Download Chapter 5 - KFUPM Faculty List

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Fundamental theorem of algebra wikipedia , lookup

Eigenvalues and eigenvectors wikipedia , lookup

Cubic function wikipedia , lookup

Quadratic equation wikipedia , lookup

Quartic function wikipedia , lookup

Elementary algebra wikipedia , lookup

History of algebra wikipedia , lookup

System of polynomial equations wikipedia , lookup

System of linear equations wikipedia , lookup

Equation wikipedia , lookup

Transcript
Chapter 6
Power Series Solutions of Linear Differential Equations
6.1
Review of Properties of Power Series
6.2
Solutions about Ordinary Points
6.3
Solutions about Regular Singular Points - The Method of Frobenius
6.4
Bessel’s Equations and Functions
6.5
Legendre’s Equations and Polynomials
6.6
Orthogonality of Functions
6.7
Sturm – Liouville Theory
6.8
Exercises
We have seen in chapter 5 that we can solve linear differential
equations of order two or more with constant coefficients. The Cauchy-Euler
equation is exception. In fact most linear differential equations of higher order
with variable coefficients cannot be solved in terms of elementary functions.
The usual strategy for solving such type of equations is to assume a solution
in the form of an infinite series and proceed in a manner similar to the method
of undetermined coefficients (Section 5.6). Since these series solutions often
turn out to be power series, it is appropriate to summarise properties of power
series in the first section of this chapter. We conclude this chapter with the
Sturm-Liouville theory dealing with eigenvalues and eigenfunctions. StrumLiouville’s differential equation includes Bessel’s and Legendre’s equations as
special cases. Examples of Strum-Liouville problems are presented.
6.1 Review of Properties of Power Series
A power series in (x-a) is an infinite series of the form

c0+ c1 (x-a) + c2 (x-a)2 +- - - - =

c n ( x  a)n
(6.1)
no
Series of (6.1) is also called a power series centered at a. The power
series centered at a=0 is often referred as the power series, that is, the

series

c n x n A power series centered at a is called convergent
at a
n o
N
specified value of x if its sequence of partial sums S N(x) =  c n ( x  a)n , that
no
is, {SN (x)} is convergent. In other words the limit of {SN (x)} exists. If the limit
does not exist the power series is called divergent. The set of points x at
which the power series is convergent is called the interval of convergence of

the power series.
For R >o, a power series

c n ( x  a)n converges if
no
x  a <R and diverges if x  a >R. If the series converges only at a then R=0,
and if it converges for all x then R=. x  a <R is equivalent to a-R<x<a+R. A
power series may or may not converge at the end points a-R and a+R of this
interval.
A power series is called absolutely convergent if the series

c
n
( x  a)n
converges. A power series converges absolutely within its
no
interval of convergence. By the Ratio test a power series centered at a, series
given in (6.1) is absolutely convergent if L= x-a lim
n 
161
c n1
cn
is less than 1,
that is, L <1, the series diverges if L>1, and test fails if L=1. A power series


defines a function f(x)=
c n ( x  a)n whose domain is the interval of
no
convergence of the series. If the radius of convergence R>o, then f is
continuous, differentiable and integrable on the interval (a-R, a+R). Moreover
f’(x) and f(x)dx can be found by term by term differentiation and integration.
Convergence at an endpoint may be either lost by differentiation or gained
through integration.

Let y =

cnxn
n o

y' =

nc n x n1
no
y” =


n(n  1)c n x n2
n o
We observe that the first term in y' and first two terms in y' are zero.
Keeping this in mind we can write

y' =

n 1

y'' =

n2

Identity property:
If

nc n x n1
(6.2)
n(n  1)c n x n2
c n ( x  a)n =0, R>o for all x in the interval of
no
convergence, then cn=0 for all n.
162
Analytic at a point.
A function f is analytic at a point a if it can be
represented by a power series in x-a with a positive or infinite radius of
f ( n ) (a )
convergence. A power series where cn=
, that is, the series of the type
n!


n o
cn
f ( n ) (a )
( x  a)n is called the Taylor series. If a=o then Taylor series is
n!
called Maclaurin series. In calculus it is shown that ex, cos x, sin x, ln (x-1)
can be written in the form of a power series more precisely in the form of
Maclaurin series. For example
x2
---2!
x3 x5
sin x  x 

---3! 5!
x2 x4 x6
cos x  1 


---2! 4! 6!
for | x |  .
e x  1 x 
Arithmetic of Power Series: Two power series can be combined through the
operation of addition, multiplication, and division. The procedures for power
series are similar to those by which two polynomials are added, multiplied,
and divided. For example:



x2 x3 x4
x3
x5
x7
e x sin x  1  x 


 - - - -  x 


 - - - - 
2
6 24
6 120 5040



1
1  5
 1 1
 1 1
 1
 (1)x  x 2      x 3      x 4  - - - - 


x  - - -  6 2
 6 6
 120 12 24 
 x  x2 
x3 x5

---3 30
Since the power series for ex and sin x converge for x<,
product series converges on the same interval.
163
the
Shifting the Summation Index: In order to discuss power series solutions of
differential equations it is advisable to learn combining two or more
summations as a single summation.

