . 1]B

()).
For any p [1 . . ], these are seminorms on X = L
p
[0 . . 1] satisfying the assumptions of
Proposition 14, hence also its conclusion. The above selection process is sure to provide
a subsequence which, for some > 0,

-converges to a certain element of


k,l
. Since
k
is nite-dimensional, this implies that the subsequence

-converges to that element for


every 0 < , hence nearly converges to it, and the proof of Proposition 13 does the
rest.
Notice that the near-limit of that subsequence of (m
n
) only depends on the polyno-
mial pieces p
j,n
for which I
j
is not trivial. Hence, if some I
j
is trivial, we may change
the corresponding p
j,n
arbitrarily without changing the near-limit. In particular, we can
arrange such changes that m
n
will fail to converge in norm, thus providing an explicit
example of the fact that near-convergence is truly more general than convergence.
03feb03 15 c _2003 Carl de Boor
The same arguments also settle existence questions when dealing with smooth piece-
wise polynomials. These are the elements in the space

k,
,
consisting of all s
k,
that are in C
(
i
)
(
i
), all i. If is just an integer, it is taken to
stand for the corresponding constant sequence. In any event, each entry of is restricted
to be no greater than k + 1 since
i
= k + 1 implies innite smoothness at
i
.
To be precise, the above arguments guarantee that any minimizing sequence from
_
0=
1
<<
l+1
=1

k,
,
leads to an element in some
k,
, but not all elements of
k,
can appear. For example,
if l = 2 and = k 1 > 0, then only elements of some

k,
can appear. It is useful to
discuss this question in the more general context of -polynomials.
-polynomials
Let : I X be a curve in the nls X, i.e., a continuous map from some (nite or
innite) interval I into X. We are interested in the approximating family
M
,l
:=
_

1
<<
l
ran[(
1
), . . . , (
l
)].
Our particular concern is the special case
: x ( x)
k
+
,
with
()
+
: R R : s (s +[s[)/2
the truncation map, and with X a normed linear space of functions on I, such as L
p
(I),
as this leads to splines. However, here are other very useful examples.
exponentials, where
(t) := e
t
;
rationals, where
(t) := 1/( t).
point functionals, where
: I X

: t
t
,
with X some space of functions on I.
The last example is dear to Numerical Analysts who like to approximate linear func-
tionals, like f
_
I
f, by linear combinations of values (and, perhaps, derivatives). It is
29jan03 16 c _2003 Carl de Boor
prototypical, since, for all the others, there exists a space Y of functions on I and a pairing
, ) so that (t) is (up to a factor) the representer of evaluation at t, i.e., so that
y(t) = y, (t)), y Y, t I.
For example, the truncated Taylor series with remainder
f =

j<k
D
j
f(a)()
j
/j! +
_
b
a
( t)
k1
+
/(k 1)! D
k
f(t) dt on [a . . b]
shows that k( )
k1
+
represents

on Y := f C
(k)
[a . . b] : f vanishes k-fold at 0
with respect to the pairing (g, f)
_
b
a
g D
k
f/k!.
In the context of existence, it becomes important to know whether M
,l
is closed. More
precisely, it becomes important to know the possible near-limits of minimizing sequences
in M
,l
. Since near convergence is a weaker concept than (norm) convergence, we may
have to go beyond the (norm-)closure of M
,l
in order to get an existence set. However,
we need to know the norm-closure M

,l
of M
,l
in any case.
If is a smooth curve, in the sense that the (norm-)limit
D(x) := lim
yx
((x) (y))/(x y)
exists, then also D(x), being a norm-limit of elements of M
,l
, must be in M

,l
. If
is even smoother, in the sense that the tangent curve D is itself smooth, then also all
points on the second derived curve D
2
must be in M

,l
.
We may pick up this discussion later, after we have recalled well-known facts about
divided dierences.
Uniqueness
The question of uniqueness is, more generally, the question of the nature of T
M
(g).
By denition,
T
M
(g) = M B

dist (g,M)
(g).
Hence, if M is convex, then so is T
M
(g), and may well contain nontrivial line segments if,
e.g., M is a linear subspace and the boundary B of the unit ball, B, contains line segments,
i.e., the norm fails to be strictly convex. The latter is the case for
p
(n) (which is F
n
with
the
p
-norm) i p = 1, . It is also the case for L
p
with p = 1, . In these spaces, we
expect nonuniqueness of best approximation even from nite-dimensional spaces. (E.g.,
X = L
1
[0 . . 1], M =
0
, g =
[1/2..1]
.) However, nonuniqueness is not guaranteed; it
depends on the attitude of M with respect to B
dist (g,M)
(g). (In the example just given,
one gets uniqueness for any g =
[x..1]
except when x = 1/2.) Draw pictures in
p
(2) for
p = 1, 2, .
03feb03 17 c _2003 Carl de Boor
A set M that provides exactly one ba for every g X is called by some a Chebyshev
set. Any Chebyshev set M in a nls X induces a map P
M
, called the metric projector
for M, by the rule
(15) T
M
(g) =: P
M
g, g X.
Any nite-dimensional Chebyshev subspace M of C[a . . b] (e.g., M =
k
) provides not
only a unique best approximation, but something called strong uniqueness. This means
that, for every g C[a . . b], there exists a positive constant c so that, for all m M,
|g m| |g P
M
g| +c|mP
M
g|.
This reects the geometric fact that the unit ball in C has corners.
(16) Proposition. The metric projector of any nite-dimensional Chebyshev subspace
M of any nls X is continuous.
Proof: Assume that g
n
g. We are to prove that P
M
g
n
P
M
g. For this, it is
sucient to prove that P
M
g is the unique limit point of (P
M
g
n
). For this, observe that
|P
M
g| |g| for any g, hence the convergence of (g
n
) implies that (P
M
g
n
) is bounded,
hence has limit points, as a bounded subset of a nite-dimensional ls. However, any such
limit point necessarily equals P
M
g, by the following Lemma.
(17) Lemma. If X is ms, and (g
n
) is a sequence in X with limit g, and (m
n
) is a
corresponding sequence with m
n
T
M
(g
n
) that converges to some m, then m T
M
(g).
Proof: By an earlier assumption, M is closed, hence m M. Therefore, if not
m T
M
(g), then there would exist m

M with d(g, m

) < d(g, m). We use the triangle


inequality to conclude that
d(m

, g
n
) d(g
n
, g) d(m

, g) < d(m, g) d(m, m


n
) +d(m
n
, g
n
) +d(g
n
, g).
Since d(g
n
, g), d(m, m
n
), and d(g
n
, g) all go to zero as n , this would imply that, for
suciently large n,
d(m

, g
n
) < d(m
n
, g
n
),
contradicting the fact that m
n
T
M
(g
n
).
Note that, even when M is a linear subspace, the metric projector is usually non-
linear hence not all that easy to construct or apply. Inner product spaces (or, equivalently,
least-squares approximation) are so popular precisely because the resulting metric projector
is linear. (In fact, one of the simpler characterizations of inner product spaces describes
them as the normed linear spaces in which the metric projector for every three-dimensional
linear subspace is linear.)
Much paper has been used in attacking the metric selection problem, which is the
problem of understanding in what circumstances an existence set permits the construction
of a continuous map P
M
satisfying (15). Since the resulting constructions are often not
03feb03 18 c _2003 Carl de Boor
practical, it is, from a practical point of view, much more interesting to construct near-best
projectors.
If M is a nonconvex existence set in the nls X, then nonuniqueness is guaranteed,
particularly if g is very far from M. Draw the picture. More than that, there are
perfectly reasonable M that have corners in the sense that #T
M
(g) > 1 for certain g
arbitrarily close to M. Here is one example.
(18) Example Consider M :=
0,2
in L
2
[1 . . 1], say, and take g = ()
2
. Since
M is closed under multiplication by scalars, we have T
M
(()
2
) = [[T
M
(()
2
), hence it is
sucient to consider g = ()
2
. Let m T
M
(g), and let be its sole breakpoint. Since g
is even, also m

: t m(t) is in T
M
(g). Hence, uniqueness would imply that m = m

,
therefore, m
0
. This would imply that, for any [1 . . 1], m + ran[m

] M,
with m

:=
[..1]
, hence, necessarily, g m must be orthogonal to m

for every such .


However, ran[m

: 1 < < 1] is dense in L


2
[1 . . 1], therefore g = m would follow,
contradicting the fact that g ,
0
.
The same argument shows that, in any L
2
-best approximation by splines with l simple
(free or variable) knots, all l knots are active (though coalescence is, of course, possible).
Note that : [a . . b] L
2
([a . . b]) : t
[t..b]
is a continuous curve which, at every
point (s), turns 90

in the sense that, for all r < s < t, the secant directions (s) (r)
and (t) (s) are perpendicular to each other. It is this counter-intuitive example that
led to the concept of -polynomials.
Characterization
The standard characterization theorems for bas are in terms of linear functionals
(which is not too surprising since the derivative of the scalar-valued map m |f m| at
some m ,= f is necessarily a linear functional if it exists).
The action of a continuous linear functional on a nls X over the real scalars is very
simple: The kernel of cuts X into two halfspaces, on one is positive, on the other it is
negative. Further, is constant on hyperplanes parallel to the kernel, i.e., on
H(, ) := x X : x = = x

+ ker
for any x

H(, ).
Let M be a closed subset of X, let g , M, hence
r := dist (g, M) > 0
and inf |M B
r
(g)| = 0, therefore
(19) sup (M) inf (B
r
(g)), X

.
07feb03 19 c _2003 Carl de Boor
Now suppose that, in fact, we have equality here, i.e., suppose that
(20) X

0 s.t. := sup (M) = inf (B


r
(g)).
Then the hyperplane H(, ) has M on its negative side and B
r
(g) on its positive side.
For that reason, such is called a separating linear functional for M and B
r
(g). More
than that, H(, ) is a supporting hyperplane for both sets, meaning that each set is
on one side of it, but with 0 distance.
Finally, suppose that m T
M
(g). Then
m sup (M) = inf (B
r
(g)) = g sup (B
r
)
= g || |g m|
g (g m) m,
hence there must be equality throughout. In particular, then
(21) m = sup (M), (g m) = || |g m|.
One says that is parallel to g m, in symbols
|g m,
if both and g m are nonzero and satisfy the second condition in (21). This language is
derived from the special situation in a Hilbert space, since then is necessarily of the form
, y) for some y X and now |g m implies that y and g m are positive multiples
of each other. Whether or not X is a Hilbert space, the condition |g m also says that
g m is an extremal for , meaning that it is a nonzero vector at which takes on its
norm.
Either way, (21) provides very useful necessary conditions for m M to be in T
M
(g).
Moreover, these conditions must hold, not only for every m T
M
(g) but also for every
satisfying (20), i.e., separating M and B
dist (g,M)
(g).
More than that, if m M satises (21) for any such , then, for any m

M,
|||g m| = g m g m

|||g m

|,
hence (since ,= 0 by assumption), m T
M
(g).
This proves
(22) Proposition. Let M be a closed subset of the nls X, g XM, hence r :=
dist (g, M) > 0, and let m M. If separates M and B
r
(g) (i.e., satises (20)), then
m T
M
(g) i m satises (21).
In general there may be no separating linear functionals. However, if we know, in
addition, that M is convex, then the Separation Theorem assures us, for each g XM,
of the existence of satisfying (20) and, with that, we have proved
07feb03 20 c _2003 Carl de Boor
(23) Characterization Theorem for ba from convex set. Let M be closed, convex
in the real nls X, let g XM and m M. Then, m T
M
(g) i there is some |(g m)
with sup (M) m.
Since dist (g, M) = r = (g inf (B
r
(g))/||, while, by (19), for any X

,
g ||r = inf (B
r
(g)) sup (M) with equality possible, it follows that
(24) dist (g, M) = max
=0
(g sup (M))/||,
which provides a useful and sharp lower bound for the distance of g from the convex set
M.
The theorem applies, in particular, to best approximation from a closed linear subspace
M of a nls X. However, for any X

0, sup (M) on a lss M can only take on the


values 0 and . For the in our theorem, this leaves only the value 0 (since sup(M)
must equal m). Thus the condition sup(M) = m is replaced by the condition
M := ker M.
(25) Characterization Theorem for ba from lss. Let M be a linear subspace of the
nls X, let g X and m M. Then, m T
M
(g) i there is some |(g m) with M.
Since sup (M) 0, in case M is a linear subspace, the lower bound (24) simplies
in that case to
(26) dist (g, M) = max
M
g/||.
Construction of ba
Characterization theorems are used in the construction of bas.
In the best of circumstances, the norm in question is smooth at the point (gm)/|g
m|, meaning that the condition |(g m) determines uniquely, up to nonnegative
multiples. On the other hand, the condition M holds i it hold for any nontrivial
scalar multiple, hence one may without loss restrict in the characterization theorem to
be of norm 1. Hence, in smooth norms (i.e., norms that are smooth every boundary point
of B, and for a nite-dimensional M, the characterizing conditions
M | (g m)
constitute nitely many equations in the coecients of the sought-for ba, m, with respect
to some convenient basis for M, i.e., in as many unknowns as there are equations.
The classical example (from which the entire geometric language used here derives) is
least-squares approximation, i.e., best approximation in an inner product space (such
as
2
or L
2
), with inner product , ). In such a space, the statement
(27) 0 < g = || |g|
implies that = c, g) for some positive constant c. The characterization theorem there-
fore specializes to the familiar
07feb03 21 c _2003 Carl de Boor
(28) Proposition. The element m of the closed linear subspace M of the inner product
space X is a ba to a given g X i g m M.
The condition g m M, when expressed in terms of a basis V = [v
1
, . . . , v
n
] for M,
becomes the so-called normal equations,

j
v
j
, v
i
)a(j) = g, v
i
), i = 1, . . . , n,
a linear system in the coecient vector a for the ba m = V a with respect to that basis. A
careless choice of the basis V may lead to numerical disaster (as would be the case if, e.g.,
X = L
2
[100 . . 101], M =
k
, and one were to choose the power basis, [()
j
: j = 0, . . . , k]).
However, if M is a spline space, then it is usually acceptable to choose for V the B-spline
basis for that space.
If V is chosen to be orthogonal, i.e., v
j
, v
i
) = 0 i i ,= j, then the normal system
becomes diagonal, and the best approximation is given by
m =

j
v
j
g, v
j
)
v
j
, v
j
)
.
In L
p
for p ,= 2 but 1 < p < , the statement (27) still determines a unique of norm
1, given in the form , g

) for some g

in the dual space, L


p
, i.e., with 1/p + 1/p

= 1.
But g

depends nonlinearly on g, and this makes the resulting normal equations harder
to solve. One successful technique consists in converting the problem into a weighted
L
2
-problem, with the weight determined iteratively. Specically, if X = L
p
[0 . . 1], then
|g m|
p
p
=
_
1
0
[g m[
2
w
g,m
,
with the weight function w
g,m
:= [g m[
p2
.
Such techniques even work for p = if one uses the Polya Algorithm, which obtains
a ba in L

as the limit of L
p
-approximations as p . Since bas in L

need not be
unique, this procedure is also a way to select a particular ba from among the possibly
many.
In L

, not only are bas in general not unique, also (27) may have many (linearly
independent) solutions, and this makes the application of the characterization theorem a
bit harder. On the other hand, if M is nite-dimensional and X = C(T), then it is possible
to restrict considerably the set of linear functionals to be considered, namely to those
which are linear combinations of no more than (dimM) + 1 point evaluations. This, and
its pleasant consequences, are detailed in the next section.
The following consequence, of an observation during the proof of Proposition 22, is
of use when the Characterization Theorem 25 is used in a nls with a non-smooth and/or
non-strictly convex norm (such as L

), in which case there may be many linear functionals


of norm 1 parallel to the error and/or many bas.
07feb03 22 c _2003 Carl de Boor
(29) Lemma. X nls, M lss, g X, m M. If M |(gm), then, for any m

T
M
(g),
also |(g m

).
In short, any characterizing linear functional, i.e., any linear functional perpendic-
ular to the approximating space and parallel to the error in some ba to the given g, must
also be parallel to the error in any other ba to that g. At times, this leads to a proof of
uniqueness of the ba.
Best approximation in C(T)
The tool here is the following representation theorem for linear functionals on nite-
dimensional linear subspaces of C(T).
(30) Theorem. If is a linear functional on a lss Y , of dimension n, of the real nls C(T),
with T compact Hausdor, then there exist U T with #U = n and w R
U
so that
=

uU
w(u)
u
Y
and || = |w|
1
.
In other words, every linear functional on an n-dimensional subspace of the real C(T)
has a norm-preserving extension to all of C(T) in the form

