Elijah's Math Notes
Elijah's Math Notes
These notes are intended to give a brief outline of the course to be used as an
aid in learning. They are not intended to be a replacement for attendance at
lectures, problem classes or tutorials. In particular, they contain few exam-
ples. Since examination questions in this course consist mainly of examples,
you will seriously compromise your chances of passing by not attending lec-
tures, problem classes and tutorials where many examples will be worked out
in detail.
2015
c School of Mathematics and Statistics, UNSW
1
TOPIC 1 – PARTIAL DIFFERENTIATION
Partial derivatives are the derivatives we obtain when we hold constant all but
one of the independent variables in a function and differentiate with respect
to that variable.
Functions of Two Variables
Suppose z = f (x, y). Define
∂f
= lim f (x+∆x,y)−f
∆x
(x,y)
∂x ∆x−→0
∂f
= lim f (x,y+∆y)−f
∆y
(x,y)
∂y ∆y−→0
These are both functions of x and y and the usual differentiation rules (prod-
uct, quotient etc) apply.
Notation
∂f ∂f
= fx = zx , = fy = zy
∂x ∂y
i.e. subscripts are used to denote differentiation with respect to the indicated
variable. Further
∂f ∂f
(x0 , y0 ) = fx (x0 , y0 ) means evaluated at the point (x0 , y0 ).
∂x ∂x
Higher-Order Derivatives
∂ 2f
!
∂ ∂f
= = (fx )x = fxx
∂x2 ∂x ∂x
∂ 2f
!
∂ ∂f
= = (fy )x = fyx
∂x∂y ∂x ∂y
∂ 2f
!
∂ ∂f
= = (fx )y = fxy
∂y∂x ∂y ∂x
∂ 2f
!
∂ ∂f
= = (fy )y = fyy
∂y 2 ∂y ∂y
2
This readily extends to higher order derivatives. In particular, if all deriva-
∂ n+m f
tives are continuous then ∂x n ∂y m can be used to denote the partial derivative
3
This extends to functions of 2 or more variables. We consider only f (x, y).
The Taylor Series of f (x, y) about the point (a, b) is
∂f ∂f
f (x, y) = f (a, b) + (x − a) (a, b) + (y − b) (a, b)
∂x ∂y
∂ 2f ∂ 2f
(
1
+ (x − a)2 2 (a, b) + 2(x − a)(y − b) (a, b)
2! ∂x ∂x∂y
∂ 2f
)
2
+(y − b) (a, b) + higher-order terms.
∂y 2
Standard Linear Approximation
If y = f (x) then a reasonable approximation when x is close to x0 is
obtained by truncating the Taylor series after the linear term. Geometrically,
we are approximating the curve y = f (x) for x near x0 by the tangent to the
curve at (x0 , f (x0 )).
This idea readily extends to functions of two or more variables. All we
do is truncate the Taylor Series after the linear terms. The standard linear
approximation of f (x, y) near (x0 , y0 ) is therefore f (x, y) ' L(x, y) where
∂f ∂f
∆f ' ∆x + ∆y.
∂x ∂y
∂f ∂f
where ,
∂x ∂y
are evaluated at (x0 , y0 ).
4
If ∆x and ∆y are known, we just substitute them in. Usually, however, all
we know are bounds on ∆x and ∆y. For example, we may only be able to
measure temperature to ±0.01◦ C. In that case we have, approximately,
∂f ∂f
|∆f | ≤ |∆x| + |∆y|.
∂x ∂y
and n
∂f
X
|∆f | ≤ |∆xk |.
∂xk
k=1
Leibniz Rule
d Z v(x) Z v(x)
∂f dv du
f (x, t)dt = dt + f (x, v(x)) − f (x, u(x)) .
dx u(x) u(x) ∂x dx dx
fx (x0 , y0 ) = fy (x0 , y0 ) = 0,
5
(ii) critical points of f
g(x, y, z) = 2x + y − z − 5 = 0
6
In this simple case, it is easy to use the constraint equation to find an explicit
expression for one of the variables (say z) in terms of the other two and to
then substitute this into f which thus becomes a function of two variables
only and then to find the extrema of f as a function of x and y. For a more
complicated constraint, it may not be possible to use the constraint equation
to obtain an explicit expression for one of the variables in terms of the others
so a more general procedure is required.
The method of Lagrange multipliers
To start off, suppose that f (x, y) and g(x, y) and their first partial deriva-
tives are continuous. To find the local minima and maxima of f subject to
the constraint g(x, y) = 0 we find the values of x, y and λ that simultaneously
satisfy the equations
∂f ∂g ∂f ∂g
−λ = 0, −λ = 0, together with g(x, y) = 0 (1)
∂x ∂x ∂y ∂y
Justification: We can, in principle, use the equation g(x, y) = 0 to write y
as a function of x although, as indicated above, this may not be possible in
practice. Hence, we may consider f to be a function of a single variable x and
look for points where df /dx = 0. Let (x, y) = (a, b) be such a point. But, by
the chain rule
df ∂f d x ∂f d y ∂f ∂f d y
= + = +
dx ∂x d x ∂y d x ∂x ∂y d x
Thus
∂f ∂f d y
+ =0 at (x, y) = (a, b) (2)
∂x ∂y d x
However, since g(x, y) = 0, dg/dx = 0 everywhere (including (a, b)). Thus
∂g ∂g d y
+ =0 at (x, y) = (a, b) (3)
∂x ∂y d x
Thus, eliminating dy/dx from (2) and (3) we obtain
∂f ∂g ∂f ∂g
− =0 at (x, y) = (a, b)
∂x ∂y ∂y ∂x
which can also be written
∂f ∂f
∂x ∂y
∂g ∂g =0 at (x, y) = (a, b)
∂x ∂y
Hence, the rows of this determinant must be linearly dependent, Thus there
exists a real number λ such that
! !
∂f ∂f ∂g ∂g
, =λ ,
∂x ∂y ∂x ∂y
7
N. B. The quantity λ is called a Lagrange multiplier and the method also
works if f and g are also functions of z. In that case we have the additional
equation ∂f /∂z = λ∂g/∂z to solve. It is also possible to introduce the so-
called Lagrangian function
The equations (1) and the constraint g(x, y) = 0 are obtained by setting to
zero the first partial derivatives of L(x, y, λ) with respect to x, y and λ.
Lagrange multipliers with two constraints
Suppose we now want to find the maxima and minima of f (x, y, z) subject
to
g1 (x, y, z) = 0 and g2 (x, y, z) = 0.
To do this, we introduce two Lagrange multipliers (one for each constraint)
and the Lagrangian function for this situation
a1 = q 1 − p 1 , a 2 = q 2 − p 2 , a 3 = q 3 − p 3
−→ −→ −→
and
q a = P Q= a 1 i + a 2 j + a 3 k = OQ − OP . The length of a is |a| =
2 2 2
a1 + a2 + a3 . The position vector of a typical point with coordinates (x, y, z)
is usually written r = xi + yj + zk.
