LectureNotesMATH2048 PDF
LectureNotesMATH2048 PDF
LectureNotesMATH2048 PDF
David Gammack
Kostas Skenderis
2016-17
Contents
2 Fourier Series 19
2.1 Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2 Orthogonality . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3 Fourier Coefficients . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3.1 Even and odd functions . . . . . . . . . . . . . . . . . . . 22
2.3.2 Even and odd extensions: half range series . . . . . . . . 27
2.4 Fourier’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.5 Calculus and Fourier Series . . . . . . . . . . . . . . . . . . . . . 31
2.5.1 Differentiation of Fourier Series . . . . . . . . . . . . . . 31
2.5.2 Integration of Fourier Series . . . . . . . . . . . . . . . . 33
2.6 Complex form of Fourier Series . . . . . . . . . . . . . . . . . . 34
2.7 Parseval’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.8 Summary of results . . . . . . . . . . . . . . . . . . . . . . . . . 36
3 Fourier transform 39
3.1 An example . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3 Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4 The response function . . . . . . . . . . . . . . . . . . . . . . . . 45
3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
1
CONTENTS CONTENTS
4 Laplace transforms 50
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.2 Definition and existence . . . . . . . . . . . . . . . . . . . . . . 51
4.2.1 Existence . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.3 General properties . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.3.1 Transforming derivatives . . . . . . . . . . . . . . . . . . 53
4.4 Inverse Transform . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.5 Special functions . . . . . . . . . . . . . . . . . . . . . . . . . . 54
4.5.1 Heaviside function . . . . . . . . . . . . . . . . . . . . . 54
4.5.2 The Dirac δ-function . . . . . . . . . . . . . . . . . . . . 56
4.6 Convolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.7 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.7.1 Solving ODEs . . . . . . . . . . . . . . . . . . . . . . . 58
4.7.2 Systems of ODEs . . . . . . . . . . . . . . . . . . . . . . 59
4.7.3 PDEs . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5 Wave equation 62
5.1 Classification of PDEs . . . . . . . . . . . . . . . . . . . . . . . 62
5.2 Small amplitude vibrations . . . . . . . . . . . . . . . . . . . . . 63
5.2.1 Equations of motion . . . . . . . . . . . . . . . . . . . . 64
5.2.2 Energy of vibrating string . . . . . . . . . . . . . . . . . 65
5.3 D’Alembert’s solution . . . . . . . . . . . . . . . . . . . . . . . 66
5.3.1 Travelling waves . . . . . . . . . . . . . . . . . . . . . . 67
5.3.2 Wave reflection and transmission . . . . . . . . . . . . . . 67
5.4 Separation of variables . . . . . . . . . . . . . . . . . . . . . . . 69
5.4.1 Ansatz . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.4.2 Solving the separated form: spatial dependence . . . . . . 70
5.4.3 Solving the separated form: time dependence . . . . . . . 72
5.4.4 Normal modes . . . . . . . . . . . . . . . . . . . . . . . 72
5.4.5 Superposition . . . . . . . . . . . . . . . . . . . . . . . . 73
5.4.6 Additional notes . . . . . . . . . . . . . . . . . . . . . . 74
6 Heat equation 76
6.1 Heat conduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
6.2 Separation of Variables . . . . . . . . . . . . . . . . . . . . . . . 78
6.3 More complex material properties . . . . . . . . . . . . . . . . . 80
6.3.1 Separation of variables . . . . . . . . . . . . . . . . . . . 80
6.4 More complex boundary conditions . . . . . . . . . . . . . . . . 81
6.5 Inhomogeneous problems . . . . . . . . . . . . . . . . . . . . . . 83
6.5.1 Inhomogeneous equations . . . . . . . . . . . . . . . . . 83
6.5.2 Inhomogeneous boundary conditions . . . . . . . . . . . 84
2
CONTENTS CONTENTS
7 Laplace’s equation 87
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.1.1 Boundary conditions . . . . . . . . . . . . . . . . . . . . 88
7.2 Separable solutions . . . . . . . . . . . . . . . . . . . . . . . . . 88
7.2.1 Cartesians . . . . . . . . . . . . . . . . . . . . . . . . . . 89
7.2.2 Other domains . . . . . . . . . . . . . . . . . . . . . . . 91
8 Vector Calculus 93
8.1 Vectors and scalars . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.1.1 Curves and vector valued functions of 1-variable . . . . . 93
8.1.2 Scalar Fields . . . . . . . . . . . . . . . . . . . . . . . . 95
8.1.3 Vector fields . . . . . . . . . . . . . . . . . . . . . . . . 98
8.1.4 Directional derivatives . . . . . . . . . . . . . . . . . . . 99
8.1.5 Gradient and steepest ascent . . . . . . . . . . . . . . . . 101
8.1.6 Gradients and level surfaces . . . . . . . . . . . . . . . . 101
8.1.7 Properties of gradφ . . . . . . . . . . . . . . . . . . . . . 104
8.1.8 The divergence of a vector field . . . . . . . . . . . . . . 104
8.1.9 The Laplacian . . . . . . . . . . . . . . . . . . . . . . . . 106
8.1.10 The curl of a vector field . . . . . . . . . . . . . . . . . . 107
8.1.11 Vector Identities . . . . . . . . . . . . . . . . . . . . . . 110
8.2 Integrals of vector and scalar fields . . . . . . . . . . . . . . . . . 112
8.2.1 Line integrals . . . . . . . . . . . . . . . . . . . . . . . . 112
8.2.2 Conservative vector fields . . . . . . . . . . . . . . . . . 118
8.2.3 Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
8.2.4 The normal to a surface . . . . . . . . . . . . . . . . . . . 121
8.2.5 Surface integrals . . . . . . . . . . . . . . . . . . . . . . 122
8.2.6 Flux integrals . . . . . . . . . . . . . . . . . . . . . . . . 125
8.3 Integral Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . 130
8.3.1 Stokes’s Theorem . . . . . . . . . . . . . . . . . . . . . . 130
8.3.2 Conservative Vector Fields . . . . . . . . . . . . . . . . . 134
8.3.3 Finding potentials for conservative vector fields . . . . . . 137
8.3.4 Stokes’s theorem in the plane . . . . . . . . . . . . . . . 138
8.3.5 The proof of Green’s theorem . . . . . . . . . . . . . . . 141
8.3.6 The divergence theorem . . . . . . . . . . . . . . . . . . 142
8.3.7 Vector potential . . . . . . . . . . . . . . . . . . . . . . . 146
3
Chapter 1
Homogeneous Equations
In the case where r(x) ≡ 0 the equation becomes
d2 y dy
2
+ p(x) + q(x)y = 0. (1.3)
dx dx
and is called homogeneous. For such homogeneous equations the solution will be
of the form
y(x) = c1 y1 (x) + c2 y2 (x) (1.4)
where y1 (x) and y2 (x) are linearly independent solutions of the homogeneous
equation (1.3) and c1 and c2 are constants. The functions y1 (x) and y2 (x) are
linearly independent if
c1 y1 (x) + c2 y2 (x) ≡ 0 ⇔ c1 = 0 and c2 = 0. (1.5)
4
ORDINARY DIFFERENTIAL EQUATIONS SECOND ORDER EQUATIONS
We rewrite this using the Wronskian determinant W [y1 , y2 ](x) which is defined
as follows:
Using the Wronskian one can show that two solutions y1 and y2 are linearly inde-
pendent if, and only if, W (x) 6= 0.
In addition, we can straightforwardly check by direct substitution into (1.3)
that the Wronskian obeys
dW
+ p(x)W = 0. (1.8)
dx
This follows from the explicit calculation below:
dW
= (y1 y200 + y10 y20 ) − (y100 y2 + y10 y20 ) (1.9a)
dx
= y1 y200 − y100 y2 (1.9b)
0 0
= −p(x) (y1 y2 − y2 y1 ) (1.9c)
where we have used that y1,2 are solutions of the homogeneous equation (1.3)
= −p(x)W. (1.9d)
As the simple equation (1.8) for the Wronskian is separable we can write the
solution as
Z Z Z
dW
= − p(x)dx ⇒ log W = − p(x)dx + C (1.10)
W
5
ORDINARY DIFFERENTIAL EQUATIONS SECOND ORDER EQUATIONS
6
ORDINARY DIFFERENTIAL EQUATIONS SECOND ORDER EQUATIONS
Euler equations
The Euler type of linear, homogeneous equation are also straightforward. These
have the form
d2 y dy
x2 2 + cx + d y = 0 (1.20)
dx dx
where c, d are constants. Again we need to find two independent solutions. We
guess that such solutions are y ∝ xk which, on substituting into equation (1.20),
gives the condition
k 2 + (c − 1)k + d = 0. (1.21)
If equation (1.21) has distinct real roots k1 and k2 then the general solution to
the Euler equation (1.20) can be written
If equation (1.21) has repeated or complex roots it is better to use the change
of variables t = log(x) to transform equation (1.20) to the constant coefficient
problem
d2 y dy
2
+ (c − 1) + d y = 0 (1.23)
dt dt
which can be solved using the techniques of section 1.1.1.
7
ORDINARY DIFFERENTIAL EQUATIONS SECOND ORDER EQUATIONS
and from this we can compute the Wronskian determinant (see section 1.1.1)
y1 vy1
W [y1 , y2 ](x) = 0 0 0
(1.26a)
y1 v y1 + vy1
= v 0 y12 (1.26b)
r
= Ke− p dx
(1.26c)
where the last equation follows from equation (1.11).
This allows us to directly compute v, the function relating the two solutions,
as
w e− r p dx
v=K dx + D (1.27)
y12
and hence the second solution is given by
y2 (x) = v(x)y1 (x) (1.28a)
r
w e− p dx
= Dy1 (x) + Ky1 (x) dx (1.28b)
y12
and, as we do not care about proportionality constants, we can write this as
w e− r p dx
y2 (x) = y1 (x) dx. (1.28c)
y12
d2 y dy
2
+ p(x) + q(x)y = r(x) (1.29)
dx dx
We often call r(x) a source term for reasons that will be clear from the examples.
In order to find the general solution to an inhomogeneous equation we first
need to find a particular solution yp (x) (sometimes called a particular integral in
some textbooks). That is some (any) solution of the inhomogeneous equation. If
one then looks at the function ỹ(x) = y(x) − yp (x) then it follows that it satisfies
the homogeneous equation
d2 ỹ dỹ
2
+ p(x) + q(x)ỹ = 0. (1.30)
dx dx
We can therefore write the general solution in the form
ỹ(x) = c1 y1 (x) + c2 y2 (x) (1.31)
8
ORDINARY DIFFERENTIAL EQUATIONS SECOND ORDER EQUATIONS
Note that if the trial function can be written as a linear combination of the
solutions to the homogeneous equation then, of course, the trial function will only
be a solution of the homogeneous equation. In this case it is usual to multiply the
trial solution by the independent variable until the trial solution is independent of
the complementary functions.
Examples!
Variation of parameters
(Not examined)
The method of undetermined coefficients is simple and effective when the in-
homogeneous term r(x) is simple, but useless in general. A more general method
can be found by assuming that the particular integral yp (x) is related to the solu-
tions y1 and y2 of the homogeneous problem, by
yp (x) = v1 (x)y1 (x) + v2 (x)y2 (x). (1.34)
9
ORDINARY DIFFERENTIAL EQUATIONS SECOND ORDER EQUATIONS
Here v1 and v2 are functions to be determined that allow us to compute the un-
known particular integral. Note that we have introduced two functions v1 and v2 in
order to find a single function yp , so that we have an additional degree of freedom.
We will use this later on to impose an additional condition on v1 and v2 which
will reduce things back to one degree of freedom in such a way as to simplify the
calculation.
To find v1 and v2 We calculate the first and second derivatives of yp and sub-
stitute into (1.2).
yp0 = (v1 y10 + v2 y20 ) + (v10 y1 + v20 y2 ). (1.35)
In order to simplify the expression for the second derivative we now use our ad-
ditional degree of freedom to impose the condition that the term in the second
bracket vanishes so that
v10 y1 + v20 y2 = 0. (1.36)
With this condition (1.35) becomes
Substituting into the inhomogeneous ODE (1.29) and remembering that y1 and y2
are solutions of the homogeneous ODE (1.32) we can see that
Hence the unknowns v1 and v2 are completely determined from our assumption
in equation (1.36) and our condition in equation (1.39) which can be written as
The unknown functions v1,2 can thus be determined by solving equation (1.40)
0
as a linear system for v1,2 and integrating the results. We find that
ry2
v10 = − (1.41a)
y1 y20
− y10 y2
r(x)y2 (x)
=− , (1.41b)
W [y1 , y2 ](x)
10
ORDINARY DIFFERENTIAL EQUATIONS BVPS AND IVPS
and similarly
r(x)y1 (x)
v20 = . (1.42)
W [y1 , y2 ](x)
This thus gives
w r(x)y2 (x)
v1 = − dx, (1.43a)
W [y1 , y2 ](x)
w r(x)y1 (x)
v2 = dx (1.43b)
W [y1 , y2 ](x)
and the particular integral of the original inhomogeneous ODE (1.29) follows
from equation (1.34).
d2 y dy
+ p(x) + q(x)y = r(x) (1.44)
dx2 dx
has general solution y(x) = yp (x) + c1 y1 (x) + c2 y2 (x) where yp (x) is a particular
solution of the equation and y1 (x) and y2 (x) are independent solutions of the
homogeneous equation and c1 and c2 are constants.
In order to obtain a unique solution to the differential equation we need some
way of fixing the constants c1 and c2 . One common way of doing this is to fix the
value of both y(x) and y 0 (x) at some point x0 . So that
y(x0 ) = C, (1.45a)
y 0 (x0 ) = D. (1.45b)
where C and D are given. Equation (1.44) together with the initial conditions
(1.45) is called an initial value problem.
Rather than give the value of the function y and its derivative y 0 at a single point
x0 one can give alternatively give the value of y at two point x0 and x1 , so that
the solution satisfies the differential equation (1.44) together with the boundary
conditions
y(x0 ) = α, (1.46a)
y(x1 ) = β. (1.46b)
11
ORDINARY DIFFERENTIAL EQUATIONS BVPS AND IVPS
One can then show that the boundary value problem of solving the differential
equation (1.44) on the region x0 ≤ x ≤ x1 subject to the boundary conditions
(1.47) has a unique solution provided the associated homogeneous boundary prob-
lem (i.e. the case where r(x) ≡ 0, α = 0, β = 0) only has the trivial solution (this
result is known as the Fredholm Alternative)
The unknown parameter λ may take any value. However, it may not be true that
the BVP given by equation (1.48) has a solution for any value of λ. The task of
finding a solution of the BVP involves finding not just solutions y that satisfy the
differential equation (1.48a), but also of finding which values of λ allow such a
solution to satisfy the boundary conditions given by equations (1.48b-1.48c).
