SG Pdf#section 1 5
SG Pdf#section 1 5
Study Guide
202330 (version 2.0)
Last updated: Thursday 11th May, 2023
Ordinary Differential Equations
Written by
L Lobb, B Wilkes, D Demskoy
Produced by School of Computing and Mathematics, Charles Sturt University, Albury -
Bathurst - Wagga Wagga, New South Wales, Australia.
First Published March 2021
ii
1
0.1 Differentials
If y = f (x) then
dy
= f ′ (x)
dx
and we could write
dy = f ′ (x)dx
but
dy = d[f (x)]
so we have the rule
d[f (x)] = f ′ (x)dx
Example 1
d
d[sin x] = (sin x)dx = cos xdx
dx
Example 2
d (x3 ) = 3x2 dx
Function of a function
dy df dz dz
= = f ′ (z)
dx dz dx dx
Then
′ dz ′
∫ f (z) dx dx = ∫ f (z)dz = f (z) + C.
Example 1
2 2
∫ 2x cos x dx = ∫ cos x ⋅ 2xdx.
3
4 CHAPTER 0. REVIEW OF INTEGRATION TECHNIQUES
2 2
∫ 2x cos x dx = ∫ cos zdz = sin z + C = sin x + C
Example 2
ln x 1
∫ dx = ∫ ln x × dx
x x
and we note that
1
d(ln x) = dx
x
giving
ln x 1
∫ dx = ∫ ln xd[ln x] = (ln x)2 + C.
x 2
Changing variables
Similar to previous, but any variable formula can be tried. There may be more than one
way to work the integral. Trying to find the easiest substitution can be interesting. Given
∫ f (x)dx
′
∫ f (x)dx = ∫ f [g(z)]g (z)dz
Example 1
Calculate
1
∫ √ dx
x+1
Try substitution z = x + 1 or x = z − 1 then
1 1
dx = dz; √ = √ = z −1/2
x+1 z
giving
Example 2
again.
Example 3
Calculate π
2
∫0 sin x cos xdx.
Try z = sin x then dz = d[sin x] = cos xdx; z(0) = sin 0 = 0 and z ( π2 ) = sin π2 = 1 giving
π
2
1 1 1 1 1 1
∫0 sin x cos xdx = ∫ zdz = [ z 2 ] = × 12 − × 02 = .
0 2 0 2 2 2
Note: We can check this using sin 2x = 2 sin x cos x. Then
π π π
2 1 2 1 2
1 1 1 1 1 1 1
= − cos π + cos 0 = − × (−1) + × 1 = + =
4 4 4 4 4 4 2
as expected.
Integration by parts
Example 1
Calculate
x
∫ xe dx.
Notice that the u in the first integral appears on the RHS differentiated. The method can
be repeated. This means that if we choose u as a polynomial (if it is possible) we will
eventually obtain a du of a constant (giving 0 ) and the method will terminate.
So we choose dv = ex dx then v = ∫ ex dx = ex ; u = x giving du = dx and we obtain
x x x x x
∫ xe dx = ∫ udv = uv − ∫ vdu = xe − ∫ e dx = xe − e + C
Example 2
Calculate
∫ ln xdx.
This is already in the form ∫ udv so we simply choose ln x = u and dv = dx so v = x and
d 1
du = (ln x)dx = dx
dx x
giving
1
∫ ln xdx = uv − ∫ vdu = (ln x) × x − ∫ x × x dx = x ln x − ∫ 1dx = x ln x − x + C
Check:
d 1
(x ln x − x + C) = ln x + x × − 1 + 0 = ln x + 1 − 1 = ln x. Amazing!
dx x
Note. Using differentials we could have done the by-parts step as
∫ ln xdx = x ln x − ∫ xd[ln x]
and so on.
Consider
x
∫ e sin xdx.
0.2. INTEGRATION TECHNIQUES 7
I = ex sin x − ex cos x − I
So
2I = ex (sin x − cos x)
giving
ex
I = ∫ ex sin xdx = (sin x − cos x).
2
Check:
d ex ex ex
[ (sin x − cos x)] = (sin x − cos x + cos x + sin x) = × 2 sin x = ex sin x. Yes!
dx 2 2 2
n sinn−1 x cos x n − 1
∫ sin xdx = − + sinn−2 xdx
n n ∫
and
n cosn−1 x sin x n − 1
∫ cos xdx = + cosn−2 xdx
n n ∫
Mixtures:
n m
∫ sin x cos xdx
If n and/or m are odd; simply let z = cos x or z = sin x if n or m is odd. If n and m are both
even, then use the identities
1
sin2 x = (1 − cos 2x)
2
and
1
cos2 x = (1 + cos 2x).
2
Example 1
Calculate
3
∫ sin xdx.
Using n = 3 gives
n sin3−1 x cos x 3 − 1
∫ sin xdx = − + sin3−2 xdx
3 3 ∫
8 CHAPTER 0. REVIEW OF INTEGRATION TECHNIQUES
sin2 x cos x 2 1 2
=− + ∫ sin xdx = − sin2 x cos x − cos x + C
3 3 3 3
Example 2
2 3
∫ sin x cos xdx
3 is odd, so try z = sin x and dz = d[sin x] = cos xdx giving
2 3 2 2 2 2 2 4
∫ sin x cos xdx = ∫ sin x cos x cos xdx = ∫ z (1 − z ) dz = ∫ z − z dz
1 1 1 1
= z 3 − z 5 + C = sin3 x − sin5 x + C
3 5 3 5
Example 3
2
2 4 1 1
∫ sin x cos xdx = ∫ 2 (1 − cos 2x) [ (1 + cos 2x)] dx
2
1 1
= ∫ (1 − cos2 2x) (1 + cos 2x)dx = ∫ sin2 2x(1 + cos 2x)dx
8 8
1 1 1 1 1
= ∫ sin2 2x + sin2 2x cos 2xdx = ∫ (1 − cos 4x)dx + ∫ sin2 2x × d[sin 2x]
8 8 2 8 2
1 1 1
= x− sin 4x + sin3 2x + C
16 64 48
√
Integrals involving a2 ± x2 , a2 ± x2
Example 1
Calculate
dx
∫ √ 2 .
a − x2
Try x = a sin θ then dx = a cos θdθ giving
a cos θdθ a cos θdθ a cos θdθ cos θdθ
I =∫ √ =∫ √ =∫ √ =∫
a2 − (a sin θ)2 a2 − a2 sin2 θ a 1 − sin2 θ cos θ
x x
= ∫ dθ = θ + C = sin−1 ( ) + C = arcsin ( ) + C.
a a
Example 2
Calculate
dx
∫ a2 + x 2 .
Try x = a tan θ then dx = a sec2 θdθ giving
a sec2 θdθ a sec2 θdθ sec2 θdθ 1 1
I =∫ 2 2 2 = ∫ 2 2 = ∫ 2
= ∫ dθ = θ + C
a + a tan θ a (1 + tan θ) a sec θ a a
0.2. INTEGRATION TECHNIQUES 9
1 x 1 x
= arctan ( ) + C = arctan ( ) + C
a a a a
Example 3
Calculate
xdx
∫ √ 2 .
x −4
Try x = 2 sec θ then dx = 2 sec θ tan θdθ giving
2 sec θ ⋅ 2 sec θ tan θdθ sec2 θ tan θdθ 4 sec2 θ tan θdθ
I =∫ √ = 4∫ √ =√ ∫
(2 sec θ)2 − 4 4 (sec2 θ − 1) 4 tan θ
4 √ √
= ∫ sec2 θdθ = 2 tan θ + C = 2 tan2 θ + C = 2 sec2 θ − 1 + C
2√
x 2 √
= 2 ( ) − 1 + C = x2 − 4 + C
2
Example
dx dx dx
∫ x2 + 2x + 5 = ∫ x2 + 2x + 1 + 4 = ∫ (x + 1)2 + 22
1 x+1
arctan ( ) + C(a = 2)
2 2
f (x) A1 A2 Am
= + + ... +
(x − r1 ) (x − r2 ) ⋯ (x − rm ) x − r1 x − r2 x − rm
Quadratic terms
These theoretically will reduce to
A1 A2
+
x − r1 x − r2
but this may be messy if we have something like
1
x2 + 4x + 6
which does not have real factors. Instead, add
A1 A2 A1 (x − r2 ) + A2 (x − r1 ) (A1 + A2 ) x + (−A1 r2 − A2 r1 )
+ = =
x − r1 x − r2 (x − r1 ) (x − r2 ) (x − r1 ) (x − r2 )
which suggests trying Ax + B in the numerator. Similarly, for a cubic term, we would try
Ax2 + Bx + C in the numerator. And so on.
Example 1
Calculate
3x − 4 3x − 4
∫ x2 − 3x + 2 dx = ∫ (x − 1)(x − 2) dx.
Try
3x − 4 A B
= + .
(x − 1)(x − 2) x − 1 x − 2
This gives 3x − 4 = A(x − 2) + B(x − 1). By equating coefficients
3 = A + B and − 4 = −2A − B
Example 2
Calculate
x2 − x + 4
∫ (x − 1)(x + 1)2 dx.
Try
x2 − x + 4 A B C
2
= + + .
(x − 1)(x + 1) x − 1 x + 1 (x + 1)2
For the single linear factor x − 1 we can use the quick method:
x2 − x + 4 12 − 1 + 4 1 + 3 4
A= ∣ = = 2 = =1
(x + 1)2 x=1 (1 + 1)2 2 4
x2 − x + 4 1 3 3
∫ (x − 1)(x + 1)2 dx = ∫ ( x − 1 − (x + 1)2 ) dx = ln(x − 1) + x + 1 + C.
Example 3
Calculate
x2 − 4
∫ (x + 2) (x2 + 5) dx.
We try
x2 − 4 A Bx + C
2
= + 2
(x + 2) (x + 5) x + 2 x +5
Again,
x2 − 4 (−2)2 − 4 4 − 4
A= ∣ = = =0
x2 + 5 x=−2 (−2)2 + 5 4 + 5
Then we have x2 − 4 = (Bx + C)(x + 2) giving straight away B = 1, 2C = −4 so C = −2
and we obtain
1
x−2 × 2x dx
I =∫ 2 dx = ∫ 2 2 dx − 2 ∫ √
x +5 x +5 x2 + ( 5)2
1 d (x2 + 5) 2 x
= ∫ 2
− √ arctan ( √ ) + C
2 x +5 5 5
1 2 x
= ln (x2 + 5) − √ arctan ( √ ) + C.
2 5 5
12 CHAPTER 0. REVIEW OF INTEGRATION TECHNIQUES
Try
x
z = tan
2
which leads to the substitution formulae:
2z 1 − z2 2dz
sin x = 2
, cos x = 2
, dx = .
1+z 1+z 1 + z2
Example
dx 1 2dz 2dz
∫ 2 + cos x = ∫ 2 + (1 − z 2 ) / (1 + z 2 ) ⋅ 1 + z 2 = ∫ 2 (1 + z 2 ) + 1 − z 2
Infinite limits
∞ R
∫a f (x)dx = lim ∫ f (x)dx
R→∞ a
If the limit exists, and is finite, the improper integral is said to converge, otherwise it is
divergent.
Example 1
∞ dx R dx R π
∫0 2
= lim ∫ 2
= lim [tan−1 x]0 = lim tan−1 R − 0 =
1+x R→∞ 0 1+x R→∞ R→∞ 2
and the integral is convergent.
Example 2
∞
dx R dx
∫1 = lim ∫ = lim [ln x]R
1 = lim ln R = ∞
x R→∞ 1 x R→∞ R→∞
1 2
sech u = = u −u ;
cosh u e + e
1 2
cosech u = = u −u .
sinh u e − e
Connection with trigonometric functions
From Euler’s formula
and
e−iθ = cos θ − i sin θ
giving
eiθ − e−iθ
sinh iθ = = i sin θ
2
and
eiθ + e−iθ
cosh iθ = = cos θ
2
and
sinh iθ i sin θ
tanh iθ = = = i tan θ
cosh iθ cos θ
So
1
sin θ = sinh iθ; cos θ = cosh iθ
i
and
1
tan θ = tanh iθ.
i
Identities
From
sin2 θ + cos2 θ = 1
we find
− sinh2 iθ + cosh2 iθ = 1
or
cosh2 u − sinh2 u = 1
Similarly
1 − tanh2 u = sech2 u
1 − coth2 u = − cosech2 u
Derivatives
d
(sinh x) = cosh x
dx
d
(cosh x) = sinh x
dx
d
(tanh x) = sech2 x
dx
d
(coth x) = − cosech2 x
dx
d
(sech x) = − sech x tanh x
dx
d
(cosech x) = − cosech x coth x
dx
Inverse hyperbolic functions
√
sinh−1 x = ln (x + x2 + 1) −∞<x<∞
√
cosh−1 x = ln (x + x2 − 1) ∣x∣ > 1
1 1+x
tanh−1 x = ln ( ) ∣x∣ < 1
2 1−x
1
sech−1 x = cosh−1 ( ) 0 < x ≤ 1
x
1
cosech−1 x = sinh−1 ( ) x ≠ 0
x
1
coth−1 x = tanh−1 ( ) ∣x∣ > 1
x
Derivatives of inverse hyperbolic functions
d 1
(sinh−1 x) = √
dx 1 + x2
d 1
(cosh−1 x) = √
dx x2 − 1
d 1
(tanh−1 x) = ∣x∣ < 1
dx 1 − x2
d 1
(coth−1 x) = ∣x∣ > 1
dx 1 − x2
d
(sech−1 x) = √
−1
dx x 1 − x2
d
(cosech−1 x) = √
−1
dx ∣x∣ 1 + x2
Example
4 √ √
= ∫ cosh udu = 2 sinh u + C = 2 sinh u + C = 2 cosh2 u − 1 + C
2
2
√
x 2 √
= 2 ( ) − 1 + C = x2 − 4 + C
2
as before.
16 CHAPTER 0. REVIEW OF INTEGRATION TECHNIQUES
Chapter 1
Chapter objectives
Chapter summary
4. Integrating factors
17
18 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
In the past 100 years, work on differential equations has gone on rapidly on many fronts.
The theoretical base of the topic has been strengthened; the set of classes into which dif-
ferential equations can be classified has been enlarged; the use of numerical methods has
been developed greatly, especially with the availability of the computer; the investigation
of partial differential equations has grown; the range of applications of differential equa-
tions has been extended and so on. Like all of mathematics, there has been an explosion
of knowledge. Today no one can be a master of the whole topic as could Euler, Laplace,
Cauchy.
Let us in this subject be content if we can master some of the first steps in the whole field
of Differential Equations!
1.2 Introduction
Read textbook sections 1.1 and 1.2 (Background, Solutions and initial value
problems)
y = ∫ f (x)dx,
dy
= f (x, y), y (x0 ) = y0 .
dx
If f and ∂f /∂y are continuous functions in some rectangle
that contains the point (x0 , y0 ), then the initial value problem has a unique solution in some
interval containing x0 .
Fortunately, for the specific types of “basic” (but very useful) D.E.s that we will examine in
this subject, we can always assume (correctly) that a solution exists. Thus the third question
is the only important one for us: we must develop and learn the appropriate methods for
solving such D.E.s.
Before proceeding to examine exact D.E.s, we will consider very briefly the problem of
finding the “derivative” (the partial derivative) of a function of more than one variable.
Thus far in Calculus, we have dealt only with functions of one variable. Such functions
may be given explicitly in the form of y = f (x) or implicitly by an equation of the form
F (x, y) = 0. In both cases, x is taken to be the independent variable and y is called the
dependent variable. Thus y varies as x varies, and we evaluate the rate of change of y with
respect to (w.r.t.) x, using the derivative, In general, it is possible to consider functions of
two (or more) variables which can be written explicitly in the form
z = f (x, y).
Note that here x and y are both independent variables, and that z is the dependent variable.
Thus z varies as we vary x and/or y independently, and we will again want to evaluate the
rate of change of z, but this time (separately) with respect to x or w.r.t. y. The notation
∂z
used for these two rates of change is ∂x ∂z
and ∂y (The notations: ∂f ∂f
∂x and ∂y or fx and fy are
also used sometimes.)
Some remarks
In general, functions of more than one variable are very important since real-life models
we want to use often have to take into account a number of “factors” or variables.
22 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
For example, the distance travelled by a car depends on both the time and the acceleration,
i.e. we have s = f (a, t) Similarly, the volume of gas depends on the pressure and the
temperature, i.e. V = g(P, T ) The volume of a circular cylindrical tank is a function of the
radius and the height, i.e. V = F (r, h).
Such functions are studied in detail in the subject MTH218 Multivariable Calculus. There
we will define the concepts of “limit” and “continuity” for such functions; we will try to
draw the surface represented by z = f (x, y); we will try to find the tangent plane to such
a surface; we will examine maxima/minima problems for such functions, and so on!
For now, we only need to consider the question of finding the partial derivatives of a
∂z
function given by z = f (x, y). We say that ∂x represents the rate of change of z (or “f ”)
with respect to x, keeping y constant, and define this “partial derivative of z w.r.t. x” as
∂z f (x + h, y) − f (x, y)
= lim
∂x h→0 h
Similarly,
∂z f (x, y + h) − f (x, y)
= lim
∂y h→0 h
is the “partial derivative of z w.r.t. y” and represents the rate of change of z w.r.t. y, keeping
x constant.
Of course, as usual, we do not use these definitions to evaluate partial derivatives! Instead
we simply differentiate according to all the ordinary rules for differentia-tion, keeping all
∂ ∂
the other (independent) variables constant. i.e. ∂x y = 0 and ∂y x = 0, and so on for more
than 2 variables.
Example 1
∂z ∂z
Let z = x2 + y 4 + 3. Find and .
∂x ∂y
Solution
∂z ∂ 4 ∂
Here, = 2x. (Note that y and 3 both equal 0, since both are constant w.r.t. x.)
∂x ∂x ∂x
∂z
= 4y 3 .
∂y
Example 2
Similarly,
∂z
= 2xy + x2 cos y.
∂y
Example 3
Find
∂ x2 + y
{ }
∂x cos (x3 y)
24 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
Solution
∂ ∂
∂ x2 + y cos (x3 y) ∂x (x2 + y) − (x2 + y) ∂x cos (x3 y)
( ) =
∂x cos (x3 y) cos2 (x3 y)
2x cos (x3 y) + 3x2 y (x2 + y) sin (x3 y)
=
cos2 (x3 y)
Example 4
Let
√ 1
f (u, r, t) = u2 + r2 − 3ur t + 2
t
∂f ∂f ∂f
Find , and .
∂u ∂r ∂t
Solution
Here we have a function of three independent variables and hence we can find three (first)
partial derivatives.
∂f √
∴ = 2u − 3r t
∂u
∂f √
= 2r − 3u t
∂r
∂f 3 2
= − urt− 2 − 3 .
1
∂t 2 t
Note
Whole new fields of investigation and applications open up from the consideration of such
functions and their partial derivatives. We will not make much use of these techniques
in this subject. The topic has really been introduced to demonstrate how the concept of
“derivative” and the process of “differentiation” can be “generalised”, and to enable us to
consider exact D.E.’s.
Tutorial 1 Question 1
1.3. EXACT DIFFERENTIAL EQUATIONS 25
Higher order derivatives can also be defined for functions of more than one variable. If
z = f (x, y), then we can find two first order partial derivatives:
∂z ∂z
and
∂x ∂y
and four second order partial derivatives:
∂ 2z ∂ 2z
,
∂x2 ∂y 2
and the 2 “mixed partials”
∂ 2z ∂ ∂z
≡ ( )
∂x∂y ∂x ∂y
and
∂ 2z ∂ ∂z
≡ ( ),
∂y∂x ∂y ∂x
and so on.
Note
In Multivariable Calculus we will find that for many functions of two variables, the mixed
partial derivatives of the same order and “type” are equal. That is,
∂ 2f ∂ 2f
=
∂x∂y ∂y∂x
for such functions.
Example 1
as “predicted”.
y+1
(Carry out the same calculation for z = 3x4 y 2 + .)
x2 − 2
Total derivative
Another important definition is that of the total derivative. The total derivative of z(x, y),
a function of both x and y, is written dz, and is defined by
∂z ∂z
dz = dx + dy.
∂x ∂y
Example
∂C ∂C
=0= .
∂x ∂y
Thus for such a function,
dC = 0.
If we have the equation
z(x, y) = C (C a constant),
the total derivative of both sides can be found, and these must be equal.
dz = dC
or
∂z ∂z
dx + dy = 0.
∂x ∂y
Such an equation is called an exact D.E.
For example, if x3 + 4xy + 6y 2 = 4 (in this example z = x3 + 4xy + 6y 2 ), then
1. How do we know whether a D.E. is exact or not? (What test, if any, can we use?)
Examples
Determine which of the following D.E.s are exact:
4xy
1. y ′ =
y − 2x2
2. y 2 dx = 2xydy
3. (x + sin y)dx + (y 3 + x cos y)dy = 0
Solutions
1. The equation
4xy
y′ =
y − 2x2
needs to be put into the form: M dx + N dy = 0.
Solution of exact D.E.s To obtain the general solution of an exact D.E., either of two
methods can be employed.
The first method involves integration, differentiation, and integration again: its use is illus-
trated in Examples 1, 2, 3.
The second method involved “the grouping of terms”: its use is illustrated in Examples 4,
5, 6.
Method 1
Recall that a D.E. of the form M (x, y)dx + N (x, y)dy = 0 is called exact if there exists a
function f (x, y) such that its total differential,
∂f ∂f
df = dx + dy,
∂x ∂y
is equal to M (x, y)dx + N (x, y)dy.
∂f ∂f
Thus, to solve an exact D.E., we simply use that function f for which ∂x = M and ∂y = N,
where M and N are given in the D.E.
(The way in which we then actually find f is demonstrated in the examples.)
Since M dx + N dy = 0 for an exact D.E., once we have found the appropriate function f ,
then we must have that df = 0, and so the solution to the exact D.E. is
Example 1
Solve the D.E.
(2x sin 3y)dx + (3x2 cos 3y)dy = 0.
Solution
Here,
∂ ∂
(2x sin 3y) = 6x cos 3y = (3x2 cos 3y),
∂y ∂x
and so the D.E. is exact.
f (x, y) must be found so that
∂f
= M (x, y) = 2x sin 3y
∂x
and
∂f
= 3x2 cos 3y.
∂y
Now consider
∂f
= 2x sin 3y.
∂x
To have obtained 2x sin 3y by partially differentiating f with respect to x, f must include
∂
a term in x2 sin 3y. (Then ∂x (x2 sin 3y) = 2x sin 3y as required.) We have
∂
where ϕ(y) is a “constant function” with respect to x. ( ∂x (x2 sin 3y+ϕ(y)) = 2x sin 3y+0 =
2x sin 3y.)
Now, for
∂f
f (x, y) = x2 sin 3y + ϕ(y), = 3x2 cos 3y + ϕ′ (y)
∂y
and this must equal
N (x, y) = 3x2 cos 3y.
Hence ϕ′ (y) must equal 0. i.e. ϕ(y) = a constant, c. Thus
f (x, y) = x2 sin 3y + c.
We have found the appropriate function for the exact D.E. and so the solution is given by
(x2 sin 3y + c) = K.
i.e.
x2 sin 3y = C.
Note
To provide some assurance that this method has worked, let us check that we have indeed
found the solution to the original D.E., which can be written as
dy 2 sin 3y
=− .
dx 3x cos 3y
The solution we have found gives
C
sin 3y =
x2
and so y is given explicitly by
1 C
y= arcsin ( 2 )
3 x
dy 1 1 −2C
= √ ⋅( 3 )
dx 3 1 − C 2 x
x4
C
2 2
=− √ x
3x 1 − C 2
x4
2 sin 3y
= − ,
3x cos 3y
C
since sin 3y = x2 and
√
√ C2
cos 3y = 1 − sin2 3y = 1−
x4
dy
Thus, dx , for the solution function, is equal to the required value and so the solution checks.
(The check can also be made by differentiating the solution x2 sin 3y = C implicitly with
respect to x. Try it!)
30 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
Example 2
Show that
∂f
= ex+y + 3x2 − 2y
∂x
Hence
f = ex+y + x3 − 2xy + ϕ(y)
by inspection.
Next,
∂f
= ex+y − 2x + ϕ′ (y)
∂y
and this must equal
N (x, y) = ex+y − 2x + 4.
We have
ϕ′ (y) = 4
ϕ(y) = 4y + c
Hence
f = ex+y + x3 − 2xy + 4y + c,
and the solution is
ex+y + x3 − 2xy + 4y = C.
Example 3
Solve
(cos x + y sin x)dx = cos xdy,
Now,
∂
(cos x + y sin x) = sin x
∂y
and
∂
(− cos x) = sin x
∂x
which tells us that the equation is exact.
f must be such that
∂f
= cos x + y sin x
∂x
f = sin x − y cos x + ϕ(y)
.
∂f ∂
Now = (sin x − y cos x + ϕ(y))
∂y ∂y
= − cos x + ϕ′ (y)
f = sin x − y cos x + c
sin π − 0 cos π = C.
i.e. C = 0. This particular solution is sin x − y cos x = 0. (i.e. y = tan x. Check that this is a
solution to the given D.E.)
Method 2
In solving an exact D.E. by the grouping method, we use three facts:
∂z ∂z
1. If z is a function of x alone, = 0, and dz = dx.
∂y ∂x
For example, if z = x3 , dz = 3x2 dx.
∂z ∂z
2. If z is a function of y alone, = 0, and dz = dy.
∂x ∂y
For example, if z = sin y 2 , dz = 2y cos y 2 dy.
Example 4
Solve
(2xy 3 + 6x)dx + 3x2 y 2 dy = 0.
Solution
This is an exact D.E. since
∂ ∂
(2xy 3 + 6x) = 6xy 2 = (3x2 y 2 ).
∂y ∂x
In this example, the term 6xdx cannot be grouped with any term in y. Hence, we have the
grouping:
6xdx + (2xy 3 dx + 3x2 y 2 dy) = 0.
The two terms grouped together in the above equation were paired-off because both terms
involved both x and y.
The given D.E. can be written as:
or
d(3x2 + x2 y 3 ) = 0.
Hence
3x2 + x2 y 3 = constant.
The solution of the given D.E. is
3x2 + x2 y 3 = C.
Example 5
In the second pass the x-y terms are paired-off. Which of the terms 4x3 ydy or 6x2 ydy is to
be grouped with 6x2 y 2 dx? The answer should be obvious: a term like 6x2 y 2 dx is obtained
from differentiating a function involving x3 .
This means that the terms 4x3 ydy and 6x2 y 2 dx are to be grouped together, leaving the pair
6xy 2 dx and 6x2 ydy to be grouped together.
1.3. EXACT DIFFERENTIAL EQUATIONS 33
Example 6
Solution
This looks to be a very complicated D.E. However, in trying to solve it, we en-counter no
problems in pairing-off terms. Terms involving exponentials go together, and so on, leading
to the grouping:
1 1
ex dx + 8 cos y sin ydy + (yexy dx + xexy dy) + ( dx + dy)
x+y x+y
+ (2y sin xy cos xydx + 2x sin xy cos xydy) = 0
Once the grouping has been completed, the solution is obtained by integration.
(It should be emphasised that, regardless of the method of solution used, essentially the
same integration is required to obtain the solution.)
In this case, the solution is derived as follows:
Hence,
ex + 4 sin2 y + exy + log(x + y) + sin2 xy = C
Tutorial 1 Question 2
34 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
y ′ = f (x)g(y)
or
M (x)dx + N (y)dy = 0.
Clearly, written in the last form, a separable D.E. is an exact D.E., since
∂M (x) ∂N (y)
=0= .
∂y ∂x
Example 1
Solve:
(x3 + 4)dx + sin ydy = 0.
Solution
This D.E. can be solved using grouping of terms:
d ( 41 x4 + 4x) + d(− cos y) = 0
i.e. d ( 41 x4 + 4x − cos y) = 0
1 4
∴ 4x + 4x − cos y = C, (C being constant)
is the general solution.
Note that this same result can be obtained in the following way
(x3 + 4) dx = − sin ydy
3
∴ ∫ (x + 4) dx = ∫ (− sin y)dy
1 4
x + 4x = cos y + C
4
(only one arbitrary constant is required for this first order D.E.), or
1 4
x + 4x − cos y = C.
4
Note
This second approach is usually the most suitable for separable D.E.s once they have been
properly classified, i.e. writing the D.E. in the form
H(y)dy = K(x)dx,
1.4. SEPARABLE DIFFERENTIAL EQUATIONS 35
∫ H(y)dy = ∫ K(x)dx,
Example 2
Solve
dy x
=−
dx y
subject to the initial condition y(0) = 4.
Solution
Here, the D.E. can be written as
ydy = −xdx
∴ ∫ ydy = ∫ (−x)dx
1 1
i.e. y 2 = − x2 + c
2 2
Now, when x = 0, y = 4 and so
8 = 0 + c i.e. c = 8.
The required particular solution is x2 + y 2 = 16.
Similarly, by integrating both sides of
1
dy = 2dx,
y
we get
ln y = 2x + c
i.e. y = e2x+c = e2x ec
= Ce2x as the solution of
dy
= 2y.
dx
Example 3
Solve
5
dy x2 (2x3 − 3)
=
dx (2y − 1)3
Solution
We can separate this D.E. as
5
(2y − 1)3 dy = x2 (2x3 − 3) dx
3 2 3 5
∫ (2y − 1) dy = ∫ x (2x − 3) dx
1 1 6
∴ (2y − 1)4 = (2x3 − 3) + C
8 36
36 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
The solution, y, is given here implicitly. The solution can be simplified but this is not
necessary.
For example, we could write
2 6
(2y − 1)4 = (2x3 − 3) + k
9
or even, explicitly,
√
1⎛ 4 2 6 ⎞
y= 1± (2x3 − 3) + k .
2⎝ 9 ⎠
Example 4
Solve
(x2 + 1)y ′ + y 2 + 1 = 0.
Solution
Here, y ′ has been used instead of dy
dx . We can separate the variables to get
and then
dy dx
2
=−
1+y 1 + x2
Hence, arctan y = − arctan x + c, i.e. arctan y + arctan x = c.
The solution can be simplified, if required, as follows:
tan(arctan y) + tan(arctan x)
∴ =k
1 − tan(arctan y) ⋅ tan(arctan x)
y+x
i.e. =k since (tan(arctan θ) = θ)
1 − xy
and even then
k−x
y= .
1 + kx
Note
Textbooks often put the solutions to D.E.s in the most “simplified” form. It is not neces-
sary for you to do so unless specifically required in the problem. You should concentrate
on solving the D.E. correctly, by first deciding on its type and then by carrying out the
appropriate method of solution accurately!
Tutorial 1 Question 3
1.5. INTEGRATING FACTORS 37
If a D.E. M (x, y)dx + N (x, y)dy = 0 is not exact, but we can find a function I(x, y) such
that
I(x, y)M (x, y)dx + I(x, y)N (x, y)dy = 0.
is exact, then we call I(x,y) and integrating factor for the D.E. The new equation can be
solved using the method developed in the section on Exact Differential Equations. For
a general first-order D.E., it may be difficult (or impossible) to find an integrating factor.
However, for the important class of linear first-order D.E.s, an integrating factor can always
be found in a standard way.
Example
Consider
−y dy −y
( )+ = 0, i.e. ( ) dx + dy = 0.
x dx x
This equation is not exact, since
∂ −y −1 ∂
( )= while (1) = 0.
∂y x x ∂x
However, if each term in the equation is multiplied by 1/x, the equation becomes
−y 1
( ) dx + ( ) dy = 0.
x2 x
This D.E. is exact since
∂ −y −1 ∂ 1 −1
( 2 ) = 2 and ( )= 2.
∂y x x ∂x x x
Hence, to solve the new D.E., we must find f such that
∂f −y y
= 2 , i.e. f = + ϕ(y).
∂x x x
Then,
∂f 1
= + ϕ′ (y)
∂y x
which must equal 1/x and so ϕ](y) = 0, i.e. ϕ(y) = c.
y
∴ f= + c.
x
Thus the solution of the exact D.E. is y/x = C or y = Cx. Hence, the solution of the given
D.E. is also y = Cx. (check this!)
