Lectures On Approximation by Polynomials: J.G. Burkill
Lectures On Approximation by Polynomials: J.G. Burkill
Lectures On Approximation by Polynomials: J.G. Burkill
Approximation By Polynomials
By
J.G. Burkill
1 Weierstrass’s Theorem 1
1 Approximation by Polynomials . . . . . . . . . . . . . . 1
2 Singular Integrals and Landau’s Proof . . . . . . . . . . 4
3 Bernstein Polynomials . . . . . . . . . . . . . . . . . . 7
3 Approximations to |x| 19
7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
8 Oscillating polynomials . . . . . . . . . . . . . . . . . . 20
9 Approximation to |x| . . . . . . . . . . . . . . . . . . . 25
4 Trigonometric Polynomials 29
10 Trigonometric polynomials. Modulus of Continuity . . . 29
11 Fourier and Fejer Sums . . . . . . . . . . . . . . . . . . 33
5 Inequalities, etc. 41
12 Bernstein’s and Markoff’s Inequalities . . . . . . . . . . 41
13 Structural Properties Depend on the... . . . . . . . . . . 45
14 Divergence of the Lagrange Sequence . . . . . . . . . . 47
iii
iv Contents
1 Approximation by Polynomials
n
ar xr is that its value for 1
P
A basic property of a polynomial P(x) =
0
a given x can be calculated (e.g. by a machine) in a finite number of
steps. A central problem of mathematical analysis is the approximation
to more general functions by polynomials an the estimation of how small
the discrepancy can be made. A discussion of this problem should be
included in any University course of analysis. Not only are the results
important, but their proofs admirably illustrate a number of powerful
methods.
This account will be confined to the leading theorems, stated in their
fundamental rather than their most general forms. There are many ex-
cellent systematic presentations in the literature, to which this may serve
as an introduction.
Variables and functions will be real. We say that f (x) is C(a, b)
meaning that f (x) is continuous for a ≤ x ≤ b.p(x) or q(x) always
denotes a polynomial; pn (x) is a polynomial of degree at most n. In this
course, the goodness (or badness!) of the fit of a particular polynomial
p(x) to the function f (x) will always be measured by
1
2 1. Weierstrass’s Theorem
We can prove that this series, which certainly converges for |u| < 1,
also converges for u = 1. This follows either from Gauss’s test applied
to !
cn 3 1
=1+ +0 2
cn+1 2n n
or by proving (on the lines of the Lemma following Theorem 2) that
cn ∼ n√An .
√
By Abel’s limit theorem, the series for (1 − u) converges uniformly
for 0 ≤ u ≤ 1, i.e., |x| is uniform limit of a sequence of polynomials for
−1 ≤ x ≤ 1.
Corollary. Let
g(x) = 0 for x < 0
g(x) = 0 for 0 ≤ x ≤ k.
Then g(x) is the limit of a uniformly convergent sequence of polyno-
mials in −k ≤ x ≤ k.
Proof. Changing the variable by a factor k, we may suppose that k is 1.
Then
1
2 g(x) = (x + |x|).
2
Proof of theorem 1. Given ε, we can find a function l(x) whose graph
is a polygon with vertices at (a, y0 ), (x1 , y1 ), . . . , (xi , yi ), . . . , (b, yn ) such
that
1
| f (x) − l(x)| < ε.
2
Now l(x) is the sum of constant multiples of functions of the type 4
g(x − xi ) defined in the Corollary, namely,
n−1
X
l(x) = y0 + ci g(x − xi ).
0
For the right hand side is linear in each (xi , xi+1 ), and the ci give the
right value of l(x) at the vertices if
y1 =y0 + c0 (x1 − x0 )
4 1. Weierstrass’s Theorem
······
i−1
X
yi =y0 + c(xi − xk ).
k=0
By the lemma and corollary, we can find a polynomial p(x) such that
1
|l(x) − p(x)| < ε, a ≤ x ≤ b
2
and this gives | f (x) − p(x)| < ε, a ≤ x ≤ b.
1 − (t − x)2 n
5 which (for large n) falls away rapidly from the value 1 as t moves away
form x.We need a theorem about the convergence of singular integrals,
and this is best stated for a general kernel Kn (t − x).
Theorem 2. Let
Z 1
Jn = Kn (u)du
−1
Z δ
Ln (δ) = Kn (u)du (0 < δ < 1)
−δ
2. Singular Integrals and Landau’s Proof 5
Suppose that
(i) Kn (u) ≥ 0
(ii) for each fixed δ, Ln (δ)/Jn → 1, as n → ∞.
Suppose that f (x) is C(0, 1) and 0 < a < b < 1. Then, as n → ∞,
Z1
1
In (x) = Kn (t − x) f (t)dt → f (x)
Jn
0
uniformly for a ≤ x ≤ b.
Proof. In In (x), we shall split up the integral over (0, 1)
(Z x−δ Z x+δ Z 1 )
1
(1) In (x) = + + ,
Jn 0 x−δ x+δ
where 0 < x − δ < x + δ < 1. Consider first the integral over (x − δ, x + δ).
Given ε, we can, by the continuity of f (x), find δ = δ(ε) such that
| f (t) − f (x)| < ε if a ≤ x ≤ b, |t − x| ≤ δ.
