Calculus Notes
Calculus Notes
D. T. Papageorgiou
Department of Mathematics
Imperial College London
London SW7 2AZ, UK
©Demetrios T. Papageorgiou (2023) These notes are provided for the personal
study of students taking this module. The distribution of copies in part or whole
is not permitted.
2
Contents
2 Derivative of a Function 19
2.1 Definition with limits, examples . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2 General rules, chain rule, rates of change . . . . . . . . . . . . . . . . . . . 22
2.3 Implicit differentiation, related rates of change . . . . . . . . . . . . . . . 25
4 Inverse Functions 35
II Integration 57
6 Anti-derivatives and Geometrical Interpretation 59
6.1 Area under a curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.2 Application to a first order differential equation . . . . . . . . . . . . . . . 62
9 Improper Integrals 73
11 Techniques of Integration 79
3
4 CONTENTS
This set of notes is not extensive (still not exhaustive) but not everything in them
will be worked through in the face-to-face lectures. The student should strive to work
through the notes and also to solve problems assigned in the notes and also in the seven
problem sheets (one for each week of lectures).
A brief roadmap and guide of use is in order. There are six Parts. Part I - Fundamen-
tals of Differential Calculus - is mostly familiar from mathematics courses at A-level/IB.
The rigorous ε − δ definitions of limits are included for completeness and are covered in
detail in Analysis I. I include the brief discussion here to set the stage for the terminology
that will be used later. Part I will be covered very quickly due to its familiarity. Read
through and complete exercises. Part II is concerned with integration including im-
proper integrals and methods to test convergence, in addition to simple applications on
volumes and surfaces of revolution, centres of mass and moments of inertia (again a lot
of the material is probably familiar). Part III covers sequences and series (convergence
tests) from a practical viewpoint; theorems will be stated and proved for completeness
but again details are left for Analysis courses. Part III also covers power series and
Taylor series - applications to series solutions of differential equations are also included.
Part IV considered trigonometric approximation of single-variable functions and Fourier
series approximations of periodic functions. Part V has the general title Further Appli-
cations in Geometry and Physics and contains sections on the theory of plane curves,
Laplace transforms and applications (integral equations will be encountered here for the
first time), and a self-contained section on Newton’s laws of gravitation and their us in
describing planetary motion.
I will assume throughout that students have mastered the material in Maths and
Further Maths (or equivalent). In particular things like: differentiation of standard func-
tions (including product and chain rule); curve sketching (maxima/minima, inflection
points); integration of standard functions; techniques of integration (including integra-
tion by parts, integration by substitution, partial fractions); elementary mechanics. We
will be using all of these and I will assume full and masterful knowledge.
At the end of the notes I have part VI called Miscellaneous Problems. These problems
are mostly related to Parts I to IV and are additional to problems appearing in the notes
and in the seven problem sheets. You can figure out which part of the notes they refer to
but the line breaks sort of indicate approximate groupings. Some of these problems are
designed to be a bit more straightforward in an effort to allow the student to practice
before attacking more challenging problems.
Part I
Fundamentals of differential
calculus
7
Chapter 1
All this material will be covered in a lot of detail in Analysis. I am including some basic
results that are useful for us going forward, and will not spend any time on going over
the material. Please go over this in your own time.
Given a function f (x) we are concerned with the behaviour near a point x = x0 , and
in particular the meaning of the statement limx→x0 f (x) = `, i.e. the limit of f (x) as x
tends to x0 , exists and is equal to `. The precise ε − δ definition is the following.
Let f be a function defined at all points near x0 , except possibly at x0 , and let
` be a real number. We say that ` is the limit of f (x) as x approaches x0 , if for
every ε > 0 there exists δ > 0 such that |f (x) − `| < ε whenever |x − x0 | < δ and
x 6= x0 . We write limx→x0 f (x) = `.
Note: The definition implies that limx→x0 is unambiguous regarding taking limits
from above or below. If you see limx→x0 then it means that we must take the limit both
from above, i.e. limx→x0 + , and from below, i.e. limx→x0 − .
√ √
Example Prove that limx→2 x = 2. Before doing the ε − δ proof, we should note
√
that what we are asked to show appears trivially √ obvious. The function f (x) = x near
x = 2 is perfectly well-behaved and so f (2) = 2 is immediate. I am using the example
to illustrate the rigorous definition in anticipation of its use in less obvious situations.
√ √
Solution: Here f (x) = x, x0 = 2√and we want to prove that ` = 2. Given ε > 0 we
√
need to
√ √ find δ > 0 so√that |√ x − 2| < ε whenever |x − 2| < δ. For all x > 0 we have
x − 2 = (x − 2)/( x + 2), hence
√ √ |x − 2| |x − 2|
| x − 2| = √ √ ≤ √ .
x+ 2 2
√
Hence picking δ = 2 ε will do.
9
10 CHAPTER 1. LIMITS OF FUNCTIONS, CONTINUITY
Instead we use the following laws of limits which can be proven easily using the ε − δ
definition.
Assume that limx→x0 f (x) and limx→x0 g(x) BOTH EXIST. Then
Example 1
Calculate limx→1 √x−1
x−1
.
Solution. Of the form “0/0”. Rationalise, i.e.
√ √x + 1)
x−1 (x − 1)( x + 1) −1)(
(x
lim √ = lim √ √ = lim = 2.
x→1 x−1 x→1 ( x − 1)( x + 1) x→1 −1)
(x
Example 2
Sketch the function f (x) = x/|x|. Do this by considering x > 0 and x < 0 separately.
What happens when x = 0?
Solution. If x > 0, then f (x) = 1. If x < 0, then f (x) = |x| x
= −|x|
|x| = −1.
At x = 0 the function is undefined. Hence this function is defined everywhere except
at x = 0.A plot is given in Figure 1.1. The function is discontinuous at x = 0 as can be
seen from the jump in going from left to right (it is called a “step function” and can be
written as a linear combination of Heaviside functions - see below).
We have a problem, however. How do we define the function at x = 0? The answer
is that we can define it to be anything we want, and irrespective of our choice we can
11
1.5
0.5
f(x)=x/|x|
-0.5
-1
-1.5
-2
-1 -0.8 -0.6 -0.4 -0.2 0 0.2 0.4 0.6 0.8 1
x
x
Figure 1.1: The function f (x) = |x| .
never make the function continuous there! In figure 1.1 it is chosen to be zero (i.e. the
average of the values as I approach from the right and the left - we will encounter this
later when we do Fourier series). Here is a summary of what figure 1.1 depicts:
x
|x| x 6= 0
f (x) = (1.6)
0 x=0
Note that:
1. limx→0+ f (x) = 1
2. limx→0− f (x) = −1
where x → 0± means take the limit as x tends to 0 from above/below. Of course
limx→0 f (x) 6= f (0) for this function, and so the function is not continuous. The defini-
tion of continuity is summarised below in Definition 5.
The properties given above also hold as x becomes large and positive or negative.
For example if f (x) = 1/x then we know that limx→±∞ f (x) = 0. Lets make this precise.
Let f (x) be defined on a domain containing the interval (a, ∞). A real number
` is the limit of f (x) as x approaches ∞ if, for every ε > 0 there exists a A > a,
such that |f (x) − `| < ε whenever x > A. We write limx→∞ f (x) = `. [Similarly
for limx→−∞ f (x) = `.]
NOTE: The limit properties (1.1)-(1.5) hold for limits of f (x) as x → ±∞, when the
limits are defined.
Consider next the limits limx→0 sin(1/x) and limx→0 (1/x2 ). The limits do not exist (I
cannot plug x = 0 into the functions). Sketch them and determine that they behave
12 CHAPTER 1. LIMITS OF FUNCTIONS, CONTINUITY
differently: the former is bounded, the latter is unbounded. In fact limx→0 (1/x2 ) = ∞.
More precisely we have:
In the example f (x) = 1/x2 we found that as x → 0, f (x) → ∞ - the function is even,
so it does not matter if I approach the limit from the right (i.e. through positive values
of x) or the left (through negative x values). What about f (x) = 1/x? It is not hard to
see that as x tends to 0 through positive values then f → +∞, whereas as x tends to 0
through negative values we have f → −∞.
Hence, we need to define one-sided limits.
Let f (x) be defined for all x in an interval (x0 , a). We say that f (x) approaches
` as x approaches x0 from the right if, for any ε > 0, there exists a δ > 0, such
that for all x0 < x < x0 + δ we have |f (x) − `| < ε. We write limx→x0 + f (x) = `.
[Analogous definition for the left-sided limit, i.e. limx→x0 − f (x) = `.]
and x = 2 and a horizontal asymptote y = 0. Sketch the graph without using the
differentiation methods of finding critical points etc., that you are familiar with. Use
intuition and estimation.
In addition to the basic properties (1.1)-(1.5), there is another powerful test which is
very useful in calculations:
1. If limx→x0 f (x) = 0 and |g(x)| ≤ |f (x)| for all x near x0 with x 6= x0 , then
limx→x0 g(x) = 0.
2. If limx→∞ f (x) = 0 and |g(x)| ≤ |f (x)| for all large enough x, then
limx→∞ g(x) = 0.
Example
13
Figure 1.2: Geometrical construction in the proof of the trigonometric limits (1.7)-(1.8).
Solution
(i) Since limx→x0 f (x) = 0, then given ε > 0, there exists a δ > 0 such that |f (x)| < ε
when |x − x0 | < δ. For the same ε and δ, we also have |g(x)| < ε when |x − x0 | < δ,
since |g(x)| ≤ |f (x)|. Hence limx→x0 g(x) = 0 also.
(ii) Take g(x) = x sin(1/x) and f (x) = x. Then |g(x)| ≤ |x| for all x 6= 0, so the
comparison test applies. Clearly limx→0 x = 0, hence the result follows.
We will need the following results in finding derivatives of sin x and cos x from
first principles.
sin h
lim =1 (1.7)
h→0 h
cos h − 1
lim =0 (1.8)
h→0 h
The proof of the former is geometrical and the construction is given in Figure 1.2.
OBC is the sector of a circle of radius 1 with subtended angle h. The two triangles
OAB and OCD are constructed as shown with BD the extension of OB. Considering
triangles OAB and OCD we have
AB DC
sin h = = s, tan h = = t.
OB OC
From geometry we have the following inequality
area of triangle OAB < area of sector OCB < area of triangle OCD,
which in turn provides
1 h 1
sin h cos h < < tan h.
2 2 2
14 CHAPTER 1. LIMITS OF FUNCTIONS, CONTINUITY
The middle quantity follows by noting that the area of the sector OCB is equal to h/2π
times the area of a circle of unit radius which is π. Considering the first inequality (after
canceling the 1/2 factor throughout) we have
sin h 1
sin h cos h < h ⇒ < .
h cos h
The above is fine since h and cos h are positive and non-zero so I can divide by them.
The second inequality gives
sin h sin h
h< ⇒ cos h < .
cos h h
Putting these together gives
sin h 1
cos h < < .
h cos h
As h tends to zero cos h tends to 1, hence sin h/h is squeezed between two numbers that
tend to 1. By the Squeezing Property we get the desired result.
To prove the second result we write
: −1
*0
: 1/2
sin h sin h −sinh 1
lim − = lim
lim sin
h lim = 0.
h→0 h cos h + 1 h→0 h
h→0
cos h + 1
h→0
Continuity
Looking back at Definition 1, the ε − δ definition of a limit, we can see that it is
equivalent to the statement
lim f (x0 + h) = `.
h→0
Definition 5. Continuity
x2
1 x>0 x 6= 0
f (x) = g(x) =
0 x≤0 1 x=0
Removable discontinuities
We will encounter functions that are perfectly nice everywhere except at a point or
points. Here are some examples.
1
f (x) = sin (1.9)
x
1
f (x) = x sin (1.10)
x
Both of these functions are well behaved everywhere except at x = 0. We say that the
functions are not defined at x = 0.
The question is: Can we define f (0) by a number of our choice so as to make the
functions (1.9) and (1.10) continuous?
For the former function this is impossible - the function oscillates infinitely many
times as x → 0 and takes on all values between −1 and 1. We can never satisfy the ε − δ
definition of continuity.
However, for the function in (1.10) we can remove the discontinuity by picking f (0) =
0, i.e.
1
x sin x x 6= 0
f (x) = (1.11)
0 x=0
f(x)=sin(1/x)
1
0.5
-0.5
-1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
f(x)=x sin(1/x)
0.8
0.6
0.4
0.2
0
-0.2
-0.4
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Figure 1.3: The functions sin(1/x) and x sin(1/x). Only x > 0 is shown, the functions
are odd and even, respectively, with respect to x.
1 1−x2
3. Find (a) limx→1 (x−1)2
, and (b) limx→∞ x3/2
.
For (a) as x → 1 (from above or below), then (x − 1)2 becomes arbitrarily small.
1
Its inverse becomes arbitrarily large, so limx→1 (x−1)2 = ∞
Intuitive answer is: For (b) as x becomes very large then x2 x3/2 , hence the
limit is −∞. Can formalize as follows
1 − x2
lim = lim (x−3/2 − x1/2 ) = −∞.
x→∞ x3/2 x→∞
Problems
1. For the function
x
x 6= 0
f (x) = |x|
c x=0
where c ∈ R, use the ε − δ definition to prove that the function is not continuous
at x = 0.
2. Consider the function
x1/2
x≥0
f (x) =
x2 x<0
Clearly f (0) = 0 and the function is continuous there. Prove continuity using ε−δ.
[Hint: The task here is, given a ε to find one δ that will work for both x > 0 and
x < 0.]
18 CHAPTER 1. LIMITS OF FUNCTIONS, CONTINUITY
Chapter 2
Derivative of a Function
Figure 2.1: Slope of f at P is the slope of the line QP as Q tends to P . Note: The Qs
are to the right of P , the definition is the same is Q1 , Q2 etc are to the left of P .
Definition of Differentiability
f (x + h) − f (x)
lim
h→0 h
exists. We call this f 0 (x), the derivative of f at point x.
19
20 CHAPTER 2. DERIVATIVE OF A FUNCTION
Examples
(x + h)2 − x2 x2 + 2xh + h2 − x2
f 0 (x) = lim = lim
h→0 h h→0 h
= lim (2x + h) = 2x ⇒ YES
h→0
Need to check if the limit exists and the values are equal as we approach 0 from
above or below.
(a)
f (0 + h) − f (0) h−0
lim = lim =1
h→0, h>0 h h→0, h>0 h
(b)
f (0 + h) − f (0) −h − 0
lim = lim = −1
h→0, h<0 h h→0, h<0 h
Right and left derivatives exist but are not equal.
Example: Consider (
x if 0 < x ≤ 1
f (x) =
x−1 if 1 < x ≤ 2
Here is the graph:
f (1 + h) − f (1) 1+h−1
lim = lim =1
h→0, h<0 h h→0, h<0 h
f (1 + h) − f (1) (1 + h − 1) − 1
lim = lim as f (1) = 1
h→0, h>0 h h→0, h>0 h
1
= lim (1 − )
h→0, h>0 h
which does not exist. In fact, → −∞. Function has no right derivative.
2.1.1 Polynomials
Theorem 1
Let n be an integer ≥ 1 and let f (x) = xn . Then
df
f 0 (x) = = nxn−1 .
dx
22 CHAPTER 2. DERIVATIVE OF A FUNCTION
Proof.
f (x + h) = (x + h)n = xn + nxn−1 h + h2 g(x, h)
where g(x, h) involves powers of x and h with some numerical coefficients. We don’t
care what it is exactly but limh→0 g(x, h) = some number. Then
df xn + nxn−1 h + h2 g − xn
= lim = nxn−1 .
dx h→0 h
Theorem 2
Let f (x) = xa , where a is any real number and x > 0. Then f 0 (x) = axa−1 .
If a is a negative integer then this is easy. General case is different from proof
above.
f (x + h)g(x + h) − f (x)g(x)
(f g)0 = lim
h→0 h
=0
z }| {
f (x + h)g(x + h) − f (x)g(x) −f (x)g(x + h) + f (x)g(x + h)
= lim
h→0 h
(f (x + h) − f (x)) g(x + h) + (g(x + h) − g(x)) f (x)
= lim
h→0 h
= g(x)f 0 (x) + f (x)g 0 (x)
2.2. GENERAL RULES, CHAIN RULE, RATES OF CHANGE 23
Proof.
f (g(x + h)) − f (g(x))
(f ◦ g)0 (x) = lim
h→0 h
f (g(x + h)) − f (g(x)) g(x + h) − g(x)
= lim ·
h→0 g(x + h) − g(x) h
Let k = g(x + h) − g(x), (if h 6= 0, k 6= 0), and write u = g(x). Then
f (u + k) − f (u) g(x + h) − g(x)
(f ◦ g)0 (x) = lim ·
h→0
k h
f (u + k) − f (u) g(x + h) − g(x)
= lim lim
h→0 k h→0 h
0 0 0 0
= f (u)g (x) = f (g(x))g (x)
24 CHAPTER 2. DERIVATIVE OF A FUNCTION
f (y) − f (x)
f 0 (x) = lim .
y→x y−x
To see equivalence write y = x + h.
f (t2 ) − f (t1 )
average speed = .
t2 − t1
So instantaneous speed at any time t is
f (t) − f (t0 )
f 0 (t) = lim rate of change.
t→t0 t − t0
d2 f
v 0 (t) = is the acceleration.
dt2
Can define higher derivatives (if they exist) by continuing this process.
Theorem 4
If f (x) is differentiable at x = x0 , then it is also continuous there.
Can prove these using implicit differentiation. Start with (i) y = x1/n , n integer.
Assume x1/n is defined.
Then
d n d
yn = x ⇒ (y ) = (x) ⇒
dx dx
dy dy 1 1 1−n 1 1
ny n−1 =1 ⇒ = y 1−n = x n = x n −1 .
dx dx n n n
e.g.
d 1 1 4
x 5 = x− 5
dx 5
p 1
(ii) y = x q . Let g(x) = x q where q is an integer. Then y = (g(x))p with p an integer.
Use chain rule.
dy 1 1 −1 p p −1
= pg p−1 x q = x q
dx q q
These of course generalize to powers of the function.
e.g.
d
(f (x))r = rf r−1 f 0 , r rational.
dx
6
Its equation is y − 2 = 23 (x − 1).
dx dy dy x dx
2x + 2y =0⇒ =−
dt dt dt y dt
dy/dt x
i.e =− .
dx/dt y
dy
√
This is the derivative of dx if we think of y = ± 1 − x2 as a function of x. (Another
dy dy dy/dt
way is x2 + y 2 = 1 → 2x + 2y dx = 0 → dx = − xy = dx/dt .)
In general, if x = f (t) and y = g(t), describe a curve in the plane called a parametric
curve, then the slope of the tangent line to the curve is given by
dy dy/dt dx
= if 6= 0.
dx dx/dt dt
To prove this, note that the curve can be defined (piecewise) as the graph of a
function y = h(x) or x = H(y). Chain rule gives dy dy dx
dt = dx dt as required.
Example
The surface area of a cube is growing at a constant rate of 4cm2 /s. How fast is the
length of a side growing when the cube sides are 2cm long? Find the side length when
the rate of change of the volume exceeds that of the area.
2.3. IMPLICIT DIFFERENTIATION, RELATED RATES OF CHANGE 27
Solution
dA dx dx 1 dA 4
A = 6x2 ⇒ = 12x ⇒ = =
dt dt dt 12x dt 12x
dx 1
If x = 2 = cm s−1
dt 6
dV dx 3x2 dA
V = x3 ⇒ = 3x2 = .
dt dt 12x dt
dV dA
So if x > 4, dt > dt in numerical value.
28 CHAPTER 2. DERIVATIVE OF A FUNCTION
Chapter 3
e.g.
Theorem 1
Let f be a function which is defined and differentiable on the open interval (a, b).
Let c be a number in the interval which is a maximum for the function.
29
30 CHAPTER 3. MEAN VALUE AND INTERMEDIATE VALUE THEOREMS
Detailed proof:
f (c) ≥ f (c + h) ⇒ f (c + h) − f (c) ≤ 0.
f (c + h) − f (c)
i.e. lim ≤0
h→0, h>0 h
Definition
f (x) is said to be continuous on an interval [a, b] if limx→x0 f (x) = f (x0 ) for all
x0 in [a, b]. Analogously, limh→0 f (x0 + h) = f (x0 ) a ≤ x0 ≤ b.