38 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
Notes
1. The equation
−y dy
( )+
x dx
can also be considered as a separable D.E., since
dy dx
=
y x
∴ ln y = ln x + c = ln x + ln C = ln Cx
∴ y = Cx again.
Method
A simplified method for finding integrating factors follows. Consider the DE:
R(y) = e∫ P dy ,
where
∂M /∂y − ∂N /∂x ∂N /∂x − ∂M /∂y
P (x) = or P (y) = .
N M
Example
Find an integrating factor for the following DE and hence find the general solution:
(1 − y)dx + (x2 y 2 + x) dy = 0.
−2 (xy 2 + 1) −2
= =
x (xy 2 + 1) x
which is indeed a function of x and we proceed:
−2
∫ P dx = ∫ dx = −2 ln x = ln (x−2 )
x
giving
−2 ) 1
R = e∫ P dx = eln(x = x−2 = .
x2
We could also look at
∂N /∂x − ∂M /∂y 2xy 2 + 1 − (−1) 2xy 2 + 2
P (y) = = =
M 1−y 1−y
which does not simplify any further nor is a function of y alone, including also x. So
we have only been able to find one integrating factor as a function of x. Sometimes a
differential equation will give two possible integrating factors, but not in this example.
Continuing we now have, using R from P (x):
1 1
2
(1 − y)dx + 2 (x2 y 2 + x) dy = 0
x x
giving
1−y 1
2
dx + (y 2 + ) dy = 0.
x x
Check that
∂M −1 ∂N 1
= 2 and =− 2
∂y x ∂x x
and the DE is now exact.
40 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
Tutorial 1 Question 4
dy d2 y dy
= 3y + sin x or − 2 + 3y = ex
dx dx2 dx
are both linear, while
dy dy 2
= 3y 2 + sin x or ( ) = 3y + ex
dx dx
1.6. FIRST ORDER LINEAR DIFFERENTIAL EQUATIONS 41
1. The method we will develop will work for any first order linear D.E.. The equation
may also be exact or separable (or some other type). In such cases we have a choice
of methods of solution!
2. In Chapter 2 we will consider second order (and higher order) linear D.E.s with
constant coefficients, which can be written in the form
d2 y dy
a2 + a1 + a0 y = F (x)
dx dx
where a0 , a1 , a2 are constants.
(Such D.E.s can be solved by a very clever “trick”!)
However, in Chapter 4 we will consider the solution of such second order linear D.E.s
where the coefficients a0 , a1 , a2 are functions of x.
(Such D.E.s can be solved by the use of series!)
y ′ + P (x)y = Q(x).
Such D.E.s are not (usually) exact, but can be made exact by multiplying the D.E. by an
appropriate integrating factor. One way to derive this integrating factor is as follows:
Multiply the D.E. by R(x), an I.F. which is assumed to depend on x alone and which must
be found, giving
Ry ′ + RP y = RQ,
where the left-hand side of the D.E. must be the derivative of Ry.
d
i.e., Ry ′ + RP y = (Ry) = Ry ′ + R′ y
dx
∴ RP y = R′ y (subtracting Ry ′ from both sides)
R′
∴ = P (x)
R
∴ ln R = ∫ P (x)dx( integrating both sides w.r.t. x)
Hence,
R = e∫ P (x)dx .
42 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
Step 3: Multiply every term in the D.E. by the integrating factor, and hence write the D.E.
in the form
d ∫ P (x)dx
(e ⋅ y) = e∫ P (x)dx ⋅ Q(x).
dx
Step 4: Integrate both sides of the D.E., i.e.
∫ d(Ry) = Ry = ∫ RQdx + C.
Example 1
Solve
dy
sin x + 4y cos x = cot x
dx
Solution
The D.E. can be re-written as
dy cos x cot x
+ (4 )y =
dx sin x sin x
which is in first order linear form.
1.6. FIRST ORDER LINEAR DIFFERENTIAL EQUATIONS 43
Step 1:
cos x 4
∫ 4 sin x dx = 4 ln sin x = ln(sin x)
Step 2:
4
eln(sin x) = (sin x)4
(Note the important use of a property of the logarithm and exponential functions.)
Step 3: By multiplying through by the I.F., the D.E. can be written as
dy cot x
(sin x)4 + (4 sin3 x cos x) y = (sin x)4 = cos x sin2 x
dx sin x
d
i.e of the form {y(sin x)4 } = cos x sin2 x
dx
Step 4:
y sin4 x = ∫ cos x sin2 xdx
1 3
= sin x + C
3
Step 5: Thus the solution is given by
1
y= csc x + C csc4 x
3
Example 2
Solve
dy
− y = ex ,
dx
subject to the condition y(0) = 2.
Solution
(This D.E. can be solved by the methods developed in Chapters 2 and 3: try it again after
completing those chapters!)
For now, we recognise that this is a linear first order D.E. with P (x) = −1 and Q(x) = ex .
Step 1: ∫ (−1)dx = −x
Step 2: I.F. is e(−x) = e−x
Step 3: (slightly modified): Multiply through by the I.F.
dy
∴ e−x − ye−x = ex ⋅ e−x = e0 = 1
dx
Now, the left-hand side, as expected, can be written as
d −x
(e ⋅ y)
dx
(Check that this is true by carrying out this differentiation!)
Therefore the D.E. is now of the form
d −x
(e ⋅ y) = 1
dx
44 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
e−x ⋅ y = ∫ dx
=x+C
Thus the general solution is given by
1
y= (x + C) = xex + Cex
e−x
Now, the particular solution required has y = 2 when x = 0.
∴ 2 = 0 ⋅ e0 + C ⋅ e0
∴ C=2
The required particular solution is:
y = xex + 2ex or y = ex (x + 2)
Tutorial 1 Question 5
y ′ + P (x)y = Q(x)y n , n ≠ 0 or 1
z = y 1−n
Example
x2
Multiplying the D.E. by this integrating factor gives
d 1 2
( 2 ⋅ z) = − 4
dx x x
1 2 −3
∴ z = x +C
x2 3
1 1 2 −3
i.e., = x +C
x2 y 2 3
Hence,
2
y −2 = x−1 + Cx2
3
is the solution of the given Bernoulli D.E.
Tutorial 1 Question 6
Example 1
(x2 + y 2 ) dx − 2xydy = 0
46 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
− ln (1 − v 2 ) = ln x + C
i.e., ln x (1 − v 2 ) = −C
x (1 − v 2 ) = (e−C ) = A, say
y2
∴ x (1 − 2 ) = A
x
2 2
Hence, x − y = Ax is the solution of the given D.E.
Example 2
Solve
xy ′ = x + y.
Solution We can write this D.E. in the form
dy y
=1+
dx x
which we can then classify as a homogeneous type.
y dy dv
Let v = , then, y = xv and = x + v.
x dx dx
The D.E. is transformed to
dv
x +v =1+v
dx
dv
∴ x =1
dx
dx
i.e. dv = , which is a separable D.E.
x
∴ v = ln ∣x∣ + C
y
i.e. = ln ∣x∣ + C
x
Thus the required solution is
y = x ln ∣x∣ + Cx.
1.9. APPLICATIONS OF FIRST ORDER DIFFERENTIAL EQUATIONS 47
Notes
2. There are many other types of D.E.s which, by an appropriate substitution as in the
Bernoulli or Homogeneous cases can be transformed into a known type of D.E. with
a “standard” method of solution. Such other transformations are beyond the scope of
this subject, but are very important in many applications of D.E.s.
Tutorial 1 Question 7
dV
= λV,
dt
where λ is some constant. This is a linear first order differential equation
dV
− λV = 0.
dt
48 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
2V1 = V0 eλT
Hence
V0 eλT = 2V0 eλt1
eλT = 2eλt1
λT = ln (2eλt1 ) = ln 2 + ln eλt1
λT = ln 2 + λt1
ln 2
T − t1 =
λ
Hence the time to double its volume (T − t1 ) is ln 2/λ.
Solid tumours, however, differ from “free-living” dividing cells in their growth. As the
tumour grows the time for its volume to double becomes longer and longer. That is, as it
gets bigger its rate of growth decreases.
Several researchers have found that the data for tumour growth is fitted extremely well by
the following exponential function:
λ
V = V0 exp { (1 − exp(−αt))}
α
where α and λ are positive constants. This is usually referred to as the Gompertzian func-
tion. You will notice that at t = 0, we have
λ λ
V = V0 exp { (1 − exp(0))} = V0 exp { (1 − 1)}
α α
= V0 exp{0} = V0
and that as t → ∞, V → V0 eλ/α . Consequently a solid tumour usually has a limiting size
(fortunately). Ordinary exponential growth has no limit as t → ∞, so how can this growth
behaviour of tumours be explained?
1.9. APPLICATIONS OF FIRST ORDER DIFFERENTIAL EQUATIONS 49
We can gain some insight into this question by finding out the differential equation which
is satisfied by the Gompertzian function. Differentiating both sides of the equation gives
dV λ d λ
= V0 exp { (1 − exp(−αt))} { (1 − exp(−αt))}
dt α dt α
λ λαe−αt
= V0 exp { (1 − exp(−αt))} ⋅ [0 + ]
α α
λ
= λe−αt V0 exp { (1 − exp(−αt))}
α
and so
dV
= λe−αt V
dt
is the differential equation governing tumour growth. Notice that at t = 0 we have dV
dt = λV ,
dV −αt
a simple exponential growth model, but as t → ∞, we have dt → 0 (since e → 0) which
reflects the fact that the rate of growth is slowing down.
Two different theories have been put forward to explain the dynamics of tumour growth. It
has been suggested that the retarding of tumour growth is due to either:
i. decrease in the reproductive rate of the cells, the proportion of reproducing cells re-
maining constant, and
ii. a drop in the proportion of reproducing cells, the reproductive rate remaining constant.
p is a continuous (and differentiable) function of t so that the rate of change dp/dt has
meaning. In the simplest case, if the growth rate, a, is constant for each individual (i.e.
each individual reproduces a new individuals per unit time), then the differential equation
for population growth is simply
dp
= ap
dt
where a is a constant and p = p(t). This is a first order linear differential equation as seen
in the last section and so if p(t0 ) = p0 we can solve to get the solution
p(t) = p0 ea(t−t0 ) .
The above model for population growth was postulated by the English economist Malthus
(1766-1834). His model has been applied to human populations to explain the rapid in-
crease over recent years. The population of the world has been increasing at about 2%
per year and so the time to double can be determined from the above equation. In fact if
p(t) = 2p0 , then
p(t) = 2p0 = p0 ea(t−t0 ) , where a = 0.02
Hence
e0.02(t−t0 ) = 2
0.02 (t − t0 ) = ln 2
and
t − t0 = ln 2/0 ⋅ 02 ≈ 34.7 years.
In fact it has been observed that the world’s population has been doubling every 35 years.
The model, however, cannot be realistic in the long term.
Example
The population of the world in 1976 was 4 billion. Given a growth rate of 2%, find the
population predicted by the Malthusian model in the years 2000 and 2510 and 2670. Fur-
thermore, if the area of the earth’s surface is 170 000 billion square metres, how much room
will people have in these years?
Solution
dp
The equation dt = ap has solution
p = p0 ea(t−t0 )
Hence
i. t0 = 1976, t = 2000, a = 0.02, p0 = 4 billion
p = 4e0.02(2000−1976) = 4e0.48 ≈ 6.5 billion.
Each person then has 170000 ÷ 6.5 = 26000 square metres of room.
Each person would then have 170000 ÷ 174000 ≈ 1 square metre of room (a bit cramped!).
How much room then? We would all be standing at least two deep on each other’s shoul-
ders.
1.9. APPLICATIONS OF FIRST ORDER DIFFERENTIAL EQUATIONS 51
So Malthus’ model is obviously not accurate for large values of p. Should we then discard
it? Consider the next example taken from an actual experiment.
Example
A certain species of rodent has a reproduction rate of 40% per month. Initially there were
2 rodents present (one of each sex) and after 2, 6 and 10 months respectively there were
observed 5, 20 and 109 rodents present.
How well does this compare with the Malthus model?
Solution
dp
Since dt = 0.4p, p(0) = 2 we have the solution p = 2e0.4t where t is in months.
As you can easily verify, at t = 2, 6 and 10 respectively, the values of p are 4.5, 22.0 and
109.2, which shows close agreement with the predictions of the Malthus model. Hence
the Malthus model is accurate in certain cases. In fact it seems to be very useful provided
that the population p does not get too large. Obviously as p increases, other factors such
as competition for resources become important and the model is then too simple to explain
actual growth behaviour. This will be considered in the next section.
The logistic equation (Getting more complicated!)
The Dutch biomathematician Verhulst introduced in 1837 an improved version of the pop-
ulation model which included a competition term. He suggested the equation
dp
= ap − bp2 = (a − bp)p
dt
where a is still the birth rate per individual and where bp represents the deathrate per in-
dividual, b being a constant. You will notice that as p increases, the deathrate increases in
proportion, which is much as you might expect due to competition, finite resources, limited
food supply and so on. The term −bp2 on the right-hand side has a levelling out effect on
the solution - as we shall see. In practice, the constant b is small compared with a. Also
when p is small, the term −bp2 will be negligible compared with ap, and so the equation
resembles the Malthus model. However, when p is large, the term −bp2 will be quite signif-
icant and serves to slow down the growth in population. The above equation is called the
Logistic Equation.
Fortunately, we can even solve the Logistic equation, since it is of separable type.
Separating the variables we get
dp
= dt
p(a − bp)
so that
dp
∫ p(a − bp) = ∫ dt
1
To integrate p(a−bp) we use partial fractions.
Let
1 A B
= +
p(a − bp) p a − bp
so that 1 = A(a − bp) + Bp.
At p = 0 we have 1 = Aa, giving A = 1/a.
52 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
dp dp
(Note: Since is the total rate of change of the population, then ÷ p is the rate of
dt dt
change per individual which in this case is 2%.
Hence from above, the limiting population
a 0.029
= ≈ 10 billion.
b 2.94 × 10−12
Verhulst applied his methods to a study of populations in European countries with some
interesting results. In 1845 he predicted a maximum population of 6,600,000 for Belgium
and a maximum population of 40,000,000 for France. By 1930, the population of France
was steady around this figure, but Belgium already had over 8 million people. Perhaps the
answer is that Belgium had the Congo and at the same time was enjoying a remarkable rise
in industry. This allowed them to support a much greater population.
Example
In 1790 the population of the U.S. was 3.9 million, the birth rate was 3.134% and the
competition constant b was 1.5887 × 10−10 . The population in 1800 was in fact 5.308
million. How close is this to the figure predicted by the logistic model?
Solution
The logistic function found above is
ap0 a
p= =
bp0 + (a − bp0 ) e−at b + (ap−1
0 − b) e
−at
(0.03134)
p=
(1.5887)10−10 0.03134
+ [( 3.9×106 ) − 1.5887 × 10
−10 ] e−0.03134t
Note At this point, ignore any reference in the text to second order DEs. We will cover
these in the next chapter. Preliminary electrical theory In this section we apply some of
the methods for solving differential equations (as studies in earlier chapters) to the solution
of some electrical circuit problems. The notes are based partly on material from the book
Applied Differential Equations by M. Spiegel, Prentice-Hall, 1967, which incidentally is an
excellent reference book for you if you wish to brush-up on standard methods for solving
differential equations.
The current I in an electrical circuit is said to be the time rate of change of the quantity
of electricity Q (or charge) flowing from one part of the circuit to another. Hence we have
dQ
I= .
dt
A very simple electric circuit would consist of a source of electromotive force, E (such
as a battery or generator), and a resistor R (such as an electric light element or heating
element or electric motor. The electromotive force (or voltage) produces the current I by
forcing charge Q to flow in a similar manner to the way in which a pump causes water to
flow in a circuit of pipe. The diagram below illustrates a simple circuit,
the direction of the current flow I being as shown. The symbols shown on the circuit are
used for resistors R and sources of electromotive force E. Note that the wire is assumed to
have no resistance. If it had any at all we could include it in R.
Two other electrical elements will be considered, the inductor and the capacitor. The in-
ductor L is essentially a coil of wire which has the property that it opposes any change in
current I. If a steady current is flowing around the circuit, the inductor will have no effect.
But if the current increases or decreases, the inductor will resist such change. Hence it can
be seen that the inductor has an inertia effect on current - much like trying to speed up or
slow down a moving train, the mass of the train and its inertia resists such a change.
The capacitor C stores electrical energy (or charge) in the circuit. A capacitor consists of
parallel plates separated by an insulating medium. Charge then collects on the plates and
can be stored.
1.9. APPLICATIONS OF FIRST ORDER DIFFERENTIAL EQUATIONS 55
The circuit below shows the symbols used for all of these elements:
We can measure the voltage drop (like a change in pressure) across any of these elements
using voltmeter. It is found that the following three laws can be verified by experiment:
1. Voltage drop across a resistor Er is directly proportional to the current I passing through
the resistor. Hence Er ∝ I and defining the constant of proportionality to be R - the
resistance, we get
Er = RI.
2. Voltage drop across an inductor El is directly proportional to the rate of change of the
current I.
Hence El ∝ dIdt
and defining the constant of proportionality to be L - the inductance, we get
dI
El = L
dt
3. Voltage drop across a capacitor Ec is directly proportional to the charge Q on the capac-
itor.
Hence Ec ∝ Q and defining the constant of proportionality to be 1/C - where C is called
the capacitance, we get
Q
Ec = .
C
• Time - t - seconds
• Charge - Q - coulombs
• Current - I - amps
• Resistance - R - ohms
• Inductance - L - henrys
• Capacitance - C - farads.
56 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
Kirchoff’s Law
This states that the algebraic sum of all the voltage drops around a closed circuit is
zero.
This is just another way of saying that the voltage supplied by the source of electromotive
force (e.g. the battery) is equal to the sum of the voltage drops in the rest of the circuit.
We will use Kirchoff’s Law in order to establish a mathematical formulation for certain
types of circuits. In fact we will get differential equations which can be solved for the
current flowing at any particular time.
Note that in the next section you will not be required to carry out all the details of sketching
circuits and setting up the D.E.s. However, you must be able to solve the resulting D.E.
First order linear differential equations for RL and RC circuits We will take the ex-
ample of the RL circuit first of all to illustrate the use of Kirchoff’s Law. The diagram of
the circuit is shown below.
Now by Kirchoff’s Law, the voltage supplied E is equal to the voltage drop across the
resistor (Er = RI) plus the voltage drop across the inductor
dI
El = L
dt
Hence,
El + Er = E
giving
dI
L + RI = E
dt
the corresponding differential equation for the RL circuit. This is an example of the well-
known linear first order differential equation.
Example 1
A battery with electromotive force 100 volts is connected in series with a 20 ohm resistor
and an inductor of 4 henrys. The switch is closed at time t = 0 and the current starts to flow.
Sketch the circuit, set up the differential equation for the current and solve it for the current
at any time t.
1.9. APPLICATIONS OF FIRST ORDER DIFFERENTIAL EQUATIONS 57
Solution
e5t I = 5 (e5t − 1)
and so
I = 5 (1 − e−5t ) .
Notice that at t = 0, I = 0 and as t → ∞, I → 5. A sketch of the solution is shown below.
Hence the current starts off at zero and quickly builds up towards its limiting value of 5
amps.
Note
58 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
Since the linear differential equation had constant coefficients, it could have been solved
by the method of separation of variables.
Example 2
Now integrating by parts twice the right-hand side gives (u = 5e5t , v ′ = sin 5t)
5t 5t 5t
∫ 5e sin 5tdt = −e cos 5t + ∫ 5e cos 5tdt
(u = 5e5t , v ′ = cos 5t)
5t 5t 5t 5t
∫ 5e sin 5tdt = −e cos 5t + [e sin 5t − ∫ 5e sin 5tdt]
Hence
5t 5t sin 5t − cos 5t
∫ 5e sin 5tdt = e [ 2
]+K
By Kirchoff’s Law, the voltage supplied E is equal to the voltage drop across the resistor
(Er = RI) plus the voltage drop across the capacitor (Ec = Q
C ). Hence Er + Ec = E, giving
Q
RI + = E.
C
dQ
Now the current I is defined to be the rate of change of charge dt , so that
dQ 1
R + ⋅Q=E
dt C
a first order linear differential equation, this time in terms of Q instead of I.
Example 3
e5t Q = 10t + K
At t = 0, Q = 0 so that K = 0.
Hence
Q = 10te−5t
the charge at any time t as represented by the sketch below.
Now
dQ
= 10e−5t (1 − 5t)
dt
which is zero when t = 51 seconds. (You should show that this is a maximum and not a
minimum, but that is fairly obvious anyway.) At t = 15 , Q = (10) ( 51 ) e−1 ≈ 0.74 coulombs.
It is interesting to compare the two curves for charge and current. If we sketch
Q = 10te−5t , and
I = 10e−5t (1 − 5t), we get:
It can be seen that at t = 0 the current is high, so the charge starts to quickly build up on
the capacitor. When the current becomes negative (reverses direction) the charge will of
1.9. APPLICATIONS OF FIRST ORDER DIFFERENTIAL EQUATIONS 61
course decrease on the plate of the capacitor and eventually will discharge completely as
there is no current remaining in the circuit (the battery has gone flat!)
Tutorial 1 Question 8
62 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
1.10 Tutorial 1
Back to Chapter 1
Question 1
∂z ∂z
In each of the following, find and :
∂x ∂y
i. z = xy
ii. z = x2 y + ln y
iii. z = (x2 + y 2 )3
Answers:
∂z ∂z
i. = y, =x
∂x ∂y
∂z ∂z 1
ii. = 2xy, = x2 +
∂x ∂y y
∂z 2 ∂z 2
iii. = 6x (x2 + y 2 ) ; = 6y (x2 + y 2 )
∂x ∂y
∂z ∂z 1
iv. = cos x cos y + xey , = − sin x sin y + x2 ey
∂x ∂y 2
Back to Chapter 1
Question 2
Part 1
Test for each of the following D.E.s for exactness, and solve, if possible:
i. 3xdx + 4ydy = 0
x−y
ii. y ′ =
x+y
x
iii. y ′ =
x+y
iv. (ye−x − sin x) dx − (e−x + 2y) dy = 0
Part 2
Solve each of the following, subject to the given condition:
Part 3
Show that the D.E. ydx + (4x − y 2 )dy = 0 is not exact, but becomes exact in multiplying
every term by the integrating factor, y 3 . Hence solve the D.E.
Short answers:
Pt 1 i. 3x2 + 4y 2 = C
ii. x2 − 2xy − y 2 = C
iii. Not exact
iv. ye−x − cos x + y 2 = C
v. Not exact
Pt 2 i. x2 y + y − 6 = 0
ii. x2 y + y 2 e2x − 1 = 0
Pt 3 6xy 4 − y 6 = C
Question 3
i. 3x (y 2 + 1) dx + y (x2 + 2) dy = 0
Short answers:
64 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
3
i. (x2 + 2) (y 2 + 1) = C
ii. 2e3x + 3 ln y = C
iii. x2 − 4 ln (1 + y 2 ) = 1
iv. ln x − arctan y = C
1
v. − y 2 − 2 ln y = C.
x2
vi. see textbook or detailed solutions
Question 4
Question 5
dy
i. x + 5y − 2x2 + 7 = 0
dx
ii. dy + 2x(y − 1)dx = 0
iii. xy ′ + 3y = x2
Short answers:
2 7
i. y = x2 − + Cx−5
7 5
ii. y = 1 + Ce−x
2
(Also solve this as a separable D.E.)
1.10. TUTORIAL 1 65
iii. x5 − 5x3 y = C
Question 6
ii. Textbook exercise set 2.6. Problems 21, 23, 25, 27.
Short answers
i.
1
z=
1 − x + Ce−x
ii. see textbook or detailed solutions
Question 7
dz z z 2
i. = + z(1) = 1
dx x x2
ii. xz ′ = 2x + 3z
2
′
z + x (cos xz ) π
iii. z = z(1) =
x 4
Short answers:
x
i. z =
1 − ln x
ii. z = Cx3 − x (can be solved as linear )
66 CHAPTER 1. FIRST ORDER DIFFERENTIAL EQUATIONS
z
iii. 1 + ln x = tan
x
Question 8
Chapter objectives
• find the particular solution yp for LDEs using various methods; and
Chapter summary
67
68 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
where an (x), an−1 (x), . . . , a1 (x), a0 (x) and F (x) are functions of x alone.
The D.E. is of nth order since the highest derivative of y is the nth derivative (that is, y (n)
or dn y/dxn . It is a linear D.E. since the function y and/or its derivatives are not multiplied
together (that is, no term of the form yy ′′′ , (y ′ ) , y ′ y ′′ , y 4 , etc. occurs). Non-linear D.E.s
2
can be very difficult to solve or may only yield to numerical methods beyond the scope of
this subject.
Examples of linear D.E.s
Note
Example 1 is beyond the scope of this subject; examples 2 and 3 can be solved using the
methods developed in this Chapter (and by a number of other important methods beyond
the scope of this subject); and example 4 is a type that will be dealt with in Chapter 5.
Examples 2 and 3 are of a type that is generally easier to deal with because they are Linear
D.E.s with constant coefficients, i.e. the ak ’s are all constants.
Examples of non-linear D.E.s
1. Such linear D.E.s (L.D.E.s) can be written more conveniently using the differential
d d2
operator, D. Thus, D, is used for ; D2 for 2
and, in general, Dn is used for
dx dx
dn
. Hence the L.D.E. y ′′ + 3y ′ + 4y = sin x can be written as
dxn
D2 y + 3Dy + 4y = sin x
or
(D2 + 3D + 4) y = sin x
or
ϕ(D)y = sin x, where ϕ(D) = D2 + 3D + 4
2. We will not make great use of the linear operator ϕ(D) in this subject. However, it is
a very important concept in the general study of D.E.s, and opens up a whole range
of methods for the solution of L.D.E.s.
ϕ(D)y = F (x)
Note that this is where the “double use” of the word “homogeneous” occurs! (In the previ-
ous Chapter we considered homogeneous first order D.E.s which were non-linear.)
The study notes will use the term “homogeneous” here too, as this is the term commonly
used in texts. There should (hopefully) be no confusion as the context should always make
the usage clear!
General and particular solutions
The method of approaching the solution of L.D.E.s depends on a crucial theorem:
Consider a L.D.E. of the form
ϕ(D)y = F (x)
Theorem
The general solution of a L.D.E. is given by y = yh + yp , where yp is a particular solution of
the given D.E. (i.e. does not involve any arbitrary constants), and yh is the general solution
of the corresponding (reduced) homogeneous D.E. ϕ(D)y = 0.
70 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
Notes
1. Of course, if the given L.D.E. is, in fact, homogeneous, then we need only to find yh ,
which is the required general solution. (Here yp = 0, in fact. Why?)
2. Clearly there are two major problems confronting us:
• How do we find yh ?
• How do we find yp ?
In the case of a linear D.E. with constant coefficients, i.e. where ϕ(D) is a polynomial,
then there is a very simple standard procedure for finding yh . (See the following section.)
However, for linear D.E.s with non-constant coefficients (i.e. where the coefficients are
functions of x, ak (x)) , then the procedure for finding yh is more complicated. (We will
examine some “elementary” cases in Chapter 5.)
The methods for finding a particular solution yp are numerous and often specialised, de-
pending on the form of the function F (x) on the right-hand side of the D.E.
Notes
In this Chapter, we will consider only three methods:
(In Chapter 6 we will see how Laplace Transforms can be used to transform a given differ-
ential equation to an algebraic equation which can then be solved by alternate methods.)
Read textbook sections 4.2 (Homogeneous linear equations: the general solu-
tion), 6.2 (Homogeneous linear equations with constant coefficients)
Introduction
The solution of a homogeneous L.D.E., ϕ(D)y = 0, is aided by the following two theorems:
1. If y1 (x) and y2 (x) are solutions of this D.E., then so is the linear combination
y(x) = c1 y1 (x) + c2 y2 (x)
2. The solution to any homogeneous linear D.E. of order n consists of a linear combi-
nation of n linearly independent solutions y1 (x), y2 (x), . . . , yn (x) i.e. any solution
of ϕ(D)y = 0 (of order n ) is of the form
y(x) = c1 y1 (x) + c2 y2 (x) + . . . + cn yn (x)
where c1 , c2 , . . . , cn are arbitrary constants.
2.3. HOMOGENEOUS SOLUTION 71
Second order L.D.E. Because of the Theorem above, the problem of finding the solution
of a homogeneous L.D.E. has been reduced to the problem of finding the correct number of
linearly independent solutions. The general solution is immediately a linear combination
of these solutions. We will now develop the method for finding such solutions for L.D.E.s
with constant coefficients.
Note
We will concentrate on the case of a second-order L.D.E. with constant coefficients, but the
method is entirely general.
Consider the homogeneous L.D.E.
dn y dy
an + . . . + a 1 + a0 y = 0
dxn dx
i.e.
(an Dn + . . . + a1 D + a0 ) y = 0
where an , . . . , a1 , a0 are fixed constants.
The general solution to such a D.E. is found using the following steps.
an rn + . . . + a1 r + a0 = 0
Note
To solve homogeneous L.D.E.s with constant coefficients, you must learn these three steps
and the three Cases. (The three “rules” in these cases mathematically guarantee the re-
quired linear independence. The only real difficulty is finding the roots of the characteristic
equation!)
72 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
Example 1
Solve
y ′′ − 3y ′ + 2y = 0
Solution
This linear D.E. can be written in the form
(D2 − 3D + 2) y = 0
∴ Let r2 − 3r + 2 = 0 i.e. (r − 2)(r − 1) = 0, and so the roots are r = 2 and r = 1, which are
distinct (Case I).
∴ The required general solution is given by
y = c1 e2x + c2 ex
(Check by differentiation that this is indeed the solution of the given D.E.!)
Example 2
Solve
d3 y d2 y dy
3
− 4 2
+ 5 − 2y = 0
dx dx dx
Solution
The operator form is usually more “helpful”: i.e.
(D3 − 4D2 + 5D − 2) y = 0
∴ Let r3 − 4r2 + 5r − 2 = 0. After some effort, we find that the roots are r = 1, 1, 2 since the
equation can be written as (r − 2)(r − 1)2 = 0
∴ Using Cases I and II, the general solution of this D.E. is given by
Example 3
Solve
(D2 + 2D + 5) y = 0
Solution
Here we let r2 + 2r + 5 = 0. The roots of this equation are
√
−2 ± −16
r= = −1 ± 2i (Case III)
2
∴ The general solution of the D.E. is given by
Example 4
2.4. METHOD OF UNDETERMINED COEFFICIENTS 73
Solve
(D4 − 5D2 + 12D + 28) y = 0, given that
(D4 − 5D2 + 12D + 28) = (D + 2)2 (D2 − 4D + 7)
Solution
Here the characteristic equation is
Tutorial 2 Question 1
Note
For any F (x) made up of a sum (or product) of the above, just try for yp the corresponding
sum (or product) of the yp forms e.g. if F (x) = (x2 − 4) e3x , try yp = (ax2 + bx + c) e3x ,
and try to find a, b, c.
Note that a trial yp = (ax2 + bx + c) dex is superfluous, since dex is already covered in the
first trial. Note further that, if the trial yp “fails to work”, i.e. if you cannot find the values
of the undermined coefficients, then it may be necessary to “adjust” yp , as shown later.
Example 1
y = yp + yh
9 3x
i.e. y = A cos 2x + B sin 2x + 13 e
(As “required”, this solution of the given second order L.D.E contains two arbitrary con-
stants and satisfies the given D.E. Check!)
Example 2
Find yp for
(D2 + 2D − 3) y = x2 + 4
Solution
Here F (x) is a quadratic polynomial. Clearly polynomials when differentiated yield poly-
nomials. In general (apart from some “exceptions”), when F (x) is a polynomial of degree
k, we assume that yp is a polynomial of degree k whose coefficients must be determined.
∴ Here, let yp = ax2 + bx + c
∴ y ′ = 2ax + b
and y ′′ = 2a.
Now substituting in the given D.E., y ′′ + 2y ′ − 3y = x2 + 4, we get
1 4 50
y p = − x2 − x − .
