You are on page 1of 120

KENYATTA UNIVERSITY

INSTITUTE OF OPEN DISTANCE & e-LEARNING

IN COLLABORATION WITH

SCHOOL OF PURE & APPLIED SCIENCES

DEPARTMENT: PHYSICS DEPARTMENT

UNIT CODE & NAME SPH 300 – WAVE THEORY

WRITTEN BY:
Eng. John Namale

i
Copyright © Kenyatta University, 2009
All Rights Reserved
Published By:
KENYATTA UNIVERSITY PRESS

ii
TABLE OF CONTENTS PAGE
LECTURE ONE............................................................................................................................1
PERIODIC MOTION AND THEIR SUPERPOSITION......................................................1
LECTURE TWO.........................................................................................................................21
FORCE & ENERGY...............................................................................................................21
LECTURE THREE.....................................................................................................................28
COMPLEX VARIABLES REPRESENTATION OSCILLATIONS..................................28
LECTURE FOUR........................................................................................................................36
FREE AND FORCED VIBRATIONS...................................................................................36
LECTURE FIVE.........................................................................................................................70
FOURIER ANALYSIS............................................................................................................70
LECTURE SIX............................................................................................................................82
COUPLED OSCILLATORS AND NORMAL MODES......................................................82
LECTURE SEVEN......................................................................................................................97
THE FREE VIBRATIONS OF STRETCHED STRINGS...................................................97
LECTURE EIGHT....................................................................................................................104
DIFRACTION OF LIGHT AND WATER WAVES............................................................104
LECTURE NINE.......................................................................................................................110
REFLECTIONS OF MECHANICAL (SHEAR) WAVES AT A DISCONTINUITY......110

iii
LECTURE ONE

PERIODIC MOTION AND THEIR SUPERPOSITION

1.1 Introduction

This deals with periodicity of oscillatory motion of bodies. In this case sinusoidal kind of
motion is considered. Also discussed here is the combination of motions of one or more
bodies especially in one medium.

1.2 Lecture Objectives

By the end of this lecture, you should be able to:


 Differentiate between periodic, oscillatory and simple harmonic motions
 Understand the concept of wave superposition in one dimension
 Derive equations of superposed equations with equal frequencies, different
frequencies (beats) and for many superposed vibrations of same frequency.
 Combine two vibrations at right angle.
 Draw Lissajous figures.
 Compare perpendicular and parallel superposition.

1.3 Periodic motion and their superposition.

Periodicity is a pattern of movement or displacement that repeats itself over and over again. This
pattern may be simple or complex. On the other hand when a body moves back and forth
repeatedly about a mean position, its motion is called Oscillatory.
The motion of the planets around the sun is periodic motion i.e. it takes equal time to go once
around the sun. But the motion is not oscillatory as the planets move in one direction only.
Alternatively oscillatory motion is exemplified in the pendulum motion. The pendulum motion
moves in one direction and then turns back, crosses the mean point (equilibrium) position or
mean position and then goes to the other side, and turns back again. This motion goes on
repeating indefinitely.

1
The oscillatory motion may also be periodic if it takes equal time to come back to its original
position after completing the to and fro motion once. Such a periodic motion is called harmonic
motion. Thus, a harmonic motions is periodic but not all periodic motion are harmonic.
If a system is disturbed from its position of equilibrium and left to itself, in the absence of
friction the harmonic oscillatory it executes about the mean position of equilibrium, is called
simple harmonic oscillation.

1.3.1 Sinusoidal Vibrations


Our attention will be focused to sinusoidal vibration. For, instance we have a body attached to a
spring, the force exerted on it at a displacement x from equilibrium is
F (x) = - (k1x + k2x2 + k3x3 + …………..
Where k1 ,k2 ,k3 etc are a set of constants, and we can always find a range of values for X within
which the sum of the terms x2, x3,----- xn is negligible, according to some stated criterion (e.g., 1
part in 103, or 1 part in 106) compared to the term – kx, unless k, is equal to zero. If the body of
mass m and that of the spring is negligible, then SHM restoring force on the particle is expressed
as
F = - kx (1.1)
where x denotes the displacement.
But applying Newton’s second law to the motion of the body
F = ma
Where a is acceleration of the particle executing SHM
d 2x
Since a= 2
dt
d 2x
F m (1.2)
dt 2

From (i) and (ii)

d 2x
m   kx
dt 2
d 2x k
or 2
 x
dt m
k
Putting  2
m

d 2x
  2 x
dt 2

2
d 2x
Or 2
 2x  0 (1.3)
dt
Eqn. (1.3) is the differential equation of a SHM.
A simple harmonic motion can be expressed in terms of sine or cosine function. So the
displacement of a body executing a simple harmonic motion can be represented either by a
cosine curve or a sine curve as shown in Fig.1.1

Fig.1.1:Displacement time curve of a simple harmonic motion. (a) Cosine function (b) sine
function

For Fig 1.1 (a), we can write a simple relation for y and t
t
y = a cos 2π (1.4)
T
and for Fig. 1 (b), we can write similar expression,

t
y = a sin 2 π (1.5)
T
The equation (1.2) can also be written as:

 2t  
y = a sin    (1.6)
 T 2
Referring to Fig (1.1) we see that we may shift the curve figure 1.1(a) by T/4 forward in the
time-scale to make it fit with the curve figure 1.1(b). It will be explained later that T/4 in the time

3
scale is related to a quantity called phase π/2, that is, π/2 is the phase change corresponding to
time change T/4. So a harmonic equation may be represented in general by: -

 2t 
y  a sin    (1.7)
 T 
Solution of equation 1.3 comes in the form
y = a sin (ω t + α) (1.8)
That equation (1.8) is a solution of equation (1.3) can be checked as follows:
Differentiating equation (1.6) twice, we get
d 2x
= - ω 2y (1.9)
dt 2
which is the same as the equation (1.3)
In all types of simple harmonic motion the displacement time curve gives a wave pattern as in
Fig.1.1 and the displacement is described by equation 1.7 or 1.8. This wave pattern has certain
distinctive characteristics defining the harmonic oscillation, example:

Amplitude: The maximum magnitude of the displacement of the oscillating body from its mean
position is known as the amplitude of oscillation. In Figure1.1(a) the displacement varies
between the limits + 5 and – 5. So the quantity 5 is the amplitude of harmonic oscillation.
Referring to equation 1.7 and 1.8, we find that y max = ± a as the value of the sine function varies
between + 1 and – 1. Thus the amplitude is a.

Period: The minimum interval of time at which the motion repeats itself is called the periodic
time or simply period. It is the time taken for completing one oscillation (T).

Frequency: The number of oscillations completed per unit time is called frequency, denoted by
v.
1
v =
T
The unit is Hertz (Hz) i.e. expressed as cycles per second or cps.
i.e. 1cps = 1Hz

Phase: In equation 1.8, the angle (ωt + α) is called angle or simple phase angle of the vibrating
particle at the instant t. It is denoted by ф.

Ф = (ω t + α)
For a change in phase angle 2π, the motion repeats itself. Note: A phase change of 2π
corresponds to a change in time scale by T.
Further at t = 0, ωt + α = α

4
Therefore, the quantity α is the initial phase of the vibrating particle and is called the phase
constant or the epoch.

Angular Frequency: It is obvious from the foregoing discussion that the phase change of
2π corresponds to one cycle or time T. Therefore, the quantity
2
 2v (1.10)
T
The unit of measurement of angular frequency is rad/ s.

1.3.2 Example
Write the displacement’s equation representing the following conditions obtained in a simple
harmonic motion: Amplitude = 0.01; Frequency = 600Hz and initial phase π/6.

Solution:
The general equation of SHM
x  asin  2/t  α 

We know frequency ν=1/T


Thus the general equation can also be written as;
x  asin  2t  α 
 0.01 sin  2 600 t  /6 
 0.01sin  3768t  π/6  metres

1.4 Superposition of waves


When the number of waves reaches simultaneously in any part of the space, their “effect” add
together according to the principle of superposition. The principle of superposition is
     
stated as: if y  

y1  y 2  y 3  ........... 

are displacement vectors due to the waves 1, 2, 3 --- at
point when they act separately then the net displacement at the point when all the waves act
together is given by the vector sum of the individual displacements i.e.

     
y   y1  y 2  y 3  ........... 
 
(1.11)

1.4.1 Superposed vibrations in one dimension


Superposition of vibrations is the resultant of two or more harmonic vibrations combined to give
a sum of the individual vibrations. If simple addition holds true, the system is said to be linear,
and most of our discussions will be confined to such systems

1.4.2 Two superposed vibrations of equal frequency


Suppose we have SHM’s described by the following equations:

5
x1 = A1cos (ω t + α1)
x2 = A2cos (ω t + α2)
There combinations are then as follows:
x = x1 + x2 = A1cos (ω t + α1) + A2cos (ω t + α2) (1.12)
The displacement can be expressed as a single harmonic vibration: (note A = A1 + A2 and
α = α1 + α2)
Therefore
x = A cos (ω t + α)
The result can be obtained geometrically as follows:

Fig 1.2: (a) Superposition of two rotating vectors of the same period. (b) Vector triangle for
constructing resultant rotating vector.

From Figure 1.2 (a) let OP1 be a rotating vector of length A1 making the angle (ω t + α1) with the
x axis at time t. Let OP2 be a rotating vector of length A2 at the angle
(ωt + α2). The sum of these is then the vector OP as defined by the parallelogram law of vector
addition. As OP1 and OP2 rotate at the same speed ω, we can think of the parallelogram OP1PP2
as the rigid figure that rotates bodily at this same speed. The vector OP can be obtained as the
vector sum of OP1 and P1P (the later being equal to OP 2). Since ∠N1OP1 = ωt + α1, and ∠KP1P =
ωt + α2, the angle between OP1 and P1P is just α1–α2. Hence we have
A2 = A12 + A22 + 2A1A2 cos (α2 – α1)
The vector OP makes an angle β [see Figure 1.2 (b)] with the vector OP1, such that
A sin β = A2 sin(α2 - α1)
and the phase constant α of the combined vibration is given by
α = α1 + β
NOTE: You may skip this part to read Chapter 3: Complex variable representation- if
you are not familiar with complex representation of variables.

6
Use of the complex exponential expression leads directly to the same results. The rotating vector
OP1 and OP2 are described by the following equation:

Z1  A1e j  t 1  Z 2  A1e j  t  2 


Hence the resultant is given by;
j  t   2 
Z  Z1  Z 2  A1e j  t  1   A2 e   (1.13)
Using the exponential form, which allows us to take out the common factor exp j(ωt + α1).


Z  e j  t 1  A1  A2 e j   2 1   2

Remembering that jθ is just an instruction to apply a positive rotation through the angle θ, we see
that the combination of terms in square brackets specifies that a vector of length A2 is to be
added at an angle (α2 - α1) to a vector of length A1, and the initial factor
exp[j(ωt + α1)] tells us that the whole diagram is to be turned to the orientation shown in Figure
1.2(b)
In general the values of A and α for the resultant disturbance cannot be further simplified, but the
special case in which the combining amplitudes are equal is worth noting. If we denote the phase
difference (α2 - α1) between the two vibrations and δ, then from the geometry of the vector
triangular in Fig1.2 (b) one can read off, more or less by inspection, the following results
β = 
2
A = 2A1 Cosβ = 2A1 Cos  2  (1.14)

Example.
Two waves x1 = A1cos (t + 1) and x2 = A2cos(t + 2) having the same frequencies are
combined together. Find the resultant equation, amplitude and phase angle.
Solution
Let the resultant wave be x.
i.e. x = x1 + x2
= A1cos (t + 1) + A2cos(t + 2) …………… (a)
= A1[costcos1- sintsin1] + A2[costcos2- sintsin2]
Expanding the brackets
= A1 costcos1- A1sintsin1 + A2costcos2- A2sintsin2
Collecting like terms together
= (A1cos1 + A2cos2) cost – (A1 sin1 + A2 sin2) sint …….. (b)

7
Let
A cos  = A1cos1 + A2cos2 ………………….……………… (c)

A sin = A1 sin1 + A2 sin2 …………………………........... (d)


Substituting Eqs. (c) and (d) into Eq. (2), we get
x = A cos  cost - A sin sint ……………………….. (e)
Eq. 5 is the resultant equation and can be rewritten as: (check rules below)
x = A cos(t + ) (resultant equation)
To get the resultant amplitude, we square Eqs. (c) and (d) and add them:
A2 = A12 +A22 + 2A1A2 cos(2 - 1) (resultant amplitude)
And dividing equation (d) by (c) we get the phase angle.
Asin  Acos = A1 sin1 + A2 sin2  A1 cos1 + A2 cos2
tan = (A1sin1 + A2sin2 )  (A1cos1 + A2cos2) (resultant phase angle)
Rules to note
sin(A + B) = sinA cos B + cosA sinB
sin(A - B) = sinA cos B - cosA sinB
cos(A +B) = cosAcosB – sinAsinB
cos(A - B) = cosAcosB + sinAsinB

1.4.3 Superposed vibration of different frequency; beats

Lets us now imagine that we have two vibrations of different amplitudes A 1, A2, and also of
different angular frequencies ω 1, ω 2. Clearly, in contrast to the preceding example, the phase
difference between the vibration is continually changing. The specification of some initial non-
zero phase difference is in general not major significance in this case. To simplify the
mathematics, let us suppose, therefore, that the individual vibrations have zero initial phase, and
hence can be written as follows:
x1 = A1 cos ω 1t
x2 = A2 cos ω 2t

8
Fig 1.3: Superposition of rotating vector of different periods.

At some arbitrary instant the combined displacement will then be as shown (OX) in Figure 1.3.
The length of OP of the combined vector must always lie somewhere between the sum and the
difference of A1 and A2; the magnitude of the displacement OX itself may be anywhere between
zero and A1 + A2.
Unless there is some simple relation between ω1 and ω2, the resultant displacement will be a
complicated function of time, perhaps even to the point of never repeating itself. The motion for
any sort of true periodicity is the combined motion is that the periods of the component motions
be commensurable – i.e., there exist two integers n1 and n2 such that

T = n1T1 + n2 T2 (1.15)
The period of the combined motion is then the value of T as obtained above, using the smallest
integral values of n1 and n2 for which the relation can be written

NOTE: If for example, the ration ω 1/ω 2 were an irrational (e.g. √ 2), there would exist
no time, however long, after which the preceding pattern of displacement would be
repeated.

Even if the periods or frequencies are expressible as a ratio of two fairly small integers, the
general appearance of the motion is not particularly simple.
If two SHM’s are quite close in frequency, the combined disturbance exhibits what are called
beats. This phenomenon can be described as are in which the combined vibrations is basically a
disturbance having a frequency equal to the average of the two combining frequencies, but with
an amplitude that varies periodically with time – one cycle of this variation including many
cycles of the basic vibrations. The beating effect is most easily analyzed if we consider the
addition of two SHM’s of equal amplitude:
x1 = A Cos ω 1t
x2 = A Cos ω 2t

9
Then by adding we get
   2     2 
x  2 A cos 1 t  cos 1 t (1.16)
 2   2 

NOTE: You may wish to recall the following trigonometric results:


cos (θ + φ) = cosθ cosφ - sinθ sinφ
cos (θ - φ) = cos θ cosφ + sinθ sin φ
Therefore,
Cos (θ + φ) + cos(θ - φ ) = 2cos θ cosφ
In this identity, let θ + φ = α, θ - φ = β, Then
    
Cosα + Cos β = 2cos   cos 
 2   2 
In the case under discussion, we put α = ω 1t, β = ω 2t.

Clearly this addition, as a purely mathematical result, can be carried out for any value of ω 1 and
ω 2. but its description as a beat phenomenon is physically meaningful only if ‫׀‬ω 1 – ω 2‫ << ׀‬ω 1 +
ω 2; i.e. if, over some substantial number of cycles, the vibration approximates to sinusoidal
vibration with constant amplitude and with angular frequency (ω 1 + ω 2)/2.
It may be seen that the combined displacement can be fitted within an envelope defined by the
pair of equations.
  2 
x = ± 2A cos  1 t (1.17)
 2 
because the rapidly oscillation factor in eq. (1.16) i.e. cos  (1   2 )t 2  – always lies between
the limits ±1, and eq. (1.17) describes a relatively slow amplitude modulation of this oscillation.
Therefore, the time between successive zeros of the modulating disturbance is one half –period
of the modulating factor described by Eq. (1.17), i.e., a time equal to
2π/(‫׀‬ω 1 – ω 2‫)׀‬. This has the consequence that the beat frequency – as observed, e.g. from two
tuning forks- is simply the difference of their individual frequencies and not half this frequency,
as might be suggested by a first glance at Eq.(1.4). Thus to take a specific case, if two tuning
forks side by side are vibrating at 255 and 257 vibrations per second their combined effect would
be that of middle C (256 vibrations per second) passing through a maximum of loudness twice
every second.
1.4.4 Many superposed vibrations of the same frequency

This is a study of superposition of a number of SHM’s all of the same frequency and amplitude,
and with equal successive phase differences. This problem has special relevance to the analysis
of multi-source interference effects in optics and other wave processes. Suppose that there are N

10
combining vibrations, each of amplitude A0 and differing in phase from the next one by an angle
δ. The situation is represented in Figure1.4

Fig.1.4: Superposition of several rotating vector of same period constant increment, phase
differences.

Let the first of the component vibrations be described, for simplicity, by the equation.
x = A0 cos ωt
The resultant disturbance will be given by the equation

x = Acos (ωt + α)
From the geometry of Fig.1.4, we can see that the combining vector form successive sides of an
(incomplete) regular polygon. Any such polygon can be imagined to be inscribed in a circle,
having some radius R and with its centre at a point C. All the corners (as, for example the points
K and L lie on the circle, and the angle subtended at C by any individual amplitude A0 (e.g., KL)
is equal to OCP1 subtended at C by the resultant vector A, is equal to Nδ. We can then write the
following geometrical statement.
 N 
A = 2Rsin  
 2 
 
A0 = 2R sin  
 2

11
Therefore

 2
sin N
sin  
A = A0 2 (1.18)
Also, for the phase angle α through which the resultant A is rotated relative to the first
component vector, we have

α = ∠COB - ∠COP
with
 
∠COB = 900 -  
 2

 N 
∠COP = 900 -  
 2 
Therefore
 N  1
α= (1.19)
2
Hence the resultant vibration along the x axis is described by the following equation

 2 cos t   N  1 
sin N
sin   
 2 
x = A0 2 (1.20)
This equation is basic to the analysis of the behaviour of diffraction grafting, which acts
precisely as a device to obtain from a single beam of light a very large number of equal
disturbances with equal phase differences.

1.4.5 Combination of two vibrations at right angles

Everything we have discussed so far has been concerned with harmonic motion along one
physical dimension only. We shall now discuss the essential difference problem of combining
two real harmonic vibrations that take place along perpendicular directions, so that the resultant
real motion is a true two-dimensional motion.