U,w
:=

uU
w(u)
u
of a linear combination of no more than n point evaluations.
This representation theorem is germane because our characterization of the elements
m of T
M
(g) involves linear functionals only as they act on Y := M + ran[g]. Indeed, all
the characterization demands is that M |(g m). Hence, whatever C(T)

the
characterization theorem might have dragged in here, we may replace it by the extension
of
Y
to C(T) guaranteed by Theorem 30. Since the original took on its norm on
g m Y , therefore |
Y
| = ||, hence the replacement functional has the same norm
as the original one. This gives the following.
(31) Characterization Theorem for ba in C(T). Let X = C(T) with F = R and T
compact Hausdor, let M be an n-dimensional lss of X, let g X and m M. Then,
m T
M
(g) i there exists U T with #U n+1 and w R
U
so that M
U,w
| g m.
Before exploiting this theorem for the construction of such bas, here is a useful aspect
of such linear functionals
U,w
.
(32) de LaVallee-Poissin lower bound. If 0 ,=
U,w
M and m M with (g
m)(u)w(u) 0, all u U, then
min[(g m)(U)[ dist (g, M) |g m|.
Proof: Set :=
U,w
and recall, e.g. from (26), that M implies [g[
|| dist (g, M), and certainly also (g) = (g m). Hence, with e := g m, we have
min
uU
[e(u)[||

u
[e(u)[[w(u)[ = [e[ = [g[ || dist (g, M),
and division by || = |w|
1
does the rest.
12feb03 23 c _2003 Carl de Boor
This lower bound is the more eective, of course, the closer min[(g m)(U)[ is to
|g m|. The process of constructing such U for given m and, further, such m for given U
is formalized in the Remes Algorithm below. For it, we now discuss related consequences
of the characterization theorem 31.
The fact that the linear functional :=
U,w
in the theorem is to take on its norm on
the error, e := g m, forces equality in the following string of inequalities:
e =

u
w(u)e(u)

u
[w(u)e(u)[ |w|
1
max
uU
[e(u)[ || |e|.
In particular, assuming as we may WLOG that none of the w(u) is zero, this implies that
(33) e(u) = |e| signumw(u), u U.
This says that the error must take on its norm at every point in U, with the sign determined
by the signature of the corresponding weight. On the other hand, these weights are not
arbitrary. Rather, they are determined by the condition that M . Since M is n-
dimensional, the statement M is, in eect, a homogeneous linear system of n equations,
namely the linear system
(34) w Q
U
V = 0,
with
Q
U
: f f
U
and
V := [v
1
, . . . , v
n
]
any basis for M, and with the weight vector w the solution sought. In particular, there
are nontrivial solutions for any choice of U with #U = n+1 (since then this homogeneous
system has more unknowns than equations).
The theorem suggests the following numerical procedure (associated with the name
Remes).
Remes Algorithm
(i) Pick any (n + 1)-set U in T.
Usually, one picks U as the set at which some approximation m M to the given g has
its absolutely largest local extrema.
(ii) Compute a best approximation m
U
to g from M in the discrete norm
||
U
: g max [g(U)[.
Suppose that (a, r) R
n
R solves the linear system
(35) Q
U
V a +r = Q
U
g.
12feb03 24 c _2003 Carl de Boor
Then [r[ = |g V a|
U
, and (g V a) = r|w|
1
, hence, by Theorem 31 applied to C(U),
m
U
:= V a is then a ba to g wrto the norm ||
U
, with |g m
U
|
U
= [r[. (If r happens to
be negative, simply replace w by w, as that wont change the fact that it is a solution of
(35) but will change r to [r[.)
(iii) Update U, making it the set of n+1 points at which g m
U
has its absolutely largest
local extrema, and go to (ii).
The algorithm is quite attractive since, at the end of step (iii), we know (as in (32))
that
[r[ dist (g, M) |g m
U
|,
hence can gauge whether or not it is worth our while to continue. Indeed, the second
inequality is obvious; the rst follows from the fact that [r[ is the error in the ba to g from
M when we only consider the maximum absolute error on the subset U of T.
The only y in the ointment is uncertainty about the solvability of (35).
The Haar condition
Since (35) is a square system, its solvability is equivalent to its unique solvability.
Hence, as g is arbitrary, we are asking that the pointset U be total for M, i.e. Q
U
be
1-1 on M. But we are asking more, we are asking for the solvability of (35), i.e., for the
invertibility of the matrix [Q
U
V, ] and this is the same as its being 1-1 since it is square.
(36) Proposition. If the n + 1-set U T is total for the n-dimensional lss M of C(T),
then (35) has exactly one solution (for every g C(T)).
Proof: Let w be a nontrivial solution of the homogeneous linear system (34), set
= signum(w), and let [Q
U
V, ](a, r) = 0, with a an n-vector and r a scalar. Then
V a = r on U, hence 0 = V a = Q
U
V a = r|w|
1
, which is only possible if r = 0.
However, now Q
U
V a = 0 follows which, by the assumption that U is total for M, implies
that V a = 0, therefore a = 0 (since V is a basis).
Since Q
U
maps into the (#U)-dimensional space R
U
, it cannot be 1-1 unless #U
n = dimM. Further, if U contains exactly n points, then Q
U
also maps M onto R
U
. This
says that M contains, for each g dened at least on U, exactly one m that interpolates
g at U, i.e., that agrees with g on U.
(37) Denition. The n-dimensional linear space M of functions on some domain T is a
Haar space : every n-set U in T is total for M.
(38) Theorem (Haar). If M is a nite-dimensional lss of C(T) (with T compact Haus-
dor), then M is Haar i M is Chebyshev.
Proof: =: Let g C(T), and let m T
M
(g) (such m exists since M is a
nite-dimensional lss). By the Characterization Theorem 31, there exists U T with
#U n + 1 and w R
U
, with w(u) ,= 0 for all u U, so that
M
U,w
|(g m).
12feb03 25 c _2003 Carl de Boor
Moreover, by Lemma 29, every m T
M
(g) must satisfy this condition, hence must satisfy
Q
U
m+ dist (g, M) = Q
U
g.
In particular, if also m

T
M
(g), then Q
U
m = Q
U
m

. Now, since M is Haar, the condition


0 ,=
U,w
M cannot hold unless #U > n (see Lemma 39 below). But, since M is Haar,
this implies that U is total for M, therefore m = m

.
=: If M fails to be Haar, then there exists an n-set U in T and some m M
with |m| = 1 that vanishes on U. Further, there exists w R
U
0 with |w|
1
= 1 so that
:=
U,w
M.
It follows that, by the characterization theorem, any f in
G := g C(T) : |g| 1, Q
U
g = := signum(w)
has 0 as a best approximation from M, and, by the Tietze Extension Theorem, there is at
one such f. In fact, for any such f, also g := (1 [m[)g G, since |m| = 1 and m
U
= 0.
More than that, for any [1 . . 1], g m G since (i) it agrees with g on U, hence
on
U,w
, and (ii)
[g(t) m(t)[ [g(t)[ +[[[m(t)[ 1 [m(t)[ + [m(t)[,
hence |g m| 1. In particular, 0 is a ba to g m from M, therefore m T
M
(g).
Thus, [1 . . 1]m T
M
(g).
Note that
k
(R) is Haar, hence we now know that it is also Chebyshev.
The proof took for granted the following
(39) Lemma. If the n-dimensional linear subspace M of the C(T) is Haar, then 0 ,=

U,w
M implies #U > n.
Proof: If #U < n + 1, then Q
U
V would have full row rank since we could then
extend U to an n-set U

making Q
U
V a submatrix of the matrix Q
U
V which is invertible
since M is Haar and, with that, w Q
U
V = 0 would imply w = 0.
Since the characterizing
U,w
can be chosen with #U dimM +1, it follows that, for
a Chebyshev space M, #U must be equal to dimM +1, meaning in particular that all the
n+1 entries of the weight vector w must be nonzero. Further, once we know this, we dont
really care about the weight vector itself anymore, all we need for checking a proposed ba
m M is the sign vector, := signum(w), since, as we saw earlier, the characterizing
condition merely demands that the error, e := g m, satisfy
(40) e(u) = |e|(u), u U,
with some xed (nontrivial) sign. We now investigate the possible sign vectors .
For this, we now think of U as a sequence, i.e., order the points in U somehow:
U = (u
1
, . . . , u
n+1
) and write, correspondingly, w
j
:= w(u
j
). Then, for the specic nor-
malization w
n+1
= 1, the n-vector (w
j
: j = 1, . . . , n) is the unique solution of the linear
system
? Q
u
1
,...,u
n
V = Q
u
n+1
V.
12feb03 26 c _2003 Carl de Boor
By Cramers rule, this implies that
w
j
= det(Q
u
1
,...,u
j1
,u
n+1
,u
j+1
,...,u
n
V )/ det(Q
u
1
,...,u
n
V ),
or, using column interchanges to restore order here,
(41) w
j
= (1)
nj
det(Q
u
1
,...,u
j1
,u
j+1
,...,u
n+1
V )/ det(Q
u
1
,...,u
n
V ).
If now T is an interval or, more generally, interval-like, i.e., a connected totally ordered
set, then it makes sense to choose the ordering of the points in U accordingly, i.e., to choose
u
1
< < u
n+1
. Moreover, each of the ordered sequences
(u
1
, . . . , u
j1
, u
j+1
, . . . , u
n+1
)
can be connected to the sequence (u
1
, . . . , u
n
) by a continuous transformation [0 . . 1] : t
(t) := (
1
(t) < <
n
(t)), with (1) the former and (0) the latter. It follows that the
corresponding determinants det(Q
(t)
V ) depend continuously on t and none can be zero
since M is Haar. Hence, if the v
j
are all real, then all the determinants in (41) have the
same (positive or negative) sign. In particular, in this case w
j
w
j+1
< 0, all j. This gives
(42) Chebyshevs Alternation Theorem. If M is an n-dimensional Chebyshev sub-
space of the real C([a . . b]), then m T
M
(g) i the error, g m, alternates at least n
times, i.e., i there are points u
1
< < u
n+1
in [a . . b] and an 1, 1 so that
(g m)(u
j
) = (1)
j
|g m|, j = 1, . . . , n + 1.
The argument leading up to this classical theorem brings with it a somber consequence,
called by some frivolous people the Loss of Haar. As soon as T contains a fork, i.e.,
contains three open arcs which have exactly one point in common, then no linear subspace
of dimension > 1 can be Haar. For, in such a setting, we can arrange the continuous map
[0 . . 1]t (
1
(t), . . . ,
n
(t)), from the unit interval to n-sequences with n distinct entries
in T, in such a way that
(0) = (r, s, u
3
, . . . , u
n
), (1) = (s, r, u
3
, . . . , u
n
).
This implies that the determinant corresponding to (0) is the negative of the determinant
corresponding to (1). Hence, if the v
j
are real, then the determinant must be zero for
some (t), and the corresponding n-set fails to be total for M. We have proved
(43) Loss of Haar. If M is a nite-dimensional real linear subspace of dimension > 1 of
C(T) (with T compact Hausdor), and T contains a fork, then M cannot be Chebyshev
(since it cannot be Haar).
In particular, no polynomial space of dimension > 1 in more than one variable can be
Haar. This has made the construction of uniform best approximations to functions of sev-
eral arguments, even by polynomials of low degree, something of an art. Correspondingly,
it has encouraged the development of near-best methods of approximation.
12feb03 27 c _2003 Carl de Boor
(44) Theorem (Newman-Shapiro). Best approximation from a nite-dimensional lin-
ear subspace M of the real X := C(T) is strongly unique. Explicitly, for every g X,
there exists > 0 so that
(45) f M |f g| |f P
M
g| +|f P
M
g|.
Proof: For g M, = 1 will do. For g XM, let M |g P
M
g, with
=
U,w
for some U T with #U n +1. Set := signum(w). For any f M and any
u U,
|g f| (u)(g f)(u) = (u)(g P
M
g)(u) +(u)(P
M
g f)(u),
hence
|g f| |g P
M
g| +K

(P
M
g f),
with
K

(f) := max
uU
(u)f(u).
Since K

is positive homogeneous, (45) follows with


:= infK

(m) : m M, |m| = 1.
It remains to show that > 0. Since K

is continuous and M is nite-dimensional,


can nd f M with |f| = 1 and = K

(f). Since 0 =
U,w
f =

uU
w(u)f(u) =

uU
[w(u)[(u)f(u), K

(f) = 0 would imply f(u) = 0 for all u U. However, M is


Chebyshev, hence Haar, hence, by Corollary 39, U is total for M, and now f = 0 would
follow, contradicting the assumption that |f| = 1.
Complex C(T)
We pointed out earlier that the general characterization theorem for bas involves
linear functionals parallel to the error e := g m because they provide the gradient of the
norm at the point e. Kolmogorovs characterization theorem is more explicitly based on
this idea of a directional derivative of the map f |e f|. The criterion is of interest
because it is also valid when F = C. It formalizes the following idea: if, in the max-norm,
we want |e +f| < |e|, then, at all points t at which e takes on its norm, e(t) +f(t) needs
to be less than [e(t)[ in absolute value. Hence, if e and f are complex-valued, we need that
Re e(t)f(t) < 0. Moreover, this must be so uniformly over the set
extr(e) := t T : [e(t)[ = |e| = argmax[e(T)[.
(46) Lemma. Let X = C(T) with T compact Hausdor, and F = R or C, and e, f X0.
Then, |e +f| < |e| for some positive i
K
e
(f) := max
textr(e)
Re(e(t)f(t)) < 0.
17feb03 28 c _2003 Carl de Boor
Proof: |e +f|
2
|e +f|
2
(extr(e)) |e|
2
+2K
e
(f), hence, as K
e
is real homo-
geneous, |e +f| < |f| for some > 0 implies K
e
(f) < 0.
Conversely, if K
e
(f) < 0, then
G := t T : Re(e(t)f(t)) < K
e
(f)/2
is open and contains extr(e) since, for t extr(e), we even have Re(e(t)f(t)) K
e
(f).
Therefore, |e|(TG) < |e|, and this implies that
|e +f|(TG) |e|(TG) +[[|f|(TG)
is less than |e| for all close to 0. But, by the strict negativity of K
e
(f), also
|e +f|
2
(G) |e|
2
+ 2K
e
(f)/2 +[[
2
|f|
2
is less than |e|
2
for all positive close to 0. So, altogether, |e +f| < |e| for all positive
near 0.
Since |g (m f)| = |(g m) + f|, this lemma has the following very useful
consequence.
(47) Kolmogorov Criterion. Let X = C(T) with T compact Hausdor and F = R or
C, and let M be a lss of X, let g X and m M. Then, m T
M
(g) if and only if
f M, K
gm
(f) 0.
Here is Alpers example: T = z C : [z[ 1 is the unit disc in the complex plane;
we take M =
k
, and g = 1/( ), with CT. Consider the function
m : z
1
z
cz
k
z 1
z
=
1 cz
k
(z 1)
z
.
This is a polynomial (in z) i the numerator of the last expression vanishes at z = , i.e.,
i c =
k
/([[
2
1). With that choice, m
k
, and
e := g m = z
k
z 1
z
,
hence extr e = z : [z[ = 1. Let f
k
and consider K
e
(f). Having it nonnegative
requires that
Arg(e(z)f(z)) = Arg(e(z)) Arg(f(z)) [/2 . . /2]
for some [z[ = 1. Now, Arg(e(z)) = const +k Arg(z) +Arg(z
1
) Arg(z ), and this
increases by 2(k + 1) as we circumnavigate the unit disc. On the other hand, f, being a
polynomial of degree k, can have at most k zeros in the unit disc, hence Arg(f(z)) can
increase by at most 2k. And that does it.
17feb03 29 c _2003 Carl de Boor
L
1
The continuous linear functionals on X := L
1
(T; ) are of the form f
_
T
hf, with
h L

(T). Further,
_
T
hf |h|

|f|
1
,
with equality i h = |h|

signum(f) o the set


Z
f
:= t T : f(t) = 0.
(In particular, most continuous linear functionals on L
1
(T) do not take on their norm,
and, even on
1
, some dont (cf. p.11).) In general, the zero-set Z
f
is only determined
up to sets of measure 0 since it depends, of course, on the particular representer of fs
equivalence class used for its determination.
(48) Characterization Theorem. Let M be a linear subspace of the nls X = L
1
(T; ),
with a non-atomic measure, let g X and m M. Then, m T
M
(g) i there exists a
h L

(T) with |h|

= 1 that is perpendicular to M and agrees with signum(g m) o


Z
gm
.
In particular, if Z
gm
has measure zero, then m T
M
(g) i signum(g m) M.
The theorem suggests a quick try at constructing bas from an n-dimensional lss of
X = L
1
([a . . b]): Assuming that Z
gm
has measure zero, we look for a sign function,
i.e., a real function h with [h[ = 1, the simpler the better, i.e., the fewer breakpoints the
better. The condition h M is, in eect, a system of equations for the breakpoints of h,
hence, in general, we would not expect to succeed with fewer than n (interior) breakpoints.
Now suppose we have succeeded, and now have in hand such a sign function with exactly
n interior breakpoints, a <
1
< <
n
< b say. If g and the elements of M are
continuous (or, at least piecewise continuous), then we can now hope to interpolate to g at
by elements of M. Let m be the resulting interpolant. Then, usually, the error, g m,
changes sign at the points of interpolation, hence usually, signum(g m) = h, and we
have found a ba to g from M.
The fact that, for every n-dimensional lss M of L
1
([a. . b]), there exists a sign function
orthogonal to it with n breakpoints is called the Hobby-Rice Theorem. Its original proof
was quite involved. While a postdoc here in Madison, Allan Pinkus pointed out that it is
a ready consequence of Borsuks Theorem. Since the latter has played a major role in the
discussion of n-width, here is its statement.
(49) Borsuks Theorem. Let f be a continuous map from the unit sphere S
n
in R
n+1
to R
n
. If f is odd, i.e., f(x) = f(x), all x S
n
, then 0 f(S
n
).
(50) Hobby-Rice Theorem. For every n-dimensional lss M of X = L
1
([a . . b]), there
exists a sign function orthogonal to it with n breakpoints.
Proof: Assume WLOG that [a . . b] = [0 . . 1], and, for s in
S
n
:= (s
0
, . . . , s
n
) R
n+1
:

j
s
2
j
= 1,
17feb03 30 c _2003 Carl de Boor
dene the linear functional
s
on X as follows: With
j
(s) :=

i<j
s
2
i
,

s
f :=
n

j=0
signum(s
j
)
_

j+1
(s)

j
(s)
f .
Then
s
f =
_
1
0
h
s
f, with h
s
a sign function with at most n breakpoints (since we may
ignore
j
(s) if s
j1
s
j
0). Further,
s
=
s
and, for xed f, the map S
n
R : s

s
f is continuous, since, for any f L
1
[0 . . 1], the map [0 . . 1]
2
F : (u, v)
_
v
u
f is
continuous. Hence, with V = [v
1
, . . . , v
n
] any basis for M, the map
f : S
n
R
n
: s (
s
v
j
: j = 1, . . . , n)
is continuous and odd, therefore Borsuk tells us that f(s) = 0 for some s S
n
. The
corresponding sign function, h
s
, provides what we are looking for.
(51) Corollary (Krein). No (nontrivial) nite-dimensional lss M of X = L
1
([a . . b]) is
a Chebyshev space.
Proof: Let h be a sign function orthogonal to M. Take any f M0 and set
g := h[f[. Then, for any [0 . . 1],
h(g f) = [f[ hf (1 )[f[ 0,
showing that M
_
h | (g f), hence [0 . . 1]g T
M
(g).
Of course, this does not imply that every g XM has many bas from M. E.g.,
g :=
[..1]
has exactly one ba from M :=
0
, except when = 1/2 in which case
[0 . . 1]()
0
T
M
(g).
near-best approximation
While the questions concerning best approximations raised in the rst lecture (such
as existence, uniqueness and characterization) occupy a good part of a standard course in
Approximation Theory, best approximations are hardly ever calculated because they can
usually only be obtained as the limit of a sequence of approximations, and because of the
easy availability of near-best approximations. To recall, the bounded linear map A on the
nls X provides near-best approximations from the subset M if
|g Ag| const dist (g, M), g X.
Such A is, necessarily, a linear projector onto M, i.e., ranA = M and A
2
= A, i.e.,
A
M
= 1. In particular, M is a linear subspace.
Conversely, if A is a bounded linear projector on X with range M, then g Ag =
(1 A)g = (1 A)(g m) for all m M, hence
(1) dist (g, M) |g Ag| |1 A| dist (g, M), g X.
17feb03 31 c _2003 Carl de Boor
This is Lebesgues inequality.
This raises (at least) two questions:
(i) How small can one make |1 A| by proper choice of A?
(ii) What can be said about dist (g, M) in terms of some information we might have about
g?
Example: broken line interpolation With X = C[a. . b] for some nite interval
[a . . b], recall the broken line interpolant A = P

g for given = (a =
1
< <
l+1
= b).
Its range is the space
M =
0
1,
of broken lines with break sequence . Since |P

| = 1, and always |1 A| 1 +|A|, we


get in this case
dist (g,
0
1,
) |g P

g|

2 dist (g,
0
1,
).
In particular, the construction of a ba from
0
1,
will, at best, cut the maximum error in
half.
Incidentally, the bound used here, |1 P

| 1 + |P

|, is sharp as can be seen by


looking at an g that is the broken line with breakpoints
1
, . . . ,
l+1
at which it has the
value 1 and one additional breakpoint at which it has the value 1.
In particular, we can get a good estimate for dist (g,
0
1,
) by looking at |g P

g|

.
In Lebesgues proof of Weierstra, we already observed that
|g P

g|


g
([[).
In particular, if g L
(1)

[a . . b], i.e., g is absolutely continuous with essentially bounded


rst derivative, then
g
(h) h|Dg|

, hence
|g P

g|

[[|Dg|

, g L
(1)

[a . . b].
Further, since P

reproduces all elements of


0
1,
, we can replace here g by an arbitrary
element of
0
1,
and so obtain, more precisely,
|g P

g|

[[ dist (Dg,
0,
), g L
(1)