In this simple case we can solve the BVP by inspection. However, to proceed
more systematically we split the problem into three separate cases depending on
the sign of the parameter λ:
1. λ = −α2 < 0. We first find solutions of equation (1.48a). As the differential
equation is homogeneous, these combine to form √ the general solution. In
this case the auxiliary equation has real roots ± −λ = ±α. It follows that
the general solution is
12
ORDINARY DIFFERENTIAL EQUATIONS BVPS AND IVPS
2. λ = 0. In this case the equation (1.48a) obviously has the general solution
y(x) = b1 x + b2 . (1.50)
Again, the only solution compatible with the boundary conditions (1.48b-
1.48c) is the trivial solution y ≡ 0 when b1 = 0 = b2 .
From the left boundary condition, equation (1.48b), we must have that c2 =
0 and so this reduces to
This condition is obviously that µ = n for any integer n1 ; given our previous
definitions, this gives the condition on λ that
λ = n2 , n = 1, 2, . . . (1.54)
The values of λ for which a solution of the BVP can be obtained are denoted λn ,
in this case are
λn = n2 , n = 1, 2, . . . , (1.56)
and are called the eigenvalues of the problem. The associated solutions yn (x) are
called the eigenfunctions. Clearly the eigenvalues are uniquely determined but
the eigenfunctions are not; any constant multiple of an eigenfunction yn is also a
solution of the BVP, and is also an eigenfunction.
A number of points should be noted for later consideration.
1
The condition (1.53) is satisfied with µ = ±n, where n is a positive integer. Since
sin (−nπx) = − sin (nπx), we may restrict to µ = n > 0 without loss of generality.
13
ORDINARY DIFFERENTIAL EQUATIONS BVPS AND IVPS
4. The nth eigenfunction (sin(nx)) has exactly n − 1 zeros inside the interval
[0, π] on which the BVP is defined. It also vanishes at the endpoints of the
interval.
d2 y dy
2
+ b(x) + c(x)y = λd(x)y. (1.58)
dx dx
This is a minor modification of previous second order ODEs that we have looked
at. We shall later specify boundary conditions in order to turn it in to a BVP.
r
A standard form is found by multiplying the whole equation by e b(x) dx to
rewrite equation (1.58) as
r 0 r r
e b(x) dx y 0 + c(x)e b(x) dx y(x) = λd(x)e b(x) dx y(x). (1.59)
This is called the self-adjoint form of the ODE. The conventional notation is
0
(p(x)y 0 ) − q(x)y + λw(x)y = 0. (1.60)
14
ORDINARY DIFFERENTIAL EQUATIONS BVPS AND IVPS
where w(x), the weight function, is positive (w ≥ 0) and non-trivial (w 6≡ 0), and
the differential operator L is given by
0
L : y → (p(x)y 0 ) − q(x)y. (1.62)
1. y = 0: Dirichlet condition.
2. y 0 = 0: Neumann condition.
3. y + ky 0 = 0: Radiation condition.
15
ORDINARY DIFFERENTIAL EQUATIONS BVPS AND IVPS
We note that the conditions at the boundaries x = x0 , x1 need not be the same.
There is one final possibility: periodic boundary conditions, where
Any of these are Sturm-Liouville boundary conditions, and given these the differ-
ential operator L is self-adjoint.
Properties of solutions
Properties 1. The eigenvalues λn are real.
Proof. Suppose that the eigenvalue λ, and possibly also the eigenfunction y are
complex. By definition
Ly = λw(x)y. (1.67)
If we take the complex conjugate then, as both the operator L and the weight
function w(x) are real, we have
Taking the difference between these two equations and integrating over the do-
main gives us
wx1 wx1
(ȳLy − yLȳ) dx = (λ − λ̄) w(x)y ȳ dx. (1.69)
x0 x0
As the operator is self-adjoint, we have from equation (1.63) that the left hand
side of equation (1.69) vanishes identically.
As the eigenfunction solution is non-trivial we can see that
and by assumption we have that w(x) > 0. Therefore in order for the right hand
side to also vanish, we must have that λ = λ̄, and hence the eigenvalues must be
real. Without loss of generality, we can therefore also take the eigenfunctions to
be real.
16
ORDINARY DIFFERENTIAL EQUATIONS BVPS AND IVPS
Proof. This is essentially given by the previous argument, as noted above, but an
alternative is as follows.
Suppose that yn = fn (x) + jgn (x) where both f, g are real and the eigenfunc-
tion yn is non-trivial. As the differential equation (and hence operator L) are real
it follows that
which, after checking the boundary conditions and using that the original BVP is
real, gives that both fn and gn must therefore be solutions of the original Sturm-
Liouville problem. Therefore we can take the non-zero real function of fn , gn to
be the real-valued eigenfunction with eigenvalue λn (at least one must be non-zero
as yn is non-trivial).
Properties 3. Eigenfunctions corresponding to distinct eigenvalues are orthogo-
nal.
Proof. Suppose that we have distinct eigenfunction solutions ym , yn with asso-
ciated eigenvalues λm 6= λn . Again we can use the self-adjoint nature of the
problem, given by equation (1.63), to show that
wx1
0= (ym Lyn − yn Lym ) dx (1.73a)
x0
wx1
= (λn − λm ) w(x)ym yn dx. (1.73b)
x0
17
ORDINARY DIFFERENTIAL EQUATIONS BVPS AND IVPS
Properties 5. The eigenfunctions are complete in the sense that for any bounded,
piecewise continuous function F (x) on [x0 , x1 ] the error N
wx1
" N
#
X
N = w(x) F (x) − an yn (x) dx (1.75)
x0 n=1
18
Chapter 2
Fourier Series
where an , bn are constants called the Fourier Coefficients and are such that the
series on the right hand side converges for all x.
More generally suppose that f (x) is a periodic function of period 2L so that
f (x + 2L) = f (x)
Then by using the result above and rescaling the x-coordinate we may expand the
nπx
function in a Fourier series in terms of the 2L-periodic functions cos L and
19
FOURIER SERIES ORTHOGONALITY
nπx
sin L
so that we may write
∞ h
1 X nπx nπx i
f (x) = a0 + an cos + bn sin (2.2)
2 n=1
L L
The question then arises: given a 2L-periodic function f (x), how can we find the
Fourier coefficients an , bn in the Fourier Series?
2.2 Orthogonality
Throughout this section (and indeed this chapter) the following identities are re-
peatedly used; they are sufficiently important that you should know them.
sin n + 12 π = (−1)n ;
sin (nπ) = 0, (2.3a)
n 1
cos (nπ) = (−1) , cos n + 2 π = 0. (2.3b)
We can straightforwardly use the formula cos(A+B) = cos A cos B −sin A sin B
and cos(A − B) = cos A cos B + sin A sin B to show that
1w
L
(m − n)πx (m + n)πx
= cos − cos dx (2.5b)
2 L L
−L
(m−n)πx L
(m+n)πx
L sin L
sin L
= − (2.5c)
2π m−n m+n
−L
= 0 except when m = n. (2.5d)
Inn = L. (2.6)
20
FOURIER SERIES FOURIER COEFFICIENTS
Note that the cos relation given by equation (2.9) does not hold for m = 0; in that
case the result is
wL nπx
cos dx = 2Lδ0n (2.11)
L
−L
Note: If we consider
wL
hf (x), g(x)i = f (x)g(x) dx (2.12)
−L
to be an inner product between the functions f and g, then the set of functions
n nπx nπx o
1, sin , cos (2.13)
L L
form an orthogonal basis for the space of 2L-periodic “nice” functions. In fact
we know from property 3 of the general theory of Sturm-Liouville boundary value
problems (even without doing the calculations above) that these functions are or-
thogonal. For Fourier series it is just as easy to do the explicit calculations but
for other types of expansion (e.g. Fourier-Bessel series which are used to describe
vibrating plates) it is much easier to use the general theory.
21
FOURIER SERIES FOURIER COEFFICIENTS
The orthogonality relations ensure that except for the single term bm , all other
terms vanish under integration. Similarly, using the appropriate cos term and the
orthogonality relations given by equations (2.9, 2.10) gives
wL mπx
Lam = f (x) cos dx, (2.16)
L
−L
and this equation holds even for m = 0 (due to the factor of 1/2 in the definition
of the Fourier Series, and using equation (2.11)).
We can summarise this to give the Euler relations:
1 w
L mπx
am = f (x) cos dx, (2.17a)
L L
−L
1 w
L mπx
bm = f (x) sin dx. (2.17b)
L L
−L
Note: Because the function f (x) is periodic, and the range of integration is a
single period, then one could equally use any 2L-periodic range in x; for example,
r 2L
one could equally well use 0 dx.
Note: Because the formulae for the Fourier coefficients an , bn are explicit, the
Fourier Series of a function is unique.
f (−x) = f (x)
22
FOURIER SERIES FOURIER COEFFICIENTS
for all x. Graphically this means that the function is symmetric about the y-axis.
The symmetrical integral of an even function can be expressed as follows
Z` Z`
f (x)dx = 2 f (x)dx
−` 0
f (−x) = −f (x)
for all x. Graphically this means that the function is symmetric under a reflection
about the y-axis and the x-axis.
The symmetrical integral of an odd function is zero i.e.
Z`
f (x)dx = 0
−`
23
FOURIER SERIES FOURIER COEFFICIENTS
If f (x) is an even function then this gives rise to a Fourier cosine series only
since
Zπ Zπ
1 2
an = f (x) cos nxdx = f (x) cos nxdx
π π
−π 0
Zπ
1
bn = f (x) sin nxdx = 0
π
−π
and so
∞ Zπ
a0 X 2
f (x) = + an cos nx with an = f (x) cos nxdx
2 n=1
π
0
24
FOURIER SERIES FOURIER COEFFICIENTS
If f(x) is an odd function then this gives rise to a Fourier sine series only since
Zπ
1
an = f (x) cos nxdx = 0
π
−π
Zπ Zπ
1 2
bn = f (x) sin nxdx = f (x) sin nxdx
π π
−π 0
and so
∞ Zπ
X 2
f (x) = bn sin nx with bn = f (x) sin nxdx
n=1
π
0
Example.
Find the Fourier series of f (x) which is equal to x in the range −π < x < π
and periodic outside, see Fig. 2.3.
Since f (x) is an odd function, it has a Fourier sine series
∞
X
f (x) = bn sin nx
n=1
with
Zπ
2
bn = f (x) sin nxdx
π
0
Zπ
2
= x sin nxdx
π
0
Zπ
2 cos nx
= xd −
π n
0
Zπ
2 − cos nx 2
= x + cos nxdx
π n nπ
0 π
2h π n
i 2 sin nx
= − (−1) +
π n nπ n 0
2
= (−1)n−1
n
25
FOURIER SERIES FOURIER COEFFICIENTS
4
Sawtooth function
-1
-2
-3
-4
-8 -6 -4 -2 0 2 4 6 8
x
Therefore
∞
X 2
f (x) = (−1)n−1 sin nx
n=1
n
1 1
= 2[sin x − sin 2x + sin 3x − · · · ]
2 3
If we consider the contribution of the first few terms to the sum we can gain some
idea of how the Fourier series approximates to the function in the interval 0 to π.
26
FOURIER SERIES FOURIER COEFFICIENTS
This last result can be used to show that the sum of the series
1 1 1π π
1 − + − ··· = =
3 5 22 4
where we have used the fact that f (x) is odd (and hence f (x) sin mπx
L
is even)
to write the expression for bn in terms of an integral over 0 < x < L (which is
half the period). For this reason such an Fourier sine series is called a half-range
Fourier series.
Alternatively given f (x) only on 0 ≤ x ≤ L we can extend the function f
onto −L < x ≤ L by making
f (−x) = f (x).
27
FOURIER SERIES FOURIER COEFFICIENTS
and then extend this to all values of x by making the function have period 2L, see
Fig. 2.4.
This is called an even periodic extension of the function. Since the function
is even it has a Fourier series which only contains the cosine terms. Hence the
function is represented by
∞
1 X nπx
f (x) = a0 + an cos , (2.20)
2 n=1
L
bn = 0, (2.21a)
2w
L mπx
an = f (x) cos dx (2.21b)
L0 L
and we have again used the fact that the integrand is even to write the expression
for an as an integral over half the range.
Note: we could also extend the function by making f be periodic with period
L, which immediately implies that it is also periodic with period 2L, see Fig. 2.4.
However in applications to differential equations it is the odd and even extensions
that are most useful.
Example.
Find the Fourier series of (i) the odd periodic extension and (ii) the even peri-
odic extension of
0 < x < 2`
1,
f (x) = `
0, 2
<x<`
in the range (0, `).
1. In (0, `)
∞
X nπx
f (x) = bn sin
n=1
`
where
Z`
2 nπx
bn = f (x) sin dx
` `
0
`
Z2
2 nπx
= sin dx
` `
0
28
FOURIER SERIES FOURIER COEFFICIENTS
ff (x)
−L L t
ll
fu
f(x) fe (x)
even
x
L −L L t
od
d
fo (x)
interval of
periodicity −L L t
29
FOURIER SERIES FOURIER COEFFICIENTS
2 h nπx i 2`
= − cos
nπ ` 0
2 h nπ i
= 1 − cos
nπ 2
Therefore
∞
2X1h nπ i nπx
f (x) = 1 − cos sin
π n=1 n 2 `
2 πx 2πx 1 3πx 1 5πx
= sin + sin + sin + sin + ···
π ` ` 3 ` 5 `
2. In (0, `)
∞
a0 X nπx
f (x) = + an cos
2 n=1
`
where
Z`
2 nπx
an = f (x) cos dx
` `
0
`
Z2
2 nπx
= cos dx
` `
0
2 ` h nπx i 2`
= sin
` nπ ` 0
2 nπ
= sin
nπ 2
and
`
Z2
2
a0 = dx = 1
`
0
Therefore
∞
1 X 2 nπ nπx
f (x) = + sin cos
2 n=1 nπ 2 `
1 2 πx 1 3πx 1 5πx
= + cos − cos + cos − ···
2 π ` 3 ` 5 `
30
FOURIER SERIES FOURIER’S THEOREM
31
FOURIER SERIES CALCULUS AND FOURIER SERIES
2. the function has a finite number of maxima, minima and discontinuities, and
to be a good start. Under this set of conditions the Fourier convergence theorem
of section 2.4 shows at least that the original definition of equation (2.23) holds
everywhere. If in addition we assume that
Note: It is important to note that although the conditions above are sufficient
and some of them can be relaxed, the example below shows that some of these
conditions are crucial.
Example 2.