31
Theorem 2
Let f (x) be continuous on the closed interval [a, b]. Then f (x) has a maximum
and a minimum on this interval. I.e there exist c1 and c2 so that f (c1 ) ≥ f (x)
and f (c2 ) ≤ f (x) for all x in [a, b].
e.g
Proof. If f (x) = 0 then there is nothing to prove. If f (x) 6= 0 then it must have at least
one maximum or minimum or both. Denote by c such points. By Theorem 1 we have
f 0 (c) = 0.
Example:
Theorem 4
Suppose f is continuous on [a, b] and differentiable on (a, b).
Then there exists a < c < b such that
f (b) − f (a)
f 0 (c) =
b−a
Geometrical interpretation:
f (b)−f (a)
The point x = c is where the tangent has the slope b−a .
Proof. The straight line joining (a, f (a)) and (b, f (b)) has equation
y − f (a) x−a f (b) − f (a)
= ⇒ y= (x − a) + f (a).
f (b) − f (a) b−a b−a
Consider g(x) = f (x) − f (b)−f b−a
(a)
(x − a) − f (a). Then g(a) = 0, g(b) = 0, and by
Theorem 3 there exists a c with a < c < b such that g 0 (c) = 0.
But g 0 (x) = f 0 (x) − f (b)−f
b−a
(a)
, and the result follows.
Definition
We say that a continuous function f (x) is increasing over a given interval if
given x1 , x2 in the interval with x1 ≤ x2 , we have f (x1 ) ≤ f (x2 ).
Theorem 5
Let f (x) be continuous in the closed interval [a, b] and differentiable in the open
interval I = (a, b).
If f 0 (x) = 0 in I, then f is constant.
If f 0 (x) > 0 in I, then f is strictly increasing.
If f 0 (x) < 0 in I, then f is strictly decreasing.
f (x2 ) − f (x1 )
f 0 (c) = =⇒ f (x2 ) − f (x1 ) = (x2 − x1 )f 0 (c)
x2 − x2
If f 0 (x) = 0 in the interval, f 0 (c) = 0 and f (x2 ) = f (x1 ) i.e f is constant. If f 0 (x) > 0
then f 0 (c) > 0 and f (x2 ) > f (x1 ), i.e strictly increasing. If f 0 (x) < 0 then f 0 (c) < 0 and
f (x2 ) < f (x1 ), i.e strictly decreasing.
Example 1 (Do yourself ). Determine the region of increase and decrease of the function
f (x) = x3 − 2x + 1.
Picture where is does not work because the function is not continuous:
Chapter 4
Inverse Functions
Definition
Let y = f (x) be defined on some interval. Given any y0 in the range of f , if we
can find a unique value x0 in its domain such that f (x0 ) = y0 , then we can
define the inverse function
Clearly we have
f (g(y)) = y and g(f (x)) = x
or f (f −1 (y)) = y and f −1 (f (x)) = x
35
36 CHAPTER 4. INVERSE FUNCTIONS
Theorem 1
Let f (x) be strictly increasing or strictly decreasing. Then the inverse function
exists.
Theorem 2
If f (x) is continuous on [a, b] and is strictly increasing (or decreasing), and f (a) =
ya and f (b) = yb , then x = g(y) is defined on [ya , yb ].
Theorem 3
Let f (x) be differentiable on (a, b) and f 0 (x) > 0 or f 0 (x) < 0 for all x in (a, b).
Then the inverse function exists and we have
d −1 1
g 0 (y) = f (y) = 0 .
dy f (x)
Chain rule way: g(y) = x where y = f (x). (In fact x = g(y), iff y = f (x)). i.e
dg 0 1
f (x) = 1 g 0 = 0
dy f (x)
Example: Consider
y = x4 + 3x3 + x − 5, x > 0.
Find g 0 (0) - i.e d
dy f −1 (y) y=0
. Note: We will not even attempt to find x = g(y).
dg
Solution: Theorem says dy = g 0 (y) = f 01(x) where y = f (x). If y = 0 then need to
solve 0 = f (x) by inspection f (1) = 0 ⇒ f 0 (1) = 14 ⇒ g 0 (0) = 14
1
.
The solution above implicitly assumed that x = 1 is the only point where f (x) = 0.
Need to justify this. The key is the observation that f 0 (x) = 4x3 + 9x2 + 1 which is > 0
for all x > 0. Hence the function is strictly increasing in the given interval and since
f (0) = −5 < 0 and limx→+∞ = +∞, by the intermediate value theorem there must be
a unique x s.t. f (x) = 0. Of course this happens to be x = 1 by inspection.
Here are some additional things for you to consider: Suppose x < 0 now.
(i) Prove that there is at least one x < 0 such that f (x) = 0. Let this point(s) be
x = ξ, i.e. f (ξ) = 0, ξ < 0.
38 CHAPTER 4. INVERSE FUNCTIONS
(ii) Prove that there is at least one point ξ < c < 0 where f 0 (c) = 0.
1 1 π π
g 0 (y) = = > 0 x ∈ (− , ).
f 0 (x) (sin(x))0 2 2
39
dx 1 1 1
g 0 (y) = = =p 2
=p .
dy cos(x) 1 − sin (x) 1 − y2
Think of y as a dummy variable now. Then arcsin(y) = sin−1 (y) is a function with
domain [−1, 1] and range [− π2 , π2 ]. Instead of y use x, i.e. y = f (x) = sin−1 (x).
Then
dy d 1
= sin−1 (x) = √ .
dx dx 1 − x2
Once you have identified the domain and range where the inverse function exists,
there is an easier way (equivalent) to find derivatives.
Start with y = sin−1 (x) i.e y is the angle whose sin is x. Then sin(y) = x. By the
dy dy 1 1
chain rule, (cos(y)) dx = 1 ⇒ dx = cos(y) = √1−x 2
as shown above.
y = tan−1 (x)
tan(y) = x
dy dy 1
(1 + tan2 (y)) =1 ⇒ =
dx dx 1 + x2
Miscellaneous Problems
(ii) If the height at time t is h(t), find an explicit expression for it. What happens
to its rate of change as t becomes large? Explain physically/intuitively.
5. (a) Determine the regions of increase and decrease of the function f (x) = x3 −
2x + 1.
(b) Sketch functions for which the intermediate value theorem holds and:
(i) For a chosen y ∗ there are at most two values of x∗ .
(ii) For a chosen y ∗ I can choose an interval [a, b] to have as many x∗ as I
want. [Any guess as to what function this is?]
(iii) For a chosen y ∗ there does not exist a x∗ , i.e. Theorem 6 does not hold.
1732
√ 17321
√
3 3
2 1000 < 2 < 2 10000 i.e. 2 ≈ 3.322 correct to 3 decimal places.
There is a way of defining ax and deriving all its properties by using properties of
real numbers - though this is technical and we do not have time to do it.
We will do it in a more intuitive way that may seem a bit unnatural at first, namely by
defining a new function (the logarithm) and then defining the exponential as the inverse
function of the logarithm. Advantage of this - intuitive, simple and clear arguments.
Definition
The quantity log(x) is the area under the curve x1 between 1 and x if x ≥ 1; and
negative the area under the curve x1 between 1 and x if 0 < x < 1. In
particular, log(1) = 0.
43
44 CHAPTER 5. EXPONENTIALS AND LOGARITHMS
Theorem 1
d
log(x) is differentiable and dx log(x) = x1 .
log(x + h) − log(x) is the shaded area above. From geometry we have (or from the
fact that x1 is a decreasing function):
5.1. GEOMETRICAL DEFINITION, DERIVATIVE 45
1 1
· h < log(x + h) − log(x) < ·h
x+h x
1 log(x + h) − log(x) 1
⇒ < <
x+h h x
(here h > 0 so we can divide by h as done above.) As h → 0 we use the squeezing
theorem to get the required limit.
If h < 0 the picture is:
So
1 1
−h · < log(x) − log(x + h) < −h ·
x x+h
1 log(x + h) − log(x) 1
⇒ < <
x+h h x
1
Hence limit is x as h → 0.
d
log(x) is a function defined for x > 0, has log(1) = 0 and dx log(x) = x1 . (*)
We will use (*) alone in what follows. If g(x) is another such function then g(x) =
log(x) uniquely. The condition log(1) = 0 fixes this.
Theorem 2
If a, b > 0, then log(ab) = log(a) + log(b).
Theorem 3
log(x) is strictly increasing for all x > 0. Its range is (−∞, ∞).
Proof.
d 1
log(x) = > 0 for all x > 0 ⇒ strictly increasing.
dx x
To prove that it takes on arbitrarily large values, note that since it is strictly increas-
ing and log(1) = 0, we must have, for example, log(2) > 0.
From Theorem 2,
n terms
z }| {
n
log(2 ) = log(2 · 2 · · · · · 2) = log(2) + log(2) + · · · + log(2) = n log(2).
This holds for any positive integer and log(2) > 0 ⇒ as n becomes large, so does
log(2n ). To prove it takes on arbitrarily large negative values, note that
1 1
0 = log(1) = log 2 · = log(2) + log
2 2
1
⇒ log = − log(2)
2
Hence
1
log n = −n log(2) by Theorem 2
2
→ −∞ as n → ∞
Theorem 4
If n is an integer (positive or negative) then log(an ) = n log(a) for all a > 0.
Theorem 5
If x1 , x2 are two numbers, then exp(x1 + x2 ) = exp(x1 ) · exp(x2 ).
Theorem 6
d
exp(x) is differentiable and dx exp(x) = exp(x).
d
Proof. Think of dx exp(x) as the derivative of the inverse function to log(x). Hence it is
differentiable.
dy 1
We have proved that if g(y) is the inverse function of y = f (x), then dx = dx/dy i.e.
f 0 (x) = g0 (y) or g 0 (y) = f 0 (x) . Let y = exp(x), then the inverse function is x = log(y).
1 1
1 dy dy dy 1
Chain rule 1 = y dx ⇒ dx = y = exp(x). (Equivalently dx = dx/dy ).
ex+h − ex
h
d x x e −1
e = lim = e lim .
dx h→0 h h→0 h
eh −1
By Theorem 6, we have limh→0 h = 1.1
y = ax a > 0.
Can write this as y = exp(x log(a)) = ex log(a) . All the usual properties ax+y = ax ay
etc. all hold.
1
This suggests that limh→0 (1 + h)1/h = e. Prove by (1 + h)1/h = exp 1
h
log(1 + h) =
exp log(1+h)−log(1)
h
→ e.
5.3. FUNCTION ESTIMATES FOR SMALL AND LARGE ARGUMENTS 49
Theorem 7
d x
a = ax (log(a))
dx
Proof.
d x d
a = exp(x log(a)) = exp(x log(a)) · log(a) by the chain rule
dx dx
Corollary:
ah − 1
lim = log(a) for a > 0
h→0 h
Theorem 8
Let a be any number and let f (x) = xa for x > 0. Then f 0 (x) exists and
f 0 (x) = axa−1 .
Proof.
a
f (x) = xa = elog(x ) = ea log(x)
a
⇒ f 0 (x) = ea log(x) · by chain rule
x
= axa−1
Start with showing 2 < e < 4. We know that log(e) = 1 i.e. the area between x = 1
and x = e of y = 1/x, is 1.
50 CHAPTER 5. EXPONENTIALS AND LOGARITHMS
Since log(2) < 1, i.e. area A1 < 1, so we must have e > 2. Now
1
log(4) = 2 log(2) > 2 × = 1.
2
This implies e < 4 ⇒ 2 < e < 4. Accurate calculation later using Taylor’s theorem.
Theorem 9
(1+a)n
Let a be any positive number. Then n → ∞ as n → ∞. [Analogously,
n
limn→∞ (1+a) n = 0.]
n(n−1) 2
Proof. Write (1 + a)n = 1 + na + 2 a + b where b ≥ 0 is some number.
(1 + a)n 1 n−1 2 b
⇒ = +a+ a +
n n 2 n
b (1+a)n
n ≥ 0, and so for large n n becomes arbitrarily large.
en
Corollary: n → ∞ as n → ∞, since e = 1 + a for some a > 0.
Theorem 10
ex
The function f (x) = x is strictly increasing for x > 1 and limx→∞ f (x) = ∞.
exp beats x.
Proof.
xex − ex ex
f 0 (x) = = (x − 1) > 0 for x > 1
x2 x2
limn→∞ f (n) = ∞, hence result follows.
5.3. FUNCTION ESTIMATES FOR SMALL AND LARGE ARGUMENTS 51
Corollary 1.
The function x − log(x) becomes arbitrarily large as x becomes arbitrarily large.
x beats log.
Proof.
ex
log = x − log(x) > 0 for x large enough
x
Since log(t) becomes large for t large.
Corollary 2
x
The function log(x) becomes large as x becomes large. x beats log.
x y
Proof. Let y = log(x), then x = ey . So log(x) = ey . log(x) becomes large as x becomes
large, hence y also becomes large. By Theorem 10 the result follows.
Corollary 3
As x becomes large, x1/x approaches the limit 1.
Proof.
1/x log(x)
x1/x = elog(x ) = e x
log(x) 1
Now = → 0 as x becomes large by Corollary 2
x x/ log(x)
⇒ lim x1/x = 1
x→∞
Note Corollary 3 is used many times for integers. i.e. n1/n → 1 for n → ∞ being
integers.
Proof.
ex
f (x) = = ex−a log(x)
ea log(x)
a
f 0 (x) = ex−a log(x) 1 − > 0 if x > a
x
x
log(f (x)) = x − a log(x) = (log(x)) −a
log(x)
By Corollary 2, log(f (x)) → ∞ ⇒ f (x) → ∞.
52 CHAPTER 5. EXPONENTIALS AND LOGARITHMS
(2x+1)1/2
(ii) Differentiate y = (x2 +1)1/4
Theorem 11
If f , g, are differentiable on an open interval containing x0 , g(x0 ) = f (x0 ) = 0,
and g 0 (x0 ) 6= 0, then
f (x) f 0 (x0 )
lim = 0
x→x0 g(x) g (x0 )
Proof. Write
f (x)−f (x0 )
f (x) f (x) − f (x0 ) x−x0
= = g(x)−g(x0 )
g(x) g(x) − g(x0 )
x−x0
f (x)−f (x0 )
f (x) x−x0 f 0 (x0 )
⇒ lim = lim =
x→x0 g(x) x→x0 g(x)−g(x0 ) g 0 (x0 )
x−x0
Example 1
1 − cos(x) 0
lim of form
sin(x)
x→0 0
sin(x)
⇒ = lim =0
x→0 cos(x)
Example 2
0/0
z }| {
sin(x) − x cos(x) − 1
lim 3
= lim 2
= . . . carry on differentiating . . .
x→0 x x→0
| {z3x }
also 0/0
f 0 (x0 ) 0
Problem is that we need to know that if limx→x0 g 0 (x0 ) which is of the form 0 exists,
then it is equal to the limx→x0 fg(x)
(x)
. This is the useful form of L’Hôpital’s rule.
5.5. L’HÔPITAL’S RULE 53
Theorem 12
Let f (x) and g(x) be differentiable on an open interval containing x0 . Assume
that g(x) 6= 0 and g 0 (x) 6= 0 for x in an interval about x0 but with x 6= x0 . Assume
0 (x)
also that f , g are continuous at x0 with f (x0 ) = g(x0 ) = 0, and limx→x0 fg0 (x) = l.
Then also:
f (x)
lim =l
x→x0 g(x)
Example 1:
1 − cos(x) sin(x)
lim = lim if latter limit exists
x→0 x2 x→0 2x
Example 2:
x2 + 1
2x
lim 6= lim = 0.
x→0 x x→0 1
Not of form 0/0, i.e. not indeterminate.
Example 3:
sin(x)−x
Show that limx→0 tan(x)−x = − 21 .
f (x) ∞ 0
What about limx→x0 g(x) when of the form ∞? Would think that casting into 0 form
f (x) 1/g(x) 0
would help, but it doesn’t. i.e. g(x) = 1/f (x) i.e this is now of form 0. But
1/g −g 0 /g 2
lim = lim
x→x0 1/f |{z} x→x+0 −f 0 /f 2
L’Hôp
f (x) f 0 (x)
lim = lim 0
x→x0 g(x) x→x0 g (x)
Note that x0 may be replaced by ±∞ or x0 ±
Example 1
log(x) ∞
lim p
, p > 0, form
x→∞ x ∞
1/x
= lim = 0 since p > 0
x→∞ pxp−1
Example 2
Example 3
(a) Find limx→0+ xx
Of form 00 (indeterminate).
Solutions
(a) xx = ex log(x) . Have shown limx→0+ x log(x) = 0 and since exp is continuous,
limx→0+ exp(x log(x)) = exp limx→0+ (x log(x) = e0 = 1.
5.5. L’HÔPITAL’S RULE 55
Proof of L’Hôpital’s rule when limx→x0 f (x) = limx→x0 g(x) = 0 i.e. f (x0 ) = g(x0 ) =
0. The case x0 → ∞ is left as an exercise below. We need the following Theorem:
f (b) − f (a)
g 0 (c) = f 0 (c).
g(b) − g(a)
Proof. Let
f (b) − f (a)
h(x) = f (a) + (g(x) − g(a)) .
g(b) − g(a)
Then h(a) = f (a) and h(b) = f (b). For the function φ(x) = h(x) − f (x) we have
φ(a) = φ(b) = 0 and φ is differentiable on (a, b). Hence by the MVT, there exists a
c ∈ (a, b) such that φ0 (c) = 0. i.e. h0 (c) = f 0 (c), and the theorem is proved.
Proof. L’Hôpital’s Theorem
Here we prove the 0/0 version, i.e. f (x0 ) = g(x0 ) = 0.
f (x) f (x) − f (x0 ) f 0 (c)
= = 0
g(x) g(x) − g(x0 ) g (c)
where c is a number between x and x0 . As x → x0 , c → x0 also. By hypothesis
0
f 0 (c)
f (x)
lim = ` ⇒ lim =`
x→x0 g 0 (x) x→x0 g 0 (c)
f (x)
⇒ lim = ` as required
x→x0 g(x)
56 CHAPTER 5. EXPONENTIALS AND LOGARITHMS
Exercise: Prove L’Hôpital’s rule when x0 → ∞, i.e. show that if limx→∞ f (x) =
limx→∞ g(x) = 0, then limx→∞ [f (x)/g(x)] = limx→∞ [f 0 (x)/g 0 (x)] = `, say.
Proof. Here are some steps to follow:
(i) First show using Cauchy’s Mean Value Theorem that given ε > 0 there exists M
such that for all y > x > M
f (y) − f (x)
− ` < ε.
g(y) − g(x)
Integration
57
Chapter 6
Given f (x) defined over some interval, then if we can find a function F (x) defined
over the same interval such that
F 0 (x) = f (x),
R
then F (x) is the indefinite integral of f ⇒ F = f (x)dx. This is not unique. Let G
be another indefinite integral, i.e. G0 (x) = f (x). Then
d
(F − G) = 0 ⇒ F (x) = G(x) + constant.
dx
By definition, F (a) = 0.
59
60 CHAPTER 6. ANTI-DERIVATIVES AND GEOMETRICAL INTERPRETATION
Theorem 1
The function F (x) is Rdifferentiable and its derivative is equal to f (x). Another
d x
way to state this is dx a f (t)dt = f (x).
F (x+h)−F (x)
Proof. Start with the Newton quotient h .
Suppose x 6= b and also h > 0. F (x + h) − F (x) is the area under the graph between
x and x + h.
Figure 6.1:
Since f (x) is continuous on [x, x + h] and is defined there, it must have a maximum
at some point x+ , and minimum at some point x− . Hence, for all t ∈ [x, x + h]
Can also bound the area using the rectangles shown in Figure 6.1.
The constant is fixed by F (a) = 0. In other words, if I can guess a function G(x)
whose derivative is f (x) (e.g. guess log(x) for the anti-derivative of x1 ), then since F
and G differ by a constant I have
F (x) = G(x) + K.
Example 1
2 2
x3
Z
2 8 1
x dx = = − .
1 3 1 3 3
x3
Here f (x) = x2 , G(x) = 3 is the guessed anti-derivative.
If f (x) < 0 then the area is below the x-axis. Define F (x) to be minus the area.
(All very familiar). This leads to the definite integral.