3 9 27
Example 3
Solve
y ′′ + 2y ′ + y = 2 cos 2x
Solution
Here F (x) = 2 cos 2x, and so we assume a form of a sin 2x + b cos 2x for yp
−3a − 4b = 0
{
4a − 3b = 2
∴ yh = (c1 + c2 x) e−x
Example 4
Find yp for
y ′′ + 3y ′ + 2y = 4e−2x
Solution
Here, we note the form of the right-hand side as 4e−2x , and so we try letting
yp = ae−2x
or
0 = 4e−2x
an “impossible” situation since e−2x is never equal to zero! This is an “exceptional” case.
The problem is that the form we have assumed for yp is, in fact, part of yh
In fact, letting y ′′ + 3y ′ + 2y = 0, and hence solving r2 + 3r + 2 = 0, gives r = − 1,-2 and so
yh = c1 e−x + c2 e−2x
Since there is already a term with e−2x in the homogeneous part of the solution, we cannot
use this form in yp .
2.4. METHOD OF UNDETERMINED COEFFICIENTS 77
Note
The trick in such situations is to multiply the “normal” assumed term by a power of x which
is sufficiently high (but not higher) so that none of the “new” assumed terms for yp are in
yh .
Hence, here yp = x (ae−2x ) should work!
∴ y = axe−2x , y ′ = ae−2x − 2axe−2x
y ′′ = −2ae−2x − 2ae−2x + 4axe−2x = −4ae−2x + 4axe−2x
Now, on substituting in the given D.E. we get
(−4ae−2x + 4axe−2x ) + 3 (a−2x − 2axe−2x ) + 2axe−2x = 4e−2x
∴ − ae−2x = 4e−2x
Hence, a = −4 ∴ yp = −4xe−2x , and the general solution is given by
y = c1 e−x + c2 e−2x − 4xe−2x .
Notes
1. These exceptional cases can cause some problems, but these can usually be overcome
by some “experimentation”, using the given trick until no “contradiction” is reached.
It is important always to find yh first since its form, as noted above, can influence our
choice for yp .
2. You will not be required to carry out very “complicated” solutions using this method
- it is the “trial-and-error” approach that is worth noting as a technique! In any such
case, you just adjust the usual “yp ” by multiplying it by an appropriate power of x so
that no term in yh appears in the new yp , e.g. the yp for y ′′ − 2y ′ + y = ex needs to be
ax2 ex (Try it!).
Example 5
Hence
1 1
yp = x ( x2 − x + 1) = x3 − x2 + x.
3 3
The general solution is
1
y = yh + yp = A + Be−x + x3 − x2 + x.
3
Example 6
Tutorial 2 Question 2
2.5. METHOD OF VARIATION OF PARAMETERS 79
Notes
2. The disadvantage with this method of Lagrange is that it requires the finding of cer-
tain integrals which may, in practice, be quite difficult!
3. No justification for the method will be given or required: a proof will be found in any
standard text on D.E.s. You need to concentrate on the technique, and must learn the
required formulae, as detailed below.
a2 y ′′ + a1 y ′ + a0 y = R(x)
Step 1: Find the solution of the reduced homogeneous D.E., i.e. find yh for the D.E.
a2 y ′′ + a1 y ′ + a0 y = 0.
Hence, yh = Ay1 (x) + By2 (x) as usual, where A and B are the two arbitrary
constants. For simplification we will use u = y1 (x) and v = y2 (x).
Step 2: Lagrange’s “trick” now lies in the method for finding yp . Following him we assume
that yp is of the form
yp = A(x)u(x) + B(x)v(x),
80 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
where u and v are the functions found in yh in Step 1, and A(x) and B(x) are
functions of x which must be found by solving simultaneously the two equations
A′ u + B ′ v = 0
{
A′ u′ + B ′ v ′ = aR2
Note We have written A′ , y1 , etc. for convenience, instead of A′ (x), y1 (x) etc.
Step 3: The above equations, when solved, actually yield A′ (x) and B ′ (x). A(x) and
B(x) must then be found by integration.
Note. Finding A′ (x) and B ′ (x) can be simplified using determinants.
For the two equations, using determinants we can write
0 v u 0
∣ ∣ ∣ ∣
R/a2 v ′ ′
u R/a2
A′ (x) = and B ′ (x) =
u v u v
∣ ∣ ∣ ∣
u′ v ′ u′ v ′
The determinant in the denominator need only be evaluated once and this also
simplifies the working.
yp = A(x)y1 + B(x)y2
y = yh + yp
i.e
y = (Ay1 + By2 ) + (A1 (x)y1 + B1 (x)y2 )
where A and B are the 2 arbitrary constants for the given 2 nd order D.E and y1 , y2
A(x), B(x) are all functions which have been found.
Example 1
Solve
y ′′ − y = xex
Solution.
Here the homogeneous equation y ′′ − y = 0 yields yh = Aex + Be−x . Thus, if we were to
use the method of undetermined coefficients to find yp , we would have to make certain
“adjustments” as per the exceptions in the previous section. (Why?) To use the variation of
parameters method we let
yp = A(x)ex + B(x)e−x = Au + Bv
A′ ex + B ′ e−x = 0
{ xex
A′ ⋅ (ex ) + B ′ ⋅ (−e−x ) = 1
2.5. METHOD OF VARIATION OF PARAMETERS 81
∴ 2A′ ex = xex
1
∴ A′ = x
2
1
and so A(x) = x2
4
Note that no arbitrary constant is required here since we are finding a particular solution.
We can find B ′ similarly by eliminating A′ from the 2 equations or by substituting for A′
in the first equation, say. i.e.,
1
( x) ex + B ′ e−x = 0
2
1
∴ B ′ = − xe2x
2
and so
1
B(x) = ∫ (− xe2x ) dx
2
which can be found by integration by parts. i.e.
1 1
B(x) = − xe2x + e2x
4 8
1 1 1
∴ yp = ( x2 ) ex + (− xe2x + e2x ) e−x
4 4 8
or
1 1 1
yp = x2 ex − xex + ex .
4 4 8
Alternatively using determinants to find A(x) and B(x) we have u = ex , v = e−x , R = xex ,
and a2 = 1 giving
u v ex e−x
∣ ∣ = ∣ ∣ = ex ⋅ (−e−x ) − ex ⋅ (e−x ) = −ex−x − ex−x
u′ v ′ ex −e−x
= −e0 − e0 = −1 − 1 = −2
Then
0 v 0 e−x
∣ ∣ ∣ ∣
R/a2 v ′ xex /1 −e−x 0 − xex ⋅ e−x −x x
A′ (x) = = = = =
−2 −2 −2 −2 2
and
u 0 ex 0
∣ ′ ∣ ∣ ∣
u R/a2 ex xex ex ⋅ xex − 0 −xe2x
B ′ (x) = = == =
−2 −2 −2 2
as before.
The general solution to the given D.E. is
1 1 1
y = Aex + Be−x + x2 ex − xex + ex
4 4 8
or
1 1
y = Cex + Be−x + x2 ex − xex .
4 4
82 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
You should convince yourself that this method really works by checking this solution in the
given D.E. y ′′ − y = xex , and by finding yp by the method of undetermined coefficients.
Example 2
Solve
y ′′ + y = sec x.
(Note that this D.E. cannot be solved using the method of undetermined coefficients.)
Solution (Outline of steps)
1.
yh = A sin x + B cos x
2. Solve
′ ′
⎪ A sin x + B cos x = 0
⎧
⎪
⎨ ′ 1
⎪
⎪ A cos x − B ′ sin x = sec x =
⎩ cos x
∴ A′ = 1
and so A(x) = x.
(Hint: multiply the first equation by sin x and the second by cos x and add. This is
an important “trick” for such equations!)
Similarly B ′ = − tan x and so
For most practical purposes the method of undetermined coefficients supplemented by the
method of variation of parameters in cases where the former is inapplicable are enough to
solve any linear equation with constant coefficients.
However, there are a number of other methods, some of which are quite ingenious, or based
on tables of “formulas”, and so on. One particularly “elegant” method involves the use of
D-operator formulas and theorems.
We give a very brief introduction to this method in the next section. The textbook does not
cover this method in full detail, so the main source of information for this method is the
study guide.
Tutorial 2 Question 3
2.6. D-OPERATORS 83
2.6 D-operators
Video (by Dr. R. Wood): D-operators. Part 1
The concept of a differential operator is new for most of you. So, it is recommended to
spend more time on this section paying special attention to the examples.
D-operators can be used to obtain particular solutions of linear D.E.s with constant coeffi-
cients. A linear D.E. can be written in the form
ϕ(D)y = f (x),
where a0 , a1 , . . . , an are constants. Since a particular solution satisfies the given D.E., we
can write
ϕ(D)yp = f (x)
This equation can then be “solved” to give
1
yp = f (x),
ϕ(D)
1
where is an “operator” which operates on f (x). At this stage, it is not clear what
ϕ(D)
1
operator does, so we should treat carefully. However, if f (x) is a particular type of
ϕ(D)
1
function, some specific rules can be determined for the operation f (x).
ϕ(D)
1
The operator ϕ(D) is called inverse for ϕ(D). These operators “cancel” the effects of each
other, that is
1
ϕ(D) f (x) = f (x)
ϕ(D)
or
1
ϕ(D)f (x) = f (x).
ϕ(D)
Let us study some of the properties of differential operators. Suppose f (x) = emx , then
d mx
Df (x) = Demx = (e ) = memx
dx
d2 mx
D2 f (x) = D2 emx = (e ) = m2 emx
dx2
⋮
d n
Dn f (x) = Dn emx = n (emx ) = mn emx
dx
84 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
Example 1
Find
(2D3 − 3D + 4) e2x .
Here,
(2D3 − 3D + 4) e2x = 2D3 e2x − 3De2x + 4e2x
= 2 ⋅ (23 ) e2x − 3(2)e2x + 4e2x
= (2 ⋅ (23 ) − 3(2) + 4) e2x
= ϕ(2)e2x ,
where ϕ(m) = 2m3 − 3m + 4.
Note that ϕ(D)e2x = ϕ(2)e2x , where ϕ(D) is the operator 2D3 − 3D + 4 and ϕ(2) is the
polynomial 2 (23 ) − 3(2)(= 14). Hence, if ϕ(D) is a polynomial in D, then
ϕ(D)emx = ϕ(m)emx
Verify this result for (D2 + 4D + 3) e2x . It should be observed that ϕ(m) is a constant, and
not an operator.
Example 2
Determine
ϕ(D)e−2x if ϕ(D) = 3D2 + 2D + 5.
Here,
ϕ(D)e−2x = (3D2 + 2D + 5) e−2x
= 3D2 e−2x + 2De−2x + 5e−2x
= 3(−2)2 e−2x + 2(−2)e−2x + 5e−2x
= 13e−2x
= ϕ(−2)e−2x since ϕ(−2) = 13.
Now, since ϕ(D)emx = ϕ(m)emx
1 1
ϕ(D)emx = ϕ(m)emx
ϕ(D) ϕ(D)
1 mx
emx = ϕ(m) e
ϕ(D)
(since ϕ(m) is a constant ) or
1 mx 1 mx
e = e ,
ϕ(m) ϕ(D)
where we assumed ϕ(m) ≠ 0.
This result is used in obtaining a particular solution to D.E.s of the form ϕ(D)y = emx .
Example 3
y ′′ + 2y ′ + 3y = e−x .
2.6. D-OPERATORS 85
Example 4
Example 5
y ′′ + 2y ′ − 8y = e2x
Now
1 2x
yp = e ; where ϕ(D) = D2 + 2D − 8
ϕ(D)
1 2x
∴ yp = e ; but ϕ(2) = 22 + 2 ⋅ 2 − 8 = 0
ϕ(2)
1
Hence is undefined! Hence, the result
ϕ(2)
1 mx 1 mx
e = e
ϕ(D) ϕ(m)
needs modification. The statement should be
1 mx 1 mx
e = e provided ϕ(m) ≠ 0.
ϕ(D) ϕ(m)
Example 5 is typical of those problems in which an alternative method of solution is needed,
as we have seen earlier. (“Adjustment” is needed here as e2x is part of the homogeneous
solution, yh i.e. Example 5 is an example of:
“Determine a particular solution yp of the D.E. ϕ(D)y = f (x), where f (x) is a part of the
solution of the homogeneous D.E. ϕ(D)y = 0.”
For example, if ϕ(D)y = emx and ϕ(D)emx = 0 (that is, emx is a solution of the homo-
geneous D.E.), then ϕ(m) = 0. This type of problem can only be solved using advanced
86 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
techniques based on the “Shift Theorem”, covered later. (We had the same “problem” of
adjustment previously, as in Example 4 of the last section.)
Note
The D -operator method can also be used to find a particular solution of the D.E.
ϕ(D)y = f (x),
where f (x) is sin mx or cos mx. Since eiθ = cos θ + i sin θ (known fact from Algebra!) then
Hence,
cos mx = Re (eimx ) ; (the real part of eimx )
and
sin mx = Im (eimx ) ; ( the imaginary part of eimx ) .
Example 6
y ′′ + 2y ′ − 4y = cos 2x
Here,
1
yp = cos 2x
D2
+ 2D − 4
1
= 2 Re (e2ix )
D + 2D − 4
1
= Re { 2 e2ix }
D + 2D − 4
1
= Re {e2ix 2
} (by our earlier work)
(2i) + 2(2i) − 4
−8 − 4i
= Re {e2ix } (complex number arithmetic)
(−8 + 4i)(−8 − 4i)
−8 − 4i
= Re {(cos 2x + i sin 2x) ( )}
80
1 i
= Re { (−8 cos 2x + 4 sin 2x) + (−8 sin 2x − 4 cos 2x)}
80 80
and taking the real part,
1
yp = (−8 cos 2x + 4 sin 2x)
80
1 1
= sin 2x − cos 2x.
20 10
This has been found somewhat “more easily” than in finding yp by undetermined coeffi-
cients. Do you agree? Try it!
In this example, the identity
1 1 a + ib a + ib a + ib
= ⋅ = 2 2
=
a − ib a − ib a + ib a − (ib) a + b2
was used.
2.7. SHIFT THEOREM 87
Example 7
2y ′′ + 3y ′ + 15y = sin 3x
Now,
1
yp = Im {e3ix }
2(3i)2 + 3(3i) + 15
−1 − 3i
= Im {(cos 3x + i sin 3x) ( )}
30
1
= − (sin 3x + 3 cos 3x)
30
1 1
Hence a particular solution of the given D.E. is − 10 cos 3x − 30 sin 3x. (Not bad as a fast,
“recipe” method!)
Example 8
Example 10
dy
− y = x3 + 2x + 3
dx
1
yp = (x3 + 2x + 3)
D−1
−1
= (x3 + 2x + 3)
1−D
= − (1 + D + D2 + D3 + D4 + . . .) (x3 + 2x + 3)
= − [(x3 + 2x + 3) + (3x2 + 2) + (6x) + 6]
= −x3 − 3x2 − 8x − 11.
This is the same as the answer found using the method of undetermined coefficients and in
a lot less lines.
Example 11
Solution
1 1 1
yp = x= x= 2
D2 − 4D + 4 4 − 4D + D 2
4 [1 − (D − D4 )]
1 D2 D2 2
= [1 + (D − ) + (D − ) + . . .] x
4 4 4
1 1 1
= (x + 1 + 0) = x + .
4 4 4
Sneaky trick
Knowing that the series expansion is generated using powers of D that occur in the de-
nominator on the LHS, if we see that Dm R = 0 then any terms of Dm or higher can be
eliminated at the beginning.
Since R = x then D2 x = 0 and we can simplify the calculation of yn as follows:
1 1 1
yp = x= x= x
D2
− 4D + 4 −4D + 4 4(1 − D)
1 1
= (1 + D + D2 + ⋯) x = (x + 1)
4 4
as before.
Tutorial 2 Question 4
90 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
Read textbook section 3.5 (Electric circuits), section 5.7 (Electrical systems)
If the current flowing in the circuit is I and the charge on the capacitor is Q, then Kirchoff’s
Law gives
El + Er + Ec = E,
or
dI 1
L+ RI + ⋅ Q = E,
dt C
where E is the electromotive force supplied by the battery or generator.
dQ
Since I = dt , this equation becomes,
d2 Q dQ 1
L 2
+R + Q=E
dt dt C
a second order linear differential equation.
Example
2.9. APPLICATIONS OF LINEAR DIFFERENTIAL EQUATIONS 91
An RLC circuit is supplied by a generator having voltage given by E = 24 sin 10t. The
inductor is 1 henry, the resistor is 12 ohms and the capacitor is 0.01 farads. If at time t = 0
the switch is closed (i.e. at t = 0, Q = 0 and I = 0 ), find the charge Q and the current I at
time t.
Solution
The RLC circuit is as follows
El + Er + Ec = E
so that
dI 1
L + RI + ⋅ Q = E, or
dt C
2
dQ dQ 1
L 2 +R + ⋅ Q = E,
dt dt C
which in this case becomes
d2 Q dQ
2
+ 12 + 100Q = 24 sin 10t
dt dt
To solve this equation we first find the corresponding solution Qc of the homogeneous
equation
d2 Q dQ
2
+ 12 + 100Q = 0
dt dt
The auxiliary equation is
m2 + 12m + 100 = 0
which has complex roots, m = −6 ± 8i.
Hence the complementary solution is
d2 Q dQ
+ 12 + 100Q = 24 sin 10t
dt2 dt
by the method of undetermined coefficients. To do this we try a particular solution of the
equation of the form
Qp = a cos 10t + b sin 10t
92 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
(these are the only functions which can give the right-hand side when differentiated once
and twice).
Substituting Qp into the differential equation and solving for a and b gives a = − 15 and b = 0
(the solution of two simultaneous equations in a and b ). (Or try the D operator method.)
Hence
1
Qp = − cos 10t
5
and so the general solution is
1
Q = Qp + Qc = e−6t (A cos 8t + B sin 8t) − cos 10t
5
where the constants A and B can be found from the initial conditions (at t = 0, Q = 0 and
I = dQ
dt = 0). The condition Q = 0 at t = 0 gives
1
0 = e0 (A cos 0 + B sin 0) − cos 0
5
so that A = 51 .
Differentiating Q we get
dQ
I= = − 6e−6t (A cos 8t + B sin 8t)
dt
+ e−6t (−8A sin 8t + 8B cos 8t) + 2 sin 10t
e−6t 1
Q= (4 cos 8t + 3 sin 8t) − cos 10t
20 5
and the current I is
5e−6t
I =− sin 8t + 2 sin 10t.
2
The transient current, − 52 e−6t sin 8t dies away very quickly as t gets larger and the steady-
state current 2 sin 10t approximates the actual current when the transient current becomes
negligible. This is illustrated in the following sketch of the actual current I (steady-state
plus transient) and the steady-state current 2 sin 10t.
2.9. APPLICATIONS OF LINEAR DIFFERENTIAL EQUATIONS 93
The property of oscillation in the steady-state solution is very useful in certain applications.
For example, radio, television, radar and communication systems all exploit properties of
electrical oscillation (resonance). We tune our radios into certain frequencies and the oscil-
lating electronic circuits in the radio then amplify these frequencies to eventually produce
sound.
Another interesting application of electric circuits is in the analog computer. This device
simulates mechanical and physical systems by replacing the differential equation repre-
senting them by their electrical analogs, and then solving experimentally the resulting dif-
ferential equation for the electric circuits. The solution can be displayed on a cathode ray
oscilloscope in the form of a graph to those shown above for charge or current.
Tutorial 2 Question 5
94 CHAPTER 2. LINEAR DIFFERENTIAL EQUATIONS
2.10 Tutorial 2
Back to Chapter 2
Question 1
Find a general solution of each of the second-order linear differential equations below.
1. y ′′ − 3y ′ + 2y = 0
2. y ′′ + 5y ′ = 0
3. 3y ′′ − 10y ′ + 3y = 0
4. y ′′ + 25y = 0
5. y ′′ − 25y = 0
6. y ′′ − 4y ′ + 8y = 0
7. y ′′ + 10y ′ + 125y = 0
8. 9y ′′ − 6y ′ + 10y = 0
9. y ′′ = 7y
10. Textbook exercise set 4.2.
Problems 1, 5, 9, 13, 17, 19 (8,8e-th editions).
Problems 4, 13, 18, 19 (9th edition).
Answers
1. y(x) = C1 ex + C2 e2x
2. y(x) = C1 + C2 e−5x
x
3. y(x) = C1 e 3 + C2 e3x
4. y(x) = C1 cos 5x + C2 sin 5x
5. y(x) = C1 e5x + C2 e−5x
6. y(x) = e2x (C1 cos 2x + C2 sin 2x)
7. y(x) = e−5x (C1 cos 10x + C2 sin 10x)
x
8. y(x) = e 3 (C1 cos x + C2 sin x)
√ √
9. y(x) = C1 exp(x 7) + C2 exp(−x 7)
10. see textbook or detailed solutions
Question 2
Use the method of undetermined coefficients to find a particular solution for each of the
given equations:
1. y ′′ + 2y ′ + 2y = 3 cos 2x
2. y ′′ + 9y = 5 cos 3x
Answers
3 3
1. yp = − cos 2x + sin 2x
10 5
5
2. yp = x sin 3x
6
1 1
3. y = ex − e−2x − x −
2 2
4. see textbook or detailed solutions
Question 3
Solve each of the following D.E.s using the method of variation of parameters:
1. y ′′ + y = tan x
2. y ′′ − y = ex
3. y ′′ + 3y ′ + 2y = 3e−2x + x
4. y ′′ − y = x2 ex
Answers
1
2. y = Aex + Be−x + xex
2
1 3
3. y = Ae−x + Be−2x + x − − 3xe−2x
2 4
1 x
4. y = Aex + Be−x + e (2x3 − 3x2 + 3x)
12
5. see textbook or detailed solutions
Question 4
Use the inverse ϕ(D) operator method to find some of the particular solutions you have
previously found using other methods:
1. y ′′ − y = ex ;
2. y ′′ + 3y ′ + 2y = 3e−2x + x
3. (D2 − 1) y = x2 ex
4. y ′′ + 2y ′ + 2y = e−x cos x
Question 5
1. It has been found experimentally that a 6 kg mass stretches a certain spring 0.5 m.
If the mass is pulled 0.25 m below the equilibrium position and released, the D.E.
which describes the motion is
d2 x
+ 20x = 0
dt2
where x is the displacement from the equilibrium position, t is the time in seconds
after the release of the spring, and g = 10 m/s2 .
Find the value of x at any time t after the release of the spring (i.e. the solution of
the D.E.).
Hint: The first sentence is irrelevant to the solution, while the first part of the second
sentence tells you that x(0) = 0.25 and ẋ(0) = 0, where ẋ ≡ dxdt .
2.10. TUTORIAL 2 97
Answers:
1 √
1. x = cos(2 5t)
4
4
2. Q = −e−4t (cos 3t + sin 3t) + 1
3
3. see textbook
Infinite series
Chapter objectives
3.1 Introduction
Informally, an infinite series may be regarded as a sum of an infinite number of terms. But,
this raises many questions. What, if any, is the connection between the terms? Can we add
infinitely many numbers to get a sum? Indeed, can we ever add infinitely many numbers?
The ancient Greeks believed that no infinite set of numbers could possibly have a finite
sum. Because of this belief, they were caught in some logical paradoxes. We can avoid
their problems by considering first the idea of a sequence.
3.2 Sequences
Informally, a sequence is a set of numbers or terms arranged in a definite order according
to some rule, e.g.
2, 4, 8, . . . , 2n , . . .
This is a sequence whose first member is 2, second member is 4 = 22 , third member is 23
and so on. We call 2n , the general term or the n-th term.
Formally, a sequence is a function whose domain is a set of integers, usually the positive
integers, e.g. the sequence above should be denoted by
or by
{(n, 2n ) ∶ n = 1, 2, 3, . . .}
Note that the functional values are the important part of the ordered pairs. For this reason,
we just use these terms of the sequence. Thus, a sequence {(n, f (n)) ∶ n = 1, 2, 3, . . .} will
always be written simply as {f (n)}.
For instance, the sequence {n2 + 1} has its first term 2, its second term 5 and so on. Here we
are assuming that n takes on values n = 1, 2, . . . . We can also have the following sequence:
1, 2, 5, . . . , n2 + 1, . . . , where n = 0, 1, 2, . . . , if we so wish; or even
5, 10, 17, . . . , where n = 2, 3, . . . .
We will denote the general term of a sequence by an . Thus a sequence whose domain is
the positive integers will be given by a1 , a2 , a3 , . . .
For example, write down the first 4 terms of the sequence an ≡ (−1)n .
Here,
a1 = (−1)1 = −1
a2 = (−1)2 = 1
a3 = (−1)3 = −1
a4 = (−1)4 = 1
It appears that the terms are approaching 1 as n takes on larger values i.e. dn → 1 as n → ∞
In general, we say that a sequence {an } converges if and only if lim an exists.
n→∞
If a sequence does not converge, it is said to diverge e.g. {n2 + 1} is divergent since
limn→∞ (n2 + 1) = ∞, i.e. does not exist (as a real number).
More examples:
{ n12 + 1} is convergent since limn→∞ { n12 + 1} = 1
{(−1)n } diverges since limn→∞ (−1)n does not exist. ((−1)n oscillates between − 1 and
1.)
Note. To test for convergence of a sequence {an } , we evaluate limn→∞ an by any appro-
priate method. (You may need to revise the evaluation of limits.)
Tutorial 3 Question 1
Note. There are a large number of results that can be proved about sequences but they are
beyond the scope of this subject. We will turn our attention instead to infinite series whose
3.3. SERIES 101
3.3 Series
Video (by Dr. R. Wood): Infinite series (sections 3.3-7)
The sigma notation for finite sums should be familiar to you as, for example, in
n
∑ f (x∗i ) ∆xi
i=1
Note that a series is a sum, while a sequence is a function. However, a series is usually
formed by adding up the terms of a sequence, e.g.
∞
∑i = 1 + 2 + 3 + ...
i=1
∞
1 1 1 1
∑ 2
=1+ + + + ...
i=1 i 4 9 16
∞
1 1 1
∑ = 1 + + + ...
i=1 i 2 3
∞ ∞ ∞
1 1
We will show that ∑ i and ∑ do not “exist”, i.e. do not have a sum value but that ∑ 2
i=1 i=1 i i=1 i
∞
1 π2
does exist, i.e. adds up to a finite number. (In fact, it is known that the series ∑ 2 = ,
i=1 i 6
the proof being beyond the scope of this subject!) Let us consider the series
∞
∑ ai ≡ a1 + a2 + a3 + . . .
i=1
i.e. Sn , the n-th partial sum, is the sum of the first n terms of the series.
We now consider the sequence of these partial sums, {Sn } . If the sequence {Sn } converges
(i.e. limn→∞ Sn exists and equals S, say), then we say that the series is convergent and has
the value (sum) S. If {Sn } diverges, then ∑∞i=1 ai is called divergent.
Example 1
The series ∞
∑i = 1 + 2 + 3 + ...
i=1
n(n+1)
is divergent since the partial sum, Sn , = 1 + 2 + . . . + n which is known to be 2 , and
limn→∞ Sn = limn→∞ n(n+1)
2 = ∞.
Example 2
1 if n is odd
Sn = {
0 if n is even
∴ ∑∞
i=1 Sn does not exist and the series is divergent.
Example 3
Consider ∞
3 3 3 3
∑ i
= + + + ...
i=1 10 10 100 1000
Here n
1
3 3 3 3 (1 − ( 10 ) )
Sn = + + ... + n = 1 .
10 102 10 10 1 − 10
Here we used the sum of a geometric progression formula.
1 n
3 ⎛ 1 − ( 10 ) ⎞
∴ lim Sn = lim
n→∞ n→∞ 10 ⎝ 1 − 1
10 ⎠
3 1 1 n
= ( ) , since ( ) → 0
10 9/10 10
1
=
3
Hence,
∞
3 1
∑ i
=
i=1 10 3
which is very “reasonable” since
∞
3 3 3
∑ i
= + + ...
i=1 10 10 100
1
can be written as 0.3333 . . . which we know is the decimal form of 3 (We can similarly
show that 1 = 0.9999 . . .).
3.3. SERIES 103
Note. The method used in Example 3 will work for any geometric series, i.e. any series
whose terms are in a geometric progression.
Hence ∞
∑ (ari−1 ) ≡ a + ar + ar2 . . .
i=1
converges to
a
1−r
if 0 < ∣r∣ < 1 and diverges if ∣r∣ > 1.
However, the general term in a sequence of partial sums may be difficult or impossible to
determine, e.g. for the series
∞
1
∑
i=1 i
Evaluate limn→∞ an and if this does not equal 0, the series is divergent.
Be careful: if limn→∞ an = 0, the question is still open, and the series may be convergent
or divergent.
For example, the series
∞
1 1 1
1+ + ... + + ... ≡ ∑
2 n n=1 n
is divergent although
1
lim an = lim
= 0.
n→∞ n n→∞
a. If ∑ an and Σbn are convergent series and k is a constant, then Σ (an + bn ), Σ (an − bn )
and Σ (kan ) , Σ (kbn ) are also convergent. e.g. once we have shown that
∞
1
∑ 2
n=1 n
is a convergent series.
Also, the series
∞
∑ (2−n + 3−n )
n=1
104 CHAPTER 3. INFINITE SERIES
converges since
∞
∑ 1/2n
n=1
and
∞
∑ 1/3n
n=1
Note. From now on we will use n as the index of summation, as in ∑∞ n=1 an . Also, Σan
is an expression to indicate that it is immaterial whether n = 1, 2, . . . or n = 0, 1, 2, . . . or
indeed n = N, N + 1, . . .
Tutorial 3 Question 2
In order to develop tests for the convergence of series, we will first examine the case where
all of the terms of the series are positive, i.e. consider ∑∞
n=1 an and an ≥ 0 for all n
Integral test
If ∑∞
n=1 an is a series of positive terms, if f is a continuous, positive valued, decreasing
∞
function in [1, ∞) and if f (n) = an for all positive integers n, then ∑∞
n=1 an and ∫1 f (x)dx
both converge or both diverge.
This test requires us to form an appropriate improper integral, to evaluate that integral and
hence to infer at once whether the series converges or diverges.
Example 1
1 1
for convergence for various values of α. Here, we take f (x) = xα , as then f (n) ≡ nα .
For x ≥ 1, x1α is positive and continuous, and, in addition, for α > 0, it is decreasing.
Therefore we can apply the integral test to the case of α > 0 by considering the improper
∞
integral ∫1 x1α dx.(α ≤ 0 will have to be considered separately. )
3.4. SERIES WITH POSITIVE TERMS 105
Now,
∞ 1 N 1
∫1 dx = lim ∫ dx
xα N →∞ 1 xα
1 1 N
= lim [ ] , (α ≠ 1)
N →∞ −α + 1 xα−1 1
1 1 1
= lim { α−1
− }
N →∞ 1 − α N 1−α
− 1 for α > 1
= { 1−α
∞ for α < 1
Hence, the integral converges for α > 1 and diverges for 0 < α < 1. Thus the series ∑∞ 1
n=1 nα
converges for α > 1 and diverges for 0 < α < 1.
Hence, the integral converges for α > 1 and diverges for 0 < α < 1. Thus the series ∑∞ 1
n=1 nα
converges for α > 1 and diverges for 0 < α < 1.
For the case α = 1, the integral is
∞ 1
∫1 dx = lim [ln x]N
1 =∞
x N →∞
i.e. diverges.
Thus the harmonic series ∑∞ 1
n=1 n is again shown to be divergent. Also, for α < 0
∞ ∞
1
∑ α
≡ ∑ n−α ,
n=1 n n=1
where (−α) is positive. Hence, limN →∞ (n−α ) ≠ 0 and so, for α < 0, ∑∞ 1
n=1 nα diverges by
an earlier test. (What about the case α = 0?)
Example 2
Note that here n starts its values at 2 since n = 0 and 1 would yield 0 in the denominator.
1
Check that x ln x is continuous, positive and decreasing for x ≥ 2. Hence we consider
∞ dx
∫2 x ln x
which can be shown to be a divergent integral. (Try it!)
Hence,
∞
1
∑
n=2 n ln n
diverges.
Tutorial 3 Question 3
106 CHAPTER 3. INFINITE SERIES
Comparison tests
1. If we are given a series ∑ an and can find a convergent series ∑ bn such that
0 ≤ an ≤ bn for all n, then ∑ an converges.
