• No results found

Orthogonal polynomials and special functions

N/A
N/A
Protected

Academic year: 2022

Share "Orthogonal polynomials and special functions"

Copied!
34
0
0

Loading.... (view fulltext now)

Full text

(1)

U.U.D.M. Project Report 2020:29

Examensarbete i matematik, 15 hp Handledare: Jörgen Östensson

Examinator: Veronica Crispin Quinonez Juni 2020

Department of Mathematics Uppsala University

Orthogonal polynomials and special functions

Elsa Graneland

(2)
(3)

Abstract

This thesis is a brief introduction to the theory of orthogonal polynomials which is based on the theory of inner product spaces. We will also consider some special functions, especially the Bessel function.

We present definitions of orthonormal and monic orthogonal polynomials, and discuss the three-term recurrence relation, the Jacobi matrix and also a result concerning the zeros of the orthogonal polynomials.

Furthermore the Sturm-Liouville problem is presented and in particular, the Bessel function. Many polynomials and special functions are solutions to differential equations describing phenomena in nature. Lastly some applications to physics, e.g. quantum mechanics, are being presented.

(4)

Contents

1 Introduction 4

2 Inner product and inner product space 4

3 Gram-Schmidt process 6

4 Orthogonal polynomials 8

4.1 Orthonormal and monic orthogonal polynomials . . . 8

4.2 Three-term recurrence relation . . . 10

4.3 The Jacobi matrix . . . 12

4.4 Zeros of orthogonal polynomials . . . 13

5 Classical orthogonal polynomials 16 5.1 Legendre polynomials . . . 16

5.2 Hermite polynomials . . . 18

5.3 Laguerre polynomials . . . 20

6 Sturm-Liouville problems 21 7 Applications 24 7.1 Eigenvalues of the Laplace operator . . . 24

7.1.1 Eigenvalue problem in a disk . . . 25

7.1.2 Eigenvalue problem in a ball . . . 26

7.2 Schr¨odinger equation . . . 28

7.2.1 Harmonic oscillator . . . 29

7.2.2 Hydrogen atom . . . 30

8 References 33

(5)

1 Introduction

The theory of orthogonal polynomials and special functions is of intrinsic interest to many parts of mathematics. Moreover, it can be used to explain many physical and chemical phenomena. For example, the vibrations of a drum head can be explained in terms of special functions known as Bessel functions. And the solutions of the

Schr¨odinger equation for a harmonic oscillator can be described using orthogonal polynomials known as Hermite polynomials. Furthermore, the eigenfunctions for the Schr¨odinger operator associated with the hydrogen atom are described in terms orthogonal polynomials known as Laguerre polynomials.

In Section 2 the definitions of inner product and inner product space are presented, and in Section 3 the Gram-Schmidt algorithm is described. In Section 4 the orthonormal and monic orthogonal polynomials are defined. This is followed by a discussion about the three-term recurrence relation, the Jacobi matrix and the zeros of orthogonal polynomials. This thesis considers three of the classical orthogonal polynomials, which are presented in Section 5. In Section 6 we discuss Sturm-Liouville problems. The last section in this thesis, Section 7, present the applications mentioned above.

2 Inner product and inner product space

The standard scalar products of Rn and Cn satisfy some calculation rules. These are taken as axioms in a general inner product space.

Definition 2.1. The inner product h · , · i on a vector space X is a mapping of X × X in to the scalar field K (= R or C) satisfying the following. For all vectors x, y and z and scalars α we have:

1. hx + y, zi = h x, z i + h y, z i.

2. h α x, y i = α h x, y i.

3. h x, y i = h y, x i.

4. h x, x i ≥ 0 and equal if and only if x = 0.

The inner product on X defines a norm on X given by kxk =phx, xi and a distance function, or metric, on X given by

d(x, y) = kx − yk =phx − y, x − yi.

When an inner product space is complete, the space is called a Hilbert space and is usually denoted by H.

Remark. It follows from the definition above, that

h αx + βy, z i = αh x, z i + βh y, z i.

h x, αy + βz i = αh x, y i + βh x, z i.

Due to the conjugate in the second variable, one says that the inner product is sesquilinear, meaning ”112 times linear”.

(6)

Definition 2.2. Two elements x and y in an inner product space X are said to be orthogonal if h x, y i = 0. A set of vectors is called a orthonormal set if these vectors are pairwise orthogonal and of norm 1.

Example 1. Euclidean space Rn.

Given vectors x = (x1, x2, ..., xn) and y = (y1, y2, ..., yn) in Rn an inner product is defined by

h x, y i = x1y1+ x2y2+ ... + xnyn=

n

X

j=1

xjyj. (2.1)

This makes Rn into a Hilbert space.

Example 2. Unitary space Cn.

The standard inner product on Cn is given by

h x, y i =

n

X

j=1

xjyj. (2.2)

This makes Cn into a Hilbert space.

Example 3. The space C[a, b].

Let C[a, b] denote the space complex-valued continuous functions defined on the interval [a, b]. An inner product on C[a, b] is given by

h f, g i = Z b

a

f (x) g(x) dx. (2.3)

This space is not complete in the metric induced by the scalar product.

Example 4. The space L2(R, dµ).

Given a positive Borel measure dµ on R, we let L2(R, dµ) denote the equivalent classes of all Borel measurable functions such that

Z

R

|f |2dµ < ∞.

Two functions are considered equivalent if they agree µ-almost everywhere. An inner product on L2[R, dµ] is given by

h f, g i = Z

R

f (x) g(x) dµ. (2.4)

This makes L2(R, dµ) into a Hilbert space.

For the classical orthogonal polynomials it holds that dµ = ω(x) dx.

Moreover the so-called ω usually vanishes outside some interval [a, b]. In this case we write L2([a, b], ω(x)dx), and the scalar product is given by

h f, g i = Z b

a

f (x) g(x)ω(x) dx.