Addition etc.
Define 0 = 0i + 0j + 0k. This is the vector all of whose components are zero,
and is not to be confused with the scalar 0. All the usual rules apply, for
example
8
a+0 = a
ca = ca1 i + ca2 j + ca3 k
−a = (−1)a = −a1 i − a2 j − a3 k
a+b = b+a
(a + b) + c = a + (b + c) = a + b + c
a + (−a) = 0.
c(a + b) = ca + cb
Inner or Dot or Scalar Product of Vectors
a · b = a1 b1 + a2 b2 + a3 b3 = |a||b| cos γ
where γ (0 ≤ γ ≤ π) is the angle between a and b.
Then a · a = |a|2 and the dot product of two (non-zero) vectors is 0 if and
only if they are orthogonal (γ = π2 ).
Observe that i · i = j · j = k · k = 1 and
a·b a1 b 1 + a2 b 2 + a3 b 3
cos γ = =q q
|a||b| a22 + a22 + a23 b21 + b22 + b23
The component of a vector a in the direction of b (otherwise known as the
projection of a onto b) is
|a|a · b a·b
p = |a| cos γ = = .
|a||b| |b|
Vector or Cross Product of Vectors
v = a × b is a vector whose magnitude is |v| = |a||b| sin γ (where γ is the
angle (0 ≤ γ ≤ π)) between a and b. The vector v is perpendicular to the
plane defined by a and b, in such a way that a right-handed screw turn in
the direction of v turns a into b through an angle of less than π.
Properties
a × b = −b × a
a × a = 0
i j k
a × b = a1 a2 a3
b 1 b2 b3
9
Also, |a · (b × c)| is the volume of the parallelepiped defined by a, b and c.
Scalar and Vector Fields
Consider some region Ω of 3-dimensional space. Let a typical point in Ω have
coordinates (x, y, z). A scalar field is a scalar quantity f (x, y, z) defined on
Ω. It often depends on time t as well. The temperature or density in the
atmosphere are examples of scalar fields.
A vector field is a vector each of whose components is a scalar field. Thus
v = v1 i + v2 j + v3 k
10
The speed is |v| = (v · v)1/2 and the acceleration is
dv d2 r
a= = 2.
dt dt
Gradient of a Scalar Field
∂φ ∂φ ∂φ
∇φ = grad φ = i+ j+ k.
∂x ∂y ∂z
Directional Derivative
Consider a scalar field φ. What is the change in φ as we move from P (x, y, z)
to Q(x + ∆x, y + ∆y, z + ∆z) keeping t constant?
If ∆s is the distance from P to Q then
Hence, letting ∆s −→ 0,
dφ ∂φ dx ∂φ dy ∂φ dz
= + +
ds ∂x ds ∂y ds ∂z ds
= ∇φ · u
b
Now, the rate of change with respect to distance in the direction specified
by the unit vector u
b is called the directional derivative and is denoted by
Dub φ. We have shown that
b φ = ∇φ · u
Du b.
11
Normal to a Surface
Next, consider a level surface φ = C. This defines a surface S in space. For
example, meteorologists talk about surfaces of constant pressure such as the
500 millibar surface. Let P and Q be any two nearby points on S. Then
φP = φQ = C, i.e. dφ ds
= 0 at P in any direction tangential to S at P. Thus
−→
∇φ at P is orthogonal to P Q.
Since this holds for any point Q close to P, i.e. is independent of the direction
from P to Q, it follows that ∇φ at P must be orthogonal to the level surface
∇φ
φ = C. A unit normal is |∇φ| .
Equation of Tangent Plane
If P has coordinates (x0 , y0 , z0 ) and (x, y, z) is any point in the plane tan-
gent to S at P then ∇φ is normal to this tangent plane which therefore has
equation
∇φ · [(x − x0 )i + (y − y0 )j + (z − z0 )k] = 0
where ∇φ is evaluated at P.
Divergence of a Vector Field
If F = F1 i+F2 j+F3 k then ∇·F = div F = ∂F ∂x
1
+ ∂F
∂y
2
+ ∂F
∂z
3
. It may be regarded
∂ ∂ ∂
at the dot product of the vector differential operator ∇ = i ∂x + j ∂y + k ∂z and
the vector F. It is just a scalar.
N.B. ∇ · F 6= F · ∇. The latter is the differential operator
∂ ∂ ∂
F · ∇ = F1 + F2 + F3
∂x ∂y ∂z
12
! ! !
∂F3 ∂F2 ∂F1 ∂F3 ∂F2 ∂F1
=i − +j − +k −
∂y ∂z ∂z ∂x ∂x ∂y
Theorem ∇ × (∇φ) = 0.
Proof
i j k
∂ ∂ ∂
L.H.S =
∂x ∂y ∂z
∂φ ∂φ ∂φ
∂x ∂y ∂z
∂ 2φ ∂ 2φ ∂ 2φ ∂ 2φ ∂ 2φ ∂ 2φ
! ! !
=i − +j − +k −
∂y∂z ∂z∂y ∂z∂x ∂x∂z ∂x∂y ∂y∂x
= 0i + 0j + 0k = 0 = R.H.S.
Vector fields F for which ∇ × F = 0 are called irrotational or conservative
.
Line Integrals
These are used for calculating, for example, the work done in moving a particle
in a force field.
Consider a vector field F(r) and a curve C from point A to point B. Let the
equation of C be r = r(t) where t is parameter. Let t = a at point A and
t = b at point B. We define
Z b
Z
dr
F(r) · dr = F(r(t)) · dt
C a dt
In terms of components, this can be written
Z b !
Z
dx dy dz
(F1 dx + F2 dy + F3 dz) = F1 + F2 + F3 dt
C a dt dt dt
where dx, dy, dz are displacements measured along C and F is evaluated on C.
In general, this integral depends not only on F but also on the path we take
between A and B. If A and B coincide, we are integrating around a closed
curve. This is denoted by I
F · dr.
C
F · dr is the work done in moving from A to B
R
Work If F is a force field, C
along C.
Simple properties
· dr = k F · dr
R R
(i) If k is a constant C (kF) C
+ G) · dr = F · dr + G · dr.
R R R
(ii) C (F C C
F · dr = − F · dr.
R R
(iii) C1 C
where C1 is the same curve as C except that we start at B and finish
at A, i.e. reversing the order of integration changes the sign of a line
integral.
13
F · dr = F · dr + F · dr
R R R
(iv) C C1 C2
Now let N −→ ∞ in such a way that the largest linear dimension of each
∆Ωj −→ 0 as N −→ ∞. Then if IN tends to some limit as N −→ ∞ we
define Z
I = f (x, y)dA = lim IN .