[a . . b].
Since dist (g,
0,
)
g
([[), we therefore obtain
|g P

g|

[[
Dg
([[), g C
(1)
[a . . b],
and, so, nally,
|g P

g|

[[
2
|D
2
g|

, g L
(2)

[a . . b].
Actually, since, for
j
t
j+1
, g(t) P

g(t) = (t
j
)(t
j+1
)[
j
,
j+1
, t]g (with
[, . . . , ]g the divided dierence of g at the point sequence (, . . . , )), we have the sharper
estimate
|g P

g|

[[
2
|D
2
g|

/8, g L
(2)

[a . . b].
17feb03 32 c _2003 Carl de Boor
You see how, as a function of the mesh size, [[, these estimates improve, i.e., go to
zero faster as [[ 0, when we restrict g to smoother and smoother function classes.
However, further more restrictive smoothness assumptions will not lead to an increase in
the rate at which the interpolation error goes to zero with [[. E.g., if |g P

g| = o([[
2
)
for arbitrary (or even just for
N
= (a, a + h, , b h, b) with h := (b a)/N and
N N) as [[ 0, then g
1
. Proof idea: for any collection of quasi-uniform ,
i.e., sup

sup
i,j

i
/
j
< , must have [
j
, t,
j+1
]g 0 uniformly for t (
j
. .
j+1
)
and j as [[ 0 while, for any renement s of the sequence (a, t, b), the second divided
dierence [a, t, b]g can be written as a convex combination of the [s
i
, s
i+1
, s
i+2
]g, hence
must be zero.
This is a simple illustration of our next topic, degree of approximation.
If there is time, I may come back to item (i), i.e., the question of how small one can
make |A| or |1 A| for given M by appropriate choice of the linear projector A onto
M. To whet your appetite, I mention that every linear projector on C[a . . b] onto
n
has
norm no smaller than ln n. This is related to the fact that the projector s
n
providing
the truncated Fourier series (mentioned earlier) is the unique projector of minimal norm
on C(TT) onto

n
, and |s
n
| (2/) lnn + const.
Remark I am getting tired of adapting earlier notes by making sure that the given
element of X to be approximated is denoted by g. So, from now on, the given element to
be approximated will be denoted by f. Life is short.
Degree of Approximation
Given a sequence (M
n
) = (M
n
: n N) of approximating sets in the nls X and f X,
one is interested in
n E
n
(f) := dist (f, M
n
)
as n .
In this generality, nothing can be said. The following general model (from Chapter
7 of DeVore and Lorentz) covers most situations of practical, and even most situations of
theoretical, interest.
(i) M
1
= 0, and (M
n
) is increasing, i.e., (M
n
) is a ladder.
This guarantees that E
n
(f) is nonincreasing, with E
1
(f) = |f|.
(ii)
n
M
n
is dense in X.
This is equivalent to having lim
n
E
n
(f) = 0 for all f X.
In view of approximation by rationals, or exponentials with free frequencies, or splines
with free knots, it would be too much to assume that the M
n
are linear subspaces. But it
is ok to assume
(iii) FM
n
:= m : F, m M
n
M
n
.
(iv) cn M
n
+M
n
M
cn
.
Finally, although clearly not essential, the following assumption will avoid a certain
amount of epsilontics.
(v) Each M
n
is an existence set.
19feb03 33 c _2003 Carl de Boor
In particular, each M
n
is closed.
To be sure, Assumption (iv) is somewhat restrictive and precludes some practically
important ladders such as the following: for some T R
2
,
M
n
:=
k,n
T
consists of all piecewise polynomial functions, on T, of some degree k and involving no
more than n polynomial pieces. In other words, for each f
k,n
there is a partition
of T into at most n subsets and, on each such subset, f agrees with some polynomial
of degree k. For this example, we have only M
n
+ M
n
M
n
2. Finding the degree
of approximation from this ladder (
k,n
: n N) is one of the outstanding problems
in nonlinear approximation. (Strictly speaking, this particular example lacks that trivial
initial space, 0, which is really only used in the general theory to simplify notation. For
that, we might have to set here M
n
=
k,n1
. Such a switch, from n to n 1, does not
change the orders n
r
of interest here. Also, its subladder (
k,2
n : n N) does satisfy
(i)-(v).)
Even with these assumptions in place, the best we can say about E
n
(f) for given
f has already been said: E
n
(f) converges monotonely to 0. The question of interest is
just how fast or slow this convergence is. As with all measuring, these terms are made
precise by comparing E
n
(f) with certain standard sequences. The most popular of these
are the sequences (n

: n N) for some real . Thus we are looking for conditions on f


that guarantee that
E
n
(f) = O(n
r
),
i.e., limsup
n
E
n
(f)n
r
< , or, perhaps, even
E
n
(f) = o(n
r
),
i.e., lim
n
E
n
(f)n
r
= 0, or

n
n
r
E
n
(f) < .
Such conditions may single out a rather small subset of X in case X is complete but not
equal to any of the M
n
. This is certainly so if the constant in (iv) is c = 1, i.e., in case the
M
n
are linear subspaces.
(52) Proposition (H. S. Shapiro). If (M
n
) is a sequence of proper closed linear sub-
spaces (i.e., only (iii) as is, (i) is weakened to proper, and (v) is weakened to closed, but
(iv) with c = 1), then, for any real sequence (
n
) converging monotonely to 0, the set
A

:= f X : E
n
(f) = O(
n
)
is thin in the sense that it is of rst category, i.e., the countable union of nowhere dense
sets. In particular, it cannot equal X in case X is complete.
Proof: A

=
n
0
,NN

nn
0
B

N
n
(M
n
), with
B

Nr
(M
n
) := f X : dist (f, M
n
) Nr = NB

r
(M
n
)
19feb03 34 c _2003 Carl de Boor
since FM
n
M
n
. If now, for some r > 0 and some x, B
r
(x) B

n
(M
n
), then B
r
(0)
(B
r
(x) + B
r
(x))/2 (B

n
(M
n
) + B

n
(M
n
))/2 and, even if we only knew (iv), we
could now conclude that B
r
(0) B

n
(M
cn
) and if, as we actually assume, the M
n
are
proper closed linear subspaces, this implies (by Riesz Lemma) that r
n
, hence B
r
(x)

nn
0
B

n
(M
n
) implies r lim
n

n
= 0, hence A

is of rst category. However, if the


M
n
are not linear spaces, then no such conclusion can be drawn.
Example Here is a nice example, provided by Olga Holtz, to show that properties
(i)-(v) by themselves are not strong enough to imply this propositions conclusion. Let
X :=

. For any closed V F := R, let


F
V
:= f X : ran f V .
Then, for any f X, dist (f, F
V
) = |f f
V
|, with f
V
: n argmin
vV
[f(n) v[.
Therefore, for any closed W,
dist (f, F
V
) dist (f, F
W
) + dist (W, V ).
Let now
F
k
:=
#V k
F
V
,
and let (f
m
) be a minimizing sequence in F
k
for |f |. Then (ran f
m
) is a bounded se-
quence of subsets in R of cardinality k, hence, AGTASMAT (:= After Going To A Subse-
quence May Assume That) there is some V R with #V k for which lim
m
dist (ranf
m
, V ) =
0. Therefore
dist (f, F
k
) dist (f, F
V
) dist (f, F
ranf
m
) + dist (ranf
m
, V )
m
dist (f, F
k
),
showing f
V
to be a ba to f from F
k
, hence showing F
k
to be an existence set. Also, with
V
k
:= 1 (2j 1)/k : j = 1, . . . , k, we have
(53) dist (f, F
k
) dist (f, F
fV
k
) |f|/k,
and, since also F
k
is homogeneous, and is increasing with k, the sequence (F
k
) satises
conditions (i)-(v), except that, at best, F
k
+ F
k
F
k
2. But (iv) is satised (with c = 2)
by its subsequence M
n
:= F
2
n, n = 1, 2, . . ., (with M
0
= 0, of course), along with the
other conditions in (i)-(v), yet, by (53),
f X : dist (f, M
n
) = O(2
n
) = X,
and this is not of rst category since X is a Bs.
Put positively, this example also shows the power of nonlinear approximation, i.e.,
approximation from nonlinear M
n
.
03mar03 35 c _2003 Carl de Boor
Degree of Approximation quantied
As is customary in mathematics, we express suitable conditions on f as membership in
some set Y . A typical choice for Y is a semi-normed ls whose semi-norm we denote by ||
Y
that is continuously imbedded in X. The standard example has M
n
=

n
X := X
p
,
with
X
p
:=
_
L
p
(TT), p < ;
C(TT), p = ,
and
Y = X
()
p
:= f X
p
: |f|
Y
:= |D

f|
p
< ,
with some positive integer.
One hopes to choose Y and r in such a way that, simultaneously,
(54) C
J
n N, f Y E
n
(f) C
J
n
r
|f|
Y
and
(55) C
B
n N, g M
n
|g|
Y
C
B
n
r
|g|.
For historical reasons, the former is called a Jackson, or direct, inequality or estimate,
the latter a Bernstein, or inverse, inequality or estimate. The Jackson inequality gives
a lower bound on the speed with which E
n
(f) goes to zero, and, as we shall see, the
(historically earlier!) Bernstein inequality provides an upper bound, at least indirectly. The
most natural bridge between Jackson and Bernstein inequalities turns out to be Peetres
K-functional:
K(f, t) := K(f, t; X, Y ) := inf
gY
(|f g| +t|g|
Y
).
This functional measures how smooth f is in the sense that it tells us how closely we
can approximate f by smooth elements. The larger t, the more stress we lay on the
smoothness of g. As a function of t, K(f, t) is the inmum of a collection of straight lines,
all with nonnegative slope, hence K(f, .) is also weakly increasing (i.e., nondecreasing), and,
further, is concave (as would be the inmum of any collection of straight lines, increasing
or not). This implies that t K(f, t)/t is (weakly) decreasing (i.e., nonincreasing).
(56) Peetres Theorem.
(i) Jackson (54) implies sup
n
E
n
(f)/K(f, n
r
) < .
(ii) Bernstein (55) implies
(57) K(f, n
r
) const
r
n
r
|(E
k
(f) : k n)|
(r)
.
Here, we use the following somewhat complicated weighted sequence norm
|a|
(r)
:=
n

k=1
[k
r
a(k)[/k, a F
n
.
19feb03 36 c _2003 Carl de Boor
This norm is monotone, meaning that [a[ [b[ implies |a|
(r)
|b|
(r)
. Further, for the
following standard null-sequence a = (a
k
= k
(r1+)
: k n), one computes
(58) |(k
(r1+)
: k n)|
(r)
=
n

k=1
k

n
1
for 0 < 1,
with the convenient abbreviation
b
n
c
n
: b
n
= O(c
n
) and c
n
= O(b
n
).
(In fact, this suggests the nonstandard abbreviation:
b
n

<
c
n
: b
n
= O(c
n
),
which I will use occasionally, e.g., right now.) Here is a proof of (58):
n
1

<
_
n+1
1
()

k=1
k

_
n
0
()

<
n
1
.
Here is a quick comment concerning the fact that, in the description of |a|
(r)
, I did not
combine the two powers of k appearing there. The reason is the following. The argument
to follow remains valid even when || and/or ||
Y
are only quasi-norms, meaning, in
eect, that, instead of the triangle inequality, we only have
|x +y|

|x|

+|y|

for some and all x and y. In such a case, one considers, more generally, the weighted
sequence norm
|a|
(r)

:=
_
n

k=1
[k
r
a
k
[

/k
_
1/
.
With these notations now fully explained, we read o from Peetres theorem that, for
[0 . . 1), E
n
(f)

<
n
(r1+)
if and only if K(f, n
r
)

<
n
(r1+)
. In the standard
situation, the latter can be shown to be equivalent to having f C
(r1)
with D
r1
f in
Lip

.
Proof: (i): For any g Y , E
n
(f) |f g| + E
n
(g) C(|f g| + n
r
|g|
Y
),
with C := max1, C
J
, and taking the inf over g Y does it.
(ii): The trick here is the use of a telescoping series. (The proof in DeVore and Lorentz
is a touch terse.) With f
k
T
M
k
(f)
, all k, we choose a sequence
1 = m
0
< m
1
< < m

= n
and consider
h

:= f
m

f
m
1
M
cm

.
19feb03 37 c _2003 Carl de Boor
Then
|h

| |f f
m

| +|f f
m
1
| 2E
m
1
(f),
while f
n
=

=1
h

. Therefore,
K(f, n
r
) |f f
n
| +n
r
|f
n
|
Y
E
n
(f) +n
r

=1
|h

|
Y
n
r
_
n
r
E
n
(f) +C
B
c
r
2

=1
m
r

E
m
1
(f)
_
n
r
const
r

=0
m
r

E
m

(f),
with const
r
:= 2C
B
c
r
max

(m

/m
1
)
r
a constant depending only on r (and the constants
C
B
and c) if we restrict attention to sequences (m

) for which max

/m
1
const.
Now, for any positive nonincreasing sequence (a
k
) (such as a
k
= E
k
(f)),
a
J

j<kJ
k
r1

j<kJ
k
r1
a
k
a
j+1

j<kJ
k
r1
,
while

j<kJ
k
r1

_
J
j
()
r1
= J
r
1 (j/J)
r
r
= j
r
(J/j)
r
1
r
.
Hence, applying this with j = m

, J = m
+1
, all , we get
1

=0
m
r

E
m

(f) const
r
n

k=1
k
r1
E
k
(f)
provided we also insist that min

(m

/m
+1
) const < 1, i.e., the ratio m

/m
1
required
earlier to be bounded must, on the other hand, be bounded away from 1. Choosing the
sequence (m

: = 1, . . . , ) to be roughly dyadic, i.e., m

2
1
, all , will do for both
requirements. With this, we get (57), with const depending on C
B
and r only.
For later use, I record the following just proved.
(59) Lemma. If a : N R
+
is nonincreasing, then, for any r 1,
2
n

k=1
k
r1
a
k

n

=1
2
r
a
2
,
with constants that only depend on r.
We will make use of (59) in a moment, in Bernsteins proof that only smooth functions
can be approximated well by trigonometric polynomials.
19feb03 38 c _2003 Carl de Boor
Bernstein estimates for trig.pols.
The argument just given is very free and easy with constants. Its only purpose is to
establish statements about the degree of convergence, such as the assertion that o(n
r
) ,=
E
n
(f) = O(n
r
). Rather dierent arguments are used to establish statements of the sort
that
E
n
(f) const n
r
|f|
Y
, f Y,
with const the smallest possible constant independent of f and n.
The rst such theorem was proved by Bernstein, using
Bernstein Inequality. For any f

n
C(TT),
|Df| n|f|.
Proof: Here is a version of v.Golitscheks proof of Szegos stronger inequality
(60) (Df)
2
+ (nf)
2
(n|f|)
2
, f

n
.
Since

n
is translation-invariant and dierentiation commutes with translation, it is su-
cient to prove that
(61) (Df(0))
2
+ (nf(0))
2
(n|f|)
2
, f

n
.
As the inequality is homogeneous in f, we may assume that Df(0) 0. Now let r > |f|,
let be the unique point in ( . . )/(2n) at which r sin n = f(0), and consider the
trig.pol.
s := f r sin n( +).
Since |f| < r, s alternates in sign at the extrema of sin n( +) and, as there are 2n such,
s has exactly one zero between any two adjacent extrema of sin n( + ). In particular,
one of these zeros must be the point 0 since we chose to make it so. If now Df(0) >
D(r sin n( +))(0), then f would rise above r sin n( +) as we move to the right from 0,
yet is certain to be below it again by the time we reach the rst extremum of sin n( +)
and this would imply an impossible second zero of s between the two extrema that bracket
0. Consequently,
0 D(r sin n( +))(0) = Df(0) rncos n = rn
_
1 (f(0)/r)
2
,
using the fact that, by our choice of , sin n = f(0)/r. Squaring and reordering terms
gives
(Df(0))
2
+ (nf(0))
2
(rn)
2
,
and, since r > |f| is arbitrary here, (61) follows.
03mar03 39 c _2003 Carl de Boor
We are ready to prove a sample inverse theorem.
(62) Theorem. Let M
n
=

n
, as a subset of C(TT). Let f C(TT) and assume that, for
some r,
(63) |(E
n
(f) : n N)|
(r)
<
(as would be the case in case E
n
(f) = O(n
r
) for some positive ). Then f C
(r)
(TT).
Proof: Since dierentiation is a closed operator, it is sucient to show that f
is the uniform limit of a sequence that is Cauchy in C
(r)
(TT). With p
n
T
M
n
(f), all
n, we have f = lim
n
p
n
; however, we have no way of knowing that the whole sequence
(p
n
: n N) is Cauchy in C
(r)
. The sequence (p
2
n : n N), on the other hand, is seen to
be Cauchy, as follows.
|D
r
(p
2
J p
2
j )|
J

k=j+1
|D
r
(p
2
k p
2
k1)|

k=j+1
(2
k
)
r
|p
2
k p
2
k1|

k=j+1
(2
k
)
r
2E
2
k1(f)
2const
r
2
J

k=2
j
k
r1
E
k
(f),
the last inequality by (59). By assumption, this last sum goes to zero as j, J .
This theorem is quite remarkable since it says that even if f is very smooth except on
a small subinterval of TT, it will be hard to approximate f well by trig.pols.
There are similar results for approximation by algebraic polynomials on [a. . b], except
for the following hitch. In contrast to TT, the points in [a . . b] are not all equal. Taking for
simplicity the max-norm on [a . . b], one has the Jackson inequality (see below)
(64) dist

(f,
n
) =: E
n
(f)

constn
1
|Df|

.
However, the best inequality relating |f| and |Df| for f
n
C([a. . b]) is the Markov
Inequality
(65) |Df|


2
[b a[
n
2
|f|

,
(note that, because of the possibility of dilating, the interval length must gure in exactly
the position at which it appears, hence the inequality is sharp since it becomes equality
for [a . . b] = [1 . . 1] and f : t cos(ncos
1
t) the Chebyshev polynomial of degree n)
and these two inequalities do not at all match in the sense of Theorem 56. It is possible,
though, to prove that f C([a . . b]) must be in C
(r)
(I) for all closed subintervals I of
(a . . b) in case E
n
(f)