2. If we are given a series ∑ cn and can find a divergent series ∑ dn such that cn ≥ dn
for all n, then ∑ cn diverges.
Example 3
Hence ∑∞ 1
n=1 2n +n converges.
Example 4
converge or diverge?
Here we observe that
1 1
√ > √
4 n−1 4 n
for all n ≥ 1; that
1 1
∑ √ ≡ ∑ 1/2
n n
is divergent by Example 1; and that ∑ 4√1 n is thus also divergent. Hence, ∑ 4√1n−1 is diver-
gent.
A more useful comparison test is given by the following result.
The limit comparison test
an
Suppose that ∑ an and ∑ bn are two series with positive terms. If lim exists and equals
n→∞ bn
L > 0, then the two series both converge or both diverge.
Example 5
3.4. SERIES WITH POSITIVE TERMS 107
Does
∞
1
∑
n=1 3n2 − 2n + 1
converge or diverge? Here, the series itself suggests a comparison with
∞
1
∑ 2
.
n=1 n
We have
n2 1
lim = >0
n→∞ 3n2 − 2n + 1 3
Since we know that ∑ n12 converges, the series
∞
1
∑
n=1 3n2 − 2n + 1
converges too.
Example 6
The series
1
∑√ 2
3n + 1
is divergent since if we compare it with ∑ n1 in the limit comparison test, we get that
√ 1
3n2 +1 n 1 1
lim = lim √ = lim √ = √ >0
n→∞ 1/n n→∞ 3n2 + 1 n→∞ 3 + 1 3
n2
Tutorial 3 Question 4
Ratio test
• If ρ = 1, the test fails and some other test will be necessary to decide convergence or
divergence.
108 CHAPTER 3. INFINITE SERIES
Example 7
Here,
an+1 2n+1 n
lim = lim n
n→∞ an n→∞ 2 (n + 1)
2n
= lim
n→∞ n + 1
= 2 > 1.
∞
2n
∴ ∑ diverges by the ratio test.
n=1 n
Example 8
is convergent or not.
Here,
an+1 1/(n + 1)3
lim = lim
n→∞ an n→∞ 1/n3
n 3
= lim ( )
n→∞ n + 1
=1.
Thus the test is inconclusive and some other test must be used (e.g. the integral test) to
show that this series is, in fact, convergent.
Note. There are a large number of other tests for convergence, most of which are highly
specialised. We will make particular use of the ratio test when we examine Power Series.
Tutorial 3 Question 5
If
a. lim an = 0 and
n→∞
Example
1
a. lim an = lim = 0, and
n→∞ n→∞ n2
1 1
b. an+1 = 2
≤ 2 = an for all n.
(n + 1) n
∞
1
∴ ∑ (−1)n+1 is convergent.
n=1 n2
Tutorial 3 Question 6
The series ∑ ∣an ∣ has positive terms and thus the tests for convergence discussed previously
can be applied to it. If ∑ ∣an ∣ can be shown to be convergent, then we say that the series
Σan is absolutely convergent.
110 CHAPTER 3. INFINITE SERIES
Once we have shown that a series is absolutely convergent (A.C.) we can use the theorem
(here accepted without proof) that if a series is absolutely convergent, then it is convergent.
Example 1
Show that
∞ sin ( nπ )
3
∑
n=1 n2
is absolutely convergent and hence is convergent. We must consider the series
∞ sin nπ
√ √ √ √
3 3 3 3 3
∑∣ 2 ∣= + + + + ...
n=1 n 2 8 32 50
i.e. √
3
2n2 n ≠ 3k
∣an ∣ = {
0 n = 3k
∞
√ √
3 3
If we compare this series with the series ∑ 2 , then ∣an ∣ ≤ 2 for all n.
n=1 2n 2n
∴ ∑ ∣an ∣ is convergent by the dominated comparison test, since
∞
√ √ ∞
3 3 1
∑ 2= ∑ 2
n=1 2n 2 n=1 n
∞
sin ( nπ )
3
is known to be convergent. Hence ∑ is absolutely convergent and, by the theo-
n=1 n2
rem, is thus also convergent.
Example 2
Test ∞
n+1
∑ (−1)n+1
n=1 n2
for absolute convergence and for convergence.
Here,
n+1 n+1
∣an ∣ = ∣(−1)n+1 ∣= 2
n2 n
∞
n+1
for all n, and so we consider the series ∑ 2
n=1 n
∞
1
If we compare this series with the divergent series ∑ using the limit comparison test,
n=1 n
n+1
2 n+1
lim n = lim ( )=1>0
n→∞ 1 n→∞ n
n
n+1
and so ∑ is also divergent.
n2
∞
n+1
Thus ∑ (−1)n+1 2 is not absolutely convergent, and the theorem cannot be applied to
n=1 n
this series.
3.7. POWER SERIES 111
∞
n+1
However, ∑ (−1)n+1 is an alternating series which meets the criteria for the alternat-
n=1 n2
∞
n+1
ing series test. (Show this!) Hence ∑ (−1)n+1 2 is convergent.
n=1 n
Note. Example 2 exhibits a series which is not absolutely convergent but is convergent.
Such a series is called conditionally convergent (C.C.).
Tutorial 3 Question 7
Introduction
Thus far we have considered series whose terms are constants. Let us now consider series
whose terms are functions of a variable x.
We will not examine the convergence of the most general of such series i.e. ∑ fn (x) e.g.
∞
∑ (2 cos x)n but will consider only a special case, the power series.
n=1
∞
A power series has the form ∑ an (x−a)n , where the an ’s are constants which depend only
n=0
on n and a is a fixed constant, i.e.
a0 + a1 (x − a) + a2 (x − a)2 + . . .
When a value is substituted for x, this becomes an infinite series which may converge or
diverge, i.e. may or may not be said to have a “value” for that particular value of x.
Clearly, if we let x = a, then the power series becomes
a0 + a1 ⋅ 01 + a2 ⋅ 02 + . . .
lim 5n ≠ 0
n→∞
In general, we note that 1 + x + x2 + . . . is a geometric series with common ratio x and so the
1
power series converges ( to 1−x ) for any x such that ∣x∣ < 1 i.e. −1 < x < 1, and diverges
for ∣x∣ ≥ 1, i.e. x ≥ 1 or x ≤ −1.
Radius and interval of convergence
Note. The series ∑∞ n
n=1 x converges for x in the interval (-1,1) and we call this the interval
of convergence of this power series. Further, the centre of this interval is at 0 and we say
that its ”radius of convergence” is 1 . If we can determine that a series is convergent for
∣x − a∣ < r
then the radius of convergence is r. The interval of convergence can then be determined by
simplifying
−r < x − a < r
Example 1
converges. First we test the series for absolute convergence using the ratio test as follows.
Consider the series
∞
xn
∑∣ ∣,
n=0 (n + 1)2n
where the terms are positive and the general term is
xn ∣x∣n
∣ ∣ = .
(n + 1)2n (n + 1)2n
3.7. POWER SERIES 113
converges absolutely (and thus also converges) if ∣x∣ < 2, i.e. −2 < x < 2, and diverges if
∣x∣ > 2.
Hence the interval of convergence is (-2,2) (“centered” at 0 as we expected) and the radius
of convergence is 2.
We still do not know how the series behaves when ρ = 1, i.e. when ∣x∣ = 2, i.e. x = 2 or −2.
Accordingly, we must test the series
∞
xn
∑ n
n=0 (n + 1)2
which we know is the divergent harmonic series. On substituting x = −2, the series becomes
∞ ∞
(−2)n (−1)n 1 1
∑ n
= ∑ = 1 − + − ...
n=0 (n + 1)2 n=0 n + 1 2 3
which is an alternating series that we know (see the exercise in the section on Absolute
Convergence) to be convergent but not absolutely convergent.
To sum up: the series is absolutely convergent and thus convergent for x in (-2,2); it is
conditionally convergent for x = −2 (we say that the complete interval of convergence is
[-2,2) ); and it is divergent for x < −2 and x ≥ 2.
Example 2
1
= ∣x∣ lim
n→∞ n + 1
= 0 for all x
Hence, for all x, ρ = 0 < 1 and so the series converges absolutely for all x. i.e. the interval
of convergence is (−∞, ∞). (No endpoints need be checked here!) (This is a case where
the “radius of convergence”, r = ∞. )
Example 3
∞
Find the interval of convergence for ∑ n!(x + 3)n .
n=0
Here,
∣(n + 1)!(x + 3)n+1 ∣
ρ = lim = lim (n + 1)∣x + 3∣
n→∞ ∣n!(x + 3)n ∣ n→∞
∞ for all x ≠ −3
={
0 for x = −3
Hence, the series converges only when x = −3 and diverges for all other x. The interval
thus consists of the single point, −3, and the “radius of convergence” is 0.
Example 4
the (complete) interval of convergence is (-1,9) which is centered at x = 4 and has “radius”
5.
Functions defined by a power series
Since a power series has a unique sum at each point of its interval of convergence, it defines
a function given by
∞
F (x) = ∑ an (x − a)n , x ∈ I
n=0
the interval of convergence. For a particular power series, the function F may be hard or
impossible to find explicitly. However, the following properties of F, can be established
(the proofs are beyond the scope of this subject):
Example 5
Consider the power series ∑ xn . This power series has the interval of convergence (-1,1)
n=0
since
∣xn+1 ∣
ρ = lim = lim ∣x∣ = ∣x∣
n→∞ ∣xn ∣ n→∞
and the series diverges at each end point. (Check all of this.) Now, it happens that this
1
power series 1 + x + x2 + . . . is a geometric series and so we know that its sum is 1−x for
∣x∣ < 1.
∞
1
∴ Here, F (x) = ≡ ∑ xn for all x in (-1,1)
1 − x n=0
1
Clearly is continuous, differentiable and integrable for all x in (-1,1).
1−x
1 2 n!
Also, F ′ (x) = 2
, F ′′ (x) = 3
and, in general, F (n) (x) =
(1 − x) (1 − x) (1 − x)n+1
∞
Now, for this power series ∑ xn , a = 0 and an ≡ 1 for all n.
n=0
F (n) (0) n!
Hence should equal 1 for all n, which is clearly so for F (n) (x) = .
n! (1 − x)n+1
116 CHAPTER 3. INFINITE SERIES
Taylor series
We now want to consider the converse problem: given a function f (x) “expand” it to give
a power series in powers of (x − a).
It can be shown that, if a power series exists for a particular series and if f has all order
derivatives in an interval containing a, then
∞
f (n) (a)
f (x) = ∑ (x − a)n
n=0 n!
f ′′ (a)
= f (a) + f ′ (a)(x − a) + (x − a)2 + . . .
2!
for all x in the interval.
∞
f (n) (a)
The series ∑ (x−a)n is called the Taylor series for f about a. If a = 0 in the Taylor
n=0 n!
series, we get
∞
f ′′ (0) 2 f (n) (0) n
f (0) + f ′ (0)x + x + ... = ∑ x
2! n=0 n!
which is often called the Maclaurin series for f .
Note
Examine carefully what the theorem above has as its suppositions. Only once we know
that a function actually has a power series expansion do we know that this power series is
in fact the Taylor series.
We can, of course, calculate the Taylor series corresponding to any given function f . Then
we can find the interval of convergence of the resulting power (Taylor) series using the
methods developed previously. But, we do not know that the Taylor series corresponding
to f actually converges to f (x) until we have checked that a certain “Remainder” tends to
zero.
This remainder is given by Rn (x) in Taylor’s formula, which states: If f has (n + 1)
derivatives in an interval I that contains the number a, then for x in I there is a number z
strictly between x and a such that
f (n+1) (z)
where Rn (x) = (x − a)n+1 .
(n + 1)!
Fortunately, for all the elementary functions that we consider, it can be shown that the
Taylor series does converge to f (x) for all x in the interval of convergence, i.e. We will
write
∞
f (n) (a)
f (x) = ∑ (x − a)n
n=0 n!
and ignore the question of the remainder.
Example 6
Expand f (x) = cos x in powers of x and find the interval of convergence for the resulting
3.7. POWER SERIES 117
(Maclaurin) power series. Here a = 0 and we evaluate f (0), f ′ (0), f ′′ (0), . . . , in turn.
f (x) = cos x f (0) =1
f ′ (x) = − sin x f ′ (0) =0
f ′′ (x) = − cos x f ′′ (0) = −1
f ′′′ (x) = sin x f ′′′ (0) =0
f (4) (x) = cos x f (4) (0) =1
A pattern is now clearly established. Now the required series is given by
∞
f (n) (0) f ′ (0) f ′′ (0) 2
f (x) = ∑ (x)n = f (0) + x+ x + ...
n=0 n! 1! 2!
0 (−1) 2 0 3 1 4
i.e. cos x = 1 + x + x + x + x + ...
1! 2! 3! 4!
x2 x4
=1− + − ...
2! 4!
∞
x2n
= ∑ (−1)n
n=0 (2n)!
To find the interval of convergence of this power series, we consider
∣ (−1)n+1 x2(n+1) /(2(n + 1))!
ρ = lim
n→∞ ∣(−1)n x2n /(2n)!∣
1
= lim ∣x∣2
n→∞ (2n + 1)(2n + 2)
= 0 for all x.
Hence the power series converges absolutely for all x and, hence, (without examining the
appropriate remainder term) we have that, for all x
x2 x4 x6
cos x = 1 − + − + ...
2! 4! 6!
Example 7
Expand ln(1 + x) in a power series about x = 0 and find the interval of convergence of the
series.
Here a = 0 and we must find f (0), f ′ (0), . . .
f (x) = ln(1 + x)
1
f ′ (x) =
1+x
−1
f ′′ (x) =
(1 + x)2
2
f ′′′ (x) =
(1 + x)3
−2 ⋅ 3
f (4) (x) =
(1 + x)4
hence
f (0) = ln 1 = 0
f ′ (0) = 1
f ′′ (0) = −1
f ′′′ (0) = 2
f (4) (0) = −6 etc.
118 CHAPTER 3. INFINITE SERIES
which is divergent.
When x = 1, the series is ∑∞n=1 (−1)
n+1 1 which is convergent by the alternating series test,
n
but is not absolutely convergent. Hence, the complete interval of convergence for which
∞
xn
ln(1 + x) = ∑ (−1)n+1 is (−1, 1].
n=1 n
Example 8
Example 1
Prove that
sin x
lim =1
x→0 x
The Maclaurin series for sin x is easily found to be given by
x3 x5
x− + − ...,
3! 5!
which converges for all x
sin x x2 x4
=1− + ......
x 3! 5!
sin x x2 x5
∴ lim = lim (1 − + − . . .)
x→0 x x→0 3! 5!
= 1 ( all other terms → 0)
Example 2
In the section on Taylor series, Example 8 we sought the expansion of ex in powers of x+2.
It is easily shown that
∞
x xn
e =∑ for all x
n=0 n!
If we replace x by (x + 2) on the right-hand side of this result, we get a power series in
(x + 2). However, to preserve the equality, we must replace x by (x + 2) on the left-hand
side too ∞
(x + 2)n
∴ ex+2 = ∑
n=0 n!
∞
(x + 2)n
∴ ex ⋅ e2 = ∑
n=0 n!
∞
(x + 2)n
∴ ex = ∑ e−2 (as before!)
n=0 n!
Example 3
Hence, we have found the sum of the alternating harmonic series. Also, we can determine
the value of ln 2, to as many decimal places as we wish, by taking a suitable number of
terms of the infinite series!
Further, since
d 1
ln(1 + x) = ,
dx 1+x
by differentiating the given power series term-by-term, we immediately get the Maclaurin
1
series for .
1+x
Also,
1 d x2 x3 x4
= (x − + − + . . .)
1 + x dx 2 3 4
= 1 − x + x2 − x3 + . . .
1
= 1 − (−x) + (−x)2 − (−x)3 + . . .
1−x
= 1 + x + x2 + x3 + . . .
which we have seen previously. (See Section on Functions defined by a power series,
Example 5.)
1
However, the series we have derived for 1−x can easily be shown to have interval of conver-
gence (-1,1) , and not (-1,1] as for ln(1 + x). (The endpoints of the interval of convergence
must thus still be examined separately for convergence.)
Note particularly the Binomial Series.
You would have already met the Binomial Theorem, which states that for a, b any real
numbers and n a positive integer,
n
n
(a + b)n = ∑ ( ) an−k bk
k=0
k
If we put a = 1, b = x, we get
n
n n n(n−1)...(n−k+1) n
(1 + x)n = ∑ ( ) xk , where ( ) = k! , ( )=1
k=0
k k 0
Newton extended the Binomial Theorem to the case where n is no longer a positive integer.
By listing the derivatives of f (x) = (1 + x)n , we can show that if n is any real number
and ∣x∣ < 1, then we get an infinite series:
∞
n
(1 + x)n = ∑ ( ) xk .
k=0
k
n
Note that if n is a given positive integer, then ( ) = 0 for k > n, whereas for other real n,
k
n − 21 (− 12 ) (− 32 ) (− 52 ) 1⋅3⋅5 15
( ) ≠ 0 for any k, e.g. ( )= =− =− .
k 3 3! 2⋅4⋅6 48
This concludes our examination of Series which have many uses and extensions. as we
will see in other subjects! In Chapter 5 we will examine the use of Series in solving certain
classes of Differential Equations.
3.7. POWER SERIES 121
This concludes our examination of Series which have many uses and extensions, as we will
see in other subjects! In Chapter 5 we will examine the use of Series in solving certain
classes of Differential Equations.
Tutorial 3 Question 8
122 CHAPTER 3. INFINITE SERIES
3.8 Tutorial 3
Back to Chapter 3
Question 1
n2 + 1
1. { }
n3 + 2n + 1
n4 + 1
2. { }
n3 + 2n + 1
n3
3. { }
2n
n3
Hint: Use l’Hopital’s rule to evaluate lim
n→∞ 2n
5. {1 + (−1)n }
1 + (−1)n
6. { }
n
1 k
7. a2k = 1, a2k+1 = 1 − ( )
2
n2 − n + 7
8. an =
2n3 + n2
sin2 n
9. an = √
n
√
n
10. an =
ln n
2n + 1
11. an =
en
Answers
1. Convergent
2. Divergent
3. Convergent
Convergent 0≤r≤1
4. {
Divergent r>1
5. Divergent
3.8. TUTORIAL 3 123
6. Convergent
7. Convergent
0
a1 = a2⋅0+1 = 1 − ( 12 ) = 0, a2 = a2⋅1 = 1
1
a3 = a2⋅1+1 = 1 − ( 21 ) = 12 , a4 = 1
2
a5 = 1 − ( 12 ) = 34 , etc.
limn→∞ an = 1.
8. Convergent: limn→∞ an = 0
9. Convergent: limn→∞ an = 0
10. Divergent
Question 2
Determine whether a given series converges or diverges. If it converges, find its sum.
1 1 1
1. 1 + + +⋯+ n +⋯
3 9 3
2. 1 + 3 + 5 + 7 + ⋯ + (2n − 1) + ⋯
3. 1 − 2 + 4 − 8 + 16 − ⋯ + (−2)n + ⋯
4 4 4
4. 4 + + +⋯+ n +⋯
3 9 3
Answers
1. 3/2
4. 6
Question 3
Use the integral test to test the given series for convergence
1. ∞
n
∑
n=1 n2
+1
2. ∞
1
∑√
n=1 n+1
3. ∞
1
∑
n=1 n2 +1
4. ∞
ln n
∑ 2
n=1 n
Answers
1. Diverges
2. Diverges
3. Converges
4. Converges
Question 4
∞
1
1. ∑
n=1 n2 +n+1
∞
1
2. ∑ √
n=1 n + n
∞
1
3. ∑ n
n=1 1 + 3
∞
√
n
4. ∑ 2
n=1 n + n
∞
sin2 n
5. ∑ 2
n=1 n + 1
3.8. TUTORIAL 3 125
Answers
1. Converges
2. Diverges
3. Converges
4. Converges
5. Converges
Question 5
Use the ratio test to determine whether the series converge or diverge.
∞
10n
1. ∑ n
n=1 n
∞
n n
2. ∑ ( )
n=2 ln n
∞
3n
3. ∑
n=1 n!n
∞
n!n2
4. ∑
n=1 (2n)!
∞
3 n
5. ∑ n ( )
n=1 4
Question 6
∞
(−1)n+1
1. ∑
n=1 n2
∞
(−1)n+1 n
2. ∑
n=1 3n + 2
∞
(−1)n+1 n
3. ∑ √
n=1 n3 + 2
126 CHAPTER 3. INFINITE SERIES
∞
(−1)n+1 n
4. ∑
n=2 ln n
∞
(−1)n+1
5. ∑ √
n
n=1 2
Answers
1. Converges
2. Diverges
3. Converges
4. Diverges
5. Diverges
Question 7
∞ ∞
(−1)n+1 (−1)n+1
a. Show that ∑ is A.C. and hence is convergent; that ∑ is not A.C. but
n=1 3n n=1 n
∞
is convergent, i.e. it is C.C.; and that ∑ 3n is not A.C., nor C.C., i.e. it is divergent.
n=1
Answers
b.2 Diverges
Question 8
Find the interval of convergence of the power series (do not forget to investigate conver-
gence at the endpoints)
∞
1. ∑ nxn
n=1
∞
nxn
2. ∑ n
n=1 2
∞
3. ∑ n!xn
n=1
∞
(−1)n nxn
4. ∑ n 3
n=1 2 (n + 1)
∞
5. ∑ (5x − 3)n
n=1
∞
2n (x − 3)n
6. ∑
n=1 n2
7. Textbook exercise set 8.2. Problems 1,3,5
Answers
1. (−1, 1)
2. (−2, 2)
3. [0, 0]
4. [−2, 2]
5. (2/5, 4/5)
6. [5/2, 7/2]
7. see textbook
Chapter objectives
• obtain a power series solution given the point classification for a differential equation
4.1 Introduction
In this chapter, we return to the study of linear D.E.s, but now take up the question of the
solution of such D.E.s with non-constant (or variable) coefficients. We will confine our
attack to second order L.D.E.s of the form
but the method can be readily generalised. In fact, we will examine only homogeneous
L.D.E.s, i.e. we will consider D.E.s where s(x) = 0. Techniques for finding a particular
solution for the general case can be easily established, but are beyond the scope of this
subject. These study notes should be comprehensive enough for our purposes.
Notes.
1. To solve such D.E.s with non-constant coefficients, we will have to use series. Thus
some of the concepts from Chapter 4 will be important here. We will, in fact, not
have to worry too much about the convergence of our series solutions - the underly-
ing theory guarantees that such a power series solution converges in an interval of
convergence!
Further, we will only examine the “basic”’ case which yields a power series solution.
We will not examine the use of Taylor series to solve D.E.s at all, and will only
“briefly” examine more complicated cases which require a Frobenius type series for
their solution.
129
130 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
2. Until now we have been concerned with, and in fact have restricted ourselves to,
differential equations which could be solved exactly. There are certain differential
equations which are of extreme importance in higher mathematics and engineering
or other scientific applications but which cannot be solved exactly in terms of ele-
mentary functions by any methods. For example, the innocent looking differential
equation
xy ′′ + y ′ + xy = 0
cannot be solved exactly in terms of functions usually studied in elementary calculus,
such as the rational algebraic, trigonometric and inverse trigonometric, exponential,
and logarithmic functions. In such cases, short of recourse to higher transcendental
functions, series or other “approximate” methods provide the only alternatives.
Read textbook section 8.3 (Power series solutions to linear differential equa-
tions)
Video (by Dr. R. Wood): Series solution of linear ODE (sections 4.2-3)
Example 1
y = C1 e−x + C2 ex
y = f (x) = a0 + a1 x + a2 x2 + a3 x3 + . . .
From before, we can differentiate the power series term by term to obtain
dy
= a1 + 2a2 x + 3a3 x2 + 4a4 x3 + . . .
dx
and
d2 y
= 2a2 + 3 ⋅ 2a3 + 4 ⋅ 3a4 x2 + 5 ⋅ 4a5 x3 + . . .
dx2
The DE then becomes
2a2 + 3 ⋅ 2a3 x + 4 ⋅ 3a4 x2 + 5 ⋅ 4a5 x3 + . . .
= a0 + a1 x + a2 x 2 + a3 x 3 + . . .
4.2. INFORMAL SERIES APPROACH TO THE SOLUTION OF DIFFERENTIAL EQUATIONS131
For this to be true for all x in the interval of convergence will require that the coefficients
of corresponding powers of x on both sides of the equality to be equal.
Equating coefficients give
2a2 = a0
3 ⋅ 2a3 = a1
4 ⋅ 3a4 = a2
5 ⋅ 4a5 = a3
and so on. Hence,
1
a2 = a0
2
1
a3 = a1
3⋅2
1 1 1 1
a4 = a2 = ⋅ a0 = a0
4⋅3 4⋅3 2 4⋅3⋅2
1 1 1 1
a5 = a3 = ⋅ a1 = a1
5⋅4 5⋅4 3⋅2 5⋅4⋅3⋅2
and looks like
1 1 1 1
a2 = a0 , a3 = a1 , a4 = a0 and a5 = a1
2! 3! 4! 5!
There is an obvious pattern and with a0 , a1 arbitrary the series solution appears to be
y = a0 + a1 x + a2 x 2 + a3 x 3 + a4 x 4 + a5 x 5 + . . .
1 1 1 1
= a0 + a1 x + a0 x 2 + a1 x 3 + a0 x 4 + a1 x 5 + . . .
2! 3! 4! 5!
x 2 x4
= a0 [1 + + + . . .]
2! 4!
= a0 y 1 + a1 y 2 .
As expected for a second order linear DE the two solution functions y1 and y2 are linearly
independent. An alternative approach (we will use later) is to consider the general solution
of the second order linear DE as
y = C1 y 1 + C2 y 2
y 1 = a0 + a1 x + a2 x 2 + . . .
= 1 + a2 x 2 + . . .
and
y 2 = a0 + a1 x + a2 x 2 + . . .
= 0 + x + a2 x 2 + . . .
= x + a2 x 2 + . . .
will be linearly independent. Basically if we can find two solution functions y1 and y2 for
the DE which are linearly independent then the linear combination
y = C1 y 1 + C2 y 2
for the second order DE should give the complete general solution.
132 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
and
x2 x4
1+ + + . . . = cosh x
2! 4!
We then have
y = C1 cosh x + C2 sinh x.
and with the initial conditions
y = 3 cosh x + 2 sinh x
Further, by reducing this solution to exponential form matches the previously found ana-
lytic solution:
ex + e−x ex − e−x
y =3⋅ +2⋅
2 2
3ex + 3e−x + 2ex − 2e−x
=
2
5ex + e−x 5 x 1 −x
= = e + e
2 2 2
as before.
More formal approach to series solution
We consider the more general form of a second order linear differential equation as
a2 (x)y ′′ + a1 (x)y ′ + a0 (x)y = 0
The form and existence of a series solution to this DE depends on whether the series is
taken about an ordinary point or a singular point of the DE.
Singular points for the DE are found by solving a2 (x) = 0. If any x satisfy this they are
called singular points and can be further classified. All other points are ordinary points for
the given DE. For a series solution about a singular point (depending on the classification
of the point) we will try the method of Frobenius. This will be covered in the next section.
For now we will only consider the series solution for a DE about an ordinary point. If x = a
is an ordinary point of the DE, then the substitution
y = a0 + a1 (x − a) + a2 (x − a)2 + a3 (x − a)3 + . . .
∞
= ∑ an (x − a)n
n=0
will lead to two series solutions involving two arbitrary constants and hence to the general
solution.
We can also construct a recurrence relation for the series solution by using the summation
form of the general series form as follows.
We write ∞
y = a0 + a1 x + a2 x 2 + . . . = ∑ an x n
n=0
Then ∞
y ′ = a1 + 2a2 x + 3a3 x2 + . . . = ∑ nan xn−1
n=1
and
y ′′ = 2a2 + 3 ⋅ 2a3 x + 4 ⋅ 3a4 x2 + . . .
∞
= ∑ n(n − 1)an xn−2
n=2
134 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
To simplify this under one summation we need the summation starting counter values to
be the same. To simplify our working we will also set the x indexes to be the same, in this
case xn . We do this step first, before adjusting the counter starting values. For the first sum
we simply add 2 to the counter and subtract 2 from the starting value giving:
∞ ∞
∑ (n + 2)(n + 2 − 1)an+2 xn+2−2 = ∑ (n + 2)(n + 1)an+2 xn .
n=0 n=0
If this is done correctly the summations are equivalent. This can be easily checked by
generating the first 2 terms for each sum and comparing as follows:
∞
∑ n(n − 1)an xn−2 = 2(2 − 1)a2 x2−2 + 3(3 − 1)a3 x3−2 + . . .
n=2
= 2 ⋅ 1a2 x0 + 3 ⋅ 2a3 x1 + . . .
= 2a2 + 6a3 x + . . .
∞
∑ (n + 2)(n + 1)an+2 xn =(0 + 2)(0 + 1)a0+2 x0 + (1 + 2)(1 + 1)a1+2 x1 + . . .
n=0
= 2 ⋅ 1a2 + 3 ⋅ 2a3 x + . . .
= 2a2 + 6a3 x + . . .
as before.
We then have ∞ ∞
∑ (n + 2)(n + 1)an+2 xn − ∑ an xn = 0
n=0 n=0
This adjustment has also resolved the first problem as the counter start values are now
equal and the terms can now be put in one summation. This will not always happen and
the counter start values will have to be adjusted. This is done by setting all the counter start
values to the highest start value. This is done by taking terms out of each summation till
they all start at the same highest value.
For example
∞ ∞ ∞
∑ an x n + ∑ b n x n + ∑ c n x n
n=2 n=1 n=0
The highest counter start value is 2 so we take terms out of the remaining summations until
their counter start values are 2 as follows:
∞ ∞ ∞
∑ an x n + b 1 x 1 + ∑ b n x n + c 0 x 0 + c 1 x 1 + ∑ c n x n
n=2 n=2 n=2
∞
= b1 x + c0 + c1 x + ∑ (an xn + bn xn + cn xn )
n=2
All the terms must be zero on the left hand side so we have
(n + 2)(n + 1)an+2 − an = 0
giving the recurrence relation
1
an+2 = an .
(n + 2)(n + 1)
For computer programming for a numerical solution using power series the general recur-
rence relation can be used.
Finding series terms for y1 and y2
For simple cases where the recurrence relation yields series which are odd and even, the
following approach can be taken to generate the general term for each series. For the even
series we want to find ∞
y1 = ∑ a2n x2n = a0 + a2 x2 + . . .
n=0
and for the odd series ∞
y2 = ∑ a2n+1 x2n+1 = a1 x + a3 x3 + . . .
n=0
The usual approach is to let the leading coefficient in each series be 1 and write the general
solution for the DE as
y = C1 y 1 + C2 y 2
This is equivalent to using the general recurrence relation and setting a0 = 1 with a1 = 0 to
find the even series and a0 = 0 with a1 = 1 to find the odd series.
This will work but a quicker approach for us to try will be to adjust the recurrence relation
for each case and use it to construct the general term as follows. For the even series we
want to find a2n so we adjust
1
an+2 = an
(n + 2)(n + 1)
to obtain a2n = . . .
The first step is to obtain an = . . . by subtracting 2 from the counter to give
1 1
an−2+2 = an = an−2 = an−2 .
(n − 2 + 2)(n − 2 + 1) n(n − 1)
Next replace n by 2n to obtain
1
a2n = a2n−2
2n(2n − 1)
and we have so far generated the corresponding recurrence relation for the even series from
the general recurrence relation. To find the general term a2n we simply expand as follows:
1
a2n−2 = a2n−2−2
(2n − 2)(2n − 2 − 1)
1
= a2n−4
(2n − 2)(2n − 3)
1
a2n−4 = a2n−4−2
(2n − 4)(2n − 4 − 1)
1
= a2n−6
(2n − 4)(2n − 5)
136 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
So we have
1 1 1
a2n = a2n−2 = ⋅ a2n−4
2n(2n − 1) 2n(2n − 1) (2n − 2)(2n − 3)
1 1
= ⋅ a2n−6
2n(2n − 1)(2n − 2)(2n − 3) (2n − 4)(2n − 5)
1
So the last few factors will be . . . a0 and we have putting it all together
4⋅3⋅2⋅1
1
a2n = a0
2n(2n − 1)(2n − 2)(2n − 3) . . . 4 ⋅ 3 ⋅ 2 ⋅ 1
Now we have the general formula for a2n the next step is to simplify it. We look for factors
of some number (2, 3 etc. ) and any factorials (n!, 2n! etc. ) For this example we have
quickly
2n(2n − 1)(2n − 2)(2n − 3) . . . ⋅ 4 ⋅ 3 ⋅ 2 ⋅ 1 = (2n)!
and we obtain
1
a2n = a0
(2n)!
giving
∞ ∞
x2n
y1 = ∑ a2n x2n = ∑
n=0 n=0 (2n)!
x2 x4
=1+ + + ...