N −→∞
Ω
14
Simple Properties
1. If k is a constant
Z Z
kf (x, y)dA = k f (x, y)dA.
Ω Ω
R R R
2. (f (x, y) + g(x, y))dA = f (x, y)dA + g(x, y)dA.
Ω Ω Ω
3. If f (x, y) ≥ 0 on Ω then
Z
f (x, y)dA ≥ 0
Ω
15
2. If Ω is defined by c ≤ y ≤ d, g1 (y) ≤ x ≤ g2 (y) then
Z Z d Z g2 (y)
f (x, y)dA = f (x, y)dx dy
c g1 (y)
Ω
∆m
δ(x, y) = lim .
∆A−→0 ∆A
Consider a small element dΩ with area dA located at (x, y). Its distance from
the y axis is x and its distance from the x axis is y. The first moment of the
lamina about the y axis is
Z
My = xδ(x, y)dA
Ω
16
The centre of mass has coordinates (xm , ym ) defined by
My Mx
xm = ym =
M M
If δ is constant, it will cancel and the centre of mass then coincides with the
centroid or centre of area of Ω which has coordinates (x, y) defined by
R R
xdA ydA
x = ΩR y = ΩR
dA dA
Ω Ω
Moments of Inertia
The moments of inertia of the above lamina about the x and y axes are
Ix = y 2 δ(x, y)dA
R
Ω
Iy = x2 δ(x, y)dA.
R
Ω
Polar Coordinates
If the region Ω is easily described using polar coordinates (x = r cos θ, y =
r sin θ) it is often better to evaluate the double integral in polar coordinates.
The main task is to express dA in polar coordinates.
If r increases by dr and θ by dθ, the little element of area so generated is
dA = (rdθ) × dr = r dr dθ
so
Z Z Z
f (x, y)dA = f (r cos θ, r sin θ)r dr dθ
Ω Ω
Z Z
= f (r cos θ, r sin θ)rd θ dr
Ω
In either form the r and θ limits are chosen to cover the region Ω. This will
be illustrated by examples in lectures. It is essential to draw a diagram of Ω.
Jacobian Transformation
Rb
The evaluation of a f (x)dx is often facilitated by the substitution x = x(u)
to give Z b Z β
dx
f (x)dx = f (x(u)) du
a α du
17
where x(α) = a, x(β) = b. For double integrals, if x = x(u, v), y = y(u, v)
then Z Z Z Z
f (x, y)dx dy = f (x(u, v), y(u, v)) |J|du dv
Ω Ω∗
This is an implicit relation between y and x and involves only one integration
constant. If we were given some initial condition such as y(0) = 1, we would
impose it now to find c.
2. Linear
The general first-order linear o.d.e. is
dy
+ P (x)y = Q(x).
dx
18
R
Multiplication by the integrating factor I(x) = exp{ P (x)dx} reduces this
to
d
(Iy) = IQ
dx
which can immediately be solved by integration to give
R
( IQdx + c)
y= .
I
y = α1 y1 (x) + α2 y2 (x)
1. a2 > 4b
Then λ1 and λ2 are both real and the general solution is
y = α1 eλ1 x + α2 eλ2 x .
y = α1 eλ1 x + α2 xeλ1 x
= (α1 + α2 x)e−ax/2
19
3. a2 < 4b. Then the roots are complex conjugates
1 1
λ1 = − a + iw, λ2 = − a − iw
2 2
where
1√ q
w= 4b − a2 = b − a2 /4 .
2
Then
1 1
y = α1 e(− 2 a+iw)x + α2 e(− 2 a−iw)x
= e−ax/2 {α1 eiwx + α2 e−iwx }
= e−ax/2 {β1 cos wx + β2 sin wx}
Free Oscillations
Several systems (e.g. a mass oscillating at the end of a spring or an electrical
circuit) may be described by the d.e.
my 00 + cy 0 + ky = 0
mλ2 + cλ + k = 0,
y = (A + Bt)e−ct/2m .
20
√
3. c2 < 4mk. This is called “underdamping” as c is smaller than 2 mk.
Then
λ1 = −α + iΩ, λ2 = −α − iΩ
q
c k c2
where α= 2m
, Ω= m
− 4m2
Thus
y = (A cos Ωt + B sin Ωt)e−αt
= Re−αt cos(Ωt − δ)
√
where R = A2 + B 2 and tan δ = B/A. This represents decaying
qthe idealised case c = 0 (no friction), y = R cos(Ω0 t − δ)
oscillations. In
where Ω0 = k/m. This has period Ω2π0 . In reality, c > 0 and these
oscillations are killed off by friction.
21
Term in r(x) choice of yP
keσx Ceσx
Pn (x) Qn (x)
k cos θx
or k sin θx K cos θx + M sin θx
keσx cos θx or keσx sin θx σx
e (K cos θx + M sin θx)
Pn (x)eσx cos θx
or Pn (x)eσx sin θx eσx (Qn (x) cos θx + Rn (x) sin θx)
1. If a term in r(x) appears in the first column, choose the yP from the
corresponding second column and determine the unknown coefficients
by substituting into (6). In the table, Pn , Qn and Rn are polynomials
of degree n. Even if Pn (x) has only one term (xn ), Qn and Rn will in
general have all (n + 1) terms, i.e. Qn (x) = nj=0 qj xj .
P
3. If r(x) is a sum of functions, choose for yP (x) the sum of the appropriate
terms in the table.
Forced Oscillations
When simple periodic forcing is added to the mechanical or electrical
system studied earlier, we have to solve an equation like
my 00 + cy 0 + ky = F0 cos wt (8)
yP = a cos wt + b sin wt
k
w02 = and ∆ = m2 (w02 − w2 )2 + w2 c2 .
m
Thus
y = Θ cos(wt − δ)
22
where
F02
Θ2 = a2 + b2 =
∆
and
b wc
tan δ = = 2
.
a (m(w0 − w2 ))
Undamped Oscillations (c = 0)
In the ideal case of no mechanical friction or electrical resistance, the complete
solution of (8) is
F
y = A cos w0 t + B sin w0 t + cos wt
m(w02 − w2 )
q
k
for w 6= w0 . This is the sum of free oscillations with frequency w0 = m and
a forced oscillation with frequency w.
As w −→ w0 , the amplitude of the latter gets larger and larger. For
w = w0 , the above is not valid and the appropriate solution (using rule 2
above) is
F0
y = A cos w0 t + B sin w0 t + t sin w0 t
2mw0
The forced response thus consists of a sinusoid with linearly increasing ampli-
tude. This is the phenomenon of resonance and occurs when the frequency
w of the forcing is exactly equal to the natural frequency w0 with which the
system likes to oscillate.
