G Stephenson - P M Radmore - Advanced Mathematical Methods For Engineering and Science Students-Cambridge University Press (1990) PDF
G Stephenson - P M Radmore - Advanced Mathematical Methods For Engineering and Science Students-Cambridge University Press (1990) PDF
G Stephenson - P M Radmore - Advanced Mathematical Methods For Engineering and Science Students-Cambridge University Press (1990) PDF
G. STEPHENSON
Emeritus Reader in Mathematics,
Imperial College, London
and
P. M. RADMORE
Lecturer, Department of Electronic and
Electrical Engineering,
University College London
CAMBRIDGE
UNIVERSITY PRESS
Published by the Press Syndicate of the University of Cambridge
The Pitt Building, Trumpington Street, Cambridge CB2 1RP
40 West 20th Street, New York, NY 10011^1211, USA
10 Stamford Road, Oakleigh, Melbourne 3166, Australia
Preface ix
Notes to the reader xi
1 Suffix notation and tensor algebra 1
1.1 Summation convention 1
1.2 Free and dummy indices 3
1.3 Special symbols 4
1.4 Vector identities 6
1.5 Vector operators 8
1.6 Orthogonal coordinate systems 9
1.7 General coordinate transformations 13
1.8 Contra variant vectors 15
1.9 Kronecker delta symbol 15
1.10 Scalars and covariant vectors 16
1.11 Tensors of higher rank 18
Problems 1 19
2 Special functions 22
2.1 Origins 22
2.2 The gamma-function 23
2.3 The exponential integral and related functions 28
2.4 The error function 28
2.5 The Bessel function 31
2.6 The Bessel function Yv 36
2.7 Generating function 39
2.8 Modified Bessel functions 42
2.9 Summary of the main properties of special functions 44
Problems 2 50
3 Non-linear ordinary differential equations 53
3.1 Introduction 53
3.2 Equations with separable variables 54
vi Contents
5 Contour integration 98
5.1 Functions of a complex variable 98
5.2 Exponential and logarithmic functions 101
5.3 The derivative of a complex function 103
5.4 The Cauchy-Riemann equations 105
5.5 Derivatives of multi-valued functions 108
5.6 Complex integration 109
5.7 Cauchy's Theorem (First Integral Theorem) 112
5.8 Cauchy Integral Formula (Second Integral Theorem) 116
5.9 Derivatives of an analytic function 118
5.10 Taylor and Laurent series 120
5.11 Singularities and residues 125
5.12 Cauchy Residue Theorem 131
Problems 5 134
G.S.
Imperial College, London
P.M.R.
University College London
IX
Notes to the reader
XI
1
Suffix notation and tensor
algebra
a= fl/Cf, (1.2)
Example 2
Example 3 Consider the term a^bibj in which / and j both occur twice.
The summation convention implies summation over / and ;
independently. Hence, if / and j run from 1 to 2,
ttijbibj = axxbxbx + a12b1b2 + a2lb2bx + a22b2b2 (1.6)
? | a 21 ). ^ (1.7)
The usual rules apply when the summation convention is being used,
that is,
(1.8)
and
We see in (1.9) that although i occurs three times in the left-hand side
it only occurs twice in each term and therefore summation over i is
implied. Also
(atCjXbidj), (1.10)
Figure 1.1
1.2 Free and dummy indices 3
Example 4
/ 2 x 2
2 <*sxs = (a1xl + a2x2)2 an)
/
\s = l
= a\x\ #2*2- (1.12)
Using the summation convention we can write this as
where cly c2y . . . , cn are given constants. This set of equations can be
written as
02;*; = C2y
(1.15)
7= 1
where / is now the dummy index and s is the free index. If ais is
symmetric so that a^ = asi then
d(j>ldxs = 2aisxh (1.21)
whereas if a^ is skew- (or anti-) symmetric so that a^ = —asi then
0 = 0 and d(f)/dxs = 0. ^ (1.22)
1. Kronecker delta
The Kronecker delta symbol <5,y is defined by
1.3 Special symbols 5
We note that <5/yey- = dnei + <5/2e2 + <5/3e3 and that only one of these
three terms is non-zero depending on /; for example <5n = 1,
<5i2 - ^13 = 0. Hence <5/y-ey = e,-. Also, as in (1.3),
. (1.24)
Consider the quantity <5iy<5yA:. Then
dijdjk = (ei.ej)(ej.ek) (1.25)
= (ef-. eOfo. ek) + (e,. e2)(e2 . ek)
+ (e / .e 3 )(e 3 .e,). (1.26)
Now if i i= ky then at most one of the two brackets in each term can be
non-zero and hence each term is zero. If i = k then one term has both
brackets non-zero and equal to 1. Hence 6ij6Jk is zero if / =£ k and is 1
if / = k. This is just the definition of 6ik (see (1.23)) and so
6rsApqs=Apqr. (1.29)
Cjik=-eijk- (1-33)
c = a X b = ( ^ ( e , X C y ) . ek)ek (1.35)
= aibj€kiJek, (1.36)
using (1.30). In (1.36), summation over the indices i, j and k is implied
so that
a X b = aibjCujei + aibJ€2ije2 + aibje3ije3 (1.37)
and the rth component of a X b is therefore
(a X b) r = e^afij, (r = 1, 2, 3). (1.38)
For the scalar triple product
(a X b ) . c = (afijet X e 7 ). ckek (1.39)
= aibJckek.(eiXeJ) (1.40)
= ekijaibjCk. (1.41)
Using (1.32), we have finally
(aXb).c = ^ Q - (1-42)
The scalar triple product does not depend on the order of the dot
and cross operations since a . (b X c) = (b X c ) . a = €pqrbpcqar =
erpqarbpcq = (a X b ) . c.
We can use a result from vector algebra to derive an important
identity involving the Kronecker delta and the alternating symbol. We
have the standard result
(a X b ) . (c X d) = (a . c)(b . d) - (a . d)(b . c). (1.43)
Putting a = e,-, b = ey, c = ek and d = eh then
(e, X e y ). (e* X ez) = (ef-. e*)(ey • e7) - (ef-. e/)(ey. ek). (1.44)
1.4 Vector identities 7
(1.54)
Since e ^ is an antisymmetric symbol under interchange of any two
indices, quantities such as eilknink are zero because for any pair of
values / and A:, two terms (with opposite signs) result from the
summations over / and k (for example, €il2nln2 cancels with
£/2i#2rti = —^12^1^2)- Hence, since nknk = 1,
( 0 % = /1,-ziy 4- eilk€km/filnm (1.55)
= /1,-rty + eilkemjkriinm. (1.56)
as required. ^
Suffix notation and tensor algebra
giving
curl a = V X a = e ^ ^ V y ^ . (1.65)
The curl of a vector (in cartesian coordinates) can easily be written
down in terms of determinants as follows: if in (1.41) a . ( b X c ) is
written out in full using the definition (1.31) of eijk, we find
a.(bXc) = e ^ , V * ( L66 )
= a1b2c3 — alb3c2+ a2b3cl — a2bxc3 + a3blc2— a3b2clf (1.67)
which can be written as the determinant
ax a2 a3
a . (b X c) = b\ b2 b3 (1.68)
cx c2 c3
Hence
VXa= ei^iVjUk (1.69)
i e2 e3
x V2 V3 (1.70;
ax a2 a3
1.6 Orthogonal coordinate systems 9
Example 7
(1.71)
(1.72)
^ (1.73)
Example 8
(1.74)
(1.75)
= 0 c u r l b + e / ^(V</>) 7 ^ (1.76)
Example 9
curl curl A = e^V/curl A)* (1.78)
= ei<EijkVj(eklmVlAm) (1.79)
= ei€ijk€klmVlVfAm (1.80)
(1.81)
(1.82)
= eiVi(VmAm)-VjVJeiAi (1.83)
2
= grad(div A) - V A, (1.84)
where
v =v Vy =
'" a ^ + a^i + acl' ^
div A
= TVT \T~ 4dq 2
n1n2n3ld
(1.88)
hxex h2e2
1 d d d
curl A =
hxh2h3 dqx dq2 dq3
h3A3
It is important to realise that in the above expressions the vectors e,
are unit vectors which are directed along the three new coordinate
axes and point in the direction of increasing coordinate values. Further
(1.99)
dp + p dip7
3
(1.100)
ei pe 2 e3
a d d
curl A = — (1.101)
P dp dip dz
Al pA2 A3
Figure 1.2
and
(1.102)
p I dp \ dp) p dd>2
Hence
(1.104)
(1.110)
r 96 r sin 6 9(f)
l) + r—(smeA2) + , (LIU)
r sin
5i re 2 r sin 6 e 3
1
curl A = (1.112)
r2 sin 6
Ax rA2
and
where f\,f2, . . . ,fn are functions specifying how the coordinates xt are
related to the coordinates xt. Using a free index /, (1.114) and (1.115)
may be written
Xi =fi(x\> x2, . . . , xn)y (i = 1, 2, . . . , n)y (1.116)
14 Suffix notation and tensor algebra
or more compactly,
= fk> (1.119)
or
dx (1.120)
J = \dXi/dxj\±0. (1.121)
dx^^dxj, (I = 1 , 2 , . . . , / I ) . (1.122)
With this definition we see that the dxl of (1.123) form a contravariant
vector.
Quantities with indices in the superscript position are contravariant
objects and a quantity with one free index (A1) is called a tensor of
rank one, the rank being equal to the number of free indices.
Accordingly, if (1.124) holds, A1 is a contravariant tensor of rank one,
a rank one object of this type being a vector.
**-££* (1.I2W
dx dx dxJ
dxk
= -—AJ = 6fAJ = Ak. (1.130)
16 Suffix notation and tensor algebra
Hence
v-g* am,
is the inverse transformation of (1.124) (provided that (1.121) holds).
a^a^aF'a^af7' (L135)
Ak=^-kAi. (1.139)
1.10 Scalars and covariant vectors 17
Similarly
(1.152)
Equations (1.151) and (1.152) are of the same form as (1.142), as
required. ^
where the mixed tensor has one contravariant and one covariant
index. Tensors of higher rank may be defined in an obvious way.
The algebra of tensors is straightforward since we may readily define
the sum and difference of two tensors using the basic transformation
laws. For example,
= bkpblqgkldxpdx« (1.161)
= gkldxkdxl = ds2. (1.162)
Hence, the quantity ds is a scalar since it is invariant under a general
coordinate transformation.
Tensor algebra and its associated calculus are important tools in the
study of continuum mechanics and in the general theory of relativity.
Problems 1
1. Write out aikXiXk in expanded form, assuming aik — akiy and
i, k = 1, 2, 3.
2. Over which indices (if any) in the following expressions is
summation implied?
(i) dijbj, (ii) dijbjj, (iii) atibny (iv) au = bu.
3. Find the values of 6iy6iy, <5,y(5yVAm<5/m, ejklAkAh and dikeikmy all
indices ranging from 1 to 3.
4. Evaluate €iklejki and €ijkeijky all indices ranging from 1 to 3.
5. The object 7^; is related to 5iy by the relation
7;,. = (adijdu + PdxdjdSu,
where a and /? are constants. Find 5^ in terms of Tijy all indices
ranging from 1 to 3. (Hint: first evaluate Tu.)
6. The square matrices A and B have elements aik and bik
respectively. Use the suffix notation to show that (AB)T = BTAT,
20 Suffix notation and tensor algebra
12. Verify that, if 4>{xt) satisfies the equation V20 + K2<j) = 0, where K
is a constant, then the second-order tensor
Tik = (VfV* - SikV2)<t>,
13. The symmetric tensor gik is related to the three-index object Yijk
which is symmetric in its last two indices (that is, Tijk = TikJ) by the
relation
_ a T — Q V = 0
> ft/5x skm Bsk1- sim y}
'
Show that
= l/dgim dgkm dgik\
2 \ dxk dxt dxj
2
Special functions
2.1 Origins
This chapter is principally concerned with some of the special
functions which occur in the solution of differential equations both
ordinary and partial. These functions are more complicated than those
met in elementary mathematical methods (for example, the sine,
cosine, exponential and logarithmic functions) and often arise in the
solutions of linear second-order differential equations of the form
d2y dy
22
2.2 The gamma-function 23
=fJo
(2.3)
= f txe-tdt = [-txe-']Z + x \ 1
dt. (2.4)
Jo Jo
In (2.4), the integrated term is zero because at the upper limit the
exponential dominates the tx term, while at the lower limit tx is zero
since x > 0 . Hence, using (2.3),
(2.9)
Jo Jo
24 Special functions
(2.10)
(2.13)
we have
(2.14)
(2.15)
Figure 2.1
T(x)
4-
-l
-4
2.2 The gamma-function 25
Example 1 Consider
using (2.12). ^ i
Example 2 Consider
using (2.9). ^
Other properties of the gamma-function are also of interest.
Differentiating (2.3) with respect to x we have
Now
where y is called Euler's constant and has the value (to four decimal
places) 0-5772. This value can be obtained either by numerical
integration in (2.22) or from the alternative definition
(2.24)
J 1
0
sin2"1"1 0 cos2""1 0^0. (2.28)
Example 3 Consider
f-Jt/2
Then
J o
V(tan 6) dd. (2.29)
rJt/2
/ = f sin^ 0 cos'i 0 dd = iB(|, i). (2.30)
Jo
Using (2.27), and T(l) = 1, this can be written
= f feT'dt. (2.35)
Jo
We change the variable of integration from t to x, where t = x 4- rV*-
Then
ix (2.37)
\ p [ ( ^ ) (2.38)
—\/x L \ vx/ J
Writing (2.38) as the sum of two integrals and replacing the logarithm
term by its MacLaurin expansion in — V* <r< V*, we have
(2.40)
Ci(x) = - \ ^ d t , (2.43)
( L ^) (2.45)
(2.46)
fS^ (2.47)
2
2 r
=^- e"M du. (2.48)
\JT Jo
2.4 The error function 29
Again using the standard result (2.10), we see that erf jt-» 1 as x-^^.
The graph of the error function is shown in Figure 2.3.
The complement of the error function, denoted erfex, is defined by
(see also Figure 2.3). Both of these functions arise from the solution of
partial differential equations and are connected with the Laplace
transforms of particular functions (see Chapters 7 and 8).
It is relatively easy for these functions to generate what is called an
Figure 2.2
+-X
(a)
30 Special functions
•2)du. (2.52)
(2.53)
-+-X
2.5 The Bessel function 31
\du (2.55)
-^-"2du (2.56)
=
l?" | /3 e "" 2 r f M " (2 57)
'
Continuing in this way, we eventually find
1.3 1.3.5
(2JC2)2 (2X2)3 ' ' '
where
2
[wr^u2du- (2 59)
-
Equation (2.58) is the asymptotic series for ^V^erfcx since the ratio
of consecutive terms is (2n — 1)/2JC2 which tends to zero as x^> °° for a
given n. However, for a given x, the series diverges as /i—>o°. By
choosing a suitable n, the resulting finite series is a good approxima-
tion to the integral in (2.58). Accordingly, using (2.49) and (2.58), it is
seen that the behaviour of erfcx for large x and to first order in 1/x is
(2.60)
* 2 ^ + * £ + (*2-v2)y = 0. (2.61)
The conditions under which a series solution of this kind is valid are
discussed in Chapter 4. In (2.62), m and the ar are constants which
must be found so that (2.62) satisfies (2.61). By this method, we will
find two possible values of m, giving two solutions of (2.61). The
solutions may not be independent if the values of m are identical or
differ by an integer (see Chapter 4) but the method will always
generate one solution. The method for finding a second solution in this
case will be discussed later. The Frobenius series is more general than
a Taylor or MacLaurin series since m may be found to have
non-integer values and the resulting Frobenius series will then contain
non-integer powers of x. This contrasts with the MacLaurin series
which only contains integer powers of x.