Example 6.1 Express

n(n  1)c n x n2 +
n2


c n x n1 as one power series.
n 0
Solution: In order to add the two given series, it is necessary that both
summation indices start with the same number and the powers of x in each
series be such that if one series starts with a multiple of x to the first power,
then we want that the other series to start with the same power. In this
problem the first series starts with xo where as the second series starts with
x1. By writing the first term of the first series outside the summation notation,


n(n  1)c n x
n 2
n2

+  cn x
n 1

=2.1c2x0+

n(n  1)c n x
n 3
n 0
n 2

+

c n x n1
n 0
Both series on the right hand side start with the same power of x,
namely x1. Let k=n-2 and k=n+1 respectively in first and second series. Then
the right hand becomes


k 1
k 1
2 c2+  (k  2)(k  1)c k 2 x k   c k 1x k
(6.3)
Keeping in mind that it is the value of the summation index that is
important not the summation index which is a dummy variable say k=n -1 or
k=n+1. Now we are in position to add the series in (6.3) term by term and we
have


n2

n(n  1)c n x n2 +  c n x n1
n 0
164

=2c2+  [(k  2)(k  1)c k 2 x  c k 1 ]x k .
k 1
6.2 Solution about Ordinary Point
We look for power series solution of linear second-order differential
equation about a special point:
a 2 ( x)
d2 y
dy
 a1 ( x )
 a 0 ( x )y  0
2
dx
dx
(6.4)
where a2 (x)  0.
This can be put into the standard form
d2 y a1( x ) dy a 0 ( x )


y0
dx 2 a 2 ( x ) dx a 2 ( x )
d2 y
dy
or
 P( x )
 Q( x )y  0
2
dx
dx
(6.5)
A point xo is said to be an ordinary point of the differential equation
(6.4) if P(x) and Q (x) of (6.5) are analytic at xo, that is, P(x) and are Q(x)
represented by a power series. A point that is not an ordinary point is called a
singular point.

A solution of the form y =  c n ( x  x o )n is said to be a solution about
n0
the ordinary point x0.
Remark 6.1 It has been proved that if x=x0 is an ordinary point of (6.4)
then there exist two linearly independent solutions in the form of a power

series centered at x0, that is, y =

c n ( x  x o )n . A series solution converges
n0
165
at least on some interval defined by
x  x o <R, where R is the distance
from xo to the closest singular point.
Power series solution about an ordinary point:

Let
y=

c n x n and substitute values of y,
n 0
dy
dy 2
 y' , 2  y" in
dx
dx
(6.5)
Combine series as in Example 6.1, and then equate all coefficients to
the right hand side of the equation to determine the coefficients c n. We
illustrate the method by the following examples. We also see through these

examples how the single assumption that y=

c n x n leads to two sets of
n0
coefficients, so we have two distinct power series y1 (x) and y2(x) both
expanded about the ordinary point x=0. The general solution of the differential
equation is y=C1y1(x)+C2y2(x), infact it can been shown that C1=c o and C2=c1.
d2 y
The differential equation
 xy  0 is known as Airy’s equation and
dx 2
used in the study of diffraction of light, diffraction of radio waves around the
surface of the earth, aerodynamics etc. We discuss here power series
solution of this equation around its ordinary point x=0.
Example 6.2 Write the general solution of Airy’s equation y'+xy=0.
Solution: In view of the remark, two power series solutions centred at 0,


n o
n2
convergent for x < exist. By substituting y=  c n x n , y  =  n(n  1)c n x n2
into Airy’s differential equation we get
166


n2
n 0
y''+xy=  c n n (n  1)x n2  x  c n x n ,

=

n2

c n n (n  1)x n2   c n x n1
(6.6)
n 0
As seen in the solution of Example 6.1, (6.6) can be written as

y''+xy=2c2+  [(k+1) (k+2)ck+2+ck-1]xk=0
(6.7)
k 1
Since (6.7) is identically zero, it is necessary that coefficient of each
power of x be set equal to zero, that is,
2c2=0 (It is the coefficient y x0) and
(k+1)(k+2) ck+2+ck-1=0, k=1,2,3 - - - -- - - -..
(6.8)
The above holds in view of the identity property. It is clear that c2=0.
The expression in (6.8) is called a recurrence relation and it determines the
ck in such a manner that we can choose a certain subset of the set of
coefficients to be non-zero. Since (k+1)(k+2)0 for all values of k, we can
solve (6.8) for ck+2 in terms of ck-1.
ck+2= -
c k 1
, k  1,2,3, - - - (k  1)(k  2)
For k=1, c3 = -
co
2.3
For k = 2, c4 = -
co
3.4
For k= 3, c5 = -
c2
= 0 as c2=0
4 .5
For k= 4, c6 = -
c3
1
c0
=
2 . 3 .5 . 6 .
5.6
167
(6.9)
For k= 5, c7 = -
 c4
1

c1
6.7 3.4.6.7
For k= 6. c8 = -
 c5
 0 as c5=0
7.8
For k= 7. c9 = -
c6
1

c0
8 .9
2 . 3 .5 .6 .8 . 9 .
For k = 8, c10 = -
For k = 9, c11= -
c7
1

c1
9.10
3.4.6.7.8.10
c8
 0 as c8=0
10.11
and so on,

Substituting the coefficients just obtained into y=  c n x n
n 0
=c0+c1x+c2x2+c3x3+c4x4+c5x5+c6x6+c7x7+c8x8+c9x9+c10x10- - - we get
y=c0+c1x+0

c o 3 c1 4
c0
c1
c0
c1
x 
x 0
x6 
x7  0 
x9 
x10  0  - - - 2.3
3.4
2.3.5.6
3.4.6.7
2.3.5.6.8.9.
3.4.6.7.9.10
After grouping the terms containing co and the terms containing c1, we
obtain y=c0y1(x)+c1y2(x), where
y1(x)=1
= 1+ 
k 1
1 3
1
1
x 
x6 
x9  - - - 2 .3
2 . 3 .5 .6
2 .3 .5 . 6 . 8 . 9
( 1)k
x 3k
2.3 - - - - (3k  1)(3k )
y2(x) = x -
1 4
1
1
x 
x7 
x 10  - - - 3 .4
3 .4 .6 .7
3.4.6.7.9.10
168
1
( 1)k
x 3k
3.4 - - - - (3k )(3k  1)