32
FOURIER SERIES CALCULUS AND FOURIER SERIES
However
f (0) = 0 6= f (2π) = 2π
and therefore the series may not be differentiated term by term. If we do differen-
tiate the series term by term we obtain the series
∞
X
g(x) = −2 cos nx
n=1
which does not converge to f 0 (x) = 1 in 0 < x < 2π. Indeed the series is
divergent since the nnth term does not tend to zero as n → ∞.
The proof is straightforward but rather detailed. Of course, the constant term
in the new Fourier Series will need to be computed as it appears as the ‘constant
of integration’ when one integrates term by term.
Example.
33
FOURIER SERIES COMPLEX FORM OF FOURIER SERIES
Hence we get
∞
π2 X (−1)n
x2 = +4 2
cos nx
3 n=1
n
34
FOURIER SERIES PARSEVAL’S THEOREM
and substitute these into the definition of the Fourier Series, equation (2.2), then
we will end up with
X∞
f (x) = cn ejnπx/L , (2.27)
n=−∞
1 w
L
cn = f (x)e−jnπx/L dx. (2.29)
2L
−L
Example.
Find the complex Fourier series for
By above
Zπ
1
cn = ex e−jnx dx
2π
−π
π
1 e(1−jn)x
=
2π 1 − jn −π
1 1 (1−jn)π
− e−(1−jn)π
= . e
2π 1 − jn
(−1)n
= sinh π
π(1 − jn)
Therefore ∞
X (−1)n jnx
f (x) = e sinh π
−∞
π(1 − jn)
35
FOURIER SERIES SUMMARY OF RESULTS
If we expand out the Fourier series representation of f (x) given by (2.2) and use
the orthogonality relations given by equations (2.7, 2.9, 2.10) then we find
wL wL 1 nπx i 2
( ∞ h
)
X nπx
[f (x)]2 dx = a0 + an cos + bn sin dx
2 n=1
L L
−L −L
(2.31a)
wL
( ∞
)
1 2 Xh 2 nπx nπx i
= a0 + an cos2 + b2n sin2 dx
4 n=1
L L
−L
(2.31b)
" ∞
#
1 2 X
=L a + a2 + b2n . (2.31c)
2 0 n=1 n
This can be restated as
X
“kf k2 = squares of components of f on basis vectors”. (2.32)
The norm of f is often called the power contained in the function, and Parseval’s
Theorem or Parseval’s Identity, given by equation (2.31), relates the power to the
Fourier coefficients.
36
FOURIER SERIES SUMMARY OF RESULTS
f (x) is even −→ bn = 0
f (x) is odd −→ an = 0
(c) Non-periodic function f (x) defined over interval 0 < x < `, say.
i. Even periodic extension (half range expansion)
∞
a0 X nπx
f (x) = + an cos
2 n=1
`
where
Z`
2 nπx
an = f (x) cos dx
` `
0
where
Z`
2 nπx
bn = f (x) sin dx
` `
0
2. Differentiation
The Fourier series for f (x) in −` < x < ` may be differentiated term by
term to give a Fourier series for f 0 (x) if
3. Integration
The Fourier series for f (x) in −` < x < ` may always be integrated term
by term.
−∞
37
FOURIER SERIES SUMMARY OF RESULTS
where
Z`
1 jnπx
cn = f (x)e− ` dx
2`
−`
38
Chapter 3
Fourier transform
The Fourier series is a very powerful tool to analyse periodic functions or func-
tions defined in an interval, using their periodic extensions. However, we now
want to extend it to functions defined on the whole real line that are not periodic.
We do this by considering the case of a Fourier series of a simple function defined
on an interval of length 2L and then see what happens in the limit that L tends to
infinity.
3.1 An example
Consider the function
0 −L ≤ t < −1,
fL (t) = 1 −1 ≤ t < 1, with fL (t + 2L) = fL (t) ∀t ∈ R. (3.1)
0 1 ≤ t < L,
1 w
+L
πn 1 πn
cn = fL (t)e−j L t dt = sin (3.3)
2L πn L
−L
39
FOURIER TRANSFORM AN EXAMPLE
y fL(t)
−2L −L −1 +1 L 2L t
y
fL(t)
L increases
−L −1 +1 L t
y
f(t)
Limit of infinite L
−1 +1 t
We want to see how equation (3.3) and (3.4) change as L tends to infinity. We
introduce the variable
nπ
ωn = (3.5)
L
and look at how this changes with n by considering
(n + 1)π nπ π
∆ωn = ωn+1 − ωn = − = . (3.6)
L L L
Note that as L increases ∆ωn tends to zero and the ωn variables becomes contin-
uous, rather than discrete. With this notation we also have that
1 ∆ωn
= . (3.7)
n ωn
We start with equation (3.4) and rewrite it as
∞
X 1 ∆ωn
fL (t) = sin(ωn )ejωn t . (3.8)
ω =−∞
π ωn
n
1 w sin(ω) jωt
∞
= e dω . (3.9c)
π −∞ ω
40
FOURIER TRANSFORM DEFINITION
In this way we have expressed f (t) as the integral of the function sin(ω)/ω of the
continuous variable ω, multiplied by the complex exponential exp(jωt). We now
need to deal with equation (3.3). We define
1w
+L
c(ωn ) = fL (t)e−jωn t dt , (3.11)
2
−L
1 w
∞
1 +1 −jωt sin(ω)
Z
−jωt
c(ω) = f (t)e dt = e = . (3.12)
2 −∞ 2 −1 ω
Note that c(ω) is the function that appears as integrand in the expression (3.9c)
of f (t). Equation (3.12) is the generalisation of (3.3) to a non-periodic function,
while equation (3.9c) is the generalisation of (3.4). We will see shortly that the
former is called the Fourier transform of f (t), while the latter is its inverse-Fourier
transform.
3.2 Definition
We are now in a position to define the Fourier transform of a function of a real
variable, f (t), with t ∈ R. We assume that the function f (t) satisfies the following
conditions:
w∞
1. |f (t)| dt < ∞, (note that this requires that |f (t)| → 0 as |t| → ∞),
−∞
1 w
∞
F (ω) = F[f (t)](ω) := √ f (t)e−jωt dt. (3.14)
2π −∞
41
FOURIER TRANSFORM DEFINITION
1 w
∞
f (t) = √ F (ω)ejωt dω (3.15)
2π −∞
In this respect the function f (t) is called the inverse Fourier transform of F (ω)
and we write
1 w
∞
−1
f (t) = F [F (ω)](t) := √ F (ω)ejωt dω (3.16)
2π −∞
w∞
Remark - The requirement that |f (t)| dt < ∞ is necessary to ensure that the
−∞
integrals (3.13) and (3.15) converge and the Fourier transform and its inverse are
defined.
Remark - There are a variety of definitions of Fourier and inverse-Fourier trans-
form: they all differ by the factor in front of the integral. The definition given here
is symmetric, the factors in√front of the Fourier and inverse-Fourier transform are
identical, both equal to 1/ 2π. The “definition” given in Section 3.1 has a factor
1/2 in front of the Fourier transform, eq. (3.12), and a factor 1/π in front of the
inverse-Fourier transform, eq. (3.9c). The important point is that the product of
the two factors must be equal to 1/2π.
Example - Find the Fourier transform of the function
(
sin(t) −π ≤ t ≤ π,
f (t) = (3.17)
0 Otherwise .
42
FOURIER TRANSFORM PROPERTIES
1 w 1 w
∞ +π
−jωt
F (ω) = √ f (t)e dt = √ sin(t)e−jωt dt
2π −∞ 2π −π
w
+π
1 j
− e−j(1+ω)t dt
j(1−ω)t
=√ − e
2π 2 −π
+π
e−j(1+ω)t
j(1−ω)t
j e
=− √ −
2 2π j(1 − ω) −j(1 + ω) −π
1 1 j(1−ω)π −j(1−ω)π
1 −j(1+ω)π j(1+ω)π
=− √ e −e + e −e
2 2π 1 − ω 1+ω
1 1 jωπ −jωπ
1 jωπ −jωπ
=− √ e −e + e −e
2 2π 1 − ω 1+ω
1 1 1 jωπ −jωπ
=− √ + e −e
2 2π 1−ω 1+ω
1 2 2j
=− √ 2
2j sin(ωπ) = √ sin(ωπ) .
2 2π 1 − ω 2
(ω − 1) 2π
(3.18)
3.3 Properties
We list here some of the basic properties of Fourier transforms. We indicate with
the symbol F the Fourier transform, in the sense that
1 w
∞
F[f (t)] = √ f (t)e−jωt dt = F (ω) . (3.19)
2π −∞
Linearity
The Fourier transform is a linear operator:
43
FOURIER TRANSFORM PROPERTIES
Differentiation
If f (t) is continuous everywhere and its derivative has at most isolated jump dis-
continuities then
df
F = jωF[f (t)] . (3.21)
dt
The generalisation of this formula is
n
d f
F = (jω)n F[f (t)] . (3.22)
dtn
1 w df −jωt
∞
df
F =√ e dt . (3.23)
dt 2π −∞ dt
The next step is to integrate by parts and use the fact that |f (t)| → 0 as |t| → ∞
(as required to ensure convergence of the Fourier transform). For simplicity we
consider here the case that the function is differentiable everywhere. In this case,
integration by parts gives
1 w
∞
df 1 −jωt t→+∞
f (t) e−jωt dt =jω F (ω)
F =√ f (t)e t→−∞
+ jω √
dt 2π 2π −∞
(3.24)
Equation (3.22) can be obtained by applying (3.21) n times.
Parseval’s theorem
(not examinable)
Starting from the definition of Fourier transform and using the δ-function it is
possible to prove the following equality, that goes under the name of Parseval’s
theorem1
w∞ w∞
2
|f (t)| dt = |F (ω)|2 dω (3.25)
−∞ −∞
This relation has a clear physical meaning. We have already mentioned when
studying Parseval’s theorem for Fourier series that the moduli of the coefficients
of the series, |an |, |bn | are a measure of the strength of the frequency component
ωn . A similar interpretation can be given for the Fourier transform: if f (t) is some
physical signal, e.g. the amplitude of a sound or electromagnetic wave measured
1
Marc-Antoine Parseval des Chênes, French mathematician (1755-1836)
44
FOURIER TRANSFORM THE RESPONSE FUNCTION
as a function of time, then the modulus square of its Fourier transform, |F (ω)|2 ,
is the energy density in frequency space, i.e. |F (ω)|2 δω is the energy of the signal
in a frequency band of width δω centred around the frequency ω (in appropriate
units). The total energy of the wave is the integral of this energy density, i.e.
w∞
Total energy of wave = |F (ω)|2 dω . (3.26)
−∞
On the other hand, the total energy of a wave of amplitude f (t) is given by inte-
grating |f (t)|2 over time (in appropriate units), i.e.
w∞
Total energy of wave = |f (t)|2 dt . (3.27)
−∞
These two expressions must be equal. In this interpretation, equation (3.25), is just
a statement that the total energy measured must be the same whether we integrate
over time, equation (3.27), or over frequency, equation (3.26).
Shift properties
The Fourier transform of a shifted signal is equal to the Fourier transform of the
original signal, multiplied by a phase factor proportional to the shift:
F ejω0 t f (t) = F (ω − ω0 ) .
(3.29)
The signal processing interpretation of this result is that the multiplication of f (t)
by a constant frequency term, exp(jω0 t), is equivalent to shifting the entire fre-
quency spectrum of the signal by an amount ω0 . This is the mathematical basis
of the process of modulation, where a high frequency carrier at frequency ω0 is
modulated by a low frequency signal f (t).
45
FOURIER TRANSFORM THE RESPONSE FUNCTION
d2 y dy du
2
+ 3 + 7y(t) = 3 + 2u(t), (3.30)
dt dt dt
where u(t) is a given signal. We can use the Fourier transform to study the re-
sponse of this system to the modulation due to the function u(t). We let Y (ω) =
F[y(t)] and U (ω) = F[u(t)] denote the Foureir transform of y(t) and u(t) respec-
tively. Taking the Fourier transform of both sides of (3.30) we obtain
where
2 + 3jω
G(ω) = . (3.32)
7 − ω 2 + 3jω
The function G(ω) is the transfer function of the system: it determines how the
system responds to the various frequency components of the forcing term u(t). In
particular, we have
|Y (ω)| = |G(ω)| |U (ω)| . (3.33)
In other words, the strength of the frequency component ω of the output is equal
to that of the input multiplied by the modulus of the transfer function. In the case
of the function G(ω) defined by (3.32) we have
r
9 ω2 + 4
|G(ω)| = (3.34)
49 − 5 ω 2 + ω 4
This quantity is plotted in Figure 3.2: from the graph we can deduce that the
system acts like a band pass filter with maximal response at
√
q
1
ω= −4 + 7 85 ' 2.59 . (3.35)
3
An example of the effect of the response function on an input signal is shown in
Figure 3.3. In this case the input signal consists in a Gaussian pulse in time (solid
line, Figure 3.3 right). Its Fourier power spectrum is also Gaussian (solid line,
Figure 3.3 left). The power spectrum of the output (dashed line, Figure 3.3 left)
is the product Y (ω) = G(ω)U (ω) where G(ω) is given by equation (3.32). The
time dependent output y(t) (dashed line, Figure 3.3 right) is the inverse Fourier
transform of Y (ω). A more detailed analysis of the response of a linear system can
46
FOURIER TRANSFORM THE RESPONSE FUNCTION
1
|G(ω)|
0.5
0
0 5 10 15 20
ω
Figure 3.2: Plot of the modulus of the response function, equa-
tion (3.34), as a function of the frequency.
0.6
0.05
0.4
0
0.2
0 −0.05
−60 −40 −20 0 20 40 60 0 1 2 3 4
Frequency (ω) Time [× π]
Figure 3.3: Input and output signal in the frequency (left) and
time (right) domain for equation (3.30). The relation between
the Fourier series of the input U (ω) and output Y (ω) signal is
Y (ω) = G(ω)U (ω), with the response function G(ω) given by
equation (3.32).
47
FOURIER TRANSFORM SUMMARY
This is up to a factor just the Fourier transform where ω = js. It is used more
widely then the Fourier transform for solving differential equations as it is simpler
to use when one can claculate the inverse transform. However unlike the Fourier
transform there is no simple formula for calculating the inverse of a Lapalce trans-
form - one either has to look it up in a table (which does not always work) or else
regard it as a “complex Fourier transform” in which case one can calculate the
inverse using complex integration.
3.5 Summary
The Fourier transform is important because it we can use it to transform an ordi-
nary differential equation into an algebraic equation (or a PDE into an ODE). We
can then solve the algebraic equation and find the solution to the original ODE by
applying the inversse Fourier transform.
• Definition
The Fourier transform of a function f (t) is defined by
1 w
∞
F[f (t)](ω) = √ f (t)e−jωt dt = F (ω) .