12
Fig 1.5: Geometrical representation of the superposition of simple harmonic vibrations at right
angles.
Suppose therefore, that a point experiences the following displacement simultaneously:

x  A Cos ω1t  α1 
1
y  A 2 Cos ω 2 t  α 2 
(1.21)

We can construct this motion by means of a double application of the rotating – vector technique.
The way of doing this is displayed in Fig.1.5. We begin by drawing two circles, of radius A1 and
A2 respectively. The first is used to define the x displacement C1X of point P1. The second is used
to define the x displacement C2X of the point P2. The two displacements together describe the
instantaneous position of the point P with respect to an origin O that lies at the centre of a
rectangle of sides 2A1 and 2A2.
One feature is immediately apparent. Whatever the relation between the frequencies and the
phases of the two combining motions, the motion of the point P is always confined within the
rectangle, and also the sides of this rectangle are tangential to the path at every point at which the
path touches these boundary lines. In general if ω1 and ω2 are not commensurable, the position of
P will never repeat itself, and the path, if continued for long enough, will from a physical
standpoint even if not from a strictly mathematical one, tend to fill the whole interior of the
bounding rectangle.
The most interesting examples of these combined motions are those for which the frequencies
are in some simple numerical ratio and the difference of the initial phases is some simple fraction

13
of 2π. One then has a motion that forms a closed curve in two dimensions, with a period that is
the lowest common multiple of the individual periods. The problems we are going to discuss in
terms of specific examples are:
a) Perpendicular motions with equal frequencies
b) Perpendicular motions with different frequencies; Lissajous figures

1.5.1 Perpendicular motions with equal frequencies

By a suitable choice of what we call t = O, we can write the combining vibrations in the
following simple form:
x = A1 cosωt
y = A2 cos(ωt + δ)
where δ is the initial phase difference (and in this case the phase difference at all later times too)
between the motions.
Giving δ a particular value while the combining frequencies are equal.
δ = O. In this case,
x = A1cosωt
y = A2cosωt
Therefore
A2
y= x
A1
The motion is rectilinear, and takes place along a diagonal of the rectangle such that x and y
always have the same sign, both positive and both negative. This represents what in optics is
called linearly polarized vibration.
(b) δ=  . We now have
2
x = A1cosωt

y = A2 cos t   2  = - A sin ωt


2

The shape of this path is readily obtained by making use of the fact that sin 2ωt+ cos2ωt = 1. This
means that
x2 y2
 1
A12 A2 2
which is the equation of an ellipse whose principle axis lies along the x and y axis. Notice,
however, that the equations tell us more than this. We are dealing with kinematics, not geometry,
and the ellipse is described in a definite direction. As t begins to increase from zero, r begins to
decrease from its greatest positive value, and y immediately begins to go negative, starting from
zero. This means that the elliptical path takes place in the clockwise direction.

14
(c) δ = π We now have
x = A1cos ωt
y = A2 cos (ωt + π) = - A2 cos ωt

Therefore,
A2
y=- x
A1

This motion is like case a, but it is a long the other diagonal of the rectangle
(d) δ = 3
2. This gives us
x = A1 cos ωt
3
y = A2 cos( t  ) = A2 sin ωt
2
We have an ellipse of the same form as in case b, but the motion is now counterclockwise.
(e) δ =  π.
4
Note that we are jumping back here to the case of a phase difference between O and π/2, i.e.
intermediate between cases a and b. It is a less obvious case than those just discussed, and lends
itself to the graphical construction of Fig.1.3. The application of the method to this particular
case is shown in Figure1.4.
The positions of the points P1, P2, on the two reference circles are shown at a number of instants
separated by one eighth of a period (i.e. π/4ω ). The points are numbered in sequence, beginning
with t = O, when C1P1 (see Fig.1.4) is parallel to the x-axis and C 2P2 is at the angle δ, i.e., 450,
measured counterclockwise from the positive y-axis. The projections from these corresponding
positions of P1and P2 then give us a set of intersections as shown in Fig1.4, representing the
instantaneous positions of the point P as it moves within the rectangle. The locus defined by
these points is an ellipse, on inclined axes, described clockwise.
The analytical equation of this ellipse can be found, if desired by eliminating t from the defining
equation for x and y:
x = A1cos ωt

 
y  A2 Cos t  
 4
A A
 2 cos t  2 sin t
2 2

15
Fig.1.6: Superposition of simple harmonic vibrations at right angles with initial phase
difference of π/4.

With the help of this last example, we can see how the pattern of this combined motion develops
as we imagine the phase difference δ to increase from zero to 2π.
Fig.1.7 shows a summary of the patterns formed from δ = 0 to δ = 2π
c

16
Fig.1.7: superposition of two perpendicular simple harmonic motions of the same frequency for
various initial phase differences.

1.5.1 Simplifying the diagram


Once one is familiar with the process involved, one can make a more compact diagram by taking
the boundary rectangle and simply constructing a semicircle on two adjacent sides. To illustrate
this, let us take the case where ω1 = ω2, δ = π/4, once again. With the division of the reference
circles into even submultiples of 2π, two different points on the circle project to give the same
value of the displacement. Thus by drawing just a semicircle, one can convey as much
information as with the full circle, but many of the points are used twice over, as indicated in
Figure.1.8. Once the points on the reference circles have been numbered according to the correct
time-sequence, the intersections that define the coordinate of the actual motion are obtained are
obtained just as before. (Here a = 1, b = 2, etc)

Fig.1.8: Abbreviated construction for superposition of vibrations at right angles.

1.6 Perpendicular motions with different frequencies; Lissasjous figures


Fig.1.9 shows the construction that one can make if ω2 = 2ω and δ = π/4. We have chosen to
divide the reference cycle for the motion of frequency ω2 into eight equal time intervals i.e., into
arcs subtending 450 each.
During one complete cycle of ω2, we go through only half–cycle of ω1 and the points on the
reference circles are marked accordingly, taking account of the assumed initial phase difference
of 450. To obtain one complete period of the combined motion it is, of course necessary to go
through a complete cycle of ω1; this requires that after reaching the point marked “9,” we retrace
our steps along the lower semicircle and proceed for a second time through all the points
corresponding to a complete tour around the ω2 circle. In this way we

17
Fig. 1.9: Construction of a Lissajous figure.

end up with a closed path which crosses itself at one point and would be indefinitely repeated.
Such a curve is known as a Lissajous figure.
Figure 1.10 we show a set of such curves all for ω2 = 2ω1, with initial phase differences of
various sizes.

ω 2 = 2ω1

δ = 0 δ = π /4 δ = π /2 δ = 3 π /4 δ= π

Fig1.10: Lissajous figures for ω2 = 2ω1 with various initial phase differences.

1.7 Comparison of parallel and perpendicular superposition


If the two primary signals are given by Acos ωt and Acos (ωt + δ), we have the following results

1.7.1 Parallel superposition


y1 = A cos ωt
y2 = A cos (ωt + δ)

  
y  y1  y2  2 A cos  cos t  
2 2

NOTE: The smooth decrease of amplitude in proportion to cos(δ/2) as δ increase
from zero to π.

18
1.7.2 Perpendicular Superposition
x = Acosωt
y = Acos(ωt + δ)
Eliminating the explicit time dependence, we have

x2 – 2xycosδ + y2 = A2sin2 δ,
Defining an elliptic curve which degenerates into a straight line for δ = 0 or π, and into a circle
for δ = π/2. .

1.8 ACTIVITIES
a) Differentiate between periodic motion and oscillatory motion.
b) What is the main difference between the motion of the planets around the sun and the motion
of a pendulum?
c) Is it possible for an oscillatory motion to be equivalent to a periodic motion? Explain.
d) Explain the meaning of the statement: Harmonic motion are periodic but not all periodic
motions are harmonic.

1.9 FURTHER READING


1) Jacobson, W. J. (1972). Periodic motion. New York: American Book Co..
2) Schlegel, R. (1980). Superposition & interaction: coherence in physics. Chicago:
University of Chicago Press
3) Harris, J. (1977). INTERP: unit on wave superposition.. London: Edward Arnold.
4) Baumhauer, J. C., & Tiersten, H. F. (1972). Nonlinear Electroelastic Equations for Small
Fields Superposed on a
5) Jacobson, W. J. (1972). Periodic motion . New York: American Book Co..
6) Finkle, I. L. (1964). Recording Lissajous Figures . Ft. Belvoir: Defense Technical
Information Center.
7) Manley, R. (1945). Waveform Analysis: A Guide to the Interpretation of Periodic Waves,
including Vibration Records. London: Chapman & Hall.

1.10 SELF-TEST QUESTIONS

19
1) Show that SHM can be projected on a straight line.
2) State two features that are essential for the establishment of oscillatory motion of
the simple harmonic type.
3) A point moves in a circle at a constant speed of 0.4 m/sec. The period of one
complete revolution around the cycle is 12 seconds. At t = 0 the line to the point
from the centre of the cycle makes an angle of 30o with the x axis.
i. Obtain the equation of the x co-ordinate of the point as a function of time,
in the form x = A cos (t + ) giving the numerical values of A,  and .
ii. Find the values of x, dx/dt and dx2/dt2 at t = 4 sec.
4) Two collinear waves whose amplitudes are the same but their angular frequencies
differ by  are represented as
i. x1 = A sint
ii. x2 = A sin( +)t
5) Determine the amplitude of the resultant combination.
6) Draw the Lissasjous figure that results from two perpendicular motions of
different frequencies given that c + 2 and having initial phase angle of.
(i) =
(ii)  = 3/4

LECTURE TWO

FORCE AND ENERGY

20
2.1 Introduction

In this lecture we learn how to derive the equation of force in a mechanical wave. Conservation
of energy in a mechanical wave is also discussed.

2.2 Lecture Objectives

By the end of this lecture, you should be able to:


 State and derive the equation of force in a mechanical wave.
 State and derive the equation of conservation of energy in a mechanical wave.
 State and derive the equation of conservation of energy in a mechanical wave.

2.3 Force in a vibrating spring


Let m be the linear density, of the string, i.e. it’s mass per unit length, which is not necessarily a
constant but may vary with x, the distance along the string. Let there be a longitudinal tension T
in the string.
Consider a small element AB (Fig. 1.0.) of the string of length δx and let y be the displacement
of AB from the equilibrium position.

Fig. 2.1: Force of an element of vibrating string

21
The tension at A has a transverse component which exerts a force T (dy/dx) on the element
towards the undisplaced position and at B the transverse force is
 dy d 2 y 
T  2
x  (2.1)
 dx dx 
in the opposite direction. There will be a net force of T (d 2y/dx2) δx on the element which will
cause a lateral acceleration d2y/ dt2 given by
d2y d2y
T  x  m  x (2.2)
dx 2 dt 2

d2y d2y
T  m (2.3)
dx 2 dt 2
The above gives us the force in a vibrating spring, which literary is the wave equation for the
vibrating string.

Example
Verify that the differential equation 2y/ x2 = -k2y as its solution y = Acos(kx) + Bsin(kx) where
A and B are arbitrary constants. Show also that this solution can be written in the form y =
Ccos(kx + ) = C Re ej(kx +x) = Re Cej. ejkx and express C and  as a function of A and B.
Solution.
y = Acos(kx) + Bsin(kx)
dy/dx = -kAsin(kx) + kBsin(kx)
d y/dx2 = -k2Acos(kx) – k2Bsin(kx)
2

d2y/dx2 = -k2[A cos(kx) + B sin(kx)]


d2y/dx2 = -k2y

Let A = C cos  and B = - C sin


y = C cos cos(kx) - C sin sin(kx)
From the identity
cos(A +B) = cosA cosB – sinA sinB
Thus y = C cos(kx + )
In complex notation
y = Ccos(kx + ) = C Re ej(kx +x) = Re Cej. ejkx
C =  (A2 + B2)

tan = - B/A
 = tan-1(- B/A)

2.4 The energy in a mechanical wave

22
At any instant the particles of a medium carrying a wave are in various states of motion. Clearly
the medium is endowed with energy that it does not have in its normal resting state. There are
contributions from the potential energy of deformation as well as from the kinetic energy of the
motion. We shall calculate the total energy associated with one complete wavelength of a
sinusoidal wave on a stretched string.
By way of approaching this problem we shall consider first a small segment of the string – so
short that it can be regarded ad effectively straight – that lies between x and x+dx as shown in
Fig. 2.2. We shall make the usual assumption that the displacements of the particles in the string
are strictly transverse and that the magnitude of the tension T is not changed by the deformation
of the string from its normal length and configuration.

Fig. 2.2: Displacement and extension of a short segment of string carrying a transverse elastic
wave.

The mass of the small segment is µdx and its transverse velocity (µy) is δy/δt. Hence for this
segment we have
2
1  y 
Kinetic energy  dx  
2  t 
and we can define a kinetic energy per unit length – what is called the kinetic – energy density –
for such a one – dimensional medium:
2
dK y 
  
1
kinetic–energy density =  (2.4)
dx 2  t 

The potential energy can be calculated by finding the amount by which the string, when
deformed is longer than when it is straight. This extension, multiplied by the assumed constant
tension T, is the work done in the deformation. Thus, for the segment we have:
potential energy = T(ds – dx)
where

ds = (dx2 + dy2)½

23
1
 y 
2
 2

=dx  1    
  x  
If we assume that the transverse displacement are small, so that δy/δx <<1, we can appropriate
the above expression using the binomial expression to two terms, thus getting
2
1  y 
ds – dx ≈   dx
2  x 
Therefore,
2
1  y 
Potential energy ≈ T  dx
2  x 
Hence we have
2
dU 1  y 
Potential–energy density ≡  T  (2.5)
dx 2  x 
It is worth noting that the kinetic–energy densities and potential-energy densities as given by
Eqs.(2.4) and (2.5) are equal. For as we have seen a traveling wave on the string is of the form
y(x,t) = f(x ± vt) = f(z), say,
where
1
 T  2
v=  
 
Thus
y
 f  z 
x
y
  vf  z 
t
Therefore,
dK 1 2
 u  f  z  
2

dx 2

dU 1
 T  f  z  
2

dx 2
which are equal since T = µv2. Although this equality of the two energy densities cannot be
assumed to hold good in all conceivable situations, it is keeping with what we know about the
equal division on the average of the total energy of simple mechanical systems subject to linear
forces.
Suppose now that we have in particular a sinusoidal wave described by the equation:

 x
y(x,t) = A sin 2πv  t   (2.6)
 u
Then at any given value of x we have

24
y  x
u(x,t) =  2vA cos 2v t  
t  u

 x
= uo cos 2πv  t  
 v
where uo (=2πvA) is the maximum speed of the transverse motion. Let us consider this
distribution of transverse velocities at the time t = 0. At this instant we have
 2vx   2vx 
u(x) = uo cos     u 0 cos 
 u   v 
Since v/u = 1/ λ, this can especially well be written
 2x 
u(x) = uo cos  
  
The kinetic energy density is thus given by
dK 1 2 1  2x 
 u   0 cos 2 
2

dx 2 2   
The total kinetic energy in the segment of string between x = 0 and x = λ is thus

1 2  2x 
K = u 0  cos 
2
 dx
2 0   
i.e.
1
K=     0 2 (2.7)
4
This, then is the kinetic energy associated with one complete wavelength of the disturbance. The
potential energy over the same portion of the string must, as we have already seen be equal to the
kinetic energy. For the sake of being quite explicit, however we will carry out the calculation.
From Eq. (2.6) we have

y 2vA  x
 cos 2v t  
x u  u

Thus at t = 0 we have
 y  2vA  2vx  2A  2x 
   cos   cos 
 x  t 0 u  u     
Hence the potential-energy density {Eq.(2.5)} is given by
dU 2 2 A 2T  2x 
 cos 2  
   
2
dx
Integrating over one wavelength then gives
 2 A 2T
U=

Putting T = µu2 = µv2λ2, this gives us

25
U = π 2A2uv2λ (2.8)
which can be recognized as equal in magnitude to K, as given by Eq.(2.7), if we use
the identity µo = 2πvA.
The total energy per wavelength E, can be written

1
E= (λu)µo2 (2.9)
2
and is thus equal to the kinetic energy that apiece of string of length λ would have if all of
it were moving with the maximum transverse velocity uo associated with the wave.

2.5 ACTIVITIES
a) Express in symbols the principle of conservation of energy for a simple harmonic oscillator.
b) Explain factors that affect energy in a mechanical wave.

2.6 FURTHER READING

1) Cummings, K., & Halliday, D. (2004). Understanding physics . Hoboken, NJ: Wiley.
2) Miles, J. B., & Shelpuk, B. C. (1981). Ocean energy: waves, current, and tides. Golden,
1) Show
Colo.: thatEnergy
Solar the wave equation
Research for the vibrating string with tension T is given by.
Institute
2 2
d y d y
T 2 m 2 .
dx dt

2) Without using a trial solution, show that the block and spring system shown below
executes simple harmonic motion (ignoring the effect of friction and air resistance.).

2.7 SELF-TEST QUESTIONS

3. Show how to obtain the equation of pendulum by using the principle of conservation of
energy.

5. A simple pendulum of length 1m has oscillatory- energy equal to 0.3joule when its
amplitude is 0.02 metre. What will be its energy if (a) its length is increased to 1.5 metres,
(b) its amplitude is increased to 0.03 metre?

6) A body of mass m is attached to a spring of negligible mass which has a force exerted
on it at a displacement x from equilibrium.
Write: 26
i. An expression to depict this force
ii. The equation of motion of the body.
LECTURE THREE

COMPLEX VARIABLES REPRESENTATION OSCILLATIONS

3.1 Introduction

In this lecture we discuss how to represent wave equation a wave equation using rotating
vectors. We also learn to convert from polar coordinates given rectangular coordinates and

27
vice versa. Equations with real and imaginary parts are also discussed. Also discussed here is
the simple harmonic motion.

3.2 Lecture Objectives

By the end of this lecture, you should be able to:


 Appreciate the importance of complex representation of wave equation.
 Be able to convert from polar coordinates given rectangular coordinates and
vice versa
 Understand the meaning of i   1
 Write equations having real (Re) and imaginary (Im) parts.
 Use complex exponentials in SHM

3.3 The Rotating Vector Presentation

SHM can also be obtained by regarding it as the projection of uniform circular motion. A disk of
radius A rotates about a vertical axis at the rate of ω rad/sec. A peg P is attached to the edge of the
disk and a horizontal beam of parallel light casts a shadow of the peg on a vertical screen, as
shown in Fig.3.1 (a). Then this shadow performs SHM with period 2π/ω and amplitude A along a
horizontal line on the screen.
We can imagine SHM as being the geometric projection of uniform circular motion. (By
geometrical projection we mean simply the process of drawing a perpendicular to a given line
from the instantaneous position of the point P.)

28
Fig.3.1: SHM as the projection in its own plane of uniform circular motion.

In Fig.3.1 (b) we indicate the way in which the end point of the rotating vector OP can be
projected onto a diameter of the circle. The horizontal axis Ox is chosen as the line along which
the actual oscillation takes place. The instantaneous position of the point P is then defined by the
constant length A and the variable angle θ. Counterclockwise direction is taken as positive, then
the actual value of θ is
θ=ωt+α
where α is the value of θ at t = 0
As specified above, the displacement x of the actual motion is given by

x = Acosθ = Acos (ωt + α) (3.1)


This equation differs from our initial description of SHM according to eq. (1.3). To make the
equation identical, we know that for any angle θ we have
 
cosθ = sin    
 2
The identity of Equations. (1.1) and (3.1) requires
A sin (ωt + φ0) = A cos (ω t + α)

i.e. sin (ω t + φ0) = Sin(ωt + α + )
2
The sines of two angles are equal if the angles are equal or if they differ by any integral multiply
of 2π. Taking the simplest of these possibilities, we can thus put


φ0 = α + (3.2)
2
The equivalence of Equations (1.3) and (3.1) subject to the above conditions allows SHM to be
described equally well in terms of a sine or a cosine function.

29
3.4 Rotating vectors and complex numbers
Circular motion once it has been set up defines SHM of amplitude A and angular frequency ω
along any straight line in the plane of the circle. In particular if we imagine a y axis
perpendicular to the real physical axis Ox of the actual motion, the rotating vector OP defines for
us, in addition to the true oscillation along x, and accompanying orthogonal oscillation along y,
such that
x = A cos (ωt + α)
(3.3)
y = A sin (ωt + α)
The motion along y has no actual existence, but the motion is dealt with as a point in two
dimensions, as described in equation (3.3), provided that, at the end we extract only the x
component, because this is the physical meaningful result of the motion thus described.

Fig. 3.2: Cartesian r polar representation of a rotating vector.

A vector OP (Fig.3.2) has the plant polar coordinates (r, θ).


The rectangular (Cartesian) components (x, y) are, of course defined by the following equations:

x = r cosθ y = rsinθ
Unit vector iˆ denotes displacement along x and a unit vector ĵ denotes displacement along y.
The component vector r can then be expressed as the vector sum of the two orthogonal
components.
r = iˆ x + ĵ y
But without any sacrifice of information content, we can define the vector by means of the
following equation
r = x + jy (3.4)
Eq. (3.4) embodies the following statements:
1. A displacement, such as x, without any qualifying factors, is to be made in a direction
parallel to x axis.