Figure 6.2:
Z b
f (x)dx = F (b) − F (a) All negative areas are accounted for.
a
62 CHAPTER 6. ANTI-DERIVATIVES AND GEOMETRICAL INTERPRETATION
b−a
xi = a + ih i = 0, 1, . . . , n, h=
n
Note: The selected partition has regular spacing. Can generalise this to have a partition
defined by a sequence {xk }k=0, ... ,n and in the limit we require max |xk −xk−1 | → 0. We
k
will avoid this technical issue; it can be dealt with in a direct way and is not necessary
for what we want to do.
Take any sub-interval [xi−1 , xi ] and let x∗i ∈ [xi−1 , xi ]. Then the Riemann sum is
defined to be
Xn
RS ∗ := f (x∗i )h.
i=1
65
66 CHAPTER 7. THE RIEMANN SUM
R1
Figure 7.3: Riemann sum for 0 f (x)dx.
67
1 i
a = 0, b = 1 ⇒ h= , xi = ih =
n n
n n n
X 1 X i 1 1 X
Upper RS = f (xi ) = · = 2 i
n n n n
i=1 i=1 i=1
1 n(n + 1) 1 1 1
= 2· = + → as n → ∞
n 2 2 2n 2
1 2 3 ... n−1 n
+ n n−1 n−2 ... 2 1
Add 2S = (1 + n) + (1 + n) + · · · + (1 + n) = n(1 + n)
| {z }
n times
1
S = n(n + 1).
2
Example 2 Z 1
ex dx
0
n
X 1
Upper Riemann Sum Un = ei/n
n
i=1
Z 1 n
1 X 1/n i
ex dx = lim e =e−1
0 n→∞ n
i=1
| {z }
geometric series
68 CHAPTER 7. THE RIEMANN SUM
The last part is left as problem 2 on Problem Sheet 4. It uses the result for a geometric
series which easily follows from the algebraic identity
1 − xn
1 + x + x2 + . . . + xn−1 = .
1−x
n n
X X xi−1 + xi
Un = f (xi )h Mn = f h
2
i=1 i=1
Rb R1
We know that limn→∞ Un = limn→∞ Mn = a f (x)dx. Example for 0 ex dx = e − 1 ≈
1.71828183 := I correct to 8 decimal places.
n h Un |I − Un | Mn I − Mn
1 1 2.7183 1.0000 1.6487 0.0696
2 0.5 2.1835 0.4652 1.7005 0.0178
4 0.25 1.9420 0.2237 1.7138 0.0045
Question: Can you think of a better way still? We have to go beyond the Riemann
sum definition - Numerical Analysis Answer: Better approximation of the function, e.g.
instead of linear segments use three points and fit parabolas; other ways are cubic splines,
etc.
We will be able to calculate the accuracy of such schemes using Taylor’s Theorem - see
later.
70 CHAPTER 7. THE RIEMANN SUM
Chapter 8
Rb Rb Rb Rb
Hence a f (x)dx ≤ a |f (x)|dx and a f (x)dx ≤ a |f (x)|dx
Rb Ra
5) a f (x)dx = − b f (x)dx.
Proofs follow easily from RS definitions and the use of signed areas.
Theorem 1
Suppose g(x) is defined for all x ∈ [a, b] and is differentiable on [a, b]. Then
Z b
g 0 (x)dx = g(b) − g(a)
a
Proof. (Sketch)
71
72CHAPTER 8. PROPERTIES OF THE DEFINITE INTEGRAL; FUNDAMENTAL THEOREM OF
b−a
Let xi = a + ih, h = n .
n n
X X g(xi + h) − g(xi )
Upper RS g 0 (xi )h ≈ ·S
h = g(b + h) − g(a + h)
h{z
S
i=1 i=1 | }
telescoping series
As h → 0 result follows.
Useful Theorem:
Z g(x)
d
f (t)dt = f (g(x)) · g 0 (x).
dx a
Rx
Proof. Let F (x) = a f (t)dt. Then F 0 (x) = f (x) - already proved.
R g(x)
Now a f (t)dt = F (g(x)) by the definition of F .
Z g(x) !
d d
⇒ f (t)dt = F (g(x) = F 0 (g(x)) · g 0 (x)
dx a dx
= f (g(x)) · g 0 (x).
Example
!
Z x2
d t 2
e dt = ex · 2x
dx a
Z x2
x2 2
or et dt = et a
= ex − ea same as before
a
Chapter 9
Improper Integrals
Definition
Rb
a f (x)dx is an improper integral if
To find improper integrals we take the limit of proper integrals. If the limit is finite,
the integral converges, otherwise it diverges.
Example 1
Z ∞ Z b
1 1 1
dx = lim dx = lim − + 1 = 1
1 x2 b→∞ 1 x2 b→∞ b
Example 2 Z ∞
dx
= lim log(b) = ∞ i.e. diverges.
1 x b→∞
Geometrically:
In general
b
br+1 − 1
Z
lim xr dx = lim r 6= −1
b→∞ 1 b→∞ r+1
So need r +1 < 0 for convergence, i.e. r < −1, (r = −1 is divergent - see log example
earlier).
73
74 CHAPTER 9. IMPROPER INTEGRALS
Comparison test is useful if we cannot carry out the integral exactly. It will tell us
if it exists, then we can find it numerically etc. e.g
Z ∞
sin(x)
converges
0 (1 + x)2
75
R∞ dx
R∞ dx
First thing to note is that 0 (1+x)2
converges by comparison to 1 x2
. Why? If
1 1
x ≥ 1, (1+x)2
< x2
. By the comparison test
∞
| sin(x)|
Z
sin(x) 1 1
dx converges since ≤ < 2 for x ≥ 1 (9.1)
0 (1 + x)2 (1 + x)2 (1 + x)2 x
e.g.
Z ∞
dx
√ is divergent
1 1 + x2
Z b Z b Z b
dx dx dx
√ ≥ √ = √
1+x 2 2
x +x 2 2x
1
Z ∞1 1
Z ∞
dx dx
Now diverges ⇒ so does √
1 x 1 1 + x2
R∞ dx
e.g. 1
√
x
diverges. (Already saw this, and we can do it directly).
Example 1
Z 1
log(x)dx exists
0
Z 1 Z 1
1 1
= lim log(x)dx = lim [x log(x)] − x dx
→0 →0 x
= lim [− log() − 1 + ] = −1
→0
Example 2
R∞ −x
e√
Show that the improper integral I = 0 x
dx converges.
76 CHAPTER 9. IMPROPER INTEGRALS
∞ −x
1
e−x
Z Z
e
Write I = I1 + I2 where I1 = I2 = √ dx,√ dx
0 1x x
Z 1 −x Z 1
e 1
I1 = √ < √ dx which is convergent
0 x 0 x
Z ∞ −x Z ∞
e
√ dx < e−x dx which is also convergent
1 x 1
Example 3 √
Find the length of the curve y = 1 − x2 for x ∈ [−1, 1].
Z 1 Z 1
0 2 1/2 dx
Length L = (1 + (y ) ) dx = √
improper at both ends
−1 −1 1 − x2
Z 0 Z 0
dx dx π π
= lim sin−1 (0) − sin−1 (p) = 0 − −
√ = lim √ =
−1 1 − x2 p→−1 p 1 − x2 p→−1 2 2
Z 1 Z 1
dx π
lim sin−1 (1) − sin −1(p) =
√ = lim ⇒ π is the length
0 1 − x2 p→1 0 p→0 2
Chapter 10
Given a function f that is integrable on [a, b], we define its average hf i[a,b] by the formula.
Z b
1
hf (x)i[a,b] = f (x)dx.
b−a a
Z b Z b
f (x)dx = hf i[a,b] dx
a a
Geometrically, the area of the shaded rectangle is equal to the area under y = f (x).
77
78 CHAPTER 10. MEAN VALUE THEOREM FOR INTEGRALS
Theorem 1 Let f be continuous on [a, b]. Then there exists a point x0 ∈ (a, b)
such that Z b
1
f (x0 ) = f (x)dx.
b−a a
Rx
Proof. Define F (x) = a f (t)dt. By the Fundamental Theorem of Calculus we have
F 0 (x) = f (x) for all x ∈ (a, b). F is continuous at a and b. By MVT we have
F (b) − F (a)
F 0 (x0 ) =
b−a
i.e. Rb Ra Z b
a f (t)dt − a f (t)dt 1
f (x0 ) = = f (t)dt
b−a b−a a
Chapter 11
Techniques of Integration
For:
Z Z
n=1 substitute u = sin(x) ⇒ sin (x) cos(x)dx = um du
m
Z
cos(x)
n = 1, m = −1 dx = log | sin(x)| + c
sin(x)
m, n 6= 1 Write in terms of sinp (x) cos(x) or similar, or use double angle formulae
79
80 CHAPTER 11. TECHNIQUES OF INTEGRATION
Example 2
sin2 (2x) 1 1 − cos(4x)
Z Z Z
2 2
sin (x) cos (x) = = dx
4 4 2
Example 3 Z
I= tan3 (θ) sec3 (θ)dθ
Example 4 Z Z
1
cos(3x) cos(5x)dx = (cos(8x) + cos(2x))dx etc
2
Trigonometric substitutions
√ √
(1) If a2 − x2 appears in an integral, try x = a sin(θ), dx = a cos(θ)dθ, a2 − x2 =
a cos(θ) (a > 0, θ acute).
√ √
(2) If x2 − a2 occurs, try x = a sec(θ), dx = a tan(θ) sec(θ)dθ and x2 − a2 =
a tan(θ).
√ √
(3) If a2 + x2 or a2 + x2 occur, try x = a tan(θ), 2 2 2
√ dx = a sec (θ)dθ, a + x =
2 2
a sec(θ). Also x = a sinh(θ), dx = a cosh(θ)dθ, a + x = a cosh(θ).
81
Example 5
x2
Z
dx x = tan(θ) dx = sec2 (θ)dθ
(1 + x2 )3/2
tan2 (θ) sec2 (θ)
Z Z
⇒ dθ = tan(θ) sin(θ)dθ
sec3 (θ)
Better by parts:
Z Z
x 2 −1/2 1
2 3/2
· xdx = −(1 + x ) ·x+ √ dx
(1 + x ) 1 + x2
x
= −√ + sinh−1 (x) + c
1 + x2
R π/2 8
Example 6 Show that 0 sin5 (x) = 15
Applications of Integration:
Lengths, surfaces and volumes of
revolution
n p
X
Total length ≈ (xi − xi−1 )2 + (f (xi ) − f (xi−1 ))2
i=1
s
n 2
X f (xi ) − f (xi−1 )
= (xi − xi−1 ) 1 + .
xi − xi−1
i=1
83
84CHAPTER 12. APPLICATIONS OF INTEGRATION: LENGTHS, SURFACES AND VOLUMES O
b−a
Now let xi − xi−1 = h = n := ∆x.
s 2
X f (xi ) − f (xi−1 )
Total length = lim ∆x 1+
n→∞, (h→0, ∆x→0) xi − xi−1
Z b 1/2
L= 1 + (f 0 (x))2 dx
a
Z t1
" 2 2 #1/2
dx dy
L= + dt
t0 dt dt
√
Clearly L > 2 = 1.4142 . . . , y = f (x) = x2 , f 0 (x) = 2x.
Z 1
L= (1 + 4x2 )1/2 dx
0
(
x = 0, θ = 0
Substitution 2x = tan(θ)
x = 1, θ = a tan(2)
2dx = sec2 (θ)dθ see Fig 12.1 below
Z a tan(2)
1
L= sec3 (θ)dθ
0 2
12.1. LENGTH OF CURVES 85
Figure 12.1:
Method 1
Z Z Z
3 cos(θ) cos(θ)
sec (θ)dθ = dθ = dθ
4
cos (θ) (1 − sin2 (θ))2
Z
du
u = sin(θ) = partial fractions
(1 − u2 )2
A B C D
= + 2
+ +
1 + u (1 + u) (1 − u) (1 − u2 )
= etc . . . etc
or write in terms of x again.
Method 2
Z Z
sec3 (θ)dθ = sec(θ) sec2 (θ)dθ integrate by parts
d
(sec(θ))
Z z dθ }| {
= tan(θ) sec(θ) − tan(θ) sec(θ) tan(θ) dθ
Z Z
3 2
= tan(θ) sec(θ) − sec (θ)(1 − cos (θ))dθ move − sec3 (θ)dθ to other side
Z Z
3
2 sec (θ)dθ = tan(θ) sec(θ) + sec(θ)dθ)
Z
1
i.e. sec3 (θ)dθ = [tan(θ) sec(θ) + log(sec(θ) + tan(θ))] (see HW3)
2
Z 1
1 a tan(2)
(1 + 4x2 )1/2 dx = [tan(θ) sec(θ) + log(sec(θ) + tan(θ))]0
0 4
1h p p i1 1 h √ √ i
= 2 × 1 + 4x2 + log 2x + 1 + 4x2 = 2 5 + log 2 + 5 = 1.4789
4 0 4
Method 3 - the easiest one! Rθ
Let 2x = sinh θ so that 2dx = cosh θ dθ and L = 0 1 (1/2) cosh2 θdθ, and sinh θ1 = 2.
Now use cosh(2θ) = cosh2 θ + sinh2 θ = 2 cosh2 θ − 1 to find
Z θ1
1 + cosh 2θ θ1 1
L= dθ = + sinh 2θ1
0 4 4 8
86CHAPTER 12. APPLICATIONS OF INTEGRATION: LENGTHS, SURFACES AND VOLUMES O
A plane cuts a solid V - cross sectional area is Rx say. Then the volume of a slice is
Rx dx. So if Px is a family of parallel planes with common axis x, and the area of V cut
by Px is A(x), then the volume of V is
Z b
A(x)dx
a
OA
=r
∆OAB OB =x
AB = (r2 − x2 )1/2
Z r
4
So A(x) =π(r2 − x2 ), V = π(r2 − x2 )dx = πr3
−r 3
Example 2 Volume of conical solids, with circular base of radius r and height h.
DE AE AF DE h−x
= = → i.e. → =
OB AB AC r h
h−x
⇒ DE = r
h
r2
A(x) = π 2 (h − x)2
h
Z h 2
r
Volume = π 2 (h − x)2 dx
0 h
h
r2 (h − x)3 1
= −π = πr2 h
h2 3 0 3
Example 3 A sphere of radius r is cut into 3 pieces with the two cuts symmetrically
placed about the centre. Where should the cuts be in order to get three equal volumes?
Complete for homework.
The volume of the solid produced by revolving y = f about the x-axis (as shown) is
given by
Z b
V = π(f (x))2 dx.
a
Example 4 The region between the graphs of sin(x) and x for [0, π2 ], is revolved about
the x-axis. Sketch the resulting solid and find its volume.
12.2. VOLUMES AND VOLUMES OF REVOLUTION 89
π/2
π4 π2
Z
V = π(x2 − sin2 (x))dx = − show this
0 24 4
If we revolve about the y-axis, what is the volume? Consider a non-negative function
f (x) on [a, b].
(b)
z}|{
|{z} (x) |{z}
2πx f dx
(a) (c)
Z b
⇒ V = 2πxf (x)dx.
a
Where (a) - circumference of cylindrical shell, (b) - radius of shell, (c) - thickness of
shell. Note that this can be done by the slice method but with planes along the y-axis
and parallel to the x-axis.
90CHAPTER 12. APPLICATIONS OF INTEGRATION: LENGTHS, SURFACES AND VOLUMES O
As we revolve about the x-axis, the area of the surface area swept out is a strip of length
≈ 2πf (xi ) and thickness ∆li . Now
1/2
∆li = (xi − xi−1 )2 + (f (xi ) − f (xi−1 ))2
1/2
≈ 1 + (f 0 (xi ))2
∆x as seen earlier.
x
f+0 = − √ = −f−0
−r2 − x2
r2
⇒ 1 + f+02 = 1 + f−02 =
(r2 − x2 )
Z r h p p i r
S = 2π a + r 2 − x2 + a − r 2 − x2 · dx
r (r2 − x2 )1/2
Z r
dx
= 4πar √
−r r2− x2
Rr
Put x = r sin θ and show that √ dx = π.
−r r2 −x2
1D case - straightforward.
If centre of mass is at x = x̄, then we must have a zero total moment. i.e.
Pn
k=1 mk xk
X
mk (x̄ − xk ) = 0 i.e. x̄ = P n
k=1 mk
n masses of mass mk and coordinates (xk , yk ). Find the center of mass, assume it
is (x̄, ȳ). There are two degrees of freedom, so without loss of generality we need to
have zero moments about the x-axis and the y-axis. What do I mean by this? Here is a
schematic.
Note: If the masses are placed symmetrically and are equal, centre of mass is on
the line of symmetry. e.g.
94CHAPTER 12. APPLICATIONS OF INTEGRATION: LENGTHS, SURFACES AND VOLUMES O
Now consider a continuous mass distribution, i.e. a plate of a certain spatial density.
You will see how to work with double integrals in term 2 and further. For the
moment, we will consider the centre of mass of regions bounded by one or more graphs
y = f (x).
Case 1 Region {(x, y) : a ≤ x ≤ b, 0 ≤ y ≤ f (x)}
Consider a partition of [a, b] as shown. For rectangle Ri , the centre of mass is (by
symmetry):
1 1
x∗i = (xi−1 + xi ) yi∗ = f (x∗i )
2 2
96CHAPTER 12. APPLICATIONS OF INTEGRATION: LENGTHS, SURFACES AND VOLUMES O
Z b
⇒ My = lim ρx∗i f (x∗i )∆x
ρxf (x)dx
=
n→∞ a
1 b
Z
1
Mx = lim ρ f (x∗i )2 ∆x = ρ(f (x))2 dx.
n→∞ 2 2 a
Rb
Now for a balance of moments, if the total mass of R is m (note m = a ρf (x)dx).
Then
Rb 1 b
R 2
a xf (x)dx 2 a (f (x)) dx
x̄ = R b ȳ = R b .
a f (x)dx a f (x)dx
By symmetry x̄ = 0.
1
R1
2 −1 (1 − x2 )dx 1 2 4
ȳ = = (2 − ) = ≈ 0.424
π/2 π 3 3π
1
(Note: If we found it to be > 2 then we know it’s wrong! Why?)
12.4. CENTRES OF MASS 97
The centre of mass of “rectangle” Ri is (x∗i , 21 (f (x∗i ) + g(x∗i ))) as before. The area of
the rectangle is approximately equal to (f (x∗i ) − g(x∗i ))∆x. So as before we find
Rb 1 b
R 2 2
a x(f (x) − g(x))dx 2 a (f (x) − g(x) )dx
x̄ = R b ȳ R b
a (f (x) − g(x))dx a (f (x) − g(x))dx
Example 8 Region between y = x and y = x2 , 0 ≤ x ≤ 1. Find centre of mass
R1
x(x − x2 )dx 1/12 1
x̄ = R0 1 = =
2 1/6 2
0 (x − x )dx
1 1 2
(x − x4 )dx
R
1/15 2
ȳ = 2 0 = =
1/6 1/6 5
98CHAPTER 12. APPLICATIONS OF INTEGRATION: LENGTHS, SURFACES AND VOLUMES O
)
1 f ( 12 ) =1 1− 2 3
5 = 5
At x= , ⇒ closer to top curve
2 g( 12 ) = 1
4
2 1 3
5 − 4 = 20
Note: If f (x) = xm , g(x) = xn , for some m, n, the centre of mass could be outside the
region. (This is ok.)
Theorem of Pappus
Let R be a region that lies on one side of a line l.
A = area of R
V = Volume obtained by rotating about l
d = distance travelled by the centre of mass when R is rotated about l
Then V = Ad
1
KE = m(yω)2 .
2
The coefficient of 12 ω 2 is defined to be the moment of inertia. Hence for the single
particle considered here we define the moment of inertia I to be
I = my 2 .
The usefulness of (12.1) is that under rigid body rotation (i.e. all the masses rotate at
the same angular velocity), the kinetic energy is
1
KE = ω 2 I.
2
The moment of inertia only has information about the masses, not their velocity.
Consider next the curve y = f (x) in the x − y plane in the interval x0 ≤ x ≤ x1 , e.g.
a wire of the given shape. Let the density of the curve (wire) be ρ(x) per unit length.
