You are on page 1of 30

Piessens, R. “The Hankel Transform.


The Transforms and Applications Handbook: Second Edition.
Ed. Alexander D. Poularikas
Boca Raton: CRC Press LLC, 2000
9
The Hankel Transform

Robert Piessens 9.1 Introductory Definitions and Properties


Katholieke Universieit Leuven 9.2 Definition of the Hankel Transform
9.3 Connection with the Fourier Transform
9.4 Properties and Examples
9.5 Applications
The Electrified Disc • Heat Conduction • The Laplace Equation
in the Halfspace z > 0, with a Circularly Symmetric Dirichlet
Condition at z = 0 • An Electrostatic Problem
9.6 The Finite Hankel Transform
9.7 Related Transforms
9.8 Need of Numerical Integration Methods
9.9 Computation of Bessel Functions Integrals over a
Finite Interval
Integration between the Zeros of Jv(x) • Modified
Clenshaw–Curtis Quadrature
9.10 Computation of Bessel Function Integrals over an
Infinite Interval
Integration between the Zeros of Jv(x) and Convergence
Acceleration • Transformation into a Double Integral •
Truncation of the Infinite Interval
9.11 Tables of Hankel Transforms

ABSTRACT Hankel transforms are integral transformations whose kernels are Bessel functions. They
are sometimes referred to as Bessel transforms. When we are dealing with problems that show circular
symmetry, Hankel transforms may be very useful. Laplace’s partial differential equation in cylindrical
coordinates can be transformed into an ordinary differential equation by using the Hankel transform.
Because the Hankel transform is the two-dimensional Fourier transform of a circularly symmetric
function, it plays an important role in optical data processing.

9.1 Introductory Definitions and Properties


Bessel functions are solutions of the differential equation

x 2 y ′′ + xy ′ + (x 2 − p 2 )y = 0 (9.1)

where p is a parameter.
Equation (9.1) can be solved using series expansions. The Bessel function Jp(x) of the first kind and
of order p is defined by

© 2000 by CRC Press LLC


k
 1 2
p ∞ − x 
1   4 
J p ( x ) =  x
2  ∑k =0
k! Γ ( p + k + 1)
. (9.2)

The Bessel function Yp(x) of the second kind and of order p is another solution that satisfies

 J (x) Y p ( x ) 2
W ( x ) = det  p = .
 J ′p ( x ) Y p′ ( x ) πx

Properties of Bessel function have been studies extensively (see References 7, 22, and 26).
Elementary properties of the Bessel functions are
1. Asymptotic forms.

2  1 1 
J p (x ) ~ cos x − p π − π , x →∞. (9.3)
πx  2 4 

2. Zeros. Jp(x) and Yp(x) have an infinite number of real zeros, all of which are simple, with the
possible exception of x = 0. For nonnegative p the sth positive zero of Jp(x) is denoted by jp,s . The
distance between two consecutive zeros tends to π: lim ( jp,s+1 – jp,s ) = π.
s→∞
3. Integral representations.

p
1 
 x
2  π
Jp(x) =
π1/ 2 Γ ( p + 1 / 2) ∫0
cos( x cos θ )sin 2 p θ d θ . (9.4)

If p is a positive integer or zero, then


1
J p (x ) = cos(x sin θ − p θ)dθ
π 0
(9.5)
j −n π
=
π ∫ 0
e jx cos θ cos(p θ)dθ .

4. Recurrence relations.

2p
J p−1 ( x ) − J ( x ) + J p+1 ( x ) = 0 (9.6)
x p

J p−1 ( x ) − J p+1 ( x ) = 2 J ′p ( x ) (9.7)

p
J ′p ( x ) = J p−1 ( x ) − J (x) (9.8)
x p

p
J ′p ( x ) = − J p+1 ( x ) + J (x) . (9.9)
x p

© 2000 by CRC Press LLC


5. Hankel’s repeated integral. Let ƒ(r) be an arbitrary function of the real variable r, subject to the
condition that

∫ 0
f (r ) r dr

is absolutely convergent. Then for p ≥ –1/2

∞ ∞

∫ ∫
1
s ds f (r) J p (sr) J p (su)r dr = [ f (u +) + f (u −)] (9.10)
0 0 2
provided that ƒ(r) satisfies certain Dirichlet condtions.
For a proof, see Reference 26. The reader should also refer to Chapter 1, Section 5.6 for more
information regarding Bessel functions.

9.2 Definition of the Hankel Transform


Let ƒ(r) be a function defined for r ≥ 0. The vth order Hankel transform of ƒ(r) is defined as


Fν (s ) ≡  ν { f (r)} ≡
∫0
r f (r) J ν (sr)dr . (9.11)

If v > –1/2, Hankel’s repeated integral immediately gives the inversion formula


f (r ) =  −ν1{Fν (s )} ≡
∫ 0
sFν (s )J ν (sr )ds . (9.12)

The most important special cases of the Hankel transform correspond to v = 0 and v = 1. Sufficient but
not necessary conditions for the validity of (9.11) and (9.12) are
1. f (r) = O(r –k), r → ∞ where k > 3/2.
2. f ′ (r) is piecewise continuous over each bounded subinterval of [0, ∞).
3. f (r) is defined as [ f (r+) + f (r–)]/2.
These conditions can be relaxed.

9.3 Connection with the Fourier Transform


We consider the two-dimensional Fourier transform of a function ϕ(x,y), which shows a circular sym-
metry. This means that ϕ(r cos θ, r sin θ)  ƒ(r,θ) is independent of θ.
The Fourier transform of ϕ is

∞ ∞

∫ ∫
1
Φ(ζ, η) = f (x , y )e − j( x ζ , y η )dx dy . (9.13)
2π −∞ −∞

We introduce the polar coordinates

x = r cos θ , y = r sin θ

and

ζ = s cos ϕ , η = s sin ϕ .

© 2000 by CRC Press LLC


We have then

∞ 2π

∫ ∫
1
φ (s cos ϕ, s sin ϕ) ≡ F (s , ϕ) = r dr e − j r s cos(θ−ϕ ) f (r)d θ
2π 0 0

∞ 2π

∫ ∫
1
= rf (r)dr e − j r s cos α d α
2π 0 0


=
∫ 0
rf (r) J 0 (rs )d r .