Differentiating (2.62) we obtain
+ 2 «r* m + r + 2 = 0. (2.66)
This simplifies to
00 OO
Expanding the first summation by writing out the first two terms
explicitly we have
ao(m2 - v2)xm + ai[(m + I) 2 - v2]xm+1
oo oo
2 m+r
+ 2 ar[(m + rf - v ]x + 2 arxm+r+2 = 0. (2.68)
r=2 r=0
2.5 The Bessel function 33
where the dummy index has been changed from s to r. Hence (2.68)
can be written with the two summations combined as
aQ(m2 - v2)xm + ax[{m + I) 2 - v2]xm+l
00
ar+2 = - % 22 5 2
2) -v
for r = 0, 1, 2,
We now consider the two possible values of the constant m
separately.
and so on for all values of r. The final result for y(x) is, from (2.62),
~2 _4
In particular
x4 x6
x x3 x5 x1
The two solutions (2.79) and (2.84) are independent since (2.79)
contains only positive powers of x whereas (2.84) contains both
positive and negative powers of x. The first is zero at x = 0 (except for
the case v = 0) whilst the second is infinite at x = 0.
If v = n, an integer, then (2.73) still holds and al = 0. From (2.71),
the recurrence relation is
- ( r + 2)] = <ir. (2.85)
Writing out these in full gives
a2.2(2n-2) =
a4.4(2n-4) = (2.86)
a6. 6(2n-6) =
and similar equations for #8, a10, . . . . Only the even values of r need
to be considered since (2.85) with ax = 0 implies that all ar with r odd
are zero. From (2.86) we see that if, say, n = 3 then a4 = 0 which,
working up the list, implies a2 and hence a0 are zero. In general then
all the ar preceding that for r = 2n are zero. We then have from (2.85)
(2.87)
2(2n + 2)'
a a
2n+2 _ 2n
(2.88)
' 4{2n + 4) " 2 . 4{2n 4- 2){2n + 4)'
Figure 2.4
• • *
36 Special functions
and hence
From (2.80) we see that Jn{x) and J-n(x) are linearly dependent. We
normally choose a2n so that
J-n(x) = (-l)nJn(x). (2.92)
The general solution of the Bessel equation may be written as
y(x)=AJy(x) + BJ-v(x), (2.93)
where A and B are constants, only when v is not an integer. When
v = n w e need to obtain a second solution (Jn(x) being one solution)
using the method in the next section.
(2.101)
an
Inserting the series (2.81) for /o0O d taking only the first two terms,
we find for small x
0.4-
-•*
38 Special functions
When v=£n, this is just a linear combination of Jv(x) and J-V(x) and
so is a solution of Bessel's equation (independent of Jv(x)). When
v = n> the numerator and denominator are both zero, so the limit as
v->n must be taken using L'Hopital's rule. Hence
— n sin(v;r)/v(jc) + COS(VJT) — —
Yn{x) = lim I — — I (2.105)
v^n(^ JT COS(VJT) J
[ ( r 4 (2107)
To see why this is a solution of Bessel's equation consider the solutions
Jv and /_ v . These satisfy
-(-l)"/_v], (2.110)
ux ux
where
We now let v-*n. Then from (2.92), the right-hand side of (2.110)
tends to zero and we find that
wn = limwv (2.112)
v—*n
Yn=^"n (2.H3)
r =o 5=0
1. Recurrence relations
(a) Differentiating (2.118) with respect to x gives
M
'- 1 / 0 = 2 *"/;(*) (2.119)
rt = — oo
Hence
\ i t"+%(x)~i
n — — °°
£/-i/n(*>= i tnK(x). (2.121)
and hence
(c) We have
Hence
£[xVn(x)]=xnJn_l(x). (2.133)
Example 4
Expanding the final summation and using the relation (2.92) we find
gix sin 0 _ j / j . \ _^_ 2ry /^\ cos(20) + / (x) cos(40) + 1
m X W e n m 1S e v e n
- f COS(JC sin 0) c o s ( m 0 ) d9 = ( ^ \ (2.140)
7i jQ 1 0 when m is odd,
W e n m iS
i fsin(x sinfl)sin(,n0) d6 = f ^ ° ' ^ (2.141)
;r Jo ^m(^) when m is odd.
42 Special functions
can be found from the standard form (2.61) of the Bessel equation by
writing t = kx. Then (2.149) becomes
<2156>
which is a real function of x. Similar considerations apply to Kv{x)
which is the second solution of the modified Bessel equation (2.152).
Iv(x) and Kv(x) are called modified Bessel functions and their
properties can be obtained in a similar way to those of Jv(x) and Yv(x).
The main properties of these functions are given in the next section.
We note finally that the differential equation
d2yldx2 = (1 - l/4x2)y (2.157)
can be solved in terms of the modified Bessel functions. Putting
y =x^u, we obtain the equation
(2160)
(V>
Again putting y = x^u, we find
2
x ^ + x^-(l4x2+v2)u = 0, (2.161)
44 Special functions
'2^+'f-('2+v>=0- (2162)
The solutions, from (2.152), are therefore Iv(t) and Ky(t) so that the
solution of (2.160) is
1. Bessel functions
Bessel's equation of order v is
When v = n,
J-n(x) = (-l)nJn(x), (2.166)
(2.167)
x4 , „ x6
2 2v (2.16R)
2. 4 " 2 . 4 2 . 62 v
2 z
jJn(x)=Jn^(x)+Jn+l(x), (2.171)
(2.175)
In particular
x2 x4 x6
- —+ . . . . (2.178)
3. Legendre polynomials
Legendre's equation is
2
o d y dy
(l-x2)—^-2x-f- + 1(1 + l)y = 0. (2.179)
46 Special functions
In particular
Po(x) = l, (2.181)
The generating function is
(2.182)
Pn+l(x) = (2.183)
(2.184)
(l-x2)P'n(x) = -nxPn(x) + nPn.x{x). (2.185)
Figure 2.6
2.9 Summary 47
f1 2
Pn(x)Pm(x)dx=——dnm. (2.187)
j-i z/i + i
The solutions are bounded only ifA = - « ( « 4 - l ) and have the form
y = PjT'Ccos 6). (2.190)
Figure 2.7
48 Special functions
4. Laguerre polynomials
The Laguerre equation
x2 —^ + (1 - x) - / + ny = 0, (2.191)
dx dx
where n = 0, 1, 2, . . . , has polynomial solutions Ln{x), called the
Laguerre polynomials, given by
n! dx
In particular
2
L0(JC) = 1, L^*) = 1 - x9 L2(x) = \(x - 4x + 2). (2.193)
L™(x)=—-Ln(x) (2.197)
5. Hermite polynomials
The Hermite equation
In particular
H0(x) = 1, Hx(x) = 2x, H2(x) = 4x2 - 2. (2.201)
The generating function is
(2.202)
(2.204)
We also have the following integral:
(2.207)
dx2
where
A = In + 1. (2.208)
If A # 2« + 1 in (2.207), then y is not finite as x -> ±00.
Figure 2.8
Bi(x)
1—
!fsrf\rf\ff \Ai(x)
-jW\iJ\xjf"
50 Special functions
6. Airy functions
The Airy equation
d2y
(2209)
has solutions called the Airy functions, Ai(x) and Bi(x), where
Ai(x) = \/(xl3)[I-&c*) - mhl)l (2.210)
Bi(x) = V(x/3)[/_i(|je^ + /j(M)]. (2.211)
The integral form for Ai(x) is
Problems 2
1. Use the gamma-function to evaluate
3. Show that
» m>
f
(i) f V(sin0)d0,
Jo
f
(ii) f
Jo
V(cot(9)d0.
Problems 2 51
COSJC
and that
f /„(*) dr = 1.
Jo
where y—x ^u. Obtain the general solution when v = ±\. Use
this result to show that for x2 » v2 — \
dx2
52 Special functions
is
y = Vx[AJ1(2\/x) + BYx{2\/x)l
12. Evaluate J J\{x^) dx, using Example 4.
13. Evaluate the following integrals using (2.180) and integration by
parts:
sin u ji
d 0 < p < 2
'
3ir(§).
3
Non-linear ordinary differential
equations
3.1 Introduction
The differential equations met in Chapter 2 were linear in the sense
that they were special cases of the general nth order linear equation
where fly f2, . . . , fn, g are given functions of x. For such equations the
Principle of Superposition applies: an arbitrary linear combination of
individual solutions is also a solution.
Equations which cannot be written in the form of (3.1) are called
non-linear and, for such equations, the Principle of Superposition does
not apply. A typical first-order non-linear equation is
whereas
d2y , dy
\
G(y) dy
"I dx + C (3.6)
where C is a constant.
Example 1
dy CO
dx
(3.7)
X
x^ + u = K(u), (3.12)
ax
or
^ =^ ^ . (3.13)
dx x
This equation is now of separable form.
3.2 Equations with separable variables 55
Example 2
d JL
i
dx x—y'
(3-14)
Writing (3.14) in the form of (3.11), we have
dy _ y/x
(3.15)
dx 1 —y/x
Putting y = xu gives
du u
(3.16)
~ dx 1-u'
or
du u2
(3.17)
dx x{\ — u)
f — ^ d u = I — + ln,4, (3.18)
J u J x
where A is a constant. Integrating both sides, we have
- - - I n n = ln(i4*), (3.19)
Example 3
xg-(,-0* (3.21,
given that y = 1 and dy/dx = 1 at x = 0.
56 Non-linear ordinary differential equations
Since
'(&.& +* (3.22)
dx \ dxl dx dx
then (3.21) can be written
x w
dx\ dx) dx " 'dx' "~'
or, simplifying,
d I dy\ dy d , ~
dx \ dx) dx dx
2
X^=$y + A, (3.25)
^ = j j + \nB, (3.27)
(3 28)
ii^iJTi) ' -
or
Hence
y - l = Bx(y + l), (3.30)
or
3.3 Equations reducible to linear form 57
Example 4 Consider
y = e" 1 . (3.40)
For y = e~x(Ax + B), y + dy/dx = Ae~x. Inserting these into (3.34),
we have
*) + e~x(Ax + B) = 0. (3.41)
Cancelling e~* and using \n(Ac~x) = In A -x, we find
A(\n A - x) + Ax + B = 0, (3.42)
so that
,41n,4 + £ = 0. (3.43)
Substituting back for B> we have a solution
Example 5 Consider
p= dy/dx, (3.46)
we have
d2y/dx2=pdp/dy. (3.47)
Now (3.45) becomes
pdp/dy + 2p2 = y2. (3.48)
This can be written
Accordingly,
P = f = ^V[(r-l)2 + ^ (3.55)
the positive sign being chosen so that p = \ when y = 0. The variables
may now be separated giving
(3.56)
(3.59)
may be solved, as in Example 5 above, by multiplying by the
integrating factor exp(J P(x) dx). The non-linear equation
Example 6 Consider
dy/dx + 2y=xy3. (3.64)
Hence
1 ^ + 1 = *, (3.65)
y dx y
and therefore
1d/1 \ 1.
2 (3.66)
2dx\y ) y2
Writing z = 1/y2, we have
dz/dx-4z = -2x. (3.67)
Multiplying by the integrating factor e~4x, we find
Example 7 Consider
whence
dx x{\-xy2Y {iJl)
=y (3 73)
j+Y ~T- *
We see that (3.73) is an equation of Bernoulli type when viewed as an
equation for x in terms of y. Letting z = 1/x, then
= (3 74)
T~T ~o> '
dy 2y 2
for which the integrating factor is y~K Hence
giving
(3.76)
*- + ^ . (3.80)
dx u dx
But, since S is a solution of (3.78),
dS/dx =p(x)S2 + q(x)S + r(x), (3.81)
and (3.80) becomes
du/dx + [2p(x)S + q(x)]u + p(x) = 0. (3.82)
This equation is of standard first-order linear form which may be
solved by the integrating factor method. Here the integrating factor is
Example 8 Consider
= j-x^dx +Q (3.88)
where C is a constant. Integrating and dividing by xe 2
(3.89)
1.
In (3.78) we consider first the case of p(x) = — 1 so that
dyldx +y2 = q(x)y + r(x). (3.91)
Writing
so that
T- = - T 4 - A ( ? ) > (3-93)
dx zdx z \dxl
(3.91) becomes 2
dz dz
—2q{x)
-q(x)--r(x)z=0. (3.94)
d?- d-x
This is a linear second-order equation for z and appropriate methods
may now be applied (for example, the Frobenius series method used in
Chapter 2 if q and r are simple polynomials). We note that if
p(x) = 4-1 in (3.78), then the substitution
2.
Secondly we consider the case of q(x) = 0 in (3.78) so that
dyldx =p(x)y2 4- r(x). (3.97)
We now make a change of the independent variable x to x', where xr
is defined by
dxf/dx=p(x)y (3.98)
or
Then
dy dy dx' dy
dx dx' dx dx'
64 Non-linear ordinary differential equations
Example 9 Consider
dy/dx = 2xy2 - 2x3. (3.104)
This is a case of (3.78) with q{x) = 0. We therefore change the variable
from x to
% = %dfx=2x% (3 106)
-
and substituting this into (3.104) and cancelling 2x gives
dy/dx' = y2 - x2 = y2 - x'. (3.107)
f
This is now a Riccati equation for y in terms of x' with p(x ) in (3.78)
equal to +1. Following (3.95), we put
1 dz
y—- (3.108)
—4 = *'*- (3.109)
This is Airy's equation discussed in Section 2.9 and hence a solution
for z in terms of x' may be found in terms of Airy functions. The
solution for y is then obtained from (3.108) and finally substituting x2
for*' from (3.105). ^
3.7 The Lane-Emden equation 65
d2y 2 dy
dx xdx
y = l, dy/dx = 0 at x = 0, (3.111)
4d2u/dt2-u + u5 = 0. (3.113)
4vdv/du - M + M5 = 0. (3.114)
Using
cos 0 = 1 - 2 sin2 (0/2), cos a = 1 - 2 sin2 (or/2), (3.130)
(3.129) becomes
Now writing
sin(0/2) = £ s i n 0 , (3.132)
where k = sin(ar/2), then
(3 134)
'w4>)' "
with O^A;^1. Similarly, the elliptic integral of the second kind is
defined by
Example 10 Evaluate
d<t>
1
Jo V(l - 4 sin
Putting 4 sin2 0 = sin2 6, we obtain
rjz/2
68 Non-linear ordinary differential equations
Example 11 Evaluate
dx
(3.138)
h J rn
0
cos
(3.139)
However, 3 — 4 sin 2 <p + sin 4 <p = (3 — sin 2 (p){\ — sin 2 <p) = (3 — sin 2
cos 2 <p. H e n c e
Figure 3.1
k=0
(a)
3.9 Buffing's equation 69
=C-ay2-lby\ (3.143)
wrx+B- <3144)
where B is another constant. For particular values of a, b and C, the
left-hand side can be expressed in terms of an elliptic integral.