= x+ 
k 1
Since the recursive use of (6.9) leaves c0 and c1 completely
undetermined, they can be chosen arbitrarily.
y=c0y1(x)+c1y2(x) is the general solution of the Airy’s equation.
Example 6.3 : Find two power series solutions of the differential equation y"xy=0 about the ordinary point x=0.

Solution: Substituting y =  c n x n into the differential equation we get
n0

y"-xy=

n2

n(n  1)c n x n2   c n x n1
n 0


k 0
k 1
=  (k  2)(k  1)c k  2 x k   c k 1 x k

= 2c2 +  [(k  2)(k  1)c k 2  c k 1 ]x k
k 1
Thus c2 = 0,
(k+2)(k+1)ck+2 –ck-1= 0
and
c k 2 
1
c k 1,k  1,2,3....
(k  2)(k  1)
Choosing co= 1 and c1=0 we find
c3 
1
1
, c 4  c 5  0, c 6 
and so on.
6
180
For c0=0 and c1=1 we obtain
169
c 3  0, c 4 
1
1
, c 5  c 6  0,, c 7 
and so on. Thus two solutions
12
504
are
y1 = 1 
1 3
1 6
x 
x  - - - - and
6
180
y2  x 
1 4
1 7
x 
x ---12
504
6.3 Solutions about Regular Singular Points – The Method of Frobenius
A singular point x0 of (6.4) is called a regular singular point of this
equation
if the functions p(x) = (x-xo) P(x) and q(x)=(x-xo)2Q(x) are both
analytic at x0. A singular point that is not regular is said to be on irregular
singular point of the equation. This means that one or both of the functions
p(x)=(x-x0) P(x) and q(x) = (x-x0)2Q(x) fail to be analytic at x0.
In order to solve a differential equation given by (6.4) about a regular
singular point we employ the following theorem due to Frobenius.
Theorem 6.1 (Frobenius Theorem) If x=x0 is a regular singular point of the
differential equation (6.4), then there exists at least one solution of the form

y=(x-xo)r

n o

c n ( x  x o ) n   c n ( x  x o ) n r
n 0
where r is constant to be determined. The series will converge at least
on some interval 0<x-x0<R.
The method of Frobenius: Finding series solutions about a regular singular
point x0, is similar to the method of previous section in which we substitute y=
170


c n ( x  x o )nr into the given differential equation and determine the
n o
unknown coefficients cn by a recurrence relation. However, we have an
additional task in this procedure. Before determining coefficients we must find
unknown exponent r. Equate to 0 the coefficient of the lowest power of x. This
equation is called the indicial equation and determines the value(s) of the
index r.
If r is found to be number that is not a non negative integer, then the

corresponding solution y=

c n ( x  x o )nr is not a power series. For the sake
n o
of simplicity we assume that the regular singular point is x=0.
Example 6.4 Apply the Method of Frobenius to solve the differential equation
2x y"+3y’-y=0 about the regular singular point x=0.
Solution: Let us assume that the solution is of the form

y=

c n x nr then
no

y' =

c n (n  r )x nr 1
n o

y"=

c n (n  r )(n  r  1)x nr 2 ,
n o
Substituting these values of y', y' and y'' into 2x y''+3 y'-y=0, we get

2

n o
c n (n  r )(n  r  1)x nr 1  3


n o
171
c n (n  r )x nr 1 


n o
c n x n  r  0.
Shifting the index in the third series and combing the first two yields


n o

c n (n  r ) (2n  2r  1)x nr 1 -  c n1x nr 1 =0
n o
Writing the term corresponding to n=0 and combining the terms for n/
into one series,

cor(2r+1)xr-1+  [c n (n  r ) (2n+2r+1)-cn-1]xn+r-1 = 0
n1
Equating the coefficients of xr-1 to zero yields the indicial equation
c0r(2r+1)=0
Since c0 0, either r=0 or = -
1
2
Hence two linearly independent solutions of the given differential
equation have the form

y1 = F0 (x) =  c n x n and
n o

y2 = F 1/ 2 (x) =x-1/2

c *n xn
no
Since cn(n+r) (2n+2r+1) -cn-1=0 for all n  1, we have the following
information on the coefficients for the two series:
(i)
co is arbitrary, and for n1, cn=
1
c n1
n(2n  1)
(ii)
c*o is arbitrary, and for n1,cn*=
1
*
c n1
n(2n  1)
Iteration of the formula for cn yields
n=1, c1 =
2c
1
2
c0 
c0  0
1.3
1.2.3
3!
172
22 c 0
1
1
n= 2, c2=
c1 
c0 
2.5
2.3.5
5!
n= 3, c3 =
1
1 22 c 0 23 c 0
c2 

3.7
3.7 5!
7!
2
to make the denominator (2n+1)!. The
2
Each term of cn was multiplied by
general form of cn is then
cn =
2n c 0
(2n  1)!
Similarly, the general form of cn*is found to be cn* =
2n c 0
.
(2n)!
The two solutions are

2n
y1=co 
xn,y2= co*x-1/2
n  o ( 2n  1)!


n o
2n n
x
(2n)!
y2 is not a power series.
Example 6.5 Apply the method of Frobenius to obtain two linearly
independent series solution of the differential equation
2x y" – y'+2y= 0
about a regular singular point x=0 of the differential equation.