= O(n
r
) for some positive .
03mar03 40 c _2003 Carl de Boor
Jackson estimates for trig.pols.
Because

n
is translation-invariant, it is easier to prove Jackson inequalities for M
n
=

n
C(TT) than for
n
C([a . . b]) (
n
is translation-invariant only on R or C). The
standard argument uses integral operators of convolution type. With L
n
functions still
to be specied, one considers the approximation
L
n
f : t
_
TT
L
n
(t s)f(s) ds =
_
TT
f(t s)L
n
(s) ds.
One makes the following assumptions:
(i) L
n

n
.
Then,

n
being translation-invariant, i.e., p

n
, s TT p( s)

n
, we neces-
sarily have
L
n
(t s) =

j
(t)
j
(s),
with [
j
: n j n] any basis for

n
. Therefore
L
n
f =

j
_
TT

j
f

n
.
(ii)
n
:=
_
TT
L
n
,= 0, hence wlog (i.e., after replacing L
n
by L
n
/
n
,
_
TT
L
n
= 1, therefore
L
n
()
0
= ()
0
.
It follows that
(f L
n
f)(t) =
_
TT
_
f(t) f(t s)
_
L
n
(s) ds =
_
TT
(
s
f)(t) L
n
(s) ds,
with

h
f := f f( h) =:
h
f( h).
Note that
|
s
f| sup
0h<|s|
|
h
f| =: (f, [s[),
with (f, ) the (uniform) modulus of continuity of f. One could now try to get an
error estimate involving (f, 1/n) by making use of the fact that
(f, [s[) = (f, n[s[1/n) (1 +n[s[)(f, 1/n)
(using that (f, ) is nondecreasing and subadditive, hence (f, [s[) = (f, (n[s[)/n)
(f, n[s[/n) n[s[(f, 1/n) (1 +n[s[)(f, 1/n)), hoping that
sup
n
_
TT
(1 +n[s[)[L
n
(s)[ ds < .
03mar03 41 c _2003 Carl de Boor
However, the L
n
usually employed satisfy the following additional assumption:
(iii) L
n
even.
and, with this, we even have
f(t) L
n
f(t) =
_

0
_
f(t) f(t +s) + f(t) f(t s)
_
L
n
(s) ds
=
_

0
(
s

s
f)(t)L
n
(s) ds,
with

h
f =
2
f( h),
and
|
2
s
f| sup
h<|s|
|
2
h
f| =:
2
(f, [s[),
and
2
(f, ) called the second (uniform) modulus of smoothness of f. In this lan-
guage, (f, ) is called the rst (uniform) modulus of smoothness of f. It is clear
how one would dene the rth modulus of smoothness, for any natural r, as are the bounds

r
(f, h) := sup
k<h
|
r
k
f| 2
r
(f, h)
and

r
(f, [s[) (1 +n[s[)
r

r
(f, 1/n),
with the latter using the fact that
mh
f =

m1
j=0

h
f( +jh), hence

r
mh
f =
m1

j
1
=0

m1

j
r
=1

r
h
f( + (j
1
+ +j
r
)h).
Since, in particular,
2
(f, h) =
2
(f, (nh)/n) (1 + nh)
2

2
(f, 1/n), a natural as-
sumption now is the following moment condition:
(iv) sup
n
_

0
(ns)
k
[L
n
(s)[ ds < , i.e.,
_

0
()
k
[L
n
[ = O(n
k
), for k = 0, 1, 2.
and with it, we get the typical Jackson inequality
|f L
n
f| c
2
(f; 1/n).
It remains to nd suitable L
n
.
The simplest choice is L
n
= D
n
/
_
TT
D
n
with
D
n
(t) :=

|j|n
e
ij
/2 =
sin((n + 1/2)t)
2 sin(t/2)
Dirichlets kernel. For it, L
n

n
, D
n
()
0
= , and L
n
even. However, L
n
decays
too slowly away from 0 to satisfy (iv); in fact, already k = 0 gives trouble since |D
n
|
03mar03 42 c _2003 Carl de Boor
_
TT
[D
n
[
_

1/n
()
1
ln n, hence, as already mentioned earlier, L
n
cannot converge to 1
pointwise on C(TT), hence a more sophisticated analysis wouldnt help here, either.
The next choice is L
n
= F
n
/
_
TT
F
n
, with
F
n
(t) :=
1
2(n + 1)
_
sin((n + 1)t/2)
sin(t/2)
_
2
Fejers kernel. One checks that
F
n
=
n

0
D
k
/(n + 1),
hence F
n

n
, and F
n
()
0
= . Also, F
n
0, hence we only need
_

0
(1 +ns)
2
F
n
(s) ds = O(1).
However this, unfortunately, is not true. Thus, while, by Korovkin, F
n
f converges to f
uniformly for every f C(TT), we dont, ohand, get a Jackson estimate from it since (iv)
does not hold for it.
Jacksons choice is the Jackson kernel
J
n
(t) :=
_
sin(mt/2)
sin(t/2)
_
4
/
n
, m := n/2 + 1,
normalized to have
_
TT
J
n
= 1. It is a special case of the generalized Jackson kernels:
J
n,r
(t) :=
_
sin(mt/2)
sin(t/2)
_
2r
/
n,r
, m := n/r + 1.
Since
_
sin(mt/2)
sin(t/2)
_
2
/m = 1 + 2
m1

k=1
(1 k/m) cos kt =

k
B
2
((k/m) + 1)e
ik
(t)
(as one veries, with B
2
the cardinal B-spline of order 2, a fact to be explored later), it
follows that J
n,r
is an even, nonnegative, trig.pol. of degree n. In particular J
n,r
()
0
> 0,
hence the 0th moment condition is trivially satised. For the others, one may prove that
J
n,r
satises the moment conditions
sup
n
_

0
(nt)
k
J
n,r
(t) dt < , k = 0, 1, . . . , 2r 2,
as follows.
10mar03 43 c _2003 Carl de Boor
Since t/ sin(t/2) t/2 on [0 . . ],
_

0
t
k
_
sin(mt/2)
sin(t/2)
_
2r
dt
_

0
t
k
_
sin(mt/2)
t
_
2r
dt
=
_
2
m
_
k2r+1
_
m/2
0
u
k
_
sin u
u
_
2r
du
n
2r1k
_

0
u
k
_
sin u
u
_
2r
du
n
2r1k
as long as k 2r < 1, hence, in particular,
n,r
n
2r1
and so
_

0
(ns)
k
J
n,r
(s) ds n
k
n
2r1k
/n
2r1
1,
for k = 0, . . . , 2r 2.
Thus, for r 2, the sequence L
n
= J
n,r
satises all four conditions (i)(iv) above.
For r = 2, Jackson got in this way the Jackson Inequality. For r > 2, the additional
moment conditions provide the inequalities (due to Stechkin):
E
n
(f) |f J
n,r
f|
p
C
r

2r2
(f, 1/n)
p
,
for any f L
p
(TT) and any 1 p .
The constant in the resulting estimate E
n
(f) const
2r2
|D
2r2
f|
p
/n
2r2
is far from
sharp. Favard showed that, with S the unit ball of the semi-normed lss Y := L
(r)
p
(TT) of
X
p
(TT), with semi-norm
|f|
Y
:= |D
r
f|
p
,
sup
fS
E
n1
(f) = E
n1
(B
r
p
) K
r
/n
r
,
with this inequality exact for p = 1, , and the numbers K
r
, the so-called Favard con-
stants, given as the value of a fast converging series whose value, for large r, is indistin-
guishable from /4.
The extremal function, B
r
p
, is a Bernoulli spline, of which, perhaps, more anon.
We now know that (54)Jackson and (55)Bernstein hold for M
n
=

n
and Y =
C
(r)
(TT) C(TT) = X. Hence, (56) Peetres Theorem now tells us that
K(f, n
r
; C(TT), C
(r)
(TT)) E
n
(f) |D
r
f|

/n
r
, f C
(r)
(TT).
10mar03 44 c _2003 Carl de Boor
Jackson estimates for alg.pols.
It is standard to prove Jacksons theorem for algebraic polynomials from the one for
trigonometric polynomials. For this, one rst translates and dilates [a. . b] into the interval
[1 . . 1]. Then one considers the map
F : C[1 . . 1] C(TT) : f Ff : f(cos()).
(Draw the picture.) This is a linear map that carries C[1 . . 1] isometrically into
C(TT)
e
:= g C(TT) : g() = g,
the space of all even 2-periodic functions. In particular, F maps
n
onto

n,e
:=

n
C(TT)
e
.
Also,
(Ff, ) (f, )
since the map [1. . 1] [0. . ] : t cos
1
(t) has slope 1 in absolute value everywhere.
Hence
dist (Ff,

n
) const (Ff, 1/n) const (f, 1/n).
Therefore, with m T

n
(Ff),
(66) dist (f,
n
) |f F
1
m| = dist (Ff,

n
) const (f, 1/n),
provided F
1
m is dened and in
n
. For this, observe that the map
C(TT) C(TT)
e
: g g
e
:= (g +g())/2
is norm-reducing since g g() is an isometry. Hence, if g C(TT) and even (as is
the case for g = Ff) and m T

n
(g), then m
e
is an even trig.pol. in

n
and |Ff
m
e
| = |g
e
m
e
| |g m| = dist (g,

n
), hence also m
e
T

n
(g), i.e., we may take
m T

n
(Ff) to be even, hence in F(
n
). In fact,
n
is Haar, hence we have just proved
that the ba from

n
to an even function is even.
Note, though, that we cannot bound (f, ) in terms of (Ff, ), and this is another in-
dication that we cannot expect the same kind of paired direct/inverse theorems concerning
the degree of approximation by algebraic polynomials.
Instead, one has theorems of the following kind.
03mar03 45 c _2003 Carl de Boor
(67) Theorem (Nikolskii, Timan). For some const and for all f C[1 . . 1] and all
n, there exists p
n

n
so that
[f(t) p
n
(t)[ const (f,
n
(t)),
with

n
: t max
1
n
2
,

1 t
2
n
.
If one is not too worried about the constants involved, then Jacksons theorem provides
the right order of the degree of approximation by polynomials to smooth functions, as
follows: Starting with Jacksons theorem,
dist (f,
n
) const (f, 1/n),
and adding to it the fact that, for f C
(1)
[a . . b],
(f, h) h|Df|,
we get, for any p
n
,
dist (f,
n
) = E
n
(f p,
n
) const (1/n) |D(f p)|,
hence
dist (f,
n
) const (1/n) dist (Df,
n1
).
Thus, by induction, for n r,
dist (f,
n
) const
r
(1/n)
r
(D
r
f, 1/n).
Finally, here is a standard result relating a specic K-functional to the modulus of
smoothness of a certain order.
(68) Theorem (Johnen). If T is a closed interval (nite, innite or all of R) or TT, and
r N, then there exist positive constants c and C such that, for all p [1 . . ] and all
f L
p
(T),
(69) c
r
(f, t)
p
K(f, t
r
; L
p
, W
(r)
p
) C
r
(f, t)
p
, t > 0.
Here, to recall,

r
(f, t)
p
:= sup
ht
|
r
h
f|
p
,
with

r
h
f :=
r

k=0
_
r
k
_
(1)
rk
f( +kh).
If T has an endpoint, then
r
h
f may, ohand, only be dened on some proper subset of
T; it is taken to be zero wherever it is not dened.
03mar03 46 c _2003 Carl de Boor
Good approximation, especially by splines
As we have seen, a process of near-best approximation, from some subset M of the
nls X, is necessarily a projector (i.e., idempotent). It is particularly easy to construct and
use if M is a nite-dimensional linear subspace of X in which case the projector can even
be chosen to be linear.
Linear projectors onto nite-dimensional linear subspaces arise naturally when con-
sidering the information that is readily available in the representation of m M with
respect to a particular basis for M, i.e., any particular 1-1 linear map
V : F
n
X : a

j
a(j)v
j
=: [v
1
, . . . , v
n
]a
with ran V = M.
Given such V , the abstract equation V ? = m for the coecients V
1
m of m M
wrto V , is turned into an equivalent numerical equation

V ? =

m by any linear map

: X F
n
that is 1-1 on M. For, with that assumption,

M maps F
n
linearly and 1-1
into itself, hence is an invertible matrix, and this gives the solution
V
1
m = (

V )
1

m.
Now notice that we may assume, in addition, that

V = 1. For, if our initial choice of

does not satisfy this, simply replace it by (

V )
1

. With this additional assumption,


m = V

m for any m M, i.e.,

m is the desired coordinate vector. But any linear map

: X F
n
is necessarily of the form

f =: (
i
f : i = 1, . . . , n) =: [
1
, . . . ,
n
]

f
for certain linear functionals
i
. Hence, we conclude that, for m = V a, the coecient a(j)
gives us the value
j
m of the linear functional
j
at m.
Of course, if M is a proper subspace of X, then there is nothing unique about these

j
. Each choice of

with

V = 1 corresponds to a particular extension of the coordi-


nate functionals for the basis V . In that sense, a(j) = (V
1
m)(j) gives us all kinds of
information about m.
Assuming merely that

is 1-1 on M, it follows that, for any f X,


m = Pf := V (

V )
1

f
is the unique element of M that agrees with f on

in the sense that

m =

f.
In particular, P is a linear projector, onto M, and every linear projector onto M arises in
this way.
03mar03 47 c _2003 Carl de Boor
An example: polynomial interpolation and divided dierences
Consider the map
W
c
: F
N
0
: a

j=1
a(j)w
j
,
with
(70) w
j
: t

i<j
(t c
i
), j = 1, 2, . . . ,
the Newton polynomials for the (arbitrary) sequence c : N F, and
F
T
0
:= a F
T
: #supp(a) < .
Since any a F
N
0
has only nitely many nonzero entries, W
c
a is, indeed, well-dened for
any such a. Also, since deg w
j
= j 1 for all j, W
c
is necessarily 1-1. It is also onto since
its range contains all the w
j
and, for each n, (w
1
, . . . , w
n
) is linearly independent and in
the n-dimensional space
<n
, hence a basis for it.
Thus, each p can be uniquely written as W
c
a, and this particular representation
for p is called its Newton form with (respect to) centers c.
The question now is: what is W
1
c
p for given p ? Related to this is the question:
What information about p = W
c
a is readily provided by the coecient a(n)?
To answer these questions, observe that
p = W
c
a =: p
n
+w
n+1
q
n
,
with
p
n
:=

jn
a(j)w
j
,
and with q
n
some polynomial. It follows that p
n
is the remainder after division of p by
w
n+1
, i.e., p
n

<n
and w
n+1
divides p p
n
. But this uniquely pins down p
n
: Indeed,
if also r
<n
and also w
n+1
divides p r, then w
n+1
also divides p
n
r. But since
deg(p
n
r) < n = deg w
n+1
, this implies that p
n
= r.
Using once more the fact that deg w
j
= j 1 for all j, it follows that a(n) is the
leading coecient, i.e., the coecient of ()
n1
, in the power expansion
p
n
= a(n)()
n1
+ l.o.t. =: P
c
1
,...,c
n
p
of the unique polynomial p
n

<n
that agrees with p at c
1
, . . . , c
n
in the sense that
p p
n
vanishes at that sequence, counting multiplicities. Note that p
n
agrees with p at
c
1
, . . . , c
n
exactly when
(71) D

(p p
n
)(z) = 0, 0 < #1 i n : z = c
i
, z F.
03mar03 48 c _2003 Carl de Boor
We now know that a(n) only depends on p at c
1
, . . . , c
n
, hence a notation like a(n) =
a(p; c
1
, . . . , c
n
) would be appropriate. In fact, here are three notations,
(72) a(n) = p[c
1
, . . . , c
n
] = [c
1
, . . . , c
n
]p = (c
1
, . . . , c
n
)p,
the rst two quite standard. The last, due to V. Kahan, will be adopted here, since
[c
1
, . . . , c
n
] is otherwise occupied.
The view of (c
1
, . . . , c
n
)p as the leading coecient in the power form for the in-
terpolating polynomial P
c
1
,...,c
n
p often is the fastest way to specic results concerning
divided dierences. For example, if c
1
< < c
n
, then, by Rolles Theorem and for
j = 1:n 1, D(p P
c
1
,...,c
n
p) must vanish at some
j
(c
j
. . c
j+1
). It then follows that
DP
c
1
,...,c
n
p = P

1
,...,
n1
Dp and so, in particular,
(n 1)(c
1
, . . . , c
n
)p = (
1
, . . . ,
n1
)Dp.
By induction, this gives that if the c
j
are real, then there is in the smallest interval
containing c
1
, . . . , c
n
so that
(73) (n 1)!(c
1
, . . . , c
n
)p = D
n1
p().
Consider now the computation of specic divided dierences. If c is a constant se-
quence, c = (, , . . .) say, then, by (71) or (73),
(, . . . ,
. .
n terms
)p = D
n1
p()/(n 1)!.
For general c, the answer is a little bit more subtle. However, we already observe the
very important fact that W
c
a is a continuous, in fact an innitely dierentiable, function
of c, hence so is (c
1
, . . . , c
n
)p for each n. Indeed we get, on dierentiating the identity
W
c
W
1
c
= 1 as a function of c and rearranging, that
D(W
1
c
) = W
1
c
D(W
c
)W
1
c
is continuous and smooth.
Further, (c
1
, . . . , c
n
)p is linear in p and symmetric in c
1
, . . . , c
n
.
The ecient way to construct (c
1
, . . . , c
n
)p is obtained as a byproduct of the ecient
evaluation of the polynomial p = W
c
a from its Newton coecients a which, in turn, is
based on writing the Newton form in a nested way, using the fact that each w
j
is a factor
of each w
k
for k > j:
p(z) = a(1) + (z c
1
)(a(2) + (z c
2
)( + (z c
n2
)(a(n 1) + (z c
n1
)a(n)) ))
in case p
<n
. Evaluation of this expression from the inside out results in the following
algorithm.
03mar03 49 c _2003 Carl de Boor
Nested Multiplication aka Horners Method. Input: The sequence a F
n
of (essen-
tial) coecients in the Newton form W
c
a for some p
<n
, the relevant center sequence
c = (c
i
: i = 1, . . . , n 1), and some scalar z.
b(n) := a(n)
for j = n 1, n 2, . . . , 1
b(j) := a(j) + (z c
j
)b(j + 1)
endfor
Output: The number b(1) = p(z).
The algorithm provides the value of p at z, at a cost of only 3n ops. More than
that, the entire sequence b generated by this algorithm is valuable since it provides the
(essential) coecients in the Newton form for p centered at (z, c) = (z, c
1
, c
2
, . . .), i.e.,
(74) p = W
(z,c)
b.
This says, in particular, that
p = b(1) + ( z)

j>1
w
j1
b(j),
illustrating the fact that Horners method can be viewed as a means for dividing p by the
linear polynomial ( z). Now, to prove (74), observe that, directly from the algorithm,
(75) a(j) = b(j) +
_
0 j = n;
(c
j
z)b(j + 1) j < n.
On substituting these expressions for the a(j) into p = W
c
a, we nd
p = w
n
b(n) +w
n1
b(n 1) + (c
n1
z)b(n)) +w
n2
) +
= w
n1
( c
n1
) + (c
n1
z))b(n) +w
n1
b(n 1) +w
n2
) +
= w
n1
( z)b(n) +w
n1
b(n 1) +w
n2
) + ,
with terms 2 and 3 in the last line looking exactly like terms 1 and 2 in the rst line,
except that n is replaced by n 1, hence continuation of the process eventually leads to
(74). In particular, with c
0
:= z, we have b(j) = (c
0
. . . , c
j1
)p, all j, hence (75) can be
rewritten
(c
1
, . . . , c
j
)p = (c
0
. . . , c
j1
)p + (c
j
c
0
)(c
0
, . . . , c
j
)p.
For c
j
,= c
0
, this gives the recurrence relation
(76)
(c
1
, . . . , c
j
)p (c
0
, . . . , c
j1
)p
c
j
c
0
= (c
0
, . . . , c
j
)p
which holds for arbitrary c
0
, . . . , c
j
(as long as c
0
,= c
j
) and is the reason why (c
i
, . . . , c
j
)p
is called the divided dierence of p at c
i
, . . . , c
j
.
03mar03 50 c _2003 Carl de Boor
This recurrence relation is also the simple tool for the calculation, in a divided
dierence table, of the coecients
(77) a(j) = (c
1
, . . . , c
j
)p, j = 1, . . . , n,
needed for the Newton form W
c
a for p
<n
, starting with the numbers
(78) (, . . . ,
. .
terms
)p = D
1
p()/( 1)!, 0 < #1 i n : = c
i
.
The table is triangular and contains eventually all the numbers
t(i, j) := (c
i
, . . . , c
j
)p, 1 i j n.
Assuming for simplicity that
(79) c
i
= c
j
= c
i
= c
i+1
= = c
j
,
each such entry t(i, j) = (c
i
, . . . , c
j
)p either has c
i
= c
j
, and in that case
t(i, j) = D
ji
p(c
i
)/(j i)!,
i.e., one of the numbers we started with, or else c
i
,= c
j
, in which case
t(i, j) =
t(i + 1, j) t(i, j 1)
c
j
c
i
,
hence computable from entries t(r, s) with s r < j i.
Having generated this table in the manner described (and assuming still for simplicity
that (79) holds), we obtain from the table the coecients of the Newton form p = W
c
a of
the polynomial p for which the values
(80) y
j
= D

