You are on page 1of 33

Mechanical Vibrations

Husnain Inayat Hussain


List of Figures
2.1 A Simple Spring Mass System . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.2 Phase Angle and associated Triangle . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Kinematics of an Undamped Spring Mass System . . . . . . . . . . . . . . . . 9
2.4 A Simple Pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.5 A circular Disc . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.6 A Compound Pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.7 (a) Springs in Parallel. (b) Springs in Series. . . . . . . . . . . . . . . . . . . 19
2.8 A Network of Dierent Springs . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.9 A Torsional Pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.1 A Vertical Spring Mass System with a Damper . . . . . . . . . . . . . . . . . 24
3.2 Examples of Exponentially Decaying Envelopes for Dierent Values of . . . 28
3
List of Tables
2.1 Relationship between Magnitudes of Kinematics for Various Values of Angular Frequencies. 10
2.2 Natural Frequencies for Various Systems. . . . . . . . . . . . . . . . . . . . . 21
3.1 Eq. (3.9) Parameters and their Equivalence. . . . . . . . . . . . . . . . . . . . 26
4
Contents
1 Introduction 1
2 Simple Oscillatory Systems 2
2.1 Spring Mass System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
2.1.1 Observing Change in Systems State . . . . . . . . . . . . . . . . . . . 3
2.1.2 Equation of Motion for the Spring Mass System . . . . . . . . . . . . 4
2.1.3 Spring Mass System Solution . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Initial Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2.1 A Neat Trignometric Trick . . . . . . . . . . . . . . . . . . . . . . . . 6
2.3 Kinematics of an Undamped Spring Mass System . . . . . . . . . . . . . . . . 9
2.4 Simple Pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.4.1 Linearization of the System . . . . . . . . . . . . . . . . . . . . . . . . 12
2.4.2 Mass Moment of Inertia . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.5 Physical Pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.5.1 Center of Percussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5.2 Radius of Gyration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.6 Energy Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.7 Equivalent Spring Constant . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.8 Angular Frequency for General Cases . . . . . . . . . . . . . . . . . . . . . . . 20
2.9 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3 Damped Oscillations 22
3.1 Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.1.1 Pressure / Drag Damping . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.1.2 Viscous Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 A Spring Mass System with Damping . . . . . . . . . . . . . . . . . . . . . . 24
3.2.1 Equation of Motion from Unstretched Position . . . . . . . . . . . . . 25
3.2.2 Equation of Motion from the Stretched Position . . . . . . . . . . . . 25
3.3 Damped Oscillations Solution . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.3.1 Solution Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3.2 Roots of Dierential Equation . . . . . . . . . . . . . . . . . . . . . . . 28
3.4 The Great Battle: Damping vs. Oscillation . . . . . . . . . . . . . . . . . . . 28
5
Nations are born in the hearts
of poets, they prosper and die
in the hands of politicians.
Allama Mohammad Iqbal
(1877-1938)
1
Introduction
A periodic movement of an object around some position of rest (equilibrium) can be the sim-
plest explanation of what vibrations are. In our everyday life we see and observe vibration
on a routine basis. In fact if we for just the time being, substitute the word periodic motion
for vibration, then we will be able to realize that our entire lives, the universe we live in, the
objects we use and the actions we take exhibit a certain form of periodicity in one way or the
other.
Please consult the power point slides for this chapter.
1
With faith, discipline and self-
less devotion to duty, there
is nothing worthwhile that you
cannot achieve.
Mohammad Ali Jinnah
(1876-1948)
2
Simple Oscillatory Systems
In this chapter we will consider the most basic oscillatory systems and then go for trying to
understand them physically rst and then we will elaborate on dierent mathematical ways
of understanding the physical behaviour.
2.1 Spring Mass System
A spring is dened as a mechanical element capable of storing energy. A mass is dened as
a mechanical element capable of exhibiting inertia which is a measure of resistance of mass
when its state is being changed. By state we mean the state of rest or the state of motion.
A system which is constructed out of these two elements as shown in Fig. 2.1 can be termed
as a simple spring mass system. In the gure we have considered a spring with stiness k
and negligible mass, attached to a block of mass m and the entire assembly is placed on a
horizontal frictionless surface. x denotes the displacement of the spring from its rest position.
The blocks current position (rest or equilibrium position) is taken as x = 0. If we now pull
m
k
x
x = 0
Figure 2.1: A Simple Spring Mass System
the block and release it, it will start moving in opposite direction to the direction in which
pull was applied. This happens because the rigidity of the spring reacts against the initial pull
and develops a resistive force which always acts in opposite direction to the applied stimulus.
In the present case the applied stimulus is the action of pulling the spring and letting it go.
Let us assume that the spring was pulled to some arbitrary location x along the x-axis. Under
the conditions that the pull is not too hard and it is gentle enough that the spring is not
2
2.1. SPRING MASS SYSTEM 3
permanently deformed then Hookes Law states that:
f
s
= kx (2.1)
f
s
is the spring force whcih has a negative sign indicating the fact that spring force acts in
opposite direction to the direction of push or pull. Once the spring is released we see that
it immediately starts moving in the negative x direction towards its rest position. However,
once at the rest position, the spring continues to move and goes further back. It continues
to move back until it reaches the other extreme. At this instant it stops, but only for some
unnoticeable instant. Now the spring force causes it to move forward toward the equilibrium
position in the positive x direction again. As it once again reaches the equilibrium, the trou-
bled block of mass ms misery does not stop. Rather it continues to move in the forward
direction until it reaches the rst point where it was released.
The question at this stage is why does the block not stop at its zero position. The block does
not stop because the inertia of the block refrains it to stop arbitrarily. It should be clear by
now that behaviour of this system is governed by sprng force and inertia. Spring force
causes the system to move in opposite direction of the initial stimulus while inertia causes the
system to keep moving and not stop even if equilibrium postion is reached. If there are no
other forces such as the force of friction then the block will continue to move forever! This of
course is an idealization just to make the mathematics digestible before more realistic complex
cases. Real world systems come to halt eventually. However, regardless of this idealization,
simple spring mass system presents an interesting and fundamental case which is worth our
time and attention as the simple concepts developed in this study will help us explore more
complex cases with relative ease. Under the assumptions of only a spring force acting on the
system, if we want to know how the block moves with respect to time then we need Newtons
Second Law of motion. It states that sum of all the force be them internal or external is
equal to the product of the mass of the object and the acceleration with which it moves. This
may be expressed mathematically as:


f = ma (2.2)
Since we are dealing with one dimensional motion, we can forget about the vector notation.
The only force present in the system is the spring force. Equating Eq. (2.1) and Eq. (2.2),
one obtains the following:
kx = ma (2.3)
2.1.1 Observing Change in Systems State
We want to investigate how the system changes its state. A systems state may be dened
by many parameters. In particular if one wants to study the change in the kinematic state
of the system then there are three parameters of interest:
position
velocity
acceleration
4 2.1. SPRING MASS SYSTEM
In the spring mass system considered above, we want to study the same parameters. Let
us consider changes in this systems position with respect to time. We set two measurement
points x
a
and x
b
. The time it takes the block to move from position a to position b can be
obtained by taking the time dierence between these two points. This may be expressed as:
change
pos
=
x
b
x
a
t
b
t
a
(2.4)
We may further rene the above equation to get:
change
pos
=
x
t
(2.5)
The numerator of Eq. (2.5) is the dispacement of the block and the denominator is the elapsed
time. However, one must understand that no matter how large x is, the change will always
be the average change in the displacement with respect to time. This change has a particular
name and we know it is the velocity v of the block. This could be written as:
v =
x
t
(2.6)
But what if we want to observe the change in the position of the block at every instant of
time. If we want to observe parameters at each and every instant then what will be the value
of t. In this case we apply a limiting process on Eq. (2.6):
v
ins
= lim
t0
x
t
=
dx
dt
= x (2.7)
By following exactly the same logic we may write the instantaneous acceleration as:
a
ins
= lim
t0
v
t
=
dv
dt
= x (2.8)
2.1.2 Equation of Motion for the Spring Mass System
Now we can rewrite Eq. (2.3) by following the discussion in (2.1.1).
kx = m x (2.9)
Rearranging the above equation we get:
m x(t) + kx(t) = 0 (2.10)
2.1.3 Spring Mass System Solution
By observing Eq. (2.10) we realize that in order to solve this equation we need a function
which when dierentiated twice gives itself as the function upto some constant. We also
know that sinusoidal functions have such properties and therefore intuitively there must be a
sinusoidal function which solves the above equation. Let us make a guess at such a solution
and see if it works:
x(t) = Acos (t) (2.11)
2.2. INITIAL CONDITIONS 5
A is an arbitrary constant giving us the amplitude of the proposed solution. The we used
here has to have rad/s as its units. This is so because the product t must produce radians
as it is the argument of a sinusoidal function. For simple harmonic motion we also know such
a quantity characterizes the angular frequency of the system. From chapter 1 we know that
for simple harmonic motion is given as:
=
2
T
(2.12)
T is the time period of a complete cycle and 2 is the span of one complete cycle. Now that
we have developed some intuition about , we substitute Eq. (2.11) in the equation of motion
of a simple mass spring system.
A
_
m
2
+ k
_
cos (t) = 0 (2.13)
Eq. (2.13) can be equal to zero in one of the three ways
m
2
+ k = 0
cos (t) = 0
A = 0 (2.14)
We know that cos (t) is an oscillatory function. It is zero only at specic points. To be
more precise this function is zero at /2 and all other integer multiples of that. A is also
not zero, in fact, it can not be zero, because if it is so then we do not have any solution to