This result shows that F(s,ϕ) is independent of ϕ, so that we can write F(s) instead of F(s, ϕ). Thus, the
two-dimensional Fourier transform of a circularly symmetric function is, in fact, a Hankel transform of
order zero.
This result can be generalized: the N-dimensional Fourier transform of a circularly symmetric function
of N variables is related to the Hankel transform of order N/2 – 1. If ƒ(r,θ) depends on θ, we can expand
it into a Fourier series

f (r , θ) = ∑ f (r )e
n =−∞
n
j nθ
(9.14)

and, similarly

∞ ∞ ∞

∫ ∫ ∑ F (s)e
1
F (s , ϕ) = r dr e − j r s cos(θ−ϕ ) f (r, θ)d θ = jn ϕ
(9.15)
2π 0 0
n =−∞
n

where


1
f n(r ) = f (r , θ)e − j n θd θ (9.16)
2π 0

and


1
Fn(s ) = F (s , ϕ)e − j n ϕd ϕ . (9.17)
2π 0

Substituting (9.15) into (9.17) and using (9.14), we obtain

2π 2π ∞

∫ ∫ ∫
1
Fn (s ) = e − jn ϕ dϕ dθ f (r, θ)e j s r cos(θ−ϕ )r dr
(2 π) 2 0 0 0

2π ∞ 2π ∞

∫ ∫ ∫ ∑f
1
= e − jn ϕ dϕ r dr e j s r cos(θ−ϕ )dθ × (r)e jm θ
(2 π) 2
m
0 0 0
m=−∞

∞ 2π

∫ ∫
1
= r dr e − jn α e j s r cos α f n (r)dα
(2 π) 0 0


=
∫0
rf n (r) J n (sr)dr

=  n { f n (r)}.

© 2000 by CRC Press LLC


In a similar way, we can derive

f n (r ) =  n {Fn (s )} . (9.18)

9.4 Properties and Examples


Hankel transforms do not have as many elementary properties as do the Laplace or the Fourier transforms.
For example, because there is no simple addition formula for Bessel functions, the Hankel transform
does not satisfy any simple convolution relation.
1. Derivatives. Let

Fν (s ) =  ν { f (x )} .

Then

 ν +1 ν −1 
G ν (s ) =  ν { f ′(x )} = s  Fν−1(s ) − Fν+1(s ) . (9.19)
 2ν 2ν 

Proof


Gν ( s) =
∫ xf ′( x ) J ( sx )dx
0
ν


d
= [ xf ( x ) Jν ( sx )]∞0 − f (x) [ xJ ( sx )]dx .
0 dx ν

In general, the expression between the brackets is zero, and

d sx
[ xJ ( sx )] = [(ν + 1) Jν−1 ( sx ) − (ν − 1) Jν+1 ( sx )] .
dx ν 2ν

Hence, we have (9.19).


2. The Hankel transform of the Bessel differential operator. The Bessel differential operator

2 2
d 2 1 d  ν 1 d d  ν
∆ν ≡ 2 + −  = r − 
dr r dr  r  r dr dr  r 

is derived from the Laplacian operator

∂2 1 ∂ 1 ∂2 ∂2
∇2 = + + +
∂r 2 r ∂r r 2 ∂θ 2 ∂z 2

after separation of variables in cylindrical coordinates (r, θ, z).


Let ƒ(r) be an arbitrary function with the property that lim ƒ(r) = 0. Then
r→∞

 ν {∆ ν f (r )} = −s 2  ν { f (r )} . (9.20)

© 2000 by CRC Press LLC


This result shows that the Hankel transform may be a useful tool in solving problems with cylindrical
symmetry and involving the Laplacian operator.

Proof Integrating by parts, we have

∞  d df ν 2 
 ν {∆ ν f (r)} =

0
 r
 dr dr

r
f (r) J ν (sr)dr

∞  2 ν2 

s
= s J ν′′ (sr) + J ν′ (sr) − 2 J ν (sr) f (r)r dr
0  x r 

= −s 2
∫0
rf (r) J ν (rs )dr

= −s 2  ν { f (r)}.

This property is the principal one for applications of the Hankel transforms to solving differential
equations. See References 2, 3, 24, 25, and 28.
3. Similarity.

1 s
 ν { f (ar )} = Fν   . (9.21)
a2  a

4. Division by r.

s
 ν {r −1 f (r )} = [F (s ) + Fν+1(s )] . (9.22)
2 ν ν−1

5.

 d 
 ν r ν−1 [r 1− ν f (r)] = −sFν−1(s ) . (9.23)
 dr 

6.

 d 
 ν r − ν−1 [r ν+1 f (r )] = sFν+1(s ) . (9.24)
 dr 
7. Parseval’s theorem. Let

Fν (s ) =  ν { f (r )}

and

G ν (s ) =  ν {g (r )} .

Then

© 2000 by CRC Press LLC


∞ ∞ ∞


0
Fν(s )G ν(s )s ds =
∫ 0
Fν(s )s ds
∫ r g (r )J (sr )dr
0
ν

∞ ∞
=
∫ 0
r g (r )dr
∫ sF (s)J (sx)ds
0
ν ν (9.25)


=
∫ r g (r )f (r )dr .
0

Example 9.1

From the Fourier pair (see Chapter 2) {e −a( x


2
+y2 )
} = (π / a)e −(ζ
2
+ η2 )/4a
and the Fourier transform

relationship { f ( x 2 + y 2 )} = 2πF0 ( ζ 2 + η2 ) ≡ 2πF0 (s ) , we obtain the Hankel transform

1 − s 2 / 4a
{e −ar } =
2
e , a >0 .
2a

Example 9.2
a a
From the relationship ∫ 0 r J 0 (sr) dr = ∫ 0 (1 / s )(d / dr)[rJ 1(sr)] = [aJ1(as)]/s (see Chapter 1, Section 5.6),
we conclude that

aJ 1(as )
 0 {pa (r )} =
s

where pa(r) = 1 for |r| < a and zero otherwise.

Example 9.3

From the identity ∫ 0 J 0 (sr) dr = 1 / s , s > 0 (see Chapter 1, Section 5.6), we obtain

1  1
 0  = .
r  s

Example 9.4

Since ∫ 0 r δ(r − a ) J 0 (sr) dr = aJ 0 (as ) (see Chapter 1, Section 2.4), we obtain

 0 {δ(r − a )} = aJ 0 (as ), a >0

and because of symmetry

 0 {aJ 0 (ar)} = δ(s − a ), a >0 .

Convolution Identity
Let ƒ1(r) and ƒ2(r) have Hankel transforms F1(s) and F2(s), respectively. From Section 2.4.1 above, we have

© 2000 by CRC Press LLC


 ∞ 


∫∫ −∞
f 1( x 12 + y 12 ) f 2 ( (x − x 1)2 + (y − y 1)2 )dx 1dy 1  = 4 π 2 F1(s )F2 (s ) .