Then we can write down the moment of inertia Ix about the x−axis, and Iy about the
y−axis by using (12.1) and Riemann sums based on a partition of the curve into element
pieces. We have
Z x1 p Z x1 p
2
Ix = 02
ρ(x)y 1 + y dx = ρ(x)f (x)2 1 + f 02 dx (12.2)
x0 x0
Z x1 p Z x1 p
2
Iy = 02
ρ(x)x 1 + y dx = ρ(x)x2 1 + f 02 dx. (12.3)
x0 x0
Z b " 2 2 #1/2
dx dy
L= + dt for parametric curved (x(t), y(t)).
a dt dt
Z β 0 1/2
L= (f cos(θ) − f sin(θ))2 + (f 0 sin(θ) + f cos(θ))2 dθ
θ=α
Zβp
= (f 0 (θ))2 + (f (θ))2 dθ
α
" 2 #1/2
Z β
dr
= + r2 dθ
α dθ
Z q 2π Z 2π p
2 2
L= (1 + cos(θ)) + sin )θ)dθ = 2 + 2 cos(θ)dθ
0 0
2 θ 2 θ
Now cos(θ) = 2 cos − 1 ⇒ (1 + cos(θ)) = 2 cos
2 2
p θ
⇒ 2(1 + cos(θ) = 2 cos
2
102CHAPTER 12. APPLICATIONS OF INTEGRATION: LENGTHS, SURFACES AND VOLUMES O
We need to do this because 1 + cos(θ) can be positive but cos 2θ is negative, e.g. for
π ≤ θ ≤ 2π. So need to write this integral as
Z π π Z 2π
θ
L= 2 cos dθ + −2 cos dθ = 8
0 2 π 2
Z 2π
θ
Otherwise 2 cos dθ = 0 which is absurd!
0 2
Area in polar coordinates in a region inside the graph of f (θ) on [α, β].
Note:
Figure 12.2:
π π π π
r ≥ 0 ⇒ − ≤ 2θ ≤ − ≤θ≤
2 2 4 4
Z π Z π
1 4 1 4 1 + cos(4θ)
A= cos2 (2θ)dθ = dθ
2 − π 2 −π 2
4 4
1π π
= =
42 8
Part III
105
Chapter 13
Series
Definition
Given a sequence {an }n≥1 of real numbers, define the sequence of partial sums
by
N
X
SN = a1 + a2 + · · · + aN = an .
n=1
Example 1
The geometric series ∞ n
P
n=0 x . (x 6= 1)
SN = 1+(x + · · · + xN )
xSN = (x + · · · + xN ) + xN +1
1 − xN +1
Subtract SN =
1−x
∞
1 X
If |x| < 1, lim SN = = xn ,
N →∞ 1−x
n=1
hence the series converges. If x ≥ 1, the series diverges.
Example 2
∞ N N
X 1 X 1 X 1 1
SN = = −
n(n + 1) n(n + 1) n n+1
n=1 n=1 n=1
1
Telescoping series ⇒ SN = 1 − N +1 → 1 as N → ∞. Series converges to 1.
107
108 CHAPTER 13. SERIES
Theorem P 1
The series ∞n=1
1
n diverges to +∞.
Proof. It is enough to prove that the partial sums are not bounded above.
Consider
1 1 1
S2K = 1 + + + · · · + K
2 3 2
1 1 1 1 1 1 1 1 1
=1+ + + + + + + + ··· + + ··· + K
2 3 4 5 6 7 8 2K−1 + 1 2
1 1 1 1 1 1 1 1 1
≥1+ + + + + + + + ··· + + ··· + K
2 4 4 8 8 8 8 2K 2
1 2 4 2K−1 1
= 1 + + + + ··· + K = 1 + K
2 4 8 2 2
Partial sums unbounded for K large ⇒ series diverges.
Theorem 2
If α > 1 is a rational number, then
∞
X 1
converges.
nα
n=1
Proof. Partial sums are increasing, so enough to prove that they are bounded above.
Compare SN ≤ S2N −1 , note N ≤ 2N − 1.
1 1 1
SN ≤ S2N −1 = 1 + α
+ α + ··· + N
2 3 (2 − 1)α
1 1 1 1 1 1 1 1
=1+ + + + + + + ··· + + ··· + N
2α 3α 4α 5α 6α 7α 2(N −1)α (2 − 1)α
2 4 2N −1
≤ 1 + α + α + · · · + (N −1)α
2 4 2
2 N −1
1 1 1
= 1 + α−1 + + ··· +
2 2α−1 2α−1
1
N
1 − 2α−1 1 1
= 1
≤ 1 if α−1
<1
1 − 2α−1 1 − 2α−1 2
13.2. CAUCHY SEQUENCES AND CONVERGENCE OF SERIES 109
P∞ PN
Proof. Let the sum be S, i.e.
PN −1 1 an = S. Then SN = n=1 an → S as N → ∞, and
SN −1 = n=1 an → S as N → ∞. Now aN = SN − SN −1 → S − S = 0 as N → ∞.
Example 3
∞
X
(−1)n = −1 + 1 − 1 + 1 . . . diverges by the theorem above, an 6→ 0 as n → ∞.
n=1
P∞ 1
Note: Theorem 3 provides a necessary but not sufficient condition, e.g. n=1 n has
an → 0 but diverges.
|Sm − Sn | <
Theorem 4
Every Cauchy sequence converges. (Proof by use of the Bolzano-Weierstrass the-
orem seen in Analysis).
Proof. We will show that the sequence Sk of partial sums is a Cauchy sequence, i.e.
given any > 0 we need to find N such that for all n > m > N , |Sn − Sm | < .
Consider any n > m. Then since an is decreasing
Since an → 0 as n → ∞, given , I can find N such that for any n > N , an < .
Now for any n > m > N .
Example 4
∞
X (−1)n−1 1
converges since |an | = → 0 and it is an alternating series.
n n
n=1
∞
X (−1)n−1 1 1 1
In fact, = 1 − + − + · · · = log(2) (we will see this later)
n 2 3 4
n=1
Theorem
P∞ 6 (Comparison test)
P∞ n=1 bn be convergent with bn non-negative. If |an | ≤ bn (n = 1, 2, . . . ), then
Let
n=1 an converges.
13.3. CONVERGENCE TESTS 111
Pk P
Proof. Let sk = j=1 aj , i.e. k-partial sum of an . For n > m we have
P∞
(More precisely, given > 0, there is N such that i=m+1 bi < for all m > N ).
converges.
(−1)n P∞ 1
With an = n3n , compare with n=1 bn with bn = 3n .
1 1
n3n+1 > 3n ⇒ |an | = n
< n
n3 3
P∞
By comparison test n=1 an converges.
Example 6 Prove that if α is any positive number and |x| < 1, then the series
∞
X
nα xn converges.
n=1
and since log |x| < 0, the exponential decay term dominates over any power of n.
Hence nα+2 xn → 0 as n → ∞ and the sequence {nα+2 xn } is bounded. Hence there
exists a constant C such that
C
|nα+2 xn | ≤ C i.e. |nα xn | ≤ n≥1
n2
But ∞ 1
P P α n
n=1 n2 converges, hence so does n x by the comparison test. Note: it is
much easier to use the Ratio Test (below).
112 CHAPTER 13. SERIES
Theorem 7
Every absolutely convergent series is convergent.
is conditionally convergent.
Example 8 Discuss the convergence of
∞ √
X (−1)n n
.
n+4
n=1
√
n 1 √
Series is not absolutely convergent since n+4 = √
n+4/ n
. Now for n ≥ 1,
1 1
√ √ ≥ √
n + 4/ n 5 n
√1
P
and since n
diverges, we are done.
√
n
For the alternating series test to apply we need to show that n+4 is decreasing.
√
x 4−x
If f (x) = ⇒ f 0 (x) = √ < 0 for x > 4.
x+4 2 x(x + 4)2
So series terms decrease for n > 4. We only care about what happens beyond the 1st
three terms - all the action is in the tail. Hence the series converges by the alternating
series test, but it is not absolutely convergent.
Theorem 8
Let f (x) be a function which is defined for all x ≥ 1, and is positive and decreasing.
Then the series
∞
X
f (n)
n=1
R∞
converges if and only if the improper integral 1 f (x)dx converges.
Z 2 Z 3
f (1) ≥ f (x)dx, f (2) ≥
f (x)dx etc
1 2
Z n
⇒f (1) + f (2) + · · · + f (n − 1) ≥ f (x)dx
1
P∞So if the partial sums are bounded by L say, (we know this is true since by assumption
n=1 f (n) converges) we have
Z n
f (x)dx ≤ L (*)
1
R∞
Claim that this implies that 1 f (x)dx exists. Give me any number b, however large
you wish. Then I can find an integer n > b so that
Z b Z n
f (x)dx ≤ f (x)dx ≤ L by (*)
1 1
Rb
Hence 1 f (x)dx is bounded above for all b. Now send b to infinity.
114 CHAPTER 13. SERIES
1
Solution Take f (x) = x in the integral test. Hence
Z n+1
1 1 dx
1 + + ··· + ≥ = log(n + 1)
2 n 1 x
Z n+1 ∞
dx X 1
⇒ lim diverges, and by the integral test, so does .
n→∞ 1 x n
1
converge/diverge?
1
Solution Let f (x) = xp and consider
n
n1−p
Z
dx 1
p
= −
1 x 1−p 1−p
Z n
dx
Hence lim exists if p > 1 and diverges if p ≤ 1.
n→∞ 1 xp
∞
X 1
Hence converges for p > 1, and diverges otherwise.
np
n=1
P∞ P∞
Example 11 Show that n=2 n
√1 diverges but 1
n=2 n(log(n))2 converges.
log(n)
Solution
Use the integral test by considering
Z ∞ Z b Z b
dx dx 1
p = lim p = lim (log(x))−1/2 dx
2 x log(x) b→∞ 2 x log(x) b→∞ 2 x
Z
d 1
(of the form f 0 (g(x))g 0 (x)dx since (log(x)) = )
dx x
h ib h p i
= lim 2(log(x))1/2 = lim 2(log(b))1/2 − 2 log(2) = ∞
b→∞ 2 b→∞
13.3. CONVERGENCE TESTS 115
Theorem 9
Let ∞
P
n=1 n be a series satisfying
a
an+1
lim = L.
n→∞ an
Then:
xn an+1 |x|n+1 n!
an = , lim = lim
n! n→∞ an n→∞ (n + 1)! |x|n
|x|
= lim =0
n→∞ n + 1
P∞ xn
In fact we will see that n=0 n! = ex .
116 CHAPTER 13. SERIES
|aN +K | < (L + )|aN +K−1 | < (L + )2 |aN +K−2 | < · · · < (L + )K |aN |.
∞
X ∞
X
Hence, |aN +j | < |aN | (L + )j
j=1 j=1
P∞
which is bounded since the series j=1 (L + ) is a geometric series ( (L + ) < 1. Hence
(We only added a finite number of terms). This proves absolute convergence in case
(1) L < 1.
In case (2), L > 1, pick L + > 1 now and we have |aN +K | > (L + )N |aN | which
diverges now as a geometric series. To prove that if L = 1 the test is inconclusive, it is
sufficient to pick an example, i.e.
∞ p p
X 1 an+1 n 1
, = =
np an n+1 1 + 1/n
n=1
p
1
Since p is fixed, we have limn→∞ 1+1/n = 1p = 1. But if p > 1 we have convergence
but if p ≤ 1 divergence. Hence, test is inconclusive.
Using this proof, we have a practical way of estimating errors in truncating series.
Suppose
an
< r < 1 for n > N.
an−1
Then
∞
X N
X ∞
X
an − an = an
n=1 n=1 n=N +1
is the error made in approximating the infinite series by a N −terms finite series. But
|aN +1 | < r|aN | and generally |aN +K | < |aN |rK . So
∞ ∞ ∞
X X X r
|an | = |aN +K | ≤ |aN | rk = |aN | .
1−r
n=N +1 k=1 k=1
r
So the error is ≤ |aN | 1−r .
13.3. CONVERGENCE TESTS 117
an
Solution: We have an−1 = n1 .
an
In the example, the truncation is N = 4, so if n > 4, an−1 < 15 .
1/5 1 1 1
The error is ≤ |a4 | · 1−1/5 = 4! · 4 = 96 < 0.0105
Theorem 10
For the given series ∞ 1/n = L. Then
P
n=1 an , suppose that limn→∞ |an |
Examples:
1.
∞
X 1 1
converges, |an |1/n = → 0.
nn n
n=1
2.
∞
X 3n
diverges. (We already know this by other methods.)
n2
n=1
3n
i.e. an = 6→ 0 as n→∞
n2
3
|an |1/n = →3 as n → ∞.
(n1/n )2
3.
∞
X nn
(again we have seen this before in HW3)
n!
n=1
(n + 1)n+1 n! 1 n
an+1
Use ratio test = = 1+
an (n + 1)! nn n
an+1
lim = e > 1 diverges.
n→∞ an
4.
∞
1
X
.
− log(n) n2
n=1
X 1
Intuition: large n series ≈ <∞
n2
1 1
2
can be bounded above by with 0 < α < 1
n − log(n) αn2
n2 − log n = αn2 + (1 − α)n2 − log n > αn2
since 0 < α < 1 and (1 − α)x2 − log x ≥ (1 − α) > 0 when x ≥ 1
∞ ∞
X 1 X 1
⇒ 2
< <∞
n − log n αn2
1 1
13.3. CONVERGENCE TESTS 119
P∞
13.3.5 Testing convergence for n=1 an
120 CHAPTER 13. SERIES
Chapter 14
Power Series
is called a power series in x. The series (14.1) can be generalised to a power series in
x − x0 by
X∞
an (x − x0 )n = a0 + a1 (x − x0 ) + a2 (x − x0 )2 + · · · , (14.2)
n=0
and it is clear that (14.2) can be transformed into (14.1) by replacing x − x0 in (14.2) by
x to give (14.1). Unless stated otherwise we will be considering power series in x, and
it is understood that there is no loss of generality in doing so. We have the following
formal definition.
Definition
and {an }n≥0 be a
Let x be a real number (can extend to complex numbers also) P
sequence of numbers. Then we can form the power series ∞ n
n=0 an x . The
PN n
partial sums sN = n=1 an x are degree N polynomials.
N
X
lim an xn ,
N →∞
n=0
exists with value equal to the sum of the infinite series. Obviously (14.1) converges at
x = 0. The question is for what values of x does a power series exist? Here are three
canonical examples of power series with different behaviour:
P∞ 2 2 3
1. n=0 n! x = 1 + x + 2! x + 3! x + · · ·
121
122 CHAPTER 14. POWER SERIES
P∞ xn x2 x3
2. n=0 n! =1+x+ 2! + 3! + ···
P∞ n
3. n=0 x = 1 + x + x2 + x3 + · · ·
Considering the convergence of each of these series using the ratio test, we find that
series 1 diverges for all x 6= 0, series 2 converges for all x, and series 3 converges for
|x| < 1. In fact series 3 for |x| < 1 is a geometric series and we can conclude that
∞
1 X
= xn , if |x| < 1,
1−x
n=0
hence the power series in this case is equal to a recognisable function. This is not
always the case, in fact it is seldom the case, but what is important is that a convergent
power series defines some function f (x) where it converges. This brings us to the crucial
question “for what values of x does a given power series converge?”
n
P
Theorem 1 Assume that there is a number P∞ R >n 0 such that n=0 |an |R con-
verges. Then for all |x| < R, the series n=0 an x converges absolutely.
Proof.
|an | |x|n ≤ |an |Rn
Hence, we obtain absolute convergence by using the comparison test with the given
series.
Definition
The greatest value
Pof R fornwhich we get convergence is called the radius of
convergence and ∞ n=0 an x converges absolutely if |x| < R. Note that x = ±R
must be tested separately.
One tool we will use for testing the convergence of power series is the ratio test as
we detail below.
1
For convergence, L|x| < 1, ⇒ |x| < L = R, where R is the radius of convergence
defined earlier.
Just as with the usual infinite series we used the root test, we also use it for power
series as detailed below.
P∞ n
Proof. For n=0 an x use the root test.
1
lim |an |1/n |x| = L|x| ⇒ |x| < =R for convergence.
n→∞ L
Examples:
(i) Determine the radius of convergence of
∞
X np
xn where p > 0 is given.
(n + 1)!
n=0
np
Hence L = limn→∞ an+1 = 0 and the radius of convergence R = ∞, i.e. ∞ p
P
an n=0 (n+1)! x
converges for all x.
P∞ xn
(ii) n=0 n Ratio test - (will do it directly now, without the L intermediate step.)
xn+1 n n
lim = lim |x| = |x|.
n→∞ (n + 1) xn n→∞ n + 1
P∞ n,
Note - also mentioned earlier: Instead of n=0 an x could define power series
centered at points other than 0, i.e.
∞
X
an (x − x0 )n .
n=0
1 1
− <x+5< ,
4 4
21 19
− < x <− .
4 4
P∞
At the point x + 5 = 1/4 the series becomes √ 1
n=0 2n+5 which diverges (e.g.
comparison test or integral test).
(−1)n
On the other hand at x+5 = −1/4 the series becomes ∞
P
n=0 2n+5 which converges
√
The question is, can we do this for power series? The answer is YES if |x| < R,
i.e. we are within the radius of convergence. We have the following very important
theorems.
Conclusion: For a power series within its radius of convergence, we can differentiate or
integrate term by term.
Note f (x) = ∞ n
P
n=0 an x can be differentiated an infinite number of times as long as
|x| < R, and the derivatives will exist. The function is smooth.
The way to show this is to consider each differentiated series as a new power series.
For example
dk X n
X
an x = n(n − 1) . . . (n − (k − 1))xn−k an
dxk
X n!
= an xn−k .
(n − k)!
Ratio test
i.e. the same radius of convergence as the undifferentiated power series. The integer k
is arbitrary so the power series can be differentiated as many times as we want and the
derivatives will exist as long as |x| < R. Similarly, we can integrate as many times as
needed.
126 CHAPTER 14. POWER SERIES
Definition
If a function f (x) has a power series expansion f (x) = ∞ n
P
n=0 an (x − x0 ) in the
neighbourhood of the point x = x0 , then it is said to be analytic at x0 . In
(n)
addition an = f n!(x0 ) . The power series is called the Taylor series of f (x) at the
point x0 .
Example
Write down power series for
x
log 1 + x2 .
and
1 + x2
Solution
1
Recall the geometric series formula 1 + r + r2 + · · · = 1−r for |r| < 1. If r = −x2 we
have
1
(1 − x2 + x4 − x6 + . . . ) · x = · x.
1 + x2
Hence
∞
x X
= (−1)n x2n+1 , |x| < 1 for convergence.
1 + x2
n=0
Z
d 2x xdx
log 1 + x2 = 2
so log 1 + x =2 .
dx 1 + x2 1 + x2
Z
2
(x − x3 + x5 . . . )dx
log 1 + x =2
x4 x6 x8
= x2 − + − + ...
2 3 4
(Constant of integration is zero since log 1 = 0). We have convergence for |x| < 1 and
also at x = ±1 since we have an alternating series now. In fact, evaluating at x2 = 1 we
obtain the formula log(2) = 1 − 12 + 31 − 14 + . . . .
14.2. DIFFERENTIATION AND INTEGRATION OF POWER SERIES 127
P∞ n
Theorem 6 - Algebraic operations. Let f (x)
P∞= n=0 an x be a power series
with radius of convergence R1 , and g(x) = n=0 bn xn be another power series
with radius of convergence R2 . Let
R = min(R1 , R2 ).
Then
(2) cf (c) = ∞ n
P
n=0 can x for |x| < R1 (c 6= 0).
(3) f (x)g(x) = ∞
P Pn n
n=0 ( m=0 am bn−m ) x for |x| < R.
P∞Some nmore details of property (3) are in order. Start with writing f (x)g(x) =
n=0 cn x , i.e.
∞
! ∞ ! ∞
X X X
n n
an x bn x = cn xn , i.e.
n=0 n=0 n=0
a0 + a1 x + a2 x2 + · · · b0 + b1 x + b2 x2 + · · · = (c0 + c1 x + c2 x2 + · · · ).