2! 4!
as we found before.
For the odd series we want to find a2n+1 so we adjust
1
an+2 = an
(n + 2)(n + 1)
giving
1
a2n+1 = a1
(2n + 1)!
and if we set a1 = 1 we have
1
a2n+1 = a1
(2n + 1)!
giving
∞ ∞
x2n+1
y2 ∑ a2n+1 x2n+1 = ∑
n=0 n=0 (2n + 1)!
x x3 x5
= + + + ...
1! 3! 5!
x3 x5
=x+ + + ...
3! 5!
as we found before.
Series convergence
It is usually a good idea to check that the series solutions obtained converge (at least for
some region) especially if a numerical solution is wanted. For
∞
x2n
y1 = ∑
n=0 (2n)!
138 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
∣x2 ∣
= lim =0
n→∞ (2n + 2)(2n + 1)
∣x2 ∣
= lim =0
n→∞ (2n + 3)(2n + 2)
Step 1 We identify and classify the singular points of the D.E. as shown in the next sec-
tion.
Step 2 If x = 0 is an ordinary point (i.e. x = 0 is not a singular point), then we assume
that ∞
y = ∑ an x n
n=0
is a solution (where a0 , a1 , a2 , . . . must be found), and substitute for y and its deriva-
tives in the given D.E.
Note: If x = 0 is a singular point, then we have to assume that
∞
y = x c ∑ an x n
n=0
Step 5 If we have the correct number of arbitrary constants in our series solution, we have
at once the general solution of the D.E. (If we do not have the general solution,
then more “powerful”’ methods, beyond the scope of this subject, are required to
obtain the complete solution!)
Note that in the next section a general theorem is given which “controls” in advance
the situation that will occur in Step 5.)
Note
If the coefficient functions, a2 , a1 or a0 , contain a denominator involving x, then we must
first “clear” these denominators by multiplying each term in the D.E. by the appropriate
factor(s). (Thus, a2 (x), a1 (x) and a0 (x) will usually be polynomials.)
Difficulties can occur in the series solution of such L.D.E.s with variable coefficients if the
leading coefficient, here a2 (x) for y ′′ , is zero. Thus, having simplified the D.E. to include
no denominators, we must solve for a2 (x) = 0. The roots of this equation, if any, are called
the singular points of the D.E. All other values of x are called ordinary points i.e. x = a
is called an ordinary point of the D.E. if a2 (a) ≠ 0.
The next step is to write the DE in standard form. Providing a2 (x) ≠ 0, divide by a2 (x) to
obtain
y ′′ + p(x)y ′ + q(x)y = 0
If x = a is a singular point of the D.E. (i.e. if p(a) is not defined), then we classify the point
x = a further as follows.
If lim(x − a)p(x) and lim(x − a)2 q(x) both exist, then x = a is called a regular singular
x→a x→a
point.
If either of the limits does not exist, then x = a is called an irregular singular point.
Examples
y ′′ − 2xy ′ + 3y = 0 has no singular points. In particular, x = 0 is an ordinary point.
1 x
xy ′′ + y ′ + xy = 0 has a singular point at x = 0. Since lim x ⋅ = 1 and lim x2 = 0, hence
x→0 x x→0 x
x = 0 is a regular singular point.
2x2 y ′′ − xy ′ + (1 + x)y = 0 has a regular singular point at x = 0. (Why?)
2x2 (x + 2)y ′′ + (x + 2)y ′ + 3xy = 0 has a regular singular point at x = −2 and an irregular
singular point at x = 0. (Why?)
Tutorial 4 Question 1
140 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
∞
i. If x = a is an ordinary point of the D.E., then the substitution y = ∑ an (x − a)n will
n=0
lead to two series solutions involving two arbitrary constants and hence to the general
solution of the D.E. (See Examples 1 and 2 in the next Section.)
∞
ii. If x = a is a regular singular point, then y = ∑ an (x − a)n may lead to one series
n=0
solution. (See Example.) In general, if x = a is a regular singular point, then the
∞
substitution y = (x − a)c ∑ an (x − a)n will lead (through the method of Frobenius),
n=0
to a particular solution or to the general solution. (See Example and method of
Frobenius in section on Series solution about a regular singular point.)
iii. If x = a is an irregular singular point, then Frobenius series solutions may or may not
exist.
Note
We will consider only power series solutions in powers of x, i.e. about x = 0. Thus for our
D.E. solutions we will be concerned only about the classification of x = 0 If x = 0 is an
∞
ordinary point of the D.E., then we let y = ∑ an xn , and proceed as in the next section.
n=0
∞
If x = 0 is a regular singular point of the D.E., then we let y = ∑ an xn+c , and proceed
n=0
as in the section on Series solution about a regular singular point. (If x = 0 is an irregular
singular point, then we try the Frobenius method, but we may or may not get any solution!)
is a solution; find the necessary derivatives of the power series and substitute in the
D.E. We then try to find the value of each an , usually by the technique of “equating
coefficients”.
4.3. POWER SERIES SOLUTION ABOUT AN ORDINARY POINT 141
Notes
1. By the general theorem about the existence of series solutions stated previously, it is
known that when x = 0 is an ordinary point of the second order D.E., Step 1 (done
correctly!) will automatically yield the general solution, involving two linearly inde-
pendent series and two arbitrary constants. (See Examples 1 and 2 ) Hence Steps 2,3
and 4 are unnecessary in this case.
2. Nevertheless, it is often useful to carry out Step 2, i.e. to check each series solution in
the original D.E. (This should help you discover errors and also should improve your
confidence that the method works!)
3. The power series method may work if x = 0 is a regular singular point but all four
steps are likely to be needed. The methods for finding the other solution(s) are quite
“technical”.
4. More generally, if x = 0 is a regular singular point, it is better to use the same four
step approach but to assume that a Frobenius series of the form
∞
∑ an xn+c
n=0
is a solution, where the appropriate an and c values must be found. One such series
solution is always obtained in step one but all four steps may be necessary. Again,
quite complicated methods (beyond the scope of this subject) may be needed to find the
general solution.
Example 2
then ∞
y ′ = a1 + 2a2 x + 3a3 x2 + . . . = ∑ nan xn−1
n=1
and ∞
y ′′ = 2a2 + 2 ⋅ 3a3 x + 3 ⋅ 4a4 x2 + . . . = ∑ n(n − 1)an xn−2 .
n=2
142 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
or ∞ ∞ ∞
∑ n(n − 1)an xn−2 − ∑ 2nan xn + ∑ 3an xn = 0 .
n=2 n=1 n=0
In order to find the required values for the an ’s it is necessary to combine these infinite
series. To do this it is usually best to “adjust” each series, as necessary, to contain the same
general power of x.
Here, the last two series contain xn and so we re-write the first series as
∞ ∞
∑ n(n − 1)an xn−2 = ∑ (n + 2)(n + 1)an+2 xn
n=2 n=0
This is quite a “tricky” step and involves here replacing n by n + 2 to get the required xn
term. (Note that the new series has lower limit given by (n + 2) = 2, i.e. n = 0). A good
check is to examine the first term on each side. Here the first term on the left-hand side is
2(2 − 1)a2 x2−2 = 2a2 while on the r.h.s. it is (0 + 2) (0 + 1)a0+2 x0 = 2a2 , which checks.
Hence the D.E. can now be written as
∞ ∞ ∞
∑ (n + 2)(n + 1)an+2 xn − ∑ 2nan xn + ∑ 3an xn = 0
n=0 n=1 n=0
We now want to combine the infinite series by grouping powers of x. The series must all
“start” at the same value for n : here, n = 1 may be used.
∞ ∞ ∞
∴ {2a2 + ∑ (n + 2)(n + 1)an+2 xn } − ∑ 2nan xn + {3a0 + ∑ 3an xn } = 0
n=1 n=1 n=1
∞
∴ (2a2 + 3a0 ) + ∑ [(n + 2)(n + 1)an+2 − 2nan + 3an ] xn = 0
n=1
and so on.
All the an ’s are now known in value in terms of a0 (a2 , a4 , a6 , . . .) and a1 (a3 , a5 , a7 , . . .).
Thus a0 and a1 will be the required two arbitrary constants needed for the general solution.
The only problem remaining is to find a general expression for the coefficients in the power
series. This is often hard to do and can really only be done by “looking for patterns” in the
first coefficients. (You should thus be very careful not to destroy patterns by cancelling out
factors or by finding products too soon.)
Here we have
−3 −3
a2 = a0 = a0
2 2!
−1 −1
a3 = a1 = a1
2⋅3 3!
−3 ⋅ 1 −3 ⋅ 1
a4 = a0 = a0
2⋅3⋅4 4!
3 −1 ⋅ 3
a5 = a3 = a1
4⋅5 5!
5 −3 ⋅ 1 ⋅ 5
a6 = a4 = a0
5⋅6 6!
7 −1 ⋅ 3 ⋅ 7
a7 = a5 = a1
6⋅7 7!
⋮
The pattern in the coefficients should now be evident. (Try writing down a11 , a12 and then
calculating them to check your pattern). It is clear that the coefficients of the odd and even
powers of x are determined differently. We can find the general term using the recurrence
relation. If we want to find the coefficient of x2k , a2k , then we will have to let (n + 2) = 2k
in the recurrence relation.
4k − 7
∴ a2k = a2k−2 (n = 2k − 2)
(2k − 1)(2k)
To find the general form we construct the first few and last few factors as follows.
2(2k − 2) − 7
a2k−2 = a2k−2−2
(2k − 2 − 1)(2k − 2)
4k − 11
= a2k−4
(2k − 3)(2k − 2)
For a0 we set k = 1 giving
4−7 −3
a2 = a2−2 = a0
(2 − 1) ⋅ 2 1⋅2
and k = 2 gives
8−7 1
a4 = a4−2 = a2
(4 − 1) ⋅ 4 3⋅4
1 −3
= ⋅ a0
3⋅4 1⋅2
giving
4k − 7 4k − 11 1 ⋅ (−3)
a2k = ⋅ ... ⋅ ⋅ a0
(2k − 1)(2k) (2k − 3)(2k − 2) 3⋅4⋅1⋅2
(4k − 7)(4k − 11) . . . 1 ⋅ (−3)
= a0
(2k − 1)(2k)(2k − 3)(2k − 2) . . . 3 ⋅ 4 ⋅ 1 ⋅ 2
144 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
So we have
(4k − 7)(4k − 11) . . . ⋅ 1 ⋅ (−3)
a2k = a0
(2k)!
which only makes sense if we start from k = 1 giving
3
a2 = − a0
2
The series is ∞
∑ a2k x2k = a0 + a2 x2 + . . .
k=0
So if a0 = 1 we have
∞
y1 = 1 + ∑ a2k x2k
k=1
∞
(4k − 7)(4k − 11) . . . 1 ⋅ (−3) 2k
=1+∑ x
k=1 (2k)!
which may be a little easier to follow if we rewrite it as
∞
(−3) ⋅ 1 . . . (4k − 7) 2k
y1 = 1 + ∑ x
k=1 (2k)!
Similarly, with n = 2k − 1,
4k − 5
a2k+1 = a2k−1
(2k)(2k + 1)
As before, we construct the first few and last few factors as follow:
2(2k − 2) − 5
a2k−1 = a2k−2−1
(2k − 2)(2k − 2 + 1)
4k − 9
= a2k−3
(2k − 2)(2k − 1)
For a1 we set k = 1 to obtain
4−5 −1
a3 = a2−1 = a1
2(2 + 1) 2⋅3
and k = 2 gives
8−5 3
a5 = a4−1 = a3
4(4 + 1) 4⋅5
3 −1 3 ⋅ (−1)
= ⋅ a1 = a1
4⋅5 2⋅3 5⋅4⋅3⋅2
giving
4k − 5 4k − 9 3 ⋅ −1
a2k+1 = ⋅ ... a1
(2k)(2k + 1) (2k − 2)(2k − 1) 5⋅4⋅3⋅2
−1 ⋅ 3 . . . (4k − 5)
= a1
(2k + 1)(2k)(2k − 1)(2k − 2) . . . 5 ⋅ 4 ⋅ 3 ⋅ 2
−1 ⋅ 3 . . . (4k − 5)
= a1
(2k + 1)!
4.3. POWER SERIES SOLUTION ABOUT AN ORDINARY POINT 145
−3 −1 −3.1
∴ y = a0 + a1 x + ( a0 ) x 2 + ( a1 ) x 3 + ( a0 ) x 4
2 3! 4!
−1.3 −3.1.5
+( a1 ) x 5 + ( a0 ) x6 + . . .
5! 6!
i.e.
−3 2 −3.1 4 −3.1.5 6
y = a0 {1 + x + x + x + . . . .}
2! 4! 6!
−1 −1.3 5 −1.3.7 7
+ a1 {x + x3 + x + x + . . .}
3! 5! 7!
or
∞
−3.1 . . . (4k − 7) 2k
y = a0 {1 + ∑ x }
k=1 (2k)!
∞
−3.1 . . . (4k − 5) 2k+1
+ a1 {x + ∑ x }
k=1 (2k + 1)!
This is the general solution of the D.E. y ′′ − 2xy ′ + 3y = 0 since there are two arbitrary
constants a0 and a1 . If initial conditions were specified, then corresponding values for a0
and a1 could then be calculated. Fitting the initial conditions:
y(0) = C1 y1 (0) + C2 y2 (0) = 4
y1 (0) = 1 + 0 = 1 y2 (0) = 0 + 0
So we have
y(0) = C1 ⋅ 1 + C2 ⋅ 0 = C1 = 4
∞
−3 ⋅ 1 . . . (4k − 7)
y1′ = 0 + ∑ ⋅ 2kx2k−1
k=1 (2k)!
∞
−3 −3 ⋅ 1 . . . (4k − 7) 2k−1
= ⋅ 2x1 + ∑ x
3! k=2 (2k − 1)!
So y1′ (0) = 0
∞
−1 ⋅ 3 . . . (4k − 5)
y2′ = 1 + ∑ (2k + 1)x2k
k=1 (2k + 1)!
∞
−1 ⋅ 3 . . . (4k − 5) 2k
=1+∑ x
k=1 (2k)!
146 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
4k − 3
= lim ∣x∣2 = 0
k→∞ (2k + 2)(2k + 1)
4k − 1
= lim ∣x∣2 = 0
k→∞ (2k + 3)(2k + 2)
Example 3
Solution
Let ∞
y = ∑ an x n ,
n=0
then ∞ ∞
y ′ = ∑ nan xn−1 and y ′′ = ∑ n(n − 1)an xn−2 .
n=1 n=2
Hence,
∞ ∞ ∞ ∞
∑ n(n − 1)an xn − ∑ (n + 2)(n + 1)an+2 xn + ∑ 2nan xn − ∑ an xn = 0
n=2 n=0 n=1 n=0
and so
∞ ∞
∑ n(n − 1)an xn − 2 ⋅ 1a2 − 3 ⋅ 2a3 x − ∑ (n + 2)(n + 1)an+2 xn + 2a1 x
n=2 n=2
∞ ∞
+ ∑ 2nan xn − a0 − a1 x − ∑ an xn = 0
n=2 n=2
y = 4y1 − 3y2 .
Tutorial 4 Question 2
Video (by Dr. R. Wood): Series solution about a regular singular point
When x = 0 is a regular singular point of a given D.E., then the method of Frobenius
requires that we assume that the series solution is of the form
y = xc (a0 + a1 x + a2 x2 + . . .)
i.e. ∞
y = ∑ an xn+c
n=0
y ′′ + p(x)y ′ + q(x)y = 0
We then evaluate
p0 = lim(x − a)p(x) and q0 = lim(x − a)2 q(x)
x→a x→a
Note. These limits would already have been determined for the classification of x = a as
being a regular singular point.
The indicial equation can be constructed as
c(c − 1) + p0 c + q0 = 0
Example 1
We will consider first an example where the method of Frobenius works without any ”com-
plications”: Solve the D.E.
2x2 y ′′ − xy ′ + (1 + x)y = 0
Solution
Writing the DE in standard form we obtain
x ′ 1+x
y ′′ − y + y=0
2x2 2x2
or
1 ′ 1+x
y ′′ − y + y=0
2x 2x2
giving
−1 1+x
p(x) = and q(x) = .
2x 2x2
Then
−1 1 1
p0 = lim(x − a)p(x) = lim x ⋅ = lim − = −
x→a x→0 2x x→0 2 2
and
1+x 1+x 1+0 1
q0 = lim(x − a)2 q(x) = lim x2 ⋅ 2
= lim = = .
x→a x→0 2x x→0 2 2 2
So the indicial equation should be
c(c − 1) + (− 12 ) c + 12 = 0 multiply by 2
2c(c − 1) − c + 1 = 2c2 − 2c − c + 1
= 2c2 − 3c + 1 = 0
with roots c = 1, 21 .
Let ∞
y = a0 xc + a1 xc+1 + a2 xc+2 + . . . = ∑ an xn+c .
n=0
150 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
Then ∞
y ′ = ca0 xc−1 + (c + 1)a1 xc + . . . = ∑ (n + c)an xn+c−1 ,
n=0
and
∞
′′ c−2 c−1
y = (c − 1)c ⋅ a0 x + c(c + 1)a1 x + . . . = ∑ (n + c − 1)(n + c)an xn+c−2
n=0
Now ∞
∑ an xn+c+1
n=0
can be written as ∞
∑ an−1 xn+c .
n=1
So ∞ ∞
∑ {2(n + c − 1)(n + c) − (n + c) + 1}an xn+c + ∑ an−1 xn+c = 0,
n=0 n=1
or
∞
c
{2(c − 1)c − c + 1}a0 x + ∑ [{2(n + c − 1)(n + c) − (n + c) + 1}an + an−1 ] xn+c = 0.
n=1
The first coefficient, a0 , is always taken to be non-zero and so 2(c − 1)c − c + 1 = 0. This is
called the indicial equation (as we found before) and can now be solved for the appropriate
values of c, i.e.
2c2 − 3c + 1 = 0
(2c − 1)(c − 1) = 0
Hence the indicial roots are c1 = 1 and c2 = 21 .
Next, from (2), we obtain the recurrence relation for the coefficients, here
−1
an = an−1 , n ≥ 1,
2(n + c − 1)(n + c) − (n + c) + 1
and then substitute c1 and c2 in turn for c.
Note
The relationship between the values of c1 and c2 is crucial in determining whether only
one or two distinct series will be obtained by this substitution - see the Theorem following.
This example, in fact, demonstrates the “best” case.
Case I
4.4. SERIES SOLUTION ABOUT A REGULAR SINGULAR POINT 151
1 1 1 (−1)n
y2 = x 2 {1 − x+ x2 − . . . + xn + . . .}
1!1 2!1 ⋅ 3 n!1 ⋅ 3 . . . (2n − 1)
Thus the general solution of the given D.E. has yielded two different series solutions.
(These two series are, in fact, linearly independent (l.i.).
Thus the general solution is given by
y = Ay1 + By2
i.e.,
1 1
y =Ax {1 −
x+ x2 − . . .}
1!3 2!3 ⋅ 5
1 1 1
+Bx 2 {1 − x+ x2 − . . .}
1!1 2!1 ⋅ 3
where A and B are arbitrary constants.
Theorem A
As mentioned in the above example, the results we obtain on using the indicial roots in the
recurrence relation are determined by a Theorem which is as follows:
i. If the indicial roots differ by a constant which is not an integer or zero, the general
solution of the D.E. is always obtained.
152 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
ii. If the indicial roots differ by a non-zero integer, there are two possibilities:
a. The general solution is obtained by use of the smaller indicial root.
b. No solution is obtained by use of the smaller indicial root. However, in all cases
where the roots differ by a non-zero integer, one solution is always obtained by
using the larger root.
iii. If the indicial roots differ by zero, i.e. are equal, only one solution is obtained.
Notes
1. You will be required only to carry out “simple” solutions based on case (i).
2. To find the general solution in cases (ii)(b) and (iii) quite complicated methods must
be used, which are beyond the scope of this subject! Those methods are studied in
MTH418 (Topics in Calculus)
For completeness, a more general theorem is given below for dealing with D.E.s where
x = 0 is a regular singular point.
Theorem B
If x = 0 is a regular singular point of the D.E. p(x)y ′′ + q(x)y ′ + r(x)y = 0, then the D.E.
has two linearly independent series solutions y1 and y2 (and hence the general solution is
given by y = Ay1 + By2 ).
If c1 ≥ c2 are the indicial roots, then the l.i. solutions y1 and y2 are obtained as follows:
∞ ∞
a. If c1 − c2 is not an integer or zero, then y1 = xc1 ∑ an xn and y2 = xc2 ∑ bn xn where the
n=0 n=0
an ’s and bn ’s are obtained from the recurrence relation using c1 and c2 in turn.
∞
b. If c1 = c2 , then y1 = xc1 ∑ an xn where the an ’s are obtained from the recurrence relation
n=0
∞
and y2 = y1 ln x + xc1 ∑ bn xn where the bn ’s are obtained by substituting this form for
n=0
y2 in the given D.E. and again equating coefficients.
∞
c. If c1 − c2 = N, a positive integer, then y1 = xc1 ∑ an xn where the an ’s are obtained from
n=0
∞
the recurrence relation using c1 and y2 = ay1 ln x + xc2 ∑ bn xn where a and the bn ’s are
n=0
obtained by substituting this form for y2 in the given D.E.
Example
and so ∞
y ′ = ∑ (n + c)an xn+c−1
n=0
∞
y ′′ = ∑ (n + c)(n + c − 1)an xn+c−2
n=0
Now, on substituting into the D.E. and “adjusting” as before, we get that
∴ c(c − 1) + c = 0 i.e., c2 = 0 ∴ c = 0.
Trouble ahead: Case (ii) in the general existence theorem, and Case (b) in Theorem B!!
Coefficient of xc ∶
{(c + 1)c + (c + 1)}a1 = 0
∴ (c + 1)2 a1 = 0
But c ≠ −1 (c equals 0), and so a1 must be equal to zero here.
Recurrence relation
−1
an = an−2 , n ≥ 2
(n + c)2
Hence,
1
a2 = − a0
(c + 2)2
1
a3 = − a1 = 0
(c + 3)2
1 1
a4 = − 2
a2 = a0 ,
(c + 4) (c + 2) (c + 4)2
2
and so on.
Now, if we put c = 0, we get
1 1
a2 = − 2
a0 , a 4 = 2 2 a0 , . . . ,
2 2 ⋅4
and
a1 = a3 = a5 = . . . = 0
Thus, as expected, we do not obtain the general series solution but only the one series
solution
1 1
y = a0 (1 − 2 x2 + 2 2 x4 − . . .)
2 24
The series in brackets is known as J0 (x), the Bessel function of order zero. i.e. J0 (x) is a
“part” of the series solution of the given D.E.
154 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
Note
After considerable effort (not required by you here!) it can be shown that th second linearly
independent series solution has the “appalling” appearance
x2 x4 1 x6 1 1
y = J0 (x) ln x + { 2
− 2 2
(1 + ) + 2 2 2
(1 + + ) + . . .}
2 2 ⋅4 2 2 ⋅4 ⋅6 2 3
2. If x = 0 is an ordinary point, then determine the general solution of the D.E. using
∞
the power series, y = ∑ an xn
n=0
∞
3. If x = 0 is a regular singular point, use the Frobenius series y = xc ∑ an xn to deter-
n=0
mine the indicial roots, c1 ≥ c2
Tutorial 4 Question 3
4.6. TUTORIAL 4 155
4.6 Tutorial 4
Back to Chapter 4
Question 1
Question 2
a. y ′ = y − x
b. y ′′ + y = 0
c. y ′′ − y ′ − 2y = 0
2. Solve
(1 − x2 ) y ′′ − 2xy ′ + 6y = 0, y(1) = 1, y ′ (0) = 0
(2 + x2 ) y ′′ − xy ′ + 4y = 0, y(0) = 3, y ′ (0) = −1
is y = 3y1 − y2 , where
1 1
y1 (x) = 1 − x2 + x4 − x6 + . . .
6 30
and
1 7 5 19 7
y2 (x) = x − x3 + x − x + ....
4 160 1920
4. Textbook exercise set 8.3 Problems 19, 23, 25
Solutions
1. a.
c 2 c 3
y = (c + 1) + (c + 1)x + x + x + ...
2! 3!
= cex + x + 1
b.
x2 x4 x6 x3 x5
y = c1 (1 − + − + . . .) + c2 (x − + − . . .)
2! 4! 6! 3! 5!
= c1 cos x + c2 sin x
156 CHAPTER 4. SERIES SOLUTION OF DIFFERENTIAL EQUATIONS
c.
1 1
y = a0 (1 + x2 + x3 + x4 + . . .)
3 4
1 2 1 3 5 4
+ a1 (x + x + x + x + . . .)
2 2 24
This series solution must be capable of being transformed into the Taylor series for
c1 e2x + c2 e−x . (Why?) (Try to transform your solution thus!)
Question 3
a. 4xy ′′ + 2y ′ + y = 0
b. y ′′ + xy ′ + y = 0
c. xy ′′ + y = 0.
Answers
3 5
x x2 x3 1 x2 x2
a. y = A (1 − + − + . . .) + B (x 2 − + − . . .) which actually is represented
2! 4! 6! 3! 5!
√ √
by y = A cos x + B sin x
x2 x4 x6 x3 x5 x7
b. y = A (1 − + − + . . .) + B (x − + − + . . .)
2 2⋅4 2⋅4⋅6 3 3⋅5 3⋅5⋅7
(A power series solution, as expected. Why?)
x2 x3 x4
c. y = A (x − + − + . . .)
1!2! 2!3! 3!4!
(not the general solution).
d. See textbook.
Chapter objectives
5.1 Introduction
Recall. The concept of a function. The elementary functions studied in earlier subjects.
In Computer Aided Mathematics 1 with Applications (MTH101), the elementary mathe-
matical functions were studied and their main properties noted. In particular, algebraic
functions, the exponential and logarithmic and trigonometric functions were considered.
In addition to these, there are a great many special functions. Some of these are specialized
in the sense of having a very specific use. However, others have a wide use in engineering,
applied mathematics, statistics and the further general theory of mathematics.
This part of this subject is concerned almost entirely with just one of the “non-elementary”
special functions - the Gamma function.
This is one of the more important of the non-elementary functions, with uses in applied
mathematics and in mathematical statistics.
It is also called the generalized factorial function, for reasons seen in the next section.
First, consider the definition as a mathematical statement. Note that the integral is improper
because of the infinite upper limit of integration.
Recall. In evaluating, an improper integral
∞ N
∫0 f (t)dt = lim ∫ f (t)dt
N →∞ 0
The integral is, however, definite. The symbol t is an integration parameter which does not
appear in the result of the definite evaluation from zero to infinity. That is to say,
∞
∫0 e−t tx−1 dt
should be seen as a function of x [which is exactly what equation (5.2.1) says; lefthand-side
equals right-hand-side and, by notation, LHS is a function of x and nothing else].
Equation (5.2.1), then, defines a function of the real variable x. (The Gamma function of a
complex variable z can be similarly defined by the same form of equation.) It is important
to note that the definition (5.2.1) is only for x > 0. At this stage, therefore, the Gamma
function is defined only for positive values although a later extension removes this (see
section on Extended definition).
The relation between the Gamma function and the ordinary factorial
function
Recall.
n! = n(n − 1)(n − 2) . . . 3 ⋅ 2 ⋅ 1,
and if
f ′ (x)
lim = L,
x→c g ′ (x)
then
f (x)
lim =L
x→c g(x)
tx xtx−1
lim = lim = ... = 0
t→∞ et t→∞ et
The remaining integral is the Gamma function of x + 1. Hence the important recurrence
relation
Γ(x + 1)
Γ(x) =
x
or
xΓ(x) = Γ(x + 1). (5.2.2)
This is the functional relation of the Gamma function.
Example 1
Γ(2) = 1 ⋅ Γ(1) = 1
Γ(3) = 2 ⋅ Γ(2) = 2 ⋅ 1 = 2!
Γ(4) = 3 ⋅ Γ(3) = 3 ⋅ 2 ⋅ 1 = 3!
Γ(5) = 4 ⋅ Γ(4) = 4 ⋅ 3 ⋅ 2 ⋅ 1 = 4!
Γ(6) = 5 ⋅ Γ(5) = 5 ⋅ 4 ⋅ 3 ⋅ 2 ⋅ 1 = 5!
and so on. . .
In general, then, for integer n, we find that
Γ(n + 1) = n! n = 1, 2, 3, . . . (5.2.3)
Hence the connection between the Gamma function and the ordinary factorial function.
Note the change in the “argument” - the n plus one.
The ideas of the previous section can be extended in two ways: firstly to define a general-
ized factorial; secondly (and of a more immediate importance) to extend the definition of
the Gamma function.
As to the first, equation (5.2.3) can be generalized to define x! as
x! = Γ(x + 1)
160 CHAPTER 5. GAMMA FUNCTION (SPECIAL FUNCTIONS)
This can be used as a definition for fractional or even negative numbers [a concept which
obviously does not fit in with the particular case (positive integer n) of the elementary
factorial function]. Thus, 12 ! = Γ ( 23 ) and (− 12 )! = Γ ( 12 ) .
What is of interest here is that now
0! = Γ(1) = 1
a result which in previous subjects may have been stated without any further justification.
Recall, for example, the evaluation of combinations and of certain discrete probabilities in
statistics.
Now, to extend the present (x > 0) definition of the Gamma function, consider the repeated
application of the recurrence:
The Gamma function remains undefined for zero and all negative integers.
Tables of values of the Gamma function have, of course, been compiled. These cover only
a unit interval of x, for example 1 ≤ x ≤ 2. Other values can be found simply by using the
recurrence formula. A short table follows.
5.2. DEFINITION OF THE GAMMA FUNCTION 161
Example 2
By careful consideration of its definition, properties and values, the graph of the Gamma
function can be drawn as follows:
The general features of the graph of the Gamma function above should be noted. Especially
note the asymptotes at x = 0, −1, −2, −3, . . ., the value of the function increasing without
limit as these arguments are approached. For negative x, the function assumes alternately
positive and negative values in successive intervals of definition, due, of course, to the sign
of the denominator in the recurrence relation (8).
As x → −∞, the local maxima and minima tend to zero, with a flattening of the peaks. The
curve segments are asymmetric. With increasing positive x, Γ(x) increases very rapidly
indeed.
Note. The values of the Gamma function of half-integral arguments, Γ (k + 21 ) can be
calculated from
There is also a multiplication formula for Gamma functions, which often proves useful. We
state it here, without proof:
π
Γ(x) ⋅ Γ(1 − x) = , x ≠ 0, ±1, ±2, . . .
sin xπ
√
Using this, with x = 12 , confirms that Γ ( 12 ) = π. Another example would be:
1 2 π 2π
Γ( ) ⋅ Γ( ) = π = √ .
3 3 sin 3 3
5.2. DEFINITION OF THE GAMMA FUNCTION 163
Tutorial 5 Question 1
164 CHAPTER 5. GAMMA FUNCTION (SPECIAL FUNCTIONS)
5.3 Tutorial 5
Back to Chapter 5
Question 1
1. Γ(1.7)
2. Γ(3.3)
3. Γ(5.7)
4. Γ(0.6)
5. Γ(−1.4)
6. Γ(−2.3)
Answers
1. 0.90864
2. 2.6834
3. 72.528
4. 1.4892
5. 2.6593
6. –1.4471
Back to Chapter 5
Chapter 6
Laplace transforms
Chapter objectives
6.1 Introduction
In this chapter, we will examine a very important concept, the Laplace Transform, which
is an example of a useful family called Integral Transforms. Such transforms are very sig-
nificant in a number of theoretical and applied areas of mathematics, numerical analysis
and computing. We will consider just a brief introduction to the topic here, and then ex-
amine how the Laplace Transform can be used to solve some of the linear D.E.s that we
examined in Chapter 2.