Hence, we have

1
 0 { f1(r)★★f 2 (r)} =  { f (r)★★f 2 (r)} = 2 πF1(s )F2 (s ) .
2π ( 2) 1

Therefore, to find the inverse Hankel transform of 2πF1(s)F2(s), we convolve f 1( x 2 + y 2 ) with


f 2 ( x 2 + y 2 ), and in the answer we replace x 2 + y 2 by r. We can also write the above relationship
in the form

 0 {2 πf 1(r ) f 2 (r )} = F1(s )★★F2 (s ) .

Example 9.5
If ƒ1(r) = ƒ2(r) = [J1(ar)]/r then from the convolution identity above, we obtain

 J 2 (ar)  1
 0 2 π 1 2  = 2 pa (s )★★pa (s )
 r  a

where

 s s s2 
pa (s )★★pa (s ) =  2 cos −1 − 1− 2  a2 .
 2a a 4a 

Hence,

 J 2 (ar)   s s s2 
 0 2 π 1 2  =  2 cos −1 − 1 − 2  p2a (s )
 r   2a a 4a 

1 | s | ≤ 2a
p2a (s ) = 
0 otherwise .

Example 9.6
From the definition, the Hankel transform of r vh(a – r), a > 0 is given by

∫ ∫
a as
1
 ν {r νh (a − r )} = r ν+1J ν (sr )dr = x ν+1J ν (x )dx
0 s ν+ 2 0

since h(a – r) is the unit step function with value equal to 1 for r ≤ a and 0 for r > a. But ∫ t vJv–1(t) dt =
t vJv(t) + C (see Chapter 1, Section 5.6, Table 5.6.1) and hence,

(as ) ν+1 a ν+1 1


 ν {r ν h(a − r)} = J ν + (as ) = J (as ), a > 0, ν > − .
s ν+ 2
1
s ν+1 2

© 2000 by CRC Press LLC


Example 9.7
The Hankel transform of r v–1e –ar, a > 0 is given by

∞ ∞ a

∫ ∫
1 − t
 ν {r ν−1e − ar } = r νe − ar J ν(sr )dr = t ν J ν(t )e s
dt
0 s ν+1 0

1  a
= ν+1
L t ν J ν(t ); p = 
s  s

where we set t = rs and L is the Laplace transform operator (see also Chapter 5). But


(−1)n t 2n+2 ν
t ν J ν(t ) = ∑n=0
n ! Γ(n + ν + 1)22n+ ν

and, hence,


(−1) n
ν
{t J ν (t ); p} = ∑
n= 0
n! Γ(n + ν + 1)2 2n+ ν
L{t 2n+ 2 ν ; p}


(−1) n Γ(2n + 2 ν + 1)
= ∑ n! Γ(n + ν + 1)2
n= 0
2 n + ν 2 n + 2 ν+1
p
.

From Chapter 1, Section 2.5, the duplication formula of the gamma function gives the relationship

Γ(2n + 2 ν + 1) 1 2n+ 2 ν  1
= 2 Γ n + ν + 
Γ(n + ν + 1) π  2

and, therefore, the Laplace transform relation becomes

 1
∞ (−1)n Γ  n + ν +  n
ν
 2  1 

2
{t ν J ν (t ); p} =  2 .
π p 2 ν+1 n= 0
n! p 

The last series can be summed by using properties of the binomial series

∞ ∞
 − b n (−1) n Γ(n + b) n
(1 + x )− b = ∑
n= 0
 x =
n ∑
n= 0
n ! Γ(b)
x , |x| < 1

where the relation

 −b (−1)nb(b + 1)L(b + n − 1) (−1)n Γ(n + b)


 = =
n ! Γ(b)
n n!

was used. The Laplace transform now becomes

© 2000 by CRC Press LLC


 1  1
2 ν Γ ν +  2 ν Γ ν + 
 2  2
{t ν J ν (t ); p} = 1
= 1
, Re(p) > 1
ν+ ν+
π (p + 1)2 2  a   2 2

π   + 1
 s  
 

and, hence,

 1  1
2 ν Γ ν +  s ν 2 ν Γ ν + 
1  2  2 1
 ν {r ν−1e −ar } = ν+1 1
= 1
, ν>− .
s ν+ ν+ 2
 a  2  2
π (a 2 + s )
2 2

π   + 1
 s  
 

If we set v = 0 and a = 0 in the above equation, we obtain the results of Example 9.3. If we set v = 0, we
obtain

1
 0 {r −1e −ar } = , a >0 .
a +s2
2

Example 9.8
The Hankel transform  0{e –ar} is given by

d  2 − 
1
 0 {e − ar } = {r J 0 (sr); r → a } = − (s + a 2 ) 2 
da  

a
= , a >0
[s + a 2]3/ 2
2

since multiplication by r corresponds to differentiation in the Laplace transform domain.


Example 9.9
From Chapter 1, Section 5.6, we have the following identity:

d 2 rn ( x ) 1 drn ( x )
2
+ + rn ( x ) = 2nrn+1 ( x ) ,
dx x dx

where rn(x) = Jn(x)/x n.


Using the Hankel transform property of the Bessel operator, we obtain the relationship

(1 − s 2 )R n ( s ) = 2nR n+1 ( s )

or

1 − s2 (1 − s 2 )n
Rn+1(s ) = Rn(s ) = L = n R1(s ) .
2n 2 n!

© 2000 by CRC Press LLC


 J 1 (r ) 
But from Example 9.2,  0   = p1(s ) and, hence,
 r 

 J (r )  (1 − s 2 )n−1
 0  n n  = n−1 p1(s )
 r  2 (n − 1)!

where p1(s) is a pulse of width 2 centered at s = 0.


Example 9.10
If the impulse response of a linear space invariant system is h(r) and the input to the system is ƒ(r), then
its output is g (r) = ƒ (r)h(r) and, hence,

G ( s ) = 2πF ( s )H ( s ) .

Since  0{J0(ar)} = [δ(s – a)]/a (see Example 9.4) and ϕ(s)δ(s – a) = ϕ(a)δ(s – a), we conclude that if
the input is ƒ(r) = J0(ar), then

2π 2πH (a)
G ( s) = δ ( s − a) H ( s ) = δ ( s − a) .
a a

Therefore, the output is

g (r ) = 2πH (a) J0 (ar ) .