start with. However, to satisfy Eq. (2.13), we need something which is zero at all instants of
time t. Thus, the only choice we have is to put the rst equation in Eq. (2.14) equal to zero.
Doing that yields:
m
2
+ k = 0
=
2
=
k
m
=
_
k
m
(2.15)
2.2 Initial Conditions
Eq. (2.10) is a second order dierential equation. Elementary calculus tells us that solving a
second order dierential equation results in two constants of integration. Two constants of
integration mean two unknown values. This requires two equations as we want to solve for two
unknowns. The two required equations come from what are known as initial conditons.
Initial conditions are the values of the state of the system at the instant when you start
observing it. For you this is your time t = 0. One such set of conditions could be the initial
displacement and initial velocity and could be given as:
x(t) = x
o
|
t=0
x(t) = v
o
|
t=0
Now let us go back to Eq. (2.11) which represents the solution of the simple mass spring
system. We realize that there are is just one arbitrary constant to play around with. However,
6 2.2. INITIAL CONDITIONS
two roots in Eq. (2.15) suggest that there are two solutions to the original dierential equation
and there should be two corresponding arbitrary constants. Since we are dealing with linear
dierential equations we may combine the two solutions as one superimposed solution as is
shown below:
x(t) = A
1
cos(
1
t) +A
2
cos(
2
t) (2.16)
Here
1
refers to +
_
k/m and
2
denotes
_
k/m. If we dene to be equal to
_
k/m then
Eq. (2.16) may be rewritten as:
x(t) = A
1
cos(+t) +A
2
cos(t) (2.17)
Since cos() = cos(), Eq. (2.17) may be rewritten as
x(t) = A
1
cos(t) +A
2
cos(t)
x(t) = (A
1
+A
2
) cos(t)
x(t) = Acos(t) (2.18)
All of this hardwork brought us back to the original proposed solution with only one arbitrary
constant. What if we propose some other solution and see if it gives us two arbitrary constants.
So we propose the following solution:
x(t) = Asin(t) (2.19)
It turns out that Eq. (2.19) is also a solution to Eq. (2.10). Following the exact same procedure
as with Acos(t) solution, we realize that we end up with only one arbitrary constant. A way
out of this situation is to again use linearity of the dierential equation as a basis of adding
these two solutions to get two arbitrary constants. So the nal solution is given as:
x(t) = A
1
cos(t) +A
2
sin(t) (2.20)
2.2.1 A Neat Trignometric Trick
In this section we will introduce a method of rewriting Eq. (2.18). This equation represents
one way of writing the solution to Eq. (2.10). We may write the solution in other forms. This
is important as it will facilitate us in incorporating initial conditions into the solution. This
other form is obtained by considering the following:
A
1
= Acos ()
A
2
= Asin () (2.21)
Replacing the values of A
1
and A
2
from Eq. (2.21) into Eq. (2.20), the latter may be rewritten
as:
x(t) = Acos(t) cos () +Asin(t) sin () (2.22)
To simplify Eq. (2.22) we use the following trignometric identity:
cos(a b) = cos(a) cos(b) + sin(a) sin(b)
Identifying a as t and b as , Eq. (2.22) may nally be rewritten as:
x(t) = Acos (t ) (2.23)
2.2. INITIAL CONDITIONS 7
Where now the arbitrary constants are the amplitude A and the phase . Dierentiating
Eq. (2.23) we get:
x(t) = Asin (t ) (2.24)
We can now substitute the initial conditions to obtain the following:
x(0) = Acos () = x
o
x(0) = Asin () = v
o
(2.25)
Multiplying the rst equation by , squaring it and adding it to the square of the second
equation in Eq. (2.25) results in the following:

2
A
2
_
cos
2
() + sin
2
()

=
2
x
2
o
+ v
2
o
The term in square brackets is equal to 1. Simplifying the above further, one can easily obtain
the following:
A =
_

2
x
2
o
+ v
2
o

(2.26)
To obtain the phase one could simply divide the two equations in Eq. (2.25) to get:

sin ()
cos ()
=
v
o
x
o
tan =
v
o
x
o
= tan
1
_
v
o
x
o
_
(2.27)
Therefore the complete solution of a simple mass spring system can be written as:
x(t) =
_

2
x
2
o
+ v
2
o

cos
_
t tan
1
_
v
o
x
o
__
(2.28)
One must note that the above is a solution to an undamped system with initial conditions as
mentioned in Eq. (2.25). We will study the damped cases later.
Exercise 1
Show that at time t = 0, x(t) is indeed equal to x
o
.
Solution
Substitute in Eq. (2.28) the time t = 0 to get:
x(0) =
_

2
x
2
o
+ v
2
o

cos
_
tan
1
_
v
o
x
o
__
From elementary calculus we know that cos () = cos . We also know that:
= tan
1
_
v
o
x
o
_
8 2.2. INITIAL CONDITIONS

x
o

v
o
x
o
2

2
+
v
o
2
Figure 2.2: Phase Angle and associated Triangle
Using the above value of , one can write the following relation:
tan =
v
o
x
o
=
perp
base
Using now the above equation and the Pythagoras theorem we can easily construct the
triangle shown in Fig. 2.2.
Using the triangle in Fig. 2.2, we can rewrite the solution as:
x(0) =
_

2
x
2
o
+ v
2
o

cos
_
tan
1
_
v
o
x
o
__
= x(0) =
_

2
x
2
o
+ v
2
o

cos
= x(0) =
_

2
x
2
o
+ v
2
o

x
o
_

2
x
2
o
+ v
2
o
x(0) = x
o
Exercise 2
Show that at time t = 0, x(t) is indeed equal to v
o
.
Solution
Most of the work for this has already been done in Exercise 1. We will simply substitute
t = 0 in the expression for the velocity to obtain:
x(0) =
_

2
x
2
o
+ v
2
o

sin
_
tan
1
_
v
o
x
o
__
= x(0) =
_

2
x
2
o
+ v
2
o

sin
= x(0) =
_

2
x
2
o
+ v
2
o

v
o
_

2
x
2
o
+ v
2
o
x(0) = v
o
2.3. KINEMATICS OF AN UNDAMPED SPRING MASS SYSTEM 9
2.3 Kinematics of an Undamped Spring Mass System
Now that we know how to solve the dierential equation of motion of an undamped spring
mass system such as that shown in Fig. 2.1, we will as well see how the kinematics of the
system changes for a given set of initial conditions.
Let us assume that initial velocity v
o
is zero. The system is only given a displacement pull
equal to x
o
. Following from Ex. 1 we can easily determine the arbitrary constants. These are
given as:
A = x
o
= 0
We can also dierentiate with respect to time the expression of displacement, once for the
velocity and twice for the acceleration. Performing these steps we obtain the folowing:
x(t) = x
o
cos t
x(t) = x
o
sin t
x(t) =
2
x
o
cos t
Having obtained these quantities, we could plot them to assess the inter-relationship among
t
f(t)
x(t) = Acos (t) x(t) = Asin (t) x(t) =
2
Acos (t)

2

3
2
2
5
2
3
7
2
4
A
A

2
A
A
A

2
A
Figure 2.3: Kinematics of an Undamped Spring Mass System
displacement, velocity and acceleration for the initial conditions which state that x(0) = x
o
and x(0) = 0 and which have also been indicated above leading us to a phase value = 0. In
the plot it is assumed that > 1. Though the plot is not drawn to any scale, the fact that
> 1 is represented by arbitrarily increased magnitudes of velocity and accleration. This
plot is presented with the help of Fig. 2.3. It is important to note the following:
1. Displacement and velocity are 90
o
out of phase.
10 2.4. SIMPLE PENDULUM
2. Displacement and acceleration are 180
o
out of phase.
3. If the maximum amplitude of displacement is Athen the maximum amplitude of velocity
is times higher.
4. If the maximum amplitude of displacement is A then the maximum amplitude of accel-
eration is
2
times higher.
In general for varying values of Table (2.1) presents relationship between magnitudes of
displacement, velocity and acceleration.
|x(t)| | x(t)| | x(t)| Relationship
< 1
A A
2
A
|x| > | x| > | x|
> 1 |x| < | x| < | x|
= 1 |x| = | x| = | x|
Table 2.1: Relationship between Magnitudes of Kinematics for Various Values of Angular
Frequencies.
2.4 Simple Pendulum
Fig. 2.4 represents a simple pendulum. A simple pendulum consists of a string which is
attached to a support at one end and to a point mass m at the other. The string is taken
to be inextensible with negligible mass. At the equilibrium position as shown by the dotted
mass object hanging right below the origin O, there is no motion and the forces, therefore,
are ought to be equal.
O
m
T
f
g
= mg

m
l
m

T
mg
l sin
mg
h
P
Figure 2.4: A Simple Pendulum
The forces acting on this system can be investigated in two phases.
2.4. SIMPLE PENDULUM 11
1. Static Phase: At this stage the following could be written:
T f
g
= 0
T = f
g
Where f
g
is the weight of the pendulum and T is the tension in the string acting counter
to the weight of pendulum.
2. Dynamic Phase: In dynamic phase the mass is then raised from its equilibrium posi-
tion to a height h and released. We observe that the mass moves back to its equilibrium
position. Once it is at its equilibrium position, it does not stop and keeps moving until
it reaches the other extreme which is also shown by a dotted line and mass at an angle
from the center line. If there are no resistive forces acting on the system to dampen
the motion, then the mass would continue to follow the dotted circular trajectory.