Suppose further that δ < min(a, 1 − b). Then the middle term on the
R. H. S. of (1)
Z δ
1
= Kn (u) f (x + u)du
Jn −δ
Z δ
Ln (δ) 1
= f (x) + Kn (u) f (x + u) − f (x) du.
Jn Jn −δ
The first term in the last line tends to f (x), from (ii) of the hypothe- 6
sis. The second term is, by (i), numerically less than εLn (δ)/Jn , that is,
less than ε.
Now return to equation (1) and consider the first term on the R. H.
S. Let M = sup | f (x)|in (0, 1).
Z x−δ
M −δ
Z
1
| Kn (t − x) f (t)dt| ≤ Kn (u)du
Jn 0 Jn −x
6 1. Weierstrass’s Theorem
( )
Ln (δ)
≤ M 1−
Jn
→ 0 as n → ∞.
A similar estimate holds for the third term of (1).
All the inequalities in the above argument are independent of x, and,
collecting the results, we have proved that In (x) → f (x) uniformly for
a ≤ x ≤ b.
If, in Theorem 2, we take, following Landau
Kn (u) = (1 − u2 )n ,
then In (x) is a polynomial in x of degree 2n. We have, therefore, a
second proof of Theorem 1 as soon as we have proved, as we do in the
following Lemma, that this Kn (u) satisfies the conditions of Theorem 2.
Lemma. In Theorem 2, Kn (u) may be taken to be (1 − u2 )n .
Proof.
1
1 2 0π
Z Z
2 n
Jn = (1 − u ) du = 2 0 sin2n+1 θdθ
−1
= 2S 2n+1 , say.
2.4 . . . 2n
S 2n+1 =
3.5 . . . (2n + 1)
7 From the inequalities
S 2n > S 2n+1 > S 2n+2 ,
it is easily proved that
r r
π π
Jn ∼ and Jn > .
n n+1
Then
R1
Ln (δ) 2 δ (1 − u2 )n du
1− =
Jn Jn
r
n+1
< 2(1 − δ2 )n → 0 as n → ∞.
π
3. Bernstein Polynomials 7
3 Bernstein Polynomials
We shall give a third proof of Theorem 1. It has the advantage of em-
bodying a definite construction for the approximating polynomials.
Definition. Write ln,m (x) = (nm )xm (1 − x)n−m , 0 ≤ m ≤ n. The nth Bern-
stein polynomials of f (x) in (0, 1) is defined to be
n
X
Bn (x) = Bn ( f ; x) = f (m/n)ln,m (x).
m=0
Put ey = x and we have the first result. Differentiate with regard to y and
put ey = x and we have second. Differentiating again gives
X
nx + n(n − 1)x2 = m2 ln,m (x)
Multiply the three equations in turn by n2 x2 , −2x, 1 and add. This gives
the third result in the lemma.
8 1. Weierstrass’s Theorem
1
for x2 , B2 (x) is x(1 + x)
2
1
for x(1 − x), B2 (x) is x(1 − x).
2
3. Bernstein Polynomials 9
Notes at the end of a chapter may include exercises (with hints for 10
solutions), extensions of the theorem and suggestions for further read-
ing.
Hints for 1 − 4
3. Could use
R 1R 1n on n on
0 0
1 − (t − x)2 1 − (u − y)2 f (t, u)dtdu
R 2
1 2 )n dt
−1
(1 − t
4. Lorentz, 26.
For Landau, with Kn (u) = (1 − u2 )n in Theorem 2,
Z 1 Z 1 Z 1
d ∂Kn ∂Kn
Kn (t − x) f (t)dt = f (t)dt = − f (t)dt
dx 0 0 ∂x 0 ∂t
and integrate by parts.
Chapter 2
The Polynomial of Best
Approximation Chebyshev
Polynomials
x0 , x1 , . . . , xn
and n + 1 constants c0 , c1 , . . . , cn .
Q
Write (x) = (x − x0 ) · · · (x − xn ).
The polynomial p(x) of degree at most n which takes the values ci
Q
at xi is, by the partial- fraction rule for p(x)/ (x),
n
Y X 1 ci
(x) Q .
0
x − xi (xi )
11
12 2. The Polynomial of Best Approximation...
|pα (xi )| ≤ A, α in I; i = 0, . . . , n.
|aα,r | ≤ AB,
where B is independent of α.
Proof. Given ε, we have two polynomials say pα (x), qα (x) which dif-
fer by at most ε for each of the values x0 , . . . , xn . By Lemma 1, their
corresponding coefficients differ by at most Bε.
5 Best Approximation
Let Pn be the set of polynomials p(x) of degree less than or equal to n.
Then
P0 ⊂ P1 ⊂ P2 · · ·
Define, for any particular p in Pn ,
If all the ε′ s have the same sign, say +, add a small constant to p(x).
This gives a polynomial of better approximation.
Generally, suppose that there are k changes of sign in the sequence
of ε′ s, where k ≤ n. Let εi , εi+1 be different. Then li and li+1 cannot
abut (since g(x) does not vanish in either), so we can choose a value of
x; lying between them. We have thus k values of x; call them
x1 , x2 , . . . , xk .