Example 12 If
f l = - l y + y \ (3.145)
with boundary conditions y = 0 and dy/dx = l when x = 0, find the
x-value for which y = 1.
Proceeding as above, we have
= 2(—\y +y ) — , (3.146)
Hence
— J sin <p)
using the definition of the complete elliptic integral of the first
kind. ^
1. a = l, b=2
In this case (3.159) becomes
/ + y2 + w2 = 2C. (3.160)
The graph of this family of curves is shown in Figure 3.2. From
(3.160), we see that it is not possible to have C < 0 for any initial
conditions. For any value of C ^ 0, the curve is closed, which indicates
that only periodic solutions of the equation exist for these values of a
and b.
2. a = 1, b = - 2
In this case (3.159) becomes
y4-y2-w2=-2C. (3.161)
Figure 3.2
C=2
72 Non-linear ordinary differential equations
Figure 3.3
c—i-
Problems 3 73
- l ) ^ + >>=0, (3.163)
Problems 3
1. Obtain a solution of the equation
In x x In x
5. Show that the equation
y" = xy + 2y3 + \
f
E. Kreysig, Advanced Engineering Mathematics (Wiley, New York, 1988)
Section 1.10.
* D. W. Jordan and P. Smith, Nonlinear Ordinary Differential Equations,
(O.U.P, Oxford, 1987)
74 Non-linear ordinary differential equations
+
' W2
where C is a constant, satisfies Pinney's equation
y"' + p(x)y = C/y3,
where u and v are independent solutions of
z" + p(x)z = Q
and W = uv' — vu'. (For equations with no first derivative term
dz/dx, W, the Wronskian, is a constant). Hence show that
is a solution of
y" + y = 2/y3.
7. Show that
(l)
J2/V3 V(2M 4 - 3M2 + 1) " V2 \ V2 ' 3 / '
where F and £ are elliptic integrals of the first and second kinds,
respectively.
8. Show that the inhomogeneous Duffing equation
y" + ay 4- by3 = A cos(3cot)
has an exact solution
y = \-g-j cos(o>o,
provided
4
Approximate solutions of
ordinary differential equations
(4.3)
n=0
Provided f(x), p(x) and q(x) are also expressible as Taylor series
about x0, then the coefficients of like powers of x — x0 may be equated
in the differential equation and, by solving for the coefficients an, the
solutions of the equation may be obtained. We shall mostly be
concerned here with the case x0 = 0 for which the series (4.3) becomes
a MacLaurin series.
Example 1 Consider
dy/dx = 2e-x-y, (4.4)
given that>>(0) = 0.
75
76 Approximate solutions of differential equations
(4.5)
n=0
= 2-a0, (4.6)
2a2 = -2-au (4.7)
3«3 = (4.8)
2!"^
2
4« 4 (4.9)
~~~J\~a3'
and so on. Hence, since y(0) = 0, ao = 0 and therefore
ax = 2y a2 — (4.10)
giving
y(x) = 2x-~ Z X "i ^ 2>X T~ . . . . (4.11)
The coefficient of xn for n ^ 1 is easily seen to be
x2
y(x)=y(0)+xy'(0) + -y"(0) + ... (4.14)
and, in general,
Hence, at x = 0,
Example 2 Consider
given that>>(0) = 0.
ax = ~al (4.20)
(4.21)
\, (4.22)
4a 4 = — 2axa2 — 2aoa3, (4.23)
5a5 = -2ala3 — a\ — 2aoa4, (4.24)
and so on. Since y(0) = 0, we have from the original series ao = 0.
Hence, from (4.20)-(4.24),
0i = O, fl2 = i «3 = 0, 04 = O, 05 = - ^ , (4.25)
and so on. We have finally
where p(x) and q(x) are given functions, a power series solution does
not necessarily exist. However, provided p(x) and q(x) are
differentiate and single-valued at a point x0 (that is, they are regular
at x0), then x0 is called an ordinary point of the equation and a Taylor
series centred at x0 will provide the solutions. If, however, x0 is not an
ordinary point but nevertheless
and
lim (x-xo)2q(x) (4.29)
where m is some number. If (4.28) and (4.29) are not finite, then x0 is
an irregular singular point. The series (4.30) is known as a Frobenius
series and has been used in Section 2.5 to develop the solution of the
Bessel equation. In applying this method, certain special cases arise
depending on the two values mx and m2 of m. If these values are
identical or differ by an integer, then only one solution can be found
by this method. This was the case with the Bessel equation of integer
order n for which mx — m2 is an integer 2n and (see (2.92))
y(x)=A+ff(x,y)dx, (4.35)
x
0
f(x,yn)dxy (4.38)
i
Example 3 Consider
f'*-^, (4.39)
dx x
given y(0) = 0. (This is a Riccati equation - see Section 3.5.) Then
(4.40)
J0 A
Hence
x2 x4 x6 x8
(4 43)
- -
Each of these approximations is a finite series. This sequence
approaches the exact solution as n —• °°. ^
Example 4 Consider
dy/dx = JC + siny, (4.44)
with >>(0) = 7i 12. Then
Hence
rx 2
7t l 71 X
However,
*-2 \ i
(4.47)
2
cannot be integrated analytically, and so the next function in the
sequence cannot be found except, perhaps, for sufficiently small values
of x for which the sine term in (4.47) may be expanded. ^
given, say, _y(0) = yQ and z(0) = z0. Then the iterated sequence is given
by
Example 5 Consider
dy/dx=x+z, dz/dx=x-y2, (4.50)
where y(0) = 2, z(0) = 1. Then
r 1 9
yj(jc) = 2 + (JC + 1) djc = 2 + JC + 2* , (4.51)
Jo
ZI (JC) = 1 + [ (JC - 22) dx = 1 - 4JC + \x2. (4.52)
Jo
Similarly
Example 6 Consider
dy/dx = \?y, (4.55)
given y(0) = 0. Exact solutions of this equation are y = 0 and y = \x2.
Applying the Picard method, we find
- ^ | p(x)dx}. (4.74)
84 Approximate solutions of differential equations
dx2 V 2dx *A
Hence the solution of (4.70) may be obtained from (4.74) and the
solution of (4.75), which is an equation in normal form.
Example 8 Reduce
— ! + x — + \x2y = 0 (4.76)
dx dx
to normal form and hence obtain its general solution.
Proceeding as above with p(x)=x and q(x) = \x2, we find from
(4.74)
note that, as e—>0, the equation is singular in the sense that the order
decreases from second to zero order when e is set equal to zero. We
must therefore expect the solution to be singular as e—»0. For
example, consider
e2d2y/dx2 + y=0, (4.83)
subject to y(0) = 0, y(l) = 1. This equation has the exact solution
sin(x/6)
y (4 84)
^^u7y -
Both sin(jt/£) and sin(l/e) rapidly oscillate as 6-^0 for any given x in
the range and (4.84) becomes undefined at e = 0.
For equations of the type (4.82), the W.K.B. method gives an
approximation which is often close to the exact solution over much of
the range. In attempting to find an expression for y(x) as a power
series in e, we must have a series which is singular as e —» 0. We adopt
the trial expression for y(x) in the form
and
d2y 2
[±-2[S0(x) + el S^x) + 62S2(X) + ..
dx2 ]
1
+ ) + eS[(x) + €2S2(X) x). (4.87)
Hence, inserting (4.87) into (4.82) and cancelling y(x) on both sides,
[S()(x) + eSl(x) + e2S2(x) + ...]2
+ e[So(x) + eS[(x) + e2S'2{x) + . . .} =f(x). (4.88)
of these are
S2Q(x)=f(x), (4.89)
25 (x)S (x) + S'(x) = 0 (4 90)
Hence, from (4.89),
S0(x) = ±[f(x)]K (4.91)
while from (4.90)
i. (4.92)
In the expression for y(x) in (4.85), we require a term
expl^S^t) di\. The integral of Sx from (4.92) is simply -\ ln[S0(x)] so
that
using (4.91) (apart from constants which again can be combined into
the specification of the lower limit of integration).
The W.K.B. approximation consists of neglecting all terms in (4.85)
of order e and higher, so that only S0(x) and S^x) are required.
Hence, using (4.91) and (4.93), we have the two W.K.B. solutions
y(x)
- iTSjH; /*[/w|i •") + i7SF xp K f i«'»i! 4
(4.95)
where A and B are constants.
The solution (4.95) will differ from the exact solution to (4.82) by
terms of order e whenever f(x)¥z0. However, if points exist where
f(x) = 0 (called the turning points of the equation), (4.95) will diverge
at these points whereas a numerical integration of (4.82) will give a
finite solution there. The W.K.B. method, although a good ap-
proximation over much of the range, therefore exhibits the wrong
behaviour close to the turning points. We discuss this is more detail in
Section 4.8. Solutions which are more accurate away from turning
points may be found by including the higher order terms S2(x), S3(x)
and so on. This will not concern us here.
4.6 The W.K.B. (Wentzel-Kramers-Brillouin) approximation 87
Example 9 Consider
(4 97)
' -
where we have chosen the lower limits of integration to be zero.
Hence imposing y(0) = 0 gives
e2d2y/dx2 + xy = 0 (4.104)
(compare (2.209)).
Then/(jc) = -x and (4.95) gives
(4.105)
88 Approximate solutions of differential equations
whence
'3e i -2Lv 3 / 2 /3
(4- 106)
X*
— x$ + d\ (4.107)
Jx [P{X) £
with y(a) -y{b) = 0, where a and b are constants (or ±°°), has an
infinite number of real positive eigenvalues provided p(x) > 0 , q(x) ^
0 and r(x) > 0. Such a differential equation (with these boundary
conditions) is called a Sturm-Liouville system. The calculation of
exact eigenvalues is difficult, if not impossible, and approximation
techniques are valuable. The following example shows the use of the
W.K.B. method in this connection.
Example 11 Consider
d2y/dx2 + Xx2y = 0, (4.119)
given y(0) = 0, y(jz) = 0. Writing this as
\d2y
id—xy (4 120)
-
and comparing with the basic form (4.82), we see that f(x) = —x2 and
e2 = I/A. Since e is assumed to be small, we expect the method to give
the large eigenvalues A with good accuracy. The general solution of
90 Approximate solutions of differential equations
or
VkJT2/2 = njt, (4.124)
where ft is an integer (which must be large in order that A itself is
large). Hence approximate eigenvalues are given by
)2, (4.125)
for large n. ^
where 6mn is the Kronecker delta (see (1.23)). Such eigenfunctions are
said to form an orthonormal set. Since the W.K.B. method of
Example 11 gives the eigenfunctions ((4.122) with D = 0 and A given
by (4.125)) up to an arbitrary multiplicative constant, this constant
may be fixed by (4.126). Hence the eigenfunctions and eigenvalues for
large n are uniquely determined.
d2y/dx2=f(x)y (4.127)
4.8 The Liouville-Green technique 91
into another, also in normal form with the same number of turning
points, which is soluble exactly or approximately in terms of known
functions.
In (4.127), we make a transformation from x to a new variable £ by
writing x=x(%)> and from y{x) to a new dependent variable G(§),
where
y(x)= (4-i28)
W?'
with f-' = d^/dx. The inclusion of the factor (§')~* ensures that the
equation for G(£) is also in normal form, as we now show. From
(4.128),
Further,
the terms in dG/dt; cancelling. Inserting (4.131) into (4.127) and using
(4.128), we find
(f(x) , A i^ (4 132)
,2 ^ , 2
where
(4-133)
and hence
Bz~\ (4.135)
Since %'2=f(x), we have
]* dt. (4.136)
Combining (4.135) and (4.136) with (4.128) gives the W.K.B. result
(4.95) (with e formally equal to 1). However, if the equation (4.127)
has turning points, that is, if f(x) has zeros in the range of interest,
then we clearly should not choose %'2=f(x) since £' 2 is always
positive whereas/(JC) changes sign. This is the origin of the divergence
in the W.K.B. solutions at the turning points of the equation. The
correct way to proceed is to choose some function h(%) having the
same number of zeros as/(x) in the given jc-range, and then let
h(%)%f2=f(x). (4.137)
The choice of h(%) should be such that A can be neglected in (4.132)
and such that the resulting approximate equation for G(§),
d2G/d^2 = h(%)G, (4.138)
is soluble in terms of known functions. We now illustrate various
features of this technique.
Example 12 Consider
d2y/dx2 = exy, (4.139)
where -oo<jt<o°. Then performing the Liouville-Green transforma-
tion (4.128), we find
d2G I e* \
~d]=2 = \]=;2*A)G' (4.140)
We now choose
£' 2 = e*, (4.141)
which gives on integration
§ = 2e*/2. (4.142)
Hence, for -oo<x<oo, w e have 0 < £ < o o . From (4.142) the deriva-
tives up to £'" can be calculated and substituted into the expression
(4.133) for A. We find
4.8 The Liouville-Green technique 93
(4.144)
where A and B are arbitrary constants and / 0 and Ko are the modified
Bessel functions. Now, from (4.141) and (4.142), (§')"* = e~*/4 =
(2/£)i From the inverse transformation (4.128) relating G(§) to y(x),
we have finally
y(x) = AI0(2ex/2) + M 0 (2e x / 2 ), (4.146)
where A and B are arbitrary constants. ^
Example 13 Consider
d2y/dx2 = x4y (4.147)
for x > 0. Transforming as before, we find
Now choosing
%'2 = x\ (4.149)
we have
? = 3*3 (4.150)
and, from (4.133),
A = -2/9§ 2 . (4.151)
Hence (4.147) becomes
f-('-£)°-
Writing %=d/2, (4.152) becomes
d2G /. 2 \
^^ 2 \ 9^ 2 / v }
Comparing (4.153) with (2.160) and its solution (2.163), we see that
v2 - \ = —\ giving v = \, and the solution of (4.153) is therefore
(4.154)
(4.155)
94 Approximate solutions of differential equations
( ) ( Y ^ (4.159)
y—\%. (4.161)
Example 15 Consider
d2y/dx2 = N2x(l + x2fyy (4.163)
with -oo <x < oo, and TV large. This equation has a single turning point
(since the function x(l + x2)2 changes sign only at x = 0). Carrying out
4.8 The Liouville-Green technique 95
d2G/d%2=%G, (4.166)
so that on integration,
§ | ^ = N(§|*|Ml^). (4.168)
Hence
§ = Aflc(l + Wf. (4.169)
The solution of (4.166) is
+ D &"(§), (4.170)
and hence the approximate solution for y{x) is found from y(x) =
(£')~^G(§) since (£')"* can be evaluated in terms of x, and § can be
substituted using (4.169). Further, from (4.169), A can be calculated.