Effects of Friction
In reality, c > 0 so the above is modified by friction or electrical resistance.
As we showed earlier, all solutions of the homogeneous equation decay to zero
as t −→ ∞ when c > 0. These are called the transients of the system. Thus
we are ultimately left with the directly forced response
yP = Θ cos(wt − δ)
q
where Θ
F0
=M = √ 1
and tan δ = wc
m(w02 −w2 )
and w0 = k
m
is
m2 (w02 −w2 )2 +w2 c2
the natural frequency of the undamped (c = 0) oscillations. The quantity M
is the magnification ratio for the amplitude. In engineering design we might
want this to be small so as to avoid damaging resonances or to be large to
magnify weak signals (e.g. tuning an AM radio). This will be discussed in
lectures.
The Method of Variation of Parameters
This is a general method for finding yP .
Consider
y 00 + p(x)y 0 + q(x)y = f (x)
23
The general solution is
yg = yh + yp
Complementary solution assumed known
y1 A0 + y2 B 0 = 0
y10 A0 + y20 B 0 = f (x)
The variable parameters A(x), B(x) are found by solving these two equations
first for A0 (x), B 0 (x).
PROOF:
yp = Ay1 + By2
yp0 = A0 y1 + Ay10 + B 0 y2 + By20
But if
A 0 y1 + B 0 y2 = 0
N
Then
y 0 = Ay10 + By20
Now
y 00 = A0 y10 + Ay100 + B 0 y20 + By200
Substitute trial yp (x) into the ODE
N L
Thus yp (x) is a solution of the ODE provided that and are satisfied.
N L
Now solve and to find A(x) and B(x) First write in matrix form
! ! !
y1 y2 A0 0
=
y10 y20 B0 f
24
We can find a unique solution if the determinant is non-zero. However the
TOPIC 6 – MATRICES
Brief revision (including special matrices)
A matrix is a rectangular array of numbers (real or complex) in the form
A = (ajk )
in which j is the row suffix and k the column suffix, e.g., a32 is the entry in
the 3rd row, 2nd column.
If all entries of A are real, we call A a real matrix; otherwise it is a complex
matrix.
Row vector : (matrix with one row)
a = (a1 , a2 , . . . , an )
b1
..
b = . .
bm
25
Definition: Two matrices A = (ajk ) and B = (bjk ) are equal if and only if A
and B have the same number of rows and columns and
We write A = B.
Definition: Addition of matrices is defined only for matrices with the same
number of rows and columns. The sum of two m × n matrices A and B is
an m × n matrix A + B with entries
We define the zero matrix 0 to be the m × n matrix with all entries zero.
Properties of Addition:
a) A + B = B + A
b) (U + V ) + W = U + (V + W )
c) A + 0 = A
d) A + (−A) = 0, −A = (−ajk ).
or cA = (cajk ).
For any m × n matrices A and B and any scalars c and k
a) c(A + B) = cA + cB
b) (c + k)A = cA + kA
c) c(kA) = (ck)A
d) 1A = A
Matrix multiplication
Let A = (ajk ) be an m × n matrix and B = (bjk ) an r × p matrix. Then
the product AB (in this order) is defined only when r = n. (i.e. the number
of rows of B = the number of columns of A). Then AB is an m × p matrix
C = (cjk ) where
26
or
n
X
cjk = ajl blk = aj1 b1k + aj2 b2k + · · · + ajn bnk (j = 1, · · · m, k = 1, · · · p).
l=1
b) A(BC) = (AB)C
c) (A + B)C = AC + BC
d) C(A + B) = CA + CB
Important notes:
(1) If A and B are matrices such that AB and BA are defined then, in general,
AB 6= BA
Special Matrices:
(1) Triangular Matrices: A square n × n matrix whose entries above the
main diagonal are all zero is called a lower triangular matrix. Similarly, a
square matrix whose entries below the diagonal are all zero is called a upper
triangular matrix.
(2) Diagonal Matrices: A square matrix A = ajk whose entries above and
below the main diagonal are all zero (i.e. ajk = 0 for j 6= k) is called a
diagonal matrix.
A scalar matrix is a diagonal matrix whose entries on the main diagonal
are all equal.
c 0 ... 0
0 c ... 0
S = .. .. . . .
. . . ..
0 0 ... c
We then have AS = SA = cA (for any n × n matrix A).
The unit matrix
1 0 ... 0
0 1 ... 0
I= .. .. . . . ..
. . .
0 0 ... 1
We then have AI = IA = A (for any n × n matrix A).
27
Transpose of a matrix: The transpose AT of an m × n matrix A = (ajk ) is
the n × m matrix in which the k th column of A becomes the k th row of AT
(and the j th row of A becomes the j th column of AT ).
a11 a21 . . . am1
a12 a22 . . . am2
AT = (akj ) = .. .. ..
..
. . . .
a1n a2n . . . amn
The transpose of a product of matrices is given by
(AB)T = B T AT .
where the ajk are the coefficients. The system of equations (9) has m equa-
tions with n unknowns. If all bi = 0 the system is homogeneous. If at least
one of the bi 6= 0 the system is nonhomogeneous. We can write the system
(9) as
Ax = b
where
a11 a12 . . . a1n
x1 b1
a21 a22 . . . a2n
.. ..
x= . , b= . , A = .. .. ..
. ...
. .
xn bm
am1 am2 . . . amn
28
This completely determines the system (9).
The system (9) is overdetermined if it has more equations than un-
knowns (m > n). It is determined if m = n and is undetermined if
m < n (i.e. the system has fewer equations than unknowns).
Theorem The system (9) has
3. Infinitely many solutions if the ranks of A and à are equal and < n.
Recall that the rank of a matrix is the maximum number of linearly inde-
pendent rows of the matrix. Somewhat surprisingly, this is also the maximum
number of linearly independent columns of the matrix or, in other words, row-
rank equals column-rank.
The solutions of (9) may be found by Gaussian elimination which is a
systematic process of elimination to reduce the matrix to its echelon form,
followed by back-substitution. Gaussian elimination is done using elementary
row operations:
AA−1 = A−1 A = I
29
Useful formula: For a nonsingular 2 × 2 matrix
where det A = a11 a22 − a12 a21 is the determinant of A. Note that a matrix
is nonsingular if det A 6= 0 (this holds for any general n × n matrix.)
Inverse of a diagonal matrix:
a11
0 ... 0
1/a11 0 ... 0
0 a22 ... 0 0 1/a22 ... 0
−1
A=
.. .. .. ..
A =
.. .. .. ..
. . . . . . . .
0 0 . . . ann 0 0 . . . 1/ann
(AC)−1 = C −1 A−1 .
Ax = λx, (10)
(A − λI)x = 0.