Define h(x) = ε1 (x1 − x)(x2 − x) · · · (xk − x).h(x) has the same sign as
g(x) in each of sub-intervals l. We shall prove that, if η is small enough,
the polynomial of Pn
p(x) + ηh(x)
has better approximation to f (x) than p(x) has.
In those intervals of the original subdivision which are not l′ s,
| f − p − ηh| = |g − ηh|.
d ≥ min di
Proof. Suppose that d < di (i = 1, . . . , n + 2) and let p be the best Pn .
Then p − q = ( f − q) − ( f − p) takes alternate signs at the n + 2 values in
the hypothesis. Therefore p − q (which is in Pn ) has at least n + 1 zeros.
This is a contradiction.
Corollary. Let q be in Pn and let
sup | f − q| = d′ .
Suppose that f − q takes the values ±d′ alternately for n + 2 values
of x. Then d′ = d and q is the best Pn .
Proof. By theorem 6, d ≥ d′ . But d ≤ d′ , since d is the best approxima-
tion.
16 2. The Polynomial of Best Approximation...
6 Chebyshev polynomials
Theorem 4 guarantees the existence of the best Pn for a given f . It is
only in special cases that the explicit calculation of this polynomial is
practicable. Theorem 7 and its corollary can often be turned to use.
Easy exercises:
1 1
For x2 in (0, 1), the best P0 is , the best P1 is x + .
2 8
1
For x4 in (−1, 1), the best P3 is x2 + .
8
Consider now the general problem.
A. Among all pn (x) with coefficient of xn equal to 1, find that which de-
viates least from 0 in (−1, 1) in other words, that for which
sup |pn (x)| = d is least.
This problem can be stated in the equivalent form.
d2 − p2 and (1 − x2 )p′2
n2 (d2 − p2 ) = (1 − x2 )p′2 .
p = d cos(nθ + C)
(1 − x2 )y′′ − xy′ + n2 y = 0.
18 2. The Polynomial of Best Approximation...
Note
7
We now take up the central problem of polynomial approximation, 21
namely
Given a function f (x), how high is the degree of the polynomial
necessary to approach it with an assigned accuracy?
The answer may well depend on structural properties of f (x). For
instance, we may guess (rightly) that we can predict a lower degree if
f (x) is assumed to be differentiable instead of only continuous. The
best theorems on these matter lie fairly deep. We shall go through some
heuristic motion of finding from particular cases what truth appears to
be and then deciding how to try to establish it.
A useful function to study with care is |x| in (−1, 1). This function
was the basis of Lebesgue’s proof of Theorem 1.
From Exercises 1, 2 at the end of Chapter I, the deviation from |x|
of a polynomial of degree n of any of the three types used in proving
√
Theorem 1 is of order 1/ n.
Let us clarify our mode of speech. If, for some p(x) in Pn ,
| f (x) − p(x)| = 0{ϕ(n)}
we will say that the approximation is 0{ϕ(n)}. If, moreover, there is no
p(x) in Pn for which
| f (x) − p(x)| = 0{ϕ(n)}
19
20 3. Approximations to |x|
a0 + a1 x2 + · · · + an x2n
8 Oscillating polynomials
Definition. If 0 ≤ α0 < α1 · · · < αn and Ai , 0 (all i), we say that
Corollary. p(x) takes the values ± sup |p(x)| with + and − sign alter-
nately.
n
Ai xαi is an oscillating polynomial in (0, 1).q(x)
P
Theorem 10. p(x) =
i=0
is another polynomial Bi xαi with the same exponents. One coefficient
P
of p is the same as the corresponding one of q (say A j = B j), where
α j > 0. Then
sup |q| > sup |p|.
Proof. If not, p − q takes alternate signs (may be 0) for the n + 1 values 24
of x for which p takes its numerically greatest value. Therefore p − q
has at least n zeros in 0 ≤ x ≤ 1. But, since A j = B j it has only n terms,
22 3. Approximations to |x|
Converse of Theorem 10. p(x) and q(x) are two polynomials with the
same exponents and one coefficient the same (A j = B j, where α j > 0).
If
sup |p| < sup |q|
for every such q, then p is an oscillating polynomial.
Proof. We gives the gist of the proof, without setting out all the detail
in full. It uses a ’deformation’ argument like that of theorem 5.
We need to prove that among all the polynomials with the given
exponents
there is a unique q(x) for which sup |q(x)| attains its lower bound.
0≤x≤1
Clearly sup |q(x)| is a continuous function of the n variables (B0 · · · ,
Bk−1 , Bk+1 , . . . , Bn ). Its lower bound is greater than 0 by Corollary 1 of
Theorem 10. It is less than or equal to K, as is seen by taking the B′ s to
be small. Again by Corollary 1, we need only consider values of Bi for
which
|Bi | ≤ K|ti | (i = 0, 1, . . . , k − 1, k + 1, . . . , n),
√
where ti is the coefficient of xαi in T 2αn ( x).
The Bi lie in a bounded closed region of n space, and so they have at
least one set of values for which sup |q(x)| attains its lower bound. This
proves the existence of an oscillating polynomial. Uniqueness follows
from Theorem 10.
Theorem 12. If
q(x) − p(x) = 0
Then q(x) − p(x) has the sign of q(x) (it may be 0) for the values
xn (k = 1, . . . , n + 1) at which |q(x)| takes its maximum value. Therefore
q(x) − p(x) vanishes for n values ξ1 , . . . , ξn such that
xi ≤ ξ1 ≤ x2 ≤ ξ2 ≤ · · · ≤ ξn ≤ xn+1 .