2π −∞
1 w
∞
−1
F [F (ω)](t) = √ F (ω)ejωt dω = f (t)
2π −∞
• Linearity
F[αf (t) + βg(t)] = αF[f (t)] + βF[g(t)] .
48
FOURIER TRANSFORM SUMMARY
F ejω0 t f (t) = F (ω − ω0 ) .
49
Chapter 4
Laplace transforms
4.1 Introduction
A standard method for solving differential equations is to tranform the problem
into an equivalent, simpler problem. For a broad range of differential equations
(especially linear differential equations) the integral transforms
wβ
ỹ(s) = K(s, x)y(x) dx (4.1)
α
convert the problem from a complex differential equation for y(x) to a (hopefully)
simpler equation for ỹ(s). Here the kernel K(s, x) is a given function which
together with the real integration limits α, β defines the transform.
The broad outline of the method is the same for all types of integral transform:
1. apply the transform to the original equation(s) to get a simpler problem for
ỹ;
In certain cases it may only be formally possible to invert the transform; that is,
we may not be able to compute the solution in closed form, but only as an integral.
Here we will only study one type of integral transform, the Laplace transform.
50
LAPLACE TRANSFORMS DEFINITION AND EXISTENCE
4.2.1 Existence
Strictly the definition is not complete. For unrestricted functions f it is not at
all obvious that the integral in the definition (4.2) converges for any given value
of s. In fact, it is simple to check (using the strict definition of the unbounded
integral) that the integral need not converge for some values of s for certain simple
functions.
We look at ecx where c is some real nonzero constant. We therefore have
w∞ wA
cx
e dx = lim ecx dx (4.3a)
A→∞
0 0
A
ecx
= lim (4.3b)
A→∞ c 0
1 cA
= lim e −1 . (4.3c)
A→∞ c
We immediately see that the limit is finite, and hence the integral converges, if
c < 0, and that the integral diverges if c > 0. The case c = 0 is not covered by
this, but it is trivial to check that the integral diverges in this case as well.
This immediately motivates:
Theorem 4.2.1. If f (x) is a piecewise continuous function on 0 ≤ x ≤ A for
all A > 0, and |f (x)| ≤ Keax for all x ≥ M , where K, a, M are some real
constants and K, M are both positive, then the Laplace transform L [f (x)] = f˜(s)
as defined by (4.2) exists for s > a.
The proof is given, for example, in Boyce & DiPrima. We note that if the
conditions of the theorem hold then
lim f (x)e−sx = 0. (4.4)
x→∞
51
LAPLACE TRANSFORMS GENERAL PROPERTIES
2. If L [f1 (x)] = f˜1 (s) and L [f2 (x)] = f˜2 (s) then
This is the first shift theorem and is a simple application of the defini-
tion (4.2). It is extremely useful when inverting transforms.
df˜
L [xf (x)] = − (s). (4.8)
ds
This follows using the definition (4.2) and integration by parts. Less useful
in practice, it illustrates the difficulties that arise when transforming nonlin-
ear equations.
∂ f˜
∂f
L = . (4.9)
∂a ∂a
52
LAPLACE TRANSFORMS GENERAL PROPERTIES
We then recall from section 4.2.1 that in order for the Laplace transform to exist
we have
lim f (x)e−sx = 0. (4.4)
x→∞
We therefore get
df
L = sf˜(s) − f (0). (4.12)
dx
This allows us to transform first derivatives of f into algebraic expressions involv-
ing f˜ and the initial data f (0).
We can repeat this exercise for the second derivative, finding that
2
df d df
L 2
=L (4.13a)
dx dx dx
df
= sL − f 0 (0) (4.13b)
dx
h i
= s sf˜(s) − f (0) − f 0 (0) (4.13c)
= s2 f˜(s) − sf (0) − f 0 (0). (4.13d)
We note that in taking these steps we have assumed that the Laplace transform for
f 0 exists, and hence assumed that
lim f 0 (x)e−sx = 0. (4.14)
x→∞
Clearly we can extend this result to higher order by repeating these methods,
finding
n n−1
d f (x) n˜
X
L = s f (s) − sn−k−1 f (k) (0), (4.15)
dxn k=0
where we require
lim f (k) (x)e−sx = 0, k = 0, . . . , n. (4.16)
x→∞
53
LAPLACE TRANSFORMS INVERSE TRANSFORM
and hence
w∞
0= e−sx (f1 (x) − f2 (x)) dx (4.18a)
0
⇒ 0 = f1 (x) − f2 (x) (4.18b)
⇒ f1 (x) = f2 (x). (4.18c)
1 w ˜
γ+j∞
54
LAPLACE TRANSFORMS SPECIAL FUNCTIONS
=e −sa
f˜(s). (4.22e)
This is the second shift theorem, which should be compared to the first shift
theorem given in section 4.3.
These theorems are most useful to invert Laplace transforms, as they show that
h i
L−1 f˜(s + a) = e−sa f (x), (4.23a)
h i
L−1 e−sa f˜(s) = H(x − a)f (x − a). (4.23b)
55
LAPLACE TRANSFORMS SPECIAL FUNCTIONS
In particular, we note that to be strict the δ-function only makes sense when inte-
grated.
One way of thinking about the δ-function is that it takes the values
(
0, x 6= 0,
δ(x) = (4.25)
∞, x = 0.
This is obviously an unsatisfactory definition at x = 0. It is more usual to rely on
definitions based on limits, such as
δ(x) = lim D(x, ) (4.26a)
→0
where
(
1/ − 2 ≤ x ≤
2
D(x, ) = , (4.26b)
0 otherwise,
which gives
w∞ w2 1
D(x, ) dx = dx (4.26c)
−∞
2
=1 ∀. (4.26d)
Note that lim→0 D(0, ) = ∞.
We see that the δ-function, when multiplied by f and integrated, “picks out”
the value of f at the value of x where the argument of the δ-function vanishes.
Thus the Laplace transform has the form
w∞
L [δ(x − a)] = δ(x − a)e−sx dx (4.27a)
0
−as
=e , (4.27b)
56
LAPLACE TRANSFORMS CONVOLUTION
4.6 Convolution
(Not examined)
One point that we have not emphasized but which should be clear is that
and say that the Laplace transform of the convolution is the product of the Laplace
transforms.
Convolution Theorem.
57
LAPLACE TRANSFORMS APPLICATIONS
and, by introducing z = x − u
w∞ w∞
!
= f (u)e−su e−sz g(z) dz du (4.33e)
0 z=0
w∞ w∞
! !
= f (u)e−su du e−sz g(z) dz (4.33f)
0 0
= f˜(s)g̃(s) (4.33g)
as required.
Convolution is often used to write the final solution y(x) in terms of a formal
integral solution.
4.7 Applications
4.7.1 Solving ODEs
Laplace transforms may be used to solve ODEs. To do this we follow the follow-
ing steps.
Using linearity this amounts to rewriting the equation in terms of the Laplace
transform of the function we want to obtain.
58
LAPLACE TRANSFORMS APPLICATIONS
2. We use the formula for the Laplace transform of derivatives and insert the
initial condition.
3. This leads to an algebraic equation for the Laplace transform of the solution
we want to find, which we solve.
0 = L [y 00 + y] = L [y 00 ] + L [y] (4.35)
where in the last equality we used the boundary conditions. Inserting (4.36) in
(4.35) (step 3) yields an algebraic equation for ỹ(s), which we solve:
1
0 = s2 ỹ(s) − 1 + ỹ(s) ⇒ ỹ(s) = (4.37)
1 + s2
Finally (step 4), we need to determine the inverse Laplace transform of ỹ(s).
Looking at the Table of Laplace transforms in the Formula Sheet we find
−1 1
y(t) = L (t) = sin t (4.38)
1 + s2
59
LAPLACE TRANSFORMS APPLICATIONS
Given the appropriate initial conditions for y we could then solve this vector equa-
tion for fixed s, as it is a standard linear algebra problem. In particular we note
that the solution
−1 ˜
ỹ(s) = (sI − A) f (s) + y(0) (4.43)
only exists when sI − A is non-singular, which is when s is not an eigenvalue of
A.
4.7.3 PDEs
As Laplace transforms are integral transforms they affect only one variable at a
time. We can therefore use a Laplace transform to convert a PDE to an ODE. For
example, if we consider the wave equation in spherical symmetry
1 ∂ 2y 1 ∂ 2 (ry)
(r, t) = (r, t), (4.44)
c2 ∂t2 r ∂r2
we can perform a Laplace transform with respect to t to find
1 ∂ 2 (rỹ)
1 2 ∂y
s ỹ(s, r) − sy(r, 0) − (r, 0) = (s, r). (4.45)
c2 ∂t r ∂r2
Once we have initial conditions for y and its time derivative then we will have
an ODE for ỹ. In particular if we make the simplest choice that y and its time
derivative vanish at t = 0 then we are left with the ODE
d2 (rỹ) s 2
− (rỹ) = 0, (4.46)
dr2 c
which obviously has the solution
sr sr
rỹ = De c + Ee− c . (4.47)
As we are assuming that c > 0, and r, s > 0 by definition, in order for the solution
to be regular as r → ∞ we must have D ≡ 0.
60
LAPLACE TRANSFORMS APPLICATIONS
In order to compute the final solution we cannot directly invert the result that
follows from equation (4.47),
sr
e− c
ỹ(s) = E , (4.48)
r
as the “constant” E that follows from solving equation (4.46) may be E ≡ E(s).
Instead we have to use an appropriate boundary condition in order to fix E. If we
have a boundary condition such as
∂y
= g(t) (4.49)
∂r r=a
a2 cg̃(s) s(a−r)
ỹ(s) = − e c , (4.51)
r (a2 s + c)
61
Chapter 5
Wave equation
∂ 2u ∂ 2u ∂ 2u
a(x, y) + 2b(x, y) + c(x, y)
∂x2 ∂x∂y ∂y 2
(5.1)
∂u ∂u
+ d(x, y) + e(x, y) + f (x, y)u = 0.
∂x ∂y
We assume that c 6= 0 and perform the general change of variables
This implies that the coefficients of uξξ and uηη in equation (5.3) vanish, so that it
simplifies to the canonical form
4 ∂ 2u
ac − b2 = G. (5.5)
c ∂ξ∂η
62
WAVE EQUATION SMALL AMPLITUDE VIBRATIONS
As we will see with the wave equation, when a PDE can be reduced to this form it
has simple solutions (at least locally). Clearly the type of the solution will depend
strongly on the qualitative form of the values β, δ used to change to the canonical
form. In particular if
• b2 > ac: the roots of equation (5.4), β and δ are real and distinct. The
equation is called hyperbolic.
• b2 = ac: the roots of equation (5.4), β and δ are real and repeated. The
equation is called parabolic.
• b2 < ac: the roots of equation (5.4), β and δ are complex conjugates. The
equation is called elliptic.
Given an initial shape and velocity, i.e. prescribed initial conditions (or boundary
conditions in time)
∂y
y(x, 0) = p(x), (x, 0) = q(x), (5.7)
∂t
what is the shape at subsequent times?
63
WAVE EQUATION SMALL AMPLITUDE VIBRATIONS
Figure 5.2: A small section of the string and the forces acting on
it. By examining the forces over this small section and insisting
that the string not break we find the wave equation holds for small
displacements.
as for large slopes the methods used below will not work and the assumptions
made above do not hold. In the case of small slopes then we can use the small
angle expansion of the trigonometric functions to see that
∂y
α≈ . (5.9)
∂x x
But we can again use the small angle expansion to see that this implies that T1 ≈
T2 , i.e. the tension T is constant along the length of the string (as assumed above).
64
WAVE EQUATION SMALL AMPLITUDE VIBRATIONS
as the slopes, and hence the derivatives, are small (equation (5.8)).
Hence the total energy E is
wL 1 ∂y 2
" 2 #
∂y
E= ρ + c2 dx. (5.16)
0
2 ∂t ∂x
As there is no friction in the system, the total energy must be constant in time.
It is a straightforward exercise (check!) to prove from the wave equation and the
boundary conditions that the energy is constant.
65
WAVE EQUATION D’ALEMBERT’S SOLUTION
∂ 2y 2
2∂ y
= c (5.17)
∂t2 ∂x2
in the absence of boundaries.
We introduce a new coordinate system, sometimes called characteristic coor-
dinates:
ξ = x + ct, (5.18a)
η = x − ct. (5.18b)
The motivation for these coordinates should be clear from section 5.1, but for the
moment we will just rewrite the wave equation (5.17) in terms of y = y(ξ, η) and
see where it takes us.
Firstly we note that the chain rule gives us
∂ ∂ ∂
=1· +1· , (5.19a)
∂x ∂ξ ∂η
∂ ∂ ∂
=c· −c· . (5.19b)
∂t ∂ξ ∂η
Thus the wave equation (5.17) becomes
2 2
∂ ∂ ∂ ∂
+ y= − y (5.20)
∂ξ ∂η ∂ξ ∂η
which gives
∂ 2y
= 0. (5.21)
∂ξ∂η
This can immediately be solved one coordinate at a time to give, e.g.,
∂y
= h(ξ) (5.22)
∂ξ
and hence
y = f (ξ) + g(η) (5.23)
where the functions f, g are arbitrary. Therefore the general solution of the wave
equation (5.17), ignoring initial and boundary conditions, can be written as
66
WAVE EQUATION D’ALEMBERT’S SOLUTION
y = Aejω(t−x/c) . (5.25)
Here A is a complex constant, and it is understood that we always take the real
part of the right hand side to get the solution. If we write the complex constant A
as A = |A|ejφ then we see that
The key quantity is typically the frequency of the wave, ω, which implicitly defines
its wavelength 2πc
ω
.
67
WAVE EQUATION D’ALEMBERT’S SOLUTION
From our assumption that the string does not break we have that y is contin-
uous at x = 0. This tells us that the sum of the incident wave and the reflected
wave is the transmitted wave at x = 0 for all times t.
We can also use the force balance at x = 0 and that the tension in the string is
constant to give
∂y ∂y
T =T , (5.29)
∂x x=0− ∂x x=0+
∂y
which implies that ∂x is continuous at x = 0 as well.
In order to get a solution we have to assume a specific form for the transmitted
and reflected waves. We look for harmonic travelling wave solutions for these as
well, where
f = Dejω(t−x/c+ ) , (5.30a)
g = Bejω(t+x/c− ) . (5.30b)
The continuity conditions at x = 0 give
y continuous: A+B =D (5.31a)
∂y jωA jωB jωD
continuous: − + =− . (5.31b)
∂x c− c− c+
We can solve these equations to get the unknown coefficients B, D in terms of the
incident wave defined by A to get
c+ − c−
B= A, (5.32a)
c+ + c−
2c+
D= A. (5.32b)
c+ + c−
Given that the problem is linear we could have taken A to be 1 from the outset,
which would give the amplitudes and phases of the reflected (B) and transmitted
(D) waves relative to the incident wave.