30
2. The term jy is to be read as an instruction to make the displacement y in a direction
parallel to the y axis.
A quantity z, understood to be the result of adding jy to x i.e. identical with r as defined above is
the usual vector symbolism. Thus

z = x + jy

3.5 The meaning of j


Symbol j is read as an instruction to perform a counterclockwise rotation of 90 0 upon whatever it
precedes.
Consider the following specific examples:
a) To form the quantity jb, we step off a distance b along x axis and then rotate through 90 0
so as to end up with a displacement of length b along y.
b) To form the quantity j2b we first form jb, as above, and the apply to it a further 90 0
rotation – i.e., we identify j 2b as j(jb). Bit this at onces leads to an important identity. Two
successive 900 rotations in the same sense convert a displacement b (along the positive x
direction) into the displacement –b. Hence we set up the algebraic identity.
j2 = -1 (3.5)
Therefore j= 1

Fig. 3.3: (a) Representation of a vector in the complex plane. (b) Multiplication of Z by j is
equivalent to a 900 rotation.

(c) Suppose we take a vector z having an x component of length a and a y component of


length b (Fig.3.3a)
What is jz?
We have
z = a + jb
jz = ja + j2b

31
= ja + (-b)
The summation of the new vector components on the right of the above equation is shown in
Fig.3.3 (b). The resultant vector jz is obtained from the original vector z just by the extra rotation
of 900.
Thus just the combination z = a + jb is what is known as complex number. But in geometry it can
be regarded as a displacement along an axis at some angle θ to the x axis, such that tan θ = b/a,
as is clear from Fig.3.3 (a).
If, after solving an oscillatory motion problem in these terms, we obtain a final answer in the
form z = a + jb, where a and b are both real numbers, then the quantity a is the wanted quantity,
and b can be discarded. Thus the quantity of the form jb alone (with b real) is called imaginary.

3.6 Complex variables representation of oscillations


We begin by taking the series expansions, of the sine and cosine functions:

3 5
Sin θ = θ -  ……. (3.6)
3! 5!

2 4
Cos θ = 1 –  …….. (3.7)
2! 5!

These expansions, if not already familiar are readily developed with the help of Taylor’s
theorem.

By Taylor’s Theorem,
x2
f(x) = f (0)  xf (0)  f  0   -------
2!
Therefore,
2
  sin 0    cos 0 --------
3
sinθ = Sin O + θ cosO +
2! 3!

  cos 0    sin 0 -------
2 3
cosθ = cosO + θ (- sinO) +
2! 3!

Let us now form the following combination:


2 3 4
cosθ + jsinθ = 1 + jθ –  j  + ------ (3.8)
2! 3! 4!

We now know that –1 = j2, so the above equation can be rewritten as follows:

32
Cos θ + jsin θ = 1 + jθ +
 j  2   j  3  ......   j  n (3.9)
2! 3! n!
But the right-hand side of this equation has precisely the form of the exponential series, with the
exponent set equal to jθ. Thus we are enabled to write the following identity:

cosθ + jsinθ = ejθ (3.10)


This expression provides a clear connection between plane geometry (as represented by
trigonometric function) and algebra (as represented by the exponential function)

3.7 Using the complex exponential in SHM

If, as often happens, the basic equation of motion contains term proportional to velocity and
acceleration, as well as to displacement itself, then the use of a simple trigonometric function to
describe the motion leads to an awkward mixture of sine and cosine terms. For example
x = A cos (ωt + α)
then
dx
= - ωAsin(ωt + α)
dt

d 2x
2
= - ω 2Acos(ωt + α)
dt
On the other hand, if we work with the combination x + jy, with x and y as given by equation
(3.3), we have the following:
z = Acos(ωt + α) + jAsin (ωt + α)
i.e.
z = Ae j(ωt + α)
with
x = real part of z (often abbreviated Re(z))
Then
dz α)
= jωAe j(ωt + = jωz
dt
d 2z
= (jω)2Aej(ωt + α) = -ω 2z
dt 2
These three vectors are shown in Fig 3.4 (using three separate diagram, because quantities of
three physically different kinds displacement, velocity, acceleration are being described). In each
case the physically relevant component is recognizable as being the real component of the vector
in question, and the phase relationships are visible at a glance (given the result that each factor of
j is to be read as an advance in phase angle by π/4.)

33
Fig. 3.4: (a) Displacement vector z and its real projection x (b) velocity vector dz/dt and its real
projection dx /dt. (c) Acceleration vector d2z / dt2 and its real projection d2x /dt2.

3.8 ACTIVITIES

Find the Re and Im parts of


i. (4 + 2j)/4-j
ii. [cos(/4) + sin(/4)]3

3.9 FURTHER READING

1) Silva, C. W. (2000). Vibration: fundamentals and practice. Boca Raton, FL: CRC Press.
2) Silva, C. W. (2000). Vibration: fundamentals and practice. Boca Raton, FL: CRC Press.
3) Kneubühl, F. K. (1997). Oscillations and waves . Berlin: Springer.
4) Miles, J. B., & Shelpuk, B. C. (1981). Ocean energy: waves, current, and tides. Golden,
Colo.: Solar Energy Research Institute.

3.10 SELF-TEST QUESTIONS

34
1. Derive the equations
Sin  =(ej - e-j)/2j and cos  =(ej +e-j)/2j
2. Find the Re and Im parts of
(i) (3 + 2j)/4-j
(ii) [cos(/4) + sin(/4)]8

3. Using the exponential representation for the sin and cos, verify the following
trigonometric identities
(i) sin2 = 1 – cos2
(ii) 2sin cos = sin2
(iii) sin θ + sin 2θ + sin 3θ = sin 2θ(1 + 2cos θ)
(iv) sin3θ - sin θ = cos 2θ

LECTURE FOUR

35
FREE AND FORCED VIBRATIONS

4.1 Introduction

In this topic we learn more about free oscillation of bodies the corresponding equation that
represents such motion. Equations of forced vibrations are also discussed here. Equations on
floating objects and pendulums are discussed here too. Damping of forces is discussed also in
this topic.

4.2 Lecture Objectives

By the end of this lecture, you should be able to:


 Derive the free vibration of a wave equation in complex exponential.
 Understand the effect of damping in waves.
 Derive equations of forced vibrations and resonance for both mechanical and
electrical vibrations.
 Compare and construct vibrations in mechanical and electrical systems.

4.3 ( A) Free vibrations


If a body capable of oscillating is distributed from its equilibrium position and then left to itself,
the body oscillates with as frequency determined by the physical dimension and the elastic
properties of the body. This oscillation is called free oscillation because the system is left free to
itself after feeding some energy initially. The frequency of free oscillation is called the natural
frequency of the system.

4.3.1 Free oscillation of mass – spring.


A basic mass – spring consists of a single object of mass m acted on by a spring or some
equivalent device e.g. a thin wire that supplies a restoring force equal to some constant K times
the displacement from the equilibrium. This identifies, in terms of a system of a particularly
simple kind the oscillating motion:
a) An inertial component, capable of carrying kinetic energy.
b) An elastic component capable of storing elastic potential.

36
By assuming that Hooke’s law holds we obtain a potential energy proportional to the square of
the displacement of the body from equilibrium. By assuming that the whole inertia of the system
is localized in the mass at the end of the spring we obtain a kinetic energy equal to just mv 2/2,
where v is the speed of the attached object. It should be noted that both of these assumptions are
specializations of the general conditions 1 and 2, and there will be many instances of oscillatory
system to which these special conditions do not apply. If however a system can be regarded as
being effectively a concentrated mass at the end of a linear spring (“linear” referring to its elastic
property rather than to its geometry), then we can write its equation of motion in either of two
ways:
(a) By Newton’s law (F = ma)
-kx = ma
(b) By conservation of total mechanical energy (E)

1 1
mv 2  kx 2  E
2 2
The second is of course, the result of integrating the first with respect to the displacement x, but
both of them are differential equations for the motion of the system. In explicit differential form,
they may be written as follows:
d2x
m 2  kx  0 (4.1)
dt
2
1  dx  1 2
m   kx  E (4.2)
2  dt  2
whenever one sees an equation analogous to either of the above, one can conclude that
the displacement x as a function of time is of the form:
x = A cos (ω t + α) (4.3)
2
where ω is the ratio (k / m) of the spring constant k to the inertia constant m. (A =
amplitude α = initial phase angle).
The initial statement of Newton’s law in Eq. (4.1) contains no adjustable constants. Eq. (4.2),
often referred to as the “first integral” of Eq. (4.1) is mathematically intermediate between Eqs.
(4.1) and (4.3) and contains one adjustable constant (the total energy E, which is equal to kA 2/2.
The introduction of one more constant at each stage of integration of the original differential
equation (Newton’s law) is always necessary, even though in particular case the constant may
turn to be zero. One can think of this as the transverse of the process whereby, in any
differentiation, constant term will disappear from sight.

4.3.2 Solving the harmonic oscillator equation using complex exponentials.


Since it is not k and m individually, but only the ratio k/m that enters in any essential
way, we began by rewriting Eq. (4.1) in the following more compact form:
d 2x
2
 2x  0 (4.4)
dt

37
This states that x and its second time derivation are linearly combined to give zero or
equivalently that d2x / dt2 is a multiple of x itself. The exponential function is known to have this
latter property; let us therefore put
x = Cept (4.5)
We have introduced a coefficient C of the dimension of distance and a coefficient P such
that pt is dimensionless. i.e. p has the dimension of (time)-1.
Then by substituting in Eq. (4.4) we have
p2Cept + ω 2Cept = 0
which can be satisfied for any t, and for any value of C1 provided that
p2 + ω 2 = 0
Therefore,
p2 = - ω 2
p = ± jω (4.6)
Each of these values of p will satisfy the original equation. Having no reason to discard
either, we accept both, each with its own value of C. Thus Eq. (4.5) becomes
x = C1ejωt + C2e-jωt (4.7)
Let us interpret Eq. (4.7) in terms of the rotational vector description of SHM. The first term on
the right corresponds to a vector C1 rotating counterclockwise at angular speed ω, the second
term to a vector C2 rotating clockwise at the same speed. These combine to give a harmonic
oscillation along the x axis, as shown in Fig. 4.1 (a) if the lengths of C1 and C2 are equal.

Fig 4.1: (a) Superposition of complex solutions of Eq. (4.4) with C 1 = C2 (b) Superposition of
complex solutions of Eq. (4.7) for nonzero initial phase angle.

38
But C1 and C2 as they appear in Eq. (4.7), do not have to be real. We can satisfy Eq. (4.7) just as
well if C1 is rotated through some angle α with respect to the direction defined by t, provided that
C2 is rotated through –α with respect to –ωt, again making the vector of equal length, as shown in
Fig. 4.1(b). This restrictive condition leads to the results below:
x = Cej(ωt+α) + Ce-j(ωt + α)
= 2Ccos (ωt + α)
≡A cos (ω t + α)
The above analysis reveals that a rectilinear harmonic motion can be produced by the
superposition of two equal and opposite real circular Lissajous figure from two equal and
perpendicular linear oscillation. Having arrived at the final equation, we can see that x can be
described as the real part of a rotating vector corresponding just to the first term alone in Eq.
(4.7). Thus we shall assume that:
x = real part of z where z = Aej(ωt + α) (4.8)

4.4 Floating objects


If a floating object is slightly depressed or raised from it’s normal position of equilibrium, there
is called into play a restoring force equal to the increase or decrease in the weight of liquid
displaced by the object, and periodic motion ensues. The situation becomes especially simple if
the floating body has a constant cross-sectional area over the part that interests the liquid. A
hydrometer (Figure 4.3), as used to measure the specific gravity of battery acid or antifreeze, is a
nice practical example of this.
Let the mass of the hydrometer be m, and let the liquid density be ρ. Denote the area of cross
section as A. Then if the hydrometer is at distance y above it’s normal floating level, the volume
of liquid displaced is equal to Ay and the equation of motion (Newton’s law) becomes
d2y
m 2   gAy
dt
giving
gA m
ω= and T = 2π (4.9)
m gA

39
Fig. 4.2: Simple hydrometer, capable of vertical oscillations when displacement from normal
floating.
On a much larger scale, one can consider such motion occurring with a slip. To some
approximation the rider of a big ship are almost vertical, and it’s bottom more or less flat, in this
case the mass of the ship is express in terms of it’s draft, h as
m = ρAh
Where ρ is the density of water and A is the horizontal cross- sectional area of the ship at
the waterline.
Substituting in Eq. (4.9) we find
m
T = 2π (4.10)
g
which is thus exactly like the simple pendulum equation.

4.5 Pendulums
Referring to Fig. 4.3 (a), if the angle θ is small we have y << x and hence from the geometry of
the figure,
x2
y ≈
2l
where l is the length of the string. The statement of conversation of energy is
2 2
1  dx   dy 
mv 2  mgy = E where v2 =    
2  dt   dt 
given the approximation already given, it is thus very nearly correct to put
2
1  dx  1 mg 2
m   x = E
2  dt  2 l

4.5 Pendulums

40
(b)
Fig. 4.3: (a). Simple pendulum (b) Suspended mass of arbitrary shape on a horizontal axis (rigid
pendulum).

Which w recognize, according to Eq. (4.2), as defining simple harmonic motion with
ω = g l.
By way of preparation for more complicated pendulums, note the alternative statement of
the problem in terms of the angular displacement θ. Using this, we have

 d 
r l  exactly
 dt 
1 2
y = l ( 1 – cosθ ) ≈ lθ
2
so that our approximate statement of energy conservation is now
2
1  d  1
ml    mgl 2  E
2  dt  2
Consider now an arbitrary object that is free to swing in a vertical plane. Let its centre of mass C
be a distance h from the point of suspension, as shown in Fig. 4.3 (b). Then the gain of potential
energy for an angular deflection θ is mghθ 2/2. The kinetic energy is the energy of rotation of the
body as a whole about O.
Since every point in the body has angular speed dθ/dt, this kinetic energy can be written
I(dθ/dt)2 /2 where I is the moment of inertia about the horizontal axis through O. Hence we have
2
1  d  1
I   mgh 2  E
2  dt  2

41
It is in many instances convenient to introduce the moment of inertia about a parallel axis
through the centre of mass. If this is written as mk 2 , where k is the “radius of gyration” of the
body then the kinetic energy of rotation with respect to the centre of mass is
mk2(dθ/dt)2/2, to which must be added the kinetic energy associated with the instantaneous linear
speed h(dθ/dt) of the centre of mass itself. Thus the energy- conservation equation may also be
written as follows:-
2 2
1  d  1  d  1
mk 2    m h   mgh 2  E
2  dt  2  dt  2
from which we have
gh
2  gh
h2  k 2
1 (4.11)
 h2  k 2  2
T  2  

 gh 

NOTE: In the triangle ONP (Fig. 4.3a), we have (by Pythagoras’s


theorem)

l2 = (1 – y)2 + x2

Hence x2 = 2ly – y2 ≈ 2ly

4.6 Water in a u – tube

Fig. 4.4: Oscillating liquid column in a U-tube.


If a liquid is contained in a U- tube arrangement of constant cross section with vertical arms as
shown in Fig.4.4, we have a system that resembles the pendulum in that, although the motion is
two–dimensional, it can be completely described in terms of a single vertical displacement y of
the liquid surface from equilibrium. Suppose that the total length of liquid column is l and its

42
cross section is A. Then, if  is the liquid density, the total mass in of the liquid is  Al. We
shall assume that every part of the liquid moves with the same speed, dy/ dt.
The increase of gravitational potential energy in the situation shown in Fig. 4.4 corresponds to
taking a column of liquid of length y from the left–hand tube raising it through the distance y,
and placing it on the top of the right–hand column. Thus we can put

U = g  Ay2
The conservation of mechanical energy thus gives the following equation:
2
1  dy 
Al    gAy 2  E
2  dt 

Hence
2g
ω2 =
l

l 2l
T = 2  (4.12)
2g g

Note the similarity to the simple pendulum equation, but also the subtle difference – that a liquid
column in these circumstances has the same period as a simple pendulum of length l/2.

4.7 The decay of free vibrations


The free vibrations of any real physical system always die away with the passage of time. Such
system has dissipative features through which the mechanical energy of the vibration is depleted.
Our very knowledge of the existence of a vibrating system is likely to imply a loss of energy on
its part. Thus it is never strictly correct to describe these free vibrations mathematically by
sinusoidal variation of constant amplitude. We shall now consider how the equation of free
vibrations is modified by the introduction of dissipative forces.

43
Fig. 4.5: Showing decay of the vibration
The resistive force of a fluid to a moving object is some function of the velocity of the object; it’s
magnitude is described as:

R(v) = b1 v + b2 v2
Where v is the magnitude ‫׀‬v ‫ ׀‬of the velocity. This resistive force is exerted oppositely to the
direction of v itself. Provided v is small compared to the ratio b1/ b2, we can take the resistive
force to be given by the linear terms alone. In this case the statement of Newton’s law for the
moving mass can be written.
d 2x
m 2   kx  bv
dt
i.e
d 2x dx
m 2  b  kx  0 (4.13)
dt dt
or
d 2x dx
  0 x  0
2
2
dt dt

where
b k
γ= ωo2 = (4.14)
m m
It may be seen, then, that in this case the damping is characterized by the quantity γ, having the
dimension of frequency, and the constant ωo would represent the angular frequency of the system
if damping were absent.
Let us now seek a solution of Eq. (4.14). We shall do this by the complex exponential method,
by assuming that x is the real part of a rotating vector z, where z satisfies an equation like Eq.
(4.14), i.e.,
d 2z dz
  0 z  0
2
2
(4.15)
dt dt
we shall assume a solution of the form

z= Aej(pt + α) (4.16)
just like eq. (4.8), and containing the requisite two constants, A and , for the purpose of
adjusting our solution to the initial values of displacement and velocity. Substituting in Eq.
(4.15) we find
(-p2 + jpγ + ωo2) Aej(pt+) = 0
If this is to be satisfied for all values of t, we must have
-p2 + jpγ + ωo2 = 0 (4.17)

44
This condition is one involving complex numbers i.e., it really contains two conditions, applying
to the real and imaginary components separately. It cannot be satisfied if the quantity p is purely
real, because the term jpy would then be a pure imaginary quantify with nothing to conceal it.
We therefore put

p = n + js
where n and s are both real. Then

p2 = n2 + 2jns – s2
Substituting these in Eq. ( 4.17) gives the following:
-n2 – 2jns + s2 + jnγ – sγ + ωo2 = 0
We thus have two separate equations:
Real parts: - n2 + s2 – sγ + ω o2 = 0
Imaginary parts: - 2ns + nγ = 0
From the second of then we get


s
2

Substituting s  in the first equation then given
2

2 γ2
n 2  ωo 
4

Now look back to Eq. (4.16). Writing p as a complex quantity n + js, we have

z  Ae j(nt  jst   )
 Ae  st e j(nt   )
and hence

x = Ae–st cos(nt + )
Substituting the explicit value of n and s we thus find the following solution:

t
x=A  2 cos (ωt+) (4.18)
e
where
2 k b2
 2  o 2    (4.19)
4 m 4m 2

Figure 4.6 shows a plot of Eq. (4.18) for the particular case  = 0. The enveloped of the damped
oscillatory curve is also plotted in the figure. The zeros of the curve are equally spaced with a
separation of ωt = π, and so are the successive maxima and minima, but the maxima and

45
minima are only approximately halfway between the zeros. Clearly ω may be identified as the
natural angular frequency of the damped oscillator.
The curve in fig. 4.6 is drawn for a case in which the decay of the vibrations is rapid. If,
however, the damping is small, the motion approximates to SHM at constant amplitude over a
number of cycles. Under these conditions one can express the effect of the damping in terms of
an exponential decay of the total mechanical energy, E. For, if γ << ω, we can say that around
time t the oscillations are well described over several cycles by SHM of constant amplitude A
such that:-
A (t) = Ao e–γt/2 (4.20)

Fig. 4.6: Rapidly damped harmonic oscillation.


The curve in fig. 4.6 is drawn for a case in which the decay of the vibrations is rapid. If,
however, the damping is small, the motion approximates to SHM at constant amplitude over a
number of cycles. Under these conditions one can express the effect of the damping in terms of
an exponential decay of the total mechanical energy, E. For, if γ << ω, we can say that around
time t the oscillations are well described over several cycles by SHM of constant amplitude A
such that:-
A (t) = Ao e–γt/2 (4.20)
Now the total mechanical energy of a simple harmonic oscillator is given by:-

1
E= kA2
2
Hence, using the above value of A, we have
1
E(t) = kAo2e-γt
2
i.e.

46
E(t) = Eoe-γt (4.21)
This decay of the total energy is illustrated in Fig. 4.4.

Fig. 4.7: Exponential decay of total energy during the damping of harmonic oscillations.