Theorem 12 (Extension). If
0 ≤ α0 < β0 < · · · < αi−1 < βi−1 < m < βi+1 < αi+1 · · · < βn < αn ,
then
sup |p(x)| > sup |q(x)|.
9 Approximation to |x|
Theorem 13. If
p(x) = x + a1 x2 + a2 x4 + · · · + an x2n
x + b1 x3 + · · ·
Then, if µ > 0,
!2 !2n
x + a x
· · · + an
x
≤ m.
1 + µ 1
1+µ 1+µ
Therefore
Therefore
n o
|µ(x + c2 x4 + · · · + cn x2n )| ≤ m (1 + µ)2 + 1
n o
and so |x + c2 x4 + · · · + cn x2n | ≤ m (1 + µ)2 + 1 /µ
29 This is true for all positive values of µ, and so √ we can replace the
right-hand side by its minimum, which is 2m(1 + 2).
As we said, the maximum modulus of a polynomial with exponents
1, 4, 6, . . . , 2n is greater than 1/(2n − 1) and therefore
1 1
m> √ .
2(1 + 2) 2n − 1
We have now all the material for the final result announced at the
end of §7.
a0 + a1 x2 + · · · + an x2n ,
1 1 1
then > d2n > √ · .
2n + 1 4(1 + 2) 2n − 1
A0 + x + A1 x2 + A2 x4 + · · · + An x2n .
sup |q(x) − A0 |
Notes
A0 ϕ0 + · · · + An ϕn (x),
Hint
√
1. If k, n are both even or both odd, consider T n (x), otherwise T 2n ( x).
Chapter 4
Trigonometric Polynomials
29
30 4. Trigonometric Polynomials
Illustrations:
1) If f (x) = tn−1 (x) + (an cos nx + bn sin nx). then tn−1 (x) gives the
best approximation in T n−1 to f (x).
p
Proof. f − tn−1 takes the values ± (a2n + b2n ) alternately at 2n points.
where 0 < a < 1, b is an odd integer and ab > 1. We shall prove that
the best approximation in T n to f (x) is
k
X
t(x) = ar cos br x, where bk ≤ n < bk+1 .
r=0
P∞
Proof. f (x) − t(x) = k+1 ar cos br x.
∞
ar at x = 0. Cos bk+1 x takes the
P
This takes its greatest value
k+1
values ±1 alternately at integral multiples of π/bk+1 , of which there are
2bk+1 in a period.
Since b is an odd integer, cos br x for r > k + 1 takes the same values
at those points as cos bk+1 x.
Now 2bk+1 ≥ 2n + 2 and so f (x) − t(x) takes its numerically greatest
value for at least 2n + 2 values ofx.
Corollary. The approximation given by this t(x) is A/nα , where α =
log(1/a)/ log b.
Proof. The approximation is
ak+1 b−α(k+1) 1 1
= ∼ · α
1−a 1−a 1−a n
32 4. Trigonometric Polynomials
(2) | f (x) − S n (x)| < M(A log n + B), where M = sup | f (x)| and A, B are
constants.
1 P∞ 1
(4) 2
= −∞ (t , kπ).
sin t (t + kπ)2
The result (2), which cannot be improved, shows that, in general,
the Fourier series of a function gives a poor approximation in the sense
measured by sup | f (x) − S n (x)|. As the R.H.S. of (2) tends to infinity
with n, (2) does not include Weierstrass’s Theorem 15. The sense in
which the Fourier series does give the best approximation is the mean-
square sense (omitted here). It is known that the Fejer sums σn of (3)
behave more regularly than the Fourier sums S n ; this is due to the kernel
(sin nt/ sin t)2 in the integral for σn being positive, whereas the kernel in
S n takes both signs. The next theorem gives the approximation to f (x)
afforded by (σn (x).
34 4. Trigonometric Polynomials
Proof. We first put σn (x) into a form more convenient than that of The-
orem 17(3). Since f is periodic
Z π
sin2 nt
Z (k+1)π
sin2 nt
f (x + 2t) dt = f (x + 2t) dt.
0 (t + kπ)2 kπ (t2 )
1 ∞ 2t sin2 t
Z
σn = f (x + ) 2 dt
π −∞ n t
) 2
1 ∞
Z (
2t sin t
σn − f = f (x + ) − f (x) dt.
π −∞ n t2
Therefore
∞
sin2 t
Z
2
| f − σn | ≤ ω(2t/n) dt.
π 0 t2
The integral on the R.H.S. is the sum of the integrals over (0, 1),
(1, X) and (X, ∞). This gives
( Z X Z ∞ )
2 dt dt
| f − σn | ≤ ω(2/n) + ω(2t/n) 2 + ω(π) 2
π 1 t X t
( Z X )
2 t+1 ω(π)
≤ ω(2/n) + ω(2/n) 2
dt +
π 1 t X
( )
2 ω(π)
≤ ω(2/n)(2 + log X) + .
π X
Choose X = 1/ω(2/n) and we have a result equivalent to that stated.