Note three special cases:
1. c+ = c− : In this case B = 0 and D = A, which is exactly what we would
expect. The material properties do not change and information is perfectly
transmitted with no reflection.
2. ρ+ /ρ− → ∞, or equivalently c+ /c− → 0: In this case D = 0 and B = −A.
This corresponds to no transmission or perfect reflection; the reflected wave
has the same amplitude but is out of phase by π. This is “reflection from a
solid wall”.
3. ρ+ /ρ− → 0: In this case D = 2A and B = A. This corresponds to the
string becoming “massless”.
68
WAVE EQUATION SEPARATION OF VARIABLES
i.e. with variables x and t separated. We substitute this ansatz into the wave equa-
tion (5.33) finding
X T̈ = c2 X 00 T (5.37a)
or, on dividing by y = XT ,
1 T̈ X 00
= . (5.37b)
c2 T X
Now we note that the left hand side of this equation depends solely on T and its
derivatives with respect to time; that is, the left hand side is solely a function of t.
However, the right hand side depends solely on X and its derivatives with respect
to space; that is, the right hand side is solely a function of x.
How is it possible for a function solely of t to equal a function solely of x?
If both functions are trivial, i.e. equal to zero, then it is obviously true. If both
functions are constant then this gives us a constraint; the two functions must be
equal to the same constant. However, if either function varies then this equation
cannot hold, as (say) the left hand side will vary as t varies whilst the right hand
side stays constant. Therefore this variables separable form immediately implies
that both sides of the equation must be separately constant. In general, if the
equation can be written as the sum of functions of different variables then each
function must be separately constant.
69
WAVE EQUATION SEPARATION OF VARIABLES
In this case we have, from equation (5.37), that both sides must be equal to the
same constant. Let us call it λ. Then we can re-arrange our variables separable
form of equation (5.37) to give
X 00 − λX = 0, (5.38a)
T̈ − λT = 0. (5.38b)
So in this very special case we have reduced the partial DE defining the wave
equation, equation (5.33), into a pair of ordinary DEs which are coupled through
the separation constant λ.
This is a second order ODE with one additional unknown, the value of the
separation constant λ. So we do not expect to be able to completely solve the
problem with the information provided, as we only have two boundary conditions.
We do need to ensure that we find the value of the separation constant λ, however,
as it is this that couples the two ODEs that we are trying to solve in equation (5.38).
We find that it is best to consider the solution of the ODE given by equa-
tion (5.39) in terms of three separate cases.
1. λ > 0. In this case the general solution to the (constant coefficient, linear,
second order, homogeneous) ODE is
√ √
X(x) = Ae λx
+ Be− λx
. (5.40)
70
WAVE EQUATION SEPARATION OF VARIABLES
x = 0, L we immediately get
x=0: A + B = 0, (5.41a)
√
2 λL
x=L: Ae + B = 0, (5.41b)
⇒ A = B = 0. (5.41c)
Hence we only have the solution X ≡ 0 which implies that y ≡ 0, the trivial
solution. This is of no use.
x=0: A = 0, (5.43a)
x=L: A + BL = 0, (5.43b)
⇒ A = B = 0. (5.43c)
Hence we only have the solution X ≡ 0 which implies that y ≡ 0, the trivial
solution. This is of no use.
3. λ < 0. In this case the general solution to the (constant coefficient, linear,
second order, homogeneous) ODE is
√ √
X(x) = A cos −λx + B sin −λx . (5.44)
If this were to imply that B = 0 then once again we would only have the
trivial solution y = 0. However, there is another possibility. It could be
that the sin term vanishes instead. We know that sin vanishes whenever its
argument is an integer multiple of π. In other words, there is a non-trivial
solution to this equation if
√
−λL = nπ, n ∈ Z. (5.46)
71
WAVE EQUATION SEPARATION OF VARIABLES
Hence, by looking at all the possible cases, we see that the only non-trivial solu-
tions to the ODE for the spatial dependence, equation (5.39), are
nπx nπ 2
Xn (x) = An sin , λn = − , n ∈ Z. (5.47)
L L
Here An is an undetermined constant.
We have not yet used any conditions that could determine the value of the constant
coefficients Cn , Dn .
72
WAVE EQUATION SEPARATION OF VARIABLES
The set of solutions given by equation (5.50) satisfy the wave equation itself,
and the boundary conditions, but not the initial conditions. They are a special set
of solutions that have the property of retaining a constant shape of the form of
a sin function with only the amplitude varying in time. These are called normal
modes.
For reasons that are related to Sturm-Liouville theory, the spatial shape sin nπx
L
are also called the eigenfunctions, whilst the separation constants λn are called the
eigenvalues.
5.4.5 Superposition
We still do not have the solution to the problem that we really want, which is a
complete solution of the wave equation (5.33) satisfying the boundary conditions
of equation (5.34) and the initial conditions of equation (5.35). The normal mode
solutions given by equation (5.50) solve the wave equation and the boundary con-
ditions, but say nothing about the initial conditions.
However, we can exploit the fact that the wave equation and the boundary
conditions are linear. Therefore if we have two solutions y1 , y2 that satisfy the
wave equation and the boundary conditions then any linear combination
y = A1 y1 + A2 y2 (5.51)
will also satisfy both the wave equation and the boundary conditions. Therefore
we can take a general solution of the wave equation to be a linear combination of
the normal mode solutions given by equation (5.50):
∞
X nπct nπct nπx
y(x, t) = Cn cos + Dn sin sin . (5.52)
n=1
L L L
This is a superposition of the normal mode solutions. This general solution satis-
fies the wave equation and the boundary conditions, but has sufficient freedom in
the choice of the Cn , Dn coefficients to also satisfy the initial condition given by
equation (5.35).
To actually use this in practise we simply set t = 0 in our general solution
which, when matched against the initial data of equation (5.35), gives
∞
X nπx
p(x) = Cn sin . (5.53)
n=1
L
This is just a Fourier sine series for the Cn coefficients. Similarly, by taking the
time derivative of our general solution given by equation (5.52) and then setting
73
WAVE EQUATION SEPARATION OF VARIABLES
t = 0, we find
∞
X nπc nπx
q(x) = Dn sin . (5.54)
n=1
L L
This is again just a Fourier sine series for the Dn coefficients.
Using the standard Euler formulas for the coefficients of Fourier series we can
explicitly write
2w
L nπx
Cn = p(x) sin dx, (5.55a)
L0 L
2 w
L nπx
Dn = q(x) sin dx. (5.55b)
nπc 0 L
and using the compound angle formulas for sin we write this as
∞
1X nπ(x + ct) nπ(x − ct)
= Cn sin + sin (5.56b)
2 n=1 L L
1
= [p(x + ct) + p(x − ct)] , (5.56c)
2
where in the last step we have noted that
∞
X nπx
p(x) = Cn sin . (5.57)
n=1
L
We can interpret this by noting that p(x) was the initial “shape” of the wave,
i.e. y(x, 0) = p(x). Therefore p(x + ct) is just the original shape of the wave with
the origin of the x coordinate moved back from x = 0 to x = −ct. Therefore
the solution given by equation (5.56) represents the initial data splitting into two
travelling waves, one travelling to the left and one to the right, both being the
same “shape” as the initial data but with half the amplitude. Each wave moves
with speed c, giving the meaning of the wave speed.
74
WAVE EQUATION SEPARATION OF VARIABLES
It should be noted that this solution is not genuinely in the separation of vari-
ables form, however, this form, suitably re-interpreted, is d’Alembert’s solution
as seen in section 5.3.
Energy
We showed earlier in section 5.2.2 that the energy of the wave is
wL 1 ∂y 2
" 2 #
∂y
E= ρ + c2 dx. (5.58)
0
2 ∂t ∂x
75
Chapter 6
Heat equation
T x0 + d2 , t − T x0 − d2 , t
H(x0 , t) = − lim κA (6.2a)
d→0 d
∂T
= −κA (x0 , t). (6.2b)
∂x
76
HEAT EQUATION HEAT CONDUCTION
This indicates that there will be positive heat flow into the section from the left if
the temperature is greater to the left of x = x0 , as we would expect. Repeating
the calculation for heat flow out of the section to the right through the surface at
x = x0 + δx gives
∂T
H(x0 + δx, t) = −κA (x0 + δx, t). (6.3)
∂x
Therefore the net rate at which heat flows into (or out of) the section of the bar
between x0 and x0 + δx is given by
Q = H(x0 , t) − H(x0 + δx, t) (6.4a)
∂T ∂T
= κA (x0 , t) − (x0 + δx, t) , (6.4b)
∂x ∂x
and hence the amount of heat entering the section in time δt is
∂T ∂T
Q δt = κA (x0 , t) − (x0 + δx, t) δt. (6.5)
∂x ∂x
This has to be balanced by the amount of heat actually absorbed by this section
of the bar. The average change in temperature δT in time δt is proportional to the
heat that flows into the section, Q δt and inversely proportional to the mass of the
section. As the mass is proportional to the volume of the section A δx, we write
this as
Q δt
δT = , (6.6)
sA δx
where s is related to (but is not exactly) the specific heat of the material of the bar.
Now, for a sufficiently small section of bar the average temperature change in the
section will be
∂T
δT ≈ δt (6.7a)
∂t
Q δt
= (6.7b)
sA δx
77
HEAT EQUATION SEPARATION OF VARIABLES
by using equation (6.6). Matching this against the expression for the heat flowing
in to the bar, equation (6.5), gives
κ ∂T (x0 , t) − ∂T
∂T ∂x ∂x
(x0 + δx, t) δt
δt ≈ . (6.8)
∂t s δx δt
Taking the limits δx → 0, δt → 0 gives
∂T ∂ 2T
= α2 2 . (6.9)
∂t ∂x
We note that the standard notation that we will use from now on is
∂y ∂ 2y
(x, t) = κ2 2 (x, t). (6.10)
∂t ∂x
Here κ is some constant related to the thermal conductivity of the material, or in
general the diffusivity of the problem that we are interested in.
∂y ∂ 2y
(x, t) = κ2 2 (x, t) (6.10)
∂t ∂x
with Dirichlet boundary conditions
X Ṫ = κ2 X 00 T (6.14a)
or, on dividing by y = XT ,
1 Ṫ X 00
= . (6.14b)
κ2 T X
78
HEAT EQUATION SEPARATION OF VARIABLES
Again the two sides are separately constant, and by the same argument made
as for the wave equation in section 5.4.1 we can write the separated equations as
X 00 − λX = 0, (6.15a)
Ṫ − κ2 λT = 0. (6.15b)
Again, λ is some constant whose value is to be determined.
At this point we need boundary conditions in order to solve the ODEs give by
equation (6.15). Exactly as in the wave equation case the boundary conditions in
equation (6.11) imply Dirichlet boundary conditions for X at x = 0, L, giving the
fully-posed ODE problem
X 00 − λX = 0, X(0) = 0, X(L) = 0. (6.16)
This is identical to the problem in the wave equation case (equation (5.39) in
section 5.4.2). Hence we can write down the solution as
nπx nπ 2
Xn (x) = An sin , λn = − , n ∈ Z. (6.17)
L L
Here An is an undetermined constant.
We now need to solve for the time dependence, which is given by equa-
tion (6.15b). Once the eigenvalue λ is known, this is a simple first order ODE.
Given that the results for the spatial dependence, summarized in equation (6.17),
tell us that λ is negative, we can write down the solution to the time dependence
as
Tn (t) = Cn exp λn κ2 t
(6.18a)
nπ 2 2
= Cn exp − κt . (6.18b)
L
Again, Cn is an undetermined constant.
Combining our solutions and superposing gives the general solution
∞ nπx
X nπ 2 2
y(x, t) = Cn sin exp − κt . (6.19)
n=1
L L
Once again we have abused notation to absorb the free constants into the single
set of constants Cn .
Again we determine the free constants by noting that our initial conditions
given by equation (6.12) must match our general solution of equation (6.19) when
evaluated at t = 0, to give
X∞ nπx
p(x) = Cn sin . (6.20)
n=1
L
This is exactly the same Fourier Series problem as in the wave equation case.
79
HEAT EQUATION MORE COMPLEX MATERIAL PROPERTIES
80
HEAT EQUATION MORE COMPLEX BOUNDARY CONDITIONS
Once again we can note that both sides must be separately constant, leading to the
ODEs
Ṫ + λT = 0, (6.28a)
0 0
(p(x)X (x)) − q(x)X(x) + λr(x)X(x) = 0. (6.28b)
Once again we have an eigenvalue problem for the unknown separation con-
stant λ. However, the spatial dependence governed by equation (6.28b) is clearly
far more complex than before. However, we can note that equation (6.28b) is ex-
actly the form of a Sturm-Liouville problem as given in section 1.2.2; in particular,
equation (6.28b) should be compared to equation (1.60).
Particular results from Sturm-Liouville theory can immediately be used in the
solution of more complex boundary value problems such as this. For example,
the theory immediately tells us which boundary conditions guarantee a solution
(see section 1.2.2), that eigenvalues and eigenfunctions with qualitatively similar
properties to those seen earlier exist (see section 1.2.2), and that there are orthog-
onality conditions that can be used in place of the Fourier Series techniques to
determine the resulting free coefficients. Therefore once the solution to the spatial
dependence given by equation (6.28b) is found, all of the standard separation of
variables steps can be applied.
∂y ∂ 2y
(x, t) = κ2 2 (x, t) (6.10)
∂t ∂x
with a broader range of boundary conditions than the simple Dirichlet conditions
81
HEAT EQUATION MORE COMPLEX BOUNDARY CONDITIONS
we can compare the results of the separation of variables ansatz in section 6.3.1
to those of Sturm-Liouville theory in section 1.2.2 to see that any appropriate
combination of
1. y = 0: Dirichlet condition;
2. y 0 = 0: Neumann condition;
3. y + ky 0 = 0: Radiation condition;
5. y periodic;
We look at the case λ = µ2 > 0 where the auxiliary equation has distinct
real roots ±µ. Rather than writing the general solution to the ODE in terms of
exponentials we instead write it as
It is straightforward to check that the only solution compatible with these bound-
ary conditions is the trivial solution. The same is true if we attempt to impose that
82
HEAT EQUATION INHOMOGENEOUS PROBLEMS
the solution y is periodic; this implies that X is periodic, which is not possible for
the hyperbolic trigonometric functions that make up this solution.