It is found that A is bounded (that is, finite for all x) and that it has an
overall factor N~% which is a small quantity for large N. It is therefore
small over the whole range of x> and may be neglected in (4.164). ^
Problems 4
1. Find the first three non-zero terms in the series solution of the Airy
equation
d2y/dx2 = xy,
dx2 dx
in the form
dx2 dx
to normal form, and hence find its general solution.
7. Use the W.K.B. method to find approximate solutions of the
equation
d2y/dx2 = N2x4y.
8. Use the Liouville-Green transformation to obtain solutions of
d2y/dx2 = xny
for x > 0 , where n is a positive constant, in terms of the modified
Bessel functions. (See Example 13.)
5
Contour integration
Figure 5.1
100 Contour integration
or equivalently
(5.9)
co = yz = (5.10)
(o2 = V7
i(0/2+jr)
7e = cos(0/2 + JT) + i sin(0/2 + (5.12)
These two functions are called the two branches of co = y/z, and
a) = y/z is termed a multi-valued function. If we give r and 0 a
succession of different values then o)x traces out some curve in the
complex plane; a)2 is obtained by rotating the curve through n about
the origin according to (5.11) and (5.12) (see Figure 5.2). Two values
of \lz exist for all z except z = 0, which is called the branch point. It is
important to realize that the two curves in Figure 5.2 are not the
graphs of co = \jz but only the paths along which z moves as we
change r and 6 (or equivalently x and y). By analogy with the above,
co = z^ will have three branches with a branch point at z = 0, while
(o = y/[(z — a){z — b)] will have two branches with two branch points
at z = a and z = b. We shall expand the discussion of branch points
and multi-valued functions in Chapter 6.
Figure 5.2
- • • * •
5.2 Exponential and logarithmic functions 101
ez = l + z + ^ + ^ + . . . . (5.14)
siny = i ( e * - e - * ) , (5.18)
sinz = ^ ( e i z - e - i z ) . (5.20)
with the other identities of real variable theory remaining true when
the real variable x is replaced by the complex variable z.
Closely related to the exponential function e2 is the logarithmic
function In z. If we write
z = eM (5.25)
then
w = \nz = \n(r^e+2njt)), (5.26)
from (5.3). Hence
a) = In r + i(0 + Inn), (5.27)
where n = 0, ±1, ±2, . . . , and we see that co = In z is a multivalued
function. Writing co = u + iv, we have
u = \nr, v = d + 2nji (5.28)
for the real and imaginary parts of co. When n = 0 and —Jt<6^jz, co
takes its principal value and co = In z is then a single-valued function.
Since r = \z\ and 6 = arg z, we can write
a> = In \z\ + i(arg z + 2AZJT). (5.29)
Sometimes, when the principal value of co is taken, we write w = Lnz,
the capital letter L meaning that n has been put equal to zero in
(5.29). We now give examples and include some elementary results.
Example 1
(a) ln(-2) = In | - 2 | + i(arg(-2) + Inn). (5.30)
Hence, since —2 = 2e™, |—2| = 2 and arg(—2) = ny and therefore
ln(-2) = In 2 + i(jr 4- 2AZJT). (5.31)
(b) Ln(-i) = In | - i | + i arg(-i). (5.32)
72
Hence, since —i = le"™ , |—i| = 1 and arg(—i) = — jr/2, and therefore
Ln(-i) = In 1 + i(-jr/2) = -jri/2. (5.33)
(Similarly Ln(i) = jri/2).
(c) Complex powers a of z can be established using the relation
za = ealnz, (5.34)
where In z is defined by (5.27). For example the principal value of i1 is
given by
ii = eiLni = ei(.i/2) = e - ^ (535)
2 s iin f t , ^ ^ (5.37)
Hence
e 2 i w -2ie i f t > z-l = 0. (5.38)
lft
This is a quadratic equation for the quantity e \ so that
eia> = ^[2iz ± V(~4z 2 + 4)] = iz± V(l - z 2 ). (5.39)
Taking the logarithm of (5.39), we find
\o) = ln[iz ± V(l - z 2 )]. (5.40)
Now, putting z = 2,
ia) = ln[(2±V3)i]. (5.41)
Hence
Since(2-V3)(2
ln(2 - V3) + ln(2 + V3) = 0 (5.44)
and therefore (5.43) may be written
^ ^i (5.45)
(5.46)
oz
104 Contour integration
Equivalently, we have
(5.47)
Z — Zo
along any path. Functions which do not satisfy (5.48) are discon-
tinuous at z0; z0 is then said to be a singularity of the function. We
shall discuss this concept in more detail later.
Figure 5.3
5.4 The Cauchy-Riemann equations 105
Figure 5.4
it
2, + dz
c,
c, c2
2,, c2
o
106 Contour integration
f-ff. ^ (5.61)
dz dco dz
Example 4 Consider
(0 — e2 = ex+iy = e*(cos y + i sin y). (5.62)
Then
u=ex cos y, v = e* sin y (5.63)
and
du/dx = e* cosy = dv/dy, (5.64)
x
du/dy = —Q siny = —dv/dx. (5.65)
5.4 The Cauchy-Riemann equations 107
Example 5 Consider
o) = l/z. (5.69)
Then
= == =
0) ; ^ ~ ^ ^— 1 ^ r. (5./U)
2 2
x + iy x +y x2 + y2 x2-\-y2 v
Hence
and
du j?->? dv
2 2 2 J
dx {x + y ) dy' "
P j
3y (x + / ) 5JC'
Consequently, du/dx^ dv/dy and the function is not analytic for any
x and y.
Similarly, consider
z — z0
*: 7-*\
(5 79)
v i^tr -
H e n c e , if z — z 0 = re 1 0 ,
Z ZQ Z ZQ A C
and
\z\2~\z,\2
= z* + z0[cos(2^) - i sin(20)]. (5.81)
z-z0
As z^z0, the limit depends on 6, the angle at which z approaches
z0. ^
= e(° (5.82)
which implies
da)/dz = e-M = l/z, (5.83)
except at z = 0 (In z is not analytic at z = 0).
5.6 Complex integration 109
Each integral on the right of (5.84) is a real line integral and may be
evaluated by integrating from A to B along the specified path C (the
equation of which will, in general, be given as y = y(x), say). Since the
complex integral can be put in the form (5.84), the usual rules of
integration which apply to real integrals must also apply to complex
ones. Accordingly, we have
Figure 5.5
B(z2)
- • *
O
110 Contour integration
containing the points A(zx) and B(z2). In this case we can show that
which implies that the integral is independent of the path C since the
result depends only on the values of / ( z ) at the two end points.
Consider
(5 89)
-
f
using the form for f (z) given in (5.53). Then
= j 2 d u + i[2dv, (5.93)
= [/(z)E=/fe)-/(*i). (5-94)
This proves the result that the integral is independent of the path. The
following examples illustrate these ideas.
Example 7 Evaluate
[ z2dz, (5.95)
Jc
where
(i) C = Cx is the straight line y = x from A(0, 0) to B(l, 1),
(ii) C = C2 is the x-axis from (0,0) to A(l, 0) and the straight line
from A(l, 0) to B(l, 1),
(see Figure 5.6).
5.6 Complex integration 111
=J [(x2-y2)dx-2xydy]
cx
+ i I [2xy dx + (x2 - y2) dy]. (5.96)
I z2dz= I [(x2-x2)dx-2x2dx]
Jcx Jo
f z2dz=i \{x2-y2)dx-2xydy]
Jc2 J(o,o)
+r i) [(^->' 2 )^-^^]
J(l,0)
r(l,l)
J (1,0)
[2xy dx + (x2 - y2) dy], (5.100)
Figure 5.6
jt
C2
0(0,0) C2
112 Contour integration
where the integrals are along the straight lines joining the end points.
Along OA, y = 0 and hence dy =0. The first two integrals in (5.100)
therefore give
f xx22dx
dx ++ i\i I 0dx = l (5.101)
Jo Jo
Along AB, x = 1 and dx = 0 and the second pair of integrals in (5.100)
gives
The results (5.98) and (5.103) are the same. This is to be expected
since z 3 /3 (the function which differentiates to give z2) is an
analytic function and hence the value of the integral is \{\ 4- i)3 =
if(z)dz = 0, (5.104)
Figure 5.7
-•*
5.7 Cauchy's theorem {first integral theorem) 113
Now if, in general, P(x, y) and Q(x, y) are two continuous functions
with continuous first partial derivatives (at least), then
Example 8 Evaluate
I=lzzzdz, (5.108)
Example 9 Consider
(5.109)
— Z
Writing
(5.111)
I=4>zndz, (5.113)
Figure 5.9
5.7 Cauchy's theorem (first integral theorem) 115
f (5.115)
Hence
Figure 5.10
-•*
o
116 Contour integration
£^-dz, (5.119)
Jc z - z0
where C is any closed curve which lies in R and includes the point z0.
To prove this, we surround the point z0 by a small circle Cx, radius a.
By the argument at the end of Section 5.7, we have
f f(z) f f(z)
4> jL±-L dz = 4) - ^ ^ - dz, (5.120)
z
Jc — z0 Jcx z — z 0
since the whole function f(z)/(z — z0) is analytic between C and Cx.
Now let z = z0 4- aeld describe the circle Cx so that on the right-hand
side of (5.120), dz = iaewd6 and z-zo = aeie. Then
IS L
4 - dz=\ f{zo + aee)—td8, (5.121)
J Cl z - z0 Jo ae
which, as the small circle Cx shrinks to zero, becomes
Example 11 Evaluate
Example 12 Evaluate
2z-l
/= -dz9
where (i) C = Cx is the circle \z\ = 4 and (ii) C = C 2 is the circle \z\ = \
(see Figure 5.12).
Figure 5.11
118 Contour integration
2z - 1 _ 1 1
+ (5.126,
(z + l ) ( z - 2 ) ~ z + l z-2'
Then
Z
I _2\dz==j) ~ ^ 7 + ^ ~Zy' (5.127)
(i) For Clf both singular points z = —1 and z = 2 lie within Cx.
Hence, by the Cauchy Integral Formula,
r 2jri.l. (5.128)
v 7
l 2 z-2
Hence
/ = 2jri(l + l) = 4jri. (5.129)
(ii) For C 2 , only the singular point z = — 1 lies within C 2 . Hence the
first integral is
^ ^ 1. (5.130)
Jc2 z + 1
• = 0. (5.131)
c2 z ~ z
Hence
^ (5.132)
Now, by (5.119),
1 r f(7\
(5.134)
Jcz-zo
and
<5 135)
-
5.9 Derivatives of an analytic function 119
dz (5.136)
z -zo-h z - zj h
<5 138)
^?*- -
In a similar way, we can obtain the general result
V2w=!^ + ^ = 0. (5.141)
Similarly
d2v 32v
V 2 u=—o2 + ^ = 0. (5.142)
dx dy2 v
'
Example 13 Evaluate
Q
I=j> _U2dz> (5.143)
C
<f> 2 dz = 2mf'(z0 = 1) = 2;ri . 2e2 = 4;rie2. ^ (5.144)
Jc {z — 1)
120 Contour integration
1. Taylor's Theorem
If /(z) is analytic within and on a circle C of radius R and centre
z0, then
Kz ~ zzo)
o) • (5.145)
n=0
This series converges within the circle, centre z0, of radius R equal to
the distance from z0 to the nearest singularity of /(z). In general,
/ (n) (z 0 ) can be found from (5.139).
Figure 5.13
>• x
5.10 Taylor and Laurent series 121
2. Laurent's Theorem
This is an extension of the preceding theorem and states the following:
If/(z) is analytic inside and on the boundaries of the circular annulus
defined by \z — zo\=Rr and \z — zo\ = R2 with R2<R\ (see Figure
5.14), then/(z) can be represented by the Laurent series
dz (5.154)
and
(5.155)
Example 16 Consider
(5.156)
which has an isolated singularity at z = 0. Then since
£; zn
e - + z + —+ . . . + — + . . . , (5.157)
Figure 5.14
5.10 Taylor and Laurent series 123
Example 17 Not all principal parts of Laurent series are finite series.
Consider, for example,
( ^ ^ ^ ) (5.160)
giving
(5 166)
[(^(^ij] ^ -
Example 19 Consider
/(z) = - ^ - ( l - z ) - \ (5.168)
124 Contour integration
Now, since 0 < | z | < l , the second term can be expanded by the
Binomial Theorem to give
- - ( l + z + z2 + z3 + . . . ) • (5.169)
Again, since \z\ > 1, the second term can be expanded by the Binomial
Theorem giving
=
? + ? + 7 + -'-- (5 172)
-
In the above we have chosen z0 = 0 as the centre of expansion. If we
choose instead zo=l, then we obtain two series in powers of z — 1, one
for 0 < \z — 1| < 1 and one for \z — 1| > 1. The first of these is obtained
by putting z — 1 = u so that
X
^ {l-u + u2-...)y (5.173)
z(z - 1 ) u 1+w u
since \u\ < 1, giving
-174)
Example 20 Consider
(^) (5.175)
for |z| > 1 (recalling that the capital letter indicates the principal value
of the logarithm). Then
Hence
l/z 2 ). (5.177)
5.11 Singularities and residues 125
z2 2z 4 3z e
1 1 1
(5-178)
r3i^+5i^-" (5 180)
-
has an essential singularity at z = 0.
126 Contour integration
1. Simple poles
Clearly if we have a simple pole at z = z0, then the Laurent series
(5.153) will have the form
\ \ -l, (5.187)
z-*oL z(z — 1)J
as before.
5.11 Singularities and residues 127
For z = 1,
a
-^diz-l)^hj]=l' (5 188)
-
as above. ^
cos(w + 3JT/2)
sin
We now expand the functions cosw and sinw as power series in u
so that
2! 4!
We may further expand e2" as a power series and use the Binomial
Theorem to expand the last factor. Then
x ^ l + - - . . . j. (5.192)
Since all the expansions involve only positive (or zero) powers of u>
the only term with a negative power of u is —e3jr/w. Hence the
principal part of the Laurent series of / ( z ) in the neighbourhood of
z = 3JT/2 is
(5.193)
128 Contour integration
= lim [ ( z - 3 j r / 2 ) e 2 z ~ l . (5.194)
Z^>3JT/2L COSZJ
Now both z — 3JT/2 and cos z tend to zero as z—> 3;r/2. Hence
m (5 197)
'W) -
and that z = z0 is a simple zero of g(z) (that is, g(z) behaves like z — z0
as Z ^ ' Z Q ) . Then z = z0 is a simple pole of / ( z ) provided p(z 0 ) =£0 and
that p(z) is not singular at z = z0. Using (5.183) we can now write
, (5.198)
a_x = lim i
z^z 0
z -z0 J
5.11 Singularities and residues 129
a_l=p(z0)/qf(z0). (5.200)
Example 23 Consider
/(z) = jrp- (5-201)
The pole at z = 1 is simple (since z = 1 is a simple zero of
1 - z 3 = (z - 1 ) ( - 1 - z - z 2 ) ) . Hence, since p(z) = l and q(z) =
1-z 3 ,
2. Po/es 0 / order m
The Laurent series (5.153) for a function f(z) with a pole of order m
at z - z0 has the form
Example 24 Consider
1 2
(5.209)
(5.210)
as in method 1. ^k
Example 25 Consider
(5 213)
-?-)+-- -
From the Laurent series (5.213), we see that the residue of / ( z ) at
z = 0 is —1. Alternatively, using (5.206), we have
<5214)
as follows:
-z3-|z4
lim (5.217)
z—•()
where the singularity is termed a pole of order m for any finite m, and
an essential singularity for m infinite. Suppose we require
=1 f{z)dz, (5.220)
-dz.