In case h f, g i = 0 we say that f and g are orthogonal with respect to the weight ω on [a, b].

For more about the theory of inner product spaces and Hilbert spaces, see [3].

(7)

3 Gram-Schmidt process

The Gram Schmidt process is an algorithm that is used to turn any linearly independent set of vectors into a orthonormal set of vectors.

We denote the linearly independent set of vectors by xj, and the resulting orthonormal set of vectors by ej. The steps of the algorithm are as follows:

• First step: The first element of the orthonormal sequence, e1, will be obtained from

e1 = 1 kx1kx1.

• Second step: All the following steps will include two parts: first create a vector orthogonal to the previous vector(s), then normalize it. We create v2 as

v2 = x2− h x2, e1ie1, and then normalize it:

e2 = 1 kv2kv2.

• Third step: We create v3 as

v3 = x3− h x3, e1ie1 − h x3, e2ie2 and then normalize:

e3 = 1 kv3kv3. The algorithm proceeds by induction.

• nth step: Suppose that {e1, ..., en−1} is an orthonormal set such that span{e1, ..., en−1} = span{x1, ..., xn−1}. The vector vn is defined by:

vn = xn

n−1

X

k=1

h xn, ekiek.

Note that vn is a non-zero vector. Otherwise xn would belong to the span of {x1, ..., xn−1} and the set {x1, ..., xn} would be linearly dependent. Note also that vn is orthogonal to all the vectors e1, ..., en−1.

Normalizing vn:

en = 1 kvnkvn,

we therefore obtain an orthonormal set {e1, ..., en} with span{e1, ..., en} = span{x1, ..., xn}.

Example 5. Gram-Schmidt procedure on vectors in R3

Consider the two vectors x1 = (1, 1, 0) and x2 = (1, 2, 1). The Gram-Schmidt procedure can be used to obtain a set {e1, e2} that is orthonormal with respect to the standard scalar product in R3.

(8)

• First step: The vector e1 is obtained by normalizing x1: e1 = 1

kx1kx1 = 1

√2



1, 1, 0 .

• Second step: We create v2 as:

v2 = x2− h x2, e1ie1 = (1, 2, 1) − 3

2(1, 1, 0) = 1

2(−1, 1, 2).

And now we normalize:

e2 = 1

kv2kv2 = 1

√6(−1, 1, 2).

Example 6. Gram-Schmidt process on polynomials

Consider the set u = {1, x, x2}, and let u1 = 1, u2 = x and u3 = x2. The Gram-Schmidt process can be used to obtain a set {e1, e2, e3} that is orthonormal with respect to the inner product

h f, g i = Z 1

−1

f (x) g(x) dx.

• First step: The first element of the orthonormal sequence, e1, will be obtained from

e1 = 1

ku1ku1 = 1

√2

• Second step: We create v2 as:

v2 = u2− h u2, e1ie1 = x − hx, 1i = x − Z 1

−1

x · 1 dx = x.

And then make it normalized:

e2 = 1

kv2kv2 = x q2

3

= r3

2x.

• Third step: We create v3 as:

v3 = u3− h u3, e1ie1− h u3, e2ie2

= x2

 Z 1

−1

x2· 1

√2dx



· 1

√2 −

 Z 1

−1

x2· r3

2x dx



· r3

2x

= x2−1 2 · 2

3− 0 = x2−1 3 Note that

kv3k2 = Z 1

−1

 x2 −1

3

2

dx = 8 45, and therefore:

e3 = 1

kv3kv3 = 3√ 5 2√

2

 x2− 1

3

 .

These are, up to a multiplicative factor, the first three so-called Legendre polynomials.

(9)

4 Orthogonal polynomials

4.1 Orthonormal and monic orthogonal polynomials

Let dµ be a positive Borel measure on R having finite moments, that is Z

R

|x|mdµ < ∞ (4.1)

for all integers m ≥ 0. Furthermore assume that supp µ contains infinitely many points, i.e. #(supp µ) = ∞. Then the set {1, x, x2, ...} is linearly independent. The

orthonormal polynomials, pn(x), satisfy Z

pn(x) pm(x) dµ = δnm (4.2)

where

δnm=

(0, n 6= m, 1, n = m.

We assume that the coefficient of the highest order term is positive:

pn(x) = γnxn+ ..., γn> 0. (4.3) This makes the pn’s unique. They can be obtained by applying the Gram-Schmidt process to the set {1, x, x2, ...}.

The orthogonal polynomials having leading coefficient 1 are called monic orthogonal polynomials. We denote them by πn(x), i.e. πn(x) = pn(x)/γn.

Let M = Mn be the (n + 1) × (n + 1) matrix where all the elements are M = (Mij) =

 Z

xi+jdµ(x)



0≤i,j≤n

, and let Dn denote its determinant:

Dn = det M =

R s0dµ R sdµ · · · R sndµ R sdµ R s2dµ · · · R sn+1

... ... ... ... R sn−1dµ R sndµ · · · R s2n−1

R sndµ R sn+1dµ · · · R s2ndµ .

Note that

0 ≤

n

X

i=0

tixi

2

L2(R,dµ) =DXn

i=0

tixi,

n

X

j=0

tjxjE

L2(R,dµ)

=

n

X

i,j=0

titj Z

R

xi+jdµ =

n

X

i,j=0

titjMi,j = hM t, tiRn+1,

(4.4)

where t = {t0, ..., tn}T ∈ Rn+1. Since #(supp µ) = ∞ it follows that the matrix M is strictly positive definite, and therefore

Dn> 0.

(10)

For notational convenience we set D−1 = 1. Now set D0(x) = 1 and for n ≥ 1

Dn(x) =

R s0dµ R sdµ · · · R sndµ R sdµ R s2dµ · · · R sn+1

... ... ... ... R sn−1dµ R sndµ · · · R s2n−1

1 x · · · xn

.

There is an explicit formula for the orthonormal polynomials, due to Heine.