9.5 Applications

9.5.1 The Electrified Disc


Let υ be the electric potential due to a flat circular electrified disc, with radius R = 1, the center of the
disc being at the origin of the three-dimensional space and its axis along the z-axis.
In polar coordinates, the potential satisfies Laplace’s equation

∂ 2υ 1 ∂υ ∂ 2υ
∇ 2υ ≡ + + =0. (9.26)
∂r 2 r ∂r ∂z 2

The boundary conditions are

υ (r, 0) = υ 0 , 0 ≤ r <1 (9.27)

∂υ
(r, 0) = 0, r >1 . (9.28)
∂z

In (9.27), υ0 is the potential of the disc. Condition (9.28) arises from the symmetry about the plane z = 0.
Let

V (s , z ) =  0 {υ (r, z )}

© 2000 by CRC Press LLC


FIGURE 9.1 Electrical potential due to an electrified disc.

so that

∂ 2V
 0 {∇ 2υ } = −s 2V (s , z ) + (s , z ) = 0 .
∂z 2

The solution of this differential equation is

V ( s , z ) = A ( s )e− sz + B ( s )e sz

where A and B are functions that we have to determine using the boundary conditions.
Because the potential vanishes as z tends to infinity, we have B(s)  0. By inverting the Hankel
transform, we have


υ (r , z ) =
∫ 0
sA(s )e − sz J 0 (sr)ds . (9.29)

The boundary conditions are now


υ (r, 0) =
∫ 0
sA(s ) J 0 (rs )ds = υ 0 , 0 ≤ r <1 (9.30)

∂υ ∞

∂z
(r, 0) =
∫ 0
s 2 A(s ) J 0 (rs )ds = 0, r >1 . (9.31)

Using entries (9.8) and (9.9) of Table 9.1 (see Section 9.11), we see that A(s) = sin s/s 2 so that

2υ 0 ∞


sin s − sz
υ (r , z ) = e J 0 (sr)ds . (9.32)
π 0 s

In Figure 9.1, the graphical representation of υ (r,z) for υ0 = 1 is depicted on the domain 0 ≤ r ≤ 2, 0 ≤
z ≤ 1. The evaluation of υ (r,z) requires numerical integration.
Equations (9.30) and (9.31) are special cases of the more general pair of equations

© 2000 by CRC Press LLC



0
f (t )t 2α Jν ( xt )dt = a( x ), 0≤ x <1 (9.33)

∫ 0
f (t ) Jν ( xt )dt = 0, x >1 (9.34)

where a(x) is given and ƒ(x) is to be determined.


The solution of (9.29) can be expressed as a repeated integral:12

2− α x 1−α
∫ ∫ a( t ) t
1 s
d
f (x) = s − ν−α Jν+α ( xs ) ν+1
( s 2 − t 2 )α dt ds , − 1 < α < 0 (9.35)
Γ (α + 1) 0 ds 0

(2 x )1−α
∫ ∫
1 s
f (x) = s − ν−α +1 Jν+α ( xs ) a(t )t ν+1 ( s 2 − t 2 )α −1 dt ds , 0 < α < 1 . (9.36)
Γ (α ) 0 0

If a(x) = x β, and α < 1, 2α + β > –3/2, α + v > –1, v > –1, then

 β + ν  −( 2α +β+1)
Γ 1 + x
 2 

x
f (x ) = t α +β+1 J ν+α (t )dt . (9.37)
 β + ν 0
2 α Γ 1 + α + 
 2 

With β = v and α < 1, α + v > – 1, v > – 1 further simplification is possible:

Γ (ν + 1)
f (x) = Jν+α +1 ( x ) . (9.38)
(2 x )α Γ (ν + α + 1)

9.5.2 Heat Conduction


Heat is supplied at a constant rate Q per unit area and per unit time through a circular disc of radius a
in the plane z = 0, to the semi-infinite space z > 0. The thermal conductivity of the space is K. The plane
z = 0 outside the disc is insulated. The mathematical model of this problem is very similar to that of
Section 5.1. The temperature is denoted by υ (r,z). We have again the Laplace Equation (9.26) in polar
coordinates, but the boundary conditions are now

∂υ (r, z )
−K = Q, r < a, z = 0
∂z (9.39)
= 0, r > a, z = 0 .

The Hankel transform of the differential equation is again

∂2V
( s, z ) − s 2V ( s, z ) = 0 . (9.40)
∂z 2

We can now transform also the boundary condition, using formula (3) in Table 9.1:

© 2000 by CRC Press LLC


∂V
−K ( s , 0) = Qa J1 (as )/ s . (9.41)
∂z
The solution of (9.39) must remain finite as z tends to infinity. We have

V ( s , z ) = A ( s )e− sz .

Using condition (9.41) we can determine

A ( s ) = Qa J1 (as )/(Ks 2 ) .

Consequently, the temperature is given by


Qa
υ (r , z ) = e − sz J 1(as ) J 0 (rs )s −1ds . (9.42)
K 0

9.5.3 The Laplace Equation in the Halfspace z > 0, with a Circularly


Symmetric Dirichlet Condition at z = 0
We try to find the solution υ (r,z) of the boundary value problem

 ∂ 2υ 1 ∂υ ∂ 2υ
 + + = 0, z > 0, 0 < r < ∞
 ∂r 2 r ∂r ∂z 2 (9.43)
υ (r, 0) = f (r).

Taking the Hankel transform of order 0 yields

∂2V
( s, z ) − s 2V ( s, z ) = 0
∂z 2

and


V ( s , 0) =
∫ rf (r ) J ( sr )dr .
0
0

The solution is


V ( s , z ) = e− sz
∫ rf (r ) J ( sr )dr
0
0

so that

∞ ∞
υ (r , z ) =
∫0
s e − sz J 0 (sr)ds
∫ 0
p f (p) J 0 (sp)dp . (9.44)

For the special case

f (r ) = h(a − r )

© 2000 by CRC Press LLC


where h(r) is the unit step function, we have the solution


υ (r , z ) = a

0
e − sz J 0 (sr) J 1(as )ds . (9.45)

9.5.4 An Electrostatic Problem


The electrostatic potential Q(r,z) generated in the space between two grounded horizontal plates at z =
±  by a point charge q at r = 0, z = 0 shows a singular behavior at the origin. It is given by

υ (r, z ) = ϕ(r, z ) + q(r 2 + z 2 )−1/ 2 (9.46)

where ϕ(r,z) satisfies Laplace’s Equation (9.26). The boundary conditions are

ϕ(r , ± l ) + q(r 2 + l 2 )−1/ 2 = 0 . (9.47)

Taking the Hankel transform of order 0, we obtain

∂ 2Φ
( s , z ) − s 2Φ ( s , z ) = 0 (9.48)
∂z 2

qe− sl
Φ( s , ± l ) = − (9.49)
s

(see formula (18) in Table 9.1).