1 (5.226)
T y (z - z 0 ) Jo
Accordingly,
i f(z)dz = 2 (5.227)
Jc
From Section 5.11, a_x is the residue of f(z) at the pole of order m at
z = z0. Since we know how to calculate a_x for a given function/(z), it
follows that we can now calculate $cf(z) dz around any closed
contour C surrounding a singularity at z0. The result (5.227), known as
the Cauchy Residue Theorem, may be readily extended to the case
where /(z) has a number of poles z0, zx, z2, . . . , zk within some closed
contour C (the case of three singularities is shown in Figure 5.15).
Figure 5.15
-•*
5.12 Cauchy residue theorem 133
Forming the cuts in the usual way, we see that the integral around C is
just the sum of the integrals around circles centred on each singularity.
Hence
where a_ly b_u c_i, . . . are the residues at z0, zx, z2, . . . . We there-
fore have
•--Sl'STl)]- 1 - (5.231)
(5.232)
-dz, (5.233)
Jc, z(z - 1)
where Cx is some closed curve not containing a pole. Then the
Figure 5.16
134 Contour integration
I e2
4> — — dz=2;ri(e-l), (5.235)
Jc,z(z-1)
Problems 5
3. Show that the function e2 has no zeros, and locate the zeros of
cosh z.
4. Given that
Use this result to find the most general forms of g(z) and h{z)
given that g(z) and h(z) are analytic for all z, and that
(p = |g(z)| 2 + \h(z)\2 is harmonic (that is, it satisfies Laplace's
equation V20 = 0).
Problems 5 135
6. Evaluate J (z 2 + 1) dz
(i) along the straight line joining the origin, z = 0, to z = 1 + i,
(ii) along the parabola x = t> y = t2 where 0 =^ t ^ 1,
(iii) along the straight line from the origin, z = 0, to z = 1, and
then along the straight line from z = l to z = 1 + i.
Why are the results the same?
7. (i) Evaluate
dz
_ z-2'
where C is (a) the circle \z\ — 1; (b) the circle \z + i| = 4.
(ii) By writing z = e10, evaluate jcLnzdz and J c z 3 L n z d z ,
where C is the curve defined by —JZ<Q<JZ.
8. Evaluate the following integrals using the Cauchy Integral For-
mula, where C is the circle \z\ = 2:
r sinz
(o
9. Evaluate the following integrals using the Cauchy Integral For-
mula, where C is the boundary of a square the sides of which are
defined by x = ±2, y = ±2:
tan(z/2)
• (z - J T / 2 ) 2 Z
'
10. Find the Taylor series about z = 0 and its radius of convergence
for each of the following functions:
(Ui)
z\z - l) 2 (z - 2) a t a U t h C P lle S
° *
lz
14. Locate the poles of z/(l — ae~ ), where a > 1 is a real constant,
and hence deduce that
6.1 Introduction
Using the basic theorems of the last chapter, in particular the Cauchy
Residue Theorem (5.229), we may now apply contour integration to
two specific problems: (i) the calculation of real integrals, and (ii) the
summation of infinite series.
Example 1 Evaluate
dd
137
138 Applications of contour integration
Choosing the unit circle C (as above) for which z = e10, we have
dO = dz/iz. (6.4)
Using (6.1) and (6.4), (6.3) becomes
dz/iz dz
/ = 2 (6.5)
2-\(z + Hz) iJcz -4z
Now the poles of the integrand occur where z2 — 4z + 1 = 0, that is, at
z = 2 ± V3. Both are simple poles since z2 — 4z + 1 = [z — (2 —
V3)][z - (2 + V3)], but only the pole at z = 2 - V3 is within C (see
Figure 6.1). Hence, to evaluate the integral by the Cauchy Residue
Theorem, we require only the residue at z = 2 - V3 which is (by
(5.183))
. (6.6)
(2-V3)-(2 2V3
Hence
dz
(6.7)
Example 2 Evaluate
** cos(2^)
(6.9)
l0 5 - 4 c o s ^ '
6.2 Calculation of real integrals 139
z2+l/z2 dz If z4+l
rfz, (6.11)
iz 4i Jc z\z - i)(z - 2)
where the contour C is the circle \z\ = 1. The poles of the integrand
are as follows:
(i) a pole of order 2 at z = 0;
(ii) a simple pole at z = \\
(iii) a simple pole at z = 2.
Of these, only the first two lie within C (see Figure 6.2) and hence we
need the residues at these two poles only.
-1
Figure 6.2
140 Applications of contour integration
Residue at z =\ Here the pole is simple and the residue is (by (5.183))
(6.13)
Figure 6.3
6.2 Calculation of real integrals 141
Further
Hence, provided
f(x) dx = 2m
c
^ lim [ \f(Rci6)\Rdd
fn At
^ lim -1Rd6 = 0, (6.20)
R-+°c JQ R
where we have used the fact that the modulus of the integral of a
function is less than or equal to the integral of the modulus of that
function. The vanishing of the modulus of the integral o n C ' a s / ? ^ o o
implies that the integral itself is zero.
Example 3 To evaluate
r
-rrn> (6-21)
where C is the closed contour of Figure 6.3. The conditions (i) and (ii)
above are satisfied since f(z) = (z4 + a4)~l has simple poles where
z 4 = -a\ and |/(z)| behaves like R~4 for large \z\=R. The four poles
142 Applications of contour integration
are at
zk = ae{jt+2kn)l\ k = 0, 1, 2, 3, (6.23)
and only two of these lie in the upper half-plane:
z 0 = aelJT/4, zx = ae3jtl/4. (6.24)
To find the residues at these poles, we use (5.200) with p(z) = 1 and
q(z) = z 4 + a4. Hence the residue at z0 = azxnl4 is
_J_ 1
3 (6.25)
~4z
whilst that at zx = ae3jTl/4 is
Piz) J_ (6.26)
:
4z 3 4a3
Finally, using (6.19), we find
f
i)1
4a L 3
V2 V2 J
n
(6.27)
at
lim
R
f(z)elkz dz
I \f(Reie)\ \eikReie\ R dd
•nil
— Q ~kRsind Rdd = lim —kR sin
(6.28)
6.2 Calculation of real integrals 143
Example 4 Evaluate
Here
where Re stands for the real part of the integral. The integral in (6.32)
now has the form of the left-hand side of (6.30). Consequently we
consider
where C is the contour of Figure 6.3, as before. The conditions (i) and
(ii) are satisfied since \f(z)\ = |l/(z 2 + a2)\ behaves like R~2 on \z\ = R
as /?—> ». The poles of / ( z ) are at z = ±\a of which only z — \a lies in
the upper half-plane. The residue at this pole is (using (5.183))
T eikz I e~ka
lim v(z - in) — — =—-. v(6.34)
^i«L (z-ifl)(z + ifl)J 2\a '
Hence
r zxkx (z~ka\ n v
"1 5 ^ = 2;ri —— = - e , (6.35)
Loo x2 + a2 \2\a) a v }
l = ^~ka. (6.36)
144 Applications of contour integration
[C^dX=Yai (6.39)
Figure 6.4
• • x
6.3 An alternative contour 145
dz. (6.41)
+ ^2 dx +i e-(-"1+ij;)2 i dy = 0, (6.42)
•'Vi
(6.43)
Figure 6.5
4-iy,
Cj •f 1
146 Applications of contour integration
and hence
\ ^ (6.47)
J — oo
(6.53)
6.4 Poles on the real axis: the principal value of integrals 147
(6.54)
= 2n\ x (sum of residues at the poles of f(z) inside C). (6.55)
To evaluate the last integral in (6.54), we use the Laurent expansion
of f(z) about z =x0 in the form
where a_x is the residue at z =x0. The basic results (6.54) and (6.55)
become, as /?—»o° and <5—»0, using (6.58),
r rxo-6 rR -i
-R
148 Applications of contour integration
(a-\\
poles of/(z) in the upper half-plane) + 2m\— I, (6.59)
(a-\\
in the upper half-plane) + 2jri( ). (6.60)
If more than one pole exists on the real axis, (6.60) generalises to
pf f(x)dx = 2m
x [(sum of the residues at the poles of /(z) in the upper half-plane)
+ ^(sum of the residues at the poles of /(z) on the real axis)]. (6.61)
— dx, (6.62)
(6.64)
z—0 \
6.4 Poles on the real axis: the principal value of integrals 149
and
•£ -dx = 0 (6.66)
sin(mx)
£ dx = n. (6.67)
( ) la1
= P(l-e-"»). (6.74)
Taking the imaginary part of each side and noticing that the integrand
is an even function, we have
as in (6.36). ^
f xj(x)dx, (6.77)
6.5 Branch points and integrals of many-valued functions 151
izj(z)dz, (6.78)
Jc
we notice that z a is a multi-valued function and that z = 0 is a branch
point. In Section 5.1, we discussed the functions z^ and z^ and found
that if the argument of z is changed by 2JZ (in other words, if a branch
point is encircled) then we obtain another branch of the function and a
different value from the starting value. In using the contour integration
technique, therefore, the contour must be chosen to exclude the
branch point so avoiding multiple values. Some integrals involving
multi-valued functions can be evaluated using a suitably indented
semicircle, and we illustrate this with two examples.
Example 9 Consider
;dx, (6.79)
1+jt 2
(6.80)
where C is the indented closed contour shown in Figure 6.7. The small
circle, radius d, centred on z = 0 ensures that the branch point z = 0 is
outside the contour. As /?-><» on the large semicircle z = Rew,
f \[za/(l + z2)]dz\ behaves like Ra~l which tends to zero. On the
small semicircle z = deld
r Xa
1+jt 2
~ dx +
f™ (xe*1)"
Jo 1+jt 2
T- dx = 2K'\
/e^ri/2\a
2i
:— . (6.83)
152 Applications of contour integration
Finally then
Example 10 Consider
where a > 0.
f x +a
(6.86)
j> zln_Ja2dz- ( 6 - 87 )
On the large semicircle, J |[(ln z)/(z 2 + a2)] dz\ behaves like R~1\nR
which tends to zero as /?—»oo. On the small semicircle z = dew,
lnz
2 * ^ . (,90,
Jo Jt2 + fl2 Jo JC2 + « 2
Expanding the logarithms into real and imaginary parts, and taking the
principal value of ln(ia) since all arguments within C lie between 0 and
jt, we have
lnfl+i
2 -2 -2dx + n\\ — 2=- T I (6.91)
Jo JC2 + a 2 J o Jt2 + a 2 aV 2/
The real part of (6.91) gives
^lna. ^ (6.92)
v
2a '
6.5 Branch points and integrals of many-valued functions 153
It should be noted in the above two examples that the integral from
-oo to 0 included, or was a multiple of, the required integral from 0 to
oo. This arose because the single-valued parts of the integrands
(1/(1 + x2) and Il(x2 + a2)y respectively) were even functions of x. If
this is not the case then we require an alternative contour which also
excludes the branch point. A suitable contour is shown in Figure 6.9.
The contour C is made up of the large broken circle (radius R), the
two straight lines along the x-axis separated by an infinitesimal
distance (the 'branch cut'), and the small broken circle (radius 5)
surrounding the branch point z = 0. We shall see that, due to the
multi-valued nature of the integrand, the contributions above and
below the branch cut do not cancel. The following example illustrates
the use of this contour.
Example 11 Evaluate
/ = :dx. (6.93)
We therefore evaluate
V2
(6.94)
Hence
r o dz = 21 = 2jri
r
c 1 + z3
x [sum of the residues of Vz/(1 + z3) at all its poles]. (6.96)
zx = ^ ' \
z2 = e ^ = - l , (6.97)
z3 — e
= ^ ( - i + i-i) = | . ^ (6.99)
jt COS(JTZ)/(Z)
= /(/!). (6.100)
(d/dz) sin(jrz)
We now integrate n cot(jzz)f(z) around a square S with corners at the
points z = (N + 2)(±1 ± i)> where N is a positive integer (see Figure
6.6 Summation of series 155
where the first term is the sum of the residues of n cot(jrz)/(z) at the
poles of cot(Jtz). On the horizontal sides of S> z=x + iy where
y = AT + i > i. Then
\JIZ
|C0t(7Tz)| = e~H
— e"ijrz
2 fin)
n = —°°
= -[sum of the residues of n cot(jrz)/(z) at the poles of/(z)],
(6.104)
Figure 6.10
i
\ Q
156 Applications of contour integration
Example 12 Consider
i-2XT=1 +i+i+
* + ---- <6-106)
n=on +1
n
=-- coth n, (6.109)
whilst that at z = - i is
n
= - - coth n. (6.110)
v-^ 1 + 71
2 (6.111)
n%n +l " 2
Similarly for the series
(6.112)
6.6 Summation of series 157
we write
(6113)
i^^=i+^ ? Sri-
Using (6.104) we have
(6.116)
Often the series we wish to sum begins with the n = 1 term, and the
function / ( z ) has a pole at z = 0, corresponding to the n = 0 term. In
this case, since the n = 0 term diverges in the sums on the left-hand
sides of (6.104) and (6.105), we omit the term n = 0 and the residue at
z = 0 is included on the right-hand sides. We illustrate this with an
example.
Example 13 Consider
i ^ = l +H U f 6 + .... (6.117)
n=\ n
Then / ( z ) = 1/z2 has a pole at z = 0. Hence
2 - 2 = - r e s i d u e of ^—-atz=0 , (6.118)
n=-^n L z J
since z = 0 is the only pole of/(z). To find the residue here, we expand
n cot(jrz)/z2 as a Laurent series about z = 0:
n cot(Jtz) n COS(JTZ)
z2 z 2 sin(jrz) z2/ JTV
(
158 Applications of contour integration
The residue (the coefficient of 1/z) at this pole (of order 3) is therefore
- J T 2 / 3 . From (6.118) we have
so that
oc
X l / « 2 = i 2 /6. (6.121)
we have
oc
(6.124)
/
Hence the residue (the coefficient of 1/z) is JT2/6 and consequently
S ^ - = " 2 2 ^ = ^ - -^ (6-125)
n =i n rt=_oc n 12
Problems 6
1. Evaluate by contour integration around the unit circle \z\ = 1:
Jo 5 - 4 c o s 0 Jo 2 - c o s 0
2. Show by contour integration that
2
f2"" dd In
Jo l -
Problems 6 159
4. By integrating
(z + 4)2 2
f zdz
/=
rx^nx_dx = jr
Jo 5 - 4 cos x 2
9. Evaluate
(i)
I (l+x2)2(l+x)dX =
2\/2\l~^2)'
11. Evaluate
7.1 Introduction
In the mathematical analysis of many linear problems it is often useful
to define an integral transform of a function f(x). The general form of
such a transform is given by
I}, (7.5)
161
162 Laplace and Fourier transforms
where a and j3 are arbitrary constants. Equations (7.3) and (7.5) show
that /{ } is a linear operator. In order that f(x) may be obtained if
f(s) is given, we now introduce the inverse operator I~1{ } which is
such that if
•} = /(*) (7.6)
then
f(x) = rl{f(s)}. (7.7)
{ ]=
__,, i i \
\s — a s + a/ s —a
Table 7.1
2{f(x)} = /(*)
1 1/5 (5>0)
x n ( / i = O , 1,2, . . . )
1
5 —a
5 / ^ A\
COS(OJC)
sinh(ajc)
5 2 -a 2 (
^|J|)
s
cosh(fljc)
s2-a2 (j
'|a|)
(7.12)
Jo
2.