Solution: Substituting y =  c n x nr ,
n o

y' =

c n (n  r )x nr 1
n o
and
n  r 1

y" =
c
no
n
(n  r )(n  r  1)x
173
into the differential equation and collecting terms, we obtain

2x y''- y'+2y=(2r2-3r)c0xr-1+ 
[2(k+r-1)(k+r)ck -(k+r)ck+2ck-1]xk+r-1=0,
k 1
which implies that
2r2-3r=r(2r-3)=0
and
(k+r)(2k+2r-3)ck+2ck-1=0.
3
The indicial roots are r=0 and r= .For r=0 the recurrence relation is
2
2c k 1
, k= 1,2,3, - - - k( 2k  3 )
ck = and
c1 = 2c0, c2= - 2c0, c3=
4
c0
9
For r=
3
the recurrence relation is
2
ck = -
2c k 1
, k=1,2,3,- - - ( 2k  3 )k
and
c1= -
2
2
4
c 0 ,c 2 
c 0 ,c 3  
c 0.
5
35
945
The general solution is
y = C1 (1+2x-2x2+
+C2 x3/2 (1-
4 3
x +- - - - )
9
2 2 2 4 3
+
xx +- - - -)
5 35
945
174
6.4 Bessel's equation
x2 y''+x y'+(x2-v2)y=0
(6.10)
(6.10) is called Bessel's equation.
Solution of Bessel's Equation:
Because x=0 is a regular singular point of Bessel's equation we know that

there exists at least one solution of the form y=  c n x nr . Substituting the last
no
expression into (6.10) gives



n o
n o
n o
x2 y"+x y'+(x2-v2)y=  c n (n  r )(n  r  1)x nr +  c n (n  r )x nr +  c n x nr  2

-v2  c n x nr = c0(r2-r+r-v2)xr
no

+xr

n 1

c n [(n  r )(n  r  1)  (n  r )  v 2 ]x n  x r  c n x n 2
n o


n 1
no
r
= c 0 (r 2  v 2 )x  x r  c n [(n  r )2 ]  v 2 ]x n  x r  c n x n  2
(6.11)
From (6.11) we see that the indicial equation is r2-v2=0, so the indicial roots
are r1=v and r2 = -v. When r1=v, (6.11) becomes

xv

n 1
c nn(n  2v )x  x
n

v

c n x n 2
n o




=xv (1  2v )c 1x   c nn(n  2v )x n   c n x n2 
n2
n 0





=xv (1  2v )c 1x   [(k  2)(k  2  2v )c k 2  c k ]x k 2   0
k 0


Therefore by the usual argument we can write (1+2v)c1=0 and
175
(k+2) (k+2+2v)ck+2+ck=0
or ck+2=
 ck
, k  0,1,2,- - - (k  2)(k  2  2v )
(6.12)
The choice c1=0 in (6.12) implies c3=c5=c7= - - - - = 0, so for k=0,2,4, - - - - we
find, after letting k +2 = 2n, n = 1,2,3, - - - - that
c2n = -
c 2n2
(6.13)
2 n(n  v )
2
Thus c2 = -
c0
2 .1(1  v )
2
c4 = -
c0
c2
 4
2 2(2  v ) 2 .2.1(1  v )( 2  v )
c6 = -
c0
c4
 6
2 .3(3  v )
2 .1.2.3(1  v )( 2  v )(3  v )
2
2
:
:
c2n =
( 1)n c 0
,n  1,2,3,- - - 2 2 n! (1  v )( 2  v )...( n  v )
(6.14)
It is standard practice to choose c0 to be specific value – namely.
c0 =
1
2 (1  v )
v
where  (1+v) is the gamma function. (See Appendix) Since this latter
function possesses the convenient property  (1+) = (), we can reduce
the indicated product in the denominator of (6.14) to one term.
For example:
 (1+v+1)= (1+v)  (1+v)
176
 (1+v+2)= (2+v)  (2+v)= (2+v)(1+v)(1+v).
Hence we can write (6.14) as
c 2n 
( 1)n
( 1)n

2 2n v n! (1  v )( 2  v )...( n  v )(1  v ) 2 2n v n! (1  v  n)
for n=0,1,2, - - - Bessel Function of the First Kind: Using the coefficients c2n just obtained

and r=v, a series solution of (6.10) is y=  c 2n x 2n v This solution is usually
n0
denoted by Jv ( x) :

( 1)n
x
 
n0 n! (1  v  n)  2 
Jv ( x )  
2n v
(6.15)
.
If v0, the series converges at least on the interval [o,  ). Also, for the second
exponent r2= -v we obtain, in exactly the same manner,