i
F
(x)
i
= T sin

j
F
(y)
j
= T cos mg
By writing the balance of forces in the x and y directions we are implying that motion is taking
place in both x and y directions. This is true but the motion as we observe it is actually
taking place in a circular direction and we ask ourselves: Is it possile to use Newtons
Laws for circular motion?. The answer is yes and here is how it is expressed. Writing
Newtons Second Law for circular motion, we may state that the sum of all the torques
is equal to the product of mass moment of inertia of the object multiplied by its angular
acceleration. Mathematically this could be written as:

k
= J

(2.29)
Here represents the net torque on the system, J is the mass moment of inertia of the pen-
dulum (J is dened with respect to the point of rotation of the pendulum which is O in the
present case) and

is the angular acceleration of the pendulum, which is also the double time
derivative of the angular coordinate . It is implied that the motion is not due to a resultant
torque or moment which is causing the pendulum to move.
The only moment in the system around the point of rotation O is that which is caused by
the weight of the pendulum. We can write this torque as the cross product of the vector

OP with the weight vector



f
g
. This is given as:

OPx

f
g
(2.30)
Since the motion is taking place in the xy plane only and the only non-zero torque is acting
along z axis only, we can drop the vectorial notation. However, we still need a magnitude of
the above expression as well as the sign that goes along with the expression. This expression
can be obtained as follows:
|

| =

OPx

f
g

= |

| =

OP

f
g

sin
= lmg sin (2.31)
12 2.4. SIMPLE PENDULUM

OP

is nothing but the length of the pendulum, which is equal to l and


f
g

is the weight
of the pendulum mg. Also what we know is that this torque is a restoring torque and hence
it always acts in a direction opposite to the applied stimulus. Taking this into account, we
could state that the sign of the above torque is coordinate dependent. Not only that but also
that it will always act in a direction opposite to the direction of motion and therefore it will
always be accompanied by a negative sign. Substituting in Eq. (2.29) we obtain:
mgl sin = J

+ mgl sin = 0 (2.32)


This is the equation of motion of a simple pendulum.
2.4.1 Linearization of the System
Eq. (2.32) is a non linear dierential equation in . Although it is possible to solve it using nu-
merical or some other complex methods, we still want to solve it using the standard methods
for linear equations that we have learnt so far. Therefore, we need to linearize it. In order to do
that, we rst want to investigate the Taylors Series expansion of sine and cosine functions.
A Taylors series for a function is its approximations using a summation of innite number
of terms. This approximation is performed at a single point. If this single point happens to
be zero, then this series becomes Maclaurins Series. Without going into further details,
we will simply state the approximation for the sine and cosine functions:
sin =

n=0
(1)
n
(2n + 1)!

2n+1
=

3
3!
+

5
5!
. . .
cos =

n=0
(1)
n
(2n)!

2n
= 1

2
2!
+

4
4!
. . .
If is small then we can approximate the above functions as:
sin
cos 1
If we now use these approximations in Eq. (2.32), we will get the following:
J

+ mgl = 0
=

+
mgl
J
= 0 (2.33)
Comparing it with the equation of motion of a spring mass system x+
2
x = 0, we can easily
realize that:

2
=
mgl
J
=
_
mgl
J
(2.34)
2.4. SIMPLE PENDULUM 13
Where is the angular frequency of the system measured in rad/s. Since = 2/T, therefore,
the time period of a simple pendulum is given as:
T = 2

J
mgl
(2.35)
2.4.2 Mass Moment of Inertia
Eq. (2.34) and Eq. (2.35) give us the natural angular frequency and time period of a simple
pendulum in terms of I which is the mass moment of inertia of the system.
r
i
m
i
x
y
z
Figure 2.5: A circular Disc
Mass Moment of Inertia of an object is a measure of its resistance against circular motion.
It plays the same role as is played by mass of an object in rectilinear motion. To obtain a
general expression of this quantity we consider a solid disc as shown in Fig. 2.5. In Fig. 2.5
is shown a portion of the disc and we are interested in nding out the kinetic energy of this
smaller portion. If KE
i
denotes this energy then we can express it as follows:
KE
i
=
1
2
m
i
v
2
i
(2.36)
And therefore the total energy can be obtained by summing the individual energies as given
below:
KE =
1
2
n

i=1
m
i
v
2
i
KE =
1
2

m
1
v
2
1
+ m
2
v
2
2
+ m
3
v
2
3
+ . . . + m
n
v
2
n
KE =
1
2
mv
2
(2.37)
14 2.5. PHYSICAL PENDULUM
Now let us reuse Eq. (2.36) and replace the rectilinear velocity with the angular velocity.
Therefore, rewriting it in angular terms gives us:
KE
i
=
1
2
m
i
r
2
i