30
For a 2 × 2 matrix we get a quadratic in λ:
a11 − λ a12
det(A − λI) = = (a11 − λ)(a22 − λ) − a12 a21
a
21 a22 − λ
For a 3 × 3 matrix we get a cubic in λ. For a general n × n matrix A the
characteristic equation is a polynomial of order n in λ. Even if A is real, some
of the eigenvalues may be complex and hence so will be the corresponding
eigenvectors.
31
x2 y2 z
elliptic paraboloid : + = ,
a2 b2 c
x2 y2 z2
elliptic cone : + 2 = 2,
a2 b c
x2 y2 z2
hyperboloid (one sheet) : + 2 − 2 = 1,
a2 b c
z2 x2 y2
hyperboloid (two sheet) : − 2 − 2 = 1.
c2 a b
Quadratic forms: Quadric surfaces can be written in terms of a quadratic
form. Consider a real n × n matrix A and real vector x. Then
n X
n
xT Ax = ajk xj xk = a11 x21 + a12 x1 x2 + · · · + a1n x1 xn
X
j=1 k=1
AT = A−1
 = T −1 AT
Ax = λx
−1
⇒ T Ax = λT −1 x
⇒ T AT T −1 x
−1
= λT −1 x
⇒ Â(T −1 x) = λ(T −1 x)
32
Theorem: Let λ1 , λ2 , . . . λk be distinct eigenvalues of an n × n matrix. Then
the corresponding eigenvectors x1 , x2 , . . . , xk form a linearly independent set.
This is basically just Property 7 above.
Theorem: If an n × n matrix A has n distinct eigenvalues, then A has n
linearly independent eigenvectors which therefore constitute a basis for Rn
(or C n ). This follows from Property 7 above.
Orthonormal Sets of Vectors: Consider the set of vectors v1 , v2 , . . . , vn
This set is said to be orthonormal if vjT vj = 1 for j = 1, . . . , n and vjT vk = 0
for j 6= k. The matrix with these vectors as column vectors is orthogonal i.e..
if P = (v1 , v2 , . . . , vn ) then P −1 = P T .
Theorem: The eigenvalues of a real, symmetric matrix are all real. Hence, we
may take the eigenvectors to be real also.
Theorem: A real, symmetric matrix has an orthonormal basis of eigenvec-
tors in Rn .
Hence, even though some of the eigenvalues of a real, symmetric matrix
may be equal, we can always find n orthogonal eigenvectors. By normalising
each of these to have length 1, we can construct an orthonormal basis for Rn .
Diagonalisation of a Matrix
Theorem: If an n × n matrix A has a basis of eigenvectors (i.e. n linearly
independent eigenvectors) then
D = P −1 AP
is diagonal, with the eigenvalues of A as the entries on main diagonal. P is
the matrix with the eigenvectors of A as column vectors. Also
Dm = P −1 Am P.
Diagonalisation of a Symmetric Matrix
Not all matrices can be diagonalised - it relies upon us being able to find n
linearly independent eigenvectors. Fortunately, a real symmetric matrix can
always be diagonalised (because of Property 8 above). Even better, because
of Property 9, the columns of P are orthogonal to one another. If we further
normalise each eigenvector by dividing it by its length (the square root of the
sum of the squares of its n components), the matrix P is then an orthogonal
matrix and so its inverse is just its transpose. This saves a lot of work.
Transformation of a quadratic form to principal axes
Suppose we have a quadratic form
Q = xT Ax (12)
where A is real–symmetric. Then by a previous theorem, A has an orthonor-
mal basis of n eigenvectors. Hence the P matrix with eigenvectors of A as
column vectors is orthogonal. Now
D = P −1 AP ⇒ A = P DP −1 ⇒ A = P DP T
33
Substitution of this into (12) gives
Q = xT P DP T x
to the principal axis form (13) where λ1 , . . . , λn are the eigenvalues (not
necessarily distinct) of the symmetric matrix A and P is the orthogonal ma-
trix with the corresponding eigenvectors p1 , . . . , pn as column vectors.
(for simplicity we will assume that the coefficients ajk are constant).
Note: any higher (nth) order o.d.e can be reduced to a system of n first order
o.d.e’s.
Example:
d3 x d2 x
+ 2 2 + x = 0.
dt3 dt
Define
x = y1 ,
dx dy1
= = y2 ,
dt dt
d2 x dy2
= = y3 ,
dt2 dt
d3 x dy3 d2 x
= = −2 2 − x = −2y3 − y1 .
dt3 dt dt
Then
dy1 dy2 dy3
= y2 , = y3 , = −2y3 − y1 .
dt dt dt
Returning to (14): Rewrite this in matrix form
dy
= Ay + h, A = (ajk ). (15)
dt
34
Assume A has a basis of eigenvectors p1 , . . . pn . Then P = (p1 , . . . , pn ) is
nonsingular. To diagonalise (15) we write
y = Pz
Then assuming that all the ajk are constant (⇒ P is constant) we have
y0 = P z0 and (15) becomes
P z0 = AP z + h
⇒ z0 = P −1 AP z + P −1 h,
⇒ z0 = Dz + P −1 h,
where D is the diagonal matrix with the eigenvalues of A along the diagonal.
Writing the last expression in component form gives
zj0 − λj zj = rj (t),
Laplace Transforms
35
The original function f (t) in (17) is called the inverse transform of F (s); we
denote it by L−1 (F ), so
f (t) = L−1 (F ).
Linearity: The Laplace transformation is a linear operation. For any functions
f (t) and g(t) whose transforms exist and any constants a and b
L {af (t) + bg(t)} = aL {f (t)} + bL {g(t)} . (18)
Tables of Laplace Transforms: A short table of Laplace Transforms is at the
end of these notes.
Existence theorem: Let f (t) be a function that is piecewise continuous (i.e.
a function which is continuous except at a finite number of points) on every
finite interval in the range t ≥ 0 and satisfies
|f (t)| ≤ M eγt for all t ≥ 0, (19)
for some constants γ and M . Then the Laplace transform of f (t) exists for
all s > γ.
Transform of derivatives
Theorem: Suppose that f (t) is continuous for all t ≥ 0, satisfies (19) (for
some γ and M ) and f 0 (t) is piecewise continuous on every finite interval in
the range t ≥ 0. Then the Laplace transform of f 0 (t) exists when s > γ and
L(f 0 ) = sL(f ) − f (0) (for s > γ) (20)
Higher derivatives
36
Taking the inverse Laplace transform gives
! Zt
−1 F (s)
L = f (τ )dτ.
s
0
then
L(eat f (t)) = F (s − a), (21)
So if we know the transform F (s) of a function f (t), (21) gives us the trans-
form of exp(at)f (t) just by shifting on the s axis i.e. replacing s with s − a
to give F (s − a). Taking the inverse transform in (21) gives
0, if t < a;
f˜(t) =
f (t − a), if t > a
37
can now be written as
f˜(t) = f (t − a)u(t − a)
(strictly, we should define f˜(a) = f (0)/2) and our shifting on the t–axis result
is
L {f (t − a)u(t − a)} = e−as F (s) for a ≥ 0
and the inverse
n o
L−1 e−as F (s) = f (t − a)u(t − a) for a ≥ 0
Cases 1–4 lead to a general form of partial fraction for Y = F/G which then
have a corresponding inverse Laplace transform.