Recall that the damping process has an assumption that dissipation is due to a resistive force
proportional to the velocity. The situation would be quite difficult if some other resistive law
applied – e.g., R(v) ~ v2 . It is worth pointing out, however, that the exponential decay of
energy as described by Eq.(4.20) may and does arise from any diverse kinds of dissipative
processes.
From the foregoing analysis, it is clear that the damped oscillator is characterized by two
parameter; ωo and γ (=b/m). The constant ωo is the angular frequency of undamped oscillations

and γ is the reciprocal of the time required for the energy to decrease to 1
e of it’s initial value.
Thus ωo and r γ are quantities of the same dimensions. For convenience in applying our results
to diverse kinds of physical systems, we define a parameter called the Q value (Q for quality) of
the oscillatory systems given by the ratio of these two quantities.

0
Q= (4.22)

Q is a pure number, large compared to unity for oscillating systems with small rates of
dissipation of energy. In terms of the Q value Eq. (4.19) becomes.

 1 
 2  o 2  1   (4.23)
 4Q 2 
If Q is large compared to unity, and this important case is the one with which we shall be mainly
concerned, Eq. (4.23) gives ω ≈ ω o and the motion of the oscillator (Eq.4.18) is given very
nearly by

 0 t
x = Aoe 2Q cos (ωot + α) (4.24)

47
It may be noted that Q is closely related to the number of cycle of oscillation over which
the amplitude of oscillation falls by a factor l. For according to Eq.(4.24) we have

 0 t
A (t) = Aoe 2Q

Let us measure the time t in terms of the number of complex cycles of oscillations, n. Then given
the approximation that ω ≈ ω o, we can put t ≈ 2πn/ωo. In terms of the number of cycle elapsed,
therefore, we can put
 n
A(n) ≈ Aoe Q (4.25)
so that the amplitude falls by a factor e in about Q/π cycles of free oscillations.
In terms of ωo and Q, we can rewrite e.g. (4.15) in the form.
d 2 x  0 dx 2
2
  0 x  0 (4.26)
dt Q dt
and this will, in many cases, be a highly convenient form of the basic differential equation of free
oscillations, including damping, of a great variety of physical systems, both mechanical and non-
mechanical.

Example.
An oscillating block and spring system is found to experience a small frictional force which is
proportional to the velocity at which the block moves. Taking the coefficient of the velocity as b,
 1 
show that the solution to this motion is given as 
 2
  a t 
 where  =b/m
xe

The effect of very large damping


We are now familiar with equation

2
 2  o 2 
4

 
k
  m
1
2 

   b 2m  ?
But what if ωo is less than 2
In this case the motion is no longer oscillatory at all. We found that the differential
equation of motion Eq. (4.15) is satisfied by a solution of the form

 t
 e j  nt   
x = Re  Ae 2
 

where
2
n2 = ω o2 –
4

48
Suppose now that ωo2 < 
2
. Then we can put
4

 2 
n2 = -    0 
2

 4 
And if we proceed to solve for n we have
1
 2 2 
2
n = ±j    0    j , say
 4 
Thus we have ejnt = e± βt, which would define an exponential decay of x with t according to one or
other of two possible exponents:

     t      t
e  2  or e  2 

A vigorous analysis shows that both exponentials are in general necessary, and that the
complete variation of x with t is given by the following of equation.

     t      t
x = A e  2 
 A2 e  2  (4.27)
1
where
1
 2 2 
2
β=    0  4
 4 
The two adjustable constants A1 and A2 (which may be of either sign) allow for the
solution to be fitted to any given values of x and dx/dt at a given instant, eg., t = 0
What happens if ω 0 and γ/2 are exactly equal to one another? In this case the right side of Eq.
(4.27) would reduce to two terms of exactly the same type and only one adjustable constant
would remain. This is not, however, an acceptable solution any longer: we still need two
adjustable constants. It turns out that the appropriate form of solution for this case is

t
x = (A + βt )e  2 (4.28)
You can verify by substituting that this satisfy the basic equation of motion Equation 4.15 if
ω0 = γ/2 or γ = 2ω0 exactly. This very special condition corresponds to what is called critical
damping. In real mechanical systems the value of the damping constant γ is often deliberately
adjusted to meet this condition. Because, under conditions of critical damping, a constant force
suddenly applied to the system will be followed by a smooth approach to a new displaced
position of equilibrium with no oscillation or overshoot. Such behavior is advantageous in the
moving parts of electrical meters and the like, with which one may want to take a steady reading
as soon as possible after the meter has been connected or a switch closed.

4.8 Forced vibrations and resonance

49
Undamped oscillator with harmonic forcing
A sinusoidal driving force F = F ocos ωt is applied to a spring of mass in with a spring constant k.
The value of k / m , representing the natural angular frequency of the system is denoted by ωo.
Then the statement of the equation of motion, in the form ma = net force, is

d 2x
m  kx  F0 cos t
dt 2

or

d 2x
m  kx  F0 cos t (4.29)
dt 2
If the oscillator is driven form its equilibrium position and then left to itself, it will oscillate with
its natural frequency ωo. A periodic driving force will, however, try to impose its own frequency
ω on the oscillator. Therefore the actual motion in this case is of superposition of oscillations of
the two frequencies ω and ωo. The initial stage, in which the two types of motion are both
prominent, is called transient.
After some long time, however, the only motion effect present is the forced oscillation, which
will continue undiminished at a frequency ω. When this condition has been achieved we have
what we call a steady–state motion of the driven oscillator.
The most striking feature of the motion will be the large response near ω = ω o. If the driving
force is of very low frequency relative to the natural frequency of free oscillations, the particle
will move with an amplitude of Fo/k (= Fo/mwo2), this means that the term m(d2x/ dt2) in Equation
4.29 plays a relatively small role compared to the term kx at very low frequencies, or in other
words the response is controlled by the stiffness of the spring. On the other hand if frequencies of
the driving force very large compared to the natural frequency of free oscillation, the opposite
situation holds. The term kx become small compared to m(d 2x/ dt2) because of the large
acceleration associated with high frequencies, so that the response is controlled by the inertia. In
this case we expect a relatively small amplitude of oscillation and this oscillation should be
opposite in phase to the driving force, because the acceleration of a particle in harmonic motion
is 1800 out of phase with displacement.
To obtain the steady–state solution of Equation 4.29 we set

x = cos ωt (4.30)
We are assuming that the motion is harmonic, of the same frequency and phase and the driving
force, and that the natural oscillations of the system are not present. Differentiating Equation
4.30 twice with respect to t, we get
d 2x
  2 C cos t
dt 2
substituting in Equation 3.9 we thus have
mω 2cos ωt + kC cosωt = Fo cosωt

50
and hence
F0 F m
C=  20 2 (4.31)
k  m  0  
2

Equation 4.31 satisfactorily defines C in such a way that Equation 4.29 is always satisfied. Thus
we can take it that the forced motion is indeed described by Equation 4.30, with C depending on
ω according to equation 4.31. This dependence is shown graphically in Figure 4.8. Notice how C
switches abruptly from large positive to large negative values as ω passes through ωo. The
resonance phenomenon itself is represented by the result that the magnitude of C, without regard
to sign, becomes infinitely large at ω = ω o exactly.

Fig.4.8: Amplitude of forced oscillations as a function of the driving frequency (assuming zero
damping.) The negative sign of the amplitude for ω > ω o corresponds to a phase lag π of
displacement with respect to driving force.
To express x in terms of a sinusoidal vibration having an amplitude A, by definition a positive
quantity, and a phase α at t = 0.

x = A cos (ω t + α) (4.32)
It is not difficult to see that this implies putting A = ‫׀‬C‫ ׀‬and giving α one or other of two values,
according to whether the driving frequency w is less or greater than ωo
ω < ωo : α = O
ω > ωo : α = π

51
The response of the system over the whole range of ω is then represented by separate curves for
the amplitude A and the phase α, as shown in Figure 4.9. The infinite value of A at ω = ωo, and
the discontinuous jump from zero to π in the value of α as one passes through ωo, must be
unphysical,

Fig. 4.9: (a) Absolute amplitude of forced oscillations as a function of the driving frequency, for
zero damping (b) Phase lag of the displacement with respect to the driving force as a function of
frequency.

Once the steady state has been established, the pendulum behaves as though if were suspended
from a fixed point corresponding to a length greater than its true length l for
ω < ω o and less than l for ω > ωo

4.9 The complex exponential method for forced oscillations


1. We start with the physical equation of motion as given by Equation 4.29
d 2x
m 2  kx  F0 cos t
dt
2. We imagine the driving force FoCost as being the projection on the x axis of a rotating vector
Fo exp(jωt), and we imagine x as being the projection of a vector z that rotates at the same
frequency ω.
1. We then write differential equation that governs z:

52
d 2z
m 2
 kz  F0 e jt (4.33)
dt
We try the solution
z = Aej(ωt + α)
Substituting in Equation 4.33 this gives us
(- mω 2A + kA) ej(ω t + α) = Foejω t

which can be rewritten as follows:

 ω 2  ω 2  A  Fo e  jα
 o  m
(4.34)
F F
 o cosα  j o sinα
m m
This contains two conditions, corresponding to the real and imaginary parts on the two sides of
the equation.
F0
(ωo2 – ω 2) A = cos 
m
F0
O=- sin α
m
This clearly lead at once to the solution represented by the two graphs in Figure 4.9.

4.10 Forced oscillations with damping


The statement of Newton’s law is
d 2x dx
m 2  kx  b  F0 cos t
dt dt
or
d 2 x b dx k F
2
  x  0 cos t
dt m dt m m

Putting k/m = wo2, b/m = γ, this can be written

d 2x dx F
   0 x  0 cos t
2
2
(4.35)
dt dt m

53
Let us now look for a steady state solution to this equation. Our basic equation then becomes in
complete – exponential.
d 2z dz 2 F
2
   0 z  0 ejωt (4.36)
dt dt m
We assume the following solution:
z = A ej(ω t – δ) (4.37)
With
x = Re(z)
Notice that we have assumed a slightly different equation for z than we did in the previous
section; the initial phase of z is now – δ instead of + α. This due is in equation 4.34 right-hand
side of the equation can be read, in geometrical terms, as an instruction to take a vector of length
Fo/m and rotate it through the angle –α with respect to the real axis. We are going to get a similar
equation now, and it will simplify things if we define our angle, formally at least, as representing
a positive (counterclockwise) rotation. That is, δ is formally a positive phase angle by which the
driving force leads the displacement. Substituting for equation 4.36 into Equation 4.36 we thus
get
  F
  2 A  jA   0 A e j  t    0 e jt
2

m
Therefore
F
(ωo2 – ω 2)A + jγωA = 0 e j (4.38)
m
Now the elegance and perspicuity of the complex exponential method are really displayed we
can equation 4.38 as a geometrical statement. The left-hand side tells us to draw a vector of
length (ωo2 – ω 2)A, and then at right angles to it a vector length γωA. The right-hand side tells us
to draw a vector of length Fo/m at an angle δ to the real axis. The equation requires that there two
operations bring us to the same point so that the vectors form a closed triangle as shown in
Figure 4.10(a)

54
Note: The left-handed side Eq.(3.18) may we written (in terms of its origins)
as a sum of three vectors.
ωo2A + jγωA + (j)2ω 2A
as shown in Fig. 4.2.3 (b)

Fig. 4.10: Geometrical representation of equation 4.38


Clearly, we have
F
(ωo2 – ω 2) A = 0 cos δ
m
F0
γωA = sin δ
m
Therefore,
F0 m
A( ) 
 
1
  2   2 2     2  2 (4.39)
 0 

55

tan δ(ω) =
0   2
2

These same results can of course be obtained without introducing complex exponentials. One
simply assumes a solution of the form.

x = A cos (ω t – δ) (4.40)
And substituting this equation 4.35, leads to:

F0
(ωo2 – ω 2)A cos(ωt – δ) – γωA sin(ωt – δ) = cos ωt
m
The type of dependence of amplitude A and phase angle δ upon frequency ω1 for an assumed
constant magnitude of Fo, is shown in Figure 4.11. (Remember that δ is the angle by which the
driving force leads the displacement, or by which the displacement lags behind the driving
force.)
These curves have a clear general resemblance to those in Figure 4.9 for the undamped oscillator.
As can be seen from the expression for tan δ in equation 4.39, the phase lag increases
continuously form zero (at ω = 0) to 1800 (in the limit ω → ∞); it passes through 900 as precisely
the frequency ωo. Less obvious is the fact that the maximum amplitude is attained at a frequency
ωm some what less than ωo; in most cases of any practical interest however, the difference
between ωm and ωo is negligibly small.

56
Fig. 4.11 (4.2.4): (a) Dependence of amplitude upon driving frequency for forced oscillations
with damping (b) phase of displacement with respect to driving force as a function of the driving
frequency.

4.11 Effect of varying the resistive term


We show how the behaviour of the resonant system changes as the Q of the system is changed,
other things remaining constant.
Substituting γ = ωo/Q into equation 4.39 we get
F0 m
A(ω( 
 
1
 ω 2  ω 2 2   ωω Q  2  2 (4.41)
 0 0 

 0 Q
tan δ(ω) = ωωo/Q
0   2
2

ωo2 – ω 2

Using the ration ω/ωo, rather than ω as a variable, equation 4.41 takes the form:

57
F0 0 
A
m0 2  1
1 
2 2
   
  0    2
   0  Q 

OR
(4.42)
F 0 
A 0
k 1
  2  2
  0     1 
   0  Q 2 
  

And
1Q
tan  
0 

 0

In Figure 4.12 we show curves calculated from equation 4.42 to show the variations with
frequency of amplitude A and phase lag δ for different value of Q. most of the change of δ takes
place over a range of frequencies roughly from ωo(1 – 1/Q) to ωo (1 + 1/Q), i.e. a band of width
2ωo/Q centred on ωo. In the limit Q → ω the phase lag jumps abruptly from zero to π as one
passes through ωo. Clearly the frequency ωo is an important property of the resonant system, even
through it is not (except for zero damping) the frequency with which the system would oscillate
when left to itself.
If we denote by Ao the amplitude Fo/k obtained for ω →0 then one can readily show that the
following results hold.
1
 1  2
 1  
2Q 2
ωm = ωo   (1 – 1) ½
2Q2

Q
Am  Ao 1
 1  2
 1  
 4Q 2  )

58
Fig. 4.12 : (a) Amplitude as function of driving frequency for different values of Q assuming
driving force of constant magnitude but variable frequency, (b)Phase difference δ as a function of
driving frequency for different values of Q.

4.12 Transient Phenomena


Our discussion so far has dealt on the steady state. But we now consider the transient state at
which initially t = 0, is at rest. As t = 0 the driving force is turned and thereafter the motion
governed by equation 4.29.
d 2x
m 2  kx  Fo cosωt
dt
or
d 2x F
  0 x  0 cos t
2
2
(4.43)
dt m
We have already seen how this differential equation of the forced motion leads to the following
equation for x:

59
F0 m
x cos t (4.44)
0 2   2
Suppose that we have found that we have also found a solution – call it x 1 – to equation 4.43 so
that
d 2 x1 F
  0 x1  0 cos t
2
2
dt m
And now suppose that we have also found a solution – call it x 2 to the equation of free vibration,
so that
d 2 x2
  0 x2  0
2
2
dt

By adding these two equations we get


d 2  x1  x2  F
  0  x1  x2   0 cos t
2
2
dt m
The equation of free vibration of frequency ωo does contain two adjustable constants – amplitude
and an initial phase. Let us call then B and β because we are using them to fit conditions at the
beginning of the forced motion.
We propose that the complete solution of the forced motion equation is as follows:

x = B cos(ωot + β) + C cos ωt (4.45)


Where
F0 m
C
0 2   2
Equation 4.45 can be tailored to fit the initial conditions (in this case) that x = 0 and dx/dt = 0 at t
= 0. For the condition on x itself we have

0 = B cos β + C
Also, differentiating equation 4.45 we have
dx
= - ω o B sin (ωot + β) – ωC sin ωt
dt
Hence, at t = 0, we have
0 = - ω oB sin β
The second condition requires that β = 0 or π. Taking the former (the final result is the same in
either case) we get B = - C, so equation 4.45 becomes;

x = C (cos ω t – cos ωot) (4.46)


which is typical example of beats. In the complete absence of damping these beats would
continue indefinitely; no steady state corresponding to equation 4.44 along would ever be

60
reached. It is perhaps worth noting that the condition just after t = 0 now make excellent sense. If
ωt, ω1t <<1, we can put

 2t 2
cos ωt ≈ 1 –
2

cos ω0t ≈ 1 –
 0 2t 2
2
Therefore

x ≈

F0 m 0   2 t 2 1 Fo 2
 t
2

2
0   2 2 2m
Thus the restoring forces have been called into play the mass starts out in the direction of the
applied force with acceleration Fo/m
In reality damping is usually assumed to be present. We therefore postulate the following
combination of free and steady state motions:

x = B e-γt/2 cos (ω1t + β) + A cos(ωt – δ) (4.47)


where
1
 2  2
 1  0 2
 

 4 
and A, δ are given elsewhere
Fitting the value of B and β to the value of x and dx/dt at t = 0 is just a complicated version of
what we did to the undamped oscillator.

4.13 The power absorbed by a driven oscillator


We are interested in knowing at what rate energy must be fed into a driven oscillator to maintain
its oscillations at fixed amplitude. Power input P, is the driving force times the velocity:

dW dx
P= F  Fv
dt dt
The undamped oscillator has no dissipative effects thus the mean power input must come out to
be zero. Assuming the steady state solution, we have
F = Focos ωt
F0 m
x = cos ωt = C cos ωt
0   2
2

ω o2 – ω 2

Therefore

61
v = - ωC sin ωt
P = - ωCFo sin ωt cos ωt

This power dx
input,v being proportional
Note: = and P =toFvsin 2ωt, is positive half the time and negative for the
dt
other half, averaging out to zero over any integral number of half-periods of oscillation. That is,
energy is fed into the system during one quarter cycle and is taken out again during the next
quarter –cycle.
Coming now to the forced oscillator with damping we have
x = A cos(ωt – δ)
Therefore
v = - ω Asin (ωt – δ)
We can only write this as

v = - vo sin (ωt – δ)
Where vo is the maximum value of v for any given value of F o and ω. Taking the value of A from
equation 4.42 we have
F00 k
v0 ( )  1
   
2
1 
2
(4.48)
  0    2
   0  Q 

The value v0 passes through a maximum at ω = ωo, exactly, a phenomenon that we call velocity
resonance. Now let us consider the work and power needed to maintain the forced oscillations.
We have.
P = - Fovo cos ωt sin (ωt – δ)
= Fovo cos ωt (sin ωt cos δ – cos ωt sin δ)
P = -(Fovo cosδ) sin ωt cos ωt + (Fovo sin δ) cos2ωt (4.49)
If we average the power input over any integral number of cycles, the first term in equation 4.49
gives zero. The average of cos2ω t, however, is ½, so that the average power input is given by

1 1
P = Fovo sin δ = ωAFo sin δ
2 2

1
NOTE: Recall for example that cos 2 t  1  cos 2t  and that  cos 2t  av  0
2
over a complete cycle.

With the help of equations 4.42 and 4.48 this becomes

62
F 
2
1
P ( )  0 0 2
2kQ     1 (4.50)
 0
   2
  0  Q
We see that this power input, like the velocity, through passes through a maximum of at precisely
ω = ωo for any Q. the maximum power is given by
F0 20Q QF0 2
Pm   (4.51)
2k 2m0
The dependence of P on ω for various Q is shown in Figure 4.13 (a). It may be noted that the
power input drops off toward zero for very low and high frequencies, and that, except for low Q,
the curve are nearly symmetrical about the maximum. It is convenient to define a width for these
power resonance curves by taking the difference between those values of ω for which the power
input is half of the maximum value. This can be done in a particularly clear and useful way if (as
in most cases of interest) Q is large. This means that the resonance is effectively contained within
a narrow band of frequencies close to ωo. It is then possible to write on approximate form of the
equation for P (ω), based on the following price of algebra:

Fig.4.13: (a) Mean power absorbed by a forced oscillator as a function of frequency for different
values of Q. (b) Sharpness of resonance curve determine in terms of power curve.