The solution is

A ( s )e− sz + B ( s )e sz

where A(s) and B(s) must satisfy

q e − sl
A(s )e + sl + B(s )e − sl = −
s
q e − sl
A(s )e − sl + B(s )e sl = − .
s

Hence,

q e − sl
A(s ) = B(s ) = −
2s cosh(sl)

and

q e− sl cosh( sz )
Φ( s , z ) = − .
s cosh( sl )

Hence,

© 2000 by CRC Press LLC


∫e
q − sl cosh( sz )
ϕ(r , z ) = −q J ( sr )ds . (9.50)
r +z
2 2 0 cosh( sl ) 0

9.6 The Finite Hankel Transform


We consider the integral transformation

∫ rf (r ) J (αr )dr .
1
Fν (α ) = H ν { f , α } = ν (9.51)
0

A property of this transformation is that

H ν (∆ ν f , α ) = −α 2 Fν (α ) + [ Jν (α ) f ′ (1) − αJν′ (α ) f (1)]

where v is the Bessel differential operator.


If α is equal to the sth positive zero jv,s of Jv(x), we have

H ν (∆ ν f , jν , s ) = − jν2, s H ν ( f , jν , s ) + jν , s J ν+1( jν , s ) f (1) .

If α is equal to the sth positive root βv,s of

hJ ν(x ) + x J ν′ (x ) = 0

where h is a nonnegative constant, we have

H ν (∆ ν f , β ν , s ) = −β 2ν , s H ν ( f , β ν , s ) + Jν (β ν , s )[hf (1) + f ′ (1)] .

The transformation (9.51) with α = j v,s , s = 1, 2, … is the finite Hankel transform. It maps the function
ƒ(r) into the vector (Fv (jv,1), Fv(jv, 2), Fv(j v,3) …). The inversion formula can be obtained from the well-
known theory of Fourier-Bessel series

∑J
Fν ( jν , s )
f (r ) = 2 2
J ν ( jν , s r ) . (9.52)
s =1 ν+1 ( jν , s )

The transformation (9.51) with α = βv ,s , s = 1, 2, … is the modified finite Hankel transform. The inversion
formula is


β 2ν , s Fν (β ν , s ) Jν (β ν , s r )
f (r ) = 2 ∑ h +β
s =1
2 2
ν,s − ν 2 Jν2 (β ν , s )
. (9.53)

If h = 0, βv,s is the sth positive zero of J ν′ (x ), denoted by j ν,′ s .


Formulas for the computation of jv,s and j ν,′ s are given by Olver.15 Values of jv,s and j ν,′ s are tabulated
in Reference 1. A Fortran program for the computation of jv,s and j ν,′ s is given in Reference 18.

© 2000 by CRC Press LLC


Application
We calculate the temperature υ(r,t) at time t of a long solid cylinder of unit radius. The initial temperature
is unity and radiation takes place at the surface into the surrounding medium maintained at zero
temperature.
The mathematical model of this problem is the diffusion equation in polar coordinates

∂ 2υ 1 ∂υ ∂υ
+ = , 0 ≤ r < 1, t > 0 (9.54)
∂r 2 r ∂r ∂t

The initial condition is

υ (r, 0) = 1, 0 ≤ r ≤1 . (9.55)

The radiation at the surface of the cylinder is described by the mixed boundary condition

∂υ
(1, t ) = − hυ (1, t ) (9.56)
∂r

where h is a positive constant.


Transformation of (9.54) by the modified finite Hankel transform yields

dV
(β , t ) = −β 20, s V (β 0, s , t ) (9.57)
dt 0, s

where


1
V (α, t ) = rυ (r, t ) J 0 (αr)dr
0

so that

J1 (α )
∫ rJ (αr )dr =
1
V (α , 0) = . (9.58)
0
0
α

The solution of (9.57), with the initial condition (9.58), is

J1 (β 0, s ) −β20 , s t
V (β 0, s , t ) = e .
β 0, s

Using the inversion formula, we obtain


β 0, s J 1(β 0, s ) J 0 (β 0, s r)
υ (r , t ) = 2 ∑e
j=1
− β 20 , s t

h 2 + β 20, s J 02 (β 0, s )
. (9.59)

© 2000 by CRC Press LLC


9.7 Related Transforms
For some applications, Hankel transforms with a more general kernel may be useful. We give one example.
We consider the cylinder function

Z ν ( s , r ) = Jν ( sr )Yν ( s ) − Yν ( sr ) Jν ( s ) . (9.60)

Using this function as a kernel, we can construct the following transform pair:


Fν ( s ) =
∫ rf (r )Z ( s, r )dr
1
ν (9.61)

∫ sF ( s) J ( s) + Y ( s) ds
Zν ( s, r )
f (r ) = ν 2 2
(9.62)
0
ν ν

The inversion formula follows immediately from Weber’s integral theorem (see Watson26):

∞ ∞

∫ ∫
1
u du f ( s )Z ν (r , u)Z ν ( s , u) s ds = [ Jν2 (r ) + Yν2 (r )][ f (r + ) + f (r − )] . (9.63)
1 0 2
For this reason, we will refer to (9.61) and (9.62) as the Weber transform. This transform has the following
important property:
If

1 ν2
f ( x ) = g ′′ ( x ) + g′( x ) − 2 g( x ) (9.64)
x x
then

2
Fν ( s ) = − s 2 G ν ( s ) − g (1) . (9.65)
π
We may expect that this transform is useful for solving Laplace’s equation in cylindrical coordinates, with
a boundary condition at r = 1.
Example
We want to compute the steady-state temperature u(r,z) in a horizontal infinite homogeneous slab of
thickness 2, through which there is a vertical circular hole of radius 1. The horizontal faces are held at
temperature zero and the circular surface in the hole is at temperature T0.
The mathematical model is

∂2u 1 ∂u ∂2u
+ + =0
∂r 2 r ∂r ∂z 2
u (r , l) = u (r , − l) = 0 (9.66)
u (1, z ) = T0 .