Video (by Dr. D. Demskoy): Laplace transform: definition and simplest properties
An integral transform is a relation of the form
β
F (s) = ∫ K(s, t)f (t)dt
α
Transforms are useful in making the solution of a problem, especially linear differential
equations, simpler.
We shall see that the Laplace transform reduces a differential equation to an algebraic
equation, and so simplifies its solution. If f (t) is defined for t ≥ 0, then the Laplace
transform of f is a function F defined by
∞
F (s) = ∫ e−st f (t)dt
0
(Here, e−st is the kernel.) The domain of the transform function F is the set of real values
of s for which the (improper) integral exists. Since e−st is a decreasing function of s, this
domain is s > s0 for some s0 determined by f .
The Laplace transform is usually represented by the operator L. Thus
∞
L{f (t)} = F (s) = ∫ e−st f (t)dt .
0
Example 1
ii. f approaches a finite limit as the end-points of each sub-interval are approached from
within the sub-interval.
We now state the Theorem for the existence of Laplace transforms. Theorem 1 Suppose:
b. ∣f (t)∣ ≤ M es0 t when t ≥ 0. Then the Laplace transform exists for s > s0 .
We have already seen what is meant by piecewise continuous. To understand the second
condition, we look at some examples of functions satisfying and not satisfying (b).
i. f (t) = sin t
−1 ≤ sin t ≤ 1 i.e. ∣ sin t∣ ≤ 1 = 1e0⋅t
∴ M = 1, s0 = 0.
∴ f (t) = sin t satisfies condition b.
Example 2
Example 3
Linearity of L
The Laplace transform is a linear operator.
Theorem 2
Suppose L{f (t)} = F (s) for s > s1 , and L{g(t)} = G(s) for s > s2 . Then,
Proof ∞
L {c1 f (t) + c2 g(t)} = ∫ e−st [c1 f (t) + c2 g(t)] dt
0
∞ ∞
= c1 ∫ e−st f (t)dt + c2 ∫ e−st g(t)dt
0 0
= c1 L{f (t)} + c2 L{g(t)}
= c1 F (s) + c2 G(s) for s > max (s1 , s2 ) .
We apply the linearity property in the following example.
Example 4
Example 5
Tutorial 6 Question 1
Read textbook sections 7.3 and 7.4 (Properties of the Laplace transform and
Inverse Laplace transform)
In order to use Laplace transforms, it is usually necessary to be able to recover the function
f (t) when L{f (t)} = F (s) is known.
This is called inverting the transform, and can be regarded as the action on F (s) of an
operator L−1 which is the inverse of L.
A general method of inversion involves a complex integral which is beyond the scope of
this subject. We will therefore use known transforms as a basis for inversion. This is
illustrated in the following example.
Example 6
If
1
F (s) = , a ≠ b,
(s − a)(s − b)
find L−1 {F }.
We use partial fractions on
1
.
(s − a)(s − b)
Let
1 A B
= +
(s − a)(s − b) s − a s − b
A(s − b) + B(s − a)
=
(s − a)(s − b)
∴ 1 ≡ A(s − b) + B(s − a)
1
s = a gives A =
a−b
−1
s = b gives B =
a−b
6.4. TRANSFORMS OF DERIVATIVES 171
1 1 1 1
∴ = ( − )
(s − a)(s − b) a − b s − a s − b
1 1 1
∴ L−1 {F (s)} = L−1 { ( − )}
a−b s−a s−b
1 −1 1 1 −1 1
= L ( )− L ( )
a−b s−a a−b s−b
1 at 1 bt
= e − e (from Example 2)
a−b a−b
1
= (eat − ebt ) .
a−b
We intend to use Laplace transforms to solve differential equations. Hence we need to con-
sider what happens when we find L {f ′ (t)} . Roughly speaking, differentiation of functions
corresponds to the multiplication of transforms by s Hence the Laplace transform reduces
a calculus problem to an algebraic problem.
To be more precise, the theorem that describes the transform of derivatives is as follows.
Theorem 3
Suppose the following conditions are satisfied
Then L {f ′ (t)} exists for s > s0 and L {f ′ (t)} = sL{f (t)} − f (0).
We will prove this theorem for the case of f ′ (t) continuous. (The proof for f ′ (t) piecewise
continuous is similar, but requires the range of integration to be broken up into parts.)
Proof
∞
L {f ′ (t)} = ∫ e−st f ′ (t)dt (definition)
0
T
= lim ∫ e−st f ′ (t)dt
T →∞ 0
∞
= lim [e−st f (t)]0 + s ∫ e−st f (t)dt (using integration by parts)
T
T →∞ 0
= lim [e−sT f (T ) − f (0)] + sL{f (t)}
T →∞
= sL{f (t)} − f (0) for s > s0
as required.
We can go on to find the transforms of higher derivatives.
and
L {f ′′′ (t)} = sL {f ′′ (t)} − f ′′ (0)
= s [s2 L{f (t)} − sf (0) − f ′ (0)] − f ′′ (0)
= s3 L{f (t)} − s2 f (0) − sf ′ (0) − f ′′ (0).
Using mathematical induction, we can show, that, for n a positive integer,
L {f (n) (t)} = sn L{f (t)} − sn−1 f (0) . . . − sf (n−2) (0) − f (n−1) (0)
Example 7
Example 8
Example 9
Table 1
1
1 1 1/s 6 eat
s−a
s
2 t 1/s2 7 cos ωt
s + ω2
2
ω
3 t2 2!/s3 8 sin ωt
s + ω2
2
n! s
4 tn n+1
9 cosh at 2
s s − a2
Γ(a + 1) a
5 ta 10 sinh at
sa+1 s 2 − a2
Tutorial 6 Question 2
174 CHAPTER 6. LAPLACE TRANSFORMS
Example 10
We assume the solution y = y(t), and its first two derivatives satisfy the conditions for
applying the Laplace transform. Then
L (y ′′ − y ′ − 2y) = L(0)
∴ L (y ′′ ) − L (y ′ ) − 2L(y) = 0 (linearity property)
∴ s2 L(y) − sy(0) − y ′ (0) − sL(y) + y(0) − 2L(y) = 0 (using transform of derivatives)
∴ (s2 − s − 2) L(y) + (1 − s)y(0) − y ′ (0) = 0
∴ (s2 − s − 2) Y (s) = s − 1
s−1
∴ Y (s) =
s2 −s−2
Let
s−1 A B A(s + 1) + B(s − 2)
= + =
s2 − s − 2 s − 2 s + 1 (s − 2)(s + 1)
∴ s − 1 ≡ A(s + 1) + B(s − 2)
1
s = 2 gives A =
3
2
s = −1 gives B = .
3
1 2
∴ Y (s) = +
3(s − 2) 3(s + 1)
1 2 1
∴ y(t) = e2t + e−t since L (eat ) =
3 3 s−a
(Table 1, No. 6)
We now look at a second order non-homogeneous differential equation.
Example 11
Example 11
∴ r = ±1, ±i
∴ y(t) = Aet + Be−t + C cos t + D sin t
176 CHAPTER 6. LAPLACE TRANSFORMS
We then use the initial conditions to find the values of A, B, C and D. Alternatively, we
can take Laplace transforms.
L (y (4) − y) = 0
∴ L (y (4) ) − L(y) = 0
∴ s4 Y (s) − s3 y(0) − s2 y ′ (0) − sy ′′ (0) − y ′′′ (0) − Y (s) = 0
∴ s4 Y (s) − s2 − Y (s) = 0
∴ (s4 − 1) Y (s) = s2
s2 s2
∴ Y (s) = 4 = 2
s − 1 (s + 1) (s2 − 1)
1 1
2 2
= +
s2 + 1 s2 − 1
and using partial fractions we obtain
1
y(t) = (sin t + sinh t) (Table 1, Nos. 8, 10).
2
From these examples, we see the advantages of using the Laplace transform in initial value
problems. There is no need to determine a general solution, and subsequently find values
for arbitrary constants. The given initial conditions are included automatically when the
transform is taken.
Tutorial 6 Question 3
Example 13
Find f (t) if
3
L{f (t)} = .
s2 +s
Now
3 1 3
= ( )
s2 + s s s + 1
and
3
L−1 ( ) = 3e−t
s+1
therefore
1 3 t
L−1 { ( )} = ∫ 3e−τ dτ
s s+1 0
= −3 [e−τ ]0
t
= −3e−t + 3.
Tutorial 6 Question 4
To obtain a better idea of the power of the Laplace transform, we look at properties which
concern shifting on the s-axis. We can also have shifting on the t-axis, but we leave treat-
ment of that to the subject MTH418 (Topics in Calculus).
Theorem 5 (s-shifting)
If
L{f (t)} = F (s) where s > s0
then
L {eat f (t)} = F (s − a)
and
L−1 {F (s − a)} = eat f (t) (inverse).
Proof ∞
F (s) = ∫ e−st f (t)dt (definition)
0
∞
∴ F (s − a) = ∫ e−(s−a)t f (t)dt
0
∞
=∫ e−st eat f (t)dt
0
= L {eat f (t)}
178 CHAPTER 6. LAPLACE TRANSFORMS
Example 14
Find L (eat tn ) .
We know that
n!
L (tn ) =
sn+1
n!
∴ L (eat tn ) = F (s − a) =
(s − a)n+1
Example 15
1
∴ y(t) = L−1 { }
(s + 1)2 + 1
1
= L−1 {F [s − (−1)]} where F (s) =
s2 +1
= e−t sin t
Tutorial 6 Question 5
6.8. TUTORIAL 6 179
6.8 Tutorial 6
Back to Chapter 6
Question 1
1. Sketch the graphs of the following functions, and determine whether f is continuous,
piecewise continuous, or neither on the interval 0 ≤ t ≤ 3 :
⎧
⎪
⎪ t2 , 0 ≤ t ≤ 1
⎪
⎪
i. f (t) = ⎨2 + t,1 < t ≤ 2
⎪
⎪
⎪
⎩ 6 − t, 2 < t ≤ 3
⎪
⎧
⎪
⎪ t2 , 0 ≤ t ≤ 1
⎪
ii. f (t) = ⎨ 1 , 1 < t ≤ 2
⎪
⎪
⎩ 3−t , 2<t≤3
⎪
i. 3t + 4
ii. eat+b (a, b constants)
iii. cos(at) (a constant)
Answers
Piecewise continuous
1. i.
Continuous
ii.
3 4
2. i. +
s2 s
eb
ii.
s−a
s
iii. 2
s + a2
Detailed solutions to Question 1.2
Back to Chapter 6
Question 2
1
ii.
(s − 4)(s − 1)
3s
iii. 2
s + 2s − 8
s2 − 6s + 4
iv. 3
s − 3s2 + 2s
2. Use the table of Laplace transforms to find the transforms of the following functions:
i. t2 + at + b (a, b constants )
ii. cos(ωt + θ) (ω, θ constants)
iii. cos2 ωt (ω constant )
iv. 3t5
v. cosh 4t
Answers
1. i. 1 − e−t
1
ii. (e4t − et )
3
iii. 2e−4t + e2t
iv. 2 + et − 2e2t .
2 a b
2. i. 3
+ 2+
s s s
s cos θ − ω sin θ
ii.
s2 + ω 2
1 s
iii. Use: cos2 ωt = 21 + 12 cos 2ωt. Ans: + 2
2s 2s + 8ω 2
3 ⋅ 5!
iv.
s6
s
v. 2 .
s − 16
2ω 2
3.
s (s2 + 4ω 2 )
s2 − 1
4. 2
(s2 + 1)
5. see textbook
Question 3
1. 4y ′′ + π 2 y = 0, y(0) = 0, y ′ (0) = 1.
2. y ′′ + 5y ′ + 6y = 0, y(0) = 0, y ′ (0) = 1.
Answers
2
1. y = π sin π2 t
2. y = e−2t − e−3t
1
3. y = cos 5t + 25 t.
4. see textbook
Question 4
Use
1 t
L−1 { F (s)} = ∫ f (τ )dτ
s 0
4
1.
s3 − 4s
1
2. here a is a constant
s2 + as
8
3. Hint: Apply the formula twice.
s4 − 4s2
Answers
1. cosh 2t − 1
2. 1
a (1 − e−at )
3. sinh 2t − 2t.
Question 5
i. t2 e−2t
ii. e−t cos t
iii. e−2t sin 3t
iv. 2t3 e−t/2
2. Use the inverse form of the s -shifting theorem to find f (t) if L(f ) equals:
1
i. 3
(s + 12 )
1
ii.
s2 + 2s + 5
6
iii. 2
s − 4s − 5
3. Using Laplace transforms, and the s -shifting theorem, solve the differential equa-
tions:
i. y ′′ − 4y ′ + 5y = 0, y(0) = 1, y ′ (0) = 2
ii. 9y ′′ − 6y ′ + y = 0, y(0) = 3, y ′ (0) = 1
iii. y ′′ + 2y ′ + 5y = 9 cosh 2t + 4 sinh 2t, y(0) = 1, y ′ (0) = 2
Answers
2
1. i.
(s + 2)3
s+1
ii. 2
s + 2s + 2
3
iii.
(s + 2)2 + 9
12
iv. 4
(s + 12 )
1 2 −t/2
2. i. 2t e
1 −t
ii. 2 e sin 2t
iii. 2e2t sinh 3t
3. i. y = e2t cos t
ii. y = 3et/3
iii. y = e−t sin 2t + cosh 2t.
4. see textbook
184 CHAPTER 6. LAPLACE TRANSFORMS
5. see textbook
7.1 Tutorial 1
Question 2
Part 1.i
The equation is
3xdx + 4ydy = 0.
We denote
M = 3x, N = 4y.
Calculating the derivatives, we obtain
∂y M = 0, ∂x N = 0.
We see that they are equal, so the equation is exact. Further, we calculate
F = ∫ M dx = ∫ 3xdx = 23 x2 + ϕ(y),
Part 1.ii
(y − x) dx + (x + y) dy = 0.
So, we denote
M = y − x, N = x + y.
Then, calculating the derivatives, we obtain
∂M ∂N
= 1, = 1.
∂y ∂x
We see that they are equal, so the equation is exact. Further, we calculate
F = ∫ M dx = ∫ (y − x) dx = − 21 x2 + yx + ϕ(y),
∂F
= N,
∂y
we obtain
x + ϕ′ = x + y
which implies that
ϕ = 21 y 2 + C.
Therefore the general solution is − 21 x2 + yx + 12 y 2 + C = 0 which can be re-written as
x2 − 2xy − y 2 = C
Part 1.iii
−xdx + (x + y) dy = 0.
So, we denote
M = −x, N = x + y.
Calculating the derivatives, we obtain
∂M ∂N
= 0, = 1.
∂y ∂x
We see that they are not equal, so the equation is not exact!
Back to Tutorial 1 Question 2
7.1. TUTORIAL 1 187
Part 1.iv
Then, we denote
M = ye−x − sin x, N = −e−x − 2y.
Calculating the derivatives, we obtain
∂M ∂N
= e−x , = e−x .
∂y ∂x
We see that the derivatives are equal, so the equation is exact. Further, we calculate
∂F
= N,
∂y
we obtain
−e−x + ϕ′ = −e−x − 2y
which implies that
ϕ = −y 2 + C.
Therefore the general solution is
−ye−x + cos x − y 2 + C = 0
or
ye−x − cos x + y 2 = C
Part 1.v
The equation is
(x2 + x) dy + (2yx + 1 + 2 cos x) dx = 0.
So, we denote
M = 2yx + 1 + 2 cos x, N = x2 + x.
Calculating the derivatives, we obtain
∂M ∂N
= 2x, = 2x + 1.
∂y ∂x
We see that they are not equal, so the equation is not exact!
Back to Tutorial 1 Question 2
188 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
Part 2.i
The equation is
2xydx + (1 + x2 ) dy = 0.
We denote:
M = 2yx, N = x2 + 1.
Calculating the derivatives, we obtain
∂M ∂N
= 2x, = 2x.
∂y ∂x
Hence the derivatives are equal and the equation is exact. Further, we calculate
yx2 + y + C = 0.
Part 2.ii
The equation is
(2yx + 2y 2 e2x ) dx + (x2 + 2ye2x ) dy = 0.
So, we denote
M = 2yx + 2y 2 e2x , N = x2 + 2ye2x .
Calculating the derivatives, we obtain
∂M ∂N
= 2x + 4ye2x , = 2x + 4ye2x .
∂y ∂x
We see that they are equal, so the equation is exact. Further, we calculate
yx2 + y 2 e2x + C = 0.
Part 3
The equation is
ydx + (4x − y 2 ) dy = 0
So, we find
M = y, N = 4x − y 2 .
We note that
∂x N − ∂y M 3
Q= =
M y
is a function of y only, so the integrating factor can be calculated:
3dy
IF = exp (∫ Qdy) = exp (∫ ) = y3.
y
Multiplying the initial equation by the integrating factor we calculate an exact equation!
The equation takes the form
y 4 dx + (4y 3 x − y 5 ) dy = 0.
Again, we denote
M = y 4 , N = 4y 3 x − y 5 .
Calculating the derivatives, we obtain
∂M ∂N
= 4y 3 , = 4y 3 .
∂y ∂x
Now we see that the equation is exact. Further, we calculate
F = ∫ M dx = ∫ y 4 dx = y 4 x + ϕ(y),
we obtain
4y 3 x + ϕ′ = 4y 3 x − y 5
which implies that
ϕ = − 61 y 6 + C.
Therefore the general solution is y 4 x − 61 y 6 + C = 0 which is equivalent to
6y 4 x − y 6 = C.
Question 3
Part 3.i
x Substitution: u = x2 + 2
−3 ∫ 2 dx = du = 2xdx
x +2
du
−3 ∫ 12 = Constant multiple rule: ∫ cf (x)dx = c ∫ f (x)dx
u
du
− 32 ∫ =
u
− 23 ln u + C
= − 23 ln (x2 + 2) + C
7.1. TUTORIAL 1 191
Part 3.ii
Part 3.iii
4 ln (y 2 + 1) − x2 = C,
C = −1
hence
4 ln (y 2 + 1) − x2 = −1
or
x2 − 4 ln (y 2 + 1) = 1.
Question 3.iv
where C is a constant of integration. Further, solving for y we get the explicit solution
Part 3.v
Part 3.vi
hence
dy dx
∫ sec2 y = ∫ 1 + x2 .
7.1. TUTORIAL 1 193
dy
Use formula: 1
= cos2 y
∫ sec2 y = sec2 y
2 cos y 2 = 1
cos (2y) + 1
∫ cos ydy = Use formula: 2 2
1 1
∫ ( 2 cos (2y) + 2 )dy =
1 1
∫ 2 cos (2y) dy + ∫ 2 dy
1
=∫ 2 cos (2y) dy + 21 y =
1 1 Substitution: u = 2y
2 ∫ cos (2y) dy + 2 y = du = 2dy
1
4 sin (2y) + 12 y + C
Back to Tutorial 1 Question 3
Separable equation:
dy
= − exp (cos x) sin xdx
y2
hence
dy
∫ y 2 = ∫ − exp (cos x) sin xdx.
Integrating the latter (using substitution u = cos x) we get
1
− = exp (cos x) + C,
y
where C is a constant of integration. Further, solving for y we get the explicit solution
1 1
y(x) = − = ,
exp (cos x) + C A − exp (cos x)
where A = −C.
Back to Tutorial 1 Question 3
hence
dy
∫ √ = ∫ 2 cos xdx.
y+1
Integrating the latter we have
√
2 y + 1 = 2 sin x + C,
where C is a constant of integration. Further, solving for y we find the explicit solution
C 2
y(x) = (sin x + ) − 1.
2
Question 4
The equation is
2xydx + (y 2 − 3x2 ) dy = 0
7.1. TUTORIAL 1 195
Then, we denote
M = 2xy, N = y 2 − 3x2 .
Note that
∂x N − ∂y M 4
Q= =−
M y
is a function of y only, so the integrating factor can be calculated:
4dy 1
IF = exp (∫ Qdy) = exp (∫ − ) = 4.
y y
Multiplying the initial equation by the integrating factor we get an exact equation! The
equation is
2xdx (−y 2 + 3x2 ) dy
− =0
y3 y4
So, we have
2x −y 2 + 3x2
M= , N = − .
y3 y4
Calculating the derivatives, we obtain
∂M 6x ∂N 6x
=− 4, =− 4.
∂y y ∂x y
We see that they are equal, so the equation is exact. Further, we calculate
2xdx x2
F = ∫ M dx = ∫ = 3 + ϕ(y),
y3 y
∂F
= N,
∂y
we obtain
3x2 ′ −y 2 + 3x2
− + ϕ = −
y4 y4
which implies that
1
ϕ = − + C.
y
Therefore the general solution is
x2 1
− + C = 0.
y3 y
The equation1 is
(x4 − x + y) dx − xdy = 0
So, we get
M = x4 − x + y, N = −x.
We note that
∂y M − ∂x N 2
P= =−
N x
is a function of x only, so the integrating factor is
2 1
IF = exp (∫ P dx) = exp (∫ − dx) = 2 .
x x
Multiplying the initial equation by the integrating factor we obtain an exact equation! The
equation is
(x4 − x + y) dx dy
− =0
x2 x
So, we have
x4 − x + y 1
M= 2
, N =− .
x x
Calculating the derivatives, we obtain
∂M 1 ∂N 1
= 2, = 2.
∂y x ∂x x
We see that they are equal, so the equation is exact. Further, we calculate
(x4 − x + y) dx 1 3 y
F = ∫ M dx = ∫ = 3 x − − ln ∣x∣ + ϕ(y),
x2 x
where ϕ(y) is a function of y only. Substituting F into
∂F
= N,
∂y
we obtain
1 1
− + ϕ′ = −
x x
which implies that
ϕ = C.
Therefore the general solution is
1 3 y
3x − − ln ∣x∣ + C = 0
x
or after multiplication by x it takes the form
1 4
3x − y − x ln ∣x∣ + Cx = 0.
1
It is slightly different in 9th ed
7.1. TUTORIAL 1 197
The equation is
(y 2 + 2xy)dx − x2 dy = 0
hence
M = y 2 + 2xy, N = −x2 .
We note that
∂x N − ∂y M 2
Q= =−
M y
is a function of y only, so the integrating factor can be calculated:
2dy 1
IF = exp (∫ Qdy) = exp (∫ − ) = 2.
y y
(y + 2x) dx x2 dy
− 2 =0
y y
Then, we have
y + 2x x2
M= , N = − 2.
y y
Calculating the derivatives, we obtain
∂M 2x ∂N 2x
=− 2, =− 2.
∂y y ∂x y
We see that they are equal, so the equation is exact. Further, we calculate
(y + 2x) dx x2
F = ∫ M dx = ∫ =x+ + ϕ(y),
y y
∂F
= N,
∂y
we obtain
x2 ′ x2
− + ϕ = −
y2 y2
which implies that
ϕ = C.
Therefore the general solution is
x2
x+ + C = 0.
y
Question 5
Part 5.i
What we’re given is a linear first order ODE whose standard form is
5y 7
y′ + = 2x − .
x x
The integrating factor for this equation is
IF = exp (∫ pdx)
5dx
= exp (∫ )
x
= x5 .
After multiplication by the integrating factor we can re-write the equation as
d
(x5 y) = (2x2 − 7) x4 ,
dx
and therefore the solution is
1 1 C
y = 5 ∫ (2x2 − 7) x4 dx = 5 ( 27 x7 − 75 x5 + C) = 27 x2 − 75 + 5 ,
x x x
where C is a constant of integration.
Back to Tutorial 1 Question 5
Part 5.ii
What we’re given is a linear first order ODE whose standard form is
y ′ + 2xy = 2x.
The integrating factor for this equation is
IF = exp (∫ pdx)
= exp (∫ 2xdx)
2
= ex .
After multiplication by the integrating factor we can re-write the equation as
d x2 2
(e y) = 2xex ,
dx
and therefore the solution is
y = e−x ∫ 2xex dx = e−x (ex + C) = 1 + Ce−x ,
2 2 2 2 2
Part 5.iii
IF = exp (∫ pdx)
3dx
= exp (∫ )
x
= x3 .
After multiplication by the integrating factor we can bring the equation to the form
d
(x3 y) = x4 ,
dx
and therefore the solution is
1 1 C
y = 3 ∫ x4 dx = 3 ( 15 x5 + C) = 15 x2 + 3 ,
x x x
where C is a constant of integration.
Back to Tutorial 1 Question 5
Part 5.iv
IF = exp (∫ pdx)
= exp (∫ 3dx)
= e3x .
After multiplication by the integrating factor the equation can be written as
d 3x
(e y) = ex ,
dx
and therefore the solution is
y = e−3x ∫ ex dx = e−3x (ex + C) = e−2x + Ce−3x ,
Part 5.v
IF = exp (∫ pdx)
1 dx
= exp (∫ 2 )
√ x
= x.
After multiplication by the integrating factor the equation can be written as
d √ √
( xy) = x3 ,
dx
and therefore the solution is
√ √ √ √ √
y = 1/ x ∫ x3 dx = 1/ x ( 25 x5 + C) = 25 x2 + C/ x,
Part 5.vi
IF = exp (∫ pdx)
= exp (∫ −2dx)
= e−2x .
After multiplication by the integrating factor we can re-write the equation as
d −2x
(e y) = 0,
dx
and therefore the solution is
y = e2x ∫ 0 = e2x (C) = Ce2x ,
Part 5.vii
Textbook exercise set 2.3. Problem 7 (8th ed); Problem 12 (9th ed)
Taking terms with y on the left we get the equation
y
y ′ − = 2x + 1.
x
The integrating factor for this equation is
IF = exp (∫ pdx)
dx
= exp (∫ − )
x
1
= .
x
After multiplication by the integrating factor we can bring the equation to the form
d 1 2x + 1
( y) = ,
dx x x
and therefore the solution is
(2x + 1) dx
y = x∫ = x (2x + ln ∣x∣ + C) = 2x2 + x ln ∣x∣ + Cx,
x
where C is a constant of integration.
Back to Tutorial 1 Question 5
IF = exp (∫ pdx)
2dx
= exp (∫ )
x
= x2 .
After multiplication by the integrating factor we can cast the equation to the form
d 1
(x2 y) = 2 ,
dx x
and therefore the solution is
1 dx 1 −1 + Cx 1 C
y= 2 ∫ 2
= 2( ) = − 3 + 2,
x x x x x x
where C is a constant of integration.
Back to Tutorial 1 Question 5
202 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
y ′ − y = t + 1.
IF = exp (∫ pdt)
= exp (∫ −dt)
= e−t .
After multiplication by the integrating factor we can re-write the equation as
d −t
(e y) = e−t (t + 1) ,
dt
and therefore the solution is
−t
RRR u = t + 1
RRR u′ = 1 RRR
RRR = −e−t t − e−t + −t
∫ e (t + 1) dt = RRR RRR ∫ e dt
RR dv = e−t dt v = −e−t RR
= −2e−t − e−t t.
Back to Tutorial 1 Question 5
Question 6
Part 6.i
The equation is
z ′ − z = xz 2
This is a Bernoulli equation. Let us introduce the new variable y = z −1 or equivalently
z = y −1 . Then using the chain rule we calculate the derivative
dy 1 dz
=− 2
dx z dx
which implies
dz 1 dy
=− 2 .
dx y dx
Substituting this and z in the equation, we find
−y ′ − y = x.
y ′ + y = −x,
7.1. TUTORIAL 1 203
IF = exp (∫ pdx)
= exp (∫ dx)
= ex .
After multiplication by the integrating factor we can cast the equation to the form
d x
(e y) = −ex x,
dx
and therefore the solution is
−x
RRR u = −x
RRR u′ = −1 RRR
RRR = xe−x − −x
∫ −xe dx = RRR RRR ∫ e dx
RR dv = e−x dx v = −e−x RR
= xe−x + e−x .
Part 6.ii
Remark. The solutions below use slightly different notation for the unknown function and
independent variables. It is assumed everywhere that the unknown z is a function of x:
z(x). This should be taken into account when comparing solutions given here with the
ones in the textbook.
The equation is
z
z′ + = x2 z 2
x
We introduce the new variable y = z −1 or equivalently z = y −1 . Then we calculate the
derivative
dy 1 dz
=− 2
dx z dx
204 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
which implies
dz 1 dy
=− 2 .
dx y dx
Substituting this and z in the equation, we have
y
−y ′ + = x2 .
x
We got a linear first order ODE whose standard form is
y
y′ − = −x2 ,
x
where the coefficient in front of y is
1
p=− .
x
The integrating factor for this equation is
IF = exp (∫ pdx)
dx
= exp (∫ − )
x
1
= .
x
After multiplication by the integrating factor we can re-write the equation as
d y
( ) = −x,
dx x
and therefore the solution is
y = x ∫ −xdx = x (− 21 x2 + C) = − 12 x3 + Cx,
The equation is
2z
z′ − = −x2 z 2
x
Let us introduce the new variable y = z −1 or equivalently z = y −1 , then the derivative is
dy 1 dz
=− 2
dx z dx
or
dz 1 dy
=− 2 .
dx y dx
7.1. TUTORIAL 1 205
d
(x2 y) = x4 ,
dx
and therefore the solution is
1 4 1 1 5 1 3 C
y= ∫ x dx = ( 5 x + C) = 5 x + ,
x2 x2 x2
where C is a constant of integration. Finally, substituting y = z −1 into the obtained solution
and solving for z, we find the solution of the Bernoulli equation
5x2
z(x) = .
x5 + 5C
Back to Tutorial 1 Question 6
The equation is
z
z′ + = −xz 3
x
1
The new variable is y = z −2 hence z = y − 2 . The derivative is given by
dy 2 dz
=− 3
dx z dx
hence
dz dy
= − 2√1y3 .
dx dx
Substituting this and z in the equation, we get
y
− 12 y ′ + = −x.
x
206 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
IF = exp (∫ pdx)
2dx
= exp (∫ − )
x
1
= .
x2
After multiplication by the integrating factor the equation can be written in the form
d y 2
( 2) = ,
dx x x
and therefore the solution is
2dx
y = x2 ∫ = x2 (2 ln ∣x∣ + C) = 2x2 ln ∣x∣ + Cx2 ,
x
where C is a constant of integration. Finally, substituting y = z −2 into the obtained solution
and solving for z, we obtain the solution (implicit) of the Bernoulli equation
1
− x2 (2 ln ∣x∣ + C) = 0.
z2
Back to Tutorial 1 Question 6
The equation is
2z z 2
z′ − =
x x2
We introduce the new variable y = z −1 or equivalently z = y −1 . Then we calculate the
derivative
dy 1 dz
=− 2
dx z dx
hence
dz 1 dy
=− 2 .
dx y dx
Substituting those in the equation, we get the linear equation
2y 1
−y ′ − =
x x2
whose standard form is
2y 1
y′ + = − 2,
x x
7.1. TUTORIAL 1 207
IF = exp (∫ pdx)
2dx
= exp (∫ )
x
= x2 .
d
(x2 y) = −1,
dx
and therefore the solution is
1 1 1 C
y= 2 ∫ −dx = 2 (−x + C) = − + 2 ,
x x x x
where C is a constant of integration. Finally, substituting y = z −1 into the obtained solution
and solving for z, we obtain the solution of the Bernoulli equation
x2
z(x) = .
C −x
Back to Tutorial 1 Question 6
Question 7
Part 7.i
dz z z 2
= + .
dx x x2
We see that it has the form dz/dx = f ( xz ) . Introducing the new variable y = z
x (or z = yx)
we calculate
dz dy
=y+x .
dx dx
Substituting these into the equation we obtain
dy
y+x = y (1 + y)
dx
or
dy y 2
=
dx x
This is a separable equation. Indeed, it can be re-written as
dy dx
=
y2 x
208 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
hence
dy dx
∫ y2 = ∫ x .