I used n as a summation variable in all series. Now we ask what terms need to be
multiplied on the left to give me a term xn on the right. Here it goes: a0 with bn xn , or
a1 with bn−1 xn−1 , or · · · , to give me precisely the sum of n products with
n
X
cn = a0 bn + a1 bn−1 + · · · + an−1 b1 + an b0 = am bn−m , (14.3)
m=0
which is the form that appears in property (3) above. There is an alternative form of
P0
Pn that arises from putting k = n − m in the series in (14.3) to give k=n an−k bk =
(14.3)
k=0 an−k bk . Hence
X n Xn
cn = am bn−m = an−m bm .
m=0 m=0
128 CHAPTER 14. POWER SERIES
Chapter 15
The objective of this chapter is to show how the power series theory developed can be
applied to construct solutions of differential equations. We will be exclusively concerned
with linear equations of first and second order 1 :
y 0 = F (x)y, y 00 + P (x)y 0 + Q(x)y = 0, (15.1)
where the functions F (x), P (x) and Q(x) are given. The objective is to find general
power series solutions of these equations once the form of F, P, Q is known. Of course,
the equation y 0 = F (x)y can be solved by separation of variables and integration, so
a development of power series solutions is only illustrative in this case. For 2nd order
equations with non-constant coefficients as is the case when P (x), Q(x) are not constant,
the situation is seldom that easy. In fact closed form solutions in terms of recognisable
elementary functions are not available. The applications of such equations are vast and
include diffraction theory, wave propagation, and quantum mechanics to mention a few.
We begin by some formal calculations to set the stage for the theory to follow.
converges for |x| < R for some positive radius of convergence R > 0. This assumption is
equivalent to assuming that y(x) is analytic at x = 0. Because of our assumptions, we
can differentiate the power series term by term to find
∞
X
0 2 n
y = a1 + 2a2 x + 3a3 x + · · · + (n + 1)an+1 x + · · · = nan xn−1 . (15.4)
n=1
dk y
1
The notation y (k) (x) = dxk
will be used with y (1) = y 0 , y (2) = y 00 , etc.
129
130CHAPTER 15. POWER SERIES SOLUTIONS OF DIFFERENTIAL EQUATIONS
Hence equation (15.2) is satisfied if the power series (15.3) and (15.4) are equal, i.e. the
coefficients of different powers of x must match. By inspection
where the last equality just switches the summation variable from m to n. Now we are
in a position to write equation (15.3) into the more useful power series form
∞
X an
[(n + 1)an+1 − an ] xn = 0 ⇒ an+1 = , (15.6)
n+1
n=0
a0
from which it follows readily that an+1 = (n+1)! , i.e. an = an!0 . Substituting this result
into the power series solution (15.3) gives
x2 xn
y(x) = a0 1 + x + + ··· + + ··· , (15.7)
2! n!
y(x) = a0 ex . (15.8)
Hence, the power series in the square bracket of (15.7) is equal to ex as long as we are
within the radius of convergence. It remains to prove that our formal calculation of
(15.7) is valid by calculating R. Using the ratio test in (15.7) we find
|xn+1 | n! |x|
lim = lim = 0, (15.9)
n→∞ (n + 1)! |xn | n→∞ n + 1
and so R = ∞, hence the power series converges for all x and ex is analytic everywhere,
a result we already know.
Equation (15.3) had constant coefficients. The next example is
The question is to find the power series expansion of (1 + x)p for arbitrary p, not just
integers (in fact p could be negative, irrational etc.). This is the Binomial Theorem 2
that we will encounter later also. We will do it by Psolving (15.10) assuming there exists
a power series solution with R > 0. Writing y = ∞ a
n=0 n x n , and picking the coefficient
of xn for each of the terms y 0 , xy 0 and py, we find that equation (15.11) is satisfied if
(n + 1)an+1 + nan = pan , n = 0, 1, 2, · · · (15.12)
hence, noting that a0 = 1 from the initial condition
a1 (p − 1) p(p − 1) a2 (p − 2) p(p − 1)(p − 2)
a1 = p, a2 = = , a3 = = ,
2 2 3 2·3
an−1 (p − n + 1) p(p − 1)(p − 2) · · · (p − n + 1)
an = = .
n n!
From these formulas we see that if p is a positive integer the series will terminate after
p + 1 terms as seen from the Binomial Theorem in footnote 2.
It remains to show that the power series
p(p − 1) 2 p(p − 1)(p − 2) · · · (p − n + 1) n
y = 1 + px + x + ··· + x + ··· , (15.13)
2! n!
has a positive radius of convergence. Using the ratio test it is easy to show that R = 1.
Additional Problems
where
n n! n(n − 1) · · · (n − m + 1)
= = .
m m!(n − m)! m!
132CHAPTER 15. POWER SERIES SOLUTIONS OF DIFFERENTIAL EQUATIONS
(b) Now get the result above by repeated differentiation of (**) and use of the
f (n) (0)
formula an = n! .
The first sum starts from n = 2 while the second one starts from n = 0. Hence, we
introduce a new variable k = n − 2 in the first sum as follows
∞ ∞
n=k+2
X X
n(n − 1)an xn−2 = (k + 1)(k + 2)ak+2 xk . (15.19)
n=2 k=0
Now the fact that there is a k in the sum in (15.19) is immaterial, I can change it back
to an n since it is a summation variable. Doing this and substituting into (15.18) gives
∞
X
[(n + 2)(n + 1)an+2 + an ] xn = 0. (15.20)
n=0
15.2. SECOND ORDER EQUATIONS 133
Equation (15.20) is equivalent to the differential equation (15.14). We know from the
theory of vector spaces that 1, x, x2 , · · · , are a linearly independent basis for polynomi-
als.3 Hence each coefficient must be zero, yielding
x2 x4 x3 x5
y(x) = a0 1 − + − . . . + a1 x − + − ... (15.22)
2! 4! 3! 5!
∞ ∞
X x2n X x2n+1
= a0 (−1)n + a1 (−1)n .
(2n)! (2n + 1)!
n=0 n=0
It will be shown in Chapter 16 that the first bracket/sum is the Taylor series of cos x
while the second bracket is that for sin x. In fact these series converge absolutely for
all x and hence the power series solution (15.22) is absolutely convergent. This is not
always the case of course. To show that the radius of convergence is R = ∞ we use the
ratio test as before. (Do this as an exercise.)
Equation (15.14) has constant coefficients and hence produced elementary solutions.
This kind of outcome is the exception rather than the rule! In applications (e.g. potential
theory, heat flow, electrostatics etc.) the equations are not as simple.
For general P (x), Q(x), elementary solutions are not available. As expected the solutions
will depend on the form and behaviour of the functions P and Q. Equation (15.14) is
the simplest we could have hoped for - it has P = 0 and Q = const. In this section we
concentrate on situations where solutions are sought near a point x = x0 and in addition
we take P (x) and Q(x) to be nice well behaved functions near x0 , i.e. they are analytic
at x = x0 (all their derivatives exist and are bounded). Such points are called ordinary
points of the differential equation.
3
If you have not seen this yet, reason as follows: the representation (15.20) is assumed to be valid
for all x, and so the only way the left hand side is zero is if the coefficients of x0 , x1 , etc are identically
equal to 0.
134CHAPTER 15. POWER SERIES SOLUTIONS OF DIFFERENTIAL EQUATIONS
2x p(p + 1)
P (x) = − , Q(x) = .
1 − x2 1 − x2
We see that P (x) and Q(x) are analytic near x = 0, and so x = 0 is an ordinary point.
In fact they are not analytic at x = ±1, they become infinite there. Hence, we cannot
expect our series solution to have an infinite radius of convergence - a good guess would
be that it is R = 1, but we will prove this. P
Seeking a series solution of the form y = ∞ n
n=0 an x , substituting this into (15.24)
and collecting all terms contributing to the coefficient of xn (see previous detailed cal-
culations) gives
(p − n)(p + n + 1)
an+2 = − an . (15.26)
(n + 1)(n + 2)
It can be seen that all even index terms a2 , a4 , · · · are expressible in terms of a0 , and all
odd ones a3 , a5 , · · · are expressible in terms of a1 . Since a0 and a1 are arbitrary constants,
the construction gives us two linearly independent solutions as desired. Calculating gives
p(p + 1) (p − 1)(p + 2)
a2 = − a0 , a3 = − a1 ,
1·2 2·3
If p is not an integer then the radius of convergence of each series in the brackets is
R = 1. The easiest way to prove this is to use the recursion formula (15.26) and the
ratio test. Considering the even terms first, we can replace n in (15.26) by 2n to find
hence we have convergence for |x| < 1, i.e. R = 1. The calculation for the odd terms is
completely analogous and is left as an exercise. This justifies the assumption of a power
series and the solution is analytic as long as |x| < 1.
Note: The solutions (15.27) for general p are not elementary. If p is a non-negative
integer, then one of the series in (15.27) terminates after a finite number of terms and
hence is a polynomial. The first series terminates if p is even and the second one if p is
odd. In either case the remaining series remains infinite and non-elementary. The result-
ing polynomials are called Legendre polynomials and play a huge role in mathematical
physics.
We have the following important theorem that connects the properties of P and Q
to those of the solutions.
Theorem
Let x0 be an ordinary point of
and let a0 , a1 be arbitrary constants. Then there exists a unique function y(x) that
is analytic at x0 and solves the differential equation (15.28) with initial conditions
y(x0 ) = a0 , y 0 (x0 ) = a1 . Furthermore, if the power series expansions of P (x) and
Q(x) are valid for |x − x0 | < R with R > 0, then the power series expansion of
y(x) is also valid on the same interval.
The proof, albeit straightforward, is beyond the scope of this course and is omitted.
The important aspect of the theorem is that once we know the radius of convergence for
P and Q we know that for the solution with no further calculation. Recall the Legendre
equation solved earlier. Since P and Q contain 1/(1 − x2 ) = 1 + x2 + x4 + · · · , we
immediately have that R = 1 since the series is geometric. Hence the series solution
(15.27) also has radius of convergence R = 1 without the need of the ratio test that was
carried out.
have
lim |P (x)| = ∞ = lim |Q(x)|,
x→±1 x→±1
and so we have a fundamentally different problem.
Here is another example from mathematical physics, the Bessel equation of order p
1 0 x2 − p2
x2 y 00 + xy 0 + (x2 − p2 ) = 0 ⇒ y 00 + y + y = 0, (15.30)
x x2
which clearly has (for p 6= 0)
lim |P (x)| = ∞ = lim |Q(x)|.
x→0 x→0
Notice that the leading order terms have been underlined. If I write down an equation
that involves only the underlined terms then this is
and we drop the −1 index from the constants for simplicity to write
x2 y 00 + bxy 0 + cy = 0. (15.34)
Equations such as (15.34) are called Euler differential equations. They have non-constant
coefficients but have very simple solutions that can be found by making the transforma-
tion z = log x, assuming x > 0 for the moment. Using the chain rule we have
dy dy dz dy 1
= = ,
dx dz dx dz x
d2 y 1 d2 y
d 1 dy 1 dy 1 d dy 1 dy
2
= =− 2 + =− 2 + 2 2.
dx dx x dz x dz x dx dz x dz x dz
Substituting these expressions into (15.34) gives
d2 y dy
2
+ (b − 1) + cy = 0, (15.35)
dz dz
which can be solved by looking for a solution of the form y = eλz so that the indicial
equation is
λ2 + (b − 1)λ + c = 0. (15.36)
If the roots are λ1 and λ2 then we have the following linearly independent solutions
e λ1 z and eλ2 z λ2 6= λ1 ,
e λ1 z and zeλ1 z λ2 = λ1 .
Note that if we look for solutions in x < 0, it is enough to put t = −x and then work
with t > 0 so that the results above hold. Note also that λ1 and/or λ2 can be negative.
The Euler solutions (15.37) or (15.38) are suggestive of the solutions when x0 is a
regular singular point. The leading singular part(s) are multiplied by power series of the
form
y = xλ (a0 + a1 x + a2 x2 + · · · ), (15.39)
where λ must be found and if the roots are equal a second solution (see (15.38)) of the
form
y = xλ log x(a0 + a1 x + a2 x2 + · · · ). (15.40)
Such series solutions are known as Frobenius solutions.
A worked example:
138CHAPTER 15. POWER SERIES SOLUTIONS OF DIFFERENTIAL EQUATIONS
(1/2) + x (−1/2)
2x2 y 00 + x(2x + 1)y 0 − y = 0 or y 00 + + y = 0. (15.41)
x x2
Since xP (x) = (1/2) + x and x2 Q(x) = −1/2, the point x = 0 is a regular singular point.
Hence, we look for a Frobenius solution of the form
y = xλ (a0 + a1 x + a2 x2 + · · · ), (15.42)
and substitute this into the first of equations (15.41). A common factor xλ−2 can be
cancelled to yield
The constant a0 can be taken to be non-zero - if it is, then the series solution (15.39)
has a common factor x which can be taken out to leave xλ+1 (a1 + a2 x + · · · ), which on
redefinition of λ+1 → λ and the constants ak → ak−1 brings us back to the form (15.39).
Hence a0 6= 0 will be assumed throughout. Hence, the indicial equation becomes
λ 1 1
λ(λ − 1) + − =0 ⇒ λ1 = 1, λ2 = − .
2 2 2
The two roots are distinct and one is negative, hence the solution will be singular. The
Frobenius method enables us to find the precise structure of the singularity. The two
independent series solutions are constructed by picking each of the indicial roots and
working out the coefficients using equations (15.44), (15.45), etc.
Starting with λ = 1 we find
a0 2
a1 = − 1 1 = − a0 ,
2·1+ 2 ·2− 2
5
2a1 4
a2 = − = a0 ,
3 · 2 + 21 · 3 − 1
2
35
15.2. SECOND ORDER EQUATIONS 139
and so on.
For λ = −1/2 we find
a0
2
a1 = 1 1
1 1 1 = −a0 ,
2 −2 + 2 · 2 − 2
a1
2 1
a2 = 3 1 1 3 1 = a0 ,
2 · 2 + 2 · 2 − 2
2
and so on.
Even though both independent solutions have been calculated in terms of a0 , it
should be understood that a0 is an arbitrary constant in each case, it is not the same.
Hence the solutions can be written as (picking c1 and c2 for a0 in each independent
solution)
2 4 1
y = c1 x 1 − x + x2 + · · · + c2 x−1/2 1 − x + x2 + · · · . (15.47)
5 35 2
We will conclude this section with a theorem that determines the radius of conver-
gence for the series solutions.
Theorem
Assume x = 0 is a singular point x0 of equation (15.23), and the power series
expansions of xP (x) and x2 Q(x) have radius of convergence R > 0, i.e. the power
series are valid on the interval |x| < R. If the roots of the indicial equation are
real and λ2 ≤ λ1 , then equation (15.23) has at least one solution
∞
X
λ1
y1 = x an xn (a0 6= 0),
n=0
where
P∞ then coefficients an are again found from the recursion formulas and
n=0 an x converges for |x| < R.
140CHAPTER 15. POWER SERIES SOLUTIONS OF DIFFERENTIAL EQUATIONS
Chapter 16
Taylor Series
This is a power series that represents a function f (x) by using its derivatives at a single
point. Intuitive construction: Assume the power series exists and identify the coefficients.
Take a fixed point x = x0 . If
∞
X
f (x) = an (x − x0 )n
n=0
f (x) = a0 + a1 (x − x0 ) + a2 (x − x0 )2 + a3 (x − x0 )3 + . . .
f 0 (x) = a1 + 2a2 (x − x0 ) + 3a3 (x − x0 )2 + . . .
f 00 (x) = 2a2 + 3 · 2a3 (x − x0 ) + . . .
f 000 (x) = 3 · 2 · 1a3 + . . .
dk f
f (k) (x) := = k!ak + O(x − x0 ).
dxk | {z }
(*)
So we can see immediately that by putting x = x0 in the formula of f (k) (x) we find
1 (k)
ak = f (x0 ) ⇒
k!
∞
X f (n) (x0 )
f (x) = (x − x0 )n (Note 0! = 1).
n!
n=0
This is the Taylor series about the point x = x0 . If x0 = 0 we get the Maclaurin
series
∞
X f (n) (0) n
f (x) = x .
n!
n=0
141
142 CHAPTER 16. TAYLOR SERIES
Example 2
f (2) (x0 )
f (x) = f (x0 )+f 0 (x0 )(x − x0 ) + (x − x0 ) + . . .
2!
f (n) (x0 )
+ (x − x0 )n + Rn
n!
where the remainder Rn is given by
Z x
(x − t)n (n+1)
Rn = f (t) dt.
x0 n!
Proof. Use integration by parts. From the fundamental theorem of calculus we have
Z x
f (x) = f (x0 ) + f 0 (t)dt. (16.1)
x0
16.1. TAYLOR’S THEOREM WITH REMAINDER 143
Now Z x Z x
0
f (t)dt = f 0 (t) d(−(x − t)) (16.2)
x0 x0
One more
x Z x
(x − t)2 (2) x (x − t)2 (3)
Z
(x − t)f (2) (t)dt = − f (t)|x0 + f (t)dt
x0 2 x0 2
Z x
(x − x0 )2 (2) (x − t)2 (3)
= f (x0 ) + f (t)dt.
2 x0 2
The choice in the integration by parts giving (16.2) may seem a little strange at first.
To motivate the choice and to understand why it is the only way to get something useful,
let us start with (16.1) and integrate by parts as we normally would, i.e. think of 1 as
the differential of t. This is what we would obtain
Z x Z x Z x
0
0 x 0 0 0
f (t)dt = tf (t) x0 − tf (t)dt = xf (x) − x0 f (x0 ) − tf 0 (t)dt, (16.3)
x0 x0 x0
The identity (16.4) is of course perfectly valid but it is not useful. The reason is that we
want to generate a power series whose coefficients are known in terms of the function and
its derivatives at x = x0 (see the statement of the Theorem). Going one step further, I
can also think of 1 as the differential with respect to t of (t − α) where α is any number
independent of t. Using this to carry out the integration by parts casts (16.1) into (verify
this)
Z x
0 0
f (x) = f (x0 ) + (x − α)f (x) − (x0 − α)f (x0 ) − (t − α)f 00 (t)dt. (16.5)
x0
We can see now that picking α = x will remove the second term on the RHS and lay the
ground for the development of the known power series and the proof developed above.
Here is a very useful alternative of Taylor’s theorem that we use in Numerical Anal-
ysis. Put x = x0 + h (and after that x0 → x if you want)
h2 hn (n)
f (x0 + h) = f (x0 ) + hf 0 (x0 ) + f (2) (x0 ) + · · · + f (x0 ) + Rn (x0 , h)
2! n!
Z x0 +h
(x0 + h − t)n (n+1)
Rn = f (t)dt
x0 n!
h(n+1) (n+1)
= f (c)
(n + 1)!
where c is between x0 and x0 + h.
with c is a number between 0 and x. If |f (n+1) (x0 )| ≤ Mn+1 for all x0 between 0 and x.
Then
|x|n+1
|Rn | ≤ Mn+1
(n + 1)!
Example 3
x3 x2n+1
f (x) = sin(x) = x − + · · · + (−1)n + R2n+3
3! (2n + 1)!
f (2n+3) (c) 2n+3 |x|2n+3
where |R2n+3 | = x ≤ .
(2n + 3)! (2n + 3)!
10−3
Hence sin(0.1) ≈ 0.1 − 6 with an error which is less than
(0.1)7 10−7
= < 10−10 .
7! 5040
π
+ 0.2 to an accuracy of 10−4 .
Example 4 Compute sin 6
Solution 3 5
Even though sin(x) = x − x3! + x5! + . . . will converge if we take enough terms, since
π
6 + 0.2 is not small, we will need a lot πof terms to get to the required accuracy.
It is much better to expand about 6 using the formula
h2 00 hn (n)
f (x + h) = f (x) + hf 0 (x) + f (x) + · · · + f (x) + Rn
2! n!
f (n+1) (c) n+1
where Rn = h with c between x and x + h.
(n + 1)!
Example 5
Compute e to 3 decimals. Showed earlier (see chapter on logarithms) that 2 < e < 4.
e 4
From results above, |Rn | ≤ (n+1)! ≤ (n+1)! .
Need 4
(n+1)! to be less than 10−3 .
4 1 1
Try n = 4, 5, 6 → = = > 10−3
5! 5×3×2 30
4 1 1
= = > 10−3
6! 6×5×3×2 180
4 1 1
|R6 | ≤ = = < 10−3 .
7! 7×6×5×3×2 1260
So
1 1 1 1 1
e≈1+1+ + + + + + R6 .