63
0  0 2   2
 
 0 0
     0   
 0
0

Hence, if ω ≈ ω0, we can put


 0  2 0      2 0   
  
 0 0
2
0
Substituting this in the denominator of equation 4.50, we have
2
F  1
P( )  0 0
2kQ 4 0    2 1
2

0 Q2
F0  0 Q 
2
1


2 k 0
2
 4 0      0 Q 
2 2

We have met the quantity ωo/Q before. It is the damping constant γ  b m was found to decay

in the absence of a driving force:


E = Eoe    0 Q  t = Eoe-γt (4.52)

Thus the above equation for P can be written (remember also that k = mωo2) in the following
simplified form:
F0
2
1
(Approximate) P    (4.53)
2m 4  0    2   2
The frequencies ωo ± Δω at which P (ω) falls to half of the maximum value P (ωo) are thus
defined by

4(Δω) 2 = γ2
ie
0
2  (4.54)
Q
Thus we find that the width of the resonance curve for the driven oscillator, as measured by the
power input (Figure 4.13(b), is equal to the reciprocal of the aims needed for the free
oscillations to decay to 1/e of their initial energy.

64
We can say that the resonance is narrow if the width is only a small fraction of the resonant
frequency, i.e., if

2
 1 (4.55a)
0
and we can say that the decay of free oscillations is slow if the oscillator losses only a small
fraction of its energy in one period of oscillation. Now from equation 4.52 we have

E
 t
E
If Δt we put the time 2π /wo, which is approximately equal to the period of the free damped
oscillation, we have
E 2

E 0
Thus a slow decay mean
2
 1 (4.55b)
0
Since γ = 2Δω = ωo /Q, the conditions described by equation 4.55a and 4.55b can both be
expressed by saying that the dimensionless quantity Q must be large.

4.13 ELECTRICAL RESONANCE


The resonant systems of electrical system made of a capacitor and a coil figure 4.14 shows
similarities to the mechanical systems with which we have already tackled. The capacitor is a
device for storing electric charge and the associated electrostatic potential energy. Its capacitance
C is defined as the measure of the charge q applied to the capacitor plates divided by the measure
of the voltage difference that this charge produces:
C = q/Vc
Vc = q/C

Fig 4.14: Capacitor and inductor in series: the basic electrical resonance system.

Under D-C conditions the coil offers no opposition to the flow of current, but if the current is
charging with time it is found that the coil (which we shall henceforth call an inductor) acts to
oppose the charge (Lenz’s law). Under these circumstances there is a voltage difference VL

65
between the ends of the inductor, and this voltage is proportional to the rate of change of the
current i. the inductance L is defined the relation
VL = L (di/dt)
This equation says that the voltage VL must be applied between the ends of the inductor in order
to make the current change at the rate di/dt. In a circuit made up of just these two components,
the sum VC and VL must be zero, because an imaginary journey through the capacitor and then
through the inductor brings us back to the same point on the circuit. Thus we have

q/C + L di/dt = 0 (4.56)

Now there is an intimate connection between q and i, because the current in the circuit is just the
rate of flow of charge past any point. A current i flowing for a time dt in the wire connected to a
capacitor plate will increase the charge on that plate by the amount dq = idt, so we have
i = dq/dt
di/dt = d2q/dt2
Hence equation (1) can be written

L(d2q/dt2) + (1/C)q = 0 (4.57)

But this is precisely like the basic differential equation of SHM for a mass-spring system, with q
playing the role of x, L appearing in the place of m, and 1/C replacing the spring constant k. We
can confidently assume the existence of free electrical oscillations such that
ωo = 1/√(LC)

(a) (b)
Fig. 4.15 : (a). Capacitor, inductor, and resistor in series. (b). Capacitor, inductor, and resistor in
series Driven by a sinusoidal voltage.

Now let us consider the effect of introducing a resistor, of resistance R, as in Figure 4.15(a). At
current i it is necessary to have a voltage VR ( = iR) applied between the ends of the resistor. Thus
the statement of zero net voltage drops in one complete tour of the circuit is as follows:
q/C + Ldi/dt =0
i.e.,
L(d2q/dt2) +R(dq/dt) + (1/C)q = 0

or
(d2q/dt2) +R/L(dq/dt) + (1/LC)q = 0 (4.58)

66
In this equation, R/L plays exactly the role of the damping constant γ, and such a circuit the
charge on the capacitor plates (and the voltage VC) will undergo exponentially damped harmonic
oscillations.

Example.
A 1.0 μf capacitor is charged to 50volts. The charging battery is then disconnected and a 10 mh
coil is connected across the capacitor, so that LC oscillations occur. What is the maximum
current in the coil? Assume that the circuit contains no resistance.

Solution.
The maximum stored energy in the capacitor must equal the maximum stored energy in the
inductor, from the conservation-of -energy principal. This leads to
½ qm2 = ½ Lim2
where qm is the maximum charge and im is the maximum current. Not that the maximum charge
and the maximum current do not occur at the same time but one-fourth of a cycle a part. Solving
for im and substituting CVo for qm gives

im = Vo√(C/L) = (50 volts) √(1.0x10-6farad/10x10-3henry) = 0.50 amp.

Finally, if the circuit is driven by an alternating applied voltage, we have a typical forced-
oscillator equation:

(d2q/dt2) + R/L(dq/dt) + (1/LC)q = (Vo/L) cos ωt (4.59)

Compare:

(d2x/dt2) + b/m(dx/dt) + (k/m)x = (Fo/m) cos ωt (4.60)

The connection between Equation 4.59 and 4.60 becomes even closer if one considers the energy
of the system. Just as Fdx is the amount of work done by the driving force F in a displacement
dx, so V dq is the amount of work done by the driving voltage V when an amount of charge dq
passes hrough the circuit. One can regard the oscillations as involving the periodic transfer of
energy between the capacitor and the inductor, with a continual dissipation of energy in the
resistor. Comparison of the mechanical and electrical equations a displacement is shown in the
table (1) below.

67
Table1: Comparison of the mechanical and electrical resonance parameters.

Mechanical system Electrical system

Displacement x Charge q

Driving force F Driving voltage V

Viscous force constant b Resistance R

Spring constant k Reciprocal capacitance 1/C

Resonant frequency √k/m Resonant frequency 1√L/C

Resonance width γ = b/m Resonance width γ = R/L

Potential energy ½ kx2 Energy of static charge ½ q2/C

Kinetic energy ½ m(dx/dt)2 = ½ mv2 Electromagnetic energy of moving


charge ½ L(dq/dt)2 = ½ Li2
Power absorbed at resonance Fo2/2b Power absorbed at resonance Vo2/2R

Example
Using the conservation-of-energy principal, the total energy in an LC circuit present at any
instant is given by
U = UB + UE = ½ Li2 + ½ q2/C.
where UB =1/2 Li2 is the energy stored in the magnetic field in the inductor and UE = ½
2
q /C is the energy stored in the electrical field in the capacitor at that instant. Assuming the
resistance to be zero, there is no energy transfer to Joules heat and U remains constant with
time, even though i and q vary. Derive the differential equation of oscillations for this circuit.

Solution.
Because U remains constant with time it shows
dU/dt = 0
therefore
dU/dt = d/dt (½ Li2 + ½ q2/C) = 0
= Li di/dt + q/C dq/dt = 0 (i)
Now, q and i are not independent variable, being related by
i = dq/dt (ii)
Differentiating yields
di/dt = d2q/dt2 (iii)
Substituting (ii) and (iii) into (i) leads:

68
L d2q/dt2 + 1/C q = 0 (iv)
which is the differential equation that describes the oscillations of a (resistanceless) LC circuit.
4.14 ACTIVITIES

a) The power input to maintain forced vibrations can be calculated by recognizing that this
power is the mean rate of doing work against the resistive force –bν. Satisfy yourself that
the instantaneous rate of doing work against this force is equal to bν 2. Using x=A cos (ωt-
δ), show that the mean rate of doing work is bω2A2/2.
b) Solve the steady-state motion of a forced oscillator if the driving force is of the form
F=Fosin ωt instead of Focos ωt.

4.15 FURTHER READING

1. (a). In an oscillating LC circuit, what value of charge, expressed in terms of the maximum
charge, is present on the capacitor when the energy is shared equally between the electric and
magnetic field? (b)How much time is required for this condition to arise, assuming the
capacitor to be fully charged initially? Assume that L = 10mh and C = 1.0μf.

2. Derive the expression for the oscillation energy of an LC electrical oscillator with the
potential amplitude Vo and the current amplitude Io.

4.6 SELF-TEST QUESTIONS

1. The power input to maintain forced vibrations can be calculated by recognizing that
this power is the mean rate of doing work against the resistive force –bν. Satisfy
yourself that the instantaneous rate of doing work against this force is equal to bν 2.
Using x=A cos (ωt-δ), show that the mean rate of doing work is bω2A2/2.
2. Paul a KU student constructed a damped oscillator with the following properties: m =
0.5kg, b = 5N-m-1sec and k = 100N/m. He made the oscillator to be driven by a force
F = Fo cos ωt, where Fo = 4N and ω = 20sec-1.
(a) What are the values A and δ of the steady-state response described by
x = A cos (ωt - δ)?
(b) How much energy is dissipated against the resistive force in one cycle?
(c) What is the power input?
3. Electrons in a CRT are deflected by the mutually perpendicular electic fields in a such a
way that the displacement at any time is given by x = E1sin (ωt + δ) and y = E2sin ωt. (i) If
E1 and E2 = 10cm and the phase angle δ is π/2 rad., show that the path of the electron is a
cycle whose radius is 10cm.
69
4. (i ) Show that the steady state complex amplitude of a damped oscillator driven by an
external force Feiωt is gives by the expression
A = F/[M(ωo2 - ω2) + iωb]
(ii) Explain the meaning of the symbols used and the conditions for resonance.
(iii) A machine of total mass 100kg is supported by a spring resting on the floor and its
motion is constrained to be in the vertical direction only. The system is lightly damped with
a damping constant of 942Nsm-1. The machine contains an eccentrically mounted shaft
which, when rotating at angular frequency ω, produces a vertical force on the system of
Foω2cos ωt, where Fo is a constant. It is found that resonance occurs at 1200r.p.m
(revolutions per minute) and the amplitude of vibrations in the steady state when the
driving frequency is (a) 2400r.p.m. and (b) very large. You may assume that the gravity is a
negligible effect on the motion.
/(k - mω2 + iωc)

3. 5. A particle P of mass 2 moves along the x axis attracted toward origin 0 by a force whose
magnitude is numerically equal to 8x. The particle P has also a damping force whose
magnitude is numerically equal to 8 times the instantaneous speed. (Figure below). Find
(a) the position and (b) the velocity of the particle at any time; (c) illustrate graphically
the position of the particle as a function of time t.

6. The differential equation that describes the oscillations of a mechanical (spring-mass)


system is given by
m d2x/dt2 + kx = 0
assuming no damping. (i) Show the equivalent equation of an electrical LC circuit
assuming no resistance. (ii). If the solution of the mechanical system is given by
x = Acos (ωt + α). Prove mathematical that the equivalent solution for the LC
circuit solution above is given by q = qm cos (ωt + Ф) where ω is the angular frequency and
Ф is the initial phase angle.

7. (a). In an oscillating LC circuit, what value of charge, expressed in terms of the


maximum charge, is present on the capacitor when the energy is shared equally between
the electric and magnetic field? (b)How much time is required for this condition to arise,
assuming the capacitor to be fully charged initially? Assume that L = 10mh and C = 1.0μf.

8. Derive the expression for the oscillation energy of an LC electrical oscillator with the
potential amplitude Vo and the current amplitude Io.

70
1. The power input to maintain forced vibrations can be calculated by recognizing that this
power is the mean rate of doing work against the resistive force –bν. Satisfy yourself that the
instantaneous rate of doing work against this force is equal to bν 2. Using x=A cos (ωt-δ),
show that the mean rate of doing work is bω2A2/2.
2. Paul a KU student constructed a damped oscillator with the following properties:
m = 0.5kg, b = 5N-m-1sec and k = 100N/m. He made the oscillator to be driven by a force
F = Fo cos ωt, where Fo = 4N and ω = 20sec-1.
(a) What are the values A and δ of the steady-state response described by
x = A cos (ωt - δ)?
(b) How much energy is dissipated against the resistive force in one cycle?

71
LECTURE FIVE

FOURIER ANALYSIS

5.1 Introduction

In this lecture we discuss Fourier Theorem and its application in the analysis of non-periodic
functions. We also discuss in detail on the vibrations of a string in any infinite choice for
both amplitude and phase of given mode.

5.2 Lecture Objectives

By the end of this lecture, you should be able to:


 Appreciate the concept of Fourier Theorem in the analysis of non-periodic
functions.
 Analyze vibrations of a string in any infinite choice for both amplitude and
phase of given mode.

5.3 FOURIER ANALYSIS.


Fourier’s theorem states that any periodic function can be expressed as the summation of a series
of simple harmonic terms having frequencies that are multiples of that of the given function.
(Actually the function must not have an infinite number of discontinuities during one period).
One refers to the function as being analyzed into its simple harmonic components; the function is
the synthesis of those components. Lt can be shown that the rectangular wave form illustrated in
Fig.1(a) (which extends from
x = -∞ to x = +∞) can be analyzed into a series of sine waves, and can be expressed as the
following summation:
y = 4a/π(sinx + 1/3sinx + 1/5sin5x + -----).
Each of the first three terms is represented by a dotted curve in the figure and the sum of these
three terms can be seen by inspection to be represented by the full curve. Fig.1 (b) shows the
resultant of the first fifteen terms.

5.4 FOURIER’S THEOREM

72
Fourier’s theorem states that any periodic function can be expressed as the summation of a series
of simple harmonic terms having frequencies that are multiples of that of the given function.
(Actually the function must not have an infinite number of discontinuities during one period).
One refers to the function as being analyzed into its simple harmonic components; the function is
the synthesis of those components. It can be shown that the rectangular wave form illustrated in
Fig.1(a) (which extends from
x = -∞ to x = +∞) can be analyzed into a series of sine waves, and can be expressed as the
following summation:
y = 4a/π (sinx + 1/3sinx + 1/5sin5x + -----).
Each of the first three terms is represented by a dotted curve in the figure and the sum of these
three terms can be seen by inspection to be represented by the full curve. Fig.1 (b) shows the
resultant of the first fifteen terms.

Fig. 5.1: Analysis of a square wave.

It can be seen that as more terms in the series are included in the summation, the resultant
approaches the rectangular wave form more closely. In this example, the Fourier analysis of the
function contains only sine terms which are initially in phase and which have frequencies which
are odd multiples of that of the function represented. In general the series also contains cosine
terms (which are equivalent to saying that it contains sine terms with various initial phases) and
there are terms with every integral multiple of the original frequency. In general there is also a
constant term.

73
This way of analyzing a periodic function into simple harmonic components can be extended to
non-periodic functions in the following manner. Suppose that the non-periodic function y = f(x)
has the form represented by the full line in Fig.2. Let Ф(x) be the periodic function which is
identical to f(x) in the region –π ≤ x ≤ +π but which repeats itself outside the region as indicated
by the dotted curve in the figure. Then, according to Fourier’s theorem as stated above, the
function Ф(x) can be analyzed into a series of simple harmonic terms. It is obvious that within
the range –π ≤ x ≤ +π the series also represents f(x). By proceeding in this way one can represent
any non-periodic function over any desired range by an appropriate series of simple harmonic
terms.

Fig. 5.2: Analysis of non-periodic function over a limited range.

These terms have frequencies which are multiples of that of Ф(x). in particular one can say that
in the range –π to +π the function f(X) can be written

n n
f ( X )  ao   an cos nX   bn sin nX (5.1)
n 1 n 1

π

 cosn1Xcosn 2 XdX  0 if n1  n 2
Now π and


 cos n1 X cos n2 XdX  if n1  n2




In addition

∫ cos n1X sin n2XdX = 0 for all values of n1 and n2. Hence by integrating each
–π

74
side (1) between –π and +π one has

ao = 1/2π ∫ f(X) dX.


–π

Similarly by multiplying each side of (1) by cos nX and integrating, one has

an = 1/n ∫ f(X) cos nX dX.


–π

And, by multiplying each side of (1) by sin nX and integrating, one obtains

bn = 1/ π ∫ f(X) sin nX dX.


–π

It is convenient to transform (1) into a summation of complex-exponential terms. One has


cos nX =1/2(einX + e-inX) and sin nX =1/2(e-inX - einX)
Therefore
an cos nX + bn sin nX = an/2 einX + an/2e-inX - ibn/2 einX + ibn/2e-inX
= Cn einX + C-ne-inX

where Cn = ½(an - ibn) and C-n = ½(an + ibn)


Hence, putting ao = co, one can write

n
f (X )   Cn einX
n 1

(5.2)

1
 f(X) e
 inX
Cn  dx (5.3)
2 

5.5 Extension of the Range. The Fourier Integral


Eq. (5.2) shows that the range –π ≤ X ≤ +π the non-periodic function f(X) may be represented as
the summation of infinite series of simple harmonic terms whose amplitudes are given by
equation 5.3. It will now be shown that the range can be extended as far as may be desired and
that it can be made to extend from -∞ to +∞. In the later case the function is expressed as a
continuous series of simple harmonic terms whose frequencies have every possible value
between -∞ to +∞.
To show that the range can be extended and that f(x) can be represented over the range
–l ≤ x ≤ +l , put x = lX/π. Then equation 5.2 and equation 5.3 give.
n=+∞

f(x) = ∑ Cn e(inπx/l) , (5.4)

75
n=-∞

where
+l

Cn = 1/2l ∫ f(x) e- inπx/l dx. (5.5)


–l

If one introduces the dummy variable x/ in place of x in equation (5.5)

n=+∞ +l

f(x) = ∑ 1/2l ∫ f(x/) einπ/l(x- x/) d x/. (5.6)


n=-∞ –l

and the summation on the right-hand side represents f(x) in the range –l ≤ x ≤ +l. in eq.(5.6) l
may be made as large as one pleases and it can be shown that one can extend the range in this
way to ±∞. Now the nth harmonic in the series has wavelength constant k = 2π/λ one has k n =
nπ/l .Similarly kn+1 = (n+1)π/l. Hence kn+1 – kn = π/l = ∆k (say).
One may then write eq. (5.6) in the form.

n=+∞ +l

f(x) = 1/2π ∑ ∆k ∫ f(x/) ei{ln∆k(x- x/)} d x/. (5.7)


n=-∞ –l

Now n∆k is simply kn and as l → ∆k → 0. Eq. (5.7) then leads to the Fourier integral:

+∞ +∞

f(x) = 1/2π ∫ dk ∫ f(x/) eik(x- x/) d x/. (5.8)


-∞ -∞

Hence one can write


+∞

f(x) = 1/√2π ∫ g(k) eikx dx. (5.9)


-∞

where
+∞

g(k) = 1/√2π ∫ f(x) e-ikx dx. (5.10)


-∞

This shows that the function f(x) is represented by a continuous series of simple harmonic terms
whose amplitudes are given by eq. (5.10). When two functions f and g are related in this way,
each is said to be the Fourier Transform of the other.
By writing eq. (5.4) as

76
n=+∞

∑ C-n e-inπx/l , one can reverse the sign in the exponents of eq. (5.9) and eq.
n=-∞

(5.10), but there must always be +I in one and –I in the other. The factor 1/√2π can be replaced
by 1/2π in eq. (5.9) and omitted from eq. (5.10). If one then puts k = k/2π,
dk = 2πdk/ , and one has
+∞

f(x) = ∫ g(k-/) e-2πikx dx/. (5.11)


-∞

where
+∞

g(k ) = ∫ f(x) e-2πikx dx/.