p(c
j
)/!, = maxj i : c
i
= c
j
, j = 1, . . . , n,
are as entered into the table. But this means that, by entering arbitrary numbers y
j
at
these places in the table, and then completing the table via the recurrence, we obtain the
unique polynomial p in
<n
that satises (80). In other words, we have completely solved
the problem of polynomial interpolation, including osculatory (or Hermite) interpolation.
If the y
j
are computed from some (suciently smooth) function f (polynomial or not),
then the table entries are denoted by
t(i, j) =: (c
i
, . . . , c
j
)f
and called divided dierences of the function f. Thus the
03mar03 51 c _2003 Carl de Boor
(81) Denition. The divided dierence of f at c
i
, . . . , c
j
, denoted by
(c
i
, . . . , c
j
)f,
is dened for any suciently smooth function f as the leading coecient of the unique
polynomial of degree j i that agrees with f at the sequence (c
i
, . . . , c
j
).
(82) Theorem. For any suciently smooth function f, and any sequence (c
1
, . . . , c
n
),
p :=
n

j=1
(c
1
, . . . , c
j
)f
j1

i=1
( c
i
)
is the unique polynomial of degree < n that agrees with f at the sequence (c
1
, . . . , c
n
).
Of the many wonderful properties that divided dierences possess, I mention here only
one, as this one served as the portal through which we rst glanced a theory of multivariate
splines, and since, strangely, there is only one Numerical Analysis text (Isaacson-Keller)
that actually provides it. Also, it makes many other properties of the divided dierences
quite evident.
(83) Genocchi-Hermite formula. For every c = (c
0
, . . . , c
k
) and every p ,
(c
0
, . . . , c
k
)p =
_
(c
0
,...,c
k
)
D
k
p,
which uses the Genocchi functional
f
_
(c
0
,...,c
k
)
f :=
_

k
f(c
0
+
k

i=1
s
i
c
i
) ds
in which integration is done over the standard k-simplex,
k
:= s R
k
: 1 s
1

s
k
0.
Proof: The proof, like many that involve divided dierences, is done by induction
on k. To be sure, since
_
(c
0
,...,c
k
)
1 = vol(
k
) = 1/k!, hence
_

k
f(c
0
+
k

i=1
s
i
c
i
) ds = f()/k!
for some convc
0
, . . . , c
k
, the Genocchi formula is immediate for the special case
c
0
= = c
k
, and so, in particular, for k = 0. Assume that c
k1
,= c
k
. Then, for any t,
_
s
k1
0
D
k
g(t +s
k
(c
k
c
k1
)) ds
k
=
D
k1
g(t +s
k1
(c
k
c
k1
)) D
k1
g(t)
c
k
c
k1
.
Hence, with
t := c
0
+s
1
c
1
+ +s
k1
c
k1
,
03mar03 52 c _2003 Carl de Boor
we compute
_

k
D
k
g(c
0
+
k

i=1
s
i
c
i
) ds =
_
1
0

_
s
k1
0
D
k
g(t +s
k
c
k
) ds
k
ds
1
=
_
1
0

_
s
k2
0
D
k1
g(t +s
k1
c
k
) D
k1
g(t)
c
k
c
k1
ds
k1
ds
1
=
(c
0
, . . . , c
k2
, c
k
) (c
0
, . . . , c
k1
)
c
k
c
k1
g = (c
0
, . . . , c
k
)g,
the second last equality by induction hypothesis and since
t +s
k1
c
k
= c
0
+ +s
k2
c
k2
+s
k1
(c
k
c
k2
).
To be sure, any conclusion derived from the Genocchi formula (including the for-
mula itself) can be extended to any function g for which it makes sense and that can be
approximated suitably by polynomials, e.g., to all g C
(k)
(T) for a suitable T.
03mar03 53 c _2003 Carl de Boor
Univariate B-splines
Here is a very swift introduction to B-splines, and thereby to splines aka smooth
piecewise polynomials. (A picturesque version of this material is available, by anonymous
ftp, from the site /ftp.cs.wisc.edu in the subdirectory Approx as the postscript le
bsplbasic.ps.)
Let t := t
i1
t
i
t
i+1
be a nondecreasing sequence. This sequence may
be nite, innite, or even bi-innite. Two important (and extreme) special cases are
(i) t = Z, leading to cardinal splines;
(i) t = IB := ( , 0, 0, 0, 1, 1, 1, . . .), leading to the Bernstein-Bezier form (or, BB-form)
for polynomials (restricted to [0 . . 1]).
The associated (normalized) B-splines of order k for the knot sequence t are, by
denition, the functions
(84) B
j
:= B
j,k
:= B
j,k,t
: x (t
j
, . . . , t
j+k
)( x)
k1
+
(t
j+k
t
j
),
with (T) the divided dierence functional introduced in the preceding section. The
normalization ensures that the B
j
form a (nonnegative) partition of unity (see (98)
below). The B-spline B
j
was originally denoted by N
j
= N
j,k,t
in order to distinguish it
from the dierently normalized B-spline
M
j
= M
j,k,t
:=
k
t
j+k
t
j
B
j,k,t
,
for which
_
M
j
= 1.
The latter normalization arises naturally when applying the divided dierence to both
sides of the Taylor identity:
f =

r<k
D
r
f(a)( a)
r
/r! +
_
b
a
k( s)
k1
+
D
k
f(s) ds/k!
to obtain (under the assumption that t
j
, . . . , t
j+k
[a . . b])
(85) (t
j
, . . . , t
j+k
)f =
_
R
M
j,k,t
D
k
f/k!,
showing that M
j,k,t
is the Peano kernel for the divided dierence (t
j
, . . . , t
j+k
).
By using Leibniz Rule
(
0
, . . . ,
k
)(fg) =
k

i=0
(
0
, . . . ,
i
)f (
i
, . . . ,
k
)g
05sep00 54 c _2003 Carl de Boor
for the divided dierence of a product, applied to the particular product ( x)
k1
+
=
( x)( x)
k2
+
, one readily obtains the recurrence relations
(86a) B
i,k
=
i,k
B
i,k1
+ (1
i+1,k
)B
i+1,k1
with
(86b) B
i,1
:=
[t
i
..t
i+1
)
,
i,k
(x) :=
x t
i
t
i+k1
t
i
.
However, once one knows these recurrence relations, it is most ecient for the devel-
opment of the B-spline theory to take (86) as the starting point, i.e., to dene B-splines
by (86). Equivalently, the development about to be given will make no use of divided
dierences, but will rely entirely on (86).
It follows at once that B
i,k
can be written in the form
(87) B
i,k
=
i+k1

j=i
b
j,k

[t
j
..t
j+1
)
,
with each b
j,k
a polynomial of degree < k since it is the sum of products of k 1 linear
polynomials. Therefore,
B
i,k

<k,(t
i
,...,t
i+k
)
on [t
i
. . t
i+k
],
while
(88) B
i,k
= 0 o [t
i
. . t
i+k
].
Further, B
i,k
depends only on the knots t
i
, . . . , t
i+k
. For this reason, the alternative
notation
(89) B
i,k
=: B([t
i
, . . . , t
i+k
)
is also customary. The actual smoothness of B
i,k
depends on the multiplicity with which
each of the knots t
j
, i j i + k, appears in its knot sequence (t
i
, . . . , t
i+k
), as we will
see in a moment.
Since both
i,k
and 1
i+1,k
are positive on (t
i
. . t
i+k
), it follows from (88) by
induction on k that B
i,k
is positive on (t
i
. . t
i+k
).
Since at most k of the B
i,k
are nonzero at any one point x R, the denition

a
i
B
i,k
: R R : x

a
i
B
i,k
(x)
of

i
a
i
B
i,k
as a pointwise sum makes sense for arbitrary a even when the sum has innitely
many terms. We call any such function a spline of order k with knot sequence t and
denote the collection of all such functions by
S
k,t
.
05sep00 55 c _2003 Carl de Boor
We deduce from the recurrence relation that
(90)

a
i
B
i,k
=

_
a
i

i,k
+a
i1
(1
i,k
)
_
B
i,k1
,
showing that the coecients on the right are ane combinations of neighboring coecients
on the left.
Before proceeding with this, note the following technical diculty. If t has a rst
knot, e.g., t = (t
1
, . . .), then the sum on the left in (90) starts with i = 1, while the sum
on the right can only start at i = 2 since the coecient a
i1
appears in it. To avoid such
diculty, we agree in this case (and the corresponding case when t has a last entry) to
extend t in any way whatsoever to a bi-innite knot sequence, denoting the extension again
by t. However, this increases the number of available B-splines, hence also increases the
spline space. Since we are still interested only in our original spline space, we further agree
to make use of results obtained from the extended situation only to the extent that they
dont explicitly involve any of the additional knots. We can be sure none of the additional
knots matters if we restrict attention to the largest interval not intersected by the interior
of the support of any of the additional B-splines. We call this the basic interval for S
k,t
and denote it by
I
k,t
:= (t

. . t
+
)
with
t

:=
_
t
k
, if t = (t
1
, . . .);
inf
i
t
i
, otherwise,
t
+
:=
_
t
n+1
, if t = ( , t
n+k
);
sup
i
t
i
, otherwise,
In practice, the knot sequence is nite, having both a rst and a last knot. In that case,
one chooses I
k,t
to be closed, and this is ne for the left endpoint, since the denition of
B-splines makes them all continuous from the right. For this to work properly at the right
endpoint, one modies the above denition of B-splines to make them continuous from the
left at the right endpoint of I
k,t
.
In summary, even if the given knot sequence is not biinnite, we may always assume
it to be biinnite as long as we apply the results so obtained only to functions on the basic
interval I
k,t
determined by the given knot sequence.
With this, consider the special sequence
a
i
:=
i,k
() := (t
i+1
) (t
i+k1
)
(with IR). We nd for B
i,k1
,= 0, i.e., for t
i
< t
i+k1
, that
a
i

i,k
+a
i1
(1
i,k
) =
i,k1
()
_
(t
i+k1
)
i,k
+ (t
i
)(1
i,k
)
_
=
i,k1
()( )
since f(t
i+k1
)
i.k
+ f(t
i
)(1
i,k
) is the straight line that agrees with f at t
i+k1
and
t
i
. This shows that
(91)

i,k
()B
i,k
= ( )

i,k1
()B
i,k1
,
hence, by induction, that
(92)

i,k
()B
i,k
= ( )
k1

i,1
()B
i,1
.
This proves the following identity.
05sep00 56 c _2003 Carl de Boor
(93) Marsdens Identity. For any IR,
(94) ( )
k1
=

i,k
()B
i,k
on I
k,t
, with
i,k
() := (t
i+1
) (t
i+k1
).
Since in (94) is arbitrary, it follows that S
k,t
contains all polynomials of degree
< k (restricted to I
k,t
). More than that, we can even give an explicit expression for the
required coecients, as follows.
By dierentiating (94) with respect to , we obtain the identities
(95)
( )
k
(k )!
=

i
(D)
1

i,k
()
(k 1)!
B
i,k
, > 0.
On using these identities in the Taylor formula
p =
k

=1
( )
k
(k )!
D
k
p()
for a polynomial p of degree < k, we conclude that any such polynomial can be written in
the form
(96) p =

i,k
p B
i,k
,
with
i,k
given by the rule
(97)
i,k
f :=
i,k,t
f :=
k

=1
(D)
1

i,k
()
(k 1)!
D
k
f().
Here are two special cases of particular interest. For p = 1, we get
(98) 1 =

i
B
i,k
since D
k1

i,k
= (1)
k1
(k1)!, and this shows that the B
i,k
form a partition of unity.
Further, anticipating that
jk
p is independent of in case p
<k
, since D
k2

i,k
is a
linear polynomial that vanishes at
t

i,k
:= (t
i+1
+ +t
i+k1
)/(k 1),
(99) =

i
(t

i,k
)B
i,k
,
1
.
24mar03 57 c _2003 Carl de Boor
The identity (95) also gives us various piecewise polynomials contained in S
k,t
: Since
t
i
< t
j
< t
i+k
implies that D
1

i,k
(t
j
) = 0 in case #t
j
:= #s : t
s
= t
j
, the choice
= t
j
in (95) leaves only terms with support either entirely to the left or else entirely to
the right of t
j
. This implies that
(100)
( t
j
)
k
+
(k )!
=

ij
(D)
1

i,k
(t
j
)
(k 1)!
B
i,k
, 0 < #t
j
.
Consequently,
(101) ( t
j
)
k
+
S
k,t
for 1 #t
j
,
(on I
k,t
).
(102) Theorem. If t
i
< t
i+k
for all i, then the B-spline sequence (B
i,k
: i) is linearly
independent and the space S
k,t
coincides with the space S :=

<k,t
of all piecewise poly-
nomials of degree < k with breakpoints t
i
that are (i) := k 1 #t
i
times continuously
dierentiable at t
i
, all i. In particular, each f S
k,t
is in C
((i))
near t
i
, all i.
Proof: It is sucient to prove that, for any nite interval I := [a . . b], the re-
striction S
|I
of the space S to the interval I coincides with the restriction of S
k,t
to that
interval. The latter space is spanned by all the B-splines having some support in I, i.e.,
all B
i,k
with (t
i
. . t
i+k
) I ,= . The space S
|I
has a basis consisting of the functions
(103) ( a)
k
, = 1, , k; ( t
i
)
k
+
, = 1, , #t
i
, for a < t
i
< b.
This follows from the observations that (i) the sequence of functions in (103) is linearly
independent; and (ii) a piecewise polynomial function f with a breakpoint at t
i
that is
k 1 #t
i
times continuously dierentiable there can be written uniquely as
f = p +
#t
i

=1
a

( t
i
)
k
+
,
with p a suitable polynomial of degree < k and suitable coecients a

. Since each of the


functions in (103) lies in S
k,t
, by (95) and (101), we conclude that
(104) S
|I
(S
k,t
)
|I
.
On the other hand, the dimension of S
|I
, i.e., the number of functions in (103), equals the
number of B-splines with some support in I (since it equals k + #i : a < t
i
< b), hence
is an upper bound on the dimension of (S
k,t
)
|I
. This implies that equality must hold in
(104), and that the set of B-splines having some support in I must be linearly independent
over I.
24mar03 58 c _2003 Carl de Boor
(105) Corollary. The sequence (B
j,k
: B
j,k
I
,= 0) of B-splines having some support in
a given (proper) interval I is linearly independent over that interval.
(106) Corollary. For all p
<k
, D

j,k
p = 0.
Indeed, since the B-spline sequence is linearly independent, the coecients
i,k
p ap-
pearing in (96) are uniquely determined, hence cannot change with . When convenient
later on, we will choose in (97) in dependence on i, i.e., as
i
.
(107) Corollary. If

t is a renement of the knot sequence t, then S
k,t
S
k,

t
.
(108) Corollary. The derivative of a spline in S
k,t
is a spline of degree < k 1 with
respect to the same knot sequence, i.e., DS
k,t
S
k1,t
.
(109) Remark The word derivative is used here in the pp sense: The derivative
of a pp f is, by denition, the pp with the same breakpoint sequence whose polynomial
pieces are the rst derivative of the corresponding polynomial pieces of f. This makes it
possible in (108) to ignore the possibility that t
i
= t
i+k1
for some i, hence the elements
of S
k,t
will, in general, fail to be dierentiable at such a t
i
.
More generally, here and elsewhere, we do not exclude the possibility that some of the
B
i,k
are trivial, i.e., that t
i
= t
i+k
for some i. However, if we were to interpret B
i,k,t
as
distributions, we would have to proceed with more caution. For, by (85) and (73),
(110) lim
t
i
,...,t
i+k

M([t
i
, . . . , t
i+k
) =

= ()
as a distribution, and that limit is quite dierent from 0.
The identity (96) can be extended to all spline functions. For this, we agree, consistent
with (86b), that all derivatives in (97) are to be taken as limits from the right in case
coincides with a knot.
(111) Theorem. If =
i
in denition (97) of
i,k
is chosen in the interval [t
i
. . t
i+k
),
then, with the understanding that D

f() := D

f(+) for any and any pp f,


(112)
i,k
_

j
a
j
B
j,k
_
= a
i
.
It is remarkable that can be chosen arbitrarily in the interval [t
i
. . t
i+k
). The reason
behind this is Corollary (106).
Proof: Assume that [t
i
. . t
i+k
), hence [t
l
. . t
l+1
) [t
i
. . t
i+k
) for some l,
and let p
j
be the polynomial that agrees with B
j,k
on (t
l
. . t
l+1
). Then

i,k
B
j,k
=
i,k
p
j
.
On the other hand,
p
j
=
l

m=l+1k

m,k
p
j
p
m
,
since this holds by (96) on [t
l
. . t
l+1
), while, by Corollary (105) or directly from (96),
(p
l+1k
, , p
l
) is linearly independent. Therefore necessarily
i,k
B
j,k
=
i,k
p
j
equals 1 if
i = j and 0 otherwise.
05sep00 59 c _2003 Carl de Boor
use of the dual functionals: dierentiation; dependence on knots
Because of (112), the functionals
i,k
are called the dual functionals for the B-
splines. Strictly speaking, there are many such functionals (of which more anon), but
these particular ones have turned out to be quite useful in various contexts. Here are
several examples.
Compare
(113)
i,k
f =
k

=1
(D)
1

i,k
()
(k 1)!
D
k
f()
with

i,k1
Df =
k1

=1
(D)
1

i,k1
()
(k 2)!
D
k1
Df().
Since (t
i
)
i,k1
=
i1,k
and (t
i+k1
)
i,k1
=
i,k
, subtraction of the latter from
the former gives
(t
i+k1
t
i
)
i,k1
=
i,k

i1,k
.
Hence
(114)
i,k1
D =

i,k

i1,k
(t
i+k1
t
i
)/(k 1)
.
Consequently, we get the dierentiation formula
(115) D

i
a
i
B
i,k
=

i
a
i
a
i1
(t
i+k1
t
i
)/(k 1)
B
i,k1
.
We now consider how
i,k
depends on the knot sequence t. Perhaps surprisingly,
although B
i,k
involves the knots t
i
, , t
i+k
,
i,k
only depends on the interior knots,
t
i+1
, . . . , t
i+k1
. Further, since
i,k
depends linearly on
i,k
, it depends anely and sym-
metrically on the points t
i+1
, , t
i+k1
. Indeed, for any , x, y,
((x + (1 )y) ) = (x ) + (1 )(y ).
Hence, with

k
: R
k1
(C
(k1)
)