If
2{f(x)}=f(s) (7.14)
then
2{xJ(x)} = (-!)"£; f{s), (7.15)
where n = 1, 2, 3 . . . .
164 Laplace and Fourier transforms
•=/(*)«(*). (7.26)
Example 1 Since
2
^{sinClx)} = -0 , (7.27)
S! ~f~ 4
using the standard result of Table 7.1, then the Shift Theorem states
that
(7.28)
(s + a)2 + 4 '
Example 2 Since
1
(7.29)
' 5+2
Figure 7.1
•+•*
166 Laplace and Fourier transforms
from Table 7.1, then the second theorem above states that
Example 3 Given
%\\\aucos[b(x-u)]du\=f{s)g(s) = - S
^ (7.34)
KJQ J \S — a)\s -TU)
Now
(7 38)
-
7.5 Laplace transform of an error function 167
Hence
Now since (using Table 7.1) J£{JZ/2} = it 12sy we have, for JC > 0,
Clearly if x = 0 then
r *M*u)du = 0 (742)
Jo u
and if x < 0 , from (7.41),
it
du = - - . (7.43)
= - ^ - f e~u2du. (7.46)
yji Jy
The particular transform needed is
V}, (7.47)
168 Laplace and Fourier transforms
(7.51)
Now writing
u2 + ^ 42 = O - aVs/2u)2 + a> r7 ^
4M
(7.55)
Figure 7.2
i a
\
7.6 Transforms of Heaviside and Dirac functions 169
Figure 7.3
H(x-a)
i
1 i
o a
170 Laplace and Fourier transforms
= f e-sxH(x-a)y(x-a)dx
Jo
+
f Q-s(v °)y(v) dv = e~say(s), (7.65)
Jo
[ d(x-a)dx=\9 (7.67)
^ — 00
I 6(x-a)f(x)dx=f(a). (7.68)
f \/(n/jt)e-nx2dx = l, (7.70)
J — 00
then
= H(x'-a), (7.73)
using (7.58). Hence by formally differentiating each side we obtain
S(x-a)=-^-H(x-a). (7.74)
This shows that the delta-function is not a proper function in the usual
sense since the right-hand side of (7.74) is the 'differential' of a
discontinuous function.
The Laplace transform of the delta-function is easily obtained as
follows:
1. Ordinary derivatives
Suppose y =y{x) is a Laplace transformable function. Then
(7.79)
(7-80)
172 Laplace and Fourier transforms
Table 7.2
2{f(*)} =/(*)
dy
sy(s)-y(O)
dx
s2y(s)-sy(O)-y'(O)
dx~2
dy(s)
xX dy S
dx ds }(S)
d2y{s) dy{s)
xX 2 d y
dx ds2 ds
X
dx2
2 ^W dy(s)+y.
X
dx2 ds2 ds ' ^yv>)
y'{0), (7.81)
(7-83)
2. Partial derivatives
We now suppose that W(JC, t) is an arbitrary function of x and t> where
a ^ x ^ b and t ^ 0 , a and b being constants. Since f ranges from 0 to
oo, we may transform with respect to this variable as follows:
du(xf t)
s u{x, s) — su(x, 0) — (7.87)
dt t=0
Again we have assumed that u and its first derivative (with respect to
t) are of exponential order and are such that at infinity the integrated
terms are zero.
Besides the derivatives du/dt and d2u/dt2, we have the other
derivatives du/dx, d2u/dx2, and d2u/dx dt. For the first of these, we
have
= — u(x,s), (7.89)
7.8 Inversion
Before applying the Laplace transform to the solution of differential
equations, it is necessary to discuss in detail how to find/(jc) from the
transform/(s). Such a process was indicated in (7.7) but we now need
specific techniques. One of the simplest and most obvious is to read
the inversion from a list of transforms so that from Table 7.1, for
example,
fe}cos(a*)- (7 92)
-
Other methods use partial fractions, the Convolution Theorem, and
the general inversion formula based on contour integration.
174 Laplace and Fourier transforms
Example 4 Consider
2s2 + 3s-4
Jy (7-94)
' (s- 2)(s2 + 2s + 2)
Then by partial fractions we may write this as
1 s+3
(7 95)
-
Hence
Hence if some function, say F(s), is given which can be written as the
product of two functions f(s) and g(s), the inversions of which are
known, we may then use (7.99) to obtain the inversion of F(s).
Example 5 To evaluate
i (7 ioo)
W\ -
7.8 Inversion 175
we write
/ ( ) ) (
^
Then from the standard results and the Shift Theorem we find
f(x)=<e-l{f(s)} = 4x, (7.102)
1 2
g(x) = %- {g(s)}=xe- *. (7.103)
2u
Hence from (7.99), since f(x — u) = 4(x — u) and g(u) = ue~ , we
have
^i7(hf}=4[(x-u)ue~2udu' (7 io4)
-
which, by integration by parts, gives
vhfH^"2^-1- ^ (7 105)
-
Example 6 Given that
f(x) = 5£-l{f{s)}, (7.106)
we can show that
{ s } j t . (7.107)
Writing
g(s) = l/s, (7.108)
we have that g(x) = 1. Hence, using the Convolution Theorem,
= f f(x-u)du, (7.110)
Jo
since g(w) = 1. Writing x — u = ty we have finally
Figure 7.4
-•Res
Singularity •
of/(z)
Res
7.8 Inversion 111
y-axis. On this part s = Reld, where Jt/2 < 0 < 3JT/2. Hence on S
\esxf(s) ds\ = \eRx(cosd+isine)f(Reie)Reidid6\
= eRx cos e \f(Reie)\ R dd. (7.113)
By assumption / behaves as a power of R on this semicircle and
cos0<O. Hence as R-+™, (7.113) tends to zero since it includes a
decreasing exponential (with R) multiplied by a power of R.
Now consider the upper arc of the circle Ax in the first quadrant.
The contribution from this arc to the total integral is bounded as
follows:
where cos 60= y/R (see Figure 7.5). Now since 60^ d^Jt/2, it
follows that cos 6 ^ cos 60 = y/R. Hence (7.114) is less than
a f71'2 aeyx
Rx /R
- ^ e v d0 = -^(jt/2- e0). (7.115)
K JeQ R
In (7.115), the quantity Jt/2- 00 = Jt/2-cos~\y/R). For small t, the
function y = cos"11 has the expansion
Letting /?—><», the first three integrals in (7.119) tend to zero and the
finite line U becomes the Bromwich contour L. Further L is chosen to
lie to the right of all the singularities of f(s). Hence
which has simple poles at s = ±2i. Using the formula for the residue at
a simple pole (5.183), we have
Hence
f(x) = (sum of residues) = \(z2bc 4- e"2Lr) = cos(2x), (7.124)
which is the inversion of (7.121) (as can be seen from Table 7.1). ^4
(7.130)
(7.134)
•f
which tends to zero as p —»0.
/(*)
y w =- f - ^ r - — du=-^r f e - |<2 dM=-7^—, (7.139)
JTJ 0 (W/VJC)X WJCJO VW)
as found in (7.132). ^ f
(7 140)
-
(see (7.57)).
(see (2.45)).
7.10 Solution of ordinary differential equations 181
Example 10 Consider
d2y
(7.142)
(7147)
Inverting (7.147) using the Shift Theorem (7.11) and result (7.66), we
have
y(x) = 1 - e"2* + H(x - 1)[-1 + 2e" ( *- 1) - e " ^ - ^ ] . ^ (7.148)
The attraction of this method is that it enables us to solve the
differential equation by solving only an algebraic equation (7.144) and
iising known results for Laplace inversions. It is sometimes convenient
|p use the differential equation defining a special function to evaluate
its Laplace transform. Consider the following example:
- f = 0, (7.150)
which simplifies to
(7.151)
as
Separating the variables and integrating gives
(7 152)
)' -
where A is an arbitrary constant which must be evaluated. Hence the
Laplace transform of / 0 (*) is
Putting 5 = 0 gives
is a type of Volterra integral equation for y{x)> where f(x) and g(x)
are given functions. The Laplace transform approach is particularly
suited to this form of integral equation since the integral is a
convolution of y and g. Hence transforming and using the Convolution
Theorem (7.21), we have
y(s)=f(s)+y(s)g(s), (7.156)
7.12 The Fourier transform 183
whence
m=
TT?+*hm (7 159)
-
or
Inverting gives
>;(jc) = 3sinx-V2sin(V2jc). ^ (7.161)
{s)^ds. (7.163)
We note that in some books the definitions differ from those above by
having different multiplicative factors in front of the integrals. Any
two such factors the product of which is 1/2JT can be used.
1. Even functions
If f(x) is an even function (f(—x)=f(x) in the range —oo<x<oo)
then we may define the Fourier cosine transform ^c{f(x)} as
l
{Us)} =/(*) = \ [fc(s) cos(sx) ds. (7.165)
2. Odd functions
If f(x) is an odd function (/(—*)= —/(*) in the range -oo<x<oo)
then we may define the Fourier sine transform ^ s {/(*)} a s
These three transforms are linear in the sense of (7.3) and (7.5).
Functions defined for x ^ 0 which are neither even nor odd can be
extended over the whole range by defining them to be either even or
odd, as in the case of Fourier series.
Example 13 Consider the Fourier sine transform of e~* for x > 0 (so
that the function is — e* for x < 0, thus making it odd). Then
Choosing some constant value of JC, say x = k (and calling the dummy s
of integration JC) we have for k > 0
0
x sin(kx)
r°° JC sin(fct JT .
2v -dx=-e~k. (7.170)
Jo * + l
For k < 0 we have
3. Convolution Theorem
The corresponding result to (7.21) is
where /(s) and g(s) are the Fourier transforms of f(x) and g(x) as in
(7.162). As with the Laplace transform we denote the convolution
integral by
| f(x-u)g(u)du=f*g (7.173)
4. Transforms of derivatives
^ { ^ } = (w)2y(*) (7-178)
and, in general, provided y (n ~ 1) (jc)^»O as *—> ±00,
(w)"y(*)- (7-179)
(b) Fourier cosine transform Using the definition (7.164) and integrat-
ing by parts, we have
+ sys(s), (7.182)
186 Laplace and Fourier transforms
and
5 . Differential equations
The application of Fourier transforms to the solution of partial
differential equations is dealt with in the next chapter.
6. Integral equations
The integral equation
g(s)> (7-189)
whence
7.12 The Fourier transform 187
When f{s) and g(s) are known, y(s) may be inverted to give y(x).
(x-u)f(u)du=^-^. (7.191)
(7.WM)
(?TiHfe~w' (7 195)
-
and from (7.192)
-]s]'2. (7.196)
Inverting (7.196) we have, using (7.163),
r f V » e e f d s (7.197)
j(U+i)
r- ( f <?<**-*> ds+l e- ds) (7.199)
y/ji \J0 j0 I
^ ) (7- 200)
«-
188 Laplace and Fourier transforms
Problems 7
1. Evaluate (i) ${cosh{ax)cos{ax)}, (ii) #{**}.
2. Use the Shift Theorem to obtain Z£{ex sin 2 x}.
3. Find the Laplace inversions of the following functions:
s x
n D
4. Use the Laplace transform to evaluate
(iii) f t-*J0(x)dx.
Jo
5. Show that
for a > 0.
9. Solve by the Laplace transform method, where D = d/dx:
(i) (D2 + 3D + 2)y = 4e~3* with y(0) = 1, / ( 0 ) = 2,
(ii) (D 2 + % = i/(x - 2) with y(0) = 1, y'(0) = 2,
(iii) (D + l)y = 6(x - a) with y(0) = l,a>0.
10. Use the Laplace transform to solve the equations
(D - 2)x + Dy = 2y
(2D - l)x - {AD - 3)y = 7
for x(t) and y(t)y where D = d/A, with JC(O) = 0, y(0) = 1.
11. Express the solution of
Problems 7 189
13. Find, using the Convolution Theorem, the solution of the integral
equation
COS(SJC) dx — y/(ft/
Jo
18. Using the Fourier Convolution Theorem and the first result of
Problem 15, solve the integral equation
£d! • * . -
8
Partial differential equations
8.1 Introduction
Partial differential equations occur frequently in the formulation of
basic laws of nature and in the mathematical study of a wide variety of
problems. In general, since we live in a universe of three space
dimensions and one time dimension, the dependent variable, u, say, is
a function of at most four independent variables. Relative to a
cartesian coordinate system, therefore, we have
u = u{x, y, z, t). (8.1)
The relationship between u and its partial derivatives ux = du/dx,
uxy = d2u/dx dy, . . . ,
f(x, y , z, t\ u, uxy uy, uxx) uxy,...) = 0, (8.2)
where / i s some function, is called a partial differential equation. If/is
linear in each of the variables u, ux, uy, uxx, . . . and if the coefficients
of each of these variables are functions only of the independent
variables x, y, z, t, then the equation (8.2) is said to be linear.
Equations which are not of this type are said to be non-linear. In
general, the order of the partial differential equation is defined by the
order of the highest-order partial derivative in the equation.
equation
d2u l d2u
- 2,,2, (8.3)
dx 2
where c is a constant. It is easily verified that the two solutions of (8.3)
are
ux=f{x + ct), u2 = g(x-ct), (8.4)
where / and g are arbitrary functions, at least twice differentiate, oi
the arguments (x + ct) and (x — ct), respectively. Accordingly, by the
Principle of Superposition,
u=f(x + ct) + g(x - ct) (8.5)
is also a solution of (8.3). We shall show in Section 8.6 that (8.5) is, in
fact, the general solution of (8.3).
The Principle of Superposition, however, does not apply to non-
linear equations. There is no general method of obtaining analytical
solutions of non-linear partial differential equations and numerical
procedures are usually required for their solution. Because of the
difficulty in solving such equations, approximations are sometimes
made which linearise the equations. Provided the physical implications
of such approximations are understood, the analytical solutions of the
linearised equation (which are much easier to obtain) often give a
valuable insight into the solution of the original problem.
where c is a constant.
192 Partial differential equations
and A is a constant.