2 3! 4! 5! 6! |{z}
<10−3
x2 x3 xn
log(1 + x) = x − + − · · · + (−1)n−1 + Rn+1 .
2 3 n
Need to show this. One way is Taylor’s theorem - exercise. Another way is to use
the identity (telescoping product)
x2 x3 xn
log(1 + x) = x − + − · · · + (−1)n−1 + Rn
2 3 Z x n n
t
where Rn = (−1)n dt.
0 1+t
xn+1 an+1
|Rn | ≤ ≤ →0 as n → ∞.
n+1 n+1
(ii) Now take −1 < a < 0 and consider x in the interval a ≤ x ≤ 0. Hence
1 + t ≥ 1 + a > 0 since t is in the interval (x, 0).
16.3. EXPONENTIALS AND LOGARITHMS. BINOMIAL THEOREM 147
tn (−t)n
≤ (t ≤ 0 remember)
1+t 1+a
0
(−t)n (−x)n+1 |a|n+1
Z
So |Rn | ≤ dt = ≤ →0 as n → ∞.
x 1+a (n + 1)(1 + a) (n + 1)(1 + a)
α(α − 1) 2
(1 + x)α = 1 + αx + x + ...
2!
∞
X α(α − 1) . . . (α − n + 1) n
= x .
n!
n=0
Example 7
log(x)
lim put x = 1 + y
x→1 ex − e
2 3
log(1 + y) y − y2 + y3 + . . .
= lim = lim
y→0 e(ey − 1) y→0 e(1 + y + · · · − 1)
1
= .
e
f (k) (c1 )
⇒ F (a + h) = , send h → 0, c1 , c2 → a, so we get the required result.
g (k) (c1 )
f (x) and all its derivatives are continuous everywhere. (You have shown this in the
problem sheets). In addition, f (n) (0) = 0 for all n. So the (Taylor) Maclaurin expansion
is
f 00 (0) 2
f (0) + f 0 (0)x + x + ...
2!
The approximating polynomial is
f (n) (0) n
Pn (x) = f (0) + xf 0 (0) + · · · + x .
n!
But this is exactly zero for all
n. Hence the remainder Rn cannot go to zero. In fact it
2
must be equal to exp −1/x2 . Reason: e−1/z with z a complex number is not analytic.
2
In fact, z = iy gives f = e1/y → ∞ as y → 0. You will see more in Complex Analysis.
Part IV
Fourier Series
149
Chapter 17
Will see how to represent fairly arbitrary functions (e.g. they can be discontinuous) with
approximations of smooth functions.
Definition 1
If f , g are real valued functions that are Riemann integrable on [a, b], then we
define the inner product of f and g, denoted by (f, g), by
Z b
(f, g) := f (x)g(x)dx
a
Z b 1/2
1/2 2
Note (f, f ) = f dx := ||f || ≥ 0.
a
Definition 2
Let S = {φ0 , φ1 , φ2 , . . . } be a collection of functions that are Riemann integrable
on [a, b]. If
(φn , φm ) = 0 whenever m 6= n
φn
Note: Can easily go from orthogonal to orthonormal by considering ||φn || . The or-
thonormal trigonometric system will be used
S = {φ0 , φ1 , φ2 , . . . } where
1 cos(nx) sin(nx)
φ0 (x) = √ , φ2n−1 = √ , φ2n = √ (n = 1, 2, . . . )
2π π π
151
152 CHAPTER 17. ORTHOGONAL AND ORTHONORMAL FUNCTION SPACES
S defined above is orthonormal on any interval of length 2π, e.g. [0, 2π], [−π, π] etc..
(You have already shown this in HW sheet 3).
Chapter 18
Start with any continuous function f (x) in an interval a ≤ x < b. Can extend this
periodically to have period T = b − a.
Figure 18.1: Function on [a, b) extended periodically and the new function is discontin-
uous at x = a + mT for any integer m.
153
154 CHAPTER 18. PERIODIC FUNCTIONS AND PERIODIC EXTENSIONS
This is known as f (x) = x − [x], where [x] denotes the integer part of x (e.g. [1.25] =
1).
From the examples, we see that the function can be discontinuous at some points
x = ξ, i.e.
Definition
At points of discontinuity define
1
f (ξ) = [f (ξ+ ) + f (ξ− )] .
2
1
e.g. for f (x) = x − [x], f (n) = 2 for all integers n.
Conclusion: Given a function defined on a closed interval [a, b], extend it periodically
and at points of discontinuity x = ξ prescribe the value f (ξ) = 12 (f (ξ− ) + f (ξ+ )).
e.g. y = x 0≤x≤1
155
In fact,
Z β Z β+T
f (x)dx = f (x)dx.
α α+T
Proof.
Z β Z β+T Z β+T Z β+T
f (x)dx = f (y − T )dy = f (y)dy = f (x)dx
|α {z } α+T α+T α+T
sub x = y − T
156 CHAPTER 18. PERIODIC FUNCTIONS AND PERIODIC EXTENSIONS
Chapter 19
Trigonometric polynomials
and since
157
158 CHAPTER 19. TRIGONOMETRIC POLYNOMIALS
So, can represent everything as complex and find real expressions by taking real or
imaginary parts. e.g.
d iω(x−φ)
ae = aiωeiω(x−φ) .
dx
Can also integrate
Z Z
inx sin(nx) i cos(nx)
e dx = (cos(nx) + i sin(nx)) dx = −
n n
1
= einx .
in
19.1.1 Orthogonality
(
π
n 6= 0
Z
0
einx dx = .
−π 2π n=0
For any integers m, n we have
Z π (
inx −imx 0 n 6= m
e e dx = (easier than HW3.)
−π 2π n=m
where
= 21 a0
γ0
1
γk = 2 (ak − ibk ) k = 1, 2, . . . , n.
= 12 (ak + ibk )
γ−k
Notice that γk = ∗ ,
γ−k (or γk∗ = γ−k ), where ∗ denotes complex conjugate. This is not
accidental. Sn (x) in equation (19.1) is real. Hence it must equal its complex conjugate.
Calculate
19.2. COMPLEX NOTATION FOR TRIGONOMETRIC POLYNOMIALS 159
n
X n
X
Sn (x) = γk eikx , Sn (x)∗ = γk∗ e−ikx .
k=−n k=−n
In the last step above, I just changed the summation variable l to k. Comparing the
two, we see that they are equal iff
∗
γk = γ−k i.e. γk∗ = γ−k , identical statements.
∗ , i.e.
then f (x) is real if and only if γk = γ−k
1
Example 1 Take Sn (x) = cos(x) + 2 sin(x) + 3 cos(2x). Express as a complex trigono-
metric series
1 ix i ix 3 2ix
e + e−ix − e − e−ix + e + e−2ix
Sn =
2
4 2
1 i 1 i 3 3
= − eix + + e−ix + e2ix + e−2ix
2 4 2 4 2 2
2
X
= γk eikx
k=−2
where
1 i 1 i
γ1 = − γ−1 = + = γ1∗
γ0 = 0 2 4 2 4
3 3
γ2 = γ−2 = = γ2∗ .
2 2
160 CHAPTER 19. TRIGONOMETRIC POLYNOMIALS
Chapter 20
Fourier series
N
1 X
f (x) = SN (x) = a0 + an cos(nx) + bn sin(nx).
2
n=1
There are 2N + 1 coefficients to determine. Use orthogonality on the interval [−π, π] for
sin(mx), cos(nx) etc.
1 π
Z
an = f (x) cos(nx)dx
π −π
1 π
Z
bn = f (x) sin(nx)dx.
π −π
Big question is: starting with fairy arbitrary functions f (x) (e.g. they are discontinuous),
can we represent them by SN by letting N → ∞?
Orthogonality properties:
If m, n are integers, then
(
π π
m 6= n
Z Z
0
sin(mx) sin(nx)dx = cos(mx) cos(nx)dx =
−π −π π if m = n 6= 0
Z π
sin(mx) cos(nx) = 0.
−π
Complex form (
π
m 6= n
Z
imx −inx 0
e e dx =
−π 2π m = n.
161
162 CHAPTER 20. FOURIER SERIES
Theorem 1
The Fourier series 21 a0 + ∞
P P∞ inx ),
n=1 (an cos(nx) + bR
n sin(nx)) (or n=−∞ αn e
1 π
formed by the Fourier coefficients an = π −π f (x) cos(nx)dx and bn =
1 π
R
π −π f (x) sin(nx)dx converges to the value f (x) for any piecewise continuous
function f (x) of period 2π which has piecewise continuous derivatives of first and
second order.a At any discontinuities, the value of the function must be defined
by f (x) = 21 [f (x+ ) + f (x− )].
Can relax the assumption of the second derivative. It is enough to have f 0 (x) be piecewise
a
continuous, i.e. the function is piecewise smooth. If f (x) is continuous, the convergence is
absolute and uniform. If it is discontinuous, absolute and uniform convergence everywhere except
at the discontinuity.
n
1 X ikx 1 −inx
+ eix e−inx + (eix )2 e−inx + · · · + e−inx e2inx
cn (x) = e = e
2 2
k=−n
i.e. a geometric progression with ratio r = eix = cos(x) + i sin(x). Now r = 1 only if
x = 0, ±2π, . . . , i.e. the exceptional points that we excluded (treated separately). Sum
it up to find
1 1 − r2n+1 1 h −inx i 1
cn (x) = e−inx = e − ei(n+1)x .
2 1−r 2 1 − eix
1
Multiply top and bottom by e− 2 ix
h i
−i(n+ 12 )x i(n+ 12 )x
1 e − e
⇒ cn (x) = 1 1
2 e− 2 ix − e 2 ix
sin (n + 12 )x
⇒ cn (x) = .
2 sin 12 x
20.1. FOURIER SERIES THEOREM, RIEMANN-LEBESGUE LEMMA 163
Lemma 1 (Riemann-Lebesgue)
If the function g(x) is integrable on [a, b], (e.g. it is piecewise continuous), then
Z b
Iλ = g(x) sin(λx)dx
a
tends to zero as λ → ∞.
Proof. (Will do it when g 0 is also piecewise continuous. For the general case see HW
problems).
Can use integration by parts
Z b b Z b
cos(λx) cos(λx) 0
Iλ = g(x) sin(λx)dx = − g(x) + g (x)dx
a λ a a λ
Z b
1
= g(a) cos(λa) − g(b) cos(λb) + cos(λx)g 0 (x)dx
λ a
1
⇒ |Iλ | ≤ M
λ
for some constant M , and the result follows.
Lemma 2
Z ∞
sin(z) π
dz = .
0 z 2
Z N
1 1 dz 2
⇒ |IN − IM | ≤ + + 2
=
M N M z M
2
In fact, letting N → ∞, we see that |I − IM | ≤ M so IM approaches its limit
algebraically. Now take p > 0 arbitrary and pick M = λp.
Z λp z=λx Z p
sin(z) z}|{ sin(λx)
IM = Iλp = dz = (λdx)
z λx
Z 0p 0
= sin(λx)
dx
0 x
where we have now fixed the integration range to [0, p]. As M → ∞, λp → ∞ i.e.
λ → ∞, and by the estimate above
Z p
sin(λx) 2 2
I− dx ≤ =
x M λp
Z0 p
sin(λx)
i.e. lim dx = I (20.1)
λ→∞ 0 x
for all p sufficiently big. Cannot apply Riemann-Lebesgue directly. Consider the function
(
1 1
x − 2 sin(x/2) x 6= 0
h(x) =
0 x = 0.
Fact: h(x) is continuous and also has a continuous first derivative for 0 ≤ x < 2π. (Proof
see HW5). Now we use the Riemann-Lebesgue Lemma 1 to see that for 0 ≤ p < 2π
Z p
1 1
sin(λx) − dx → 0 as λ → 0.
0 x 2 sin(x/2)
Note: The convergence is uniform for 0 ≤ p ≤ π since |h(x)| and |h0 (x)| are both
bounded in this interval. From (20.1) we have immediately,
Z p
sin(λx)
lim dx = I.
λ→∞ 0 2 sin(x/2)
1
Pick λ = n + 2 and p = π, we have shown already that
sin (n + 21 )x
Z π
π
dx =
0 2 sin(x/2) 2
and substitute the formulas for ak , bk , change order of summation and integration (finite
sum, so ok), to find
n
" #
1 π
Z
1 X
Sn (x) = f (t) + (cos(kt) cos(kx) + sin(kt) sin(kx)) dt
π −π 2
k=1
n
" #
1 π
Z
1 X
= f (t) + cos(k(t − x)) dt
π −π 2
k=1
sin (n + 12 )(t − x)
1 π
Z
= f (t) dt
π −π 2 sin 12 (t − x)
1 π−x f (x + ξ) sin (n + 12 )ξ
Z
substitute ξ = t − x ⇒ = dξ
2 sin 21 ξ
π −π−x
sin (n + 12 )ξ
1 π
Z
= f (x + ξ) dξ (20.2)
π −π 2 sin 21 ξ
sin (n + 12 )ξ
1 π
Z
lim f (x + ξ) dξ = f (x).
n→∞ π −π 2 sin 12 ξ
1
f (x+ ) + f (x− ) .
f (x) =
2
We have proven already that
sin (n + 12 )t
Z π
π
1
dt =
0 2 sin 2 t 2
sin (n + 21 )t0
Z 0
π
1 0
dt0 = .
−π 2 sin 2 t 2
Hence
sin (n + 12 )t sin (n + 12 )t
π
1 0
Z Z
1 + −
f (x) = f (x ) dt + f (x ) dt.
π 0 2 sin 12 t π −π 2 sin 12 t
166 CHAPTER 20. FOURIER SERIES
This is also odd of course, but we get φ(x) from f (x) by (i) shifting the latter to the
right by π (ii) reflecting about x = 0.
⇒ φ(x) = f (−(x − π)) = f (π − x)
sin(2(π − x)) sin(3(π − x))
= 2 sin(π − x) + + + ...
2 3
sin(2x) sin(3x)
= 2 + sin(x) − + − ...
2 3
∞
X sin(kx)
=2 (−1)k+1 .
k
k=1
Convergence is uniform as long as |x| < π − for any small > 0. In particular, putting
x = π2 we recover the Leibnitz series
π 1 1 1
= 2 1 − + − + ...
2 3 5 7
d P∞
Note: Cannot differentiate dx 2 k=1 (−1)k+1 sin(kx)
k and get a convergent series. Rea-
son: derivative of φ(x) does not satisfy conditions of Fourier Theorem.
20.2. EXAMPLES, SINE AND COSINE SERIES 169
Example 2:
f (x) = |x| −π ≤x≤π
Function is now continuous but has discontinuous derivatives at a set of finite points
±nπ.
1 X
Even function ⇒ f (x) = a0 + an cos(nx)
2
n=1
and
2 π
Z
an = x cos(nx)dx
π 0
sin(nx) π cos(nx) π
2
= x +
π n 0 n2 0
2 1
= (cos(nπ) − 1) 2
π
( n
0 n even n 6= 0
⇒ an = 4
− πn2 n odd
2 π
Z
a0 = x dx = π
π 0
1 4 cos(3x) cos(5x)
⇒ |x| = π − cos(x) + + + . . . .
2 π 32 52
∞
π2 X 1 1 1
= 2
= 1 + 2 + 2 + ...
8 (2n + 1) 3 5
n=0
Example 3
−1 for − π < x < 0
f (x) = sgn(x) = 0 x=0
+1 0 < x < π
170 CHAPTER 20. FOURIER SERIES
P
Clearly f (x) is odd ⇒ f (x) = bn sin(nx)
π
cos(nx) π
Z
2 2
bn = −
sin(nx)dx =
0 π π n 0
(
2 (1 − cos(nπ)) 0 n even
= = 4
π n nπ n odd
4 sin(3x)
⇒ sgn(x) = sin(x) + + ... .
π 3
π
Check: function f (x) = 0 at x = nπ, uniform convergence elsewhere. Putting x = 2
again gives
π 1 1 1
= 2 1 − + − + . . . . Leibnitz
2 3 5 7
d
Note: dx |x| = sgn(x) two series agree everywhere except at discontinuities. Ques-
tion: why can I differentiate |x| series but not φ(x)? Former is sectionally or piecewise
continuous.
where )
1
γn = 2 (an − ibn )
1
for n = 1, 2, . . .
γ−n = 2 (an + ibn )
1 1 π
Z
1
γn = (an − ibn ) = · (f (x) cos(nx) − if (x) sin(nx)) dx
2 2 π −π
Z π
1
= f (x)e−inx dx.
2π −π
20.4. FOURIER SERIES ON 2L−PERIODIC DOMAINS 171
Similarly Z π
1
γ−n = f (x)e+inx dx.
2π −π
(Clearly γn∗ = γ−n since f (x) is real).
Hence,
∞
X
f (x) = γn einx −π <x<π
n=−∞
where Z π
1
γn = f (x)e−inx dx n = 0, ±1, ±2, . . .
2π −π
Note: If the period is 2L instead of 2π
∞ Z L
X 1
f (x) = γn e inπx/L
, |x| < L, γn = f (x)e−inπx/L dx, n = 0, ±1, ±2, . . .
−∞
2L −L
∞ h
1 X nπx nπx i
⇒ f (x) = a0 + an cos + bn sin
2 L L
n=1
1 L
Z nπx
where an = f (x) cos dx
L −L L
1 L
Z nπx
bn = f (x) sin dx.
L −L L
The complex form is
∞
X
f (x) = γn einπx/L |x| ≤ L (20.4)
n=−∞
Z L
1
γn = f (x)e−inπx/L dx n = 0, ±1, ±2, . . . (20.5)
2L −L
172 CHAPTER 20. FOURIER SERIES
then we have Z π ∞
1 1 X
f 2 dx = a20 + (a2n + b2n ).
π −π 2
n=1
as needed.
Example 4 Compute the Fourier series of cos(x/2) over (−π, π]. Use Parseval’s theorem
to deduce the value of
∞
X 1
.
(4n − 1)2
2
n=1
20.6. FOURIER TRANSFORMS AS LIMITS OF FOURIER SERIES 173
Function is even ⇒
x 1 X∞
cos = a0 + an cos(nx) − π ≤ x ≤ π
2 2
n=1
2 π
Z x 4
a0 = cos dx =
π 0 2 π
Z π
2 x
an = cos cos(nx)dx
π 0 2
2 π1
Z
1 1
= cos (n + )x + cos (n − )x dx
π 0 2 2 2
" #
1 1
1 sin (n + 2 )π sin (n − 2 )π
= 1 +
π n+ 2 n − 12
" #
(−1)n
1 cos(nπ) cos(nπ) 2 2
= − = −
π n + 12 n − 12 π 2n + 1 2n − 1
(−1)n −4
= .
π 4n2 − 1
By Parsevel’s theorem
∞
1 π
Z
x 1 X
cos2 ( )dx = a20 + a2n
π −π 2 2
n=1
∞
8 16 X 1
= +
π2 π2 (4n2 − 1)2
1
∞
X 1 π2 −8
LHS = 1 ⇒ = .
(4n2 − 1)2 16
n=1
π
where h = L. In the limit L → ∞, h → 0 but nh := ωn = O(1).
∞ Z L
1 X −iωn t
f (x) = h f (t)e dt eiωn x .
2π n=−∞ −L
P∞
This is of the form n=−∞ G(ωn )h. Now h = ωn+1 − ωn = (n + 1)h − nh := δω
∞
X Z ∞
⇒ Riemann sum G(ωn )δω → G(ω)dω.
n=−∞ −∞
Very useful in many applications. You will use them a lot to solve differential equa-
tions.
Part V
175
Chapter 21
There is a general way of representing curves in the plane as the zero level curves of
a function of several variables. You will encounter a lot more of this in Multivariable
Calculus, but for our needs we can see that the function
f (x, y) = 0, (21.2)
where we will use the shorter notation x(t), y(t) without confusion. We will assume
throughout that φ(t) and ψ(t) possess continuous derivatives.
Given an interval t0 ≤ t ≤ t1 , equation (21.3) represents a curve C in the plane. This
curve can cross itself, can be multivalued etc., see the examples of a circle (21.1) and an
ellipse whose parametrisation is
177
178 CHAPTER 21. THEORY OF PLANE CURVES
Take an arbitrary function τ = h(t) that is monotonic and continuous in the t−interval
of interest. We are interested in using τ in the parametrisation rather than t. The in-
verse of h exists, hence we can write t = σ(τ ) in a corresponding τ −interval. Then the
parametrisation of (21.3) becomes
with α(τ ), β(τ ) also continuous. Since σ is monotonic, for each t there exists a unique
τ . The role of τ is to “re-label” the points on the curve; this is very useful in Numerical
Analysis where we may want to cluster nodes in a given region of the curve.