11. Fourier and Fejer Sums 35
37 Corollary 1. Theorem 15
AB
Corollary 2. If ω(δ) < Aδα (0 < α < 1), then | f − σn | < nα , where
B = B(α) is independent of f .
Proof.
2 ∞ sin2 t
Z
| f − σn | ≤ ω(2t/n) 2 dt
π 0 t
α+1 Z ∞ 2
2 A α sin t
≤ t dt
πnα 0 t2
The estimates in Chapter III would lead us to suspect that, if we can
find a t(x) which approximates to f (x) more closely than σn (x) does, we
may get rid of the log ω(1/n) on the R.H.S. of Theorem 18. It is easy to
see how to try to do this. The logarithm arises from integrating a term
in 1/t. The Fejer sum is
Z ∞ ! !
1 2t sin t
Fr (x, n) = f x+ 2r dt
Jr −∞ n t
for r = 1 and Jr = π. If r ≥ 2, there will be no term in 1/t. We shall
achieve our purpose by taking r = 2.
R∞
sin t 4
2π
Lemma 1. (1) J2 = −∞ t dt = 3 .
(2) Reversing the steps by which F1 (x, n) was obtained in the first part
of the proof of Theorem 18. we have
Zπ
d2
!
3 3 4 1
F2 (x, n) = f (x + 2t) sin nt 2 dt.
2π 6n dt sin2 t
0
d2
! !
4 1 4 6 4
38 Then sin nt 2 = sin nt − .
dt sin2 t sin4 t sin2 t
sin nt
Now is the sum of multiples of cos kt where k ≤ n − 1. Hence
sin t !
d 2 1
sin4 nt 2 is the sum of multiples of cos kt where k ≤ 4n − 2.
dt sin2 t
Moreover, the expression is even and has period π, so k takes only even
values, 21 say, where 1 ≤ 2n − 1. Finally,
Z π
1 2π
Z
f (x + 2n) cos 21tdt = f (u) cos l(u − x)du
0 2 0
and F2 (x, n) is in T 2n−1 .
Theorem 19. | f (x) − F2 (x, n)| ≤ 3ω(1/n).
3 R∞ 2t sin t !4
Proof. F2 (x, n) − f (x) = f (x + ) − f (x) dt.
2π −∞ n t
!
2t 2|t|
Now | f (x + ) − f (x)| ≤ ω ≤ (2|t| + 1)ω(1/n) from Theorem
n n
16 (2). Therefore
!4
3 ∞
Z
sin t
| f (x) − F2 (x, n)| ≤ ω(1/n) (2t + 1) dt = Aω(1/n),
π 0 t
6 R ∞ sin4 t
where A = 1 + dt. But
π 0 t3
1 π sin3 t
Z ∞ 4
sin3 t
Z ∞ Z
sin t
dt | 2 |dt = dt = 1
0 t3 0 t 2 0 sin2 t
(again by use of Theorem 17(4)).
6
So A < 1 + < 3.
π
11. Fourier and Fejer Sums 37
A
| f (x) − F2 (x, n)| ≤ ω1 (1/n) where A < 5/2.
n
!4
3 R∞
( ! ! )
2t 2t sin t
Proof. F2 (x, n) − f (x) = f x+ + f x− − 2 f (x) dt.
2π 0 n n t
The modulus of the term within { } is 39
2 t ′
Z ( ! !)
2u ′ 2u
| f x+ − f x− du|
n 0 n n
Zt !
2 4u
≤ ω1 du
n n
0
! Zt
2 1
≤ ω1 (4u + 1)du
n n
0
!
2 1
= ω1 (2t2 + t)
n n
Therefore
!
A 1
| f (x) − F2 (x, n)| ≤ ω1 ,
n n
where
Z∞ !4 Z∞
3 sin t 3 2 3 sin4 t 3 π 5
A= (2t2 + t) dt = sin tdt + 3
dt < +1 <
π t π π t π 2 2
0 0
Notes on Chapter IV 40
38 4. Trigonometric Polynomials
4. With the notation of Illustration (2), Corollary, page 31, prove that
Weierstrass’s function ar cos br x satisfies a Lipschitz condition of
P
order α.
41 Hints
t′ (x), being continuous, attains its bounds and so, for some c, t′ (a) =
±nl and we will suppose that
t′ (c) = nl.
t′′ (c) = 0.
Define S (x) = l sin n(x − c) − t(x).
Then r(x) = S (x) = nl cos n(x − c) − t′ (x).
′
41
42 5. Inequalities, etc.
Then
Proof. Put
t(θ) = p(cos θ)
t (θ) = −p′ (cos θ) sin(θ).
′
The bound for |p′ (x)| given in Corollary 2 fails at the end-points ±1. A 44
better result, due to Markoff, is
Lemma 1. Let
(2k − 1)π
xk = cos (k = 1, 2, . . . , n)
2n
be the zeros of the Chebyschev polynomial T n (x). If q(x) is in Pn−1 , then
n
1X T n (x)
q
q(x) = (−1)k−1 (1 − x2k )q(xk ) · .
n k=1 x − xk
Proof. Both sides are in Pn−1 and so it is sufficient to show that they
agree for the n values xk . As x → xk ,
T n (x) n
→ T n′ (xk ) = q sin(n arc cos xk )
x − xk
(1 − x2k )
n(−1)k−1
= q
(1 − x2k )
Also, for x = xk , every term on the R.H.S. except the k-th vanishes.