Case 1: In Y = F/G a fraction
A
(s − a)
38
has inverse transform
Aeat
with
A = F (a)/G0 (a).
Case 2: In Y = F/G a sum of m fractions
Am Am−1 A1
+ + · · · + .
(s − a)m (s − a)m−1 (s − a)
where
(s − a)m F (s)
Am = lim
s→a G(s)
dm−k (s − a)m F (s)
" #
1
Ak = lim (k = 1, . . . , m − 1)
(m − k)! s→a dsm−k G(s)
As + B A(s − α) + αA + B
or
(s − α)2 + β 2 (s − α)2 + β 2
Using the table and s–shifting theorem gives the inverse transform
!
αt αA + B
e A cos βt + sin βt
β
Here A is the imaginary part and (αA + B)/β is the real part of
As + B Cs + D
2 2 2
+
[(s − α) + β ] (s − α)2 + β 2
39
which has the inverse transform
" #
αt A αA + B
e t sin βt + (sin βt − βt cos βt)
2β 2β 3
" #
αt αC + D
+ e C cos βt + sin βt
β
The constants are given by formulas similar to those above.
Solving Ordinary Differential Equations
One of the main uses of Laplace transforms is in solving differential equa-
tions, both ordinary and partial.
Example: Using Laplace transforms solve
Therefore
y(t) = 2 cos t + t sin t.
The number p is then called the period. From (23) we then have
f (x + 2p) = f (x + p + p) = f (x + p) = f (x)
40
Thus for any integer n
Simple examples:
Observe that a function which is constant has period p for any p > 0. If
f is periodic but not constant, the smallest positive p for which (23) holds is
called the primitive period of f .
We will want to represent functions of period p = 2π in terms of simple
trigonometric functions 1, cos x, sin x, cos 2x, sin 2x, etc. Such a representation
will give rise to a trigonometric series.
where a0 , a1 , b1 etc are real constants—the an and bn are called the coeffi-
cients of the series. We can rewrite (24) as
∞
X
a0 + (an cos nx + bn sin nx).
n=1
Each term in the series has period 2π. Hence if the series converges, its sum
will be a function with period 2π.
i.e. assume the series converges and has f (x) as its sum.
Question: How do we determine the coefficients an and bn in (25) ?
First a0 . Integrate both sides of (25) from −π to π.
Zπ Zπ " ∞
X
#
f (x) dx = a0 + (an cos nx + bn sin nx) dx.
−π −π n=1
41
Assuming that term-by-term integration is valid, we have
Zπ Zπ ∞
X Zπ ∞
X Zπ
f (x) dx = a0 dx + an cos nx dx + bn sin nx dx
−π −π n=1 −π n=1 −π
Zπ π π
1Z 1Z
sin nx cos mx dx = sin(n + m)x dx + sin(n − m)x dx
2 2
−π −π −π
and evaluating these integrals shows that all four integrals are zero except for
π
1Z 0, if m 6= n;
cos(n − m)x dx =
2 π, if m = n.
−π
We then have
Zπ
f (x) cos mx dx = πam
−π
or π
1Z
am = f (x) cos mx dx (m = 1, 2, . . .) (27)
π
−π
42
Finally to obtain the coefficients bm we multiply (25) by sin mx (where m is a
positive integer) and integrate from −π to π. Following the above procedure
we obtain π
1Z
bm = f (x) sin mx dx (m = 1, 2, . . .) (28)
π
−π
To summarise
Rπ
1
a0 = f (x) dx
2π
−π
Rπ
1
am = π
f (x) cos mx dx (m = 1, 2, . . .) (29)
−π
1 Rπ
bm = f (x) sin mx dx (m = 1, 2, . . .)
π
−π
The expressions (29) are known as the Euler formulæ. In these integrals,
the limits need not be −π and π. Since everything has period 2π, we may
take the limits to be α and α + 2π for any real number α. In particular, it
is often convenient to take them to be 0 and 2π. The a0 , an , bn are called the
Fourier coefficients of f (x) and the series
∞
X
a0 + (an cos nx + bn sin nx). (30)
n=1
Suppose f (x) = f (x+2L) for all x, i.e., f has period 2L. Then its Fourier
series is of the form
∞
X nπx nπx
f (x) = a0 + (an cos + bn sin )
n=1 L L
The proof follows our earlier derivation of the Euler formulæ for 2π–periodic
functions. We can also prove these results by a change of scale. Set v = πx/L.
43
Then x = ±L ⇒ v = ±π and f (x) = g(v) where g(v) has period of 2π in v.
So g(v) has Fourier series
∞
X
g(v) = a0 + (an cos nv + bn sin nv)
n=1
Sums of functions
(1) The Fourier coefficients of a sum f1 +f2 are the sums of the corresponding
Fourier coefficients of f1 and f2 .
(2) The Fourier coefficients of cf are c times the Fourier coefficients of f .
44
ZL Z0
= h(x) dx + (−h(−v)) dv
0 L
ZL ZL
= h(x) dx − h(v) dv
0 0
=0
with coefficients
L L
1Z 2Z mπx
a0 = f (x) dx am = f (x) cos x (m = 1, 2, . . .)
L L L
0 0
Half–Range expansions
(i) A Fourier cosine series for f (x) by extending it to an even function on the
interval −L ≤ x ≤ L. The cosine half–range expansion for f (x) is thus
∞
X nπx
f (x) = a0 + an cos
n=1 L
where
L L
1Z 2Z mπx
a0 = f (x) dx am = f (x) cos dx (m = 1, 2, . . .)
L L L
0 0
45
(ii) A Fourier sine series for f (x) by extending it to an odd function on the
interval −L ≤ x ≤ L. The sine half–range expansion for f (x) is thus
∞
X nπx
f (x) = bn sin
n=1 L
where
L
2Z mπx
bm = f (x) sin dx (m = 1, 2, . . .)
L L
0
In (0, L) both these series represent f (x). Outside (0, L) they represent differ-
ent functions – the even and odd periodic extensions of f respectively.