In contrast, the case λ = 0 may have a solution. The general solution of
equation (6.30) when λ = 0 is obviously
X(x) = Ax + B. (6.33)
In most cases we can still check that A = 0 (but this need not be true, for exam-
ple, for the radiation condition!), but the solution where B 6= 0 is compatible with
periodic boundary conditions, and with a range of Dirichlet and radiation condi-
tions. This leads to additional solutions for the time dependence that are similarly
of the form
T (t) = Ct + D. (6.34)
Typically the solution being bounded either in the past or the future implies that
C = 0. However, the constant solution is again frequently valid. It should be
noted that in more complex domains these additional solutions often lead to con-
siderably more complex behaviour that must be considered.
∂y ∂ 2y
(x, t) = κ2 2 (x, t) (6.10)
∂t ∂x
suggested by the generalized heat equation (6.24); we add the source term F to
get
∂y ∂ 2y
(x, t) = κ2 2 (x, t) + F (x, t). (6.35)
∂t ∂x
For simplicity and for later use we will assume trivial Dirichlet boundary condi-
tions
y(0, t) = 0 = y(1, t) (6.36)
on the simple domain x ∈ [0, 1].
Clearly for general source term F we cannot use separation of variables, as our
ansatz y = XT does not lead to an equation containing only separably constant
terms. Instead we make the assumption that the spatial dependence of the solution
can still be described by a Fourier Series, and because of the simple boundary
83
HEAT EQUATION INHOMOGENEOUS PROBLEMS
conditions of equation (6.36), this implies a Fourier Sine Series. This gives us the
more general ansatz
X∞
y(x, t) = yn (t) sin(nπx). (6.37)
n=1
We still cannot start solving our problem without making additional assump-
tions. However, we can substitute our ansatz into the inhomogeneous heat equa-
tion (6.35) and see what results. We find
∞
X
ẏn (t) − (nπ)2 yn (t) sin(nπx) = F (x, t).
(6.38)
n=1
At any instant in time the left hand side is a Fourier Sine Series. Therefore in order
to be consistent the source term F must also be a Fourier Sine Series. Making that
assumption,
X∞
F (x, t) = Fn (t) sin(nπx), (6.39)
n=1
This is a set of ODEs to solve for the unknown yn (t) in terms of the Fn , which
can be computed from the known source term F . In particular, this linear ODE
has the obvious solution
wt
yn (t) = Cn exp −(nπ) t + Fn (s) exp −(nπ)2 s ds.
2
(6.42)
0
Once Fn has been computed we can compute yn and our original ansatz in equa-
tion (6.37) gives the solution.
∂y ∂ 2y
(x, t) = κ2 2 (x, t) (6.10)
∂t ∂x
84
HEAT EQUATION INHOMOGENEOUS PROBLEMS
We have assumed that the domain has length 1 for simplicity and that f, g are
given functions.
It should immediately be obvious that separation of variables fails. This is
because we can no longer solve the separated problem for the spatial dependence,
as the boundary conditions are functions of time. However, we can use a simple
trick inspired by techniques for solving inhomogeneous ODE problems to reduce
this inhomogeneous PDE to a form that can be solved by separation of variables.
First we note that the heat equation (6.10) is linear so we can superpose solu-
tions, even when we have complex boundary conditions such as equation (6.43).
Therefore if we had a particular solution yP (x, t) that satisfies the heat equation
and the boundary conditions (6.43), we could superpose it with another solution
yG (x, t) of the heat equation, and the combined solution
will satisfy the heat equation. More importantly, we note that if the solution
yG (x, t) vanishes at the boundaries x = 0, 1 then the combined solution y(x, t)
also satisfies the boundary conditions (6.43).
Unfortunately it is just as difficult to find a single solution to the full inho-
mogeneous problem as it is to find the general solution. Instead we suppose we
have a particular solution yP that satisfies the boundary conditions but not the
heat equation. When we form the combined solution using equation (6.44) we
find that the combined solution satisfies the boundary conditions but not the heat
equation (6.10). It will however satisfy the inhomogeneous heat equation (6.35),
where the source term F (x, t) can be computed directly from yP .
Thus we have reduced the problem of finding the general solution of the heat
equation (6.10) with complex boundary conditions (6.43) to the problem of find-
ing a simple function yP that satisfies the complex boundary conditions (6.43),
plus the problem of solving the inhomogeneous heat equation (6.35) with trivial
Dirichlet boundaries as studied in section 6.5.1.
We still have the problem of finding a sufficiently simple particular solution
yP . However, this can itself be simplified by assuming that it separates, yP (x, t) =
XP (x)TP (t). In particular, if we choose XP to be a linear polynomial, then it
trivially vanishes under the second derivative in the heat equation. We then choose
T to satisfy the boundary conditions. In particular the choice
85
HEAT EQUATION INHOMOGENEOUS PROBLEMS
has the correct behaviour at the boundaries and vanishes under the second spatial
derivative. Therefore we can see that this solution satisfies the equation
∂yP ∂ 2 yP
(x, t) = (x, t) + xġ(t) + (1 − x)f˙(t). (6.46)
∂t ∂x2
This gives the simple source term
meaning that the general solution of the heat equation with inhomogeneous bound-
ary conditions (6.43) reduces to
where yG is the general solution of the inhomogeneous heat equation (6.35) with
source term (6.47) and trivial Dirichlet boundary conditions.
86
Chapter 7
Laplace’s equation
7.1 Introduction
Laplace’s equation is simply
∇2 φ = 0. (7.1)
In R1 and Cartesian coordinates Laplace’s equation reduces to
d2 φ
= 0, (7.2)
dx2
where φ(x) represents, e.g., the static shape of a stretched string, or the steady
temperature in a conducting bar, or something similar.
In R2 and Cartesian coordinates Laplace’s equation reduces to
2
∂2
∂
+ φ = 0, (7.3)
∂x2 ∂y 2
where φ(x, y) represents, e.g., the static shape of a stretched membrane such as a
soap film, or the steady temperature in a conducing sheet, or something similar.
In R3 and Cartesian coordinates Laplace’s equation reduces to
2
∂2 ∂2
∂
+ + φ = 0, (7.4)
∂x2 ∂y 2 ∂z 2
87
LAPLACE’S EQUATION SEPARABLE SOLUTIONS
at each point of the bounding surface. We note that this is two points in R1 , a
bounding curve in R2 , and a bounding surface in R3 . The (unit) vector n̂ is the
normal to the bounding surface.
X 00 − λX = 0, (7.6a)
Y 00 + λY = 0. (7.6b)
Both are BVPs with trivial boundary conditions. This is different to the situation
for the heat and wave equations where only one of the ODEs was a BVP.
We solve the eigenvalue problem for the x dependence as normal, finding that
the only non-trivial solution is
We are left with the solution for the y dependence, and using this value of the
separation constant λn in the ODE of equation (7.6b) gives
88
LAPLACE’S EQUATION SEPARABLE SOLUTIONS
The trivial Dirichlet boundary conditions immediately imply that the only solution
compatible with the boundary conditions is the trivial solution Yn ≡ 0 and hence
φ(x, y) ≡ 0.
This illustrates two points. One is the maximum principle: solutions of Laplace’s
equation always take their extreme values (both maxima and minima) at the bound-
aries of the domain. This holds in all dimensions and (reasonable) domains. Al-
though not proved or used in this course, it is easily motivated by considering the
stationary limit of the heat equation.
The second point is that for the vast majority of interesting problems using
Laplace’s equation we must expect them to be inhomogeneous, either through
their boundary conditions or a source term (Poisson’s equation). There is a straight-
forward method of dealing with these problems, exploiting the linearity of the
equation.
7.2.1 Cartesians
For simplicity we explain the technique using the two-dimensional Laplace equa-
tion 2
∂2
∂
+ φ=0 (7.3)
∂x2 ∂y 2
on the unit square (hence using Cartesian coordinates) with some set of boundary
conditions
where each ui satisfies Laplace’s equation in the interior, satisfies the bound-
ary condition Ui on boundary i and vanishes (i.e., satisfies a trivial Dirichlet
boundary condition) on all other boundaries.
89
LAPLACE’S EQUATION SEPARABLE SOLUTIONS
Once we have found each of the ui , we can reconstruct the full solution to the
original problem using
The values of the α coefficients is set by the restrictions given at the corners of the
domain, so that
This simple linear system can be solved by forward substitution, provided the
boundary conditions are consistent in the corners.
90
LAPLACE’S EQUATION SEPARABLE SOLUTIONS
91
LAPLACE’S EQUATION SEPARABLE SOLUTIONS
The constant R defines the radius of the disk and the constants R1,2 the inner and
outer radii of the annulus.
There are two key points to note. The first is that the boundaries are smooth,
so there are no corners to worry about as in the Cartesian case. This means that the
step of finding the difference function as in section 7.2.1 is considerably simpler
– only one boundary needs considering in the case of the annulus, and the step is
completely unnecessary for the disk.
The second point is that there is no explicit boundary condition in the angular
direction, as there is no explicit boundary. However, in order to have a single-
valued solution we need φ to be 2π-periodic. In addition, in the case of the disk
we require the solution to be finite at r = 0. These combine with the standard
separation of variables techniques to give solutions of the form
∞
1 X
φ(r, θ) = a0 + rn (an cos(nθ) + bn sin(nθ)) (7.22)
2 n=1
for the disk problem, and imposing boundary conditions at r = R and using
orthogonality will give the free constants an , bn .
92
Chapter 8
Vector Calculus
93
VECTOR CALCULUS VECTORS AND SCALARS
Figure 8.1: As t varies between a and b the vector r(t) traces out
a curve.
I˜ → R3
s 7→ r̃(s)
94
VECTOR CALCULUS VECTORS AND SCALARS
If we think of the parameter t representing time, then ṙ(t) is the velocity and
v(t) = |ṙ(t)| is the speed. This gives us the following formula for the length of
the curve between the point A with position vector r(t0 ) and the point B with
position vector r(t1 ):
Z t1 Z t1 Z t1
1/2
ẋ(t)2 + ẏ(t)2 + ż(t)2
L= v(t)dt = |ṙ(t)|dt dt
t0 t0 t0
What happens if we use some other parameterisation of the same curve? Let
s = f (t) and r̃(s) = r(f (s)) as above. Then by the chain rule
dr̃(s) dr(t) dt
=
ds dt ds
Let f (t0 ) = s0 and f (t1 ) = s1 . Then the length of the curve from A to B using
the s parameterisation is given by
Z s1
dr̃(s)
L̃ = ds ds
s
Z 0s1
dr(t) dt
= dt ds ds
s
Z 0s1
dr(t) dt
= dt ds ds
s0
Z t1
dr(t)
= dt dt
t0
=L
Hence the length of the curve is independent of the parameterisation. This feature
of the integral is called re-parameterisation invariance.
To visualise this function we can plot the graph of the surface given by
z = φ(x, y)
96
VECTOR CALCULUS VECTORS AND SCALARS
y
4
0 x
-2
-4
-4 -2 0 2 4
1 3
Figure 8.3: Contour plot of the function φ(x, y) = 12
y −y−
1 2
4
x + 72 .
97
VECTOR CALCULUS VECTORS AND SCALARS
F : R2 → R2
(x, y) 7→ F(x, y) = F1 (x, y)i + F2 (x, y)j
98
VECTOR CALCULUS VECTORS AND SCALARS
99
VECTOR CALCULUS VECTORS AND SCALARS
x = r1 + tv1
y = r2 + tv2
x = r3 + tv3
and r0 = r1 i + r2 j + r3 k and v = v1 i + v2 j + v3 k.
Differentiating (8.1) with respect to t using the chain rule gives
df ∂φ dx ∂φ dy ∂φ dz
= + +
dt ∂x dt ∂y dt ∂z dt
∂φ ∂φ ∂φ
= v1 + v2 + v3
∂x ∂x ∂z
100
VECTOR CALCULUS VECTORS AND SCALARS
∇v φ = v̂·∇φ
= |v̂||∇φ| cos θ
= |∇φ| cos θ
−|∇φ| ≤ ∇v φ ≤ |∇φ|
101
VECTOR CALCULUS VECTORS AND SCALARS
∇v φ = 0
⇒v·∇φ = 0
⇒the tangent vector v is perp. to ∇φ
⇒∇φis perpendicular to the tangent space at P
⇒∇φis normal to the level surface φ = c0
∇φ = k n̂ (8.2)
where k = |∇φ|.
On the other hand we know from Theorem 1 that the maximum value of ∇v φ is
|∇φ| and that it obtains this value when v points in the same direction as ∇φ, i.e.
when v̂ = n̂. Thus
|∇φ| = ∇n̂ φ (8.3)
Where we have taken n̂ to be the choice of normal that points in the direction of
increasing φ.
Combining formula (8.2) with formula (8.3) we obtain
∇φ = (∇n̂ φ) n̂
∂φ
To simplify notation we write to denote the directional derivative in the normal
∂n
direction ∇n̂ φ. Hence we can write
∂φ
∇φ = n̂
∂n
To summarise:
Gradient (geometric definition)
The gradient of the scalar field φ is the vector field ∇φ given by
∂φ
∇φ = n̂
∂n
102
VECTOR CALCULUS VECTORS AND SCALARS
∂φ
where n̂ is the unit normal to the surfaces φ = const and is the directional
∂n
derivative of φ in the n̂ direction.
Gradient (Cartesian definition)
If the scalar field φ is given in Cartesian coordinates by φ(x, y, z) then
∂φ ∂φ ∂φ
∇φ = i+ j+ k
∂x ∂y ∂z
Directional Derivative
The directional derivative
φ(r0 + tv) − φ(r0 )
∇v φ(r0 ) = lim
t→0 t
= v·∇φ
∂φ ∂φ ∂φ
= v1 + v2 + v3 (in Cartesian coordinates)
∂x ∂y ∂z
Geometric method:
The level surfaces of φ =const. are the surfaces r =const. which are simply
spheres centre the origin. Hence the normal to the surface points in the same
direction as the position vector. Thus n̂ = r̂. Hence
∂φ ∂φ
∇φ = n̂ = r̂ = 1.r̂
∂n ∂r
103
VECTOR CALCULUS VECTORS AND SCALARS
More generally if the scalar field φ only depends upon the distance r from the
origin and not the direction then we can write φ(r) = f (r). The level surfaces are
still spheres centre the origin and hence we still have n̂ = r̂. Thus
∂φ ∂φ df
∇φ = n̂ = r̂ = r̂ = f 0 (r)r̂
∂n ∂r dr
Example 2: Calculate the equation of the tangent plane to the surface S given by
x3 y − yz 2 + z 5 = 9 at the point P given by (3, −1, 2).