1
Lemma 2. Suppose that q(x) is in Pn−1 and |q(x)| ≤ p (−1 <
(1 − x2 )
x < 1).
π 1
p q
(1 − x2 ) ≥ (1 − x21 ) = sin ≥ .
2n n
1 X T n (x)
|q(x)| ≤ | |,
n x − xk
1
Therefore |q(x)| ≤ |T n′ (x)|.
n
′ n sin nθ
But, if x = cos θ, T n (x) = , which gives
sin θ
|T n′ (x)| ≤ n2 .
p′ (x)
q(x) = .
Mn
Corollary. p(x) = T n (x) shows that the result is the best possible.
13. Structural Properties Depend on the... 45
Theorem 23. Let f (x) be C(2π). Suppose that, for all n, the best ap-
proximation in T n to f (x) is less than A/nα , where 0 < α < 1. Then f (x)
is Lip.α.
A
| f (x) − tn (x)| ≤ .
nα
Define u0 (x) = t1 (x)
an (x)t2n (x) − t2n−1 (x) (n ≥ 1).
∞
P
Then f (x) is the sum of the uniformly convergent series un (x). 46
0
1
Choose δ with 0 < δ ≤ , and m such that
2
1
2m−1 ≤ < 2m .
δ
Suppose |x − y| ≤ δ. We have
m−1
X ∞
X ∞
X
| f (x) − f (y)| ≤ |un (x) − un (y)| + |un (x)| + |un (y)|.
0 m m
Therefore
∞ ∞
X X
α 1 A(1 + 2α ) 1
|un (x)| ≤ A(1 + 2 ) = .
m m
2nα 1 − 2−α 2mα
This gives
m−1
X B
| f (x) − f (y)| ≤ |un (x) − un (y)| + .
o
2mα
Therefore
m−1
X B
| f (x) − f (y)| ≤ A(1 + 2α )δ 2n(1−α) + .
o
2mα
1
Putting C = A(1 + 2α ) and using < δ, we have
2m
m−1
X
ω(δ) ≤ Cδ 2n(1−α) + Bδα.
o
47 If now α < 1,
m−1
X 2m(1−α) − 1 2m(1−α)
2n(1−α) = < .
o
21−α − 1 21−α − 1
2
Use now 2m ≤ and we find
δ
21−α
!
ω(δ) < 1−α C + B δα
2 −1
See Notes 1-4 at the end of the Chapter.
14. Divergence of the Lagrange Sequence 47
Lemma. Let p(x) be the Lagrange polynomial which takes the value 0
at the 2m values of x
k/m (−m ≤ k ≤ m; k , 2)
!
1
and takes the value 1/m when x = 2/m. Then, if m is odd, |p |→∞
2
as m → ∞.
1 (x + 1)(x + m−1
m ) . . . x(x − 1/m)(x − 3/m) . . . (x − 1)
m−1
m (2/m + 1)(2/m + m ) . . . 2/m(2/m − 1/m)(2/m − 3/m) . . . (2/m − 1)
!
1
This gives for |p | 48
2
3.54 . . . m
where A=
2.4 . . . (m − 1)
(m + 2)(m + 4) . . . (2m + 1)
B=
(m + 1)(m + 3) . . . 2m
! ! !
2m + 3 3m + 5 3m
C= − ... .
m+1 m+3 2m − 2
Here A > 1, B > 1, and the factors of C decrease from left to right,
the last being greater than 3/2. So C > (3/2)m−1 .
1
Note.x = has been taken for ease of calculation. The conclusion holds
2
for other values of x.
Theorem 24 (Borel). There is f (x) in C(−1, 1) whose nth Lagrange
polynomial does not converge to f (x) as n → ∞.
Proof. Define a continuous curve Ck which coincides with Ox outside
the interval (3−k−1 , 3−k ) and has maximum 3−k−1 at the midpoint of that
interval. For example we can define Ck by
n o
y = 3−k−1 sin (3k+1 x − 1)π/2 .
We shall use the Ck to construct a curve S .Pk,S (x) will denote the
Lagrange polynomial which takes the same values as S for the values
x = 1/3k , where −3k ≤ 1 (integer) ≤ 3k . !
1
49 We shall construct S so that Pk,S does not converge to the point
2
1
on S where x = . Observe first that Pk,Ck−1 is the Lagrange polynomial
2
in the Lemma with m = 3k . From the Lemma, given A, there is h1 such
that !
1
|Pk,Ck−1 | > 2A if k − 1 > h1 .
2
There are two possibilities:
!
1
(a) With h1 fixed, Pk,Ch1 does not tend to 0 as k → ∞. Then S can
2
be taken to be Ch1 ; or
14. Divergence of the Lagrange Sequence 49
5. For further ‘negative results’ like Theorem 24, see Natanson, 369
– 388. For example, the Lagrange polynomial taking the values of
|x| at n equally spaced points in (−1, 1) converges to |x| as n → ∞
for no value of x expect 0, ±1.
Chapter 6
Approximation in Terms of
Differences
15
This is the only chapter in the course, of which the results are not clas- 52
sical. The point of view here might lead to a re-orientation towards
algebraic rather than trigonometric polynomials.