2
(2.38)
Note that we have used the simple relationship that v
i
= r
i
. Also note that since we are
using a solid disc, is the same for all mass particles. Now the total kinetic energy of the
system can once again be obtained by summing up the individual energies over all elementary
mass particles. This gives:
KE =
1
2

2

i
m
i
r
2
i
(2.39)
Since the energy of the moving disc is a physical quantity and it must bear the same numerical
value whether we calculate it using rectilinear terms or angular terms, therefore the result of
Eq. (2.37) and Eq. (2.39) must be the same. Also we know that the angular counterpart of
the rectilinear velocity v is , thus the angular counterpart of mass m of the disc must be
the quantity under the summation in Eq. (2.39). This quantity is known as mass moment
of inertia and it is given as:
J =

i
m
i
r
2
i
(2.40)
In other words the role that mass plays in rectilinear motion is the same as performed by
mass moment of inertia in angular motion.
Mass moment of inertia for various solids is available in mechanics textbooks and online
resources. This text will not go into further details for calculation of the same. However, the
system at hand, namely a simple pendulum has a mass moment of inertia equal to ml
2
.
Using this value of J in Eq. (2.35), the angular frequency is given as:
=
_
mgl
ml
2
=
_
g
l
(2.41)
2.5 Physical Pendulum
A physical pendulum also known as a compound pendulum is an arbitrarily shaped rigid
body which can swing around a pivot point . A pivot can be any supporting point, e.g., a
nail onto which an irregularly shaped plate is hung. This nail performs the job of a pivot as
the plate can swing around it while the irregularly shaped plate itself becomes the compound
pendulum. In fact the entire system when studied under isolation could be termed as a com-
pound pendulum.
Such a pendulum is dened with the help of Fig. 2.6. Although, admittedly, this is quite
nicely shaped; however, the model that we will develop will be applicable to any irregular
shape also. O is the pivot point, C is the center of mass and CP is the center of percussion
which will be dened later and l is the length of the pendulum.
2.5. PHYSICAL PENDULUM 15
f
g
f
g

r
r sin
f
g
R
R
l
cp
l
h
O
C
C
CP
CP
Figure 2.6: A Compound Pendulum
So what happens physically is that the object such as that shown in Fig. 2.6 is initially in
equilibrium. At this stage we can write the following:
f
g
= R (2.42)
Here f
g
denotes the weight of the object acting at the center of mass C, whereas, R is the
reaction force which develops at the support / pivot point O. The object is then raised by
a height h = r(1 cos ) and let go. It starts oscillating about the pivot point O. We are
interested in nding the angular frequency of this compound pendulum which will then dene
the vibration characteristics of this object. From (2.4) we realize that the only moment
acting about the pivot point in this entire system is a restoring torque which is due to the
weight of the object f
g
= mg acting at the center of mass C. The magnitude of this torque
is given by:
= mgr sin (2.43)
Using Newtons law of balance of moments of torque we could write the following:
J

= mgr sin
J

+ mgr sin = 0

+
mgr
J
sin = 0

+
mgr
J
= 0 (2.44)
Where the sign in the rst line of Eq. (2.44) indicates that the RHS is a restoring torque.
Also one immediately realizes that since we assume to be small therefore, we may approxi-
mate sin . From the last line in Eq. (2.44) we can deduce the natural angular frequency
16 2.5. PHYSICAL PENDULUM
of the system to be:
=
_
mgr
J
(2.45)
2.5.1 Center of Percussion
We want to nd out the length of a simple pendulum whose point mass is equal to the mass
of the compound pendulum such as that shown in Fig. 2.5 and whose frequency is also the
same as that of this compound pendulum. The length which satises this criteria is called the
length of percussion and when this length is projected on the compound pendulum the the
point where it intersects the center line is called the center of percussion. Mathematically
we may obtain it as follows:

(s)
=
(c)
_
g
l
cp
=
_
mgr
J
(c)
g
l
cp
=
mgr
J
(c)
l
cp
=
J
(c)
mr
(2.46)
Where J
(c)
is the mass moment of inertia of the compound pendulum, m is the mass and
r is the distance from the pivot point to the center of mass C. The superscripts (s) and (c)
respectively denote the simple and compound pendulums.
The signicance of center of percussion is that it is believed that it denes a sweet spot
for racquets, bats and swords where an impact results in maximum velocity imparted to
the striking object such as a ball in case of a cricket or a baseball bat. For a sword that would
mean that the required force to cut through will be small. However, having said that, to the
best of authors knowledge, this belief is under active research consideration and some results
suggest that this theory of sweet spot is probably not right.
2.5.2 Radius of Gyration
Since the compound pendulum is undergoing circular vibrational motion then the measure of
resistance oered by this system is its mass moment of inertia.
Consider a circular hoop, the mass m of which is the same as the mass m of the compound
pendulum. If we try to set this hoop into circular motion then it will oer some resistance.
As with the compound pendulum the measure of this resistance is its mass moment of inertia.
Now we are looking for a circular hoop which oers the same resistance as does the compound
pendulum. When we nd such a hoop we term its radius as radius of gyration. This is
obtained mathematically as:
J
(h)
= J
(c)
mr
2
gy
= J
(c)
r
gy
=
_
J
(c)
m
(2.47)
2.6. ENERGY METHOD 17
From Eq. (2.46) we deduce that the mass moment of inertia of a compount pendulum can be
written as:
J
(c)
= mrl
cp
(2.48)
Substituting this value in Eq. (2.47), radius of gyration associated with a compound pendulum
can be given as:
r
gy
=
_
rl
cp
(2.49)
The signicance of radius of gyration is that it gives us an eective radius such that if the
mass of the object is known than its mass moment of inertia can be easily calculated using
J = mr
2
gy
.
2.6 Energy Method
For conservative systems the equation of motion of any oscillatory system can be obtained by
dierentiating its total energy with respect to time. This can be given mathematically as:
d
dt
TE(t) = 0 t (2.50)
The RHS of Eq. (2.50) is zero because the method is used on conservative systems. Con-
servative systems are dened as systems which do not lose their energy during the course
of motion. This is, of course, ideal because no system runs forever and eventually all the
energy in the system is exhausted. But for the undamped cases, especially the onese where
one is only interested in the natural frequencies of the system this method is used. In general,
for complex cases where it is dicult to identify forces acting on the objects or for higher
dimensional cases, equation of motion cannot be easily obtained. In such a situation energy
method is used.
However, we will use a very simple case to elaborate the method. We will use a simple
pendulum as shown in Fig. 2.4. Let TE, KE and PE respectively denote the total energy,
kinetic energy and potential energy of the system. Then at any instant of time t the total
energy can be expressed in terms of its kinetic energy and its potential energy. This can be
expressed mathematically as:
TE(t) = KE(t) + PE(t) t (2.51)
Substituting Eq. (2.51) in Eq. (2.50), we obtain the following:
d
dt
[KE(t) + PE(t)] = 0 (2.52)
Now the only thing left is to gure out the expressions for the potential and kinetic energies
of a simple pendulum. These are respectively given as:
PE = mgh
KE =
1
2
J
_

_
2
(2.53)
18 2.7. EQUIVALENT SPRING CONSTANT
Referring to Fig. 2.4 we realize that the height h by which the pendulum is raised is given by:
h = l (1 cos ) (2.54)
Substituting Eq. (2.54) and Eq. (2.53) in Eq. (2.52) and simplifying, we get the following:
d
dt
_
1
2
J
_