/
(5.12)
-∞

k/ (= 1/λ) is now the number of cycles per unit of x. thus, if one is thinking of a complicated
motion as the superposition of simple harmonic motion, x is time, λ is periodic, and k / is
frequency; f(x) is the displacement as a function of time, and g(k -/) is the amplitude of the simple
harmonic component of frequency k-/. If one is analyzing a complicated variation of some
quantity with distance, x is distance, λ is wavelength, and k / is know as spatial frequency. Eq.
(5.11) can be written

∞ ∞
/ 2πikx
f(x) = ∫ g(k ) e dk + ∫ g(-k/) e-2πikx dk/
/

o o

∞ ∞

= ∫ [g(k ) + g(-k )] cos 2πk x dk + i ∫ [f(x) - f(-x)] sin 2πk/x dk/ ,


/ / / /

o o

Similarly, eq. (5.12) gives

∞ ∞

g(k/) = ∫ [f(x) + f(-x)] cos 2πk/x dx - i ∫ [f(x) - f(-x)] sin 2πk/x dx ,


o o

and

∞ ∞

g(-k/) = ∫ [f(x) + f(-x)] cos 2πk/x dx + i ∫ [f(x) - f(-x)] sin 2πk/x dx ,


o o

Suppose, now that f(x) is real and an even function, ie. f(-x) = f(x). Then g(-k /) = g(k/) and one
has

77
f(x) = 2 ∫ g(k/) cos 2πk/x dx/. (5.13) a.
o

and

g(k ) = 2 ∫ f(x) cos 2πk/x dx.


/
(5.13) b.
o

Each is referred to as the Fourier cosine transform of the other.

5.6 Analysis of the string

Suppose we have a string of length L fixed at it two ends. Then it should be able to vibrate in any
of an infinite freedom of choice of both amplitude and phase of a given mode, we shall put
 nx 
yn(x,t) = An sin   cos  n t   n  (n π x) cos (ωnt – δ) (5.14)
 L 

Further more, we can imagine that all these modes are permitted to be present, so that the motion
of the string is completely specified by the following equation:

 nx 
y(x,t) =  An sin   cos  n t   n  (5.15)
n 1  L 
The quantities cos(ωnto – δn) can be treated as a set of fixed numbers, and the displacement of the
string at any designated value of x can be written as follows:

 nx 
y(x) =  Bn sin   (5.16)
n 1  L 
where Bn = An cos(ωnto – δn)
The frequency ωn are integral multiple of a fundamental frequency ω1. We now can write
An sin(n π x/L) as a constant coefficient Cn, and thus have.

y(t) = C
n 1
n cos (ω nt – δn) where ω n = nω1 (5.17)
This means that any possible motion of any point on the string is periodic in the time 2π/ω1,
where ω1 is the frequency of the lowest mode, and further that this periodic motion can be written
as a combination with suitable amplitudes and phases, of pure sinusoidal vibrations comprising
all possible harmonic of ω1. This then is the Fourier analysis in time, rather than in space.

5.7 Fourier analysis in action.


Consider the expansion for y (x) as given by Eq.(5.15)

78

 nx 
y(x) = B
n 1
n sin 
 L 

Suppose we want the amplitude associated with particular value of n – say n1. To find it we
multiply both sides of the equation by sin (n1πx/L) and integrate with respect to x over the
range from zero to L.
 n1x   nx   n1x 
L  L

 y ( x) sin  dx   B  sin   sin   dx


0
L  n 1 o  L   L 
(5.18)
On the right we still appear to have an infinite series of terms. But now consider the properties of
an integral whose intergrand is a product of sines. Given any two angles θ, and φ, we have
cos (θ – φ) = cos θ cos φ + sin θ sin φ
cos (θ + φ) = cos θ cos φ - sin θ sin φ

Therefore

sinθsinφ = ½ [cos (θ – φ) – cos (θ + φ)]


Here we can put
 nx   n x  1    n  n1 x    n  n1 x  
sin   sin  1    cos    cos   
 L   L  2  L   L  
Therefore,
 nx   n x  L   n  n1 x  L   n  n1 x 
 sin L 
 sin  1  dx 
 L 
sin 
2  n  n1   L   sin 
 2 (n  n1 )  L 

If we insert the limits x = 0, x = L, the values of sin(n+n 1)πx/L are all zero. Thus at first sight it
would appear that we have got rid of the right-hand side of Eq.(5.5) altogether.
But we note that the quantity (n-n1) appears in the denominator of one of the integrals. Thus if n
=n1, we have one integral of the form 0/0. And at once turns out that, although all other terms are
zero, this one is not. For if n = n1, the integral to be evaluated is the following:

 n1x  1   2n1x  
L L

0 sin  L  dx  2 0  1  cos L   dx
2

The cosine term contributes nothing between the given limits, but the other part gives us L/2.
Thus we arrive at the following identity:
 nx 
L
L
 y x  sin 
0
L 
 dx  Bn
2
i.e.
 nx 
L
2
Bn =  y  x  sin   dx (5.19)
L0  L 

79
This equation determines for us the amplitude B n associated with any given value of n in the
harmonic analysis of y(x).
The evaluation of the Bn’s then proceeds as follows:
 nx 
L
2
Bn =  kx sin   dx
L0  L 
Integrating by parts, we find
2k  L   nx  
L
L
L
 nx  
Bn =
L
 
 n  x cos L  
   

n  cos
0
 dx 
L  
0

2k    nx  
L
L   nx  
L

=    x cos     sin    
n    L   0
n   L   0 

2kL cos n
= -
 n
One recognizes that the value of Bn fall into two categories, according to whether n is odd or
even, because the value of cos nπ alternates between the values +1 and –1. We have, in fact,
n odd:

2kL
Bn =
n
n even:

2kL
Bn = -
n
If one wishes, however, can represent both sets by means of the single formula
2kL
Bn =   
n 1

n
It is now an easy matter to tabulate the various amplitudes (table 5.1). Thus our description of the
triangular profile becomes
2kL   x  1  2x  1  3x  
y(x) =  sin    sin    sin   .....
   L 2  L  3  L  

Table 5.1: Values of Bn/kL

N Bn/kL
1 2
= 0.636

80
2 1
 = -0.318

3 2
= 0.212
3

4 1
 = - 0.156
2

5 2
= 0.127
5

The sine curves in terms of which this Fourier analysis is made represents an example of what
are called orthogonal functions. The orthogonality of two sine functions in Fourier analysis is
described by the results.

 n1x   n x 
L

 sin 
0
L 
 sin  2  dx  0
 L 
for n1 ≠ n2 (5.20)

If we have two vectors, A and B , the condition that they are orthogonal (perpendicular) to
each other is that their scale product be zero. In terms of their components this can be written
AxBx + AyBy + A2B2 = 0 (5.22)
Now if we replace continuous integral Eq. (5.7) by a summation over a large number, N 1 of
separate terms (as we might do if we were evaluating the integral by numerical methods), a
particular value of x could be written as xp, where

pN
xp =
N

Thus Eq.(5.7) would be replace by the following statement.

L N
 n1p   n p 
N
 sin
p 1 N 
 sin  2   0 for n1 ≠ n2
 N 

If we write the condition for orthogonality of two ordinary vectors in the form
3

A
p 1
p Bp  0 for Δ  B

we see that, in a purely formal sense, the difference between the two statements is merely that
one of them involves quantities that are completely described by just three components, whereas
the other need N components (and, in the limit infinitely many).

81
5.8 ACTIVITIES
(a) Expand f(x) =x, 0 < x < 2, in half range (a) sin series (b) cosine series.

(b) Prove that by making suitable assumptions concerning term by term integration of
infinite series that for n=1, 2, 3, …,
l

(i) an = 1/l ∫ f(x) cos (nπx/l) dx,


-l

(ii) bn = 1/l ∫ f(x) sin (nπx/l) dx,


-l

(iii) = ao /2.

5.9 FURTHER READING

1) Grafakos, L. (2009). Modern Fourier analysis . New York: Springer.


2) Grafakos, L. (2008). Classical Fourier analysis (2nd ed.). New York: Springer.
3) Stade, E. (2005). Fourier analysis . Hoboken, N.J.: Wiley-Interscience.
4) Howell, K. B. (2001). Principles of Fourier analysis . Boca Raton, Fla.: Chapman &
Hall/CRC.
1. Find the Fourier series for the following function (0 ≥ x ≤ L):
5) Walker, J. S. (1988). Fourier analysis . New York: Oxford University Press.
a. y(x) = Ax(L - x)
b. y(x) = A sin(πx/L)
 A sin  2x L  ,0  x  L 2
c. y(x) = 
 0, L 2  x  L.
2. Find the Fourier series for the motion of a string of length L if
a) y(x,0) = Ax(L-x); (δy/δt) = 0
5.10 SELF-TEST QUESTIONS t=0
b) y(x,0) = 0; (δy/δt)t=0 = Bx(L-x).
3. (a)Find the Fourier coefficients corresponding to the function
f(x)   0
3
 5 x  0
0 x 0 Period  10
(b)Write the corresponding Fourier series.

4. A string with fixed ends is picked up at its centre a distance H from the equilibrium
position and release. Find the displacement at any position at any time.

l l

5. Prove ∫ sin (kπx/l) dx = ∫ cos (kπx/l) dx = 0 if k = 1, 2, 3,...


-l -l

82
LECTURE SIX

COUPLED OSCILLATORS AND NORMAL MODES

6.1 Introduction

A real physical system, however, is usually capable of vibrating in many different ways, and may
resonate to many different frequencies. We speak of these various characteristic vibrations as
modes or for reasons that emerges as normal modes of the system. We shall begin by discussing
in some details the properties of a system of just two coupled oscillators. We also discuss in
coupled oscillators in transverse vibrations.

6.2 Lecture Objectives

By the end of this lecture, you should be able to:


 Understand the concept of modes in vibrations
 Derive equations of free and forced coupled
83 oscillators.
 Derive equations of normal frequencies in coupled oscillators.
 Appreciate resonance in coupled oscillators in transverse vibrations
So far we have confined our analysis to the systems having natural frequency. A real physical
system, however, is usually capable of vibrating in many different ways, and may resonate to
many different frequencies. We speak of these various characteristic vibrations as modes or for
reasons that will emerge later, as normal modes of the system. We shall begin by discussing in
some details the properties of a system of just two coupled oscillators.
6.3 Two coupled pendulum
Take two identical pendulums, A and B and connect them with a spring whose relaxed length is
equal to the distance between the pendulum bobs as shown in Fig. 6.1. Draw pendulum A aside
while holding B fixed and then release both of them. What happens?
Pendulum A swings from side to side, but it’s amplitude of oscillation continuously decreases.
Pendulum B, initially undisplaced, gradually begins to oscillate and its amplitude continuously
increases. Soon, A and B have equal amplitudes. The amplitude of A still continues to decrease
and that of B to increase until eventually the displacement of B is equal (or about equal) to that
originally given to A, and the displacement of A diminishes towards zero. The starting condition
is almost reversed. The energy, originally given to A (and to the spring) does not remain
confined to the oscillation of A, but is transferred gradually to B and continued to shuffle back
and forth between A and B.

84
Fig. 6.1: (a) coupled pendulum in equilibrium portion (b)Coupled pendulums with one
pendulum displacement

6.4 Symmetry considerations


Suppose we draw both A and B aside by equal amounts [Fig. 6.2 (a)] and them release them.
The distance between them equals the relaxed length of the coupling spring and therefore the
spring exerts no force on either pendulum. A and B will oscillate in phase and with equal
amplitudes always maintaining the same separation. Each pendulum might just as well be free
(uncoupled). Each oscillates with it’s free natural frequency ωo (= g l ). The
equations of motion are: -
x A  C cos ωo t
x B  C cos ωo t
(6.1)

Where xA and xB are the displacements of each pendulum from it’s equilibrium position. This
represents a normal mode of coupled system. (Both masses vibrate at the same frequency and
each has constant amplitude (the same for both).

Fig. 6.2: (a) Lower normal mode of two coupled pendulums (b) Higher normal mode of
two coupled pendulums

How many normal modes can we find? There is only one another. Draw A and B aside by equal
amounts but in the opposite direction (Fig. 6.2 (b)) and then release them. Now, the coupling
spring is stretched; a half – cycle later it will be compressed, and it does exert forces. The
symmetry of the arrangement tells us that the motions of A and B will be mirror images of each
other.
If the pendulums were free and either one were displaced a small distance x, the restoring force
would be mω02x. But in the present situation the coupling spring is stretched (or compressed) a
distance 2x and exerts a restoring force of 2kx, where k is the spring constant.
Thus the equation of motion for A is

85
d 2 xA
m 2  m 0 x A  2kx A  0
2

dt
or
d 2 xA
dt 2
 2 2

  0  2 c x A  0

where we have let ω c2 = k/m. This is an equation for SHM of frequency ω1 given by
1

ω1 = (ωo2 + 2ω c2) ½ =  g  2k 
2

 l m

For the given starting conditions, its solution is


xA = D cos ω1t (6.2a)
The motion of B is the minor image of A, and therefore
xB = -D cos ω1t (6.2b)
Each pendulum oscillates with SHM, but the action of the coupling spring has been to increase
the restoring force and therefore to increase the frequency over that of the uncoupled oscillation.
The motion of A and B are clearly always 1800 out of phase in this type of oscillation, which
constitutes the normal mode.
It is perhaps worth pointing out that if either of the pendulum is clapped, the regular frequency of
the other, under the action of the gravity plus the coupling spring, is equal to (ω02 + ω c2) ½. Thus
if one chooses to regard this motion as being, in a send the motion characteristic of our pendulum
alone, the normal modes have frequencies that are displaced above or below the single-
pendulum value.
6.5 The super position of the normal modes
Take an arbitrary moment when pendulum A is at xA and pendulum B at xA (Fig. 6.3). The spring
is stretched an amount xA – xB and therefore pulls on A and B with a force whose magnitude is
k(xA – xB). Thus the magnitude of the restoring force on A is
mωo2xA + k(xA – xB)
and on B it is
mωo2xB + k(xA – xB)

86
Fig.6.3: Coupled pendulums in arbitrary configuration.

Therefore, the equation of motion for A and B are


d 2 xA
 m 0 x A  k  x A  x B   0
2
m 2
(6.3)
dt
d 2 xB
 m 0 x B  k  x A  x B   0
2
m 2
dt
Again letting ω c2 = k/m1 we can write these as follows:

d 2 xA
dt 2
 2 2

  0   c x A   c xB  0
2

(6.4)
d 2 xB
dt 2
2
2 2

  0   c xB   c x A  0

The first, describing the acceleration of A1 contains a term in xB. And the second equation
contains a term in xA. These two differential equations cannot be solved independently but must
be solved simultaneously. A motion given to A does not stay confined to A but affects B and vise
versa.
Adding Eqs. (6.4) together we get
d2
 x A  xB    0 2  x A  xB   0
dt 2
If we subtract the second equation from the first, we get
d2
2
 x A  xB    0  2 c  2  x A  xB   0
dt

87
These are familiar equations for simple harmonic oscillations. In the first, the variable is x A + xB
and the frequency is ωo. In the second, the variable is xA – xB and the frequency is ω1.
where ω1 = (ωo2 + 2ωc2)½
Letting xA + xB = q1 and xA –xB = q2, we get two independent equations in q1 and q2:
d 2 q1
  0 q1 = 0
2
2
dt
d 2 q2
   q1 = 0
2
2
dt
Possible solutions (although not the most general ones) are
q1 = C cos ωot
(Special case) (6.5)
q2 = D cos ω1t
where C and D are constants which upon the initial conditions. {The lack of generality in Eq.
(6.5) can be recognized in that we have set the initial phases equal to zero.}
We have two independent oscillations. They represent another description of the normal modes,
as represented by oscillations of the variables q1 and q2 respectively, and these variables are
consequently called normal coordinates. Changes in the value of q 1occurs independently of q2
and vise versa.
In terms of our original coordinates, xA and xB, the solution are
xA = ½ (q1 + q2) = ½ Ccos ωot + ½ D cos ω1t (special case) ( 6.6)
xB = ½ (q1 – q2) = ½ Ccos ωot – ½ D cos ω1t
If C = 0, both pendulums oscillate with frequency ω1, or if D = 0, with the frequency ωo. These
are the frequencies of the individual normal modes and are called normal frequencies. We see
that a characteristic of a normal frequency is that both xA and xB can oscillate with that frequency.
The initial conditions (at t = 0) we have

dx A dx B
xA = Ao,  0, xB = 0, 0
dt dt
It may be noted that the conditions on the initial velocities are automatically met by eqs. (6.6),
because differentiation with respect to t gives us terms in sin ωot and sin ω1t only, all of which go
to zero at t = 0. From the conditions on the initial displacements themselves we have
xA = Ao = ½ C + ½ D
xB = 0 = ½ C – ½ D
Therefore,
C = Ao D = Ao

88
Hence with these particular starting conditions we have, by substitution back into eqs. (6.6), the
following results:
xA = ½ Ao (cos ωot + cos ω1t)
xB = ½ Ao (cos ωot – cos ω1t)

which can be rewritten as follows:


    0      0 
x A  A0 cos t  cos t
 2   2 
(6.7)
    0      0 
xB  A0 sin  t  sin  t
 2   2 

Each of these is a sinusoidal oscillation of angular frequency (ω1 + ωo)/2. The amplitude
associated with each pendulum is zero at the instant when the amplitude associated with the
other is maximum-although the actual displacement of the latter a t any such instant
depends on the instantaneous value of (ω1 + ωo)t/2.

6.6 Normal frequency: general analytical approach

We make use of the characteristics we discussed in connection with Eqs. (6.6). Both x A and xB
can oscillate with one of the normal frequencies. Let us take therefore,

xA = C cos ωt (6.8)
1
xB = C cos ωt
and see if there are values of ω and C and C1 for which these expressions are solutions of
equations (6.4)
d 2 xA
dt 2
 2 2

  0   c x A   c xB  0
2

(6.9)

 
2
d xB
  0   c xB   c x A  0
2 2 2
2
dt
Substituting equations (6.8) into equations (6.4), we get
(-ω 2 + ω o2 + ω c2 ) C - ω c 2 C1 = 0
ω c2C + (-ω 2 + ω o2 + ω c2) C1 = 0
For an arbitrary value of ω1,these constitute two simultaneous equation for the unknown
amplitudes C and C1. If they are independent equations, there is only one solution : C = 0, C 1 = 0
– which simply means that, for an arbitrary value of ω1 equations (6.8) are not a solution to the
problem.
89
But if these two equations are not independent i.e. If the second is a multiple of the first, then we
have in effect only one equation for the two amplitudes C and C 1. In this case, C can have any
value. But once C is chosen, then C1 is fixed.
For what value of ω are the two equations not independent and thus able to yield non zero
solutions for C and C1? From the first equations, we have


2
C
 2 C2 (6.10a)
C     0  C 2
and from the second,

2 2
C   2  0  C
 (6.10b)
C C
2

If C and C1 are not both zero the right–hand sides of these equations must be equal. Thus

C   2  0  C
2 2 2

2 =
  2  0  C C
2 2

or
-ω 2 + ω02 + ω c2 = ± ω c2
ω 2 = ω o2 + ω c2 ± ω c2
We have two solutions for ω; let us call them   and   :
  2 = ω o2 + 2 ω c2
  2 = ω o2
The positive square roots of these expressions are the two normal frequencies of the system;
once again we have arrived at the now familiar results. The relation between C and C 1 is now
gotten for each of the normal modes, from equations (6.9). For ω =   .
C
= -1
C1
These are two specific forms of equations (6.8) which are solutions to the coupled differential
equations of motion {equation 7.4}:
xA = C cos ωot xA = D cos ω1t
and (6.11)
1
xB = C cos ωot xB = -D cos ω t
The differential equation are linear (only the first powers of xA, xB
d2xA/dt2 and d2xB/dt2 appear), and therefore the sum of the two solutions is also a solution:
xA = C cos ωot + D cos ω1t( Special case) (6.12)

xB = C cos ωot – D cos ω1t

90
This is the solution previously given by equations (5 – 6)

NOTE: There is a factor of 2 lacking throughout in equations 6.10 as compared with


equations (6.6), but this makes no difference at all when one fixes the values of the
coefficients via the initial values of xA and xB.
Now let us find the general solution to the equations of free oscillation of this coupled system.
Specifically, insisted of equations, (6.11) we may in general have the following:
Lower mode: xA = C cos (ωot + α)
xB = C cos (ωot + α) (6.13)
1
Higher mode: xA = D cos (ω t + β)
xB = -Dcos (ω1t + β)
The existence of four adjustable constants then allows us to fit these solutions to arbitrary value
of the initial displacement and velocities of both pendulums. This removes the restriction to zero
initial velocity that required us to label our earlier solutions as special cases.