Integrating the latter we get
1
− = ln (x) + C,
y
where C is a constant of integration. Further, solving for y we get the explicit solution
1
y(x) = − .
ln (x) + C
Part 7.ii
ln (y + 1) = 2 ln (x) + C,
7.1. TUTORIAL 1 209
where C is a constant of integration. Further, solving for y we obtain the explicit solution
y(x) = eC x2 − 1 = Ax2 − 1,
z(x) = Ax3 − x.
Part 7.iii
dz z z 2
= + cos ( ) .
dx x x
Question 8
Part 8.1
where C is a constant of integration. Solving the initial problem x(0) = 1 and substituting
a = 5 × 10−5 we obtain the solution
1
1 9 − 2000 t
x (t) = 10 + 10 e .
Further, we calculate
1
1 9 − 2000
x (1) = 10 + 10 e = 0.97.
Now, if x(t) = 0.2 then solving
1
1 9 − 2000 t
0.2 = 10 + 10 e
for t we find
t = 73.24.
Back to Tutorial 1 Question 8
in the equations
we obtain
1
p1 = 6000, A = 42000 ln (5) .
Hence the predicted limiting population is p1 = 6000. The population in 2020 is estimated
by substituting t = 30 in the formula
−p0 p1
p (t) =
−p0 − exp (−Ap1 t) p1 + exp (−Ap1 t) p0
i.e.
p (30) = 6000000/ (1000 + 5000 exp (− 30
7 ln (5))) = 5969.8458.
Part 8.2
I ′ + 2I = 10e−2t .
d 2t
(e I) = 10,
dt
and thus the solution is
where C is a constant of integration. Using the condition I(0) = 0 we find C = 0 hence the
solution is
I (t) = 10te−2t .
7.2 Tutorial 2
Question 1
Part 1
y ′′ − 3y ′ + 2y = 0.
The latter implies the quadratic equation
k 2 − 3k + 2 = 0
y = Aex + Be2x .
Part 2
y ′′ + 5y ′ = 0.
Quadratic equation
k (k + 5) = 0
has roots:
k1 = −5, k2 = 0.
The general solution, corresponding to this, reads
y = Ae−5x + B.
Part 3
3y ′′ − 10y ′ + 3y = 0.
Quadratic equation:
k 2 − 10
3 k+1=0
Part 4
y ′′ + 25y = 0.
The quadratic equation
k 2 + 25 = 0
7.2. TUTORIAL 2 213
Part 5
y ′′ − 25y = 0.
The quadratic equation
k 2 − 25 = 0
has the roots
k1 = −5, k2 = 5.
The general solution, corresponding to this, has the form
y = Ae−5x + Be5x .
Part 6
y ′′ − 4y ′ + 8y = 0.
The quadratic equation
k 2 − 4k + 8 = 0
has the roots
k1 = 2 − 2i, k2 = 2 + 2i.
The general solution, corresponding to this, reads
Part 7
y ′′ + 10y ′ + 125y = 0.
The quadratic equation
k 2 + 10k + 125 = 0
has the roots
k1 = −5 − 10i, k2 = −5 + 10i.
The general solution, corresponding to this, is
Part 8
9y ′′ − 6y ′ + 10y = 0.
The quadratic equation
k 2 − 32 k + 10
9 =0
214 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
Part 9
y ′′ = 7y.
The latter implies the quadratic equation
k2 − 7 = 0
whose roots are √ √
k1 = 7, k2 = − 7.
The general solution, corresponding to this, is given by
√ √
y = Ae 7x
+ Be− 7x
.
Back to Tutorial 2 Question 1
Part 10
y ′′ + 6y ′ + 9y = 0.
The quadratic equation
k 2 + 6k + 9 = 0
has only one root k = −3. The general solution, corresponding to this, reads
y = Ae−3x + Bxe−3x .
Note that the second term in the solution is the first one multiplied by x. Here A and B are
arbitrary constants.
Back to Tutorial 2 Question 1
y ′′ − 5y ′ + 6y = 0.
The quadratic equation
k 2 − 5k + 6 = 0
has the roots
k1 = 2, k2 = 3.
The general solution, corresponding to this, reads
y = Ae2x + Be3x .
Back to Tutorial 2 Question 1
7.2. TUTORIAL 2 215
4y ′′ − 4y ′ + y = 0.
The quadratic equation
k 2 − k + 41 = 0
has only one root k = 1/2. The general solution, corresponding to this, has the form
1 1
y = Ae 2 x + Bxe 2 x .
Note that the second term in the solution is the first one multiplied by x.
Back to Tutorial 2 Question 1
y ′′ + 2y ′ − 8y = 0.
The latter implies the quadratic equation
k 2 + 2k − 8 = 0
y = Ae−4x + Be2x .
Here A and B are arbitrary constants.Substituting the initial values y(0) = 3, y ′ (0) = −12
into the obtained solution we obtain the algebraic system
y ′′ − 2y ′ − 2y = 0.
The quadratic equation
k 2 − 2k − 2 = 0
has the roots √ √
k1 = 1 − 3, k2 = 1 + 3.
The general solution, corresponding to this, reads
√ √
y = A exp ((1 − 3) x) + B exp ((1 + 3) x) .
216 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
Here A and B are arbitrary constants. Substituting the initial values y(0) = 0, y ′ (0) = 3
into the obtained solution we get the algebraic system
√ √
0 = A + B, 3 = A + A 3 − B 3 + B.
y ′′ + 2y ′ + y = 0.
The quadratic equation
k 2 + 2k + 1 = 0
has only one root k = −1. The general solution, corresponding to this, is
y = Ae−x + Bxe−x .
Substituting the initial values y(0) = 1, y ′ (0) = −3 into the obtained solution we have the
algebraic system
−3 = −A + B, 1 = A.
Solving this system, we get
A = 1, B = −2.
So, the solution is
y (x) = e−x − 2xe−x .
Back to Tutorial 2 Question 1
Question 2
Part 1
y ′′ + 2y ′ + 2y = 3 cos (2x) .
We have an inhomogeneous second order equation with constant coefficients. We will be
looking for a particular solution in the form
(−4K − 2L) sin (2x) + (4L − 2K) cos (2x) = 3 cos (2x) .
−4K − 2L = 0, −2K + 4L − 3 = 0
which gives
3
K = − 10 , L = 35 .
Therefore a particular solution y = yp is given by
y = 35 sin (2x) − 10
3
cos (2x) .
Part 2
y ′′ + 9y = 5 cos (3x) .
We are looking for a particular solution in the form
−6K = 0, 6L − 5 = 0
which means
K = 0, L = 65 .
Therefore a particular solution y = yp has the form
y = 56 sin (3x) x.
Part 3
y ′′ + y ′ − 2y = 2x.
First consider the homogeneous equation:
y ′′ + y ′ − 2y = 0.
k2 + k − 2 = 0
218 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
Part 4
y ′′ + 2y ′ − y = 10.
We’re looking for a particular solution in the form
y p = c0 ,
where c0 is a constant. We have:
yp′ = 0, yp′′ = 0
hence
−c0 = 10
or
c0 = −10.
Back to Tutorial 2 Question 2
7.2. TUTORIAL 2 219
y ′′ − y ′ + 9y = 3 sin (3x) .
We are looking for a particular solution in the form
yp = K sin (3x) + L cos (3x) ,
where K and L are some constants to be found. The derivatives are
yp′ = 3K cos (3x) − 3L sin (3x) ,
yp′′ = −9K sin (3x) − 9L cos (3x) .
Substituting these expressions in the equation we calculate
−3K cos (3x) + 3L sin (3x) = 3 sin (3x) .
Equating the coefficients at sin and cos, we obtain
3L − 3 = 0, −3K = 0
which gives
K = 0, L = 1.
Therefore a particular solution y = yp has the form
y = cos (3x) .
Back to Tutorial 2 Question 2
y ′′ − 5y ′ + 6y = xex .
We are looking for a particular solution in the form
yp = (c0 + c1 x)ex ,
where ci are some constants. The derivatives of yp are
yp′ = (c0 + c1 x + c1 ) ex
yp′′ = (c0 + c1 x + 2c1 ) ex .
Substituting these expressions in the original equation we calculate
ex (2c0 + 2c1 x − 3c1 ) = xex .
Factoring out the exponential function, we obtain
2c0 − 3c1 = 0, 2c1 − 1 = 0.
Solving this we find
c0 = 43 , c1 = 21 .
Therefore a partial solution reads
yp = ( 34 + 12 x)ex .
Back to Tutorial 2 Question 2
220 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
y ′′ − 2y ′ + y = 8ex .
We’re looking for a particular solution in the form
yp = x2 ex c0 ,
yp′ = (2xc0 + x2 c0 ) ex
yp′′ = (2c0 + 4xc0 + x2 c0 ) ex .
2ex c0 = 8ex .
2c0 − 8 = 0
hence
c0 = 4.
Thus a partial solution has the form
yp = 4x2 ex .
Question 3
Part 1
y ′′ + y = tan (x) .
First we restrict our attention to the homogeneous equation:
y ′′ + y = 0.
k2 + 1 = 0
yh = A cos x + B sin x.
Here A and B are arbitrary constants. Now we’re looking for a particular solution of the
form
y = M (x) cos x + N (x) sin x.
7.2. TUTORIAL 2 221
M ′ cos x + N ′ sin x = 0.
sin2 x
M′ = − , N ′ = sin x.
cos x
Integrating this we get
sin2 xdx
M =∫ − = sin x − ln (sec (x) + tan (x)) ,
cos x
N = ∫ sin xdx = − cos x.
Part 2
y ′′ − y = ex .
First solve the homogeneous equation:
y ′′ − y = 0.
k2 − 1 = 0
yh = Ae−x + Bex
y ′ = M ′ e−x − M e−x + N ′ ex + N ex
y ′′ = −M ′ e−x + M e−x + N ′ ex + N ex .
−M ′ e−x + N ′ ex = ex .
M ′ e−x + N ′ ex = 0.
M ′ = − 21 e2x , N ′ = 21 .
M = ∫ − 12 e2x dx = − 14 e2x ,
1
N =∫ 2 dx = 21 x.
or
y = Ae−x + Cex + 12 xex
Back to Tutorial 2 Question 3
Part 3
y ′′ + 3y ′ + 2y = 3e−2x + x.
First solve the homogeneous equation:
y ′′ + 3y ′ + 2y = 0.
k 2 + 3k + 2 = 0
yh = Ae−2x + Be−x ,
M ′ e−2x + N ′ e−x = 0.
or
y = Ce−2x + Be−x − 3xe−2x − 43 + 12 x
Back to Tutorial 2 Question 3
Part 4
y ′′ − y = x2 ex .
First we restrict our attention to the homogeneous equation:
y ′′ − y = 0.
yh = Ae−x + Bex .
y ′ = M ′ e−x − M e−x + N ′ ex + N ex
y ′′ = −M ′ e−x + M e−x + N ′ ex + N ex .
−M ′ e−x + N ′ ex = x2 ex .
M ′ e−x + N ′ ex = 0.
M ′ = − 21 x2 e2x , N ′ = 21 x2 .
or
y = Ae−x + Cex + 14 xex − 14 x2 ex + 61 x3 ex
Back to Tutorial 2 Question 3
Part 5
7.2. TUTORIAL 2 225
y ′′ + 2y ′ + y = e−x .
First solve the homogeneous equation:
y ′′ + 2y ′ + y = 0.
M ′ e−x + N ′ xe−x = 0.
M ′ = −x, N ′ = 1.
N = ∫ dx = x.
Therefore the general solution reads
y ′′ − y = 4 + 2x.
First we solve the homogeneous equation:
y ′′ − y = 0
which implies the quadratic equation
k 2 − 1 = 0.
The roots are
k1 = −1, k2 = 1.
The general solution, corresponding to this, has the form
yh = Ae−x + Bex .
Here A and B are arbitrary constants.
Now we’re looking for a particular solution of the form
y = M (x) e−x + N (x) ex .
Firsly, we calculate the derivative:
y ′ = M ′ e−x − M e−x + N ′ ex + N ex
and assume that
M ′ e−x + N ′ ex = 0.
Next, we calculate the second derivative:
y ′′ = −M ′ e−x + M e−x + N ′ ex + N ex .
Substituting the derivatives in the equation, we calculate :
−M ′ e−x + N ′ ex = 4 + 2x.
This needs to be solved along with
M ′ e−x + N ′ ex = 0.
From the last two equations we obtain
M ′ = −ex (2 + x) , N ′ = e−x (2 + x) .
Integrating this we get
Remark. In the solutions below we use the geometric series expansion which is
1
= 1 + x + x2 + . . . .
1−x
7.2. TUTORIAL 2 227
Question 4
Part 1
We re-write the equation in the form
(D2 − 1)y = ex ,
and then as
1
y= ex .
(D − 1) (D + 1)
We denote φ(D) = (D − 1) (D + 1) and check that φ(1) = 0. Using the shifting theorem
(we have to make the substitution D → 1+D when moving the exponential function through
the D−operator), we obtain
1
y = ex 2 1,
D + 2D
which can then be re-written as
1 1 1
y = ex 1 = ex x.
D+2 D D+2
We can re-write the latter as
1
y = 12 ex x.
1 + 12 D
Using the geometric series expansion, we obtain
y = 12 ex (1 − 12 D + . . . )x
= 12 ex (x − 12 ) .
Part 2
We solve the problem in two steps: solve separately for equations
y ′′ + 3y ′ + 2y = 3e−2x , y ′′ + 3y ′ + 2y = x
and then combine the answers. We re-write the equation in the form
and then as
1
y= 3e−2x .
(D + 2) (D + 1)
We denote φ(D) = (D + 2) (D + 1) and check that φ(−2) = 0. Using the shifting theorem
(we have to make the substitution D → −2 + D when moving the exponential function
through the D−operator), we obtain
1
y = 3e−2x 1,
D2 − D
228 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
(D2 + 3D + 2)y = x,
and then as
1
y= x.
(D + 2) (D + 1)
Then we find
1
1 2
y= x = x.
D2 + 3D + 2 1 + 12 D2 + 23 D
Using the geometric series expansion, we get
yp = 21 (1 − 21 D2 − 32 D + . . . )x
= 21 x − 43 .
So, combining the solutions we get
Part 3
We re-write the equation in the form
(D2 − 1)y = x2 ex ,
and then as
1
y= x2 ex .
(D − 1) (D + 1)
We denote φ(D) = (D − 1) (D + 1) and check that φ(1) = 0. Using the shifting theorem
(we have to make the substitution D → 1+D when moving the exponential function through
the D−operator), we obtain
1
y = ex 2 x2 ,
D + 2D
which can then be re-written as
1 1 2 1 1 3
y = ex x = ex x.
D+2 D D+23
We can re-write the latter as
1
y = 61 ex x3 .
1 + 21 D
7.2. TUTORIAL 2 229
y = 16 ex (1 − 12 D + 41 D2 − 18 D3 + . . . )x3
= 61 ex (x3 − 23 x2 + 32 x − 43 ) .
Note that the last term can be absorbed into the general solution thus a particular solution
is
yp = 16 ex (x3 − 32 x2 + 32 x) .
Back to Tutorial 2 Question 4
Part 4
We re-write the equation in the form
and then as
1
y= e−x cos x .
D2 + 2D + 2
This is equivalent to
1
y = Re ( exp ((−1 + i) x))
D2
+ 2D + 2
We denote φ(D) = D2 + 2D + 2 and check that φ(−1 + i) = 0. Using the shifting theorem
(we have to make the substitution D → −1 + i + D when moving the exponential function
through the D−operator), we obtain
1
y = Re (exp ((−1 + i) x) 1) ,
D2 + 2Di
which can then be re-written as
1 1 1 1
y = Re (exp ((−1 + i) x) 1) = Re (exp ((−1 + i) x) )
D D + 2i D 2i
= Re (− 21 i exp ((−1 + i) x) x) .
yp = 12 xe−x sin x.
Question 5
Part 1
x′′ + 20x = 0.
The quadratic equation
k 2 + 20 = 0
has the roots √ √
k1 = −2i 5, k2 = 2i 5.
230 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
Substituting the initial conditions t = 0, x (0) = 41 and x′ (0) = 0 into the solution and its
derivative we find √
0 = 2B 5, 14 = A
and thus
B = 0, A = 41 .
Hence the solution reads √
x (t) = 14 cos (2 5t) .
Back to Tutorial 2 Question 5
Part 2
k 2 + 8k + 25 = 0.
Here A and B are arbitrary constants. Now we’re looking for a particular solution in the
form
Qp = c0 ,
where c0 is a constant. We have:
Q′p = 0, Q′′p = 0.
25c0 = 25.
Thus
c0 = 1.
7.2. TUTORIAL 2 231
Q′ = A (−4 cos (3t) − 3 sin (3t)) e−4t + B (−4 sin (3t) + 3 cos (3t)) e−4t .
Substituting the initial conditions t = 0, Q (0) = 0 and Q′ (0) = 0 into the solution and its
derivative we get
0 = 1 + A, 0 = 3B − 4A.
Solving for A and B we get
A = −1, B = − 34 .
Hence the solution reads
Part 3
k 2 + 25k + 25 = 0.
Q p = c0 ,
Substituting the initial conditions t = 0, Q (0) = 4 and Q′ (0) = 0 into the solution and its
derivative we get
√ √
0 = (− 25 5 25 5
2 + 2 21) B + (− 2 − 2 21) A, 4 = B + A + 5 .
1
7.3 Tutorial 3
Question 1
Part 1
1 1
n2 + 1 n + n3 0+0
= lim 2 1 = = 0 ( convergent )
h3 + 2n + 1 n→∞ 1 + n2 + n3 1 + 0 + 0
Part 2
n4 + 1 n + n13
lim = lim
n→∞ n3 + 2n + 1 n→∞ 1 + 22 + 13
n n
= lim n = ∞ ( divergent )
n→∞
Part 3
n3 ∞
lim n
= ( , l’Hopital )
n→∞ 2 ∞
d(n3 )
dn 3n2
= lim d2 n = lim
n→∞ n→∞ 2n ln 2
dn
d(3n2 )
dn 6n
= lim = lim n
n→∞ d(2n ln 2) n→∞ 2 (ln 2)2
dn
d(6n)
dn 6
= lim = lim n
n→∞ d(2n (ln 2)2 ) n→∞ 2 (ln 2)3
dn
= 0, as lim 2n = ∞
n→∞
(convergent)
Back to Tutorial 3 Question 1
7.3. TUTORIAL 3 233
Part 4 The rigorous proof is quite complicated, but obviously rn = rn−1 r < r < 1 (r < 1).
So the sequence rn stays less than 1 and decreasing. We accept without proof that it
decreasing all the way to zero.
1
If r > 1 then r < 1 hence
1
lim = 0 → lim rn = ∞
n→∞ rn n→∞
Part 6
1 + (−1)n 0, n − odd
an = ={ 2
n n , n − even
Since
2
lim =0
n→∞ n
then
1 + (−1)n
lim an = lim = 0 (convergent)
n→∞ n→∞ n
Part 7
⎧
⎪ 1, n - even
⎪
an = ⎨ 1
n−1
2
⎪
⎪
⎩ 1 − ( 2
) n - odd
n−1
Since → ∞ as n → ∞, then
2
n−1
1 2
lim ( ) = 0.
n→∞ 2
Hence
lim an = 1 (convergent)
n→∞
Part 8
1 1 7
n2 − n + 7 n − n2 + n3
lim = lim
n→∞ 2n3 + n2 n→∞ 2 + n1
0−0+0
= = 0 (convergent )
2+0
Part 9
sin2 n
an = √
n
Note that 0 < sin2 n < 1, thus 0 < an < √1n .
√
Since limn→∞ n = ∞, then limn→∞ √1n = 0 hence
sin2 n
lim √ = 0.
n→∞ n
Part 10 √
n ∞
lim = ( , l’Hopital )
n→∞ ln n ∞
√ 1
√
d n 2 n
= lim d ln n = lim 1
n→∞ n→x
n
√dn
n
= lim =∞ (divergent)
n→∞ 2
Part 11
2n + 1 ∞
lim = ( , l’Hopital )
n→∞ en ∞
d(2n +1)
dn 2n ln 2
= lim de n = lim
n→∞
dn
n→∞ en
2 n
= ln 2 lim ( ) = 0
n→∞ e
2
Since e < 1 then
2 n
lim ( ) = 0 (converges)
n→∞ e
Question 2
Part 1
∞
1 1 1 1 n−1
1+ + + ... + n + ... = ∑ ( )
3 9 3 n=1 3
1 3 1
= 1 = (Geometric sevies with q = )
1− 3 2 3
Part 2
1 + 3 + 5 + 7 + . . . + (2n − 1) + . . .
Series diverges because
lim an = lim (2n − 1) ≠ 0.
n→∞ n→∞
Part 3
1 − 2 + 4 − 8 + 16 − . . . + (−2)n + . . .
Series diverges because
Part 4
∞
4 4 4 1 n−1 4
4+ + + ... + n + ... = ∑ 4( ) = =6
3 9 3 n=1 3 1 − 31
Geometric series with a0 = 4, q = 13 .
Question 3
Part 1 We have
n 1 2
∫ n2 + 1 dn = 2 ln (n + 1)
hence ∞ n
∫1 dn = ∞ − 12 ln (2) = ∞.
n2
+1
The series diverges.
Part 2
1 √
∫ √ dn = 2 n + 1
n+1
∞ 1 √
∫1 √ dn = ∞ − 2 2 = ∞.
n+1
The series diverges.
Part 3
1
∫ n2 + 1 dn = arc tan (n)
∞ 1 1
∫1 n2 + 1 dn = 4 π.
The series converges.
Part 4
ln (n) ln (n) 1
∫ 2
dn = − −
n n n
∞ ln (n)
∫1 dn = 1.
n2
The series converges. Here is the detailed integration (by parts):
RRR u = ln (n) u′ = 1 RRR
ln (n) n RRR = − 1 ln (n) + 1
∫ dn = RRRRR RRR ∫ dn
n2 RRR dv = n12 dn v = − n1 RR n n2
1 1
= − ln (n) − .
n n
Question 4
Part 1
1 1
< 2, n ⩾ 1
n2 +n+1 n
The series ∞
1
∑ 2
n=1 n
is known to converge. So
∞
1
∑
n=1 n2 +n+1
also converges.
236 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
Part 2
1 1 1
√ ⩾ = , n⩾1
n + n n + n 2n
The series
∞
1 1 ∞ 1
∑ = ∑
n=1 2n 2 n=1 n
diverges. Thus the series
∞
1
∑ √
n=1 n + n
is divergent.
Part 3
1 1
< , n⩾1
1 + 3n 3n
The series
∞
1
∑ n
n=1 3
1
converges as a geometric series with q = 3 . Hence
∞
1
∑ n
n=1 1 + 3
is convergent.
Part 4 √ √
n n 1
2
⩽ 2 = 3, n⩾1
n +n n n2
The series
∞
1
∑ 3
n=1 n2
3
converges as an α -series with α = 2 > 1 (see Example 1). Hence
∞
√
n
∑ 2
n=1 n + n
converges.
Part 5
sin2 n 1 1
2
⩽ 2 ⩽ 2, n ⩾ 1
n +1 n +1 n
The series
∞
1
∑ 2
n=1 n
converges. Thus
∞
sin2 n
∑ 2
n=1 n + 1
converges as well.
Question 5
Part 1 We have
10n
an = .
nn
Consider the expression
an+1 10n+1 nn 10 n n 10 1
= n+1 = ( ) = n.
an (n + 1) 10 n n+1 n+1 n + 1 (1 + n1 )
Part 2 We have
n n
an = ( ) .
ln n
n
Note that > 1, so the series diverges. Let us use the ratio test to show the same.
ln n
Consider the expression
n+1 n
an+1 n+1 ln n n n+1 n n+1 ln n
=( ) ( ) =( ) ( ) .
an ln(n + 1) n n ln(n + 1) ln(n + 1)
Since
n+1 n n+1
lim ( ) = e, lim = ∞ (L’Hospital’s rule)
n→∞ n n→∞ ln(n + 1)
n
ln n
lim ( ) =1
n→∞ ln(n + 1)
then
an+1
lim =∞
n→∞ an
Part 3 We have
3n
an = .
n!n
Consider the expression
Since
3n
lim 2 =0<1
n→∞ (n + 1)
the series converges.
Back to Tutorial 3 Question 5
238 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
Part 4 We have
n!n2
an = .
(2n)!
Consider the expression
2 2
an+1 (n + 1)! (n + 1) (2n)! 1 (n + 1)
= =2 .
an (2n + 2)!n!n2 (2n + 1) n2
Since
2
1 (n + 1)
lim =0
n→∞ 2 (2n + 1) n2
Part 5 We have
n
an = n ( 43 ) .
Consider the expression
3 (n+1)
an+1 (n + 1) ( 4 ) 3n+1
= n = 4 .
an n ( 34 ) n
Since
3 n+1 3
lim =
n→∞ 4 n 4
Question 6
Part 1 We have
1
an = .
n2
Since
1
lim =0
n→∞ n2
and
1 1
an+1 = 2 < = an , for any n > 1
(n + 1) n2
we conclude that the series converges.
Part 2
n
an = .
3n + 2
We have
n 1
lim = 3 ≠0
n→∞ 3n + 2
hence the series diverges.
Back to Tutorial 3 Question 6
7.3. TUTORIAL 3 239
Part 3
n
an = √ .
n3 + 2
Since
n
lim √ =0
n→∞ n3 + 2
and
n+1 n
an+1 = √ <√ = an , for any n > 1
3
(n + 1) + 2 n3 + 2
Part 4
n
an = .
ln (n)
Since
n
lim =∞≠0
n→∞ ln (n)
Part 5
1
an = 1 .
2n
Since
1
lim 1 =1≠0
n→∞ 2n
we conclude that the series diverges.
Back to Tutorial 3 Question 6
Question 7
∞ ∞
(−1)n+1 1
a. Consider ∑ n
. The respective series with positive terms is ∑ n
which is
n=1 3 n=1 3
∞
(−1)n+1
a geometric series with q = 31 < 1 hence it converges. Thus ∑ converges
n=1 3n
absolutely.
∞
(1)n+1
Further, cousider ∑ . The respective series with positive terms is the harmonic
n=1 n
∞
1
series ∑ which is divergent.
n=1 n
On the other hand
1 1
a) > and
n n+1
1
b) lim = 0
n→∞ n
∞
(−1)n+1
thus ∑ converges conditionally.
n=1 3n
Back to Tutorial 3 Question 7
240 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
with
nn
an = n.
(n + 1)
Since
nn −1
lim n =e ≠0
n→∞ (n + 1)
where
√ √ 1
an = n+1− n= √ √ .
n+1+ n
Since
1
lim an = lim √ √ =0
n→∞ n→∞ n+1+ n
and
1 1
an+1 = √ √ <√ √ = an , for any n > 1
n+2+ n+1 n+1+ n
we see that series converges. Now consider the series with positive terms
∞
1 1 ∞ 1
∑√ √ > ∑√ .
n=1 n + 1 + n 2 n=1 n + 1
The latter series is known to be divergent hence the former series diverges as well.
Thus the series in question converges conditionally.
Back to Tutorial 3 Question 7
7.3. TUTORIAL 3 241
Question 8
1. ∞
∑ nxn
n=1
We denote
an = nxn
and consider the expression
an+1 (n + 1) xn+1 (n + 1) x
∣ ∣=∣ n
∣=∣ ∣.
an nx n
Then we calculate
(n + 1) x
lim ∣ ∣ = ∣x∣ .
n→∞ n
According to the ratio test the series converges if
x ∈ (−1, 1).
with
an = n.
Since
lim n = ∞ ≠ 0
n→∞
Since
lim n = ∞ ≠ 0
n→∞
2. ∞
nxn
∑ n
n=1 2
We denote
nxn
an =
2n
and consider the expression
an+1 (n + 1) xn+1 2n (n + 1) x
∣ ∣=∣ n+1 n
∣ = ∣ 21 ∣.
an 2 nx n
Then we calculate
(n + 1) x
lim ∣ 21 ∣ = ∣ 21 x∣ .
n→∞ n
According to the ratio test the series converges if
∣ 12 x∣ < 1, i.e.
−1 < 21 x < 1
−2 < x < 2
x ∈ (−2, 2).
with
an = n.
Since
lim n = ∞ ≠ 0
n→∞
Since
lim n = ∞ ≠ 0
n→∞
3. ∞
∑ n!xn
n=1
We denote
an = n!xn
and consider the expression
an+1 (n + 1)!xn+1
∣ ∣=∣ ∣ = ∣(n + 1) x∣ .
an n!xn
Then we calculate
lim ∣(n + 1) x∣ = ∞ for x ≠ 0.
n→∞
4.
∞ ∞
(−1)n nxn n∣x∣n
∑∣ 3 ∣ = ∑ 3
n=1 2n (n + 1) n
n=1 2 (n + 1)
We denote
n∣x∣n
an = 3
2n (n + 1)
and consider the expression
4 4
an+1 (n + 1) xn+1 2n (n + 1) x
∣ ∣=∣ 3 ∣ = ∣ 12 3 ∣.
an n+1
2 (n + 2) nx n (n + 2) n
Then we calculate
4
(n + 1) x
lim ∣ 1 ∣ = ∣ 12 x∣ .
n→∞ 2 (n + 2) n
3
∣ 12 x∣ < 1
−1 < 12 x < 1
−2 < x < 2
x ∈ (−2, 2).
x ∈ ( 25 , 45 ).
We have
an = 1.
Since
lim 1 = 1 ≠ 0
n→∞
Since
lim 1 = 1 ≠ 0
n→∞
6.
∞ n
2n (x − 3)
∑
n=1 n2
We denote n
2n (x − 3)
an =
n2
and consider the expression
n+1
an+1 2n+1 (x − 3) n2 2 (x − 3) n2
∣ ∣=∣ 2 n ∣ = ∣ 2 ∣.
an (n + 1) 2n (x − 3) (n + 1)
Then we calculate
2 (x − 3) n2
lim ∣ 2 ∣ = ∣2x − 6∣ .
n→∞ (n + 1)
According to the ratio test the series converges if
∣2x − 6∣ < 1
−1 < 2x − 6 < 1
5 < 2x < 7
5 7
2 <x< 2
x ∈ ( 52 , 72 ).
We have
1
an = .
n2
Since
1
lim =0
n→∞ n2
and
1 1
an+1 = 2 < = an , for any n > 1
(n + 1) n2
we conclude that the series converges at x = 52 .
7
Secondly we set x = 2 to obtain the series with positive terms
∞
1
∑ 2
.
n=1 n
This series is know to converge (α-series with α > 1). We also use the integral test:
∞ 1
∫1 dn = 1.
n2
Since the integral is finite, then according to the integral test the series converges at
x = 72 . The interval of convergence is thus [ 52 , 72 ].
Back to Tutorial 3 Question 8
We denote n
(x − 1)
an = n
2 (n + 1)
and consider the expression
n+1
an+1 (x − 1) 2n (n + 1) 1 (x − 1) (n + 1)
∣ ∣ = ∣ n+1 n ∣ = ∣2 ∣.
an 2 (n + 2) (x − 1) n+2
Then we calculate
(x − 1) (n + 1)
lim ∣ 1 ∣ = ∣ 21 x − 12 ∣ .
n→∞ 2 n+2
According to the ratio test the series converges if
∣ 21 x − 12 ∣ < 1
−1 < 21 x − 12 < 1
− 12 < 21 x < 3
2
−1 < x < 3
7.3. TUTORIAL 3 247
x ∈ (−1, 3).
We have
1
an = .
n+1
Since
1
lim =0
n→∞ n + 1
and
1 1
an+1 = < = an , for any n > 1
n+2 n+1
we conclude that the series converges.
Secondly we set x = 3 to obtain the series
∞
1
∑
n=0 n + 1
Then we calculate
2
(n + 1) (x + 2)
lim ∣ 21 ∣ = ∣ 21 x + 1∣ .
n→∞ n2
248 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
∣ 12 x + 1∣ < 1
−1 < 21 x + 1 < 1
−2 < 12 x < 0
−4 < x < 0
x ∈ (−4, 0).
We have
an = n 2 .
Since
lim n2 = ∞ ≠ 0
n→∞
Since
lim n2 = ∞ ≠ 0
n→∞
Then we calculate
(x − 2) n3
lim ∣ 3 ∣ = ∣x − 2∣ .
n→∞ (n + 1)
According to the ratio test the series converges if
∣x − 2∣ < 1
−1 < x − 2 < 1
1<x<3
x ∈ (1, 3).