: t
0,k
,
we have

k
(x + (1 )y, s
2
, . . . , s
k1
) =
k
(x, s
2
, . . . , s
k1
) + (1 )
k
(y, s
2
, . . . , s
k1
)
as well as

k
(s) =
k
(s )
for any permutation of order k 1.
05sep00 60 c _2003 Carl de Boor
Consider, in particular, t
1
= = t
k1
= x. Then
0,k
= (x )
k1
, hence
(D)
1

0,k
()/(k 1)! = (x )
k
/(k )!
and therefore

k
(x, . . . , x) =
k

=1
(x )
k
(k )!
[]D
k
.
In particular

k
(x, . . . , x)p = p(x), p
<k
.
This meshes entirely with the fact that, if x = t
i+1
= = t
i+k1
, then B
i,k
is the only
B
j,k
that is nonzero at x, hence B
i,k
(x) must be 1 (since the B
j,k
form a partition of
unity), and therefore

j
a
j
B
j,k
(x) = a
i
in this case.
side issue: blossoms
It is worthwhile to point out that, associated with each p
r
, there is a unique
symmetric r-ane form called its polar form (in Algebra) or its blossom (in CAGD),
denoted therefore here by

p,
for which
x R p(x) =

p (x, . . . , x).
E.g., the blossom of ( )
r

r
is s (s
1
) (s
r
). If p =

j
()
j
c
j

r
, then

p (s
1
, . . . , s
r
) :=

j
c
j

iI
s
i
: I
_
1, . . . , r
j
_
/
_
r
j
_
,
with
_
M
j
_
:= K M : #K = j. We deduce from the above that

p (t
1
, . . . , t
k1
) =
k
(t
1
, . . . , t
k1
) p, p
<k
.
In particular, the ith B-spline coecient of a kth order spline with knot sequence t is
the value at (t
i+1
, . . . , t
i+k1
) of the blossom of any of the k polynomial pieces associated
with the intervals [t
j
. . t
j+1
), j = i, . . . , i +k 1. This observation was made, in language
incomprehensible to the uninitiated, by de Casteljau in the sixties. It was discovered
independently and made plain (and given the nice name of blossom) by Lyle Ramshaw
in the early eighties.
05sep00 61 c _2003 Carl de Boor
knot insertion
This has led to the evaluation of a spline by knot sequence renement. Such a rene-
ment can always be reached adding one knot at a time. So, suppose that the knot sequence

t has been obtained from the knot sequence t by the insertion of just one additional term,
the point x say. Thus, for some j,

t
i
=
_
t
i
, for i j;
x, for i = j + 1;
t
i1
, for i > j + 1.
We saw already that, with
x = t
i+k1
+ (1 )t
i
,
i.e., with
= (x) :=
i,k
(x),
we have

k
(x, t
i+1
, . . . , t
i+k2
) =
k
(t
i+1
, . . . , t
i+k1
) + (1 )
k
(t
i
, . . . , t
i+k2
).
While this is true for arbitrary i, it matters here only when i < j +1 < i +k. Altogether,
we nd that

i,k
= (1
i,k
(x))
i1,k
+
i,k
(x)
i,k
, all i,
with
(116)
i,k
(x) := max0, min1,
i,k
(x).
Correspondingly,
(117)

i
a
i
B
i,k
=

i
a
i

B
i,k
,
with
a
i
= (1
i,k
(x))a
i1
+
i,k
(x)a
i
, all i.
Note that this is exactly the way we computed coecients by recurrence in (90). This
intimate connection between the recurrence relation and knot insertion was rst observed
by Wolfgang Boehm. Note further that k 1-fold insertion of the knot x produces even-
tually a knot sequence

t, in which all the interior knots for

B
j,k
are equal to x, hence,
correspondingly,
a
j
= (

i
a
i
B
i,k
)(x).
In CAGD, one thinks of a spline f =

j
a
j
B
j,k
in terms of the curve x (x, f(x))
that is its graph. Since x =

j
t

j,k
B
j,k
(x) by (99), we can represent this spline curve as
the vector-valued spline

j
P
j
B
j,k
with coecients
P
j
:= (t

j,k
, a
j
),
called its control points. The broken line with vertex sequence (P
j
) is called its control
polygon; Ill denote it by
C
k,t
f,
to stress its dependence on t.
05sep00 62 c _2003 Carl de Boor
(118) Proposition. If

t is obtained from t by the insertion of one additional knot, then,
for any f S
k,t
, C
k,

t
f interpolates, at its breakpoints, to C
k,t
f, specically,

P
j
= (1
j,k
(x))P
j1
+
j,k
(x)P
j
,
(and is thereby uniquely determined).
Here are two gures, to illustrate this geometric interpretation of knot insertion which,
ultimately led to an entirely new and quite dierent way to generate curves and surfaces,
namely subdivision.
t
j+1
x
t
j2
t
j1
x t
j+2
t
j
x
t
j+3
P
j1

P
j
(119) Figure. Insertion of x = 2 into the knot sequence
t = (0,0,0,0,1,3,5,5,5,5), with k = 4.
(120) Figure. Three-fold insertion of the same point.
26mar03 63 c _2003 Carl de Boor
(121) Figure. A cubic spline, its control polygon, and various straight lines
intersecting them. The control polygon exaggerates the shape
of the spline. The spline crossings are bracketed by the control
polygon crossings.
It follows that the map a f
a
:=

i
a
i
B
i,k
is variation-diminishing, meaning that
f
a
has no more variations in sign than does the sequence a, in a sense to be made precise
next.
variation diminution
For a real sequence a without any zero entries,
S(a) := #i : a(i)a(i + 1) < 0
denotes the number of sign changes in it. It is less clear what this number should be in
case a has some zero entries. The maximum number of sign changes obtainable in such
a sequence by an appropriate choice in the sign of any zero entry is called the number of
weak sign changes in it, denoted
S
+
(a).
The minimum number so obtainable is denoted by
S

(a)
and called the number of strong sign changes (an example of the Bauhaus maxim less
is more?). It equals the number of sign changes when we ignore the zeros. The number
of strong (weak) sign changes can only increase (decrease) under small perturbations.
Precisely,
S

(a) liminf
ba
S

(b) limsup
ba
S
+
(b) S
+
(a).
The following Lemma is immediate (from the geometric picture of knot insertion).
05sep00 64 c _2003 Carl de Boor
(122) Lemma (Lane, Riesenfeld). If a is the B-spline coecient sequence obtained
from the sequence a by the insertion of (zero or more) knots, then
S

( a) S

(a).
If f is a function on an interval (including the interval R), one denes
S

(f) := sup
x
1
<<x
r
S

(f(x
1
), . . . , f(x
r
)).
(123) Proposition. S

j
a
j
B
j
) S

(a).
Proof: Insert into t each of the entries in a given increasing sequence (x
i
) enough
times to have them appear in the resulting knot sequence

t at least k 1 times. Then
(f(x
1
), . . . , f(x
r
)) is a subsequence of the resulting B-spline coecient sequence a, hence
S

(f(x
1
), . . . , f(x
r
)) S

( a) S

(a).
(124) Denition. Schoenbergs (spline) operator is, by denition, the linear map
V = V
k,t
given by the rule
V g :=

j
g(t

j,k
)B
j,k
.
It is usually dened only when #t
i
< k, all i.
(125) Proposition. Schoenbergs spline operator is variation-diminishing. Precisely,
for any g and any
1
,
(126) S

(V g ) S

(g ).
Even more precisely,
(127) D
r
g 0 = D
r
V g 0, r = 0, 1, 2,
and this holds even locally.
Proof: By (99), V = (on I
k,t
) for all
1
, hence V (f ) = V f ,
therefore, by Proposition (123), and by the strict increase in the sequence t

i
: i = 1, . . . , n
S

(V f ) S

((f )(t

i
) : i) S

(f ).
However, (126) by itself fails to imply the more precise statement (127), which follows
from the nonnegativity of the B-splines along with the observation that
DV g =

j
(t

j1
, t

j
)g B
j,k1
,
hence
D
2
V g =

j
(t

j1
, t

j
)g (t

j2
, t

j1
)g
(t
j+k2
t
j
)/(k 2)
B
j,k2
.
05sep00 65 c _2003 Carl de Boor
zeros of a spline, counting multiplicities
There is a complete theory that provides an upper bound on the number of zeros of a
spline, even counting multiplicity, in terms of the number of sign changes (strong and/or
weak) in its B-spline coecients, with multiplicity of a zero dened as the maximal
number of distinct nearby zeros in a nearby spline (from the same spline space).
For the theory to give useful information, one has to assign a nite multiplicity to
zero intervals, and this adds a further complication.
Multiplicity considerations are important when one wishes to consider osculatory
spline interpolation, i.e., interpolation at possibly repeated points. Since I will not get
to that topic in this course, I am content to state and prove only the following very useful
proposition.
(128) Proposition. If f =

j
a
j
B
j,k,t
vanishes at x
1
< < x
r
, while f
t
:=

j
[a
j
[B
j,k,t
does not, then S

(a) r.
Proof: Since f
t
(x
i
) > 0, while f(x
i
) = 0, the sequence (a
j
B
j
(x
i
) : B
j
(x
i
) ,= 0)
must have at least one strong sign change, hence, so must the sequence (a
j
: B
j
(x
i
) ,= 0),
by the nonnegativity of the B-splines. This gives altogether r strong sign changes in
a, provided we can be sure that dierent x
i
generate dierent sign changes. O-hand,
this may not be so, but can be guaranteed by inserting each of the points (x
i
+ x
i+1
)/2,
i = 1, . . . , r 1, into the knot sequence k times. If

t and a are the resulting knot and
coecient sequences, respectively, then still f

t
> 0 = f on the x
i
(since, from f
t
(x
i
) > 0,
we know that x
i
is an isolated zero of f), while now j :

B
j
(x
i
) ,= 0j :

B
j
(x
h
) ,= 0 =
for i ,= h, hence
S

(a) S

( a) r.
spline interpolation
We consider spline collocation, i.e., interpolation from S
k,t
at the increasing se-
quence x = (x
i
) of points. We consider this under the assumption that t = (t
i
: i =
1, . . . , n +k), with #t
i
< k, all i, hence S
k,t
C(R). This simplifying assumption avoids
discussion otherwise needed in case some x
j
agrees with a knot of multiplicity k, in
which case one would have to specify, in addition, whether it is x
j
or x
j
+ one wants.
With the assumption, none of the B-splines B
i,k
is trivial, hence dimS
k,t
= n. For this
reason, we assume, more precisely, that
x = (x
1
< < x
n
).
Hence, for given g, we are seeking f S
k,t
with f = g on x. Equivalently, we are seeking
a solution to the linear system A? = g
x
, with
A := (B
j,k
(x
i
) : i, j = 1, . . . , n)
the so-called collocation matrix. There is exactly one interpolant to a given g i A is
invertible i A is 1-1.
05sep00 66 c _2003 Carl de Boor
(129) Proposition. If A = (B
j,k
(x
i
)) is invertible, then
t
i
< x
i
< t
i+k
, i.
Proof: If, for some i, t
i+k
x
i
, then the rst i columns of A have nonzero entries
only in the rst i 1 rows, hence A cannot be invertible. Again, if x
i
t
i
, then columns
i, . . . , n of A have nonzero entries only in rows i + 1, . . . , n, hence A cannot be invertible.
Note that the argument used nothing more than the fact that both the sequence x
and the sequence of the supports of the B-splines are increasing. In particular, we have
proved
(130) Corollary. If (B
m
j
,k
(s
i
) : i, j = 1, . . . , r) is invertible, with both (m
j
) and (s
i
)
increasing, then B
m
i
(s
i
) ,= 0, all i.
In other words, such a matrix is invertible only if its diagonal entries are nonzero. As
it turns out, the converse also holds. The converse of the Proposition is
(131) Schoenberg-Whitney Theorem. Let t = (t
1
, . . . , t
n+k
) with B
j,k
,= 0, all j, and
let x := (x
1
< < x
n
). Then, A
x
:= (B
j,k
(x
i
)) is invertible i t
i
< x
i
< t
i+k
, all i.
Proof: We only need to prove the if. Since A
x
is square, it is sucient to prove
that A
x
a = 0 implies a = 0. Consider f =

j
a
j
B
j,k
with A
x
a = 0. If a ,= 0, then, by
Corollary (105), f ,= 0. Let I = ( . . ) be a maximal open interval in
supp

j
[a
j
[B
j,k
=
_
(t
j
. . t
j+k
) : a
j
,= 0.
It follows that I = (t

. . t
+k
) for some 1 n, and that
f = f
I
:=

j=
a
j
B
j,k
on I.
In particular, f
I
has the distinct zeros x

, . . . , x

, therefore, by Proposition (128),


S

(a

, . . . , a

) + 1 ,
which is nonsense.
The Schoenberg-Whitney Theorem has been generalized in at least two directions: (i)
permission of coincidences in the sequence (x
i
) correspondingly to osculatory interpolation;
and (ii) consideration of a subsequence (B
m
j
: j = 1, . . . , n) instead of the whole sequence
(for a suitably longer knot sequence).
Any x = (x
1
< < x
n
) satisfying the Schoenberg-Whitney conditions for S
k,t
gives
rise to the corresponding projector P
x
that associates g C with the unique f = P
x
g S
k,t
that agrees with g at x. Since P
x
is a linear projector, we have
dist (g, S
k,t
) |g P
x
g| (1 +|P
x
|) dist (g, S
k,t
),
05sep00 67 c _2003 Carl de Boor
hence P
x
s a candidate for a good approximation scheme from S
k,t
to the extent that
|P
x
| is small, i.e., not much larger than 1. We pursue this question in the context of
the uniform norm, i.e., in the space
X = C[t
1
. . t
n+k
].
Since P
x
g =

j
a
j
B
j
with a = A
1
x
(g
x
), while (B
j
) is a positive partition of unity,
|P
x
g|

|A
1
x
(g
x
)|

|A
1
x
|

|(g
x
)|

|A
1
x
|

|g|

,
with |A
1
x
|

= max
i

[A
1
x
(i, j)[. Hence
|P
x
| |A
1
x
|

.
While this bound is not sharp (in general), it is the only bound readily available. Hence,
in search for a good approximation scheme from S
k,t
, we now look for x so that |A
1
x
|

is as small as possible, and this will lead us to a particularly good choice for x, namely
the Chebyshev-Demko sites x

, easily computable for given t, and, for these, |P


x
|
|A
1
x
|

k2
k
. This bound is quite small for modest k and, surprisingly, is independent
of the knot sequence t. In other words, interpolation at the Chebyshev-Demko sites is near-
best independent of the knot sequence. This makes it possible to use such interpolation
protably even when one has chosen the knot sequence quite non-uniform in order to adjust
to the varied behavior of the function being approximated.
The search for x that minimizes |A
1
x
|

is aided by the knowledge that our collocation


matrix A
x
is totally positive, to be established next.
total positivity
We recall that an (m, n)-matrix C is totally positive if, for any strictly increasing
(index) sequences i = (i
1
< < i
r
) in 1, . . . , m and j = (j
1
< < j
r
) in 1, . . . , n,
the determinant det C(i, j) of the (r, r)-submatrix
C(i, j) :=
_
C(i
p
, j
q
) : p, q = 1, . . . , r
_
is nonnegative. The most immediately important fact concerning total positivity is the
following.
(132) Fact. If C is invertible and totally positive, then its inverse is checkerboard,
meaning that C
1
(i, j)(1)
ij
0, all i, j.
Proof: By Cramers rule,
C
1
(i, j) = (1)
ij
det C(j, i)/ det C.
31mar03 68 c _2003 Carl de Boor
(133) Theorem (Karlin). For any x := (x
1
< < x
n
), any k N, and any knot
sequence t = (t
i
: i = 1, . . . , n +k), the collocation matrix
A :=
_
B
j
(x
i
) : i, j = 1, . . . , n
_
is totally positive.
Proof: If

t is obtained from t by the insertion of just one knot, and

B
j
:= B
j,k,

t
,
all j, then, by (117),
B
j
= (1
j+1
)

B
j+1
+
j

B
j
,
with all
j
[0 . . 1]. Since the determinant of a matrix is a linear function of the columns
of that matrix, we have, e.g.,
det[ , B
j
(x), . . .] = (1
j+1
) det[ ,

B
j+1
(x), . . .] +
j
det[ ,

B
j
(x), . . .],
with unchanged in their respective places. It follows that, for any i, j,
det A(i, j) =

h
det

A(i, h),
with all the
h
0, and the sum, ohand, over certain nondecreasing sequences, since
only neighboring columns of

A participate in a column of A. However, we may omit all
h that are not strictly increasing, since the corresponding determinant is trivially zero.
Therefore,
det A(i, j) =

h
det

A(i, h),
with the
h
0 and all h strictly increasing.
Now insert each of the x
i
enough times so that the resulting rened knot sequence

t
contains each x
i
exactly k 1 times. By induction, we have
det A(i, j) =

h
det

A(i, h),
with the
h
0 and all h strictly increasing. However, in each row of

A, there is exactly
one nonzero entry, namely the entry belonging to

B
m
i
with

t
m
i
< x
i
=

t
m
i
+1
, and that
entry equals 1. In other words,

A has all its entries zero except that the submatrix

A(:, m)
is the identity matrix. Thus det

A(i, h) =
h,mi
, hence det A(i, j) =
mi
0.
31mar03 69 c _2003 Carl de Boor
spline interpolation (cont.)
We continue with the spline interpolation setup introduced earlier, considering inter-
polation at x = (x
1
< < x
n
) from S
k,t
with t = (t
1
t
n+k
) and #t
i
< k, all
i.
The fact that the collocation matrix
A
x
:= (B
j
(x
i
))
is totally positive implies that
|A
1
x
|

= max
i

j
[A
1
x
(i, j)[ = max
i

j
A
1
x
(i, j)(1)
ij
= max
i
(1)
in
a
x
i
,
with a
x
the unique solution to the equation A
x
? = ((1)
nj
: j = 1, . . . , n), hence
|P
x
|

|A
1
x
|

= |a
x
|

,
with
f
x
:=

j
a
x
j
B
j
the unique element of S
k,t
satisfying
f
x
(x
i
) = (1)
ni
, i = 1, . . . , n.
Our search for argmin|A
1
x
|

therefore is the search for the x that minimizes |a


x
|

. For
this, we use, in eect, Remez (second) algorithm for the construction of a ba from
k
.
First we note that
(1)
ni
a
x
i
> 0, all i,
since a
x
i
=

j
A
1
x
(i, j)(1)
nj
= (1)
ni

j
[A
1
x
(i, j)[ with the sum necessarily posi-
tive. Further, since f
x
strictly alternates in sign at the n points x
1
, . . . , x
n
and vanishes
outside (t
1
. . t
n+k
), it follows that f
x
has n distinct local extrema y
1
< < y
n
with
(1)
ni
f
x
(y
i
) 1, i.
This implies that t
i
< y
i
< t
i+k
. all i. Indeed, if, e.g., y
i
t
i
for some i, then
B

(y

) = 0 for all i , showing that, on y


1
, . . . , y
i
, f
x
agrees with

j<i
a
x
j
B
j
. In
particular, i 2 S

(a
x
1
, . . . , a
x
i1
) S

j<i
a
x
j
B
j
) = i 1, a contradiction.
Consequently, there is exactly one f
y
=:

j
a
y
j
B
j
with f
y
(y
i
) = (1)
ni
, all i. For
any < 1, the dierence, f
x
f
y
, strictly alternates in sign on y
1
< < y
n
, hence we
must have S

(a
x
a
y
) = n 1, and therefore (1)
ni
a
x
i
(1)
ni
a
y
i
0, all i.
In this way, we obtain a sequence f
m
:=

j
a
m
j
B
j
, m = 1, 2, . . ., whose coecients
converge monotonely to some sequence a

. Its corresponding sequences y


m
1
< < y
m
n
of extrema of f
m
therefore also converge, necessarily to a strictly increasing sequence
x

1
< < x

n
since f
m
strictly alternates in sign on y
m
1
< < y
m
n
. Let f

:=

j
a

j
B
j
.
Then,
(134) 1 = (1)
ni
f

(x

i
) = |f

, i = 1, . . . , n,
since 1 = (1)
ni
f
m
(y
m1
i
), while |f
m
| = max
i
[f
m
(y
m
i
)[ and f

is the uniform limit of


(f
m
: m).
31mar03 70 c _2003 Carl de Boor
(135) Lemma. Let f

j
a

j
B
j,k,t
satisfy (134) for some (strictly) increasing x

. Then
(1)
ni
a

i
= 1/ dist (B
i
, span(B
j
: j ,= i)), i = 1, . . . , n.
In particular, f

is independent of x

(hence of the initial x in the above iteration).