Theorem 1. Let pn(x) for n ≥ 0 be the sequence of polynomials orthogonal with respect to dµ. Then

pn(x) = 1

√Dn−1DnDn(x).

Proof. To prove that

pn(x) = 1

√Dn−1DnDn(x) (4.5)

we write ˜pn(x) = √ 1

Dn−1Dn

Dn(x). For all j < n, multiplying with xj we get following expression

pDnDn−1 Z

xjn(x)dµ =

=

R s0dµ R sdµ · · · R sndµ R sdµ R s2dµ · · · R sn+1

... ... ... ... R sn−1dµ R sndµ · · · R s2n−1

R xjdµ R xj+1dµ · · · R xj+ndµ .

Note that the last row in the matrix above is identical to some other row lying above.

When two rows in a matrix are identical, the determinant will be 0, as seen e.g. using row operations. That gives us

Z

xjn(x) dµ = 0,

which also implies that ˜pn(x) ⊥ ˜pm(x) for all m < n. Furthermore we want to show that k˜pnk2 = 1. Because of the equation ˜pn(x) = √ 1

Dn−1Dn

Dn(x), we have the following

k˜pnk2 = h ˜pn, ˜pni = 1 Dn−1Dn

hDn(x), Dn(x)i.

The right hand of this equation can be expressed using the definition of the inner product:

k˜pnk2 = 1 Dn−1Dn

Z

Dn(x) Dn(x) dµ.

(11)

Expanding the second factor Dn(x) along the last row, it follows that this can be rewritten as

k˜pnk2 = 1 Dn−1Dn

Z

Dn(x) (Dn−1xn+ · · · ) dµ where the dots represent the terms of lower order.

Note that all the terms of lower order will be orthogonal to Dn(x) and therefore these integrals will be zero. What we have left is the following expression:

k˜pnk2 = 1 Dn

Z

Dn(x) xndµ.

Using expansion along the last row one realizes that this can be rewritten as:

k˜pnk2 = 1 Dn

R x0dµ R xdµ · · · R xndµ R xdµ R x2dµ · · · R xn+1

... ... ... ... R xn−1dµ R xndµ · · · R x2n−1

R xndµ R xn+1dµ · · · R x2ndµ ,

where the determinant is exactly the same thing as Dn. Thus k˜pnk2 = 1

DnDn = 1.

Finally, expanding Dn(x) along the last row one sees that the coefficient for the highest order term in ˜pn is given by

√ 1

Dn−1Dn · Dn−1 =r Dn−1

Dn > 0. (4.6)

Hence ˜pn(x) = pn(x).

4.2 Three-term recurrence relation

Since x pk(x) is a polynomial of degree k + 1, it can be written as a sum

x pk(x) =

k+1

X

j=0

cjpj(x) (4.7)

for some constants cj. Suppose that k ≥ 1 and that l < k − 1. Then, due to orthogonality, we have following expression:

Z

x pk(x) pl(x) dµ = Z

pk(x) x pl(x) dµ = 0, (4.8)

(12)

since x pl(x) is a polynomial of degree l + 1 < k. On the other hand, by (4.7) we have Z

x pk(x) pl(x) dµ =

k+1

X

j=0

cj Z

pj(x) pl(x) dµ = cl. (4.9)

By comparing equation (4.8) and equation (4.9), we see that for all l < k − 1 it holds that cl= 0. That gives us only three terms left in (4.7): when l = k − 1, l = k and l = k + 1. Therefore, the polynomial x pk(x) can be expressed as

x pk(x) = ck+1pk+1(x) + ckpk(x) + ck−1pk−1(x), k ≥ 1. (4.10) This is the famous three-term recurrence formula. Obviously, we also have

x p0(x) = c1p1(x) + c0p0(x), k = 0. (4.11) Set

ak= Z

x p2k(x) dµ bk=

Z

x pk+1(x) pk(x) dµ

for k ≥ 0. Multiplying equation (4.10) by pk(x) and integrating with respect to dµ, we get:

ak= Z

x p2k(x) dµ = 0 + ck Z

p2k(x) dµ + 0 = ck. (4.12) Similarly multiplying equation (4.10) by pk+1(x) and integrating with respect to dµ, we get:

bk = Z

x pk+1(x) pk(x) dµ = ck+1 Z

p2k+1dµ + 0 + 0 = ck+1. (4.13) Finally, multiplying equation (4.10) by pk−1(x) and integrating with respect to dµ, we get:

bk−1 = Z

x pk(x) pk−1(x) dµ = 0 + 0 + ck−1 Z

p2k−1dµ = ck−1. (4.14) Doing similar calculation for the case k = 0 we get:

x pk(x) = bkpk+1(x) + akpk(x) + bk−1pk−1(x), k ≥ 1, (4.15) x p0(x) = b0p1(x) + a0p0(x), k = 0. (4.16) The leading coefficient of the left-hand side of the equation (4.15) is γk and on the right-hand side bk· γk+1. Therefore:

γk = bk· γk+1,

(13)

so that

bk= γk

γk+1 > 0. (4.17)

If we set T to be the tridiagonal matrix

T =

a0 b0 0 · · · b0 a1 b1 . ..

0 b1 a2 . ..

... . .. ... ...

 the equations (4.15) and (4.16) can be rewritten as

T

 p0(x) p1(x)

... pk(x)

...

= x

 p0(x) p1(x)

... pk(x)

...

 .

4.3 The Jacobi matrix

Definition 4.1. The Jacobi matrix is defined as the symmetric matrix T such that:

T =

a0 b0 0 · · · 0 b0 a1 b1 . .. ... 0 b1 a2 . .. 0

... . .. ... ... bn−2 0 · · · 0 bn−2 an−1

, bj > 0, (4.18)

in the finite-dimensional case, and:

T =

a0 b0 0 · · · b0 a1 b1 . ..

0 b1 a2 . ..

... . .. ... ...

, bj > 0, (4.19)

in the infinite-dimensional case.