( 1)n
x
Jv ( x )  
 
n0 n! (1  v  n)  2 
2n v
(6.16)
.
The functions Jv(x) and J-v(x) are called Bessel functions of the first kind of
order v and –v, respectively. Depending on the value of v, (6.16) may contain
negative powers of x and hence converge on (0,  ).*
6.5 Legendre's Equation
(1-x2) y"-2x y'+n(n+1)y = 0
(6.17)
Equation (6.17) is known as Legendre's equation.
*
When we replace x by
x , the series given in (6.15) and (6.16) converge for 0< x <  .
177
Solution of Legendre's Equation: Since x=0 is an ordinary point of the

equation, we substitute the power series y=  c n x n , , shift summation indices,
n 0
and combine series to get
(1-x2) y"-2x y'+n(n+1)y=[n(n+1)c0+2c2]+[(n-1)(n+2)c1+6c3]x

+  [ j  2)( j  1)c j 2  (n  j)(n  j  1)c j ]x j  0,
j 2
which implies that
n(n+1)c0+2c2=0
(n-1)(n+2)c1+6c3=0
(j+2)(j+1)cj+2+(n-j)(n+j+1)cj=0
or
c2= -
c3 = -
c j 2 
n(n  1)
c0
2!
(n  1)(n  2)
c1
3!
(n  j)(n  j  1)
c j , j  2,3,4,- - - ( j  2)( j  1)
(6.18)
If we let j take on the values 2,3,4, - - - -, the recurrence relation (6.18) yields
c4 

(n  2)(n  3)
(n  2)n(n  1)(n  3)
c2 
c0
4.3
4!
c5  
(n  3)(n  4)
(n  3)(n  1)(n  2)(n  4)
c3 
c1
5. 4
5!
c6  
(n  4)(n  5)
(n  4)(n  2)n(n  1)(n  3)(n  5)
c4  
c0
6. 5
6!
C7  
(n  5)(n  6)
c5
7. 6
(n  5)(n  3)(n  1)(n  2)(n  4)(n  6)
c1
7!
178
and so on. Thus for at least |x| <1 we obtain two linearly independent power
series solutions:
 n(n  1) 2 (n  2)n(n  1)(n  3) 4
y 1( x )  c 0 1 
x 
x
2!
4!


(6.19)
(n  4)(n  2)n(n  1)(n  3)(n  5) 6

x  - - - -
6!


(n  1)(n  2) 3 (n  3)(n  1)(n  2)(n  4) 5
y 2 ( x)  c 1 x 
x 
x
3!
5!


(n  5)(n  3)(n  1)(n  2)(n  4)(n  6) 7

x  - - - -
7!

Notice that if n is an even integer, the first series terminates, whereas
y2(x) is an infinite series. For example, if n=4, then
35 4 
 4 . 5 2 2 .4 .5 . 7 4 

y1( x )  c 0 1 
x 
x   c 0 1  10 x 2 
x .
2!
4!
3




Similarly, when n is an odd integer, the series for y2(x) terminates with
xn; that is, when n is a nonnegative integer, we obtain an nth-degree
polynomial solution of Legendre's equation.
Since we know that a constant multiple of a solution of
Legendre's
equation is also a solution, it is traditional to choose specific values for c0 or
c1, depending on whether n is an even or odd positive integer, respectively.
For n=0 we choose c0=1, and for n = 2,4,6, - - - -,
c 0  ( 1)n / 2
1.3 - - - - (n  1)
;
2 .4 - - - - n
where as for n=1 we choose c1 = 1, and for n=3,5,7, - - - -,
179
1.3 - - - - n
.
2.4 - - - - (n  1)
c 1  ( 1)(n1) / 2
For example, when n=4, we have
y1( x )  ( 1)4 / 2


1. 3 
35 4  1
1  10 x 2 
x   35 x 4  30 x 2  3 .

2. 4 
3
 8
Legendre Polynomials These specific nth-degree polynomial solutions are
called Legendre polynomials and are denoted by Pn(x). From the series for
y1(x) and y2(x) and from the above choices of c0 and c1 we find that the first
several Legendre polynomials are
P0(x) =1
P1(x) = x
P2 ( x ) 
1
(3 x 2  1)
2
P4 ( x ) 
1
(35 x 4  30 x 2  3)
8
P3 ( x ) 
1
(5 x 3  3 x )
2
P5 ( x ) 
(6.20)
1
(63 x 5  70 x 3  15 x ).
8
Remember, P0(x), P1(x), P2(x), P3(x), - - - - are, in turn, particular solutions of
the differential equations
Properties
n = 0:
(1 - x 2 )y"-2xy'  0
n = 1:
(1  x 2 )y" 2xy' 2y  0
n = 2:
(1  x 2 )y" 2xy' 6y  0
n = 3:
(1 x 2 )y" 2xy' 12y  0
(6.21)
You are encouraged to verify the following properties using the
Legendre polynomials in (6.20)
(i)
Pn(-x)=(-1)nPn(x)
(ii)
Pn(1)=1
(iii)
Pn(-1)=(-1)n
180
(iv)
Pn(0)=0, n odd
(v)
P'n (0)  0 , n even
6.6 Orthogonality of Functions
The concept of orthogonality of functions is the generalization of the
notion of orthogonality or perpendicularity of two vectors in the plane.
Deprition 6.1 (i) (Orthogonal Function) Two functions 1 and 2 defined on
an interval (a,b) into R are said to be orthogonal if
b