Taking the Weber transform of order zero, we have

∂ 2 U0 2
( s , z ) − s 2 U0 ( s , z ) = T .
∂z 2
π 0
The solution of this ordinary differential equation, satisfying the boundary condition, is

© 2000 by CRC Press LLC


2T0  cosh sz 
U0 ( s , z ) =  − 1 .
πs 2  cosh sl 

Consequently, we have


1  cosh sz  Z 0(s , r )

2T0
u (r , z ) = −1 (r > 1) .
s  cosh sl  J 02(s ) + Y02(s )
ds (9.67)
π 0

9.8 Need of Numerical Integration Methods


When using the Hankel transform for solving partial differential equations, the solution is found as an
integral of the form

∫ J ( px ) f ( x )dx
a
I (a, p, ν) = ν (9.68)
0

where a is a positive real number or infinite. In most cases, analytical integration of (9.68) is impossible,
and numerical integration is necessary. But integrals of type (9.68) are difficult to evaluate numerically if
1. The product ap is large.
2. a is infinite.
3. ƒ (x) shows a singular or oscillatory behavior.
In cases 1 and 2, the difficulties arise from the oscillatory behavior of Jv(x) and they grow when the
oscillations become stronger.
We give here a survey of numerical methods that are especially suited for the evaluation of I (a, p, v)
when ap is large or a is infinite. We restrict ourselves to cases where ƒ(x) is smooth, or where ƒ(x) =
x α g (x) where g (x) is smooth and α is a real number.

9.9 Computation of Bessel Function Integrals Over a Finite


Interval
Integral (9.68) can be written as

∫x
1
I (a, p, ν) = aα +1 α
Jν (apx ) g (ax )dx . (9.69)
0

We assume that α + v > – 1. If (α+v) is not an integer, then there is an algebraic singularity of the
integrand at x = 0. If ap is large, then the integrand is strongly oscillatory.
If (α + v) is an integer and ap is small, classical numerical integration methods, such as Romberg
integration, Clenshaw-Curtis integration of Gauss-Legendre integration (see Davis and Rabinowitz 4 ) are
applicable. If (α + v) is not an integer and ap is small, the only difficulty is the algebraic singularity at
x = 0, and Gauss-Jacobi quadrature or Iri-Moriguti-Takesawa (IMT) integration 4 can be used. If ap is
large, special methods should be applied that take into account the oscillatory behavior of the integrand.
We describe two methods here.

9.9.1 Integration between the Zeros of Jv (x)


We denote the sth positive zero of Jv (x) by j v , s and we set jv, 0 = 0. Then

© 2000 by CRC Press LLC


N

∑ ∫
a
I (a , p , ν) = (−1)k +1I k + J ν( px )f (x )dx (9.70)
j ν, N p
k =1

where


j ν, k p
Ik = | J ν( px )| f (x )dx (9.71)
j ν , k −1 p

and where N is the largest natural number for which jv,N ≤ ap. This means that N is large when ap is large.
Using a transformation attributed to Longman,10 the summation in Equation (9.70) can be written as

S= ∑(−1)
k =1
k +1 1
2
1
4
1
I k = I 1 − ∆I 1 + ∆2I 1 + L + (−1) p−12− p ∆p−1I 1
8

1 1 1 
+ (−1)N −1  I N + ∆I N −1 + ∆2I N −2 + L + 2− p ∆p−1I n− p+1 
2 4 8 

+ 2− p(−1) p[∆pI 1 − ∆pI 2 + ∆pI 3 − L + (−1)N −1− p ∆pI N − p ].

Assuming now that N and p are large and that high-order differences are small, the last bracket may be
neglected and

1 1 1
S  I 1 − ∆I 1 + ∆2I 1 − L
2 4 8
(9.72)
N −1 1 1 1 2 
+ (−1)  2 I N + 4 ∆I N −1 + 8 ∆ I N −2 + L .
 

The summations in (9.72) may be truncated as soon as the terms are small enough. For the evaluation
of Ik , k = 1, 2, …, classical integration methods (e.g., Lobatto’s rule) can be used, but special Gauss
quadrature formulas (see Piessens16) are more efficient. If the integral I1 has an algebraic singularity at
x = 0, then the Gauss-Jacobi rules or the IMT rule are recommended.

9.9.2 Modified Clenshaw-Curtis Quadrature


The Clenshaw-Curtis quadrature method is a well-known and efficient method for the numerical eval-
uation of an integral I with a smooth integrand. This method is based on a truncated Chebyshev series
approximation of the integrand. However, when the integrand shows a singular or strongly oscillatory
behavior, the classical Clenshaw-Curtis method is not efficient or even applicable, unless it is modified
in an appropriate way, taking into account the type of difficulty of the integrand. We call this method
then a modified Clenshaw-Curtis method (MCC method). The principle of the MCC method is the
following: the integration interval is mapped onto [–1, +1] and the integrand is written as the product
of a smooth function g(x) and a weight function w(x) containing the singularities or the oscillating factors
of the integrand; that is,

+1
I=
∫ −1
w ( x ) g ( x )dx . (9.73)

© 2000 by CRC Press LLC


The smooth function is then approximated by a truncated series of Chebyshev polynomials

g (x )  ∑′ c T (x),
k =0
k k −1≤ x ≤ 1 . (9.74)

Here the symbol Σ′ indicates that the first term in the sum must be halved. For the computation of the
coefficients c k in Equation (9.74) several good algorithms, based on the Fast Fourier Transform, are
available.
The integral in (9.73) can now be approximated by

I ∑′c M
k =0
k k (9.75)

where

+1
Mk =
∫ −1
w ( x )Tk ( x )dx

are called modified moments.


The integration interval may also be mapped onto [0, 1] instead of [–1, 1], but then the shifted
Chebyshev polynomials Tk★ (x ) are to be used.
We now consider the computation of the integral (9.69),

∫x
1
α
I= Jν (ωx ) g ( x )dx . (9.76)
0

If

g (x )  ∑ ′ c T ′(x)
k =0
k k (9.77)

then

I ∑ ′ c M (ω, ν, α)
k =0
k k (9.78)

where

∫ x J (ωx)T
1
α
M k (ω, ν, α) = ν k

(x )dx . (9.79)
0

These modified moments satisfy the following homogeneous, linear, nine-term recurrence relation:

© 2000 by CRC Press LLC


ω2  ω2 
M k + 4 + (k + 3)(k + 3 + 2α) + α 2 − ν 2 − M
16  4  k + 2

+ [4(ν 2 − α 2 ) − 2(k + 2)(2α − 1)]M k +1

 3ω 2 
−  2(k 2 − 4) + 6(ν 2 − α 2 ) − 2(2α − 1) − M (9.80)
 8  k

+ [4(ν 2 − α 2 ) + 2(k − 2)(2α − 1)]M k −1

  ω2  ω2
+ (k − 3)(k − 3 − 2α) +  α 2 − ν 2 −  M + M =0.
  4   k − 2 16 k − 4

Because of the symmetry of the recurrence relation of the shifted Chebyshev polynomials, it is convenient
to define

T−★k (x ) = Tk★ (x ), k = 1, 2, 3, …

and consequently

M − k (ω , ν, α ) = M k (ω , ν, α ) .