6.7 Forced vibration and resonance for two coupled oscillator


Let us now suppose, then, that a harmonic driving force F 0cos ωt is applied to pendulum A, the
motion of pendulum B being controlled only by its own restoring force and the coupling spring.
The statement of Newton’s law for Pendulum B is thus just the same as we had in considering
the free vibrations, and the equation or A is modified only to the extend of adding the term F o cos
ωt. Our two equations of motion thus become the following:

d 2 xA
m 2  m 0 x A  k  x A  x B   F0 cos t
2

dt

d 2 xB  g
m 2  m 0 x B  k  x A  x B   0  0  
2 2

dt  l
while dividing through by m, become
d 2 xA
dt 2
 2

2 2 F
  0   c x A   c x B  0 cos t
m
d 2 xB
dt 2
 2 2

  0   c xB   c x A  0
2
c2 
k
m
We let q1 (= xA + xB) and q2 (= xA – xB), and adding the differential equation above, we get

d 2 q1
  0 q1 = Fo Cos ω t
2
2
(6.14a)
dt
Subtracting them, we get
d 2 q2
   q 2 = Fo Cos ωt
2
2
(6.14b)
dt

91
Where

 2 = ω o
2
+ 2ω c2

The equations are as though we had two harmonic oscillator, of natural frequencies ωo and   .
We can clearly describe the steady state solution by the equations.
F0 m
q1 = C cos ωt where C = 2 (6.15)
0   2

F0 m
q2 = D cos ω t where D =
 2   2
We now extract the frequency dependence of the individual amplitude A and B of the two
pendulums, for we have
xA = A cos ωt where A = ½ (C + D)
xB = B cos ωt where B = ½ (C - D)

These give us the following results:

A( ) 

F0 0 2  c 2   2 
 
m 0 2  c 2   2   2  (6.16)
F0 c 2
B( ) 
 
m 0 2  c 2   2   2 

The variation of these quantities with us is shown in figure 6.4. In the region of frequencies
dominated by the lower resonance, the displacements of A and B are always of the same sign i.e.,
in phase with one another. In the region of frequencies dominated by the higher resonance, the
displacements are of opposite sign and hence 1800 out of phase. The introduction of nonzero
damping would as with the single driven oscillator, lead to a smooth variation of phase with
frequency as one goes through the resonances.
Note from Fig.6.4 that at certain frequency ω1 between the resonances, we have A = 0 and B =
nonzero. Yet from the assumed conditions of the problem it is clear that the periodic forcing of
pendulum B depends on the motion of pendulum A. The frequency ω1 at which the anomalous
situation develops is precisely the natural frequency of a single pendulum, with coupling spring

92
attached, under the circumstance that the other pendulum is held quite fixed ω1 = (ω o2 + ω c2)½.
In the complete absence of damping forces an arbitrarily small driving force of frequency ω1,
caused by arbitrarily small vibrations of pendulum A, would cause an arbitrary large response in
pendulum B. The existence of damping forces, however small, would destroy this condition, and
would mean that the amplitude Aω), although becoming very small near ω1, would never fall
quite to zero.
The main point to be learned from this analysis is the confirmation that one can trace out the
normal modes of a coupled system by means of resonance observations, and that the steady state
motions of the component parts of resonance are just like what they would be for the same
system in free vibration at the same frequency.

Fig.6.4: Forced response of two coupled pendulums with negligible damping. The normal modes
have the frequencies ω o and ω. (a)Amplitude of first pendulum as a function of driving
frequency [ω1= (ω o2+ω 12) ½]. (b) Amplitude of second pendulum as a function of
driving frequency.

6.8 N coupled oscillators


Because transverse oscillations are easier to visualize and to display than longitudinal
oscillations, we shall analyze the transverse oscillations of a prototype system of many particles.

93
Consider a flexible elastic string to which are attached N identical particles, each of mass m,
equally spaced a distance l apart. Let us hold a string fixed at two points, one at a distance l to
the left of the first particle and the other at a distance l to the right of the Nth particle (Fig. 6.5)

Fig.6.5: N equi-distant particles along a mass less string.

The particles are labeled 1 to N, or from 0 to N + 1, if we include the two fixed ends and treat
them as if they were particles with zero displacement. If the initial tension in the string is T and if
we confine ourselves to small transverse displacement particles, then we can ignore any increase
in the tension of the string as the particles oscillate.
Suppose for example, that particle 1 is displaced to y 1 and particle 2 to y2 (Fig.6.6); then the
length of string between them becomes l1 = l/ Cosα 1. For α1 <<1 rad, then cos α1
≈ 1 – α12/2 and l1 ≈ l(1 + α12/2). The increase in length is lα12/2, and any increased tension
that is proportional to this may be ignored in the comparison to any term proportional to the first
power of α1.

Fig.6.6: Force diagram for transversely displaced masses on a long string.

From the configuration shown the resultant x component of force on particle 2 is T cosα 1 + T
cosα2 = ½ T(α12 – α22), a difference between two second power terms in α. For small values of α 1
and α2, it is exceedingly small and we shall pay it no attention in what follows.
Figure 6.3 shows a configuration of the particles at some instant of time during their transverse
motion. We shall restrict ourselves to y displacements that are small compared to l. The resultant
y component of force on a typical particle, say the pth particle, is
Fp = - T sinαp – 1 + T sinαp
The approximate value of the sines are
y p  y p 1
Sinαp – 1 =
l
y p 1  y p
Sinαp =
l

94
Therefore,
Fp = -
T
 y p  y p1   T  y p1  y p 
l l
and this must equal the mass m times the transverse acceleration of the pth particle. Thus
d 2 yp
 2 0 y p   0  y p 1  y p 1   0
2 2
2
(6.17)
dt
where we have put
T
= ωo2
ml
We can write a similar equation for each of the N particles. Thus we have a set of N differential
equations, one for each value of p from 1 to N. remember that yo = 0 and
yN+1 = 0.
You may find it helpful to consider the single special case of equation 6.17 for N = 1 and
N = 2. If N = 1 we have
d 2 y1
 2 0 y1 = 0
2
2
dt
There is transverse harmonic motion of angular frequency w o√2 = (2T/ml)½, as one can
conclude directly form a consideration of figure 6.7 (a). N= 2, we have
d 2 y1
 2 0 y1 -  0 y 2  = 0
2 2
2
dt
d 2 y2
 2 0 y 2 -  0 y 1  = 0
2 2
2
dt
These are similar to Equations 6.4 for the two coupled pendulums, but we now have the
simplification that ωo and ωc are equal, so that ωo2 + ωc2 in equations 6.4 corresponds to 2ωo2 here,
and ωc2 there becomes wo2 here. The angular frequencies of the normal modes in this case are in a
definite numerical relationship; their actual values are ω o and ω o√3. The modes of N = 2 are
illustrated in figure 6.4 (b) and (c).

Fig 6.7: Normal modes of the two simplest loaded string systems (a) N = 1, one mode only,
(b)N = 2, lower mode and (c) N = 2, higher mode

6.9 Finding the normal modes for n coupled oscillators


We look for sinusoidal solution such that each particle oscillates with same frequency. We set.

95
yp = Ap cos ωt (p = 1,2, ----,N) (6.18)
th
where Ap and ω are the amplitude and frequency of vibration of the p particle. The velocity of
any particle can be obtained from equations (6.18) and is
dy p
 - ωAp sin ωt (p = 1, 2, ------, N)
dt
Using equations (6.18) as our trial solution we are restricting ourselves to the additional
boundary condition that each particle has zero velocity at t = 0; i.e., each particle starts from rest.
Substituting equations (6.17) into differential equations (6.17), we get
(-ω 2 + 2ωo2)A1 – ω o2(A2 + Ao) = 0
(-ω 2 + 2ω o2)A2 – ωo2(A3 + A1) = 0
(-ω 2 + 2ω o2)Ap – ωo2(Ap+1+ Ap-1) = 0
(-ω 2 + 2ω o2) AN – ωo2(AN+1 – AN-1) = 0
This formidable-looking set of N simultaneous equations can be written more compactly as
follows:
(-ω 2 + 2ω o2)Ap – ωo2(Ap-1+ Ap+1) = 0 (p = 1, 2, ------, N) (6.19)
Our earlier boundary condition requiring the ends to be held fixed means that Ao = 0 and AN+1 = 0
Rewriting equation (6.18)
A p 1  A p 1   2  2 0
2

 2
(p = 1, 2, --------, N) (6.20)
Ap 0
We see that, for any particular value of ω1 the right hand side is constant, and therefore the ratio
on the left most be a constant and independent of the value of p. What values can be assigned to
the Ap’s such that this condition will be satisfied and at the same time give
Ao = 0 and AN+1 = 0?
Suppose that the amplitude of particle p is expressible in the form
Ap = C sin pθ (6.21)
where θ is some angle. If a similar equation is used to define the amplitudes of the adjacent
particles p–1 and p+1, we shall have
Ap-1 + Ap+1 = C [sin (p – 1) θ + sin (p + 1)θ ]
= 2C sin pθ cos θ

But C sin pθ is just Ap1 so that we have


A p 1  A p 1
 2 cos  (6.22)
Ap
This means that the recipe represented by Eq.(6.21) is successful. The right-hand side of Eq.
(6.21) is a constant, independent of p, which is just what we need so as to have a condition
equivalent to Eq.(6.20). It can be used to satisfy all N of the equations (6.19) from which we
started. All that remain is to find the value of θ.
Imposing the requirement that
Ap = 0 for p = 0 and p = N + 1.

96
The former condition is automatically satisfied; the latter will hold well if (N + 1)θ is set equal to
any integral multiple of π. Thus we put
(N + 1) θ = n π (n = 1, 2, 3, ----------)
n
θ= (6.23)
N 1
Substituting for θ in Eq.(6.21) we thus get
n
Ap = C sin (6.24)
N 1
The permitted frequencies of the normal modes are also determined, for from Eqs.(6.20) through
(6.23) we have
A p 1  A p 1   2  2 0
2
n
 = 2cos( )
Ap 0
2 N 1

Therefore,
  n  
ω2 = 2ω o2  1  cos  
  N  1 
 n 
= 4ωo2 sin2  
 2 N  1 

Taking the square root of this, we have


 n 
ω = 2ω o sin   (6.25)
 2 N  1 
1. A string of length 2.4m and mass 0.2kg is placed under tension of 20N.
a) requires.
just as Eq.(6.23) What is the frequency of its fundamental mode?
b) If the string is plucked transversely and is then touched at a point 1.0m from
6. 10 ACTIVITY one end, what frequency persists?
2. The figure below shows two masses m connected by springs having equal spring
Two harmonic oscillators
constant A and
k, so that B, of mass
the masses are free m and spring
to slide on constants k A and
a frictionless tablekBAB.
, respectively, are
The walls at
coupled together
A and by B atospring
whichof spring
the ends constant
of the kC.springs
Find theare
normal frequencies
attached  and
are fixed.  the
Set up and
2
describe the normal modes
differential of oscillation
equations if kC of
of the motion = kthe
AkBmasses.

6.11 FURTHER READING

1) Rousseau, O., & Blaise, P. (2011). Quantum oscillators . Hoboken, N.J.: Wiley.
2) Scudiere, A. J. (2008). Resonance . S.l: Griffyn Ink.

6.12 SELF-TEST QUESTIONS

3. Find (a) the normal frequencies and (b) the normal modes of vibration for the system
in Problem 3.
4. Suppose that in Problem (3) the first97mass is held at its equilibrium position while
the second mass is given a displacement of magnitude a>0 to the right of its
equilibrium position. The masses are then released. Find the position of each mass at
any later time.
LECTURE SEVEN

THE FREE VIBRATIONS OF STRETCHED STRINGS

7.1 Introduction

In this topic we discuss how a string, with both ends fixed, has a number of well defined states of
natural vibrations called stationary vibrations. These vibrations execute SHM. Also discussed
here is differential equations and properties of a transverse progressive wave like tension and
speed wave number.

7.2 Lecture Objectives

By the end of this lecture, you should be able to:


 Understand the concept of stationery and progressive waves.
98
 Derive the differential equations of a progressive wave.
 Derive the tension, speed wave number in a transverse progressive wave.
Our interest here is the basic mechanical fact that a string, with both ends fixed, has a number of
well defined states of natural vibrations. These are called stationary vibrations, in the sense that
each point on the string vibrates transversely in SHM with constant amplitude, the frequency of
this vibration being the same for all parts of the string.
Such stationary vibrations represent the so called normal modes of the string. In all except the
lowest mode, there exist points at which the displacements remain zero at all times. These are
nodes; the positions of maximum amplitude are called antinodes. One can thus think of these
basic states of vibration as being stationary in the additional sense that the nodes remain at fixed
points along the string.
Suppose there is a string of length L, with its ends held fixed at the points x = 0, x = L. We
suppose further that the string has a uniform linear density (mass per unit length) equal to μ and
that it is stretched with a tension T. At some instant let us configure a portion of the sting to be as
shown in figure7.1.

Fig.7.1: Force diagram for short segment of massive string in transverse vibration.

Thus for a short segment of the string, of length Δx, the net force acting on it is given by
Fy = T sin(θ + Δθ) – T sinθ

99
Fx = T cos(θ + Δθ) – T cosθ
Where θ, θ + Δθ are the directions of tangents to the string at the ends of the segment, i.e. at x
and x + Δx.
We are assuming that the transverse displacement y is small, so that θ and θ + Δθ are small
angles. In this case we have
Fy ≈ T Δθ
Fx ≈ 0

The equation governing the transverse motion of the segment is thus (very nearly)

TΔθ = (μΔx.) ay (7.1)


Now θ embodies the variation of y with x at a given value of the time t, and a y embodies the
variation of y with t at a given value of x. Therefore, in rewriting Eq.(7.1) in terms of x 1y1 and t
we must use partial derivatives, and we have the following relationship:
y
tan θ = δy
x
 2y
Sec2θΔθ = Δx
x 2
But secθ ≈ 1, and so

 2y
Δθ ≈ Δx
x 2
Also
 2y
ay =
t 2
Thus Eq.(7.1) becomes
 2y  2y
T Δx = μΔx
x 2 t 2
Therefore,
 2y   2y
= (7.2)
x 2 T t 2
T/μ has the dimension of the square of a speed, i.e. the speed with which progressive waves
travel on a long string having these values of μ and T. We define the speed v through the
equation.
1
 T  2
v=   (7.3)
 
and will then rewrite Eq.(7.2) in the following more compact form
 2y 1  2y
= (7.4)
x 2 v 2 t 2
or

100
(∂2y/∂t2) = c2(∂2y/∂x2)
(this is because in some books the speed v is usually replaced by c)

Every point on the string is moving with a time dependence of the form cosωt, but the amplitude
of this motion is a function of the distance x of that point from the end of the string. (Our
assumed time dependence would require every point of on the string to be instantaneously
stationary at t = 0. If this is not so an initial phase angle must be introduced).
Thus we assume
y(x,t) = f(x) cos ωt (7.5)
This gives us
 2y
= -ω 2f(x) cos ωt
t 2

 2y d2 f
= cos ωt
x 2 dx 2
(Note that, since f is by definition a function of x only, we can write d 2f/dx2, instead of a partial
derivative). Substituting there derivatives in Eq. (7.4) then gives us
d2 f 2
  f
dx 2 v2
But this is the familiar differential equation satisfied by a sine or cosine function. Remembering
that we have defined x = 0 as corresponding to one of the fixed ends of the string, with zero
transverse displacement at all times, we know that an acceptable solution must be of the form.

 x 
f(x) = A sin   (7.6)
 v 

But we have the further boundary condition that the displacement is always zero at x = L.
Hence we must also have

 L 
A sin   =0
 v 
Therefore,
L
= nπ (7.7)
v
where n is any (positive) integer.
It will be convenient to introduce the number of cycles per unit time v, equal to ω/2π. The
frequencies of the permitted stationary vibrations are thus given by
1
nv n  T 2
vn =    (7.8)
2L 2 L   
where n, according to this calculation, may be 1, 2, 3, ------- to infinity.
A vivid way of describing the shape of the string at any instant, in any particular mode n, is
obtained by recognizing that the total length of the string must exactly accommodate an integral

101
number of half-sine curves, as implied by Eq.(6.7). We can therefore define a wavelength, λn
associated with the mode n, such that
2L
λn = (7.9)
n
Then we can put
 n 2
 
v L n
Hence, from Eq.(7.6), the shape of the string in mode n is characterized by the following
equation.
 2x  nx
fn (x) = Ansin    An sin  
 (7.10)
  n   L 
and the complete description of the motion of the string is thus as follows:
2x
yn(x,t) = An sin cos ωnt (7.11)
n
Where
1

ω n = n  T 2
  = nω1
L  
Since all the possible frequencies of a given stretched string are, according to the above analysis,
simple integral multiples of the lowest possible frequency, ω1 a particular interest attaches to this
basic mode – the fundamental. It is the frequency of the fundamental that defines what we
recognize as the characteristic pitch of a vibrating string, and which therefore defines for us the
tension required to obtain a certain note from a string of given mass and length.

Example: The E string of a violin is to be turned to a frequency of 640 Hz. Its length and mass
(from the bridge to the end) are 33cm and 0.125g, respectively what tension is required?

Solution.
From Eq.(6-8) we have
1
1  T  2
v1 =  
2 L   
and we shall put μ = m/L where m is the total mass. This then gives us
T = 4 mLv12
= 4(1.25 x 10 –4) (0.33) (6.4 x 102)2
= 68N

7.3 Forced harmonic vibration of stretched string


We imagine that the end of the string at x = L remains firmly fixed, but that the end at x = 0 is
vibrated transversely at some arbitrary angular frequency and with an amplitude B.
We suppose a steady-state solution of the form
y(x,t) = f(x)cosωt

102
but now subject to the following conditions.
y(θ,t) = B cos ωt
y(L,t) = 0
The basic equation of motion is still Eq.(7.4) i.e.
 2y 1  2y

x 2 v 2 t 2
So that f(x) must be sinusoidal function of x. We therefore put
f(x) = A sin = (kx + α)
From Eq.(7.4) we then get k = ω /v, so that
 x 
f(x) = A sin   
 v 
This is just like Eq.(6-6) except for having an adjustable parameter α. From the boundary
condition at X = L, we then have
 L 
sin    = 0
 v 
Therefore,
L
+ α = pπ
v
where p is an integer. From the boundary condition at x = 0, we get

B = A sinα
Therefore,
B
A
 L  (7.12)
sin  p  
 v 
The implication of this result is that for a given amplitude of the forced displacement at the
extreme end, the response of the string as a whole will be very large whenever the driving
frequency is close to one of the natural frequencies defined by Eq.(7.8)

7.4 ACTIVITIES

a) Derive the equation of Problem 5 if the string is horizontal and gravity is taken into
account.

b) Assume that a continuous string, which is fixed at its endpoints and vibrates transversely,
is replaced by N particles of mass m at equal distance from each other. Determine the
equation of the particle.

7.5 FURTHER READING

1) Rousseau, O., & Blaise, P. (2011). Quantum oscillators . Hoboken, N.J.: Wiley.

103
2) Scudiere, A. J. (2008). Resonance . S.l: Griffyn Ink.
3) Rajput, R. (2002). A textbook of engineering mechanics (S.I. units) (2nd rev. ed.).
Daryaganj, New Delhi: Dhanpat Rai Publications.
4) Frohrib, D. A., & Plunkett, R. (1966). The free vibrations of stiffened drill strings with
static curvature . New York, N.Y.: ASME.
5) Magie, W. F. (1935). A source book in physics, . New York: McGraw-Hill Book
Company.

7.6 SELF-TEST QUESTIONS

1. The equation of a transverse wave traveling along a string is given by


y = 0.3 sin π(0.5x-50t), where y and x are in centimeters and t is in seconds.
a) Find the amplitude, wavelength, wave number, frequency, period, and velocity of
the wave.
b) Find the maximum transverse speed of any particle in the string.

2. A wave of frequency 30 sec-1 has a velocity of 90m/sec.


a). How far apart are two points whose displacements are 30o apart in phase?
b) At a given point, what is the phase difference between two displacements occurring
at times separated by 0.02 sec?