Proof: Let A

:= (B
j
(x

i
)) and set

i
: S
k,t
R : f

j
A
1

(i, j)f(x

j
).
Then
i
B
j
=
ij
, hence
(1)
ni
a

i
= (1)
ni

i
f

j
(1)
ni
A
1

(i, j)(1)
nj
=

j
[A
1

(i, j)[ = |
i
|,
the last equality by the fact that

j
[A

(i, j)[ is obviously an upper bound for |


i
|, yet
it equals [
i
f

[ with |f

= 1, hence is also a lower bound for |


i
|. Now, for any linear
functional on any nls X and for any f X,
[f[ = || dist (f, ker ).
Hence, 1 = dist (f

, ker
i
), while ker
i
= span(B
j
: j ,= i) and so [a

i
[ dist (B
i
, ker
i
) =
dist (f

, ker
i
) = 1.
The function f

is, by denition, the Chebyshev spline for S


k,t
, i.e., the unique
(up to scalar multiples) element that maximally equioscillates, i.e., satises (134). From
Lemma (135), we can write it as
T
k1,t
:=

j
(1)
nj
B
j
/ dist (B
j
, span(B
i
: i ,= j)).
To be sure, any f S
k,t
can have at most n 1 sign changes, hence n is indeed the
maximal number of equioscillations possible for f S
k,t
.
(136) Proposition (S. Demko). x |a
x
|

is uniquely minimized when x is the


extreme-point sequence for the Chebyshev spline.
Proof: We just saw that, starting with any x, we reach the Chebyshev spline in
the limit in a process during which the B-spline coecients decrease in absolute value.
Hence |a

|a
x
|

for any x.
31mar03 71 c _2003 Carl de Boor
The terminology Chebyshev spline, though apt in view of the fact that, for t
1
= =
t
k
= 1, t
k+1
= = t
2k
= 1 it is the Chebyshev polynomial of degree k1, unfortunately
clashes with the standard term Tschebyschean spline, meaning a piecewise function
whose pieces all come from the same Haar space. The notation T
k1,t
is meant as a
legitimate extension of the notation T
r
for the Chebyshev polynomial of degree r. Although
whole books have been written on the special case k = n, in which case T
k1,t
on I
k,t
agrees
with the (suitably scaled and translated) Chebyshev polynomial T
k1
of degree k 1, the
Chebyshev spline is largely unexplored territory (except for K. Mrkens Ph.D. Thesis).
(137) Corollary. For x = (x
1
< < x
n
) with t
i
< x
i
< t
i+k
, all i, let f
x
=

j
a
x
j
B
j
be
the unique element in S
k,t
satisfying f(x
i
) = (1)
ni
, all i. Then argmin
x
|a
x
|

equals
the extreme-point sequence of T
k1,t
.
Ohand, some of the extreme points of T
k1,t
may lie outside the basic interval
I
k,t
= [t
k
. . t
n+1
]. However, if we restrict attention to this interval, then we would choose
t
1
= t
k
and t
n+1
= t
n+k
. This violates the assumption that #t
i
< k, all i. However, assume
rst that, e.g., t
1
< t
2
= t
k
. Then f
x
is strictly monotone on [t
1
. . t
2
], hence necessarily
t
2
y
1
. By the same reasoning, y
n
t
n+k1
in case t
n+k1
< t
n+k
. This implies
that nothing in the above arguments changes if we use the interval [t
2
. . t
n+k1
] instead.
That choice made, the location of t
1
and t
n+k
becomes irrelevant to S
k,t
as restricted to
[t
2
. . t
n+k1
] (and not even the B-spline coecients will change as we vary t
1
and t
n+k
).
In particular, we may choose t
1
= t
k
and t
n+k
= t
n+1
, hence have [t
1
. . t
n+k
] = I
k,t
.
Since, for all i,
i,k
=

j
A
1

(i, j)(x

j
) on S
k,t
, and the latter linear functional, as
we have just seen, takes on its norm on T
k1,t
, it follows that the map
S
k,t
R :

j
a
j
B
j
|a|

/|

j
a
j
B
j
|
takes on its maximum at T
k1,t
, and that maximum is |a

. This maximum determines


the condition of the B-spline basis, to be discussed next.
The condition of the B-spline basis
The condition
(V ) := |V ||V
1
|
of a basis V (i.e., an invertible linear map from some F
n
to the normed linear space ranV )
measures the extent to which the relative changes in the coordinates a of an element V a
may be close to the resulting relative change in V a itself. The closer (V ) to 1, the more
closely do these two relative sizes correspond.
For the B-spline basis, using the max-norm both in R
n
and in S
k,t
, we have

([B
j
: j]) = sup
a
|a|

j
a
j
B
j
|

since the B
j
form a partition of unity, hence trivially sup
a

j
a
j
B
j

= 1.
Further,

([B
j
: j]) = max
i
|
i,k
|.
Hence the following lemma is of interest.
31mar03 72 c _2003 Carl de Boor
(138) Lemma. The number
d
k
:= sup
t
sup
i
sup
fS
k,t
[
i,k,t
f[/|f|

is nite.
Proof: Let f S
k,t
, hence

i,k
f =
k

=1
(D)
1

i
(
i
)/(k 1)! D
k
f(
i
),
and let I := [t

. . t
+1
] be any knot interval in [t
i
. . t
i+k
], and choose, as we may,
i
I.
Then, as both
i
and f are polynomials of degree < k on I, Markovs inequality (65)
implies that
[(D)
1

i
(
i
)[[D
k
f(
i
)[ const
k,
|
i
|

(I)/[I[
1
|f|

(I)/[I[
k
.
On the other hand, by choosing, as we may, I to be a largest such knot interval, we can
ensure that
[t
j

i
[ k[I[, j = i, . . . , i +k,
therefore |
i
|

(I) const
k
[I[
k1
. Therefore, altogether, [
i,k
f[ const

k
|f|

(I), which
is even stronger than the claim to be proved.
A more careful quantitative analysis shows that d
k
= O(9
k
). Better results can be
obtained with the aid of the following
(139) Claim. d
k
= sup
s
|
k1,k
S
k,s
|([1. .1]), with ||(I) := sup
fdom
[f[/|f|

(I),
and s any knot sequence of the type
s
1
= = s
k
= 1 s
k+1
s
2k3
1 = s
2k2
= = s
3k3
.
Proof: Consider any particular
i
:=
i,k
S
k,s
. If t
i+1
= t
i+k1
, then (assuming
WLOG that t
i
< t
i+1
),
i
= (t
i+1
), therefore |
i
| = 1 d
k
. In the contrary case, we
may assume, after a suitable linear change of the independent variable, that 1 = t
i+1
,
t
i+k1
= 1. Let

t be the knot sequence obtained from t by inserting both 1 and 1
enough times to increase their multiplicity to k 1, and let

i be such that

t

i+j
= t
i+j
for
j = 1, . . . , k1. The corresponding spline space S
k,

t
may well be larger than the space S
k,t
we started with, but
i
=

i
since, by (97),
i
only depends on the knots t
i+1
, . . . , t
i+k1
.
This shows that
|
i
| := sup
fS
k,t
[
i
f[/|f|

sup
fS
k,

t
[

i
f[/|f|

= |

i
| = |

i
|([1 . . 1]),
the last equality since

i
f only depends on f
[1..1]
.
Consequently, d
k
sup
s
|
k1,k
|([1 . . 1]). The opposite equality is trivial.
31mar03 73 c _2003 Carl de Boor
It is believed that d
k
2
k
. However, earlier hopes that the extremal knot congura-
tion in the Claim would have no interior knots (i.e., would have an s with all its entries
from 1, 1) were dashed by the following simple counter-example: The cubic Chebyshev
polynomial can be shown to have maximum B-coecient 5 when written as an element of
S
k,t
with t = (1, 1, 1, 1, 1, 1, 1, 1); however, the cubic Chebyshev spline for the knot
sequence (1, 1, 1, 1, 0, 1, 1, 1, 1) has B-coecients (1, 7/2, 11/2, 7/2, 1). Neverthe-
less, by considering the related extremum problem
(140) d
k,1
:= sup
t
sup
i
sup
fS
k,t
[
i,k,t
f[[t
i
t
i+k
[/|f|
1
([t
i
. . t
i+k
]),
Karl Scherer and Aleksei Shadrin were recently able to show that d
k
k2
k
.
It follows that
|a|

/d
k
|

i
B
i,k
a
i
|

|a|

for any knot sequence t and any coecient sequence a.


For such an estimate in L
p
, observe that, for any p and with 1/p + 1/p

= 1 and by
Holders Inequality,
[

j
B
j
(x)a
j
[ (

j
B
j
(x)[a
j
[
p
)
1/p
(

j
B
j
(x))
1/p

,
hence
|

j
B
j
a
j
|
p
p

j
[a
j
[
p
(t
j+k
t
j
)/k
(using the fact that
_
B
j
= (t
j+k
t
j
)/k). On the other hand, from (140) and Hahn-
Banach, we deduce the following
(141) Proposition. There exists h
j
with [h
j
[ d
k,1
/(t
j+k
t
j
) and support in [t
j
. . t
j+k
]
for which
_
R
h
j
f =
_
t
j+k
t
j
h
j
f =
j,k
f, f S
k,t
.
In particular,
d
k
d
k,1
.
Consequently, if f =

j
B
j
a
j
and with
_
j
:=
_
t
j+k
t
j
, then
a
j
=
_
j
h
j
f (
_
j
[h
j
[
p

)
1/p

(
_
j
[f[
p
)
1/p
,
while
(
_
j
[h
j
[
p

)
1/p

d
k,1
(t
j+k
t
j
)
(t
j+k
t
j
)
1/p

= d
k,1
/(t
j+k
t
j
)
1/p
.
Therefore,

j
[a
j
[
p
(t
j+k
t
j
)/k (d
k,1
)
p

j
_
j
[f[
p
/k (d
k,1
)
p
|f|
p
p
,
31mar03 74 c _2003 Carl de Boor
the last inequality since at most k of the intervals [t
j
. . t
j+k
] have some given knot interval
[t

. . t
+1
] in common. It follows that, for any 1 p ,
(142) |c|
p
/d
k,1
|

i
B
i,k,p
c
i
|
p
|c|
p
,
for any knot sequence t and any coecient sequence c, with
(143) B
i,k,p
:= (k/(t
i+k
t
i
))
1/p
B
i,k
.
Degree of approximation by splines
quasiinterpolants
It follows from Theorem (111) that the linear map
P : g

j
B
j,k

j,k
g
is a linear projector, with range S = S
k,t
=

k,t
, provided (as we have already assumed)
that, for each i, we choose in (97) equal to some
i
[t
i
. . t
i+k
). It is also local, since

j,k
g depends only on the behavior of g near
j
, hence, by our choice of
j
, only on the
behavior of g on supp B
j,k
.
However, P is dened only for suciently smooth functions. In order to get such a
projector on all of L
1
(I
k,t
), we make use of (141) in order to obtain the linear functional

j
g :=
_
h
j
g
with supp h
j
[t
j
. . t
j+k
] and |h
j
|

d
k,1
/(t
j+k
t
j
) that, on S
k,t
, agrees with
j,k,t
.
This implies that, for any g L
1
and for 1 p ,
[
j
g[
d
k,1
(t
j+k
t
j
)
1/p
|g|
p
([t
j
. . t
j+k
]).
The corresponding linear map
Q : g

j
g B
j,k
is dened for any g L
1
(I
k,t
), hence for any g L
p
(I
k,t
). Further, it is the identity on its
range, hence a linear projector, its norm is bounded by d
k,1
, and it is local, in the sense
that, for x [t
l
. . t
l+1
),
(144)
Qg(x) =
l

j=l+1k

j
g B
j
(x)
l

j=l+1k
[
j
g[ B
j
(x)
d
k,1
|g|
p
([t
l+1k
. . t
l+k
])
l

j=l+1k
B
j
(x)/(t
j+k
t
j
)
1/p
d
k,1
|g|
p
([t
l+1k
. . t
l+k
]) max
j{l+1k,...,l}
(t
j+k
t
j
)
1/p
.
8apr03 75 c _2003 Carl de Boor
Since also g Qg = (1 Q)(g q) for any q
<k
, we conclude that, for any g,
(145) |g Qg|
p
([t
l
. . t
l+1
]) (1 +d
k,1
) dist (g,
<k
)
p
([t
l+1k
. . t
l+k
])
(since (t
l+1
t
l
)/(t
j+k
t
j
) 1 for j = l + 1 k, . . . , l). In fact, this conclusion can
already be reached if we only take care that
j
agree with
j
on
<k
, all j, while still
[
j
g[ d
k,1
|g|
1
([t
j
. . t
j+k
])/(t
j+k
t
j
).
The resulting map Q is called a good quasiinterpolant of order k, with good
referring to the uniform localness expressed by (144), and order k referring to the fact
that Q reproduces
<k
. The term quasiinterpolant was chosen by nite-element people
once they realized that approximation order could be ascertained with the aid of maps
Q that did not actually interpolate at the nodal points of their elements, but merely
matched enough information to give reproduction of certain polynomial spaces.
The error bound (145) is local; it is in terms of how well g can be approximated locally
from polynomials of order k. That local distance is best estimated with the aid of
(146) Whitneys Theorem. For any nite interval I,
dist (g,
<k
)
p
(I)
k
(g, [I[)
p
.
Proof: For all f
<k
,
k
(g, [I[)
p
=
k
(g f, [I[)
p
2
k
|g f|
p
(I), hence

k
(g, [I[)
p
const
k
dist (g,
<k
)
p
(I).
For the converse inequality, let I =: [a . . b] and start with an arbitrary f W
(k)
p
(I).
With
T
k
f :=

j<k
D
j
f(a)( a)
j
/j!
its truncated Taylor series, we nd
[f(x) T
k
f(x)[ = [
_
I
(x )
k1
+
/(k 1)! D
k
f[
|D
k
f|
p
/(k 1)!
_
_
I
[(x )
k1
+
[
p

_
1/p

|D
k
f|
p
/(k 1)! [I[
k1/p
,
hence, for any 1 q ,
|f T
k
f|
q
(I) |D
k
f|
p
/(k 1)! [I[
k1/p+1/q
.
Consequently,
dist (g,
<k
)
p
(I) |g T
k
f|
p
(I) |g f|
p
+[I[
k
|D
k
f|
p
/(k 1)!,
and, as f W
(k)
p
(I) is arbitrary here, we get
dist (g,
<k
)
p
(I) const
k
K(g, [I[
k
; L
p
(I), W
(k)
p
(I)),
while, from Theorem (68), we know that
k
(g, t)
p
K(f, t
k
; L
p
, W
(k)
p
).
8apr03 76 c _2003 Carl de Boor
We conclude from Whitneys theorem and from (145) that
(147) dist (g, S
k,t
)
p
const
k

k
(g, [t[)
p
, 1 p .
More than that, since

k
(g, h)
p
(I) = O(h
k
) g W
(k)
p
(I),
with
sup
h

k
(g, h)
p
h
k
=
_
|D
k
g|
p
(I), p > 1;
Var(D
k1
g), p = 1,
we have the local bound
|g Qg|
p
([t
l
. . t
l+1
]) const
k
[t
l+k
t
l+1k
[
k
|D
k
g|
p
([t
l+1k
. . t
l+k
]).
For p = , this suggests that, in approximating some g that is smooth except for some
isolated singularities, the knot sequence t be chosen so as to make
l [t
l
t
l+1
[
k
|D
k
g|

([t
l
. . t
l+1
])
approximately constant. This is equivalent to making the map
l [t
l
t
l+1
[
_
|D
k
g|

([t
l
. . t
l+1
])
_
1/k
approximately constant, or, at least for a large knot sequence, making
l [t
l
t
l+1
[[D
k
g([t
l
. . t
l+1
])[
1/k
approximately constant. If there are to be n knot intervals, then we can achieve this
(approximately), by choosing t
0
= 0 < t
1
< < t
n
= 1 so that
l
_
t
l+1
t
l
[D
k
g[
1/k
is constant. With that choice,
1
n
k
|D
k
g|
1/k
=
_
_
t
l+1
t
l
[D
k
g[
1/k
_
k
[t
l+1
t
l
[
k
|D
k
g|

([t
l
. . t
l+1
]),
hence
|g Qg|

const
k
n
k
|D
k
g|
1/k
.
While this argument lacks some details, it makes the following essential point: In
order to achieve approximation order n
k
from the set of splines of order k with n interior
knots, it is sucient to have
|D
k
g|
1/k
nite. For example, this norm is nite for functions such as [ [
1/2
on [1 . . 1], ensuring
therefore approximations by splines of order k whose error behaves like O(n
k
), with n
the degrees of freedom used. In contrast, the error in best polynomial approximation from

n
to [ [
1/2
on [1 . . 1] cannot be better than O(n
1/2
), by Bernsteins Inverse Theorem,
hence approximation by splines with n equally spaced knots cannot be better, either, as
we show in the next section.
8apr03 77 c _2003 Carl de Boor
Jackson and Bernstein for splines with uniform knot sequence
We know from (147) that, for every g W
(k)
p
,
(148) dist (g, S
k,t
)
p
const
k
[t[
k
|D
k
g|
p
.
However, having dist (g, S
k,t
)
p
= O([t[
k
) is, in general, no guarantee that g W
(k)
p
unless
the knot sequences t involved are suciently generic. Indeed, if every knot sequence
considered contains the point 1/2, then the function g := ( 1/2)
k1
+
can be approximated
without error from S
k,t
even though g fails to have a kth derivative. But it is true that
(148) cannot hold for every knot sequence t unless g W
(k)
p
. In fact, this conclusion can
already be reached if (148) holds for every uniform knot sequence. The essential point is
that, for every point in the interval of approximation, there must be innitely many knot
sequences among those considered for which that point falls somewhere in the middle of a
knot interval. Here are the details.
(149) Theorem (Butler, DeVore, Richards). For m N, let
t
(m)
:= (. . . , 0, 0, 1/m, 2/m, . . . , 1, 1, . . .),
and set
E
m
(g)
p
:= dist (g, S
k,t
(m))
p
(I)
with I := [0 . . 1]. Then,

k
(g, )
p
const
k
_
_
1
n

2n
m=n
E
m
(g)
p
p
_
1/p
, p < ;
max
nm2n
E
m
(g)