We have
3
an = .
n3
Since
3
lim =0
n→∞ n3
and
3 3
an+1 = 3 < = an , for any n > 1
(n + 1) n3
we conclude that the series converges.
Secondly we set x = 3 to obtain the series
∞
3
∑ 3
n=1 n
7.4 Tutorial 4
Question 1
y ′ + (x + 2) y = 0
We are looking for a solution in the form
∞
y = ∑ an x n .
n=0
or ∞ ∞ ∞
∑ nan xn−1 + 2 ∑ an xn + ∑ an xn+1 = 0.
n=1 n=0 n=0
a1 + 2a0 = 0
which implies
a1 = −2a0 .
Equating the coefficients at different powers of x we find the recurrent formula
or
2an + an−1
an+1 = − .
n+1
Setting n = 1, 2, . . . in the recurrent formula we calculate
a1 = −2a0 , a2 = 23 a0 , a3 = − 13 a0 , a4 = − 24
5
a0 .
y = a0 − 2a0 x + 23 a0 x2 − 13 a0 x3 − 24
5
a0 x 4 + . . .
y ′′ − x2 y = 0
Hence ∞ ∞
y ′ = ∑ nan xn−1 , y ′′ = ∑ (n − 1) nan xn−2 .
n=1 n=2
or ∞ ∞
∑ (n − 1) nan xn−2 − ∑ an xn+2 = 0.
n=2 n=0
Setting n = 0, 1 we obtain
n=0∶ 2a2 = 0
n=1∶ 6a3 = 0
which imply
a2 = 0, a3 = 0.
Equating the coefficients at different powers of x we get the recurrent formula
(n + 1) (n + 2) an+2 − an−2 = 0
or
an−2
an+2 = .
(n + 1) (n + 2)
Setting n = 2, 3, . . . in the recurrent formula we find
1 1
a2 = 0, a3 = 0, a4 = 12 a0 , a5 = 20 a1 .
y ′′ + (x − 1) y ′ + y = 0
252 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
∞
y = ∑ an x n .
n=0
∞ ∞ ∞
∑ (n − 1) nan xn−2 + (x − 1) ∑ nan xn−1 + ∑ an xn = 0
n=2 n=1 n=0
or
∞ ∞ ∞ ∞
∑ (n − 1) nan xn−2 − ∑ nan xn−1 + ∑ an nxn + ∑ an xn = 0.
n=2 n=1 n=1 n=0
∞ ∞ ∞ ∞
∑ (n + 1) (n + 2) an+2 xn − ∑ (n + 1) an+1 xn + ∑ an nxn + ∑ an xn = 0.
n=0 n=0 n=1 n=0
2a2 − a1 + a0 = 0 .
a2 = 21 a1 − 12 a0 .
or
an+1 − an
an+2 = .
n+2
Setting n = 0, 1, . . . in the recurrent formula we calculate
a2 = 21 a1 − 12 a0 , a3 = − 61 a1 − 16 a0 , a4 = − 61 a1 + 12
1
a0 , a 5 = 1 1 1
20 a0 , 2 a1 − 2 a0 = 12 a1 − 12 a0 .
y = a0 + a1 x + ( 21 a1 − 21 a0 ) x2 + (− 16 a1 − 16 a0 ) x3 + (− 16 a1 + 12
1 1
a0 ) x4 + 20 a0 x 5 + . . .
Question 2
Part 1 a)
y′ − y + x = 0
We are looking for a solution in the form
∞
y = ∑ an x n .
n=0
or ∞ ∞
∑ nan xn−1 − ∑ an xn = −x.
n=1 n=0
2a2 − a1 = −1, n = 1
(n + 1) an+1 − an = 0, n ≠ 1
or
an
an+1 = , n ≠ 1.
n+1
Setting n = 0, 1, . . . in the recurrent formula we find
or ∞ ∞
∑ (n − 1) nan xn−2 + ∑ an xn = 0.
n=2 n=0
(n + 1) (n + 2) an+2 + an = 0
or
an
an+2 = − .
(n + 1) (n + 2)
Setting n = 0, 1, . . . in the recurrent formula we get
a2 = − 21 a0 , a3 = − 61 a1 , a4 = 1
24 a0 , a5 = 1
120 a1 .
or ∞ ∞ ∞
∑ (n − 1) nan xn−2 − ∑ nan xn−1 − 2 ∑ an xn = 0.
n=2 n=1 n=0
or
an+1 n + an+1 + 2an
an+2 = .
(n + 1) (n + 2)
Setting n = 0, 1, . . . in the recurrent formula we get
a2 = 21 a1 + a0 , a3 = 12 a1 + 13 a0 , a4 = 5
24 a1 + 14 a0 , a5 = 11
120 a1
1
+ 12 a0 .
Part 2
(1 − x2 ) y ′′ − 2xy ′ + 6y = 0
We are looking for a solution in the form
∞
y = ∑ an x n .
n=0
or
∞ ∞ ∞ ∞
∑ (n − 1) nan xn−2 − ∑ an n (n − 1) xn − 2 ∑ an nxn + 6 ∑ an xn = 0.
n=2 n=2 n=1 n=0
a2 = −3a0 , a3 = − 32 a1 .
or
an (n2 + n − 6)
an+2 = .
(n + 1) (n + 2)
Setting n = 0, 1, . . . in the recurrent formula we calculate
a2 = −3a0 , a3 = − 32 a1 , a4 = 0, a5 = − 15 a1 , a6 = 0, . . .
The solution takes the form
y = a0 + a1 x − 3a0 x2 − 32 a1 x3 − 15 a1 x5 + . . .
Back to Tutorial 4 Question 2
Part 3
(2 + x2 ) y ′′ − xy ′ + 4y = 0
We are looking for a solution in the form
∞
y = ∑ an x n .
n=0
y ′ − 2xy = 0
or ∞ ∞
∑ nan xn−1 − 2 ∑ an xn+1 = 0.
n=1 n=0
On adjusting the indices we obtain
∞ ∞
∑ (n + 1) an+1 xn − 2 ∑ an−1 xn = 0.
n=0 n=1
n=0∶ a1 = 0 .
(n + 1) an+1 − 2an−1 = 0
or
2an−1
an+1 = .
n+1
Setting n = 1, 2, . . . in the recurrent formula we obtain
a1 = 0, a2 = a0 , a3 = 0, a4 = 21 a0 .
y = a0 + a0 x2 + 21 a0 x4 + . . .
y ′′ − x2 y ′ − xy = 0
or ∞ ∞ ∞
∑ (n − 1) nan xn−2 − ∑ an nxn+1 − ∑ an xn+1 = 0.
n=2 n=1 n=0
n=0∶ 2a2 = 0
.
n=1∶ 6a3 x − a0 x = 0
a2 = 0, a3 = 61 a0 .
or
an−1 n
an+2 = .
(n + 1) (n + 2)
Setting n = 1, 2, . . . in the recurrent formula we get
a2 = 0, a3 = 61 a0 , a4 = 16 a1 , a5 = 0, 16 a0 = 16 a0 .
y = a0 + a1 x + 16 a0 x3 + 16 a1 x4 + . . .
y ′′ + 3xy ′ − y = 0
or
∞ ∞ ∞
∑ (n − 1) nan xn−2 + 3 ∑ an nxn − ∑ an xn = 0.
n=2 n=1 n=0
n=0∶ 2a2 − a0 = 0 .
a2 = 12 a0 .
(n + 1) (n + 2) an+2 + 3an n − an = 0
or
an (3n − 1)
an+2 = − .
(n + 1) (n + 2)
Setting n = 0, 1, . . . in the recurrent formula we find
a2 = 21 a0 , a3 = − 13 a1 , a4 = − 24
5
a0 , a 5 = 2 1
15 a1 , 2 a0 = 12 a0 .
y = a0 + a1 x + 21 a0 x2 − 13 a1 x3 − 24
5 2
a0 x4 + 15 a1 x 5 + . . .
Question 3
a)
4xy ′′ + 2y ′ + y = 0
We are looking for a solution in the form
∞
y = ∑ an xn+c .
n=0
or
∞ ∞ ∞
n+c−1 n+c−1
4 ∑ an (n + c) (n + c − 1) x + 2 ∑ (n + c) an x + ∑ an xn+c = 0.
n=0 n=0 n=0
4c (c − 1) + 2c = 0.
The coefficients of the series are then obtained by setting n = 0, 1, . . . in the previous
recurrent formula we obtain
a1 = − 21 a0 , a2 = 1
24 a0 ,
1
a3 = − 720 a0 , a 4 = 1
40320 a0 .
1
Secondly, substituting c = 2 in the recurrent formula we find
an
an+1 = − 21
(2n + 3) (n + 1)
a1 = − 61 a0 , a2 = 1
120 a0 ,
1
a3 = − 5040 a0 , a4 = 1
362880 a0 .
+B (1 − 12 x + 24
1 2 1
x − 720 1
x3 + 40320 x4 − . . . )
b)
y ′′ + xy ′ + y = 0.
The point x = 0 is an ordinary point of the equation. So, we are looking for a solution
in the form ∞
y = ∑ an x n .
n=0
Then the derivatives of the solution are
∞ ∞
y ′ = ∑ nan xn−1 , y ′′ = ∑ (n − 1) nan xn−2 .
n=1 n=2
or ∞ ∞ ∞
∑ (n − 1) nan xn−2 + ∑ an nxn + ∑ an xn = 0.
n=2 n=1 n=0
On adjusting the indices we calculate
∞ ∞ ∞
∑ (n + 1) (n + 2) an+2 xn + ∑ an nxn + ∑ an xn = 0.
n=0 n=1 n=0
a2 = − 21 a0 .
(n + 1) (n + 2) an+2 + an n + an = 0
or
an
an+2 = − .
n+2
Setting n = 0, 1, . . . in the recurrent formula we find
a2 = − 21 a0 , a3 = − 13 a1 , a4 = 18 a0 , a5 = 1
15 a1 , . . .
y = a0 + a1 x − 21 a0 x2 − 13 a1 x3 + 18 a0 x4 + 15
1
a1 x 5 + . . .
or ∞ ∞
∑ an (n + c − 1) (n + c) xn+c−1 + ∑ an xn+c = 0.
n=0 n=0
On adjusting the indices we obtain
∞ ∞
∑ an+1 (n + 1 + c) (n + c) xn+c + ∑ an xn+c = 0.
n=−1 n=0
c (c − 1) = 0
hence
c = 0, c = 1.
Equating the coefficients at different powers of x for n > −1 we obtain the recurrent
formula
an+1 (n + 1 + c) (n + c) + an = 0
or
an
an+1 = − .
(n + 1 + c) (n + c)
Firstly, substituting c = 0 in the recurrent formula we obtain
(n + 1) nan+1 = −an
which, however, gives a0 = 0 for n = 0 and then an = 0 for all n. Thus c = 0 does not
yield any non-trivial solutions.
Secondly, substituting c = 1 in the recurrent formula we find
an
an+1 = −
(n + 1) (n + 2)
By setting n = 0, 1, . . . in the previous recurrent formula we obtain
a1 = − 21 a0 , a2 = 1
12 a0 ,
1
a3 = − 144 a0 , . . .
y = Ax (1 − 12 x + 12
1 2 1
x − 144 x3 + . . . )
(x2 − 1) y ′′ + xy ′ + 3y = 0
Dividing through by the coefficient at the second derivative: x2 − 1 and then factorising,
we obtain
x 3y
y ′′ + y′ + = 0.
(x − 1) (x + 1) (x − 1) (x + 1)
7.4. TUTORIAL 4 263
Thus we have
x 3
p(x) = , q(x) = .
(x − 1) (x + 1) (x − 1) (x + 1)
The singular points are
x = −1, x = 1.
For x = −1 we need to calculate the limits:
x 3x + 3
lim (x + 1)p(x) = lim = 21 , lim (x + 1)2 q(x) = lim =0
x→−1 x→−1 x − 1 x→−1 x→−1 x − 1
2
(x2 − 1) y ′′ − (x − 1) y ′ + 3y = 0
Dividing through by the coefficient at the second derivative: (x2 − 1)2 and then factoris-
ing, we obtain
−(x − 1) 3y
y ′′ + 2
′
2y + 2 2 =0
(x − 1) (x + 1) (x − 1) (x + 1)
264 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
or
y′ 3y
y ′′ − 2 + 2 2 = 0.
(x − 1) (x + 1) (x − 1) (x + 1)
Thus we have
1 3
p(x) = − 2, q(x) = 2 2.
(x − 1) (x + 1) (x − 1) (x + 1)
The singular points are
x = −1, x = 1.
For x = −1 we need to calculate the limits:
1
lim (x + 1)p(x) = lim − = undef ined,
x→−1 x→−1 (x − 1) (x + 1)
3
lim (x + 1)2 q(x) = lim = 43
x→−1 x→−1 (x − 1)2
2
(x2 − x − 2) y ′′ + (x2 − 4) y ′ − xy = 0
Dividing through by the coefficient at the second derivative: (x2 − x − 2)2 and then
factorising, we obtain
(x − 2) (x + 2) −xy
y ′′ + 2 2y
′
+ 2 2 =0
(x + 1) (x − 2) (x + 1) (x − 2)
or
(x + 2) y ′ xy
y ′′ + 2 − 2 2 = 0.
(x − 2) (x + 1) (x + 1) (x − 2)
Thus we have
x+2 x
p(x) = 2, q(x) = − 2 2.
(x − 2) (x + 1) (x + 1) (x − 2)
The singular points are
x = −1, x = 2.
For x = −1 we need to calculate the limits:
x+2
lim (x + 1)p(x) = lim = undef ined,
x→−1 x→−1 (x + 1) (x − 2)
x
lim (x + 1)2 q(x) = lim − = 91
x→−1 x→−1 (x − 2)2
7.4. TUTORIAL 4 265
2+x x
lim(x − 2)p(x) = lim 2 = 94 , lim(x − 2)2 q(x) = lim − 2 = − 92 .
x→2 x→2 (x + 1) x→2 x→2 (x + 1)
Since one of the limits for the point x = −1 does not exist, then we conclude that this
point is an irregular singular point. Since both limits for the point x = 2 exist we
conclude that this point is a regular singular point.
Back to Tutorial 4 Question 3
2
(x2 + 2x − 8) y ′′ + (3x + 12) y ′ − x2 y = 0
Dividing through by the coefficient at the second derivative: (x2 + 2x − 8)2 and then
factorising, we obtain
3x + 12 −x2 y
y ′′ + 2 2 y ′
+ 2 2 =0
(4 + x) (x − 2) (4 + x) (x − 2)
or
3y ′ x2 y
y ′′ + 2 − 2 2 = 0.
(4 + x) (x − 2) (4 + x) (x − 2)
Thus we have
3 x2
p(x) = 2 , q(x) = − 2 2.
(4 + x) (x − 2) (4 + x) (x − 2)
3
lim(x − 2)p(x) = lim = undef ined,
x→2 x→2 (4 + x) (x − 2)
x2
lim(x − 2)2 q(x) = lim − 2 = − 91 .
x→2 x→2 (4 + x)
Since both limits for the point x = −4 exist we conclude that this point is a regular
singular point. Since one of the limits for the point x = 2 does not exist, we conclude
that this point is an irregular singular point.
Back to Tutorial 4 Question 3
x2 y ′′ − 2xy ′ − 10y = 0
266 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
c2 − 3c − 10 = 0.
1 ∞
y = 2 ∑ an xn
x n=0
or ∞
y = x ∑ an x n .
5
n=0
(x − 1) y ′′ + (x2 − 1) y ′ − 12y = 0
2
Dividing through by the coefficient at the second derivative: (x − 1)2 and then factoris-
ing, we obtain
(x − 1) (x + 1) ′ −12y
y ′′ + 2 y + 2 =0
(x − 1) (x − 1)
or
(x + 1) y ′ 12y
y ′′ + − 2 = 0.
x−1 (x − 1)
Thus we have
x+1 12
p(x) = , q(x) = − 2.
x−1 (x − 1)
We need to calculate the limits:
c2 + c − 12 = 0.
9x2 y ′′ + 9x2 y ′ + 2y = 0
We are looking for a solution in the form
∞
y = ∑ an xn+c .
n=0
or
∞ ∞ ∞
9 ∑ an (n + c) (n + c − 1) xn+c + 9 ∑ an (n + c) xn+1+c + 2 ∑ an xn+c = 0.
n=0 n=0 n=0
9c2 − 9c + 2 = 0.
or
9an−1 (n + c − 1)
an = − .
(3c − 1 + 3n) (3c − 2 + 3n)
1
Firstly, substituting c = 3 in the recurrent formula we calculate
an−1 (3n − 2)
an = −
n (−1 + 3n)
268 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
The coefficients of the series are then obtained by setting n = 1, 2, . . . in the previous
recurrent formula we get
a1 = − 12 a0 , a2 = 51 a0 , a3 = − 120
7
a0 , a4 = 7
528 a0 , . . .
2
Secondly, substituting c = 3 in the recurrent formula we get
a1 = − 21 a0 , a2 = 5
28 a0 ,
1
a3 = − 21 a0 , a 4 = 11
1092 a0 , . . .
x2 y ′′ + xy ′ + x2 y = 0
We are looking for a solution in the form
∞
y = ∑ an xn+c .
n=0
or ∞ ∞ ∞
∑ an (n + c) (n + c − 1) xn+c + ∑ an (n + c) xn+c + ∑ an xn+2+c = 0.
n=0 n=0 n=0
c2 = 0.
Hence
c = 0.
7.4. TUTORIAL 4 269
Equating the coefficients at different powers of x for n > 0 we get the recurrent formula
an n (n − 1) + an n + an−2 = 0
or
an−2
an = − .
n2
The coefficients of the series are then obtained by setting n = 1, 2, 3, . . . in the previous
recurrent formula we find
a1 = 0, a2 = − 41 a0 , a3 = 0, a4 = 1
64 a0 ,
1
a5 = 0, a6 = − 2304 ,....
y = a0 (1 − 14 x2 + 64
1 4 1
x − 2304 x6 + . . . ) .
x2 y ′′ + (x2 + x) y ′ − y = 0
We are looking for a solution in the form
∞
y = ∑ an xn+c .
n=0
or
∞ ∞ ∞ ∞
∑ an (n + c) (n + c − 1) xn+c + ∑ an (n + c) xn+c + ∑ an (n + c) xn+1+c − ∑ an xn+c = 0.
n=0 n=0 n=0 n=0
c2 − 1 = 0.
c = −1, c = 1.
270 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
Equating the coefficients at different powers of x for n > 0 we obtain the recurrent
formula
an (n + c) (n + c − 1) + an (n + c) + an−1 (n + c − 1) − an = 0
or
an−1
.
an = −
n+1+c
Firstly, substituting c = −1 in the recurrent formula we find
an−1
an = −
n
The coefficients of the series are then obtained by setting n = 1, 2, . . . in the previous
recurrent formula we get
(−1)n
a1 = −a0 , a2 = 21 a0 , a3 = − 16 a0 , a4 = 1
24 a0 , . . . , an = n! .
a1 = − 13 a0 , a2 = 1
12 a0 ,
1
a3 = − 60 a0 , a 4 = 1
360 a0 .
y = Ax (1 − 13 x + 12
1 2 1 3
x − 60 1
x + 360 x4 − . . . )
+Bx−1 (1 − x + 21 x2 − 16 x3 + 24
1 4
x − ...)
7.5 Tutorial 6
Question 1
Problem 2.i.
We calculate the indefinite integral (by parts):
RRR u = 3t + 4 u′ = 3 RRR
−ts
dt = RRRRR RRR = − (3t + 4)e−ts + 3 −ts
∫ (3t + 4) e RRR s∫ e dt
RRR dv = e−ts dt v = − e s s
−ts
RR
(3t + 4)e−ts 3e−ts
=− − 2 .
s s
Note that limt→∞ te−ts = 0. Thus we have
∞
∞ 4 3 3t 4 3 4 3
−ts −ts
∫0 (3t + 4) e dt = e (− − 2 − ) ∣ = 0 − (− − 2 ) = + 2 .
s s s 0 s s s s
7.5. TUTORIAL 6 271
Problem 2.ii
We have
at+b −ts b (a−s)t eb (a−s)t
∫ e e dt = e ∫ e dt =
a−s
e
hence ∞
∞ eb (a−s)t eb eb
at+b −ts
∫0 e e dt = e ∣ =0− =
a−s 0
a−s s−a
provided a − s < 0.
Problem 2.iii
This question involves two integrations by parts:
RRR u = cos (at) u′ = − sin (at) a RRR
cos −ts
e dt RRR RRR = − cos (at) e−ts − a −ts
∫ (at) = RRR −ts e−ts RRR s s ∫ e sin (at) dt
RR dv = e dt v = − s RR
RRR u = sin (at) u′ = cos (at) a RRR
= RRRRR RRR = − cos (at) e−ts + sin (at) ae−ts − a2 cos (at) e−ts dt.
e−ts a e−ts a R
R s s 2 s 2 ∫
RRR dv = − s dt v = s2 RRR
Thus we have
hence
∞
s 2 + a2 ∞ −ts cos (at) e−ts sin (at) ae−ts 1
∫ cos (at) e dt = (− + ) ∣ = .
s2 0 s s2 0
s
Finally we obtain
∞ s
∫0 cos (at) e−ts dt = .
s2 + a2
Question 2
1. i.
1
.
s (s + 1)
For this fraction the decomposition has the form
1 A B
= + .
s (s + 1) s s + 1
272 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
A = 1, B = −1.
Hence we have
f (t) = −e−t + 1.
Back to Tutorial 6 Question 2
ii.
1
.
(s − 4) (s − 1)
We’re looking for a decomposition of the form
1 A B
= + .
(s − 4) (s − 1) s − 1 s − 4
Hence we have
f (t) = 31 e4t − 13 et .
Back to Tutorial 6 Question 2
iii.
3s
.
s2 + 2s − 8
For this fraction the decomposition has the form
3s A B
= + .
(s + 4) (s − 2) s + 4 s − 2
−2A + 4B = 0, A + B = 3
whose solution is
A = 2, B = 1.
The partial fractions decomposition is given by
3s 2 1
= + .
(s + 4) (s − 2) s + 4 s − 2
Hence we have
f (t) = e2t + 2e−4t .
Back to Tutorial 6 Question 2
iv.
s2 − 6s + 4
.
s (s2 − 3s + 2)
For this fraction the decomposition has the form
s2 − 6s + 4 A B C
= + + .
s (s − 1) (s − 2) s − 1 s s − 2
Multiplying by the common denominator and collecting the coefficients at s, we
obtain
s2 − 6s + 4 = (A + C + B) s2 + (−2A − C − 3B) s + 2B
Comparing the coefficients we find the system
2B = 4, −2A − C − 3B = −6, A + C + B = 1.
A = 1, B = 2, C = −2.
2. i. Use the property of the Laplace transform: L(t2 +at+b) = L(t2 )+aL(t)+bL(1).
ii. Use the identity
L(f ′ ) = sL(f )
or
L(f ′ )
L(f ) = .
s
Further, we calculate the derivative
2ω 2
L(f ′ ) = .
s2 + 4ω 2
Hence
2ω 2
L(f ) = .
s(s2 + 4ω 2 )
Back to Tutorial 6 Question 2
which implies f (0) = 0, f ′ (0) = 1. The formula L(f ′′ ) = s2 L(f ) − sf (0) − f ′ (0)
then becomes
L(f ′′ ) = s2 L(f ) − 1.
On the other hand
2
L(f ′′ ) = −2L(sin t) − L(t cos t) = − − L(f ).
s2 +1
From the previous we get
2
s2 L(f ) − 1 = − − L(f ).
s2 + 1
Solving for L(f ) we obtain
s2 − 1
L(f ) = .
(s2 + 1)2
Back to Tutorial 6 Question 2
7.5. TUTORIAL 6 275
Question 3
1.
4y ′′ + π 2 y = 0, y(0) = 0, y ′ (0) = 1.
Applying the Laplace transform to the equation, we obtain
4s2 L (y) − 4y ′ (0) − 4sy (0) + π 2 L (y) = 0.
On substitution of the initial conditions, we get
(4s2 + π 2 ) L (y) − 4 = 0.
Solving for L(y) gives
π
4 2 2
L (y) = 2 2
= π 2.
4s + π s + ( π2 )
2
Applying the inverse Laplace transform to the right hand side, we obtain the solution
2
y= sin ( 12 πx) .
π
Back to Tutorial 6 Question 3
276 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
2.
y ′′ + 5y ′ + 6y = 0, y(0) = 0, y ′ (0) = 1.
Applying the Laplace transform to the equation, we obtain
(5s + s2 + 6) L (y) − 1 = 0.
3A + 2B = 1, A + B = 0.
A = 1, B = −1.
y = e−2t − e−3t .
3.
y ′′ + 25y = x, y(0) = 1, y ′ (0) = 1
25 .
1 1
(s2 + 25) L (y) − 25 −s= .
s2
Solving for L(y) gives
2
+ 25s3
1 25 + s
L (y) = 25 2 2
s (s + 25)
7.5. TUTORIAL 6 277
25 + s2 + 25s3 A + Bs C
= 2 + .
s2 (s2 + 25) s + 25 s2
25C = 25, A + C = 1, B = 25
which gives
A = 0, B = 25, C = 1.
Thus the required partial fractions decomposition is
2
+ 25s3
1 25 + s s 1 1
25 2 2
= 2 + 25 .
s (s + 25) s + 25 s2
Applying the inverse Laplace transform to the right hand side, we obtain the solution
1
y = cos (5t) + 25 t.
2 + 2s2 − s3 + s4 A + Bs C
= 2 + .
s3 (s2 + 1) s + 1 s3
2 + 2s2 − s3 + s4 = s3 A + s4 B + Cs2 + C
A = −1, B = 1, C = 2.
278 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
2 + 2s2 − s3 + s4 s − 1 2
3 2
= 2 + 3.
s (s + 1) s +1 s
Applying the inverse Laplace transform to the right hand side, we obtain the solution
9s + 7
s2 L (y) − y ′ (0) − sy (0) − 7sL (y) + 7y (0) + 10L (y) = .
s2 + 1
9s + 7
(s2 − 7s + 10) L (y) + 39 − 5s = .
s2 + 1
Applying the inverse Laplace transform to the right hand side, we obtain the solution
y ′′ − y = t − 2, y(2) = 3, y ′ (2) = 0.
First we make the substitution t − 2 = x. This brings the problem to the form
y ′′ − y = x, y(0) = 3, y ′ (0) = 0.
1 + 3s3 A B C
2
= + 2+ .
s (s − 1) (s + 1) s − 1 s s+1
B = −1, A + B − C = 0, A + C = 3
which yields
A = 2, B = −1, C = 1.
Thus the required partial fractions decomposition has the form
1 + 3s3 2 1 1
2
= − 2+ .
s (s − 1) (s + 1) s − 1 s s+1
Applying the inverse Laplace transform to the right hand side, we obtain the solution
y = 3 cosh (t − 2) + sinh (t − 2) − t + 2.
7 − s − s2 + s3
L (y) =
(s2 + 1) (s + 1) (s − 2)
7 − s − s2 + s3 A B + sC K
2
= + 2 + .
(s + 1) (s + 1) (s − 2) s − 2 s +1 s+1
K = −1, A = 53 , B = − 11 7
5 , C = 5.
7 − s − s2 + s3 1 7s − 11 1
2
= 35 + 15 2 − .
(s + 1) (s + 1) (s − 2) s−2 s +1 s+1
Applying the inverse Laplace transform to the right hand side, we obtain the solution
−x
y = 53 e2x + 75 cos (x) − 11
5 sin (x) − e .
Question 4
1. We have
4 2 2
= 2 .
s3 − 4s s s − 4
We know that
2
L−1 ( ) = sinh 2t
s2 −4
hence
t
2 2 t
−1
L ( 2 ) = 2 ∫ sinh 2τ dτ = cosh 2τ ∣ = cosh 2t − 1.
ss −4 0
0
282 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
2. We have
1 1 1
= .
s2 + as s s + a
We know that
1
L−1 ( ) = e−at
s+a
hence
t
1 1 t 1 1
−1
L ( ) = ∫ e−aτ dτ = − e−aτ ∣ = − (e−at − 1).
ss+a 0 a 0
a
3. We have
8 2 4
= 3 .
s4 − 4s 2 s s − 4s
Previously we’ve calculated that
4
L−1 ( ) = cosh 2t − 1
s3 − 4s
hence
t
2 4 t
−1
L ( 3 ) = ∫ (2 cosh 2τ − 2)dτ = (sinh 2τ − 2τ )∣ = sinh 2t − 2t.
s s − 4s 0
0
Question 5
1. i. We have
2
L(t2 ) =
.
s3
Applying the s-shifting theorem we obtain
2
L(t2 e−2t ) =
(s + 2)3
ii. We have
s
L(cos t) = .
s2 +1
Applying the s-shifting theorem we obtain
s+1
L(e−t cos t) = .
(s + 1)2 + 1
iii. We have
3
L(sin 3t) = .
s2 +9
Applying the s-shifting theorem we obtain
3
L(e−2 sin 3t) = .
(s + 2)2 + 9
7.5. TUTORIAL 6 283
iv. We have
6
L(t3 ) =
.
s4
Applying the s-shifting theorem we obtain
12
L(2t3 e−t/2 ) =
(s + 21 )4
Back to Tutorial 6 Question 5
2. i. We know that
2
L−1 ( 21 3
) = 12 t2
s
hence
2
L−1 ( 12 ) = 21 t2 e−t/2 .
(s + 12 )3
ii. We first complete the square
1 1
= .
s2 + 2s + 5 (s + 1)2 + 4
We know that
1
L−1 ( ) = 12 sin 2t
s2 +4
hence
1
L−1 ( ) = 21 e−t sin 2t.
(s + 1)2 + 4
iii. We first complete the square
6 3×2
= .
s2 − 4s − 5 (s − 2)2 − 9
We know that
3
L−1 ( ) = sinh 3t
s2 −9
hence
3×2
L−1 ( ) = 2e2t sinh 3t.
(s − 2)2 − 9
Back to Tutorial 6 Question 5
3. i.
y ′′ − 4y ′ + 5y = 0, y(0) = 1, y ′ (0) = 2.
Applying the Laplace transform to the equation, we obtain
s2 L (y) − y ′ (0) − sy (0) − 4sL (y) + 4y (0) + 5L (y) = 0.
On substitution of the initial conditions, we get
(−4s + s2 + 5) L (y) + 2 − s = 0.
Solving for L(y) gives
s−2 s−2
L (y) = =
s2
− 4s + 5 (s − 2)2 + 1
Applying the inverse Laplace transform to the right hand side and using the s-
shifting theorem, we obtain the solution
y = e2t cos (t) .
284 CHAPTER 7. SOLUTIONS TO SELECTED TUTORIAL QUESTIONS
ii.
9y ′′ − 6y ′ + y = 0, y(0) = 3, y ′ (0) = 1.
Applying the inverse Laplace transform to the right hand side, we obtain the
solution
1
y = 3e 3 t .
iii.
y ′′ + 2y ′ + 5y = 9 cosh (2t) + 4 sinh (2t) , y(0) = 1, y ′ (0) = 2.
9s + 8
s2 L (y) − y ′ (0) − sy (0) + 2sL (y) − 2y (0) + 5L (y) = .
s2 − 4
9s + 8
(2s + s2 + 5) L (y) − 4 − s = .
s2 − 4
5s − 8 + 4s2 + s3 5s − 8 + 4s2 + s3
L (y) = =
(s − 2) (s + 2) (2s + s2 + 5) (s − 2) (s + 2) ((s + 1)2 + 4)
5s − 8 + 4s2 + s3 A B + sC K
2
= + 2
+ .
(s − 2) (s + 2) ((s + 1) + 4) s − 2 (s + 1) + 4 s + 2
5s − 8 + 4s2 + s3 1 1 2
2
= 21 + 21 + .
(s − 2) (s + 2) ((s + 1) + 4) s+2 s − 2 (s + 1)2 + 4
Applying the inverse Laplace transform to the right hand side, we obtain the
solution
y = cosh (2t) + e−t sin (2t) .
Conclusion
287