3. A string of length L, which is clamped at both ends and has a tension T, is pulled aside a
distance h at its centre and released. What is the energy of the subsequent oscillations?

4. A uniform inextensible string of length l and total mass m is suspended vertically and
tapped at the top end so that a transverse impulse runs down it. At the same moment a body
is released from rest and fall freely from the top of the string. How far from the bottom does
the body pass the impulse?
104
5. Derive the partial differential equation for the transverse vibrations of a vibrating string.
LECTURE EIGHT

DIFFRACTION OF LIGHT AND WATER WAVES

8.1 Introduction

Here we discuss the diffraction effects in light waves. Due to this effect we also find out the
angular position of diffraction at a given range of wavelength. Effects and properties of group
waves are also discussed in this topic.

8.2 Lecture Objectives

By the end of this lecture, you should be able to:


 Appreciate diffraction and its effects in light waves,
 Calculate angular position of diffraction minima at a given range of wavelength,
 Differentiate between group and phase velocities.

105
8.3 Diffraction
Parts of a wave from a single source may interfere with each other causing effects such as
bending of waves around an obstacle. Single-source interference is referred to as diffraction and
occurs for all kinds of waves.
Light sources such as the sun are not monochromatic point source, and the diffraction patterns
due to different parts of the source and to different wavelengths usually overlap and obscure each
other. Nevertheless, diffraction patterns are visible when we look at a distant light source, such as
a street lamp, through a crack between two fingers or through a cloth umbrella.
Diffraction effects are large only when we deal with obstacles or apertures comparable in size to
the wavelength.

8.4 Diffraction of light by a Narrow slit.

Fig. 8.1: Geometry of the single-slit diffraction experiment

Fig. 8.1 shows monochromatic light passing through a slight of width in a comparison to the
wavelength  and falls on a screen at a lager distance. Huygen’s wavelets from different parts of
the slit interfere and produce the diffraction pattern shown in Fig. 8.2.

106
Fig 8.2: Diffraction pattern

To locate the minima in this diffraction pattern, we start by dividing the slight into two halves
(AB and BC in Fig. 8.3). The distance to the screen for the wavelet originating at A is greater
than that for the wavelet from B by x = ½ a sin . Wavelets originating at any point on AB and at
the corresponding point on BC will have the same path difference.

Fig.8.3: A diffraction minimum occurs when wavelets from different parts of the slit interfere
destructively.

If x is exactly ½ , all the parts of wavelets will be out of phase and cancel exactly, producing a
minimum. Hence a minimum occurs when x = ½ a sin  = ½  or

107
a sin  = 
At this point, one might suppose that, a maximum occurs when x =  since the wavelet from A to
B will then be exactly in phase. However, this is not the case, as can be seen by dividing the slit
into four equal parts (Fig.8.3). wavelets from A and D will have a path difference that is ½ x = ½
, so they will be exactly out of phase and cancel completely. The same is also true for wavelets
from any two points a/4 apart, so x =  also implies a minimum, Similarly, we can divide the
slight into 6, 8,------parts, and find that the total destructive interference occurs when x = ½ a sin
 = ½ , , 3/2 , 2, ……… Thus in general, diffraction minima occurs at angles satisfying
a sin  = m, m = 1, 2,……….
(diffraction minima)
The negative integers correspond to minima above the axis in figure 8.2. The maxima are located
approximately ( but not exactly ) midway between the maxima, or at m = 0, 3/2 , 5/2 ,
…….. The central m = 0 maximum s very bright, since all the wavelets have nearly the same
path length and are in phase. At the other maxima, the intensity is much smaller, and it
diminishes rapidly as m increases (Fig. 8.2). This happens because, except at  = 0, the wavelets
from some parts of the slit cancel even at the maxima. This partial cancellation becomes more
and more complete as the angle increase. The angular width of the diffraction pattern depends
on the size of the slit relative to the wavelength of light, as seen in the next example.

Example:
A slit is illuminated by light of wavelength. Find the angular position of the first diffraction
minima as the slit width expands from  to 5 and finally to 10.

Solution
Substituting m = 1 and a =
in a sin  = m, we find
sin  = m/a = (1) / = 1
so  = 90o. Similarly, a = 5 gives sin  = 0.2 and  = 12o and  = 12o; a = 10  gives sin  =
0.1 and  = 6o. Thus as the slit becomes wider, the central diffraction peak becomes narrower. A
similar narrowing of the diffraction pattern would be observed if the slit size were held fixed and
the wavelength decreased.

8.5 GROUP VELOCITY


In the above discussions it has been assumed that simple harmonic waves are propagated with a
speed c which is independent of the frequency-there is no dispersion. When this is so every
component of a wave group travels at the same speed. This means that the whole group also

108
travels at the speed and is completely unchanged during its passage through the medium. It can
be shown that if there is dispersion the wave group constituting the disturbance of finite length
travels at a speed which differs from that of any of the component simple harmonic waves. This
phenomenon can be observed with the wave spreading across the surface when a stone is
dropped into still water. A group of waves is observed to spread out across the water surface but,
in the group, waves can be seen to be moving faster than the group itself. These waves appear at
the rear of the group, travel through it, and die out at the front. It is found that for any type wave
motion the group velocity is less than the velocity of individual waves (the wave velocity) if the
dispersion is such that the wave velocity increases with wavelength.
Suppose one has two superposed simple harmonic waves, with equal amplitudes and slightly
different frequencies and velocities, propagated in the positive x-direction. The resultant is given
by:
Ф = a cos (2π/λ)(ct - x) + a cos (2π/λ+δλ) [(ct – δc)t - x)]

= 2a{cos π [ (c/λ + (c+ δc)/ (λ+δλ))t – (1/ λ + 1/(λ+δλ))x]


X cos π [ (c/λ - (c+ δc)/(λ+δλ))t – (1/ λ - 1/(λ+δλ))x]}

If δλ ≤ λ one can write λ2 in place of λ/(λ+δλ) so that one has


1/ λ - 1/λ+δλ = δλ/λ and c/λ – (c+δc/λ+δλ) = (cδλ- λδc)/ λ2
2

Ф = 2acos (2π/λ)(ct - x)cos πδλ/λ2 [([cδλ - λδc)/δλ]t - x)] (8.1)


At a given instant the variation of Ф with x is of the form shown in figure 8.4. The wave shown
as a full line corresponds to the first term in equation 8.1 and is modulated by the slowly varying
envelope which corresponds to the second term.

Fig. 8.4: Wave group.

The first term defines a wave of the same speed and wavelength as the original wave but the
envelope has wavelength 2λ2/δλ and speed (cδλ - λδc)/δλ. Thus the wave is divided up into
groups which travel with the group velocity u given by;

u = c – λ(dc/dλ) (8.2)

As the disturbance is propagated, the wave shown as a continuous line in Fig. (8.4) moves along
with velocity c which the envelope moves with velocity u. This example involves the
superposition of only two waves. It has already been seen that if one group and equation 8.2

109
shows that this has group velocity only if the dispersion term λ(dc/dλ) is constant. If the wave
group is mainly confined to a narrow range of frequencies, the dispersion term is usually almost
constant so that there is a unique group velocity. Equation 8.2 shows that the group velocity is
smaller than the wave velocity if the latter increases with wavelength, is greater than the wave
velocity if dc/dλ is negative, and is equal to the wave velocity if there is no dispersion.
It is sometimes convenient to have alternative expressions for the group velocity. For example, if
ω is the pulsatance and f is the frequency, one has
ω = 2πf = 2πc/λ = ck
therefore
dω/dk = c + k(dc/dk) = c - λ(dc/dλ)
u = dω/dk (8.3)
Now dω/dk = (dω/df)(df/dλ)(dλ/dk) where
dω/dk = 2π and dk = dλ = -2π/λ2
Therefore;
u = - λ2 (df/dλ) (8.4)

if co is the velocity of light in a vacuum the refractive index of the medium is defined by n = c o/c.
from eq. (3) one has
1/u = dk/dω = d/dω(ω/c) = 1/c – ( ω/c2)(dc/dω),
whence
1/u = 1/c + ( ω/co)(dn/dω), (8.5)
In practice the dispersion of a material is often in the form of a relation between the refractive
index and the vacuum wavelength λo. (it should be noted that in equation 8.2. λ is the wavelength
in the medium and it is itself determined by the refractive index). A manipulation of equation 8.2
yields a formula that is convenient in this situation.
One has
dn/dω = (dn/dλo)(dλo/dω)
Now ω = koco, or λo = 2πco/ω
Therefore
(dλo/dω) = -2πco/ω2 = - λo/ω.
Equation 8.5 then gives

1/u = 1/c - (λo/co)(dn/dλo), (8.6)

8.6 ACTIVITIES
a) Discuss effects of diffraction of light
b) Explain what happens when two waves of same frequency and are out of phase are
superimposed?

8.5 FURTHER READING

110
1) Kunik, M., & Skrzypacz, P. (2007). Diffraction of light revisited . Magdeburg: Univ., Fak.
für Mathematik.
2) Born, M., & Wolf, E. (1999). Principles of optics: electromagnetic theory of propagation,
interference and diffraction of light (7th expanded ed.). Cambridge: Cambridge
University Press
3) Meyer, C. F. (1934). The diffraction of light, . Chicago, Ill.: The University of Chicago
press

8.6 SELF-TEST QUESTIONS

1. A slit is illuminated by light of wavelength λ. Find the angular position of the first
diffraction minima as the slit width expands from λ to 5λ and finally to 10λ.
2. Explain carefully what is meant by group velocity, giving examples from various
types of wave motion. Derive an expression for the relation between group velocity
and phase velocity. The refractive index of a certain glass is given by the formula μ
= A + (B/λo2) where λo is the wavelength in cacuo. Show that the ratio of group
velocity to phase velocity (A/λo2 + B)(A/λo2 + 3B)
LECTURE NINE

REFLECTIONS OF MECHANICAL (SHEAR) WAVES AT A DISCONTINUITY

9.1 Introduction

In this topic we discuss aspects of transverse and longitudinal waves which will help in
understanding the discontinuity of waves. Also we discuss in details on the reflection and
transmission of mechanical waves.

9.2 Lecture Objectives

By the end of this lecture, you should be able to:


 Differentiate transverse and longitudinal waves
 Understand the discontinuity of waves.111
 Describe reflection and transmission of mechanical waves.
To understand the discontinuity of waves we first look at the differences between
longitudinal and transverse waves.

9.3 Transverse and longitudinal traveling waves

In the wave propagation, particles of the medium do not move along with the wave. It is the
disturbance which moves forward. The particles simple execute vibrations about their mean
positions. All waves are not necessarily periodic. We consider the simple case of periodic waves
in which the particles of the medium execute simple harmonic oscillations about their mean
positions as the waves travels. Such a wave is called a harmonic wave. There are two types of
waves (a) Transverse (b) Longitudinal.
A wave in which the particles of the medium vibrate in a plane perpendicular to the direction of
propagation of the wave of the wave is called a transverse wave. But a wave in which the
particles of the medium execute simple harmonic vibrations along the direction of propagation is
called a longitudinal wave.
There are alternate positions of maximum density (compression) and minimum density
(rarefaction). Thus in longitudinal waves, alternate compressions (high pressure) and rarefactions
(low pressure) are transmitted with the waves. Sound waves traveling in the air (or in any, gas)
are longitudinal waves. Though we do not see the air particles we can observe the motion at the
source and also the motion at the receiver. Sound is produced by the mechanical vibrations such
as the vibrations of the string of the violin, vibrations of prongs of tuning fork, vibrations of a
table membrane etc.
The stretched sting is essentially a carrier of transverse wave. A long spring on the other hand, is
capable of carrying both transverse and longitudinal disturbances. In this respect a spring is a
better analogue of a real solid, which can also carry both transverse (shear) waves and
longitudinal (compressional) waves. A column of liquid or gas, in contrast to a solid, has no
elastic resistance to change of shape, only to change of density.
Thus a column of a fluid (e.g. air) carrier only longitudinal waves, except and it is a very
important exception- when gravity or surface tension provides in effect an elastic restoring force
against transverse deformations.
Transverse waves have two different directions of polarization for the vibrations- perpendicular
to one another and to the direction of propagation. It may even be that there different
polarization states have different wave speeds associated with them.
It may be worth pointing out that, a given interface may behave differently with respect to
longitudinal and transverse waves on encountering a boundary or a barrier. Suppose, for

112
example, that water rests in a tank with smooth vertical walls. The interface between water and
wall then acts as an almost completely rigid boundary with respect to longitudinal waves, but as
a completely free end with respect to transverse waves. If standing waves were to be set up, the
wall would represent a node for longitudinal vibrations of the water but an antinode for
transverse vibrations.

9.4 Reflections of mechanical (shear) waves at a discontinuity

We shall begin the stretched string, and will consider what happens when a traveling wave on a
string encounter a discontinuity of some kind. One can set up a standing wave by agitating one
end of a string, thereby generating a traveling wave which undergoes some process of reflection
at the far end. The outgoing and returning waves then conspire to produce a standing–wave
pattern with nodes at fixed positions.
Quantitatively, a given normal mode on a string with fixed ends can be regarded as the
superposition of two sine waves of equal amplitude, wavelength and frequency, traveling in the
opposite direction. The following two statements are mathematically equivalent:
Normal node:

 nx 
y(x,t) = A sin   )cosωt
 L 
Two traveling waves:
A  nx  A  nx 
y ( x, t )  sin   t   sin   t 
2  L  2  L 
If we take the second of these statements and fix attention on the condition at x= 0 or x= L, we
have

A A
y(0,t) = y(L,t) = sin   t   sin t
2 2

A A
=- sin t  sin t
2 2
These means that these oppositely traveling waves must, at all times produce equal and opposite
displacement at the fixed ends. More so that this same condition must define the reflection
process for any traveling wave when it encounters a rigid boundary.
The point on the string at which the two pulses must remain at rest at all times! The waves pass
through in opposite directions without causing any displacement of the point at any time.
This gives us the clue as to what happens when a wave pulse is incident upon the end of a string
which is held stationary. A pulse of opposite displacement is reflected from the end and travels
back towards the source.

113
The arrival of a positive displacement will exert an upward force on the support which holds the
end fixed. By Newton’s third law the support exerts a reactions force in the opposite direction
back on the string, thus generating a pulse of opposite polarity which travels backward towards
the source.
If the end of the spring is completely free to move (e.g.; if it were tied to a massless ring on a
functionless vertical rod/) the arrival of a positive pulse will exert an upward reaction force back
on the string, generating a pulse of positive polarity. This positive pulse is then transmitted back
along the string. Reflection from a “free” end thus produces a pulse of the same polarity
traveling back towards the source.
If a string with a certain tension and mass per unit length is fastened to another string with a
different µ, in general some reflection will take place (as well as some transmission) at the
discontinuity. Consider a pulse of the form f 1(t – x/u1) moving a long a structural cord of linear
density µ1, which is joined to a cord of linear density µ2 at x = 0 (Fig 9.1).

Fig 9.1: Reflection and transmission at a junction (boundary) between dissimilar strings.
NOTE:
1. Another method of achieving a “free” termination is to tie it to ‘a very’ less massive
string.
2. Writing the pulse as f(t –x/v) instead of our usual f(x–vt) is more appropriate in this
analysis, basically because when a wave passes from one medium to another the
wavelength changes but the frequency does not. Thus we associate the changed factor
with the x co–ordinate and not with t.

Assuming partial reflection and partial transmission at the junction the transverse displacement
in the two strings can be assumed to be given by the following equations:
 x   x 
y1( x, t )  f 1 t    g  t  
 u1   u1 
(9.1)
 x 
y 2( x, t )  f 2 t  
 u 2 

114
Because the ends of the two cords remains in contact with each other, the transverse
displacements, y, at the point x = 0, must be the same for both cords. Also at each instant the
codes must join with equal slopes and have equal tensions; otherwise the element of mass
represented by the junction would be given a very large acceleration. Thus we have the following
two conditions:

y1(O,t) = y2(0,t)

y1
 0, t   y 2  0, t 
x x

ie.

f1(t) + g1(t) = f2(t) (9.2)

1 1 1
f 1 t   g 1  t   f 2 t  (9.3)
u1 u1 u2
Integrating Eq. (10.3), we have

u2f (t) – u2g1(t) = u1f2(t) (9.4)


Solving Equations. (9.2) and (9.4) for g1 and f2 in terms of f1, we find
u u
g 1 (t )  2 1 f1  t 
u 2  u1
(9.5)
2u 2
f 2 (t )  f1  t 
u2  u1

As they stand, equations (9.5) are merely a description of the state of affairs at x = 0 at any
arbitrary value of t. What equations (9.5) do is to relate the valve of f1, g1 and f2 at the same valve
of their argument.
For any given value, say  , of this argument, we have g1(  ) = const.x f1(  ) and f2 (  ) = const.
x f1(  ). But one is not restricted to interpreting the value of t at x =0. Thus the function f 1 is
defined to be the argument t – x/u1, and the function g1 is defined to be the function of the
argument t +x/u1. Suppose each of these arguments is set equal to the same value  , as required
by Eq.(9.5). Clearly we cannot use the same pair of values of x and t for both; let us therefore
label the values as xf, tf and xg, tg. Then we have
xf xg
 = tf   tg 
u1 u1
If we put tf = tg = t, then we must have
xg = -xf

115
and what this means is that the displacement associated with pulse g 1 at any given instant, at any
given value of x, is directly related to the valve of f 1 as calculated at the same time as the position
–x. Specifically, according to the first of equations (9.5) we have
  x u  u1  x
g1  t    2 f 1  t   (9.6a)
 u1  u 2  u1  u1 

And what this says is that the reflected pulse besides being scaled down by the factor (u 2 –
u1) / (u2 + u1), is reserved right to left with respect to the incident pulse. If u2< u1, it is also turned
upside down.
In a similar way we can relate the transmitted waveform f 2 to the incident waveform f1. At a
given value of t the corresponding values of x (call them x1 and x2) are defined by the relation.

x1 x
 = t t 2
u1 u2
Hence x2 = (u2/u1)x1, and the second of equations (9.5) requires us put
 u x u  2u 2  x
f 2  t  2 1   f1  t  
 u2  u 2  u1  u1 
(9.6b)

This tells us that, compared to the incident pulse, the transmitted pulse suffers not only a change
in height but also a scale change along x.
In using the above relationships, it is to be noted that if the pulse f 1 is incident from the negative
x direction and if the junction (boundary) is at x = 0, then the functions f 1 and g1 represent
physically real displacement only if x ≤ 0: Thus, for example, in using Eq.(9.6a), we find the real
displacement in the reflected pulse g1 at some negative value of x, by considering what the
displacement of the incident pulse f1 would have been if it had continued on into the region of
position x and then multiplying by the factor;
(u2 – u1) / (u2 +u1).
As extreme cases of Eq. (9.6a) we have the following:
a. String 2 infinitely massive:
u2 = 0
  x  x
g  t     f1  t  
 u1   u1 

b. String 2 massless or absent:


u2 = ∞
  x  x
g  t    f1  t  
 u1   u1 

9.5 ACTIVITIES

116
Explain the differences between transverse and longitudinal waves.

9.6 FURTHER READING

1) Cummings, K., & Halliday, D. (2004). Understanding physics . Hoboken, NJ: Wiley.
2) Beiser, A. (2003). Concepts of modern physics (6th ed.). Boston: McGraw-Hill.
3) Davis, J. L. (2000). Mathematics of wave propagation. Princeton, NJ: Princeton
University Press.
4) Davis, J. L. (2000). Mathematics of wave propagation . Princeton, NJ: Princeton
University Press.

9.7 SELF-TEST QUESTIONS

1. Two strings of tension T and mass densities μ1 and μ2 are connected together.
Consider a travelling wave incident on the boundary. Find the ratio of the
reflected amplitude to the incident amplitude and the ratio of the transmitted
amplitude to the incident amplitude, for the cases μ1/ μ1= 0, 0.25, 1.4, œ
2. Two strings of tension T and mass densities μ1 and μ2 are connected together.
Consider a travelling wave incident on the boundary. Show that the energy flux of
the reflected wave plus the energy flux of the transmitted wave equals the energy
flux of the incident wave.
[Hint: The energy flux of a wave (the energy density times the wave speed) is
proportional to A2/v where A is the amplitude and v is the wave speed.]

117

You might also like