Theorem 2. Let T be a finite n × n Jacobi matrix. Then the following holds:

1. The spectrum of T is simple. Thus, there exist n distinct eigenvalues λ1, λ2, ..., λn. 2. If f = (f (1), f (2), ..., f (n))T 6= 0 is an eigenvector of T , i.e., T f = λf , then

f (1) 6= 0 and f (n) 6= 0.

Proof. We begin by proving the second statement. Assume that:

(T − λI)f = 0 (4.20)

(14)

where f = (f (1), f (2), ..., f (n))T. The first row of equation (4.20) reads:

(a0− λ)f (1) + b0f (2) = 0. (4.21) For the proof we use proof by contradiction. Suppose that f (1) = 0. Since b0 > 0 it follows from equation (4.21) that f (2) = 0. The equations in the middle rows of (4.20) can be written as

bi−2f (i − 1) + (ai−1− λ)f (i) + bi−1f (i + 1) = 0 (4.22) for 2 ≤ i < n. From equation (4.22), with i = 2, and get following expression:

b0f (1) + (a1− λ)f (2) + b1f (3) = 0 (4.23) Since f (1) = 0, f (2) = 0 and b1 > 0 it follows that f (3) = 0, too. Continuing in this manner, setting i = 3, ..., i = n − 1, it follows f (1) = f (2) = ... = f (n) = 0. This means that, if f (1) = 0, then f = 0. Therefore f (1) 6= 0. That f (n) 6= 0 follows from a similar reasoning, starting from the last row.

For the proof of the first statement, we suppose that f and ˜f are two (non-zero) eigenvectors with the same eigenvalue λ. We can find a pair (a, ˜a) 6= (0, 0) such that g = af + ˜a ˜f has g(1) = 0. For example we may take a = 1 and ˜a = −f (1)˜

f (1). Since g is in the same eigenspace it follows, from the proof of the second statement, that g = 0. This means that f and ˜f are linearly dependent. So each eigenspace in one-dimensional.

Since the matrix T is symmetric, it follows that it must have n distinct real eigenvalues.

4.4 Zeros of orthogonal polynomials

When considering the zeros of orthogonal polynomials, there are some properties that characterise the zeros. We will focus on two of them:

Theorem 3. Let dµ be a Borel measure with finite moments and infinite support, and let pn(x) be the associated orthonormal polynomials. Then,

1. The n zeros of pn(x), are real and simple.

2. The zeros of pn+1(x) and pn(x) interlace, that is, between any two zeros of pn+1(x) lies exactly one zero of pn(x), and vice versa.

Proof of 1. When proving the first property we may assume, without loss of generality, that R dµ = 1. Let T be the Jacobi matrix such that

T =

a0 b0 0 · · · b0 a1 b1 . ..

0 b1 a2 . ..

... . .. ... ...

 .

We know that

(15)

0 = (T −λ)

 p0(λ) p1(λ)

...

=

a0− λ b0 0 · · · b0 a1− λ b1

0 b1 a2− λ . ..

. . . . .. . .. bn−2

bn−2 an−1− λ bn−1 bn−1 an− λ bn

. .. ...

 p0(λ) p1(λ)

...

.

Hence, pn(λ) = 0 if and only if

(p0(λ), ..., pn−1(λ)) = (1, p1(λ), .., pn−1(λ)) 6= 0, is an eigenvector of the n × n Jacobi matrix

a0 b0 0 · · · 0 b0 a1 b1 . .. ... 0 b1 a2 . .. 0

... . .. ... ... bn−2 0 · · · 0 bn−2 an−1

(4.24)

with eigenvalue λ. Therefore, the zeros of pn are the same as the eigenvalues of the matrix in equation (4.24). But, and due to Theorem 2 this matrix has n real and simple zeros.

Proof of 2. The roots of pn(x) and pn+1(x) are eigenvalues of the two matrices Tn−1 and Tn, where

Tn−1 =

a0 b0 0 · · · 0 b0 a1 b1 . .. ... 0 b1 a2 . .. 0

... . .. ... ... bn−2 0 · · · 0 bn−2 an−1

 ,

Tn=

a0 b0 0 · · · 0 b0 a1 b1 . .. ... 0 b1 a2 . .. 0 ... . .. ... ... bn−1 0 · · · 0 bn−1 an

 .

Consider

h en, (Tn− λI)−1eni,

the (n, n)-entry in the resolvent (Tn− λI)−1. Observe that we here use the notation e0, ..., en for the standard vectors in Rn+1, and index rows and columns in the

(n + 1) × (n + 1) matrix Tn from 0 to n. Now recall the formula A−1 = 1

det Aadj (A)

(16)

for the inverse of a matrix A in terms of the so-called adjoint of A:

adj (A) = (cof (A))T. Using this gives us the following:

h en, (Tn− λI)−1eni = det(Tn−1− λI)

det(Tn− λI) = − πn(λ)

πn+1(λ), (4.25) where π are the monic orthogonal polynomials. The spectral theorem gives us:

Tn= UnΛ UnT, (4.26)

where Un is the orthogonal matrix whose columns are the eigenvectors of Tn and Λ = (λ0, ..., λn) is the diagonal matrix of eigenvalues. Write Un = (fj(i)). Using the spectral theorem

Tn− λI = UnΛ UnT − λI = Un(Λ − λI) UnT, so that

(Tn− λI)−1 = Un(Λ − λI)−1UnT. From equation (4.25) we get

− πn(λ)

πn+1(λ) = h en, (Tn− λI)−1eni = h en, Un(Λ − λI)−1UnTeni

= h UnT en, (Λ − λI)−1UnT eni =

n

X

j=0

fj2(n) λj− λ.