1(x) 2 (x) dx = 0, 12
a
 0, 1= 2
(ii) A set of real-valued functions {1(x), 2 (x)- - - -} is said to be
orthonormal if
b

m(x) n (x) dx = 0, mn
a
0, m=n
(iii) A set of real-valued functions {0(x), 1 (x),2 (x)- - - -} is said to be
orthonormal if
b

m(x) n (x) dx = 0, mn
a
=1, m=n
In other words if {n(x)} is an orthogonal set of functions on the interval
b
[a,b] with the property that

|n(x)2dx = 1 for n= 0,1,2,3- - - -then {n(x)} is
a
orthonormal set on the interval.
181
(iv) A set of functions {0,1,2,
- - - -
} is said to be orthogonal with
b
respect to weight function p(x), if

m (x) n(x) p (x) dx = 0, mn
a
 0, m=n
Example 6.6 The set {1, cos x, cos 2x,- - - -} is orthogonal on the interval [,]
Verification: If we make the identification o(x)=1 and n(x) = cos nx, we

must then show that

0 ( x )n ( x)dx  0,if n0, and




m ( x)n ( x)dx  0,m  n
We have, in the first place,




o ( x)n ( x)dx   cos n x d x

1

  sin nx 
n
 

1
sin n  sin( n)  0, n  0.
n
In the second place




m
( x ) n ( x )dx   cos mx cos nx dx


1
  cos (m  n)x  cos (m  n)x dx, by using a well
2 
known trigonometric identity,
182

1  sin(m  n)x sin(m  n)x 
 

2 mn
m  n   
m  n.
= 0,

Example 6.7 (i) Compute
| 
n
( x ) |2 dx where

 o ( x )  1, 1 ( x )  cos x,  2 ( x )  cos 2x - - - -
 1 cos x cos 2x

Show that the set 
,
,
,- - - -


 2

is orthonormal on the interval [-,].


2
 | 0 ( x) | dx   dx
Solution: (i) For o(x) =1 we have


 x    ( )  2


 |  ( x) |
For n (x) = cosnx,n>0,
n


=
 cos
2
nxdx

1

2

 [  cos 2 nx ]dx

=

Thus for n>0,
 |  ( x) |
n
2
dx  

183

2
dx   | cos nx |2 dx

1

2
or   | n ( x ) |2 dx   
 

We are required to show that

(a)  |
o ( x)
2


(b)
|
n ( x )


|2 dx  1
|2 dx  1

Verification of (a) :
|

=
1
2
2
|2 dx 

| 
0

 1dx  1

|
n ( x )



|2 dx 
1
| n ( x ) |2 dx

 