Example 1 For the line y = x we have an obvious parametrisation
x = t, y = t, −∞ < t < ∞. (21.6)
Regularly spaced points in t give regularly spaced points on the curve (21.6). However,
regularly spaced points in τ will give points that are further apart on the line (21.7).
This is illustrated in Figure 21.1 where the marked points are ti = 0.02i, i = 0, . . . 50 on
the left panel, and mapped according to (21.7) on the right panel.
1 1
0.9 0.9
0.8 0.8
0.7 0.7
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0.2 0.2
0.1 0.1
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x x
Figure 21.1: Left panel parametrisation (21.6). Right panel parametrisation (21.7).
21.3. MOTION ALONG A CURVE WITH TIME AS A PARAMETER 179
The chosen interval for t describes the ellipse exactly once in the positive (anti-clockwise)
direction. The transformation
a(1 − τ 2 ) 2bτ
x= , y= . (21.8)
1 + τ2 1 + τ2
All points of the ellipse are recovered except the point at (−a, 0) which is recovered
asymptotically close as τ → ±∞.
Show (21.8) as an exercise. Use trigonometric identities, e.g. cos t = 2 cos2 (t/2) − 1,
etc., and write sin(t/2), cos(t/2) in terms of τ .
Example 3 We can remove singularities in a curve by an appropriate parametrisation.
Consider the function y = x2/3 . There is a cusp at x = 0, since dy/dx = (2/3)x−1/3
which is unbounded at x = 0. We can parametrise the curve by the smooth functions
x = t3 , y = t2 , −∞ < t < ∞.
3. Label the point of the circle at the origin by P and define this time as t = 0
4. Now roll the circle to the right (without any sliding) with unit angular velocity so
that at time t the circle has turned from its original orientation by an angle t.
5. This means that the centre M of the circle moves with constant speed a to the
right.
6. The cycloid is the curve traced by the point P as the circle moves to the right. A
sketch is given in Figure 21.2.
180 CHAPTER 21. THEORY OF PLANE CURVES
4.5
3.5
3
y
2.5
1.5
M
1 a
P t
a
0.5
0
0 1 N 2 3 4 5 6
x
Figure 21.2: Construction of the cycloid by a rolling circle on a flat plane.
The coordinates of M are (at, a), and noting that the angle P M N is t, we immedi-
ately have from simple trigonometry
x = a(t − sin t), y = a(1 − cos t), t ≥ 0. (21.9)
The blue branch of the cycloid drawn in Figure 21.2 corresponds to 0 ≤ t ≤ 2π. As t
increases the figure repeats identically - it is a periodic function of period 2π.
Expression (22.45) is compact and elegant, unlike the x − y version. The latter can
be calculated by noting that cos t = a−y
a , i.e. t = cos
−1 a−y . Follow this through to show
a
that
a−y
x = a cos−1
p
∓ y(2a − y). (21.10)
a
Note that (21.10) is an equation for x in terms of y unlike the usual y in terms of x. The
multivalued nature of the cos−1 function is what gives a countable infinity of x values
for a given 0 < y < 2a.
Note: If we take the solution (22.45) and send y → −y and also shift up by a units, we
obtain the curve
x = a(t − sin t), y = a cos t, t ≥ 0. (21.11)
This is now a convex curve and we can study what happens if particles are released and
slide under gravity in the bowl type region of the curve. Such problems are taken up
later.
21.3. MOTION ALONG A CURVE WITH TIME AS A PARAMETER 181
Exercise 2: Sketch the curve x = t2 − 1, y = t3 − t, for −∞ < t < ∞ and indicate with
arrows the direction of increasing t starting from −∞. To help you along with minimal
calculations, consider the first time the curve passes through the origin, the next time it
again passes through the origin, and what happens between these times.
Challenge Problem
(a) Consider a fixed circle Ca of radius a with its centre at (0, 0) in the usual Cartesian
coordinates. A second circle Cb of radius b is rolling at a uniform speed along the
circumference and outside of the first circle. Suppose that the the circle Cb is
rolling in such a way that its centre has rotated about the origin to an angle t in
time t. A point P is fixed on Cb and at t = 0 the coordinates of P are at the
contact point (a, 0).
Show that the position of P as the circle Cb rolls over the circle Ca is given by the
parametric equations
a+b
x(t) = (a + b) cos t − b cos t ,
b
a+b
y(t) = (a + b) sin t − b sin t .
b
In the special case a = b = 1, sketch the trajectory of P and analyse what happens
in the neighbourhood of (1, 0).
(b) Now take b < a but allow the circle Cb to roll as above but on the interior circum-
ference of Ca . Again start with the point P at the position t = 0, and show that
its trajectory is
a−b
x(t) = (a − b) cos t + b cos t ,
b
a−b
y(t) = (a − b) sin t − b sin t .
b
5 4
x2 + y 2 = a2 + a2 cos(3t),
9 9
and use this result to show that the trajectory meets the circle at exactly three
points. Sketch the resulting curve.
182 CHAPTER 21. THEORY OF PLANE CURVES
and equivalently dx/dy = ẋ/ẏ holds if ẏ 6= 0. There is one detail I skipped, namely the
result
dt 1
= dx .
dx dt
This is not because of the properties of fractions. It is due to the derivative of inverse
functions that we have seen before. In fact using the notation above
dt d −1 1 1
= φ (x) = 0 = .
dx dx φ (t) ẋ(t)
Recall that in taking the derivative of the inverse function and writing it in terms of the
original function, we evaluate the latter in its domain, i.e. the domain of φ−1 is x and
that of φ is t. See previous chapters.
The tangent to the curve always exists as long as ẋ and ẏ do not vanish at the same
point, i.e.
ẋ2 + ẏ 2 6= 0, (21.13)
If ẋ = 0, ẏ 6= 0, then the tangent is vertical, and if ẋ 6= 0, ẏ = 0 the tangent is horizontal.
Challenge Problem
Show the cusp structure (21.16) found above parametrically BUT starting from the
implicit equation of y as a function of x given by (21.10). You will only need the result
2
cos ξ = 1 − ξ2! + . . . given above when ξ is small.
Next consider a point on the parametric curve that corresponds to t = t0 , i.e. consider
the point P0 = (x0 , y0 ) where x0 = φ(t0 ) and y0 = ψ(t0 ). We wish to find the equations
for the tangent and normal directions to the curve at P0 . The equation for the tangent
is
dy
y − y0 = (x − x0 ). (21.17)
dx P0
At points where ẋ2 + ẏ 2 = 0 (i.e. ẋ and ẏ are zero simultaneously), equation (21.18) does
not make sense. Geometrically we would be trying to find a tangent at a point such as
a corner or a cusp.
The normal to the curve is a straight line perpendicular to the tangent, i.e. it has
slope −ẋ/ẏ at P0 . Hence the equation is
As an example, consider Exercise 2 in section 21.3, i.e. the curve x = t2 −1, y = t3 −t,
for −∞ < t < ∞. We see that when t = −1 and t = +1, we obtain x = y = 0, that is
the curve crosses itself at the origin. The tangent directions are different at t = −1 and
t = +1 as can be seen by calculation:
t = −1 x+y =0
t = +1 x − y = 0.
Finally note that if α is the angle the tangent makes with the horizontal, then
dy ẏ
= = tan α. (21.20)
dx ẋ
184 CHAPTER 21. THEORY OF PLANE CURVES
dy dy dt ẏ
= =p = sin α. (21.28)
ds dt ds ẋ + ẏ 2
2
21.5. CURVATURE 185
2 2
dx dy
+ = 1, (21.29)
ds ds
21.5 Curvature
The curvature at a point of a curve is defined to be the rate at which the angle of
inclination α (the angle that the tangent at the point makes with the horizontal) changes
with arc length. This can be expressed as
dα
κ= . (21.30)
ds
ẏ
tan α = if ẋ 6= 0, and (21.31)
ẋ
ẋ
cot α = if ẏ 6= 0. (21.32)
ẏ
186 CHAPTER 21. THEORY OF PLANE CURVES
K =
IF
.
We can differentiate either of (21.31), (21.32) and the result is the same. Will do it
for (21.31) with implicit differentiation and use of the quotient rule, i.e.
dα ẋ ÿ − ẍ ẏ ẋ ÿ − ẍ ẏ ẋ2 ẋ ÿ − ẍ ẏ ẋ ÿ − ẍ ẏ
sec2 α = ⇒ α̇ = cos2 α = = 2 .
dt ẋ2 ẋ2 2
ẋ + ẏ 2 ẋ 2 ẋ + ẏ 2
Using the formula for ṡ in equation (21.25) and the result above, we have
dα α̇ ẋ ÿ − ẍ ẏ
κ= = = 2 . (21.33)
ds ṡ (ẋ + ẏ 2 )3/2
If the arc length s is the parameter, then the dot means an s−derivative, and on use of
(21.29) we have
κ = xs yss − ys xss .
Note: The sign of κ is positive if α increases as s increases and negative if α decreases
with increasing s. Hence, κ > 0 if the function is convex and κ < 0 if it is concave. See
Figure 21.4.
Next, consider the curvature of the curve given by the function y = f (x). In this
case x is the parameter and so the expression (21.33) gives
y 00
κ= . (21.34)
(1 + y 02 )3/2
This formula shows that κ > 0 if y 00 > 0 and such curves are called convex. On the other
hand κ < 0 if y 00 < 0 and we define such curves as concave. At a point where y 00 = 0 we
have a change in sign of κ, i.e. a point of inflexion that is already familiar from earlier
work.
21.5. CURVATURE 187
T÷÷¥÷÷÷¥ I •X
at
.
CENCI
it •
Figure 21.4: Positive κ and negative κ for convex functions (left) and concave functions
(right).
Problems:
189
190CHAPTER 22. LAPLACE TRANSFORMS AND APPLICATIONS, THE BRACHISTOCHRONE A
R∞ p
8. f (x) = cosh ax, F (p) = 0 e−px cosh ax dx = p2 −a2
. Converges for p > a.
Note that we can use the linearity of L and the results above to write down Laplace
transforms of simple functions. Here is an example:
1 1 1 1 1 p
L cos2 ωx = L
+ cos 2ωx = L [1] + L [cos 2ωx] = + . (22.3)
2 2 2 2 2p 2(p2 + 4ω 2 )
Problem: Use the result (22.3) and without integrations calculate L[sin2 ωx].
Solution: Since sin2 ωx + cos2 ωx = 1, we have
1 1 p 1 p
L[sin2 ωx] = L[1 − cos2 ωx] = − − = − .
p 2p 2(p2 + 4ω 2 ) 2p 2(p2 + 4ω 2 )
Remark: Piecewise continuous functions that are also of exponential order (see below
for the definition), have Laplace transforms. The intuitive reason is that integrals of
piecewise continuous functions over finite domains exist. Therefore, we can deal with
numerous functions found in applications, e.g. the Heaviside function (or step function),
the square function found in electrical circuits etc. See problems for these.
R∞
The existence of L[f (x)] centres on the existence of the improper integral 0 e−px f (x)dx.
This has been discussed in general already and essentially we need to consider the inte-
grand e−px f (x) for large x. If f (x) is of exponential order then the integral exists as we
see next.
Definition
f (x), x > 0, is said to be of exponential order if there exist constants M and c
such that
Problems
1. Sketch each of the following functions and find their Laplace transforms (it is
understood that in all cases x > 0):
(i) f (x) = H(x − a) where a > 0 and the Heaviside function H(x) is defined by
0 x<0
H(x) =
1 x≥0
(ii) f (x) = [x] where [x] is the integer part of x (e.g. [π] = 3).
(iii)
sin x 0 ≤ x ≤ π
f (x) =
0 x>π
y 00 + ay 0 + by = f (x), (22.5)
Here is an example. Using L−1 [1/p2 ] = x we have L−1 [1/(p + 1)2 ] = e−x x.
Problems
6 p+3 1
1. Find the inverse Laplace transforms of (i) (p+2)2 +9
, (ii) p2 +2p+5
, (iii) (p+3)4
.
and differentiate with respect to p under the integral sign (permissible since the integral
is absolutely convergent) to find
Z ∞
0
F (p) = e−px (−xf (x))dx ⇒ L[−xf (x)] = F 0 (p).
0
Solved Examples
1. Find L[x sin x]. We use L[sin x] = 1/(p2 + 1) and the result above to find
d 1 2p
L[x sin x] = − 2
= 2 .
dp p + 1 (p + 1)2
Next we consider the use of Laplace transforms in solving differential equations that
do not have constant coefficients. The methods will be illustrated with Bessel’s1 equation
of order zero, namely
xy 00 + y 0 + xy = 0. (22.13)
Equation (22.13) is second order and so two independent solutions exist. One of them
can be taken to satisfy y(0) = 1 (i.e. y0 = 1 in our notation) and we will try to find it.
Taking the Laplace transform of (22.13) we need to calculate
d d 2 d 2
L[xy 00 ] = − L[y 00 ] = −
p Y − py0 − y1 = − p Y − py0 , (22.14)
dp dp dp
d d
L[xy 0 ] = − L[y 0 ] = − (pY ), (22.15)
dp dp
d dY
L[xy] = − L[y] = − . (22.16)
dp dp
dY dY p dp
(p2 + 1) = −pY ⇒ =− 2 ,
dp Y p +1
where k is a constant. The transform (22.17) is not one that we recognise! You will
learn contour integration techniques that can deal with such inverses (they include branch
cuts), but for our purposes we wish to use elementary techniques to get a solution. One
way to do this is to re-write (22.17) in the form
1 −1/2
k
Y = 1+ 2 , (22.18)
p p
formally and take the inverse transform of (22.19) term-by-term. The result is
∞
(−1)n 2n x2 x4 x6
X
y(x) = k x = k 1 − 2 + 2 2 − 3 2 2 + ··· . (22.20)
22n (n!)2 2 4 ·2 4 ·3 ·2
n=0
1
Bessel’s equation of order α is x2 y 00 + xy 0 + (x2 − α2 )y = 0, and clearly α = 0 gives (22.13). Bessel
functions arise in the theory of oscillations of hanging chains (Daniel Bernoulli), vibrations of circular
membranes (Euler), planetary motion (Bessel), as well as numerous modern applications such as wave
propagation, elasticity, fluid dynamics, potential theory and diffusion, as well in some pure mathematics
problems.
22.3. DERIVATIVES AND INTEGRALS OF LAPLACE TRANSFORMS 195
Since y(0) = 1 we have k = 1 and so we have found the power series of the zeroth order
Bessel function of the first kind, namely,
x2 x4 x6
J0 (x) = 1 − 2 + 2 2 − 3 2 2 + · · · . (22.21)
2 4 ·2 4 ·3 ·2
The function J0 (x) is plotted in figure 22.1. It is seen that it oscillates infinitely often
and decays to zero for large x. The zeros of the function form an infinite sequence
{z1 , z2 , z3 , · · · } and it can be shown that limn→∞ (zn+1 − zn ) = π, i.e. the difference
between consecutive zeros tends to π. This is similar to sin x and cos x, but in addition
we have a decay of J0 (x) as x → ∞. The radius of convergence of the power series
(22.19) and (22.21) are left as an exercise - see Problem 1 below.
0.5
J 0(x)
-0.5
0 20 40 60 80 100
x
Figure 22.1: Zero order Bessel function of the first kind.
A final remark is that the solution (22.21) can be found using a power series solution
without going in Laplace space. The alternative calculation presented here is more
straightforward algebraically but is of course fully equivalent.
Problems
1. Find the radius of convergence of the function Y (p) in equation (22.19) and of the
function J0 (x) in equation (22.21).
2. This problem is concerned with showing the result
Z ∞
f (x)
L = F (p0 )dp0 , F (p) = L[f (x)], (22.22)
x p
196CHAPTER 22. LAPLACE TRANSFORMS AND APPLICATIONS, THE BRACHISTOCHRONE A
5. If f (x) is periodic of period a (i.e. f (x + a) = f (x)), show that F (p) = L[f (x)] is
given by Z a
1
F (p) = e−px f (x)dx.
1 − e−ap 0
Find F (p) is f (x) = 1 in the intervals 0 to 1, 2 to 3, 4 to 5, etc., and f (x) = 0 in
the remaining intervals, i.e. 1 to 2, 3 to 4, etc.
This is very useful as we will see below, because given a product of Rtwo recognisable
x
Laplace transforms, their inverse transform is the convolution integral 0 f (x − t)g(t)dt.
And vice vera.
We want to prove this result. Start with
Z ∞ Z ∞
−ps −pt
F (p)G(p) = e f (s)ds e g(t)dt
0 0
Z ∞ Z ∞
−p(s+t)
= e f (s)ds g(t)dt. (22.24)
0 0
xt−plane. Hence for a given x doing the t integration first we have a range of integration
0 ≤ t ≤ x, while the x integration has a range x ≥ 0. The result follows since (22.25)
becomes
Z ∞ Z x Z x
F (p)G(p) = e−px f (x − t)g(t)dt dx = L f (x − t)g(t)dt . (22.26)
0 0 0
The result (22.23) is useful in solving integral equations. These are equations that
involve the unknown as well as its weighted integral. Here is a fairly general example of
an integral equation of the first kind. Given functions f (x) and K(x) we need to find
y(x) that satisfies Z x
y(x) + K(x − t)y(t)dt = f (x). (22.27)
0
In most integral equations arising from physical problems an added complication is that
K(x) is singular at x = 0 - we will see such an example later. Note also that it makes
no sense to differentiate (22.27) with respect to x since the integral term will not go
away2 (even assuming that K(0) < ∞). However, the convolution theorem proved
above becomes a useful tool to find solutions as the following example illustrates.
Solved Example We will find the solution of the integral equation
Z x
3
y(x) = x + sin(x − t) y(t) dt. (22.28)
0
3! 1 3! 3!
Y (p) = 4
+ Y (p) ⇒ Y (p) = + 6 (22.29)
p 1 + p2 p4 p
3! 5
y(x) = x3 + x . (22.30)
5!
The answer can be verified by plugging into (22.28).
to the material it is travelling through (e.g. water, air, glass etc.). Consider initially,
then, a ray of light entering one layer of thickness a at a point A1 , travelling along a
straight line (since the layer properties are constant) with speed v1 to reach point A2 ,
from where it enters a second layer of thickness b with different properties and travelling
along a straight line with speed v2 to reach the layer’s bottom surface at a point A3 .
The constant speeds v1 and v2 are different. A schematic of this scenario is given in
figure 22.2.
As
r
-
·
e e
a
I
⑧
M
layer 2 ix2
~ ~
!
⑭ I As
P
i
-
-
i
Next we calculate the time t it takes for the light to travel from A1 to A3 . Given the
geometry in figure 22.2 we have
√ p
a2 + x2 b2 + (c − x)2
t= + . (22.31)
v1 v2
Now we impose Fermat’s principle of least time which says that when light travels from
one point to another in layered media, it picks the path that gives the least time of travel.
For expression (22.31) this requires us to find where the point A2 should be to minimise
travel time, i.e. we need to find x and for a minimum we need to solve dt/dx = 0. This
gives
x c−x
√ = p , (22.32)
v1 a2 + x2 v2 b2 + (c − x)2
Inspection of the triangles A1 A2 P1 and A2 A3 P2 and trigonometry shows that equation
(22.32) is
sin α1 sin α2
= . (22.33)
v1 v2
This is Snell’s law named after Willebrod Snell (1591-1626) who was a Dutch mathe-
matician and astronomer.
22.5. THE BRACHISTOCHRONE 199
In a medium that has continuously variable density we use calculus to split it into
layers of smaller and smaller thickness. If we have n layers then Snell’s law says that
and in the limit as n → ∞ we have a continuous ray path. At any point on the path the
tangent is at an angle α to the vertical and the speed is v such that
sin α
= const. (22.35)
v
n
VI
limit
!
"
y a
-
-
" i wo ⑧
I
-
!