(4.27)

Therefore,

n

X

j=0

fj2(n)

λ − λj = πn(λ)

πn+1(λ). (4.28)

By looking at the left-hand side of the equation, we realise that when λ is close to λj the ratio will give us either big, positive values or big negative values, depending on whether λ gets closer to λj from the left-hand side or the right-hand side. The derivative

d dλ

Xfj2(n)/(λ − λj) = −X

fj2(n)/(λ − λj)2

is negative, so that the function in the left-hand side of equation (4.28) is decreasing everywhere.

The graph of the function in the left-hand side of equation (4.28) can be visualized as in Figure 1 below.

λ0 λ1 λ2 λn−1 λn

Figure 1: Interlacing zeros of orthogonal polynomials.

(17)

Thus, the left-hand side of equation (4.28) must have a zero between any two zeros of πn+1(λ). But this must be a zero of πn(λ).

For further theory about orthogonal polynomials, see [1].

5 Classical orthogonal polynomials

Many orthogonal polynomials are named after the ones who discovered them. This thesis will focus on three of these orthogonal polynomials.

• Legendre polynomials

• Hermite polynomials

• Laguerre polynomials

The polynomials mentioned above are sometimes called classical orthogonal polynomials.

They are polynomials orthogonal with respect to different weight functions. It turns out that they are also solutions to differential equations.

5.1 Legendre polynomials

Consider the polynomials which are orthogonal on the interval [−1, 1] with respect to the inner product

h f, g i = Z 1

−1

f (x) g(x) dx,

i.e. w.r.t. the weight function ω(x) = 1. The Legendre polynomials Pn(x) are defined by

pn(x) =

r2n + 1

2 Pn(x), (5.1)

where pn(x) denotes the orthonormal polynomial. It turns out that the Legendre polynomials are given by the formula

Pn(x) = 1 2nn!

dn

dxn[(x2− 1)n]. (5.2)

Equation (5.2) is called Rodrigues’ formula. Using the binomial theorem on (x2− 1)n and differentiating the result n times, it follows that:

Pn(x) =

N

X

j=0

(−1)j (2n − 2j)!

2nj!(n − j)!(n − 2j)!xn−2j (5.3) where N = n/2 if n is even and N = (n − 1)/2 if n is odd. The six first Legendre

polynomials will, by help of equation (5.3), therefore be:

P0(x) = 1 P1(x) = x

P2(x) = 12(3x2 − 1) P3(x) = 12(5x3− 3x)

P4(x) = 18(35x2− 30x2+ 3) P5(x) = 18(63x5− 70x3+ 15x)

(18)

Theorem 4. The polynomials pn(x) defined from equations (5.1) and (5.2) are orthonormal i.e. the Legendre polynomials satisfy

h Pn, Pmi =

(0 if m 6= n,

2

2n+1 if m = n.

Proof. For m = n, we must show that kPnk =hZ 1

−1

Pn2(x)dxi1/2

=

r 2

2n + 1. (5.4)

Let u = x2− 1. The function un and its derivatives (un)0, . . . , (un)(n−1) are zero at x = ±1, and (un)(2n) = (2n)!. Integrating by parts n times, using equation (5.2) we get:

(2nn!)2kPnk2 = Z 1

−1

(un)(n)(un)(n)dx

= (un)(n−1)(un)(n)

1

−1

− Z 1

−1

(un)(n−1)(un)(n+1)dx

= · · ·

= (−1)n(2n)!

Z 1

−1

undx

= 2(2n)!

Z 1 0

(1 − x2)ndx.

Doing the substitution x = sin τ and we get:

Z 1 0

(1 − x2)ndx = Z π/2

0

cos2n+1τ dτ = Z π/2

0

cos τ · cos2nτ dτ

Let In =Rπ/2

0 cos τ · cos2nτ dτ . Clearly I0 = 1. Doing integration by parts, we get:

In= (sin τ · cos2nτ )

π/2

0

− Z π/2

0

sin τ · 2n cos2n−1τ (− sin τ )dτ =

= 2n Z π/2

0

cos2n−1τ (1 − cos2τ )dτ

= 2n Z π/2

0

cos2n−1τ − cos2n+1

= 2n(In−1− In).

That is,

In = 2n(In−1− In), which gives

In= 2n

2n + 1In−1. (5.5)

(19)

Repeated use of equation (5.5) gives In = 2n

2n + 1 · 2(n − 1)

2(n − 1) + 1In−2 = · · · = 2n · 2(n − 1) · ... · 2 (2n + 1)(2n − 1) · ... · 3I0. Extend both numerator and denominator with the even factors 2n, (2n − 2), (2n − 4), ..., 2, it follows that

(2nn!)2kPnk2 = 2(2n)! · 22nn2(n − 1)2(n − 2)2· ... · 12

(2n + 1)(2n)! = 22n+1(n!)2 (2n + 1)(2n)!. Now dividing both sides with (2nn!)2 gives that

kPnk2 = 2 2n + 1. To complete the proof it is enough to check that

h xm, Pni = 0 for m < n. But for m < n,

2nn! h xm, Pni = Z 1

−1

xm(un)(n)dx

= xm(un)(n−1)

1

−1

− m Z 1

−1

xm−1(un)(n−1)dx

= · · ·

= (−1)mm!

Z 1

−1

(un)(n−m)dx

= (−1)mm!(un)(n−m−1)

1

−1

= 0,

since n − m − 1 is an integer k such that 0 ≤ k ≤ n − 1.

The Legendre polynomials are known to be solutions of the Legendre differential equation (5.6)

(1 − x2)Pn00− 2xPn0+ n(n + 1)Pn = 0. (5.6)

5.2 Hermite polynomials

The second class of polynomials are the Hermite polynomials. They are orthogonal on R with respect to the inner product

h f, g i = Z

−∞

f (x) g(x) e−x2dx, (5.7)

i.e. with the weight function ω(x) = e−x2.

(20)

The Hermite polynomials Hn(x) are defined by pn(x) = 1

(2nn!√

π)1/2Hn(x), (5.8)

where pn(x) is the sequence of orthonormal polynomials. It can be shown that Hn(x) = (−1)nex2 dn

dxn(e−x2). (5.9)

When developing the differentials from equation (5.10) we get:

Hn(x) = n!