In Theorem 20 (and its known extensions) the approximation attain-
able in Pn or T n to a differentiable function f (x) is expressed in terms of
its first higher derivative. We shall now give simple examples which lead
us to suppose that bounds of differences of f (x) rather than derivatives
may be more directly related to the closeness of the approximation.
Example 1. If f (x) attains its greatest value at x2 and its least at x1 , then
the best approximation in P0 is
1
{ f (x1 ) + f (x2 )}
2
1 1
and do = { f (x2 ) − f (x1 )} = sup |∆ f |
2 2
51
52 6. Approximation in Terms of Differences
Example 2. If f (x) is C(0, 1), there is a linear function p(x) for which
0 ≤ x ≤ x + 2h ≤ 1.
1
If < x1 ≤ 1 take x = 2x1 − 1, h = 1 − x1 . Then
2
|g(x + 2h) − 2g(x + h) + g(x)| = |g(1) − 2g(x1 ) + g(1 − x1 )| ≥ M.
Hn = 2n /T n ,
54 6. Approximation in Terms of Differences
n
X
where T n = T n (ho , h1 , . . . , hn ) = |ϕ′ (hi )|−1 .
i=0
by definition of Hn .
Therefore, for all x in (−1, 1),
Suppose, then, that sup |∆n | = L and that the bound L is assumed
for the values ho , h1 , . . . , hn of the independent variable. Define points
(hi , yi ) for i = 0, 1, . . . , by taking
L
yi = f (hi ) − (−1)k ,
2n
where k is i or i + 1 according as ∆n ( f, ho , . . . , hn ) is positive or negative.
Construct a p(x) of Pn−1 through the n points (hi , yi ) for i = 0, 1, . . . ,
n − 1. Write
g(x) = f (x) − p(x).
Since ∆n (p) ≡ 0, |∆n (g)| = |∆n ( f )| attains its upper bound for h0 , h1 ,
. . . , hn . From the definition of ∆n , the value of g(hn ) which makes
|∆n (g, h0 , h1 , . . . , hn )| = L is (−1)k L/2n , where k is n or n + 1 accord-
ing as ∆n ( f, ho , . . . , hn ) is positive or negative.
We prove that |g(x)| ≤ 2−n L for all x. Suppose that |g| takes values
greater than L/2n , say g(h′i ) > L/2n for a value h′i between hi−1 and hi+1
where g(hi ) = 2−n L. Then, from the definition of T n ,
T n (ho , . . . , hn ) ≥ 2n−1
and the sign = holds if and only if the hi are the Chebyshev points.
Proof. The polynomial qn (x) of degree n which takes the value (−1)i at
hi (i = 0, . . . , n) is
n
X (−1)i
qn (x) = ϕ(x) .
i=0
ϕ′ (hi ) (x − hi )
57
APPENDIX
Approximation by
Polynomials in the Complex
Domain
1 Runge’s Theorem
The problem considered till now was the approximation of a given con- 59
tinuous function on a finite closed interval by polynomials in a real vari-
able. Even for functions of two variables, we considered only the prob-
lem of approximation by polynomials in two independent real variables
x, y. in what follows, we shall consider the approximation of a function
in a domain in the plane (open connected set) by polynomials in the
complex variable z = x + iy (which are analytic functions of the variable
z).
Let pn (z) be a sequence of polynomials and suppose that G (which
we assume is not empty) is the largest open set in which pn (z) converges,
uniformly on every compact subset. (This is the type of approximation
we shall consider; the problem of approximation on closed sets more
difficult). By Weierstrass’s theorem the limit of the sequence pn (z) is an
analytic function in G. There is moreover, a purely topological restric-
tion on G, viz., every connected component D of G is simply connected:
for, if C is a sample closed curve contained in D and B is its interior, then
59
60 BIBLIOGRAPHY
The main idea in the proof of the theorem is contained in the next 61
step, which we state as a separate lemma.
Lemma. Let C be a simple are joining the points zo and z1 and K a
compact set not meeting C. Then, any rational function whose only
possible pole is at zo can be approximated, uniformly on K, by rational
functions which no poles expect possibly at z1 .
Proof of the lemma. Let ε > 0 be given and 2d be the distance
between C and K. We find points ao , . . . , aµ on C, ao = zo , aµ = z1 such
that |ak+1 − ak | ≤ d. Let R(z) be the given rational function. There are
two polynomials p and q so that
!
1
R(z) = p(z) + q
z − zo
!
1
and we have only to approximate q = f (z). Since f (z) is ana-
z − zo
lytic in |z−a1 | > d and is finite at ∞, the Laurent expansion of f (z) about
a1 contains no positive powers of z − a1 and converges uniformly on ev-
ery compact subset of |z − a1 | > d. A suitable! partial sum then gives us
1 ε
a polynomial p1 with f (z) − p1 < for z ∈ K. Repeating
z − a1 µ+1
this process, we find successively, polynomials p j , j = 1, . . . , µ with
! !
1 1 ε
p
j − p j+1 < for z ∈ K.
z − aj z − a j+1 µ+1
Then clearly
!
1
f (z) − p
µ < ε, z ∈ K
z − aµ
and the lemma follows.