Higher-order equations also arise. For example, the biharmonic wave
equation
V 4 « = V 2 (V 2 M) = - 1 | ^ , (8.15)
1. First-order equations
The method of characteristics can be applied usefully to both linear
and non-linear equations. To illustrate the technique, we take the
simplest form of (8.16) by letting n = 0, so that
du du
P(x,y) — + Q(x,y) — = R(x,y). (8.22)
ox oy
Consider now the family of curves defined by the equation
^ ^ (8.23)
(8.25)
Q2
(8.26)
Hence
dx _dy ds
=
= dt. (8.27)
~P ~Q'
Now from (8.23), the curves are defined by
dy _Q
dx~~P' (8.28)
Figure 8.1
196 Partial differential equations
=
^ V(P 2 ^-6 2 )^' (8 32)
*
Using (8.27) this becomes
du=Rdt. (8.33)
Finally then (8.23) may be written as
*-!-!<-*>•
from which w can be found by integration. This integration is clearly
along any curve C defined by (8.23). These curves form a special
family relating to the given differential equation and are called the
characteristic curves (or, simply, the characteristics). The method of
characteristics requires that the region over which (8.22) is to be
solved should be filled with characteristic curves. The following
examples illustrate the method in some simple cases.
Example 1 To solve
given u = 2 on y = x2.
The basic equation (8.34) gives (since P — xy Q =y, R =2ry)
- =^ =^ (8.36)
x y 2xy
and the characteristics are defined by
dy/dx=y/x (8.37)
8.6 Method of characteristics 197
u=xy+f(y/x). (8.42)
Since on the curve y =x2, u = 2 (given), we have
2 = *3+/(*). (8.43)
Hence
f(x) = 2-x3 (8.44)
giving
f(y/x) = 2-(y/x)3. (8.45)
Substituting (8.45) into (8.42), we see that the required solution of
(8.35) is
v 2
^x ' 2' By
^ (8.47)
Equation (8.34) becomes (since R = 0)
dx _ dy du
2 2 (8.48)
2xy (jc + y ) 0
198 Partial differential equations
(8.55)
Then
(8.57)
X /
Example 3 To solve
du du
with u = 0 on y = x2.
Following (8.34) with R =xe~u, we have
- =^ ~ - (8-62)
x y xe
As in Example 1, the characteristics are
y/x = K. (8.63)
We also have from (8.62)
dx = eudu, (8.64)
which on integration gives
eu=x+f(K). (8.65)
Hence using (8.63),
eu=x+f(y/x). (8.66)
Now u = 0 when y = x2, giving
2. Second-order equations
One of the simplest applications of characteristics to second-order
equations occurs when the equation may be factorised into two
first-order equations. We illustrate this by the following example.
Example 4 To solve
s?u d2u 2d
2
u du_
dx2 dx dy dy2 dy
given that du/dx = y2 on x = 0 and u = ey on x = 0.
Consider the operator
L = ^- + x^-. (8.71)
dx dy
Then
3\ 32u d2u 2d
2
u du
(d+ x
d
) ( d , .,. U (8.72)
\dx dy/\dx dy) 3x2 + dx dy X
dy2' dy'
and hence (8.70) may be written as
L2u = 0 (8.73)
Now writing
Lu = w (8.74)
(8.73) becomes
L 2u = L(Lu) = Lw = 0, (8.75)
or, using (8.71),
dw dw
= 0. (8.76)
dx dy
The characteristics of (8.76) are found from
whence
y = \x2 + K. (8.78)
Also dw = 0 from (8.77), which implies
*=f(K), (8-79)
where / is an arbitrary function. The general solution of (8.76) is
therefore
w=f(y-hx2). (8.80)
8.6 Method of characteristics 201
— + x — = (y - \x2fy (8.84)
whence
^ = __^£_. (g.85)
(8.93)
dx dx dy dy
This equation, known as Euler's equation, is of one of the types
described by (8.19)-(8.21). Here we take ay h and b to be constants.
Defining new independent variables
du du dE du drj du du
L
— = —r — + = p—z. + r—, (8.95)
du du dt; du drj du du
Hence
d2u d (du\ i d d\i du „.
d2u +2pr
^ stu+r 2d
2
u (8, 98)x
w wr, ^> -
d2u d (du\
—? = — I — I = I q hs—
dy dy \dy) \ dg dr
,32u
(8.100)
<jg e g urj oi\
and
d2u d (du\ ( d d\( du c
dx dy
+rs
d2u 32u (8102)
^, w
Using (8.98), (8.100) and (8.102), (8.93) becomes
^ = 0. (8.103)
dr)
8.6 Method of characteristics 203
We may now choose the constants p> q> r and s to simplify the
equation. Let p = r = 1 and choose q and s to be the two roots kx and
A2, say, of the quadratic equation
6A2 = 0. (8.104)
Then (8.103) becomes
^-*2)a0Ta (8 1O8)
-
Provided now b =£ 0 and the equation is not parabolic in nature
(ab-h2*0), (8.108) becomes
3r/=0, (8.109)
where, making use of our choices of/?, qy r and s,
% = x + X1y, r/=x + A2>^. (8.110)
Now (8.109) may be integrated immediately, first with respect to, say,
§ giving du/drj =m(rj), where m(rj) is an arbitrary function, and then
with respect to rj giving
^A-\^ =0 (8.114)
ox c at
is a special case of (8.17) with 0 = 1, h = 0 and b = -l/c2. Hence
ab — h2= — l / c 2 < 0 and the equation is therefore hyperbolic (as
mentioned in Section 8.4). Accordingly (8.104) becomes
l - A 2 / c 2 = 0, (8.115)
giving
A1 = c , A2=-c. (8.116)
The characteristics are therefore
§ = x + ct = constant, (8.117)
rj=x — ct = constant, (8.118)
and the canonic form is
whence
A = l or A = 2. (8.123)
The characteristic curves are
§ = x + 2y = constant, (8.124)
x) = x + y = constant. (8.125)
The equation reduces, by (8.108), to the canonic form
\{~i)wk=y=^~n' (8126)
or
d2u/dt;dr) = ri-t=. (8.127)
This equation may be readily integrated, first with respect to 77 giving
3u/d% = \Y]2 - %rj +/(§), (8.128)
where/is an arbitrary function, and secondly with respect to f giving
(8.129)
where F and G are arbitrary functions. Expressed in terms of x and y;
using (8.124) and (8.125), we have
(8.130)
Example 7 To solve
from which we see that the left-hand side is a function of x only, whilst
the right-hand side is a function of y only. Hence the equation has
been separated with terms dependent only on x on one side and
only on y on the other. For (8.135) to be satisfied, each side must be
equal to a constant (or, say) so that
i yt y
u-e 2J c e
(8.141;
all C
all a
8.7 Separation of variables 207
| ^ + 0 + ( * 2 + /) 2 M = O (8.147)
is not separable in x and y but is separable in plane polar coordinates
(r, 0).
Writing u=X(x)Y(y), (8.147) becomes
X" Y"
— + — + x4 + 2x2y2 + / = 0 (8.148)
JC Y
or
4
( ) V ( / ) 0 . (8.149)
The terms in square brackets are functions of r only and the remaining
term is a function of 6 only. Hence (8.152) is separable into two
ordinary differential equations for R(r) and O(0),
r d ( dR\ 6 1 d 2e
V2r = - ^ y , (8.154)
and hence that, if T is finite (bounded) at the centre of the sphere and
ST/dr = 0 on the surface r = a for t > 0, the solution is of the form
\ ^ . (8.158)
8.7 Separation of variables 209
Hence
dr2 r dr K dt
Now putting u = rT, we have
3T \du 1U
2 (8.160)
dr r dr r2 y
'
and
d T 11
(8.161)
dr2 r dr2 r2 dr r
Substituting into (8.159), we find, as required by (8.155),
d2u l du
(8 162)
-&-K*' -
and writing u(r, t) in separable form as
u(r,t) = R(r)S(t), (8.163)
we have
R" 1 5'
S=Ae±Ka)2t. (8.165)
In order that the temperature falls off exponentially with time rather
than increasing without limit (a non-physical solution), we must exclude
the solution with the positive sign in (8.165). Hence the solutions of
(8.164) for w ^ O a r e
Now for T to be finite at the centre of the sphere, we must remove any
term behaving as 1/r as r—»0 since this will diverge. Hence F = 0, and
Bn = 0 (for all n) since cos(ft>nr)/r—•<» as r—>0. Consequently
^ ) ] =0, (8.174)
or
[rcon cos(conr) - sin(conr)]r=a = 0, (8.175)
giving
tan(cona) = cona, (8.176)
as required. An infinity of con values satisfy this transcendental
equation and these values must be used in (8.173) to form the solution
T(r, t).
By imposing some initial condition in the form T(r, 0 ) = / ( r ) , we
may determine the constants E and Cn as follows. From (8.173),
inserting t = 0, we have
Hence
+ 2 Cnsm((onr)sin(wmr)dr. (8.179)
n=l h
Now integrating by parts and using (8.176), we find
r
r sm(comr) dr = 0. (8.180)
Jo
Further
m
sin((Dnr) sin(a)mr) dr = - ( 2 2)dmn, (8.181)
Jo l \ l + o) m a /
for m, « = 1, 2, 3, . . . , using sin(cona) = cona/^(l 4- ft>^«2) and
cos(a^a) = 1/V(1 + G ^ 0 2 ) from (8.176). Hence from (8.179) we find
= ?y, (8.184)
£ = 4 f r2f(r)dr. (8.185)
fl Jo
Example 10 To solve
0, (8.186)
where or is a constant and E is the energy parameter, subject to the
conditions that u is a single-valued function and is bounded (finite)
everywhere within a sphere of radius a, given u = 0 on r = a.
Expressing V2 in spherical polar coordinates (r, 0, (p), using (1.84),
(8.186) becomes
1 9 / ~ du\ 1 9 / du\ 1 d2u
-~ — \r— +-5 s i n 0 — + -; ^ ~
r 2 d r \ <9r/ r 2 sin0<90\ 9 0 / r2 sin2 d d(f>2
(8.187)
Writing
u(r, 0, <p) = / ? ( r ) B ( 0 ) * ( 0 ) , (8.188)
we have
1 d / 2dR\ 1 I d / d0\ 1 1
e
^ 7 V ) )
(8.189)
Letting
— — 2 = -m2> (8.190)
>dRX • "" [ l d
' ' — (8.192)
= A (8193)
- ^ '
where A is the separation constant. This equation is discussed in
Chapter 2 (see (2.189)) where it is noted that bounded solutions exist
only if
A = - / ( / + !), / = 0,1,2,..., (8.194)
8.7 Separation of variables 213
and that these solutions are given in terms of the associated Legendre
functions
0 = £P}m|(cos 0), (8.195)
where |ra|=^/, and B is an arbitrary constant. Finally we solve the
equation for R(r) which, from (8.192) and (8.193), is
(8.200)
thus normalising u*u = \u\2 to unity when integrated over all space.
Now if u = 0 on r = ay we have
JlH(\/(aE)a) = 0. (8.203)
For each given / value, there will be an infinity of roots rly r2y r3y . . .
such that // + i(r / ) = 0 (the values of the r, being the zeros of the Bessel
214 Partial differential equations
function). Hence
= rlr\yrl...y (8.204)
and the energy E therefore takes on an infinite set of discrete values
(this property being known as the quantisation of energy). ^
1. Laplace transforms
dx2 K dt'
for x > 0, t > 0 given that u = u0 (a constant) on x = 0 for f > 0, and
u = 0 for x > 0, f = 0.
Taking the Laplace transform with respect to t gives
U
\ 2'S =-psu(x,s), (8.206)
dx /C
using (7.86) with u(x, 0) = 0. Also
and hence
(8 2n)
iW^] -
where erfc is defined in (2.49). ^k
(8.216)
(8.217)
<8224)
(8 225)
Uv)^ -
(8 226)
-
where s' = J^5. Since the Laplace inversion of 1/V(1 + s2) is / 0 (^) (see
Chapter 7, Example 11), (8.226) gives
S h = 4 y ° W K ) - -* <8227>
-
2. Fourier transforms
Finally we illustrate the use of the Fourier transform in solving
particular types of boundary value problems. In Chapter 7, it was
found that the Fourier sine transform of a second derivative term gave
the result
, —2 sin(sjt) dx = -s2us(s, t) + sw(0, r), (8.228)
Jo ox
8.8 Integral transform techniques 217
2 C O S ^ A J UX — —S UQ{S, I) — 1 — I , yp.AZry)
o a* \dx/x=o
ws and wc being respectively the Fourier sine and cosine transforms of
u{x> t) with respect to x. The choice of a suitable transform is
therefore dependent on the nature of the given boundary conditions.
If u(0y t) is specified, but not (du/dx)x=0, then the sine transform is
appropriate, whereas if (du/dx)x=0 is specified but not u(0, t) then the
cosine transform is appropriate. We now give an example.
Example 13 To show, using the basic result
w
e""2 cos(Aw) du = — e"A2/4, (8.230)
that the Fourier sine transform of erfc(fljt) with respect to x is
(1 - e~s2/4a2)/s, and hence to solve
d2u 1 du
f-^ye-*2'4*2) (8-236)
= - (1 - e- j2/4a2 ), (8.237)
s
using (8.230).
218 Partial differential equations
Problems 8
1. Use characteristics to solve
du du
given u — cos x on x1 + y2 — 1,
(in) x — -y — = u2
ax ay
given u = 1/lnjc on JC2^ = 1,
J
dx2 XJ
' dx dy J
dy2 dy J
dx
where A, B and A are constants, and /0(*) and Y0(x) are the
zero-order Bessel functions satisfying
xy" + y' +xy = 0.
Given u(a, t) = 0 and u(r,0) = Jo(a>r/a), where a> is defined by
/0(cw) = 0, find the value of t such that
u(o, o = Mo, o).
8. An infinite cylinder with circular cross-section has radius a = 2
units. The temperature distribution within the cylinder is T(r, t).
The surface of the cylinder is kept at zero temperature for all
time, and at / = 0,
r(r,0)=/ 0 (1.2r),
where Jo is the Bessel function of zero order. Writing the heat
conduction equation
K dt
in cylindrical coordinates, show that the temperature distribution
is approximately
conditions
u(0,t) = Uo, t>0y
u(l,t) = Uu t>0,
where Uo, Ux and / are given constants, and
u(x, t) = x + - f ) ^Z^- s i 2 2
^
c2 dt2 ~ dr2* r dr
32u _ du
a—5 + b e cx — — , (JC 52 0, f^O),
5JC 5^
222 Partial differential equations
find W(JC, 0-
12. Show that the solution of the heat conduction equation
d2u l du ,
u0
"(X' S)
ss cosh[y/(s/K)l] '
By finding the residues of u(x, s)est at the poles 5 = 0 and
s = -(2n- 1)2(JZ2K/412), where n = 1, 2, 3, . . . , show that
9.1 Introduction
One of the elementary uses of differential calculus is in finding the
stationary values of functions of one or more variables. The calculus of
variations considers the more complicated problem of finding the
stationary values of integrals. A simple physical example which
illustrates the general type of mathematical problem involved is as
follows: suppose P(xly y{) and Q(x2, y2) are two given fixed points in a
cartesian coordinate system (see Figure 9.1). We wish to determine
the equation of the curve joining these two points such that when the
curve is rotated through 2n about the jc-axis to form a surface, the
surface area so generated has a minimum value. Suppose v =y(x) is
some curve joining P and Q. Then, if ds is the element of arc length,
the surface area S so formed is given by
I[y{x)]=Pf(x,y,y')dx (9.2)
Figure 9.1
-•*
226 Calculus of variations
whence
61(6) =1(6)-I
(9.8)
Assuming now that the first term of the integrand may be expanded in
powers of 6, we have (using Taylor's expansion of a function of two
variables)
6
61(6) = 6ll+ — I2 + terms in 63 and higher orders, (9.9)
where
(9.10)
Figure 9.2
9.2 Euler's equation 227
and
However r/(xj) = r/(jt2) = 0 (see (9.5)) and the first term on the right of
(9.14) is therefore zero. Inserting (9.14) into (9.13) gives
- *
Now rj(x) is an arbitrary function, apart from the conditions (9.5).