Let φ = x3 y − yz 2 + z 5 then at P , the value of φ = 9 so that S is the level surface
given by φ(x, y, z) = 9. Then
F = φc
104
VECTOR CALCULUS VECTORS AND SCALARS
where φ is a scalar field and c is a constant vector, we define divF by the formula
divF = ∇φ·c
We also require that the divergence is a linear operator, so that if F and F̃ are two
vector fields then
div(F + F̃) = divF + divF̃
Using the fact that we may write a general vector field in a Cartesian basis as
F = F1 i + F2 j + F3 k
Then
divF = div(F1 i) + div(F2 j) + div(F3 k)
= ∇(F1 )·i + ∇(F2 )·j + ∇(F3 )·k
∂F1 ∂F2 ∂F3
= + +
∂x ∂y ∂z
105
VECTOR CALCULUS VECTORS AND SCALARS
∇2 φ = div(gradφ) = ∇·(∇φ)
106
VECTOR CALCULUS VECTORS AND SCALARS
∂φ ∂φ ∂φ
∇φ = i+ j+ k
∂x ∂y ∂z
= ex y 3 sin zi + 3ex y 2 sin zj + ex y 3 cos zk
Alternatively
∂ 2φ ∂ 2φ ∂ 2φ
∇2 φ = + + 2
∂x2 ∂y 2 ∂z
∂ (e y sin z) ∂ 2 (ex y 3 sin z) ∂ 2 (ex y 3 sin z)
2 x 3
= + +
∂x2 ∂y 2 ∂z 2
x 3 x x 3
= e y sin z + 6e y sin z − e y sin z
F = φc
107
VECTOR CALCULUS VECTORS AND SCALARS
where φ is a scalar field and c is a constant vector, we define curlF by the formula
curlF = ∇φ×c
We also require that the curl is a linear operator, so that if F and F̃ are two vector
fields then
curl(F + F̃) = curlF + curlF̃
Using the fact that we may write a general vector field in a Cartesian basis as
F = F1 i + F2 j + F3 k
Then
curlF = curl(F1 i) + curl(F2 j) + curl(F3 k)
= ∇(F1 )×i + ∇(F2 )×j + ∇(F3 )×k
∂F1 ∂F1 ∂F2 ∂F2 ∂F3 ∂F3
= − k+ j + k− i + j− i
∂y ∂z ∂x ∂z ∂x ∂y
∂F3 ∂F2 ∂F3 ∂F1 ∂F2 ∂F1
= − i+ − j+ − k
∂y ∂z ∂x ∂z ∂x ∂y
108
VECTOR CALCULUS VECTORS AND SCALARS
F = ∇φ = 2xyz 3 i + x2 z 3 j + 3x2 yz 2 k
So in this example we find that ∇×∇φ = 0. In fact this is not just chance but is
a general result.
Theorem: Let F be a C 2 (ie twice differentiable with continuous second deriva-
tive) vector field then:
curl(gradφ) = ∇×∇φ = 0
Proof:
∂φ ∂φ ∂φ
∇φ = i+ j+ k
∂x ∂y ∂z
So that
i j k
∂ ∂ ∂
∇×∇φ = ∂x ∂y ∂z
∂φ ∂φ ∂φ
∂x ∂y ∂z
2
∂ 2φ
2
∂ 2φ
2
∂ 2φ
∂ φ ∂ φ ∂ φ
= − i+ − j+ − k
∂y∂z ∂z∂y ∂x∂z ∂z∂x ∂x∂y ∂y∂x
=0
∂ 2φ ∂ 2φ
where the C 2 condition is required to show that, for example, =
∂y∂z ∂z∂y
Example: Let F = x2 ey i + x ln zj + (x + z)k. Calculate G = ∇×F and also
∇·G.
i j k
∂ ∂ ∂
∇×F = ∂x = (0 − x/z)i + (0 − 1)j + (ln z − x2 ey )k
∂y ∂z
x2 ey x ln z (x + z)
109
VECTOR CALCULUS VECTORS AND SCALARS
x
Hence G = − i − j + (ln z − x2 ey )k. Calculating the divergence of this we
z
obtain
∂G1 ∂G2 ∂G3
∇·G = + +
∂x ∂y ∂z
1 1
=− +0+ =0
z z
Hence in this example ∇·(∇×F) = 0. Again this is not just chance but is a
general result:
Theorem: Let F be a C 2 vector field, then
div(curlF) = ∇·(∇×F) = 0
Proof:
Let F = F1 i + F2 j + F3 k, then
i j k
∂ ∂ ∂
G = ∇×F = ∂x ∂y
∂z
F1 F2 F3
∂F3 ∂F2 ∂F3 ∂F1 ∂F2 ∂F1
= − i+ − j+ − k
∂y ∂z ∂x ∂z ∂x ∂y
Where we have again used the C 2 condition to ensure that the mixed second
derivatives commute.
110
VECTOR CALCULUS VECTORS AND SCALARS
111
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
the line integral of F along the curve γ. We will see below that this integral
depends upon the curve γ, but not upon the parameterisation of the curve - only
upon the set of points on the curve.
In the usual way when considering integrals we can break the curve γ up into a
series of m small straight line segments.
Let δrk denote the line segment joining rk−1 to rk , and let Fk = F(rk ). Then we
may define the line integral by
Z Xm
F·dr = lim Fj ·δrj
γ m→∞
j=1
However this definition is not very useful when it comes to calculations. In oder
to calculate a line integral we convert it to an ordinary 1-dimensional integral.
112
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
γ : t 7→ r(t), a≤t≤b
Then
dr
dr = dt
dt
We also look at the value of the vector field along the curve
F = F(r(t))
113
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
which agrees with the previous formula. Hence the line integral is independent of
the parameterisation of the curve as claimed.
Example
Z 1a: Let F = 2xyi + (x2 − z 2 )j − 3xz 2 k. Evaluate the line integral
F·dr, where γ is the curve given by
γ
(2,1,3)
z
g2 y
g1
114
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
dr
= 2i + 3t2 j + 6tk
dt
Hence
Z Z 1
dr
F·dr = F(r(t))· dt
γ t=0 dt
Z 1
= (4t4 i + (4t2 − 9t4 )j − 54t5 k)·(2i + 3t2 j + 6tk)dt
Zt=0
1
= (8t4 + 12t4 − 27t6 − 324t6 )dt
t=0
1
5 351
= 4t −
7 0
323
=−
7
Example 1b: Calculate the line integral for the bf same vector field F = 2xyi +
(x2 − z 2 )j − 3xz 2 k, but this time along the straight line connecting (0, 0, 0) to
(2, 1, 3).
This time the curve is given by
Hence
x = 2t, y = t, and z = 3t
substituting into the formula for F(r) this gives
dr
= 2i + j + 3k
dt
115
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
Hence
Z Z 1
dr
F·dr = F(r(t))· dt
γ t=0 dt
Z 1
= (4t2 i − 5t2 j − 54t3 k)·(2i + j + 3k)dt
Zt=0
1
= (6t8 + 36t8 + 12t8 )dt
t=0
1
54 9
= t
9 0
=6
This example shows the important fact that in general the value of the line integral
depends upon the path taken, not just the end points.
Example 2: Let F = ∇φ where φ = xy 2 z. Calculate the line integral of F for
both of the curves used in example 1.
F = ∇φ = y 2 zi + 2xyzj + xy 2 k
For part (a)
r(t) = 2ti + t3 j + 3t2 k, 0≤t≤1
So
F(r(t) = 3t8 i + 12t6 j + 2t7 k
and
dr
= 2i + 3t2 j + 6tk
dt
Hence
Z Z 1
dr
F·dr = F(r(t))· dt
γ t=0 dt
Z 1
= (3t8 i + 12t6 j + 2t7 k)·(2i + 3t2 j + 6tk)dt
Zt=0
1
= (6t8 + 36t8 + 12t8 )dt
t=0
1
54 9
= t
9 0
=6
116
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
So that in this example the two line integrals give the same answer. In fact this is
true for the line integral of any vector field that is the form gradφ.
Proposition:
1
Z vector field so that F = ∇φ, for some C scalar field φ, then
If F is a gradient
the line integral F·dr only depends upon the end points A and B of γ, not on
γ
the path taken. In fact
Z
(∇φ)·dr = φ(rB ) − φ(rA )
γ
where φ(rA ) and φ(rB ) are the values of φ at the start and end points of the curve
γ.
Proof: Let γ be parameterised by r(t), t0 ≤ t ≤ t1 then
Z Z t1
dr
(∇φ)·dr = ∇φ(r(t))· dt
γ t=t0 dt
Now
∂φ ∂φ ∂φ
∇φ = i+ j+ k
∂x ∂y ∂z
117
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
And if we write
then
dr dx dy dz
= i+ j+ k
dt dt dt dt
Hence
dr ∂φ dx ∂φ dy ∂φ dz
∇φ· = + +
dt ∂x dt ∂y dt ∂z dt
dφ(r(t))
=
dt
Thus
Z Z t1
dφ(r(t))
(∇φ)·dr = dt
γ t=t0 dt
= [φ(r(t))]tt=t
1
0
= φ(r(t1 )) − φ(r(t0 ))
= φ(rB ) − φ(rA )
W = F·r, where r = AB
However if the force and path are changing, we have to divide the path into small
sections, and then take the limit
Xm Z
W = lim Fj δrj = F(r)·dr
m→∞ γ
j=1
in order to obtain the formula for the work done by the vector field F(r) in moving
from A to B along the path γ.
If we can find a scalar field such that
118
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
then we say that φ is a potential for the force F and the change in φ simply
represents the change in potential energy. In such a situation any loss of potential
energy results in a gain in kinetic energy (and vice-versa) so that the total energy
is conserved. For this reason a gradient vector field is often called a conservative
vector field, and we will use this name from now onwards in this course. In section
4.2 of these notes we will look at conditions which guarantee that a vector field is
conservative.
8.2.3 Surfaces
We start by giving a number of different definitions of a 2-dimensional surface in
three dimensional space R3 . There are essentially three different ways of specify-
ing a surface
(1.) As the graph of a function of 2 variables, z = f (x, y)
(2.) As a level surface of a function of 3 variables F (x, y, z) = c
(3). As a parameterised surface (s, t) 7→ (x(s, t), y(s, t), z(s, t))
For the purpose of calculating surface integrals we will mostly use definition (3)
and think of surfaces as (smooth) maps
S → R3
(s, t) 7→ r(s, t)
where S ⊂ R2 is some 2-dimensional set which specifies the region in which the
parameters s and t lie in. In many of the applications this will simply involve s0 ≤
s ≤ s1 and t0 ≤ t ≤ t1 so that S will be given by the rectangle [s0 , s1 ] × [t0 , t1 ].
If we draw a picture of this map
s1 s2
s3
s4 t5 t6
t4
s5 t3
t2
s6
t1
119
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
we see that the lines s =const. and t =const. are simply the coordinate lines on
the surface.
Example 1: A cylinder radius a and height h
r(s, t) = a sin si + a cos sj + tk, −π ≤ s < π, 0 ≤ t ≤ h
n̂
The lines t =const. are circles and the lines s =const. are straight vertical lines
on the surface of the cylinder.
Example 2: A sphere radius a centre the origin
r(s, t) = a cos s sin ti + a sin s sin tj + a cos tk, −π ≤ s < π, 0 ≤ t ≤ π
ݏൌͲ
ݐൌͲ
ݐൌ ߨȀʹ
ݐൌߨ
120
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
The lines s =const. are the lines of longitude and the lines t =const. are the
lines of latitude on the surface of the sphere. Thus the s-coordinate is just the
φ-coordinate of spherical polar coordinates, while the t-coordinate is just the θ-
coordinate of spherical polar coordinates.
¶r
¶t
¶r
s1 ¶s
s2
t4
s3 t3
t2
s4
t1
121
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
Hence
i j k
n = −a sin s a cos s 0 = a cos si + a sin sj
0 0 1
Note this is simply the (i, j)-part of r(s, t) as one would expect from the picture
above.
Example 2: A sphere radius a centre the origin
r(s, t) = a cos s sin ti + a sin s sin tj + a cos tk
∂r
= −a sin s sin ti + a cos s sin tj + a cot tk
∂s
∂r
= a cos s cos ti + a sin s cos tj − a sin tk
∂t
Hence
i j k
n = −a sin s sin t a cos s sin t a cot t
a cos s cos t a sin s cos t −a sin t
= −a2 sin t(cos s sin ti + sin s sin tj + cos tk)
= −a2 sin tr
Thus n̂ = r̂ (as one would expect!).
dr2
dA
dr1
s1
s2
t4
s3 t3
t2
s4
t1
122
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
Suppose we start at the point r(s, t) and move to the nearby point r(s + δs, t).
Then the vector connecting these two points is
δr1 = r(s + δs, t) − r(s, t)
Now by Taylor’s theorem
∂r
r(s + δs, t) = r(s, t) + δs (s, t) + O(δs2 )
∂s
Hence
∂r
δr1 = δs (s, t) + O(δs2 )
∂s
Similarly if we move in the t direction to the nearby point r(s, t + δt) then the
vector measuring the displacement from r(s, t) is
δr2 = r(s, t + δt) − r(s, t)
Again using Taylor’s theorem we have
∂r
r(s, t + δt) = r(s, t) + δt (s, t) + O(δt2 )
∂t
and hence
∂r
δr2 = δt (s, t) + O(δt2 )
∂t
Thus the area of the parallelogram with sides δr1 and δr2 is
δA = |δr1 ×δr2 |
∂r ∂r
≈ × δsδt
∂s ∂s
123
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
and hence
∂r ∂r
× = a(cos2 s + sin2 s)1/2
∂s ∂s
=a
dA = adsdt
Hence we have shown that the surface area of a cylinder of height h and radius a
is 2πah (in agreement with the standard formula).