To start the recurrence relation with k = 0, 1, 2, 3, … we need only M 0, M 1, M 2, and M 3 . Using the
explicit expressions of the shifted Chebyshev polynomials, we obtain

M 0 = G (ω , ν, α )

M1 = 2G (ω , ν, α + 1) − G (ω , ν, α )
(9.81)
M 2 = 8G (ω , ν, α + 2) − 8G (ω , ν, α + 1) + G (ω , ν, α )

M 3 = 32G (ω , ν, α + 3) − 48G (ω , ν, α + 2) + 18G (ω , ν, α + 1) − G (ω , ν, α )

where

∫x
1
α
G (ω , ν, α ) = Jν (ωx )dx . (9.82)
0

Because

ω 2 G (ω , ν, α + 2) = [ν 2 − (α + 1) 2 ]G (ω , ν, α )
(9.83)
+ (α + ν + 1) Jν (ω ) − ωJν−1 (ω )

we need only G(ω, v, α) and G(ω, v, α + 1).


Luke12 has given the following formulas:
1. A Neumann series expansion that is suitable for small ω

© 2000 by CRC Press LLC


 ν − α + 1
∞ (ν + 2k + 1) 
 2 

2
G (ω , ν, α ) = Jν+ 2k +1 (ω ) (9.84)
ω (α + ν + 1) k =0
 ν + α + 3
 
 2 k

2. An asymptotic expansion that is suitable for large ω

 ν + α + 1
α
Γ 
2  2  2
G (ω , ν, α ) = − ( g1 cos θ + g 2 sin θ ) (9.85)
ω α +1  ν − α + 1 πω 3
Γ 
 2 

where

θ = ω − νπ / 2 + π / 4

and

g1 ~ ∑(−1) a
k =0
k
2k ω −2k , ω→∞

g2 ~ ∑(−1) a
k =0
k
2 k +1 ω −2k −1, ω→∞

(1 / 2 − ν)k (1 / 2 + ν)k
ak = bk
2k k !
b0 = 1

2(k + 1)(α − k − 1 / 2)
bk +1 = 1 + b .
(ν − k − 1 / 2)(ν + k + 1 / 2) k

If α and v are integers, the following formulas are useful1:

ν−1

∫ ∫ ∑J
1 1
2
J 2 ν(ωx )dx = J 0(ωx )dx − 2 k +1 (ω)
0 0 ω k =0

ν
1 − J 0(ω) 2
∫ ∑J
1
J 2 ν+1(ωx )dx = − (ω) .
0 ω ω k =1
2k

For the evaluation of

∫ J (ωx )dx
1

0
0

© 2000 by CRC Press LLC


Chebyshev series approximations are given by Luke.11 We now discuss the numerical aspect of the
recurrence formula (9.80). The numerical stability of forward recursion depends on the asymptotic
behavior of Mk(ω, υ, α) and of eight linearly independent solutions yi,k , i = 1, 2, …, 8, k → ∞.
Using the asymptotic theory of Fourier integrals, we find

y1, k ~ k −2

y 2, k ~ k −4

y 3, k ~ k −2(α +1)−2 ν

y 4, k ~ k −2(α +1)+2 ν , if ν ≠ 0 (9.86)

~ k −2(α+1)lnk , if ν = 0
k
ω
y 5, k ~ y 6, k ~   e kk α
 4k 
k
 4k 
y 7, k ~ y 8, k ~   e − kk α
ω

and

1
M k (ω, ν, α) ~ − J (ω)k −2
2 ν
(9.87)
ων
+ (−1) 2
k −3 ν− 2α −1
Γ(ν + 1)
[ ]
cos π(α + 1) Γ(2α + 2)k −2α − 2 ν− 2 .

The asymptotically dominant solutions are y7,k and y8,k . The asymptotically minimal solutions are y5,k and
y6,k. We may conclude that forward and backward recursion are asymptotically unstable. However, the
instability of forward recursion is less pronounced if k ≤ ω/2. Indeed, practical experiments demonstrate
that Mk(ω, v, α) can be computed accurately using forward recursion for k ≤ ω/2. For k > ω/2 the loss
of significant figures increases and forward recursion is no longer applicable. In that case, Oliver’s
algorithm14 has to be used. This means that (9.80) has to be solved as a boundary value problem with
six initial values and two end values. The solution of this boundary value problem requires the solution
of a linear system of equations having a band structure.
An important advantage of the MCC method is that the function evaluations of g, needed for the
computation of the coefficients ck of the Chebyshev series expansion, are independent of the value of ω.
Consequently, the same function evaluations may be used for different values of ω, and have to be
computed only once.
Numerical examples can be found in References 19 and 20.

9.10 Computation of Bessel Function Integrals over an Infinite


Interval
In this section we consider methods for the computation of


I ( p, ν) =
∫0
Jν ( px ) f ( x )dx . (9.88)

© 2000 by CRC Press LLC


9.10.1 Integration between the Zeros of Jv (x) and Convergence Acceleration
We have

I ( p, ν) = ∑ (−1)
k =1
k +1
Ik (9.89)

where


j ν,k p
Ik = J ν( px ) f (x )dx . (9.90)
j ν , k −1 p

Using Euler’s transformation,4 the convergence of series (9.89) can be accelerated

1 1 1 1
I ( p, ν) = I1 − ∆I1 + ∆2 I1 − ∆3 I1 + L . (9.91)
2 4 8 16

It is not always desirable to start the convergence acceleration with I1, but with some later term, say Im,
so that

1 

j ν, m −1 p
1 1
I ( p , ν) = J ν( px ) f (x )dx + (−1)m −1  I m − ∆I m + ∆2I m − L .
0 2 4 8 

Other convergence accelerating methods, for example the -algorithm,23 are also applicable (for an
example, see Reference 21).

9.10.2 Transformation into a Double Integral


Substituting the integral expression

( x / 2) ν
∫ (1 − t )
1
2 ν−1/ 2
Jν ( x ) = 2 cos( xt )dt (9.92)
Γ (ν + 1 / 2) π 0

into (9.88) and changing the order of integration, we obtain

2( p / 2) ν
∫ (1 − t )
1
2 ν−1/ 2
I ( p, ν) = F (t )dt (9.93)
Γ (ν + 1 / 2) π 0

where


F (t ) =
∫x 0
ν
f ( x )cos( pxt )dx . (9.94)

We assume that the integral in (9.94) is convergent. If we want to evaluate (9.93) using an N-point Gauss-
Jacobi rule, then we have to compute the Fourier integral (9.94) for N values of t. Because F(t) shows a
peaked or even a singular behavior especially when ƒ(x) is slowly decaying, a large enough N has to be
chosen.
This method is closely related to Linz’s method,8 which is based on the Abel transformation of I (p,v).