Hence we may make use of a fundamental result which states that if
rj(x) is continuous and has continuous derivatives up to at least second
order in the range xx ^x ^x2 with r){xx) = rj(x2) = 0 and rj(x) =£0 for
xx<x<x2y and if g(x) is itself a continuous function and
X2
fl(x)g(x)dx=0 (9.16)
gfimO. (9.17)
This equation is known as Euler's equation and its solution leads to
the form of y(x) which produces the stationary value of /. The curves
y =y(x) so obtained are called extremals of (9.2).
228 Calculus of variations
or
d2y
(9.22)
= (l-x2)dx = l ^ (9.28)
9.3 Alternative forms of Euler's equation 229
dx\dy /
giving
df/dyf = constant. (9.31)
We can now apply the Euler equation to this integral and, since the
integrand is explicitly independent of y, we can use the form (9.31).
The Euler equation therefore has the form
or
y'
vYi—^ = c ( 9 * 35 )
Hence
y'=Ay (9.36)
where A is a constant, and the equation of the curve of shortest length
joining P and Q is therefore
y=Ax + By (9.37)
the constants A and B chosen so that (9.37) passes through the end
points P and Q. ^k
230 Calculus of variations
Case 2 If y' is explicitly absent from /, then /(JC, y, y') =/(*, y).
Therefore df/dy' = 0 and the Euler equation (9.17) becomes
df/dy=Q. (9.38)
Case 3 If JC is explicitly absent from / then /(JC, y, y') =f(y, y') and
hence df/dx = 0. To see the effect this has on the Euler equation
(9.17), we note that the total differential of /(JC, y, y') with respect to JC
is given by
^31 3ldy K ±
dx dx dydx dy'dxKy} K }
y y 0 (9 45)
dx dy' dxKdy'
or
£</->•§)-*
Integrating (9.46) we find
f~y'§ = C, (9.47)
where C is a constant.
9.3 Alternative forms of Euler's equation 231
The following two examples illustrate the use of this form of the
Euler equation.
8t=8sN(2gy), (9.49)
ds V(l + v' 2 K
<ijc, (9.50)
(9.51)
V(2g)
We now wish to find y(x) such that T is a minimum. The integrand of
(9.51) is explicitly independent of x and hence we may use the form
Figure 9.3
o
ds
232 Calculus of variations
or
/ / l + v' 2 \ y' 2
V --r~77 2T = C - (9-53)
V\ y ) vyy(l+y )
This simplifies to
y ( l + y ' 2 ) = D, (9.54)
where D is a constant.
Making the substitution y' = cot Q> we have
and
whence
(9.57)
Equations (9.55) and (9.57) give the parametric form of the equation
of the curve, and define a cycloid. ^
w^)=a (9 58)
-
After simplification, this becomes
y2 = C 2 ( l + / 2 ) , (9.59)
which integrates to give
(^^) (9.60)
9.5 Several dependent variables 233
where the constants A and C must be chosen so that the curve passes
through the end points P and Q. This curve is known as a catenary,
and the surface of minimum area generated by this curve is known as a
catenoid. ^
<9-67>
All of these equations must be satisfied simultaneously for (9.64) to
have a stationary value.
Figure 9.4
-•*
O
9.6 Lagrange's equations of dynamics 235
ftKft)* (9.69)
where L = L{t\ qlf q2, . . . , qn\q\, qi, - - , qn) and qt = dqjdt.
Equations (9.69) are known as Lagrange's equations and provide a
powerful technique for deriving the equations of motion of a dynami-
cal system. This is illustrated by the following two examples.
Hence
L=T-V = + mgl cos 0. (9.72)
Since there is only one generalised coordinate, the single Lagrange
equation (from (9.69)) is
9L dL(dL\_
(9.73)
~dO~~dt\d8)~ '
which, using (9.72), gives
Example 8 Three masses mly m2 and m3 lie on a straight line and are
connected by two springs of stiffness K for which the potential energy
is (by Hooke's Law) V = \K (extension)2. No other forces act. If xx, x2
and x3 are the displacements when the system is disturbed from its
equilibrium position then
L = T - V = \mxx\ + \m2x\ + hm3x\ - \K{x2 - xxf - \K(x3 - x2f.
(9.76)
The Lagrange equations (9.69) give (with qt = xt)
dL d_(dL\_
(9.77)
~dxi~Jt\dlJ~
Figure 9.5
10
9.7 Integrals involving derivatives higher than the first 237
for i = 1, 2, or 3, whence
m1x1 - K(x2 - JCO = 0, (9.78)
m2x2 + K(x2 - JCO - K(jt3 - *2) = 0, (9.79)
m3x3 + AT(JC3 - JC2) = 0. (9.80)
This coupled set of three second-order differential equations defines
the motion of the system, and elementary methods may be used to
obtain xlf x2 and x3 as functions of time. ^
( ) + ( ) + ( i r ( ) 0 (984)
~8y~ic(~6y')+is{2y")=0> (9 88)
-
or
d\ d2y
J + iJ-4y.O. (9.89)
238 Calculus of variations
I[y{x)]=\X2f(x,y,y')dx (9.92)
J[y(x)]=Pg(x,y,y')dx = C, (9.93)
where g(x, y> y') is a given function and C is a given constant. As with
the conventional Lagrange multiplier approach for finding the station-
ary points of a function subject to a constraint, we now form the linear
combination
/ + Ag, (9.94)
where A is a constant (Lagrange) multiplier, and determine the
extremals of the integral
\X2 x. (9.95)
9.8 Problems with constraints 239
with y(0) = 0 and y(l) = 0, and where y(x) is subject to the constraint
f y(x)dx = h (9.99)
^ j , (9.104)
where A is a constant.
240 Calculus of variations
) = 4(^-2) = ). (9.108)
~16 4^
Now by (9..99) we require
rf{x,y,y')dx, (9.111)
where y{xx) = a and y(x2) = /?, oc and /? being fixed. Now consider a
sequence of functions
yi(x),y2{x)y...yyn{x),..., (9.112)
each satisfying the end conditions yn{xx) = or, yn{x2) = /?, for all n, such
9.9 Direct methods and eigenvalue problems 241
that
\im I[yn(x)] = I0 (9.113)
«—•<»
and
\imyn(x)=y0(x), (9.114)
where 70 is the minimum value of / and yo(x) is the solution of the
Euler equation producing the value 70. The sequence (9.112) is called a
minimum sequence. The simplest case is if a = )3 = 0, and then the end
conditions can be satisfied if each function yn is written as
yn(x) = Cx4>i{x) + C 2 0 2 (x) + . . . + Cn<t>n(x), (9.115)
where 0I(JC), . . . , (f>n{x) are n independent functions each satisfying
0/(^i) = 0/(*2) = 0, and Clf . . . , Cn are arbitrary constants. Inserting
the nth member of the sequence (9.115) into (9.111) we find
I[yn{x)]=\Xlf(x,yn,y'n)dx (9.116)
Example 11 Consider
withy(0) = 0,
Suppose we choose (p1(x)=x(l — x) since this satisfies the end
conditions. Then the first member of a minimum sequence is
y1(x) = C1x(l-x). (9.120)
Hence
(9.121)
242 Calculus of variations
and
(9.124)
gives
(9.125)
and with this value, (9.123) gives
and
^^ =^ + ^ 2 + ^ = 0. (9.131)
*-£(2/) = 0 (9.135)
9.9 Direct methods and eigenvalue problems 243
or
y" = \x. (9.136)
Integrating we have
y' = \x2 + A (9.137)
and
y = ±x3+Ax + B. (9.138)
Applying the end conditions gives B = 0, A = — -^ and hence we find
y = Ux3~x). (9.139)
This is precisely the function found in (9.133), and hence the value of /
in (9.134) obtained by the Rayleigh-Ritz method is exact in this case.
We emphasise that this depended on our choice of (p1 and 0 2 - F ° r this
simple example, the Euler equation gives the extremal curve very
easily, but nevertheless this example illustrates the use of the
Rayleigh-Ritz method when the value of the extremal curve is zero at
both ends. ^
If the end values of the curve are not both zero, then a linear
combination of the form (9.115) can still be written but we must
ensure that each yn(x) still satisfies the end conditions. This usually
results in relationships between the arbitrary constants, thereby
reducing the number of independent constants to less than n. The
stationary value of I[yn] is then found by minimising with respect to
the independent constants remaining. We give an example to illustrate
this.
Example 12 Consider
Hence
y2(x) = C,x + (1 - Cx)x2, (9.144)
which satisfies the end conditions for all Cx. Using (9.144), we find
I[y(x)]=l\y'2-ky2)dx, (9.154)
which has (9.151) as its Euler equation. We choose some trial function
passing through the end points, say
x). (9.155)
Then
IbiM] = C2S - ^A) = F ( d ) . (9.156)
Hence
dF/dd = 0 = 2 d G " ^A), (9.157)
which gives A = 10 as an estimate for the lowest eigenvalue. This is to
be compared with the exact value A = ;r 2 ~9.87 obtained by putting
n = \ in (9.153).
In order to improve the approximation, we choose, say
y2(x) = d*(l - x) + C2x\\ - x) (9.158)
which also passes through the end points. We now find
Kyiix)] = \c\ + \cxc2 + ic22 - A[ic? + i d d + ifecl] = F(d, d).
(9.159)
Hence
5 F / 3 d = (f " * A ) d + G ~ *A)C2 = 0 (9.160)
and
9F/dC2 = G - i A ) d + ( B - ifeA)C2 = 0. (9.161)
Equations (9.160) and (9.161) are a pair of homogeneous linear
equations for Cx and C2. For a non-trivial solution we require that the
determinant of the coefficients is zero. This gives
A2 - 52A + 420 = (A - 42)(A - 10) = 0 (9.162)
and hence
A = 10, A = 42. (9.163)
The first of these is the value we found above corresponding to n = 1
(for which the exact value is n2^ 9.87), while the second corresponds
to n = 2 (and an exact value of A = 4JT2 ~ 39.48). ^
246 Calculus of variations
Problems 9
1. Find the extremal curves and the stationary values of the following
integrals:
(i)
rQ
(ii) [2ysmx + (dyldx)2]dx,
JP
Jo
J 1
o
y dx = JT/2 - 1.
Chapter 1
1. auxi + a22xl + a33xj + 2{aX3xxx3 + ai2x1x2 + a23x2x3).
2. (i) j , (ii) none, (iii) i and /', (iv) i on each side.
3. 3, 3, all three components zero, all three components zero.
4. 26,1, 6.
5. S,, = -
Chapter 2
1. (i) T(5) = 24, (ii) V(w/a).
4. 2w/3V3, using (2.25) and (2.27).
6. (i) |B(|, | ) , (ii) w/V2, using (2.25) and (2.27).
9. AJv(ex) + BYv(ex).
12.
13. (i) 0, (ii) -I
Chapter 3
2. 4/(x-3)2.
3. _y = x In Ax.
Cx + 1
In A:
4. y = I n
x+C
Chapter 4
1.
3. Jt+x 3 + | r \
4. yx = 2 + x2y zx = 3x2,
y2 = 2 + x2 + x3, z2 = 3x2 4- \x
>>3 = 2 + jt 2 4 - x 3 4-|>jt 5 4 - ^ j t 6 ,
z3 = 3JC2 4- |JC4 4- %x5 4- 4 * 7 + i
5. cos t 4- ju[|r cos t — ^ sin t — ^
6. e"*3/6[,4 COS(2JC) + 5 sin(2x)].
7. i
8-
Chapter 5
2. (i) \ In 2 4- i(-JT/4 + 2A;JT), where A: = 0, ±1, ±2, . . . ,
(ii) nil + kjt + (i/2) In 3, where k = 0, ±1, ±2,
3. i(jr/2 + A:JT), where k = 0, ±1, ±2,
4. u = 2jry + Ay - 2x 4- C, /(z) = z2 + (4 - 2i)z + C.
5. g and /i constant.
6. All parts 5 4- f i; z 2 4-1 is analytic for all z.
7. (i) (a) 0, (b) 2m,
(ii) -2jri, jri/2.
8. (i) 2;risinhl, (ii) 4^:ie/3.
9. (i) In, (ii) 2m/9, (iii) 2jri.
10. (i) l + z 2 + ^z 4 + . . . ( / ? = oo),
(ii) sin 1 4- (cos l)z + (2 cos 1 - sin l)z 2 /2 4- . . . ( / ? = 1),
(iii) In 2 4- ^z + \z2 + . . . (R = JT).
• • •] (double pole
at z = n/2).
250 Answers to problems
12. (i) -
1 2 3 4
13. (i) - 1 ,
00 1,
(iii) simple pole at z = 2, residue | ,
double pole at z = 1, residue 2,
pole of order 3 at z = 0, residue — ^ .
14. 2kjt — Una, where k = 0, ± 1 , ±2, . . . .
Chapter 6
1. (i) r, (ii) (4-2V3
3. 0) . (») JT, (iii) iTT, (iv) \n.
6.
U
JV2>
8. J = 0.
9. e™(
11. (i) i i cosech2 n + iJT COth JT,
(ii) TIC
Chapter 7
3
s
1. (i) (ii)
+ a2][(s-\-af + a2]'
i
2
is. *?y-y>.
16. (i)
Chapter 8
2. 0 = ^( ; y )
3. M = e " ^ / ^ + i>^) 4- g{x — iy)], f, g arbitrary functions.
2 2
4. w = j t V
5 . 0 = i(
7. r = fl2
in t , 2a2 " ( - l ) n . /n^r\ . /njrcA
10. w(r, 0 = - - 2 ~ Z 1 - ^ L s i n sin .
nrcr n=i n \ a / \ a )
13 M(
- ^» = 2V(b)f,coshva^^ 2 ) c o s h [ V ( 1 + s 2 ) y ] e i
14. u(x,t) = - ^
where
a )
252 Answers to problems
Chapter 9
1. (i) y=x,/ = k ( l + i^ 2 ),
(ii) y = n — x — sin xy I = §JT.
2. y = 2x.
3. y = sinh(jcsinh~12).
4. y = 1 — COSJC.
9. (i) C = 4 ( i i ) ^ = ^ ^ 2 = * .
Index