N

X

j=0

(−1)j 2n−2j

j!(n − 2j)!xn−2j (5.10)

where N = n/2 if n is even and N = (n − 1)/2 if n is odd. With help from equation (5.10) the six first polynomials can be computed:

H0(x) = 1 H1(x) = 2x

H2(x) = 4x2− 2 H3(x) = 8x3− 12x

H4(x) = 16x4− 48x2+ 12 H5(x) = 32x5−160x3+120x Theorem 5. The polynomials pn(x) defined by equation (5.8) and (5.9) are orthonormal, i.e. the Hermite polynomials satisfy

Z

−∞

e−x2Hm(x)Hn(x)dx =

(0 if m 6= n, 2nn!√

π if m = n.

Proof. Equation (5.10) can also be written as:

Hn(x) =

N

X

j=0

(−1)j

j! n(n − 1) · · · (n − 2j + 1)(2t)n−2j (5.11) when n = 2, 3, 4.... For n ≥ 1 we get following from differentiating:

Hn0(x) = 2n

M

X

j=0

(−1)j

j! (n − 1)(n − 2) · · · (n − 2j)(2t)n−1−2j = 2n Hn−1(x) (5.12) where M = (n − 2)/2 if n is even and M = (n − 1)/2 if n is odd. To prove the theorem we apply equation (5.12) to Hm, and we assume that m ≤ n. Note that, since

ω(x) = e−x2 it follows from equation (5.9) that

(−1)ne−x2Hn(x) = ω(n)(x).

(21)

Integrating by parts, m times, we get (−1)n

Z

−∞

e−x2Hm(x)Hn(x)dx = Z

−∞

Hm(x)ω(n)(x)dx

= Hm(x)ω(n−1)(x)

−∞

− Z

−∞

2mHm−1ω(n−1)(x)dx

= −2m Z

−∞

Hm−1(x)ω(n−1)(x)dx

= · · ·

= (−1)m2mm!

Z

−∞

H0(x)ω(n−m)(x)dx.

In the last row of the equation, H0(x) = 1. If m < n, integrating once more, we obtain 0 since ω and its derivatives approach 0 as x → ±∞. If m = n, the integral in the last row equals:

Z

−∞

e−x2dx =√ π.

There is a differential equation that the Hermite polynomials satisfy, called the Hermite differential equation:

Hn00− 2x Hn0 + 2n Hn= 0. (5.13)

5.3 Laguerre polynomials

Consider the polynomials which are orthogonal on [0, ∞) with respect to the inner product

h f, g i = Z

0

f (x) g(x) e−xdx, (5.14)

i.e. the weight function ω(x) = e−x. The Laguerre polynomials are defined by

pn(x) = e−x/2Ln(x) (5.15)

where pn(x) is the sequence of orthonormal polynomials. It can be shown that Ln(x) = ex

n!

dn

dxn(xne−x), n = 0, 1, 2, .... (5.16) Equation (5.16) can be expressed without the differentiation. Then it will look like

Ln(x) =

n

X

j=0

(−1)j j!

n j



xj. (5.17)

Using the formula from equation (5.17), the first five polynomials can be calculated:

L0(x) = 1 L1(x) = −x + 1

L2(x) = 12(x2− 4x + 2) L3(x) = 16(−x3+ 9x2− 18x + 6) L4(x) = 241 (x4− 16x3+ 72x2− 96x + 24)

(22)

The Laguerre polynomials are a solution to a differential equation, called the Laguerre differential equation:

xL00n+ (1 − x)L0n+ nLn= 0.

The above results concerning classical orthogonal polynomials can be found in [3].

6 Sturm-Liouville problems

Many mathematical and physical problems lead to a differential equation of the form:

(p(x)v0)0+ q(x)v + λr(x)v = 0 a < x < b. (6.1) In addition one has boundary conditions at the end points of the interval. The

differential equation (6.1) together with boundary conditions is called a Sturm-Liouville eigenvalue problem, and a solution is called an eigenfunction with eigenvalue λ.

Example 7. Consider the Sturm-Liouville problem d2v

dx2 + λv = 0 0 < x < L, (6.2)

v(0) = v(L) = 0, (6.3)

where p = r = 1 and q = 0. This boundary condition is called a Dirichlet boundary condition. The eigenfunctions and eigenvalues for this problem are

vn(x) = sinnπx

L , λn= nπ L

2

n = 1, 2, 3, ...

Example 8. Consider the Sturm-Liouville problem d2v

dx2 + λv = 0 0 < x < L, (6.4)

v0(0) = v0(L) = 0, (6.5)

where p = r = 1 and q = 0. This boundary conditions is called a Neumann boundary condition. The eigenfunctions and eigenvalues for this problem are

vn(x) = cosnπx

L , λn = nπ L

2

n = 0, 1, 2, ....

Example 9. Consider the Sturm-Liouville problem with d2v

dx2 + λv = 0 0 < x < L, (6.6)

v(0) = v(L), v0(0) = v0(L), (6.7)

where p = r = 1 and q = 0. This kind of boundary condition is called a periodic boundary condition. The eigenfunctions and eigenvalues for this problem are

vn(x) = αncos 2nπx L



+ βnsin 2nπx L



, λn = 2nπ L

2

n = 1, 2, 3, ....

The results above can all be shown by elementary computations using the fact that the eigenvalues must be real.

(23)

Example 10. The Bessel equation.

For ν ∈ [0, ∞), consider the differential equation

r2w00(r) + rw0(r) + (r2− ν2)w(r) = 0, r > 0. (6.8) The equation above is called the Bessel equation of order ν.

Divide (6.8) with r and put v(x) = w(r) where x = r/√

λ. Then (xv0(x))0+



λx − ν2 x



ν(x) = 0. (6.9)

This equation has the form of a the Sturm-Liouville equation with p(x) = r(x) = x and q(x) = −ν2/x. The bounded solution of equation (6.8), denoted Jν, is called the Bessel function (of the first kind) of order ν.