© 2000 by CRC Press LLC


9.10.3 Truncation of the Infinite Interval
If a is an arbitrary positive real number, we can write

∫ J ( px ) f ( x )dx + R ( p, a)
a
I ( p, ν) = ν (9.95)
0

where


R ( p, a) =

a
Jν ( px ) f ( x )dx . (9.96)

The first integral in the right side of (9.95) can be computed using the methods of Section 9.9.
If a is sufficiently large and ƒ is strongly decaying, then we may neglect R(p,a).
If

c1 c 2
f (x) ~ + +L (9.97)
x x2
is an asymptotic series approximation which is sufficiently accurate in the interval [a, ∞), then

∑ ∫
J ν( px )
R( p , a)  ck dx . (9.98)
k
a xk

Longman9 has tabulated the values of the integrals in (9.98) for some values of v and ap.
Using Hankel’s asymptotic expansion,1 for x → ∞

2
Jν ( px ) ~ [P ( px )cos χ − Qν ( px )sin χ ] (9.99)
πpx ν

where χ = px – (v/2+1/4)π, and where Pv(x) and Qv(x) can be expressed as a well-known asymptotic
series, R(p,a) can be written as the sum of two Fourier integrals.
Especially, if a = (8 + v/2 + 1/4)π/p, we have


2
R( p , a) = P (p(u + a))f (u + a)cos pu du
0 πp(u + a) ν
(9.100)


2
− Q (p(u + a))f (u + a)sin pu du .
0 πp(u + a) ν

For the computation of the Fourier integrals in (9.100), tailored methods are available.4

9.11 Tables of Hankel Transforms


Table 9.1 lists the Hankel transform of some particular functions for the important special case v = 0.
Table 9.2 lists Hankel transforms of general order v. In these tables, h(x) is the unit step function, Iv and
Kv are modified Bessel functions, L0 and H0 are Struve functions, and Ker and Kei are Kelvin functions
as defined in Abramowitz and Stegun.1 Extensive tables are given by Erdélyi et al.,6 Ditkin and Prudnikov,5
Luke,11 Wheelon,27 Sneddon,24 and Oberhettinger.13

© 2000 by CRC Press LLC


TABLE 9.1 Hankel Transforms of Order 0.

© 2000 by CRC Press LLC


TABLE 9.2 Hankel Transforms of General Order v.

© 2000 by CRC Press LLC


References
1. M. Abramowitz and I. N. Stegun, Handbook of Mathematical Functions, Dover Publications, New
York, 1965.
2. R. N. Bracewell, The Fourier Transform and Its Applications, McGraw-Hill, New York, 1978.
3. B. Davies, Integral Transforms and Their Applications, Springer-Verlag, New York, 1978.
4. P. J. Davis and P. Rabinowitz, Methods of Numerical Integration, Academic Press, New York,
1975.
5. V. A. Ditkin and A. P. Prudnikov, Integral Transforms and Operational Calculus, Pergamon Press,
Oxford, 1965.
6. A. Erdélyi, W. Magnus, F. Oberhettinger, and F. G. Tricomi, Tables of Integral Transforms, Vol. 2,
McGraw-Hill, New York, 1954.
7. A. Gray and G. B. Mathews, A Treatise on Bessel Functions and Their Applications to Physics, Dover
Publications, New York, 1966.
8. P. Linz, A method for computing Bessel function integrals, Math. Comp., 20, pp. 504-513, 1972.
9. I. M. Longman, A short table of ∫ ∞x J0(t)t–n dt and ∫ ∞x J1(t)t–n dt, Math. Tables Aids Comp., 13, pp.
306-311, 1959.
10. I. M. Longman, A method for the numerical evaluation of finite integrals of oscillatory functions,
Math. Comp., 14, pp. 53-59, 1960.
11. Y. L. Luke, The Special Functions and their Approximations, Vol. 2, Academic Press, New York,
1969.
12. Y. L. Luke, Integrals of Bessel Functions, McGraw-Hill, New York, 1962.
13. F. Oberhettinger, Table of Bessel Transforms, Springer-Verlag, Berlin, 1971.
14. J. Oliver, The numerical solution of linear recurrence relations, Num. Math., 11, pp. 349-360,
1968.
15. F. W. J. Olver, Bessel Functions, Part III: Zeros and Associated Values, Cambridge University Press,
Cambridge, U.K., 1960.
16. R. Piessens, Gaussian quadrature formulas for integrals involving Bessel functions, Microfiche
section of Math. Comp., 26, 1972.
17. R. Piessens, Automatic computation of Bessel function integrals, Comp. Phys. Comm., 25, pp. 289-
295, 1982.
18. R. Piessens, On the computation of zeros and turning points of Bessel functions, in E. A. Lipitakis
(Ed.), Advances on Computer Mathematics and Its Applications, World Scientific, Singapore, pp. 53-
57, 1993.
19. R. Piessens and M. Branders, Modified Clenshaw-Curtis method for the computation of Bessel
function integrals, Bit 23, pp. 370-381, 1983.
20. R. Piessens and M. Branders, A survey of numerical methods for the computation of Bessel function
integrals, Rend. Sem. Mat. Univers. Politecn. Torino, Fascicolo speciale, pp. 250-265, 1985.
21. R. Piessens, E. de Doncker-Kapenga, C. W. Ueberhuber, and D. K. Kahaner, Quadpack, A Subroutine
Package for Automatic Integration, Springer-Verlag, Berlin, 1983.
22. F. E. Relton, Applied Bessel Functions, Dover Publications, New York, 1965.
23. D. Schanks, Non-linear transformations of divergent and slowly convergent sequences, J. Math.
Phys., 34, pp. 1-42, 1955.
24. I. N. Sneddon, The Use of Integral Transforms, McGraw-Hill, New York, 1972.
25. C. J. Tranter, Integral Transforms in Mathematical Physics, Methuens & Co., London, 1951.
26. G. N. Watson, A Treatise on the Theory of Bessel Functions, Cambridge University Press, Cambridge,
U.K., 1966.
27. A. D. Wheelon, Table of Summable Series and Integrals Involving Bessel Functions, Holden-Day, San
Francisco, 1968.
28. K. B. Wolf, Integral Transforms in Science and Engineering, Plenum Press, New York, 1979.

© 2000 by CRC Press LLC

You might also like