V
Figure 22.3: Schematic of Snell’s law for n layers and the limit n → ∞.
sin α
= const., (22.36)
v
but we can also find v from the principle of conservation of energy: the bead has de-
scended a distance y and assuming it started from rest (and since there are no frictional
forces) we have
1
mgy = mv 2
p
⇒ v = 2gy. (22.37)
2
200CHAPTER 22. LAPLACE TRANSFORMS AND APPLICATIONS, THE BRACHISTOCHRONE A
A
⑧ >x
m
P
·
⑰
y~ ⑧
B
π
Looking at figure 22.4 we know that the tangent to the curve makes an angle 2 −α
to the horizontal, and hence
cos α 1 1
y 0 = tan(π/2 − α) = i.e. tan α = ⇒ sin α = p . (22.38)
sin α y0 1 + (y 0 )2
Substituting (22.37) and (22.38) into (22.36) provides a single equation for y, namely
y 1 + (y 0 )2 = k,
(22.39)
k−y
where k is a constant. To solve (22.39) we note that (y 0 )2 = y and separating variables
we have √
y
√ dy = dx. (22.40)
k−y
Change variables by introducing
√
y
√ = tan φ, (22.41)
k−y
and solving we find that y = k sin2 φ, hence dy = 2k sin φ cos φ dφ, casting (22.40) into
Integrate to find
sin 2φ
x=k φ− + `. (22.43)
2
22.6. ABEL’S MECHANICAL PROBLEM AND ABEL’S INTEGRAL EQUATION201
k k
x= (2φ − sin 2φ), y = k sin2 φ = (1 − cos 2φ). (22.44)
2 2
We have encountered the cycloid in a different context in Section 21.3 where it was
constructed as the locus of a point on the circumference of a circle that is rolling in
a straight line. The connection to the brachistochrone is that given the point B, the
constant a is chosen to obtain a cycloid (its first arch) that passes through B. This can
always be done.
In section 22.5 we asked the question: What should the shape of a wire be for a particle to
descend a given vertical height in the smallest possible time. Abel’s mechanical problem
asks a different question: For a given height specify the time of descend (without loss
of generality to the origin y = 0), and find the shape of the wire that will achieve this.
This is a very different problem as we will see, and belongs to a class of inverse problems
that do not always have solutions.
To analyse this problem begin with defining the starting point to be (x, y) and the
end point to be the origin (0, 0). What we are given is the time T (y) and we need to
find the shape of the wire y ≡ y(x) that achieves this time of descend. Think of T (y) as
a prescribed function - it is not a constant, different y give different times.
Let any point on the curve between the starting and end points, have coordinates
(ξ, η). A schematic is given in figure 22.5. As shown in the figure we define s to be the
arc length of the curve y(x) measured from the origin (or end point).
202CHAPTER 22. LAPLACE TRANSFORMS AND APPLICATIONS, THE BRACHISTOCHRONE A
1.8
1.6
1.4
1.2
1 (x,y)
y
o
0.8
0.6
( , )
0.4 o
0.2 s
0
0 0.5 1 1.5 2 2.5 3 3.5
x
Figure 22.5: Schematic of Abel’s mechanical problem.
Use conservation of energy. Since the particle of mass m, say, starts from rest, when
it gets to the point (ξ, η) it will have a kinetic energy equal to the potential energy lost
by the descend. This implies
2
1 ds
m = mg(y − η), (22.46)
2 dt
and on taking a square root and picking the minus sign (since the particle is moving
from left to right) we have
ds p ds
= − 2g(y − η) ⇒ dt = − p . (22.47)
dt 2g(y − η)
We need to integrate this equation between 0 and y; the integration variable is ξ. Hence
Z y Z y ds dη
ds 1 dη
T (y) = p =√ √ . (22.48)
0 2g(y − η) 2g 0 y−η
and equation (22.51) should be understood as defining a function f (η) or f (y) or what-
ever independent variable I choose to use.
Putting these elements together into (22.48) gives the equation
Z y
1 f (η)dη
T (y) = √ √ . (22.52)
2g 0 y−η
The easy problem is to find T (y) when y(x) is given. This immediately gives us f (η)
and the problem becomes one of carrying out a singular integral that exists.
To go the other way, that is to find y(x) given T (y) in (22.52), we use Laplace
transforms and the results we found in earlier sections. The first thing to note is that
the integral on the rhs of equation (22.52) is the convolution of the functions y −1/2 and
f (y). Taking the Laplace transform of (22.52) and using the convolution Theorem we
obtain √
1 −1/2 1 π
L[T (y)] = √ L[y ]L[f (y)] = √ √ L[f (y)], (22.53)
2g 2g p
and keeping track of what we want to calculate, namely y(x), we re-write the equation
above as r
2g √
L[f (y)] = p L[T (y)]. (22.54)
π
In principle, given the function T (y) we can find its Laplace transform and then by
inversion find f (y). With f (y) known we next need to solve the nonlinear differential
equations (22.51). All this cannot be done in general - numerical methods are needed.
We will solve, however, another classical problem that arises when T (y) = T0 where
T0 is a constant. What this says physically, is that we want to find the shape of the wire
so that the time taken to descend to the origin from any starting point is the same! This
curve has a name, the tautochrone. In this case (22.54) becomes
r
2g −1/2
L[f (y)] = T0 p , (22.55)
π
√
and since the inverse transform of p−1/2 is πy −1/2 (we did this calculation earlier), we
find √
f (y) = by −1/2 , where b = 2gT02 /π 2 . (22.56)
This in turn implies from (22.51) that
2 Z 1/2
dx b b−y
1+ = ⇒ x= dy, (22.57)
dy y y
204CHAPTER 22. LAPLACE TRANSFORMS AND APPLICATIONS, THE BRACHISTOCHRONE A
where the integral has been left as indefinite for now. Substituting y = b sin2 φ turns the
integral into Z
b
x = 2b cos2 φ dφ = (2φ + sin 2φ) + c, (22.58)
2
where c is a constant. Hence the parametric form of the curve is
b b
x = (2φ + sin 2φ) + c, y = b sin2 φ = (1 − cos 2φ). (22.59)
2 2
We have chosen the curve to pass through the origin (0, 0) and this can happen when
φ = 0 and c = 0. Putting 2φ = θ and a = (b/2) = gT02 /π 2 gives the familiar cycloid
again
x = a(θ + sin θ), y = a(1 − cos θ). (22.60)
In fact in figure 22.5 I plotted half a branch of the cycloid for a = 1. Note that any
starting point in space can be found by choosing a appropriately.
Problems
1. Check the Abel integral equation (22.52) for the special case of finding the time of
descend of a particle of mass m from the point (2, 1) to the origin along a straight
wire joining the two.
√
2. If T (y) = k y where k is a constant, compute the curve of descend.
3. Start with equation (22.54) and show that it can be written as
√ Z y
2g d T (χ)dχ
f (y) = √ .
π dy 0 y−χ
where a is a real constant. Show using Laplace transform methods that the solution
is
1 x
Z
y(x) = f (x0 ) sin[a(x − x0 )]dx0 .
a 0
Chapter 23
205
206CHAPTER 23. NEWTON’S LAW OF GRAVITATION, PLANETARY MOTION AND KEPLER’S
The force acting on the particle can be resolved into components in the r and θ directions
as follows
F = Fr er + Fθ eθ . (23.6)
From Newton’s Second Law of motion we have
ma = F,
Equations (23.7)-(23.8) are general in the sense that the nature of F has not been
specified. Some additional physics is needed to specialise these to planetary motion as
we do next. Note also that the equations are nonlinear for the unknowns r(t) and θ(t);
if a solution can be found, then the trajectory of P under the action of the force F will
have been determined.
d2 θ
dr dθ d 2 dθ
0=r 2 +2 = r , (23.9)
dt dt dt dt dt
and on integration we have
dθ
r2 = h, (23.10)
dt
where h is a constant. If the mass m moves in the positive θ−direction (i.e. counter-
clockwise), then h > 0.
From equation (23.10) we can put Kepler’s Second Law that was based on obser-
vations on a sound mathematical footing. Kepler’s Second Law states that as a planet
moves around the sun (in our notation the origin O), the radius vector r sweeps out
equal areas in equal intervals of time. This in turn implies that when a planet is closer
to the sun it will travel faster but the precise speed is dictated by Kepler’s Second Law.
Consider, then, the point P being at the position (r, θ) at a given time t. After a
small time dt it sweeps through an angle dθ to the new position (r + dr, θ + dθ). As we
saw in Section 12.6, the swept area dA is given by
1 1 dθ 1
dA ≈ r2 dθ = r2 dt = h dt, (23.11)
2 2 dt 2
23.3. NEWTON’S LAW OF GRAVITATION - CENTRAL GRAVITATIONAL FORCES207
1
A(t2 ) − A(t1 ) = h(t2 − t1 ). (23.12)
2
Equation (23.12) is exactly Kepler’s Second Law. Kepler’s First Law states that the
motion of a planet about the sun traces out an ellipse, with the sun at one of the
ellipse’s eccentricity. We will derive this mathematically also. 1
Equation (23.14) is nonlinear but we can turn it into a linear problem by introducing a
new variable z = 1/r and also using θ as the independent variable instead of t. From
(23.10) we have dθ/dt = hz 2 , and the chain rule is used to compute
dr d 1 dz 1 dz dθ dz
= (1/z) = − 2 =− 2 = −h , (23.15)
dt dt z dt z dθ dt dθ
d2 r d2 z dθ 2 2 d2z
= −h = −h z . (23.16)
dt2 dθ2 dt dθ2
1
Historical note: Observational astronomy went on for thousands of years until Kepler finally
quantifies the hidden beauty of planetary motion that had occupied so many scientists from ancient
times onwards. Kepler (1571-1630) was the assistant of the astronomer Tycho Brahe who died in 1601,
and so he had access to huge amounts of raw observational data. It took him twenty years of constant
work to extract his three elegant laws from a zoo of data. It is probably the first and largest achievement
of physics-informed data science but with no computers.
2
To get an idea we can estimate |Fr | in (23.13); the mass of the earth is m = 5.972 × 1024 kg, and
the average distance of the earth from the sun is about r = 1.5 × 1011 m, hence |Fr | ≈ 3.5 × 1022 kg m s−2
(note that 1 Newton is 1 kg m s−2 . For comparison, if you go to a depth of 99 m of water, the pressure
you feel is 10 bar which is 107 N m−2 , so |Fr | is about 1015 times larger than the force exerted by the
100 m of water on an area of 1 square meter.
208CHAPTER 23. NEWTON’S LAW OF GRAVITATION, PLANETARY MOTION AND KEPLER’S
The next component in the energy equation is the potential energy stored by the mass
m due to its presence in the gravitational field of the sun. Now at infinity (i.e. very
far away from the sun at the focus) the potential energy is zero. To find the potential
energy at a finite distance r from the focus we need to calculate the work done to move
the mass from r to infinity. The attractive gravitational force is given Fr = −km/r2 (see
(23.13)), and so the work done against this field is the potential energy and is given by
Z ∞
km km
PE = − 2
dr = − . (23.22)
r r r
By the principle of conservation of energy E = KE + P E is a constant, therefore we
have " 2 #
2
1 dθ dr km
m r2 + − = E. (23.23)
2 dt dt r
3
The polar curve r = 1+εpεcos θ is a conic section with eccentricity ε and distance p between the focus
and the directrix. Hence in our case we have ε = Bh2 /k and p = 1/B.
23.3. NEWTON’S LAW OF GRAVITATION - CENTRAL GRAVITATIONAL FORCES209
Our objective now is to connect the energy E, a constant for a given trajectory, with the
eccentricity ε. Evaluating (23.23) at the instant of the trajectory when θ = 0 and the
planet is at its closest approach to the sun, we immediately see that dr/dt = 0 there.
We also have the fundamental expression (23.10); using this gives r2 (dθ/dt)2 = h2 /r2 ,
and hence
and this can be readily solved (luckily the linear term in ε cancels) to give the eccentricity
r
2h2
ε= 1+ E, (23.25)
mk 2
To find the period of a planet, that is the time it takes to complete one full elliptical
orbit, we will use Kepler’s second law that we gave as equation (23.12). If the period is
T , then (23.12) says that hT
2 = A where A is the area of the ellipse. Writing the ellipse
in Cartesian coordinates we have
x2 y 2
+ 2 = 1. (23.26)
a2 b
From analytic geometry we have the following facts connecting the eccentricity ε and
the distance c of the focus from the Cartesian origin:
c a2 − b2
c2 = a2 − b2 , ε= ⇒ ε2 = , b2 = a2 (1 − ε2 ). (23.27)
a a2
nY
-m
b
↓
--
·
O
>
·
F i
- a-
We need to express a and b in terms of h and k. From the figure we see that
1 1
a = (rmin + rmax ) = [r(0) + r(π)]. (23.28)
2 2
The quantity a is termed the mean distance of the planet. Using the solution (23.20) we
calculate
1 h2 /k h2 /k h2 h2 a2
a= + = 2
= , (23.29)
2 1+ε 1−ε k(1 − ε ) kb2
where the last equality follows from (23.27). Solving for b we find
h2 a
b2 = . (23.30)
k
We have shown that T = 2A/h and given that for our ellipse A = πab we have
2πab 4π 2 a2 b2 4π 2 3
T = ⇒ T2 = = a , (23.31)
h h2 k
where the last equality is due to (23.30).
In the formula (23.31) the constant k = GM is the same for all planets since it only
depends on the mass of the sun (the central mass) alone. Formula (23.31) is Kepler’s
Third Law that says that the squares of the periods of the planets are proportional to
the cubes of their mean distances.
Problems
1. Show that the speed v of a planet at any point of its orbit is given by
2 2 1
v=h k − .
r a
Part VI
Miscellaneous Problems
211
213
7. (a) Let f (x) = xa for x > 0 and a any real number. Prove that f 0 (x) exists and
is equal to axa−1 . [Hint: Write as an exponential function.]
d
(b) We know that dx log x = x1 . Show that d
dx loga x = 1
x log a .
[Hint: Let y = loga x so that x = ay .]
8. Find the equation of the tangent line at (0, log 3) to the graph of the curve defined
implicitly by ey − 3 + log(x + 1) cos y = 0.
214
x2 + 1 2x
lim = lim =0
x→0 x x→0 1
10. Consider the step-ladder function f (x) = 1 + [x] for x ≥ 0, where [x] denotes the
integer part of x. (For example [0.9] = 0, [2.1] = 2, etc.)
(a) Sketch f (x).
Rx
(b) Calculate and sketch F (x) = 0 f (t)dt.
R1
11. Need to calculate 0 ex dx. (Of course the answer is e − 1.)
The Upper Riemann sum is Un = ni=1 ei/n n1 .
P
13. An inverted conical tank of base radius 6m and height 10m is full and is to be
pumped empty (from the top) by a pump that has power output of 105 joules per
hour (this is approximately 27.77 watts).
(a) What is the water level at the end of 6 minutes of pumping?
(b) How fast is the water level dropping this time?
R ∞ dx
14. Show that 0 √1+x 8
is convergent, by comparison with 1/x4 .
2
15. (a) Calculate the integral (1+xx2 )3/2 dx using (i) a trigonometric substitution, (ii)
R
integration by parts.
R π/2
(b) Calculate 0 cos7 xdx.
16. An object moves from left to right along the curve y = x3/2 at constant speed. If
the object is at (0, 0) at noon and at (1, 1) at 1:00 pm, where is it at 1:30pm?
17. The region between the graphs of sin x and x over the interval 0 ≤ x ≤ π/2, is
revolved about the y−axis. Sketch the resulting solid and find its volume.
18. Find the area of the surface of the solid formed by rotating the curve y = x2 ,
0 ≤ x ≤ 1, about (i) the x− axis, and (ii) the y-axis.
19. (i) Consider the particular case of the centre of mass of a plate of area A that
has unit density per unit area. Write down formulas for the centre of mass in
terms of double integrals.
(ii) Now take A to be the region 0 ≤ x ≤ 1, 0 ≤ y ≤ 1 (unit density also).
Compute the double integrals to show that the centre of mass is (1/2, 1/2) as
expected by intuition.
215
(iii) Now take A as in part (ii) but assume that the density in half the square
between 0 ≤ x ≤ 1/2 is 1, and the density in the remaining half of the square
is ρ0 . Compute the centre of mass in this case.
(iv) Check that your formula in (iii) agrees with your answer in (ii) when ρ0 = 1.
What happens as ρ0 tends to 0 and ∞? Explain using physical intuition.
20. (a) Consider the functions f (x) = xm and g(x) = xn where 0 < m < n. Find a
condition satisfied by m and n that guarantees that the centre of mass of the
region between the graphs y = f (x) and y = g(x) is outside the region.
(b) We know from Example 2 from Recording 20 that m = 1, n = 2 has the
centre of mass within the region. What are the smallest integer pair m, n for
which the centre of mass is outside the region?
21. Use the results developed in Recording 21 to show that a circle of radius a has
perimeter 2πa and area πa2 .
eαn
22. (a) For what real values of α (if any at all) does the series ∞
P
n=1 n converge?
(b) Calculate ∞ 1
P
n=100 n1/100 .
(c) Let P = K
P n
PK n
n=1 3 and Q = n=1 (1/3) . Find P , Q and P Q.
Which ones converge as K → ∞? Could you have anticipated this without
calculations?
(−1)n log n
23. (a) Does the series ∞
P
n=1 n2
converge?
P∞ log n
(b) Does the series n=1 n2 converge?
(−1)n log n
(c) Is the series ∞
P
n=1 n convergent? Is it absolutely convergent?
P∞ 1
24. (a) Show that n=1 n2 +1 converges by using (i) the comparison test, (ii) the
integral test.
√
(b) Does the series ∞ 3n+ n
P
n=1 2n +2 converge or diverge?
3/2
(−1)n x2n+1
26. Find the radius of convergence of the power series ∞
P
n=0 2n+1 .
28. Use Taylor’s theorem to find a second order accurate formula for the finite differ-
ence approximation of f (2) (x) using information from the points f (x), f (x ± h).
29. Use Taylor series to compute 1/e to three decimal accuracy. Compare the number
of terms needed with example 3 in Recording 29. Explain.
√
30. Find the first four non-zero terms in the Maclaurin expansion of log 1 + sin x .
216
31. The Chebyshev4 polynomials of the first kind are defined by Tn = cos n cos−1 x ,
where n = 0, 1, 2, . . ..
(a) Show that
(b) Show that the Chebyshev polynomials Tn (x) are orthogonal on the interval
1
[−1, 1] with respect to the weight function w(x) = √1−x 2
.
(c) Verify that Tn (x), n = 0, 1, 2, . . ., satisfy the Chebyshev differential equation
d2 y dy
(1 − x2 ) 2
−x + n2 y = 0.
dx dx
32. Find and sketch the periodic extensions of
(a) f (x) = x2 , defined on x ∈ [−π, π]. Call this function g1 (x).
(b) f (x) = ex , defined on x ∈ [0, 1]. Call this function g2 (x).
R 7π R 21/2
(c) Calculate 0 g1 (x)dx and 1/2 g2 (x)dx.
33. Express f (x) = sin(x + π/4) + sin2 x − cos2 x as a complex trigonometric series.
34. Give examples of functions f (x) that are periodic of period 2π and are:
(i)
Smooth (i.e. infinitely differentiable).
(ii)
Piecewise continuous.
(iii)
Piecewise continuous with piecewise continuous first derivative.
(iv)Piecewise continuous with piecewise continuous first derivative but not second
derivative.
(v) As nasty a function as you can think of whose Fourier series exists.
35. Find directly the Fourier series of φ(x) = x, −π < x < π without recourse to shifts
and reflections used in class.
36. (a) For the function f (x) = ex defined on [0, π], find the cosine series and the
sine series. Call them Sc (x) and Ss (x).
(b) Discuss convergence of Sc and Ss for all values of x ∈ [−π, π].
d d
(c) Now consider dx (Sc ) and dx (Ss ). Which one (if any) converges and why?
37. Sketch the function f (x) given below and find its Fourier series:
0 −3 ≤ x < −2
−x − 1 −2 < x ≤ 1
f (x) = 0 −1 ≤ x < 1
1 − (x − 1)2 1<x≤2
0 2≤x≤3
4
Pafnuti Chebyshev (1821-1894), Russian mathematician. Chebyshev polynomials are used widely in
solutions of partial differential equations, including aerodynamics, biofluiddynamics, materials, waves,
...
217