Complex form of the Fourier series
Recall that
eiθ = cos θ + i sin θ (31)
Taking the complex conjugate gives
e−iθ = cos θ − i sin θ (32)
Then (31)+(32) and (31)-(32) give
1 iθ 1 iθ
cos θ = e + e−iθ sin θ = e − e−iθ .
2 2i
Hence
1 inx 1 inx
cos nx = e + e−inx sin nx = e − e−inx .
2 2i
Now consider the Fourier series
∞
X
f (x) = a0 + (an cos nx + bn sin nx)
n=1
∞
an inx bn inx
e + e−inx + e − e−inx
X
= a0 +
n=1 2 2i
∞
1 1
(an − ibn )einx + (an + ibn )e−inx
X
= a0 +
n=1 2 2
∞
cn einx + kn e−inx
X
⇒ f (x) = c0 +
n=1
where
1
cn = (an − ibn ), kn = c∗n .
2
Remember that
π
1Z
an = f (x) cos nx dx
π
−π
Zπ
1
bn = f (x) sin nx dx
π
−π
46
Hence
π π
1 1 Z i Z
cn = f (x) cos nx dx − f (x) sin nx dx
2 π π
−π −π
Zπ
1
= f (x)(cos nx − i sin nx) dx
2π
−π
Zπ
1
= f (x)e−inx dx
2π
−π
Similarly
π
1 Z
kn = f (x)einx dx
2π
−π
with π
1 Z
cn = f (x)e−inx dx.
2π
−π
This is then the complex form of the Fourier series for f (x). The cn are the
complex Fourier coefficients.
Forced Oscillations
Earlier we saw that forced oscillations of a body of mass m on a spring
are governed by the equation
my 00 + cy 0 + ky = r(t)
where k is the spring modulus, c the damping constant. If r(t) is a cos or sin
we can solve this equation by the method of undetermined coefficients.
What if r(t) is some other periodic function? We can represent r(t) by a
Fourier series; we should then be able to find a particular solution as a similar
Fourier series. To fix our ideas we consider a simple example
47
Example: Find the general solution to
y 00 + ω 2 y = r(t)
where ∞
X 1
r(t) = 2
sin(2n − 1)t.
n=1 (2n − 1)
Solution: We have
∞
00 2
X 1
y +ω y = 2
sin(2n − 1)t (33)
n=1 (2n − 1)
48
• The order of the highest derivative is called the order of the equation.
Some examples
∂ 2u 2
2∂ u
(1) = c one dimensional wave equation
∂t2 ∂x2
2
∂u 2∂ u
(2) =c one dimensional heat equation
∂t ∂x2
∂ 2u ∂ 2u
(3) + =0 two dimensional Laplace equation
∂x2 ∂y 2
∂ 2u ∂ 2u
(4) + = f (x, y) two dimensional Poisson equation
∂x2 ∂y 2
(1)-(3) are homogeneous, linear. (4) is inhomogeneous, linear. All are second
order.
We may, depending on the problem, have boundary conditions (the
solution has some given value on the boundary of some domain) or initial
conditions (where the value of the solution will be given at some initial time,
e.g. t = 0).
Superposition Theorem: If u1 and u2 are any solutions of a linear homogeneous
p.d.e in some region, then
u = c1 u1 + c2 u2
Vibrating Strings
As shown in Section 2 of Chapter 11 of the book by Kreyszig, the small
transverse oscillations of a tightly stretched string satisfy
∂ 2u 2
2∂ u
= c where c2 = T /ρ,
∂t2 ∂x2
T is the tension in the string, ρ is the constant mass per unit length of the
string and u(x, t) is the displacement of the string.
49
D’Alembert’s solution of the wave equation
The equation
∂ 2u 2
2∂ u
= c (35)
∂t2 ∂x2
is called the one-dimensional wave equation.
Introduce new variables
v = x + ct, z = x − ct
Thus
uvz = 0. (36)
Integrating (36) gives
∂u
= h(v)
∂v Z
⇒ u = h(v)dv + ψ(z)
i.e.
u = φ(v) + ψ(z)
R
where φ = h(v)dv. Thus
50
The functions φ and ψ can be determined from the initial conditions.
Suppose we have to solve (35) for −∞ < x < ∞ and t > 0 subject to the
initial conditions
∂u(x, 0)
u(x, 0) = f (x), = g(x) for −∞<x<∞
∂t
From (37) we have
∂u
= cφ0 (x + ct) − cψ 0 (x − ct)
∂t
where the 0 denotes differentiation with respect to the entire argument x + ct
and x − ct respectively. Then
2ψ(x) = f (x) − k
1 1
⇒ ψ = [f (x) − k] , φ= [f (x) + k]
2 2
Hence
1
u(x, t) = [f (x + ct) + f (x − ct)] .
2
If g(x) 6= 0 then
1 1 Z x+ct
u(x, t) = [f (x + ct) + f (x − ct)] + g(s)ds.
2 2c x−ct
Note: In the solution (37), ψ(x − ct) represents a disturbance travelling
to the right with speed c without changing its shape and φ(x + ct)
represents a disturbance travelling to the left with speed c without
changing its shape.
SEPARATION OF VARIABLES
This is an important technique which can be used to find the solutions of
many linear partial differential equations and is best illustrated by examples.
Wave equation
The one–dimensional wave equation
∂ 2u 2
2∂ u
= c (40)
∂t2 ∂x2
51
(e.g. a model of the deflection u(x, t) of an elastic string) can be solved by
this technique. Physically, conditions will be imposed on the solution of (40).
Boundary conditions (string fixed at ends x = 0 and x = L). Then
Initial conditions (the motion of the string will depend upon its initial
deflection and its initial velocity i.e. the deflection and velocity at t = 0).
u(x, t) = F (x)G(t).
Differentiating gives
∂ 2u d2 G ∂ 2u d2 F
= F , = G
∂t2 dt2 ∂x2 dx2
Substituting into equation (40) gives
d2 G 2
2 d F
F = c G
dt2 dx2
rearranging (divide by c2 F G)
1 d2 G 1 d2 F
=
c2 G dt2 F dx2
1 d2 G 1 d2 F
⇒ = = k (k constant).
c2 G dt2 F dx2
This then gives us two ordinary differential equations:
d2 F d2 G
− kF = 0, − kc2 G = 0 . (44)
dx2 dt 2
52
Step 2: Satisfy the boundary conditions. With u = F G
F = Aeµx + Be−µx
F (0) = 0 ⇒ C=0
nπ
F (L) = 0 ⇒ D sin pL = 0 ⇒ sin pL = 0 ⇒ p= (n positive integer)
L
So p = nπ/L and D is arbitrary (take D = 1). We then have an infinite
number of solutions
nπx
Fn (x) = sin for n = 1, 2, 3, · · ·
L
Now with k = −p2 = −(nπ/L)2 we have
d2 G
2
cnπ
2
+ G = 0.
dt L
which has the general solution
or
nπx
un (x, t) = (Bn cos λn t + Cn sin λn t) sin for n = 1, 2, 3, · · ·
L
53
are solutions of (40) satisfying the boundary conditions (41) . We call the
un ’s the eigenfunctions and the λn ’s the eigenvalues. The set {λ1 , λ2 , · · ·}
is called the spectrum.