Theorem 6. Consider the Sturm-Liouville equation

(p(x)v0)0+ q(x)v + λr(x)v = 0, a < x < b, (6.10) together with Dirichlet, Neumann or periodic boundary conditions. Then the

eigenfunctions corresponding to different eigenvalues are orthogonal with respect to a inner product

h u, v ir = Z b

a

u(x) v(x) r(x) dx.

Proof. Let us write

Lv := (pv0)0+ qv. (6.11)

Using the product rule, we get:

uLv − vLu = u(pv0)0+ uqv − v(pu0)0− vqu

= (upv0)0− u0pv0− (vpu0)0+ u0pv0. (6.12) Equation (6.12) can be simplified to

uLv − vLu = [p(uv0 − vu0)0]0 (6.13) Equation (6.13) is called the Lagrange identity. Integrating equation (6.13) over the interval [a, b], we get following equation, which is called Green’s formula:

Z b a

(uLv − vLu) dx = p (uv0− vu0)

b

a. (6.14)

Assuming that u and v satisfy the boundary conditions, then p (uv0− vu0)

b

a = 0, (6.15)

which leads to

Z b a

(uLv − vLu) dx = 0. (6.16)

(24)

Now assume that vn and vm are two eigenfunctions with two different eigenvalues, λn and λm. Then

− Lvn = λnr vn, (6.17)

− Lvm = λmr vm. (6.18)

If we multiply equation (6.17) with vm and equation (6.18) with vn, integrating them over the interval [a, b] and finally taking the difference of these two equations, we obtain the following expression:

− Z b

a

(vmLvn− vnLvm) dx = (λn− λm) Z b

a

vnvmr dx (6.19) Applying Green’s formula we get that

n− λm) Z b

a

vnvmr dx = 0. (6.20)

Since λn 6= λm it follows that

h vn, vmir = Z b

a

vnvmr dx = 0, (6.21)

proving orthogonality.

For further theory about Sturm-Liouvlle problems, see [4, Pinchover].

The three classical orthogonal polynomials were all solutions to differential equations:

• The Legendre differential equation

(1 − x2)v00− 2xv0+ λv = 0 (6.22)

• The Hermite differential equation

v00− 2xv0+ λv = 0 (6.23)

• The Laguerre differential equation

xv00+ (1 − x)v0+ λv = 0 (6.24) They can all be written in the form of Sturm-Liouville equation for different choices of p, q and r.

The Legendre differential equation can, clearly, be written in the form (6.1), where p = (1 − x2), q = 0 and r = 1. Here [a, b] = [−1, 1].

Now consider the Hermite differential equation. After multiplying both sides of equation (6.1) with e−x2 we see that

e−x2v00− 2xe−x2 + λe−x2v = 0, or

(e−x2v0)0+ λe−x2v = 0.

(25)

This gives us: p = e−x2, q = 0 and r = e−x2. Here (a, b) = (−∞, ∞).

Finally consider the Laguerre differential equation. Multiplying both sides of (6.1) with e−x, we find

e−xx v00+ e−x(1 − x)v0+ e−xλv = 0. (6.25) It is easy to see that this can be rewritten as

(e−xxv0)0+ e−xλv = 0, (6.26) that is p = x e−x, q = 0 and r = e−x. Here [a, b) = [0, ∞).

It is interesting to note that the orthogonality of the classical orthogonal polynomials with respect to the weights indicated in the previous work can also be seen from the proof of Theorem 6. Indeed formula (6.15) still holds true, even in the absence of boundary conditions. Simply note that p vanishes at the ends of the intervals (and, in case the intervals are unbounded, p decays faster than the polynomials).

For more about the theory of Sturm-Liouville problems, see [4].

7 Applications

7.1 Eigenvalues of the Laplace operator

Let Ω be a bounded open subset of Rn. The eigenvalue problem for the Dirichlet Laplace operator

∆ = ∂2

∂x21 + · · · + ∂2

∂x2n is then given by the equation

− ∆u = λu, (7.1)

where u is a function in Ω such that u = 0 on ∂Ω.

One can show that the Dirichlet Laplace operator has discrete eigenvalues (λj)1 , 0 < λ1 ≤ λ2 ≤ λ3 ≤ ...,

accumulating at infinity. They are important in many applications. For example, consider the wave equation

2v

∂t2 − c2∆v = 0

for the amplitude v(x, t) of a wave at the position x = (x1, ..., xn) at time t. Separating variables by writing v(x, t) = u(x)T (t) we find that

u T00− c2∆u T = 0.

Dividing both sides by c2uT , it follows that

∆u u = T00

c2 .

Both sides must be constant, i.e. independent of (x, t), and denoting the constant by

−λ we find that u must satisfy the equation (7.1), whereas

T00+ λc2T = 0. (7.2)

(26)

The solutions of equation (7.2) describe periodic motion depending on the frequency ω, where ω2 = λc2. Note that the frequencies are determined by the eigenvalues of the Laplacian.

7.1.1 Eigenvalue problem in a disk

To find the frequencies generated by a circular drum we must find the eigenvalues for the Dirichlet Laplace operator in a disk. Let Ω be the disk defined by

{0 ≤ r < a, 0 ≤ θ ≤ 2π}.