, p = ,
with
n := 1/.
Proof: Since
[
k
h
g(x)[ = [
k
h
(g f)(x)[ 2
k
|g f|

(x . . x +kh)
for every f
<k
, the heart of the proof is in the (nontrivial) observation (known as
a mixing lemma, see(150) below) that, for all n, there exists m n, . . . , 2n so that
dist (x, t
(m)
) 1/(16n), hence, for our n and for all 0 < h
1
:= 1/(16kn), since
kh k
1
16kn
= 1/(16n), there is m n, . . . , 2n so that [x . . x +kh] t
(m)
= , implying
that any f S
k,t
(m) is a polynomial on [x . . x +kh], hence, with f a ba to g from S
k,t
(m) ,
we get
[
k
h
g(x)[ 2
k
E
m
(g)

.
It follows that

k
(g,
1
) 2
k
max
nm2n
E
m
(g)

.
8apr03 78 c _2003 Carl de Boor
Since 1/n = 16k
1
, hence
k
(g, )
k
(g, 1/n) (16k)
k

k
(g,
1
), this implies the
claimed result for p = .
As to p < , the mixing lemma (150) implies that
2n

m=n

I
m
(x)
n
64
on [0 . . 1 kh],
with
I
m
:= x [0 . . 1 kh] : [x . . x + 1/(16n)] t
(m)
= .
Therefore,
1
64
_
1kh
0
[
k
h
g(x)[
p
dx
1
n
2n

m=n
_
I
m
[
k
h
g(x)[
p
dx
2
kp
n
2n

m=n
E
m
(g)
p
p
.
This proves the result for p < .
For the record, here is the afore-mentioned
(150) Mixing Lemma. For any x [0 . . 1] and any n N,
#n m 2n : dist (x, t
(m)
) 1/(16n) n/64.
Its proof (see, e.g., [DeVore and Lorentz, Constructive Approximation: pp. 356-7])
relies on the fact that, for any i, N N with i < N,
#N m 2N : dist (i/N, t
(m)
) 1/(6N) N/16.
We conclude from the Theorem that
E
n
(g)
p

k
(g, 1/n)
p
.
In particular,
E
n
(g)
p
= O(n
k
) g
_
W
(k)
p
1 < p ;
W
(k1)
(BV ) p = 1.
Further, we get the saturation result:
E
n
(g)
p
= o(n
k
) g
<k
.
For the characterization of other rates of convergence, we make use of the following stan-
dard way to measure convergence behavior.
11apr03 79 c _2003 Carl de Boor
Approximation Spaces
It has become standard to measure the decay of E
n
(g) := dist (g, M
n
) of the distance
of g X from a given sequence (M
n
: n N) of subsets by comparing it to the sequence
(n

: n N) for various values of . Most simply, one might ask to have


E
n
(g) = O(n

),
with the supremum over all such then being the approximation order for g provided
by (M
n
: n N). However, in order to be able to characterize the class of functions g with
a given approximation order, a somewhat more subtle way of measuring approximation
order turns out to be often needed.
Dene
|a|
()
q
:=
_
_
n
(n

[a(n)[)
q
/n
_
1/q
0 < q < ;
sup
n
n

[a(n)[ q = .
With this, we dene
A
()
q
:= a R
N
: [a[ 0, |a|
()
q
< ,
and make the following observations. First,
|a|
()

< a(n) = O(n

),
i.e., A
()

consists exactly of all the antitone sequences a that go to zero (at least) to
order . Further, |a|
()
q
< for some nite q implies that a(n) = o(n

). (Indeed, if
a(n) ,= o(n

), then m(n)

[a(m(n))[ M > 0 for some strictly increasing m : N N, and


AGTASMAT m(n 1) < m(n)/2; thus, (|a|
()
q
)
q

m(n)/2jm(n)
(j

[a(j)[)
q
/j

n
(2

M)
q

m(n)/2jm(n)
1/j (since j

[a(j)[ = (j/m(n))

m(n)

[a(m(n))[
2

M for m(n)/2 j < m(n), and

m(n)/2jm(n)
1/j ln2), a contradiction.) In
particular
N := |()

a|

< ,
hence
> (|a|
()
q
)
q
=: N
q

n
[b(n)[
q
/n N
q

n
[b(n)[
r
/n
for any r > q since [b(n)[ 1, all n. In particular,
|a|
()
r
= N
_

n
[b(n)[
r
/n
_
1/r
is nite for any r > q. In other words,
q < r = A
()
q
A
()
r
.
11apr03 80 c _2003 Carl de Boor
Finally, under the same assumption, and for any positive and any r,

n
[n

a(n)[
r
/n =

n
[n

a(n)[
r
/n
1+r
N
r

n
n
1r
< .
Hence,
< = A
()
q
A
()

A
()
r
, all q, r.
For a given sequence (M
n
: n N) of subsets of the nls X, one denes, correspondingly,
the approximation classes
A
()
q
(X, (M
n
)) := g X : |(E
n
(g) : n N)|
()
q
<
that single out all the elements g of X for which E
n
(g) goes to zero in a certain way. By
the earlier discussion of (, q) |a|
()
q
, we conclude that (, q) A
()
q
is antitone in
for arbitrary q and isotone in q for xed , i.e.,
< = A
()
q
A
()
r
,
q < r = A
()
q
A
()
r
.
There is a corresponding way to quantify and compare the speed at which, for a
function f, f(t) approaches 0 as t 0. Precisely, for such a function, one denes
|f|
()
q
:=
_
_ _
I
[t

f(t)[
q
dt/t
_
1/q
, q < ;
sup
I
[t

f(t)[, q = ,
with I some suitable interval, e.g., I = [0 . . 1]. Usually, f is isotone. In that case,
|f|
()
q
is equivalent to (i.e., bounded above and below by certain f-independent positive
multiples of) the following discrete versions, in which I is replaced by the sequence (a)
I = (1/n : n N) or (b) I = (2
n
: n = 0, 1, . . .), and, correspondingly, on the interval
[t
n+1
. . t
n
], dt/t is replaced by (t
n
t
n+1
)/t
n
, which, for (a), is 1/(n +1), and, for (b),
is a constant. For q < , this gives the equivalent discrete versions
|f|
()
q
(1/N) := |(f(1/n) : n N)|
()
q
and
|f|
()
q
(2
N
) :=
_

n
(2
n
f(2
n
))
q
_
1/q
,
respectively.
With these denitions in place, consider again the earlier result that
E
n
(g)
p
:= dist (g, S
k,t
(n))
p

k
(g, 1/n)
p
,
hence
g A
()
q
(L
p
, (S
k,t
(n) : n N)) |
k
(g, )
p
|
()
q
< .
11apr03 81 c _2003 Carl de Boor
The latter condition appeared some time ago in the study of approximation order from
trigonometric polynomials, leading to what we now call Besov spaces. Precisely,
B
()
q
(L
p
) := g L
p
: [g[
B
()
q
(L
p
)
:= |
+1
(g, )
p
|
()
q
< , 0 < , 0 < q .
These are complete metric spaces with the metric given by the (quasi-)norm
|g|
B
()
q
(L
p
)
:= |g|
p
+[g[
B
()
q
(L
p
)
.
(This is only a quasi-norm when p < 1 since then one only has |a+b| const(|a| +|b|)
for some const instead of the triangle inequality.)
(151) Fact. For any r > ,
g |
r
(g, )
p
|
()
q
provides a (quasi-)seminorm on B
()
q
(L
p
) equivalent to [ [
B
()
q
(L
p
)
.
One direction of this claim is obvious since

r+m
(g, t)
p
2
m

r
(g, t)
p
.
For the other direction, one needs a result like
(152) Marchauds Theorem. For g L
p
(I),

r
(g, t)
p
const
r
t
r
_
_
|I|
t

r+m
(g, s)
p
s
r+1
ds +
|g|
p
[I[
r
_
.
along with
(153) Hardys Inequality. For > 0, and 1 q , and f any positive measurable
function,
_

0
_
t

_
t
0
f(s)
ds
s

q dt
t

1

q
_

0
_
t

f(t)

q dt
t
.
With this, we reach the conclusion that
(154) A
()
q
(L
p
(0 . . 1), (S
k,t
(n) : n N)) = B
()
q
(L
p
(0 . . 1)), 0 < q , 0 < < k.
But we have to question just what we have gained by this somewhat formal exercise. The
gain is substantial to the extent that we understand
11apr03 82 c _2003 Carl de Boor
Besov spaces
Such understanding comes from knowing more about typical elements and/or from
knowing alternative characterizations of such spaces. Among these will be those derived
from the fact that Besov spaces turn out to be approximation spaces for various other se-
quences of approximating sets, among these S
k,n
:= splines with n free knots and
n
/
n
:=
rational functions with numerator and denominator degree n, both classical examples
of nonlinear approximating sets, to be discussed next.
As for specic examples, consider the Heaviside function
g := ()
0
+
,
as an element of L
p
[1 . . 1], say. For 0 < h < 1, |
h
g|
p
p
= h, hence

1
(g, h)
p
= h
1/p
.
Therefore, for 0 < < 1 and 0 < q < ,
(|(g, )
p
|
()
q
)
q

_
1
0
(t

t
1/p
)
q
dt/t =
_
1
0
()
(+1/p)q1
,
and this is nite i < 1/p. For q = , we look instead at
sup
t
t

t
1/p
,
and this is nite i 1/p. Consequently,
()
0
+
B
()
q
(L
p
)
_
< 1/p, if 0 < q < ;
1/p, if q = .
Note the minor, yet decisive, role played by the parameter q here. Note also that can
be large provided we are willing to consider p < 1. Finally, note the implication that
any function with nitely many jump discontinuities but that is otherwise smooth lies in
B
()
q
(L
p
) for exactly the same triples (, q, p).
More generally,
(
k
h
()
k1
+
)(x) = k!h
k
(0, h, . . . , kh)(x +)
k1
+
= (k 1)!h
k1
B(x[0, h, . . . , kh),
hence, |
k
h
()
k1
+
|
p
h
k1+1/p
, and so, for 0 < q < ,
(|
k
(()
k1
+
, )
p
|
()
q
)
q

_
1
0
(t

t
k1+1/p
)
q
dt/t =
_
1
0
()
(+k1+1/p)q1
while
|
k
(()
k1
+
, )

|
()

sup
t
t
+k1+1/p
.
18apr03 83 c _2003 Carl de Boor
Therefore,
()
k1
+
B
()
q
(L
p
)
_
< k 1 + 1/p, if 0 < q < ;
k 1 + 1/p, if q = .
Other examples include:
B
(r)

(L
p
(I))
_
W
(r)
p
(I), if 1 < p < ;
W
(r1)
BV (I), if p = 1,
, r N,
i.e., for 1 p < and integer r, B
(r)

(L
p
(I)) contains the Sobolev space of all functions
on I with absolutely continuous (r 1)st derivative and (a) rth derivative in L
p
if p > 1;
(b) (r 1)st derivative of bounded variation, if p = 1. There is no equality here, the only
related equality being
B
(r)
2
(L
2
) = W
(r)
2
.
For p = , one has
B
(r)

(C(I)) = Lip(r + 1, C(I)),


with the special case B
(1)

(C(I)) equal the Zygmund space, i.e., slightly larger than


Lip
1
(I). Note the somewhat more subtle description in the extreme cases p = 1, .
More generally, for nonintegral , and with =: r 1 + for some r N and
(0 . . 1),
B
()

(L
p
(I)) = Lip(, X
p
(I)),
the space of all functions on I with absolutely continuous (r 1)st derivative and rth
derivative in Lip

(I)
p
. In view of the fact that
Lip

(I)
p
:= g L
p
(I) : sup
h
|
h
g|
p
/h

< ,
this is certainly just a tautology.
Besov spaces are helpful also because they appear as the right spaces in interpolation
between standard spaces and in the Sobolev embedding theorem and its generalization.
For the discussion of these matters, it is very helpful to follow Ron DeVores advice and
view the whole situation by representing all the spaces (B
()
q
(L
p
) : 0 < q ) by the
point
(1/p, )
in R
2
+
, as in Figure (155).
Besov spaces arise naturally in interpolation between the spaces L
p
and W
(r)
p
. Briey,
for any pair (X
0
, X
1
) of (quasi-)normed spaces with
X
1
X
0
,
i.e., X
1
continuously imbedded in X
0
, one obtains a two-parameter continuum of spaces
X
1
(X
0
, X
1
)
,q
X
0
11apr03 84 c _2003 Carl de Boor
1
C
()
+
1
p
B
()
(L
1/
)
L
p
L
(1/p)
1
B
()

(L
p
)
L
(1/p+1)
1
1
C
(155) Figure. DeVores diagram associates with the point (
1
p
, ) the whole
family B
()

(L
p
) := (B
()
q
(L
p
[0 . . 1]) : 0 < q ) of Besov
spaces, to facilitate discussion (and retention) of basic facts
about embeddings, interpolation in smoothness spaces, as well
as the essential dierence between linear vs nonlinear approx-
imation.
The shaded triangle comprises all the Besov spaces into
which the marked space B
()
q
(L
p
) is continuously embedded,
with the precise choice of the secondary parameter, q, of import
only along the slanted edge.
as
X
,q
:= (X
0
, X
1
)
,q
:= g X
0
: [g[
,q
:= |K(g, )|
()
q
< ,
with
K(g, t) := K(g, t; X
0
, X
1
) := inf
fX
1
(|g f|
X
0
+t|f|
X
1
)
the K-functional for the pair (X
0
, X
1
).
The main result concerning this interpolation of spaces is the following
(156) Theorem. Let X
1
X
0
and Y
1
Y
0
be pairs of complete (quasi-)normed
spaces and assume that the linear map U maps X
i
boundedly into Y
i
, i = 0, 1. Then, for
0 < q and 0 < < 1, U also maps each X
,q
boundedly into Y
,q
and, with
M
i
:= |U : X
i
Y
i
|, i = 0, 1,
one has
|U : X
,q
Y
,q
| M
1
0
M

1
.
(See Theorem 7.1 in Chapter 6 of DeVore-Lorentz.)
11apr03 85 c _2003 Carl de Boor
As an application, take X
0
= L
p
(I) and X
1
= W
(k)
p
(I). Then, for any 0 < < k,
(X
0
, X
1
)
/k,q
= B
()
q
(L
p
),
(using Theorem (68)). Further, with Y
0
= L
p
= Y
1
, we take for U the error 1 Q in
the quasi-interpolant Q introduced earlier. From (145) and Whitneys Theorem (146), we
know that then
M
1
:= |(1 Q) : X
1
X
0
| const
k
[t[
k
,
while
M
0
:= |(1 Q) : X
0
X
0
| 1 +d
k,1
< .
Therefore, for any 0 < < k,
|(1 Q) : B
()
q
(L
p
) L
p
| const
k
[t[

,
thus providing (a lower bound on) the approximation order from S
k,t
to elements of
B
()
q
(L
p
), hence, in particular, of L
()
p
(I) in case is an integer.
Nonlinear approximation
We know from Theorem (149) that we cannot have dist (g, S
k,t
)
p
= O(t
k
) for all t
unless g W
(k)
p
. Yet, we already observed that some functions without a kth derivative,
like g :=
1/2
on I := [0 . . 1], can nevertheless be approximated to O(n
k
) by a spline of
order k with n suitably chosen interior knots.
The basic results here are the following. Let
M
n
:= S
k,n
be the space of all splines of order k on I with < n interior knots, hence with at most
n polynomial pieces. Then, M
n
is scale-invariant but fails to be closed under addition.
However,
M
n
+M
n
M
2n
,
hence Theorem (56) is applicable here provided we can produce compatible Jackson and
Bernstein inequalities. These were obtained not all that long ago by Petrushev.
(157) Theorem (Petrushev). Let 0 < p < and, correspondingly,
B
()
:= B
()

(L

(0 . . 1)), := 1/( + 1/p).


then, for 0 < < k,
(i) g L
p
, dist (g, S
k,n
) const
k
n

[g[
B
() .
(ii) m S
k,n
, |m|
B
() const
k
n

|m|
p
.
(See Theorem 8.2 in Chapter 12 of DeVore-Lorentz.) In light of Peetres Theorem (56),
this says that these particular Besov spaces consist exactly of those functions that can be
11apr03 86 c _2003 Carl de Boor
approximated in L
p
to O(n

) by splines with n polynomial pieces. To get that kind of


approximation order for splines with n polynomial pieces and a uniform knot sequence
(i.e., linear approximation), we need g to lie in B
()

(L
p
). In the DeVore diagram (155),
both spaces lie on the same horizontal line, but the latter is much further to the left (i.e.,
much smaller) than the former, and so nicely illustrates the gain available to nonlinear
approximation.
As a quick example, consider the simplest possible case, that of approximation from
M
n
:= S
1,n
, the (nonlinear) space of all step functions on [0 . . 1] with at most n dierent
values. Already in 1961, Kahane proved the following neat result:
(158) Proposition (Kahane). For g C([0 . . 1]), dist (g, S
1,n
) M/(2n) for all n N
if and only if Var(g) M.
Proof: There is nothing to prove if M or Var(g) are innite, hence assume that
both are nite.
Choose 0 = t
0
< < t
n
= 1 so that Var(g)(t
i
. . t
i+1
) Var(g)/n, all i. Let f S
1,n
be such that, on (t
i
. . t
i+1
), f equals the midpoint of the interval g([t
i
. . t
i+1
]), all i. Then,
|g f|


1
2
Var(g)/n.
Conversely, for any f S
1,n
, and any 0 = x
0
< < x
m
= 1,
[g(x
i
) g(x
i1
)[ 2|g f|

#f([x
i1
. . x
i
]),
therefore

i
[g(x
i1
) g(x
i
)[ 2|g f|

i=1
#f([x
i1
. . x
i
]) 2(n +m)|g f|

.
Hence, if dist (g, S
1,n
) M/(2n), then, for any > 0,

i
[g(x
i1
) g(x
i
)[ (M +)(1 +m/n),
therefore, by letting n , Var(g) M +.
Thus, while dist (g, S
1,t
(n))

= O(1/n) requires g to lie in W


(1)

, getting dist (g, S


1,n
) =
O(1/n) only requires g BV, i.e., in B
(1)

(L
1
). In the DeVore diagram (155), W
(1)

lies vertically above C, while the latter space lies on the 45-degree line emanating from
C. Petrushevs result shows this to hold for splines of general order k. Namely, the
approximation order from splines of order k cannot exceed k, but that order is achieved
by nonlinear approximation (i.e., approximation from (S
k,n
: n N)) to much rougher
functions than is possible by linear approximation (i.e., approximation from (S
k,t
(n) : n
N)).
See DeVores survey article on Nonlinear Approximation, in Acta Numerica, 1998.
21apr03 87 c _2003 Carl de Boor

You might also like