Proof of Theorem. Let D be any domain and f analytic in D. Let Gn
be the sequence of regions exhausting D described above. There is a 62
rational function rn with poles at most on Cn+1 such that
f (z) − r (z) < 1 on G .
n 2n n
62 BIBLIOGRAPHY
the sum being over those (finitely many) aν which lie in Gn . Since fn+2 −
fn+1 is analytic in Gn+1 , we can find a rational function Rn+1 with no
poles in D such that
f (z) − f (z) − R (z) < 1 for z in G .
n+2 n+1 n+1
2n
n
Since the poles of the Rn+1 lie outside D and the series
∞
X
( fn+1 − fn − Rn )
n=no +1
∞
X
f (z) = f2 (z) + ( fn+1 (z) − fn (z) − Rn (z)).
n=2
2 Interpolation
For functions f of a real variable, if pn is the (Lagrange) polynomial p
of degree n which agrees with f at n + 1 equally spaces points on an
interval, the sequence pn in general diverges as n → ∞. The behaviour
for functions of a complex variable is more satisfactory. We proceed to
prove two of the main theorems.
64 BIBLIOGRAPHY
THEOREM B. Let f (z) be analytic for |z| < R, R > 1 and pn (z) the
polynomial of degree n with pn (zi ) = f (zi ), i = 0, 1, . . . , n, where the zi
are the (n + 1)th roots of unity.
Then
pn (z) → f (z) as n → ∞
uniformly for |z| ≤ ̺ < R.
65 Proof. Let C be the circle |z| = ρ′ , ρ′ > ρ, ρ > 1. Then, for |z| ≤ ρ,
Z n+1
1 z − 1 f (t)
f (z) − pn (z) = dt,
2πi c tn+1 − 1 t − z
so that
zn+1 − 1 f (t)
Z
1
| f (z) − pn (z)| = | dt|
2πc tn+1 − 1 t − z
1 + ρn+1
≤ ′n+1 M (M = sup | f (z)|)
(ρ − 1)(ρ′ − ρ) z∈C
→ 0 as n → ∞ since ρ′ > ρ, ρ′ > 1.
1 Pn
z(w)−z(wi ) 1
R2π z(w)−z(eiθ )
2. lim n+1 i=o log w−wi = 2π log w−eiθ dθ
n→∞
o
and the limit is uniform for w in a compact set in |w| > 1.
z(w) − z(ζ)
Now is an analytic function of ζ for |ζ| > 1 and fixed w,
w−ζ
including ζ = ∞, ζ = w. Hence the integral in (2) is equal to
z(w) − z(ζ)
lim log = log A, say
ζ→∞ w−ζ
(we have only to make the substitution ζ → 1/ζ and use the Poison
integral).
From (1), it follows that
n
Q 1
n 1 |z(w) − z(wi )| (n+1)
Y |z(w) − z(wi )| (n+1)
i=0
lim = lim
n→∞
i=o
|w − wi |1/(n+1) n→∞ |wn+1 − 1|1/(n+1)
66 BIBLIOGRAPHY
n
Q 1
lim |z(w) − z(wi ) (n+1)
n→∞ i=0
= =A
|w|
67 Proof of Theorem. Let CR (R > 1) denote the image under z(w) of the
circle |w| = R; we can choose R > 1 such that f is analytic inside and on
CR . Let 1 < r1 < r2 < R; and put
n
Y
πn (z) = (z − α(n)
i ).
i=o
We have Z
1 πn (z) f (t)
f (z) − pn (z) = dt.
2πi πn (t) t − z
Cr2
3 Best Approximation
In this section we shall consider the problem of best approximation.
68 Let K be a compact set containing infinitely many points and f (z) a
continuous function on K. Our aim is to prove the existence and unique-
ness of a polynomial pn (z) of degree n such that
d( f, pn ) = sup f (z) − pn (z)
z∈K
3. Best Approximation 67
d ≤ d( f, p) ≤ d( f, p(νk ) ) + d(p(νk ) , p) → d
so that d( f, p) = d.
[If K contains a circle |z − a| < r, r > o, the existence of a sequence
p(νk ) converging uniformly on any compact set follows at once, as in the
case (Ch.II, T h.4, p.14) if we use the Cauchy inequalities.]
68 BIBLIOGRAPHY
Let 1 > η > o be sufficiently small. Consider g(z) − ηq(z); then for
|z − zi | < δ, |g(z) − g(zi )| < ε, |ηq(z) − ηq(zi )| < ηε, so that |g(z) − ηq(z)| =
|η(g(zi ) − g(z)) + ηg(z) − g(z) + η(q(z) − q(zi ))| (since q(zi ) = g(zi )) <
ηε + ηε + (1 − η)d < d if 2ε < d
If we choose η so small that
Then
1 1
| f (z) − r(z)| = | ( f (z) − p(z)) + ( f (z) − q(z))| ≤ d
2 2
Let z1 , . . . , zn+2 be points at which | f (zi ) − r(zi )| = d. Then, unless 71
1
f (zi ) − p(zi ) = f (zi ) − q(zi ) = wi with |wi | = d, | f (zi ) − p(zi ) + f (zi ) −
2
q(zi )| < d. Hence, p(z) and q(z) take the same value at the n + 2 points
zi and since they are polynomials of degree n, p(z) = q(z).
Bibliography
71