The eigenvalue problem (7.1), written in polar coordinates, becomes

urr+1

rur+ 1

r2uθθ = −λu (7.3)

where 0 < r < a, 0 ≤ θ ≤ 2π and u(a, θ) = 0. To find the solutions we separate variables: u(r, θ) = R(r)Θ(θ). Then we get two Sturm-Liouville problems:

Θ00(θ) + µ Θ(θ) = 0 (7.4)

where Θ(0) = Θ(2π), and

R00(r) +1

rR0(r) +

 λ − µ

r2



R(r) = 0. (7.5)

The solutions of the first problem is

Θn(θ) = Ancos n θ + Bnsin n θ (7.6) where µn = n2 for n = 1, 2, 3, .... Using this, the second Sturm-Liouville problem

(equation (7.5)) can be written as:

R00(r) +1

rR0(r) + λ − n2

r2



R(r) = 0. (7.7)

Doing the change of variables s =√

λr and R(r) = Ψ(s) we get:

R0(r) = d

drR(r) = d drΨ(√

λr) = Ψ0(√

λr) ·√

λ (7.8)

and

R00(r) = Ψ00(s) · λ. (7.9)

Therefore equation (7.7)

λ Ψ00(s) +

√ λ

s Ψ0(s)√

λ + λ − n2λ

s2 Ψ(s) = 0. (7.10)

Dividing both sides with λ, it follows Ψ00(s) + 1

0(s) + 1 − n2

s2



Ψ(s) = 0. (7.11)

(27)

After multiplying both sides of equation (7.11) by s2, we see that Ψ satisfies the Bessel equation of order n, compare (6.8). Since Ψ is bounded, we have that Ψ(s) = Jn(s).

The function R(r) = Jn(s) = Jn(√

λr) must vanish when r = a. If we let (αn,m)m=1 denote the zeros of the function Jn on the positive real line, we see that the eigenvalues λ are given by

λa = αn,m or

λ = (αn,m)2 a2 .

This implies that the frequencies that a circular drum can generate are given by ω =√

λc = αn,m a c.

7.1.2 Eigenvalue problem in a ball

Imagine a solid ball of radius a given in spherical coordinates. Let us introduce the spherical coordinates

x = r sin φ cos θ, y = r sin φ sin θ, z = r cos φ, see Figure 2 below.

y z

x

r

θ φ

Figure 2: Notation for a spherical coordinate system.

We define Ba and S2 by

Ba = {0 ≤ r < a, 0 ≤ φ ≤ π, 0 ≤ θ ≤ 2π} and S2 = {0 ≤ φ ≤ π, 0 ≤ θ ≤ 2π}.

The eigenvalue problem for the Dirichlet Laplace operator in Ba, written in spherical coordinates, becomes

1 r2

∂r

 r2∂u

∂r

 + 1

r2

 1 sin φ

∂φ



sin φ∂u

∂φ



+ 1

sin2φ

2u

∂θ2



(7.12)

= −λu, (r, φ, θ) ∈ Ba,

u(a, φ, θ) = 0, (φ, θ) ∈ S2. (7.13)

(28)

The function u(r, φ, θ) can be expressed, using the method of separating variables, as u(r, φ, θ) = R(r)Y (φ, θ).

We obtain two eigenvalue problems, one over the unit sphere S2 and the one for the radial function R(r). The eigenvalue problem over the unit sphere takes the form

1 sin φ

∂φ



sin φ∂Y

∂φ



+ 1

sin2φ

2Y

∂θ2 = −µY (φ, θ) ∈ S2, (7.14) with the boundary condition

Y (φ, 0) = Y (φ, 2π), Yθ(φ, 0) = Yθ(φ, 2π). (7.15) The second eigenvalue problem takes the form

1 r2

d dr

 r2dR

dr



= µ r2 − λ



R 0 < r < a, (7.16) with boundary condition R(a) = 0. The function Y can be expressed, using the method of separating variables, as

Y (φ, θ) = Φ(φ)Θ(θ).

Substituting this form of Y into equation (7.14), we obtain

Θ00(θ) + νΘ(θ) = 0, (7.17)

where 0 < θ < 2π, and sin φ d



sin φdΦ(φ) dφ



+ (µ sin2φ − ν)Φ(φ) = 0, (7.18) where 0 < φ < π. The periodic boundary condition from equation (7.15) gives us the eigenvalues νm = m2 and eigenfunctions

Θm(θ) = Amcos mθ + Bmsin mθ

for m = 0, 1, 2, .... Substituting ν = m2 into the equation (7.18) for Φ(φ) and doing the change of variables x = cos φ, P (x) = Φ(φ(x)), we get

(1 − x2) d dx



(1 − x2)dP dx



+(1 − x2)µ − m2P = 0 (7.19) where −1 < x < 1 and m = 0, 1, 2, .... For m = 0 we obtain

d dx



(1 − x2)dP dx



+ µP = 0, −1 < x < 1, (7.20) i.e. the Legendre equation. It turns out that equation (7.20) has bounded solutions if and only if µ = n(n + 1) where n = 0, 1, 2, ...; they are the Legendre polynomials, Pn. For m > 0, equation (7.19) is called the associated Legendre equation of order m. This equation has got bounded solutions when µ = n(n + 1) given by

Pnm(x) = (1 − x2)m/2dmPn

dxm . (7.21)

References

Related documents

[r]

Each pair (b, i) corresponds to a certain Ehrhart polynomial, and that Ehrhart polynomial is said to be realized by a lattice polygon having b boundary points and i interior

Lam´ e equation, Schr¨ odinger equation, spectral polynomials, asymp- totic root distribution, hypergeometric functions, elliptic functions, finite gap

In this thesis it will be shown that the hydrogen atom has a SU (2) × SU (2) symmetry generated by the quantum mechanical angular momentum and Runge-Lenz vector

Syftet med denna uppsats är att beskriva hur det praktiska arbetet med intern kontroll går till i ett företag och på så sätt bidra med en ökad förståelse för intern

Majoriteten av revisorerna i denna studie anser att standarden inte är anpassad till mindre ägarledda kunder, vilket kan göra det svårare för non-Big X byråer att anpassa sig

Alexandersson conjectured an instance of the cyclic sieving phenomenon on rectangular coinversion–free fillings and the group action is given by cyclically shifting the columns....

This topic is of particular interest in the field of algebraic dynamics as recent results show that there exists a connection between the geo- metric location of periodic points