Step 3: Satisfy the initial conditions. The initial conditions are given by
(42) and (43) . Since the wave equation is linear and homogeneous, a sum of
solutions is also a solution. Consider
∞ ∞
X X nπx
u(x, t) = un (x, t) = (Bn cos λn t + Cn sin λn t) sin
n=1 n=1 L
We have ∞
∂u X nπx
= λn (−Bn sin λn t + Cn cos λn t) sin
∂t n=1 L
From (43)
∞
∂u X nπx
= g(x) = λn Cn sin
∂t t=0
n=1 L
Again, this is the Fourier–sine half–range expansion for g(x). Thus
L
2Z nπx
λn Cn = g(x) sin dx
L L
0
or
L
2 Z nπx
Cn = g(x) sin dx (for n = 1, 2, . . .) (47)
cnπ L
0
Summary
∂ 2u 2
2∂ u
= c
∂t2 ∂x2
∂u
u(x, 0) = f (x) = g(x)
∂t t=0
has solution (see note below)
∞
X nπx
u(x, t) = (Bn cos λn t + Cn sin λn t) sin
n=1 L
54
where
L
2Z nπx
Bn = f (x) sin dx (for n = 1, 2, . . .)
L L
0
L
2 Z nπx
Cn = g(x) sin dx (for n = 1, 2, . . .)
cnπ L
0
Note: it is a solution if the infinite series converges and the series obtained
by differentiating twice with respect to x and t converges with sums uxx and
utt (which are continuous).
For the case g(x) = 0, we have Cn = 0 and
∞
X nπx
u(x, t) = Bn cos λn t sin
n=1 L
We can write
cnπt nπx 1 nπ nπ
cos sin = sin (x − ct) + sin (x + ct)
L L 2 L L
Then
∞ ∞
1X nπ 1X nπ
u(x, t) = Bn sin (x − ct) + Bn sin (x + ct)
2 n=1 L 2 n=1 L
but ∞
X nπx
Bn sin = f (x)
n=1 L
(the Fourier sine series for f (x)). Thus
1 ∗
u(x, t) = {f (x − ct) + f ∗ (x + ct)}
2
where f ∗ is the odd periodic extension of f with period 2L. Since f (x) is
continuous on 0 ≤ x ≤ L and zero at endpoints ⇒ u(x, t) is continuous for
all x, t. Differentiating ⇒ u(x, t) is a solution of (40) (provided f 00 exists
for 0 < x < L and f has one–sided derivatives at x = 0, L). Under these
conditions u(x, t) is a solution of (40) satisfying the conditions (41), (42) and
(43) .
Note: compare the above solution with D’Alembert’s solution.
Heat equation
The heat flow in a body of homogeneous material is governed by the heat
equation
∂u κ
= c2 ∇2 u c2 = .
∂t σρ
55
where u(x, y, z, t) is the temperature in the body, κ =thermal conductivity,
σ =specific heat and ρ =density. Here
∂ 2u ∂ 2u ∂ 2u
∇2 u = + +
∂x2 ∂y 2 ∂z 2
Consider the temperature in a long thin wire of constant cross section
(homogeneous material) that is perfectly insulated laterally (so that heat
flows only in the x direction). Then u depends only on x and t. The heat
equation becomes
∂u ∂ 2u
= c2 2 (48)
∂t ∂x
Consider the case when the ends of the bar are kept at zero temperature and
suppose the initial temperature is f (x). This then gives us the conditions
boundary conditions:
initial conditions:
u(x, 0) = f (x) for 0 < x < L (50)
We want to determine the solution of (48) subject to (49) and (50). Use
separation of variables.
Step 1: Separation of variables. Look for a solution in the form
1 dG 1 d2 F
= = −p2 = constant.
c2 G dt F dx2
Thus
dG
+ c2 p 2 G = 0 (52)
dt
d2 F
+ p2 F = 0 (53)
dx2
Step 2: Satisfy the boundary conditions (49).
56
(setting B = 1 we have)
nπx
Fn = sin (n = 1, 2, . . .).
L
Now equation (53) becomes
dG c2 n2 π 2
+ G=0
dt L2
which has the general solution
2
Gn = Bn e−λn t , λn = cnπ/L.
So the functions
nπx −λ2n t
un = Bn sin e .
L
are solutions of (48) satisfying (49).
Step 3: Satisfy the initial conditions (50). As the heat equation is linear
∞
X nπx −λ2n t
u(x, t) = Bn sin e , (55)
n=1 L
which is just the half–range expansion for f (x) (i.e. the Fourier sine series).
Hence
L
2Z nπx
Bn = f (x) sin dx (56)
L L
0
So (55) with coefficients (56) is the solution to the heat equation (48) satisfy-
ing (49) and (50) (provided the series converges). Note: The presence of the
2
e−λn t terms in (55) means that all terms in (55) approach zero as t −→ ∞
(i.e. the temperature tends to zero as t −→ ∞).
57
Mixed problem if u is prescribed on a part of C and un on the rest of C.
We will consider only the Dirichlet problem on the rectangle 0 ≤ x ≤
a, 0 ≤ y ≤ b with u = 0 along the boundary except for the edge y = b where
u(x, b) = f (x) for 0 ≤ x ≤ a.
To solve (57) we use separation of variables. Substituting u(x, y) = F (x)G(y)
into (57) gives
1 d2 F 1 d2 G
= − = −k = (constant).
F dx2 G dy 2
Our boundary conditions give
F (0) = 0, F (a) = 0, G(0) = 0
Solving for F gives
nπx
Fn (x) = sin k = (nπ/a)2
a
Solving for G subject to G(0) = 0 gives
nπy
Gn = 2An sinh
a
Finally we need to satisfy the boundary condition
u(x, b) = f (x).
Write ∞
X nπx nπy
u(x, y) = 2 An sin sinh
n=1 a a
Then
∞
X nπx nπb
f (x) = 2 An sin sinh
n=1 a a
∞
X nπx
= Cn sin
n=1 a
where
nπb
Cn = 2An sinh .
a
These are the Fourier coefficients of f (x)
a
nπb 2Z nπx
2An sinh = f (x) sin dx
a a a
0
Thus
Za
1 nπx
An = nπb f (x) sin dx
a sinh a a
0
Therefore the solution of the problem is
∞
X nπx nπy
u(x, y) = 2 An sin sinh
n=1 a a
with the An ’s given above.
58