Example 2: We now use the formula for dA to calculate the surface area of a
sphere. Recall that a parameterised sphere is given by
So that
i j k
∂r ∂r
× = −a sin s sin t a cos s sin t a cot t
∂s ∂s
a cos s cos t a sin s cos t −a sin t
= −a2 sin t(cos s sin ti + sin s sin tj + cos tk)
and hence
∂r ∂r
× = a2 sin t cos2 s sin2 t + sin2 s sin2 t + cos2 t 1/2
∂s ∂s
1/2
= a2 sin t (cos2 s + sin2 s) sin2 t + cos2 t
= a2 sin t
dA = a2 sin tdsdt
124
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
Note that a is the radius of the sphere and (s, t) correspond to the usual spherical
polar coordinates (φ, θ), so that the above formula agrees with the standard result
that
dA = r2 sin θdθdφ
We now use this to calculate the area of a sphere
Z π Z π
A= a2 sin tdsdt
s=−π t=0
Z π
2
= 2πa sin tdt
t=0
= 2πa [− cos t]πt=0
2
= 4πa2
Hence we have shown that the area of a sphere radius a is 4πa2 (in agreement with
the standard formula).
dS = n̂dA
125
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
Then dS points in the same direction as n and has modulus |dS| = dA. From
this we see that
∂r ∂r
dS = × dsdt
∂s ∂t
This is because the above expression points in the same direction as n, since
∂r ∂r
n= ×
∂s ∂t
and has modulus dA since
∂r ∂r
dA = ×
∂s ∂t
Using this formula for dS we may write the flux integral as
ZZ ZZ
∂r ∂r
F·dS = F(r(s, t))· × dsdt
∂s ∂t
S S
126
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
(a twisted ribbon or “helicoid”). Let F(r) ZbeZ the vector field given by F = xi +
yj + (z − 2y)k. Evaluate the flux integral F·dS. We first evaluate the vector
S
field on the surface
r(s, t) = s cos ti + s sin tj + tk
Now on the surface
127
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
∂r
= cos ti + sin tj + 0k
∂s
∂r
= −s sin ti + s cos tj + k
∂t
Now
∂r ∂r
dS = × dsdt
∂s ∂t
i j k
= cos t sin t 0 dsdt
−s sin t s cos t 1
= sin ti + cos tj + (s cos2 t + s sin2 t)k dsdt
Hence
ZZ Z 1 Z 2π
F·dS = (s cos ti + s sin tj + (t − 2s sin t)k)·(sin ti − cos tj + sk)dsdt
s=0 t=0
S
Z 1 Z 2π
= (st − 2s2 sin t)dsdt
s=0 t=0
Z 1 2π
1 2 2
= st + 2s cos t ds
s=0 2 t=0
Z 1
= 2sπ 2 ds
s=0
2 2 1
= s π s=0
= π2
Hence ZZ
F·dS = π 2
S
We end this section by giving the expression for a flux integral in the special case
that the surface is the graph of a function z = f (x, y). To calculate the flux integral
in this case we take x and y themselves to be the paramters used to describe the
surface. Hence the parameterised surface is given by
128
VECTOR CALCULUS INTEGRALS OF VECTOR AND SCALAR FIELDS
Then
∂r ∂f
= i + 0j +
∂x ∂x
∂r ∂f
= 0i + j +
∂y ∂y
Hence
∂r ∂r
dS = × dxdy
∂x ∂y
i j k
1 0 ∂f
= ∂x dxdy
∂f
0 1 ∂y
∂f ∂f
= − i− j + k dxdy
∂x ∂y
Thus for the case of a suface which is the graph of a function we have the formula
ZZ ZZ
∂f ∂f
F·dS = F(x, y, f (x, y)· − i − j + k dxdy
∂x ∂y
S
As a final remark in this section we point out that it often saves time to think a bit
about the geometry of a flux integral before going ahead and mechanically doing
the calculation.
2 2 2 2
Example: Let F = xi+yj Z+zk,
Z and S the surface of the sphere x +y +z = a .
Calculate the flux integral F·dS.
S
In this example a little thought shows that F = r and that dS = n̂dA = r̂dA.
Hence
F·dS = r·r̂dA = r(r̂)·(r̂)dA = rdA
Hence ZZ ZZ ZZ
F·dS = rdA = a dA = a.4πa2 = 4πa3
S S S
(since r = a on S).
129
VECTOR CALCULUS INTEGRAL THEOREMS
n̂
¶S
Note 2: In the picture above the boundary to S consists of just one component.
However the theorem allows for the boundary to consist of several components
(e.g. three as shown in the picture below).
130
VECTOR CALCULUS INTEGRAL THEOREMS
n̂
¶S2
¶S3
¶S1
In such a case the line integral is the sum of the line integral for the various com-
ponents, so in the example above
I I I I
F·dr = F·dr + F·dr + F·dr
∂S ∂S1 ∂S2 ∂S3
where the direction of integration round all three loops ∂Si is clockwise relative
to the normal n to S.
Note 3: Stokes’s Theorem contains some fine print concerning the nature of the
surface “piecewise smooth, orientable”, the nature of the boundary “finite number
of piecewise C 1 simple closed curves” and concerning the nature of the vector
field F “C 1 on domain that includes S”. All these restrictions are important and
the theorem is not valid without them (or equivalent) conditions. For further ex-
planation on the precise meaning of these terms see the Appendix.
Example: Let S be the paraboloid
z = 9 − x2 − y 2 , z≥0
(0,0,9)
z=9- x2-y2
(0,3,0)
(-3,0,0) (3,0,0)
(0,-3,0)
S: x2+y2=9
131
VECTOR CALCULUS INTEGRAL THEOREMS
r(x, y) = xi + yj + (9 − x2 − y 2 )k, x2 + y 2 ≤ 9
Then
∂r ∂r
dS = × dxdy
∂x ∂y
i j k
= 1 0 −2x dxdy
0 1 −2y
= (2xi + 2yj + k) dxdy
Hence
ZZ Z Z
(∇×F)·dS = (−3i − j + 2k)·(2xi + 2yj + k)dxdy
S x2 +y 2 ≤9
Z Z
= (−6x − 2y + 2)dxdy
x2 +y 2 ≤9
Z Z
= 2dxdy, (since the other terms cancel by symmetry)
x2 +y 2 ≤9
To verify Stokes’s theorem we now calculate the line integral. From the diagram
we see that C = ∂S is a circle radius 3 centre the origin and lying in the xy-plane..
Since the normal n to S point upwards we parameterise C as:
132
VECTOR CALCULUS INTEGRAL THEOREMS
Then
F(r(t)) = −3 sin ti + 3 cos tj + (9 cos t − 6 sin t)k
and
dr
= −3 sin ti + 3 cos tj + 0k
dt
Thus
I Z 2π
dr(t)
F·dr = F(r(t))· dt
∂S t=0 dt
Z 2π
= ((−3 sin ti + 3 cos tj + (9 cos t − 6 sin t)k)·(−3 sin ti + 3 cos tj + 0k)) dt
t=0
Z 2π
= (9 sin2 t + 9 cos2 t)dt
Zt=0
2π
= 9dt
t=0
= [9t]2π
t=0 = 18π
But S and S̃ both have the same boundary given by C, a circle radius 3 in the
xy-plane. So that ∂ S̃ = ∂S. Hence
ZZ I I
(∇×F) ·dS = F·dr = F·dr = 18π
∂ S̃ ∂S
S̃
The method used in this example is often useful so we state it in the general case
as a consequence of Stokes’s theorem.
133
VECTOR CALCULUS INTEGRAL THEOREMS
for any 2 piecewise C 1 , oriented curves γ1 and γ2 which lie in the set U and both
start at the same point A, and end at both end at the same point B.
Proposition: Let F be a continuous vector field. Then F has path-independent
integrals in U if and only if I
F·dr = 0
γ
B
g1
g2
g3
g4
A
134
VECTOR CALCULUS INTEGRAL THEOREMS
Conversely suppose the line integral round any closed loop vanishes and γ1 and
γ2 are two curves that start and finish at the same point. Then γ1 followed by −γ2
(ie γ2 traversed in the opposite direction) is just γ so that
Z Z I
F·dr − F·dr = F·dr = 0
γ1 γ2 γ
where γ is some path from some (arbitrary) fixed point A to the point P with po-
sition vector r. Note that because the vector field F has path independent integral
this is well defined without having to specify the path γ.
We now show by direct computation that
∂φ ∂φ ∂φ
= F1 , = F1 , = F1 , ie ∇φ = F.
∂x ∂x ∂x
Let Z
φ(x, y, z) = F·dr
γ
Then Z Z
φ(x + h, y, z) = F·dr + F·dr
γ γ1
135
VECTOR CALCULUS INTEGRAL THEOREMS
Hence
∂φ φ(x + h, y, z) − φ(x, y, z)
= lim = lim (F1 (x, y, z) + O(h)) = F1 (x, y, z)
∂x h→0 h h→0
Similarly
∂φ ∂φ
= F2 , and = F3
∂y ∂z
Note that different choices of the arbitrary initial point A simply change φ by a
constant, but this has no effect on ∇φ which is unchanged.
The above theorem tells us that a conservative vector field is equivalent to one
with path-independent integral. Unfortunately this condition is not very easy to
check as we need to know it is true for any possible path! The following theorem
gives a more useful criterion, which is easy to check.
Theorem: Let F be a C 1 vector field defined on the simply connected region U
in R3 . Then
There exists some C 2 scalar field φ such that F = ∇ if and only if ∇×F = 0
Note: The fine print in this theorem is again important to note. In particular the
simply connected condition (which means we can continuously shrink any closed
path down to a point) is required - see the exercise sheets for a counterexample to
this theorem when U is not simply connected.
Proof:
We know from chapter 2 that ∇×(∇φ) = 0, so if F = ∇φ then we know that
∇×F = 0.
On the other hand if ∇×F = 0 then we know from Stokes’s theorem that
I ZZ
F·dr = (∇×F)·dS = 0
C
S
136
VECTOR CALCULUS INTEGRAL THEOREMS
where S is some spanning surface with boundary C. (Note that such a spanning
surface S exists precisely because of the simply-connected condition.)
Hence by the first proposition in this section F has path-independent integral, and
thus by the previous theorem F = ∇φ, as required.
f 0 (y) = 2y ⇒ f (y) = y 2 + c
137
VECTOR CALCULUS INTEGRAL THEOREMS
φ(x, y) = x2 y + x cos 2y + y 2 + c
In this section we look at this in the case that the vector field F depends only
upon x and y and both it and the surface S lie in the xy-plane. Since F lies in the
xy-plane and only depends upon x and y, then
138
VECTOR CALCULUS INTEGRAL THEOREMS
Then
i j k
∂ ∂ ∂
∇×F = ∂x ∂y ∂z
F1 (x, y) F2 (x, y) 0
∂F2 ∂F1
= 0i + 0j + − k
∂x ∂y
∂F2 ∂F1
= − k
∂x ∂y
dS = n̂dA = kdxdy
Hence ZZ ZZ
∂F2 ∂F1
(∇×F) ·dS = − dxdy
∂x ∂y
S S
r = xi + yj ⇒ dr = dxi + dyj
So that I I
F·dr = (F1 dx + F2 dy)
∂S ∂S
Hence in the xy-plane Stokes’s theorem becomes
ZZ I
∂F2 ∂F1
− dxdy = (F1 dx + F2 dy)
∂x ∂y ∂S
S
We have seen that one can obtain Green’s theorem from Stokes’s theorem by
restricting things to be in a 2-dimensional space. However it is possible to derive
139
VECTOR CALCULUS INTEGRAL THEOREMS
Stokes’ theorem from Green’s theorem. This follows from the fact that if one
parameterises the surface S in Stokes’s theorem in terms of s and t by
∂r ∂r
where Fs = F· and Ft = F· . It then follows from Green’s theorem in the
∂s ∂t
st-plane that
ZZ I
∂Ft ∂Fs
− dudv = (Fs ds + Ft dt)
∂s ∂t ∂ S̃
S̃
= F·dr
∂S
Hence we have used Green’s theorem in the st-plane to prove Stokes’s theorem:
ZZ I
(∇×F) ·dS = F·dr
∂S
S
140
VECTOR CALCULUS INTEGRAL THEOREMS
(x,y+)
C
+
¶S=C+-C-
C-
(x,y-)
Let C + be the upper curve and C − be the lower curve in the above diagram. Let
(x, y + (x)) denote points on the graph of the upper curve and (x, y − (x)) denote
points on the graph of the lower curve. Then if we traverse the closed curve in a
clockwise direction we move from left to right on the bottom curve and from right
to left on the upper curve. Hence
Z Z b
F1 dx = F1 (x, y − (x))dx
−
ZC Zx=a
a
F1 dx = F1 (x, y + (x))dx
C+ x=b
Z b
=− F1 (x, y + (x))dx
x=a
141
VECTOR CALCULUS INTEGRAL THEOREMS
Thus
I Z Z
F1 dx = F1 dx + F1 dx
C C− C+
Z b
F1 (x, y + (x)) − F1 (x, y − (x)) dx
=−
x=a
Zb
y + (x)
=− [F1 (x, y)]y=y− (x) dx
x=a
Z (Z y+ (x) )
∂F1
=− dy dx
y=y − (x) ∂y
Hence I ZZ
∂F1
F1 dx = − dxdy
C ∂y
S
Similarly I ZZ
∂F2
F2 dy = dxdy
C ∂x
S
Hence I I ZZ
∂F2 ∂F1
F1 dx + F2 dy = − dxdy dxdy
C C ∂x ∂y
S
as required.
142
VECTOR CALCULUS INTEGRAL THEOREMS
{
where the symbol is used to indicate that the integral is over a closed surface.
S1 a
S3
h
S2
The hard way is to calculate the flux integral directly. To do this we observe that
S consists of three surfaces: the top S1 , the bottom S2 and the curved surface S3
so that { x x x
F·dS = F·dS + F·dS + F·dS
S S1 S2 S3
143
VECTOR CALCULUS INTEGRAL THEOREMS
Hence {
F = F·dS = πa2 h + 0 + 2πa2 h = 3πa2 h
S
The easy way is to use the divergence theorem
{ y
F = F·dS = (∇·F)dv
S V
Now
∂F1 ∂F2 ∂F3
∇·F = + +
∂x ∂y ∂z
=1+1+1=3
Hence
{ y y
F = F·dS = (∇·F)dv = 3dv = 3.volume of V = 3πa2 h
S V V
(0,a,0)
~
S
(-a,0,0) (a,0,0)
(0,-a,0) ~
S : x2 + y2 £ a2; z = 0
S : x2 + y2 + z 2 = a2
144
VECTOR CALCULUS INTEGRAL THEOREMS
Since S is not a closed surface we cannot immediately apply the divergence theo-
rem. However if we add on the disc S̃ given by x2 +y 2 ≤ a2 , z = 0 (see diagram),
then S ∪ S̃ is a closed surface. Hence
x x y
F·dS + F·dS = (∇·F)dv (∗)
S S̃ V
y Z a
1
2
2y dydxdz = 2y 2 π(a2 − y 2 )dy
−a 2
V
Z a
=π (a2 y 2 − y 4 )dy
−a
4π 5
= a
15
Now x x x
F·dS = F·kdA = x2 ydxdy = 0 (by symmetry)
S̃ S̃ S̃
Hence substituting into (*) gives
x x y
F·dS + F·dS = (∇·F)dv
S S̃ V
x 4π 5
F·dS + 0 = a
15
S
Hence x 4π 5
F·dS = a
15
S
145
VECTOR CALCULUS INTEGRAL THEOREMS
∇·(∇×F) = 0
G = ∇×F ⇒ ∇·G = 0
However in a simply connected region one can also show (although we do not
give the proof in this course) that
where S1 and S2 are two open surfaces with the same boundary C = ∂Si . Note,
that we have already come across this result when we considered Stokes’s theorem
in section 4.1.
146