1
=  | cos nx |2 dx
 

=  cos 2 nx dx


1
[1 cos 2 nx ]dx
2 

1
2


Verification of (b):
0 ( x)
2
1
2
Orthogonal Series Expansion
184
( x ) |2 dx
Let {n (x)} be an infinite orthonormal set of functions on interval
[a,b]. and f(x) be a function defined on [a,b]. Then f(x) can be written as
f(x)=coo(x)+c12(x)+c22(x)+- - - - +cnn(x)+ - - - -
(6.22)
b
where c n 
 f ( x ) ( x )dx  1
 |  ( x ) | dx
n
a
b
a
(6.23)
2
n
n=0, 1,2,3 …
The series on the right hand side of (6.22) is called orthogonal
expansion of f(x) defined on [a,b] in terms of the orthonormal set of functions
{n(x)} defined on [a,b]. cn's given by (6.23) are called coefficients of
orthogonal expansion of f. If orthonormal set of Example 6.6 is considered we
get cosine Fourier expansion of f(x), that is, (6.22) will be cosine Fourier
series and (6.23) will give cosine Fourier coefficients. One can consider
expansion of a function in terms of Bessel's orthonormal set of functions and
Legendre's orthonormal set of functions.
6.7
Sturm-Liouville Theory
Consider the linear differential equation of order two
y"+R(x) y'+(Q(x)+ P(x)) y=0
(6.24)
Given an interval on which the coefficients R(x) and (Q(x)+ P(x) are
continuous we seek values of  for which (6.24) has non-trivial solutions.
We can also seek values of  when (6.24) is given with boundary
conditions. Let us put this differential equation in a more convenient form.
Multiply (6.24) by r=e  R( x )dx
to get
185
y''e  R( x )dx + R(x) y'e  R( x )dx +(Q(x)+  P (x)) ye  R( x )dx = 0
(6.25)
Since r(x)0, equation (6.25) has the same solutions as (6.24). This
equation can be written as
(r y')' + (q+ p) y=0
(6.26)
where q(x)=Q(x) e  R( x )dx , p(x) = P(x) e  R( x )dx
Equation (6.26) is called the Sturm-Liouville differential equation,
or the Sturm-Liouville form of equation (6.24). Through out this section we
assume that p.q, and r and r’ are continuous on [a,b] or at least on (a,b), and
p(x) >0 and r(x) >0 on (a,b).
Remark 6.2 Bessel’s equation given by (6.10)
and Legendre’s equation given by (6.17) are special cases of the
Sturm-Liouville differential equation (6.26).
For Bessel’s equation we can choose r(x)=
x2
, q(x)=x2,
2
p(x)=1, = -v2 in (6.26).
For Legendre’s equation we take r(x) = 1-x2
q(x)=0, p(x)=1, and  = n(n+1) in (6.26)
The Regular Sturm-Liouville Problem: Find numbers  for which there are
non-trivial solutions of (6.26) subject to the regular boundary conditions
having the following form
A1y(a) + A2 y' (a)=0, B1y(b) + B2 y' (b)=0
where A1 and A2 are given constants, at least one must be non-zero. Similarly
B1 and B2 are given constants, at least one must be non-zero.
186
The Periodic Sturm-Liouville Problem:
Find numbers  for which there are non-trivial solutions of (6.26) on an
interval [a,b] where r(a)=r(b) and subject to the periodic boundary conditions
y(a)=y(b), y' (a)= y' (b)
The Singular Sturm-Liouville Problem:
Find numbers  for which there are non-trivial solutions of the SturmLiouville equation on (a,b), subject
to one of the following three kinds of
boundary conditions:
Case I.
r(a)=0 and there is no boundary condition at a, while at b the
boundary condition is
B1y(b)+B2 y' (b)=0,
where B1 and B2 are not both zero.
Case 2. r(b)=0 and there is no boundary conditions at b, while at a the
condition is
A1y(a)+A2 y' (a)=0,
with A1 and A2 not both zero.
Case 3. r(a)=r(b)=0, and no boundary condition is specified at a or b. We seek
solutions that are bounded functions on [a,b].
Definition 6.2 A number  for which Sturm-Liouville differential equation,
(6.26), subject to boundary conditions of one of these three problems, has
nontrivial solution is called an eigenvalue of the problem. A corresponding
nontrivial solution is called an eigenfunction associated with this eigenvalue.
187
Remark 6.3 (i) The zero function cannot be an eigenfunctiion. Any nonzero
constant multiple of an eigenfunction is an eigenfunction.
(ii) In mathematical models of real systems, eigenvalues have some
physical meaning. For example in the study of wave motion the eigenvalues
are fundamental frequencies of vibration of the system.
The fundamental properties of Sturm-Liouville problems are described
by the following theorem which is considered as the heart of Sturm-Liouville
theory.
Theorem 6.2 (a) Each regular and each periodic Sturm-Liouville problem has
an infinite number of distinct real eigenvalues. If these are labeled 1, 2. - - -,
so that n<n+1, then lim  n=.
n .
(b) If n and m are distinct eigenvalues of any of the three kinds of SturmLiouville problems defined on an interval (a,b) and n and m are
corresponding eigenfunctions, then
b
 p( x) ( x)
n
m
( x )dx  0
a
(c) All eigenvalues of a Sturm-Liouville problem are real numbers.
(d)
For
a
regular
Sturm-Liouville
problem,
any two
eigenfunctions
corresponding to a single eigenvalue are constant multiples of each other.
Interested readers will find the proof of this theorem in references [.,.]
or at website no. [
].
Remark 6.4 (i) Part (a) assures existence of eigenvalues, at least for regular
and periodic problems. A singular problem may also have an infinite
188
sequence of eigenvalues say for example, for Bessel’s equation. This part
also asserts that the eigenvalues spread out so that if arranged in increasing
order, they increase without bound. For example, numbers 1-
1
could not be
n
eigenvalues of a Sturm-Liouville problem, since these numbers approach 1 as
n
(ii) Part (b) can be stated as “Eigenfunctions associated with distinct
eigenvalues are orthogonal on [a,b], with weight function p(x). The weight
function p is the coefficient of  in the Sturm-Liouville equation.
This orthogonality provides the possibility of expansion of functions in
series of eigenfunctions of a Sturm-Liouville problem, analogue of equation
(6.22) is possible for eigenfunctions.
(iii) Part (c) states that a Sturm Liouville problem can have no complex
eigenvalue.
(iv) Part (d) applies only to regular Sturm-Liouville problems.
Example 6.8 Discuss solutions of regular Sturm-Liouville problem:
y''+ y=0, y(0)=y(l)=0
on an interval [0,l] in cases (i)  = 0,
(ii)  is negative number, and (iii)  is positive number
Case (i) Let =0, then y''=0 and integrating it twice we get y(x)=cx+d for
some constants c and d. Now y(0)=d=0, and y(l)=cl=0 implies c=0. This
means that y(x)=cx+d must be the trivial solution. In the absence of a nontrivial solution, =0 is not an eigenvalue of this problem.
189
Case (ii) Suppose that  is negative, say =-k2 for k>0
Now y"-k2y=0. This is homogeneous linear differential equation with
constant coefficients. The auxiliary equation is m 2-k2=0. Roots are m1=k,m2= k. The general solution is
y(x)=c1ekx+c2e-kx
Since y(0)=c1+c2=0, then
c2=-c1, so y=c1(ekx-e-kx). since
sin hkx =
ekx  ekx
, we have
2
y=2c1 sinh kx. But then
y(l)=2c1 sinh kl=0
Since kl>0, sinh kl>0, so c1, = 0
This case also leads to the trivial solution, so this Sturm-Liouville
problem has no negative eigenvalue.
Case (iii)  is positive, say =k2
Now y''+k2y=0. The auxiliary equation of this homogeneous linear
differential equation with constant coefficients is
m2+k2=0. Roots are m1=ik, m2=-ik.
As discussed in Section 5.5, equation (5.18) the general solution is
y(x)=c1cos (kx) +c2 sin(kx)
Now
y(o)=c11 +c2.0=0 or c1=0
y(x)=c2 sin (kx). Finally, we need
190
y(l)=c2 sin kl=0
To avoid trivial solution, we need c20.
Then we must choose k so that sin kl=0, which means that kl must be a
positive multiple of , say kl = n.Then
n =
n2 2
for n=1,2,3,- - - -- - - -.
l2
Each of these numbers is an eigenvalue of this Sturm-Liouville
problem. Corresponding to each n, the eigenfunctions are
 nx 
yn(x) = c sin 
,
 l 
where c is any non-zero real number.
Example 6.9 Discuss solution of periodic Sturm-Liouville problem:
y''+y=0, y(-l)=y(l), y'(-l)= y'(l)
on an interval [-l,l] for cases
(i)
=0
(ii)  <0 (ii) >0
Solution: Case (i) =0 Then y=cx+d. (See example 6.8) Now
y(-l) = - cl+d=y(l)=cl+d implies c=0. The constant function y=d satisfies both
boundary conditions. Thus =0 is an eigenvalue with nonzero constant
eigenfunctions.
Case (ii) <0, say = -k2
Solving as in case (ii) of Example 6.8
y(x)=c1ekx+c2e-kx is the general solution.
Since y(-l)=y(l), then
191
c1e-kl+c2ekl=c1ekl+c2e-kl
(6.27)
And y' (-l)= y' (l) gives us after dividing out the common factor k
c1e-kl-c2ekl=c1ekl-c2e-kl
(6.28)
Rewrite equation (6.27), as
c1 (e-kl-ekl)=c2 (e-kl-ekl)
This implies that c1=c2. The equation (6.28) becomes
c1(e-kl-ekl) = c1(ekl-e-kl)
But this implies that c1=-c1, hence c1=0. This solution is therefore trivial, hence
this problem has no negative eigenvalue.
Case (iii)  >0, say =k2
Now as in Example 6.8 (case iii) the general solution is
y(x)=c1cos (kx)+c2 sin (kx)
Now
y(-l)=c1cos kl-c2 sin (kl)=y(l)=c1cos (kl)+c2sin (kl)
But this implies that
-c2sin (kl)=c2 sin (kl)
or –c2=c2 implying c2=0
Also y' (-l)=kc1sin (kl) + kc2 cos (kl)
= y' (l)=-kc1 sin (kl) +kc2 cos (kl).
Then
kc1 sin (kl)=0
192
If sin (kl)0, then c1=c2=0, leaving the trivial solution. Thus we assume
that sin kl=0 which requires that kl=n for some positive integer n. Therefore,
the numbers
n=
n2 2
ll
2
are eigonvalues for n=1,2,- - - - , with corresponding eigenfunctions
yn(x)=c1 cos (
nx
nx
) +c2 sin (
)
l
l
where c1 and c2 are not both zero
6.8
Exercises:
Review of Power Series
1.
Write excos x in the form of a power series. Examine whether this
power series is convergent.
Solution About Ordinary Points:
Find the general solution of the following differential equations about
an ordinary point in terms of two power series
2.
y''-(1+x)y=0
3.
y''+(cos x) y = 0
4.
y''+x2y=0
5.
y''+y=ex
6.
y'+xy=x2-2x
7.
(x2-1) y'+y=0
8.
y"-(x+1) y'-y=0
193
Use the power series method to solve the following initial value
problems
9.
(x-1) y''-x y'+y=0
y(0)=-2, y' (0)=6
10.
(x2+1) y''+2x y'=0, y(0)=0, y' (0)=1
11.
y"+xy=0, y(0)=1, y' (0)=1
12.
xy"+y+x=0, y(1)=1, y' (1)=1
Solution About Regular Singular Point: The method of Frobenius
Use the method of Frobenius to solve the following differential equations
13.
x y''-x y'+y=0
14.
y''+
15.
x y''+ y'+y=0
16.
2x y'-3 y'-
17.
4x2y''+(3x+1)y=0
18.
x y''-(x+5) y'+3y=0
19.
x y''+(x-5) y'+3y=0
20.
x y''+ y'+xy=0
3
y'-2y=0
x
3x
y=0
x
Bessel’s Equation
Find the general solution of the following equations
21.
x2 y''+x y'+(x2-1)y=0
22.
x y''+x y'+xy=0
23.
Verify that y=xnJn(x) is a particular solution of x y"+(1-2n) y'+xy=0, x>0
194
Legendre’s Equation
Solve the following equations
24.
(1-x2) y''-2x y'=0
25.
(1-x2) y''-2x y'+12y=0 subject to initial conditions
y(0)=0, y' (0)=1.
Sturm Liouville Theory
In each of problems 26 through 35, classify the Sturm-Liouville problem
as regular, periodic, or singular; state the relevant interval, find the
eigenvalues; corresponding to each eigenvalue, find an eigenfunction.
26.
y''+ y=0;
y' (0)= y' (l)=0
27.
y''+ y=0;
y(0)=0, 3y(1)+ y' (1)=0
28.
y''+ y=0;
y(0)=0, y' (l)=0
29.
y''+ y=0;
y' (0)=0, y' (l)=0
30.
y''+ y=0;
y' (0)=y(4)=0
31.
y''+ y=0;
y(0)=y(),y' (0)= y' ()
32.
y''+ y=0;
y(-3)=y(3), y' (-3)= y' (3)
33.
y''+ y=0;
y(0)=0, y()+2 y' ()=0
34.
y'+ y=0;
y(0)-2 y' (0)=0, y' (1)=0
35.
y''+2 y'+(1+)y=0,
y(0)=y(1)=0.
195