_
2
+ mgl (1 cos )
_
= 0
_
J

+ mgl sin
_

= 0
J

+ mgl sin = 0

+
mgl
J
sin = 0

+
mgl
J
= 0 (2.55)
The method described here is a simplication of the general Lagrangian Energy Method
which is used for complex systems where the application of Newtonss Laws to obtain the
equation of motion becomes dicult.
2.7 Equivalent Spring Constant
Sometimes a spring with desired rigidity is not available in which case springs can be combined
in one of the following two ways:
1. in parallel
2. in series
When the springs are joined in parallel the equivalent spring constant is given by
k
eq
= k
1
+ k
2
and when the springs are connected in series the equivalent stiness is given by:
k
eq
=
k
1
k
2
k
1
+ k
2
The way springs can be combined is explained with the help of Fig. 2.7. In Fig. 2.7 (a) two
springs of unequal stiness, namely, k
1
and k
2
are combined in parallel while in the part (b),
two springs having rigidities k
3
and k
4
are combined in series.
2.7. EQUIVALENT SPRING CONSTANT 19
k
1
k
2
k
eq
= k
1
+ k
2
k
3
k
4
k
eq
=
k
3
k
4
k
3
+ k
4
(a)
(b)
Figure 2.7: (a) Springs in Parallel. (b) Springs in Series.
Exercise 3
Calculate equivalent stiness for the network of springs shown in Fig. 2.8.
m
k
a
k
b
k
c
k
1
k
2
k
3
k
4
k
5
Figure 2.8: A Network of Dierent Springs
20 2.8. ANGULAR FREQUENCY FOR GENERAL CASES
Solution
The best way to solve Ex. 3 is to isolate subsystems of springs which can be combined using
rules given in (2.7). Therefore, three regions are identied and the equivalent stiness values
are termed respectively as k
a
, k
b
and k
c
. The values of these equivalent spring constants are
given as:
k
a
=
k
1
k
2
k
1
+ k
2
k
b
= k
a
+ k
3
k
c
= k
4
+ k
5
The equivalent spring constant for the top half of the spring mass system is k
b
and the that
for the lower part of the system is k
c
At rst glance one might think that the nal spring
constant would be the result of combining k
b
and k
c
in series, however, this is not true and
these two springs are connecting the mass at the same potential level (or same height if you
will). Therefore, the upper and lower systems in fact are in parallel to each other with respect
to the mass m and thus the nal spring constant will be given as:
k
eq
= k
b
+ k
c
2.8 Angular Frequency for General Cases
We now have developped the idea that in nding out the natural frequency of a system, we
either need the restoring force or the restoring torque. In case where forces are dicult to
nd, we resort to energy method (energy method was dealt with here at a very basic level).
Having gured out the above we always nd, atleast for the simple systems that we have
studied so far, that the natural frequency is of the following form:
=
_
Restoring Force Coecient
Measure of Inertia
Following this basic idea, we may investigate the natural frequency of the torsional pendulum
as shown in Fig. 2.9. The disc is given an initial rotation about the z axis. Because of the
torsional rigidity the disc rod twists back by a restoring torque whose magnitude is given
by:
=
Where is given in terms of G the shear modulus of the rod, J
p
the polar moment of inertia
of the rod and l the length of the rod as shown in Fig. 2.9. The fact that the resistance to
motion is oered by the mass moment of inertia of the circular disk which has a mass m and
radius r, the angular frequency of the system can be given as:
=
_

J
(d)
(2.56)
And the mass moment of inertia for the disc in Fig. 2.9 is given by:
J
(d)
=
1
2
mr
2
2.9. CONCLUSION 21
m, r
l
=
GJ
p
l

x
y
z
Figure 2.9: A Torsional Pendulum
We have gured out the natural frequency of the torsional pendulum without explicitly solving
for the equation of motion. This is left as an exercise and the reader may conrm this. Based
on this simplication, we can now generalize the method for nding angular frequencies for
simple systems studied so far. The natural frequencies for various systems are given with the
help of Table (2.2). The superscripts (s), (c) and (d)respectively denote simple pendulum,
System Restoring Force Inertia Angular Frequency
Spring Mass kx m
_
k
m
Pendulum mgl J
(s)
_
mgl
J
(s)
Compound Pendulum mgr J
(c)
_
mgr
J
(c)
Torsional Disc J
(d)
_

J
(d)
Table 2.2: Natural Frequencies for Various Systems.
compound pendulum and torsional disc.
2.9 Conclusion
In this chapter we studied very simple oscillatory systems and obtained very fundamental
characteristics such as the natural frequencies of vibrational systems and time periods. We
used Newtonss Laws for balance of forces and torques to obtain the equation of motion. We
then solved those equations to obtain the kinematics of those systems. We then introduced
energy method for nding equation of motion for systems with complex forces. The treatment
presented was quite basic. Then equivalent spring constants were introduced and nally the
method to nd out the angular frequencies was generalized.
Personal beauty is a greater
recommendation than any let-
ter of introduction.
Aristotle (384 BC-322 BC)
3
Damped Oscillations
In this chapter, we will build upon the concepts developed so far and gain insight into real
physical systems. A real oscillatory system is one which will eventually die out because it
will have used up all its energy. The question is where has this energy come from and more
importantly how does the system use up this energy. Energy may come from pulling the block
of mass in case of simple spring mass system and in case of a simple pendulum it might come
from raising the bob of pendulum and letting it go. It is not the only method of imparting
energy, in fact, nstead of pulling the mass, if it is just given a push in any direction along
the axis of motion, this will also be considered as imparting energy to the system. In the
rst case, we are giving the system a displacement and in the second case we are giving the
system a velocity. Note that the system can be given both a displacement and a velocity at
the same time. This is how energy is imparted to the system. But why do these systems die
out eventually? The reason for these systems to eventually stop is that all the energy that
was imparted to the system initially is ultimately dissipated as heat energy and that energy
disseminates in the surrounding atmosphere. In vibration this loss of energy by the system
is attributed to a physical process which is called damping.
3.1 Damping
Damping is an inherent characteristic of any oscillatory system. This is to say that whenever
a system such as a spring mass system is excited by any of the methods above, we will see
that after some time the system will stop. Although we did not specically try to stop it
by applying opposing forces for example, but even then the system comes to a halt. All real
systems behave in this manner. This means that the ideal oscillatory systems that we studied
in Chapter 2 need renement. Sometimes the time it takes to completely exhaust all the
energy is very long and it takes more than just a couple of minutes for the oscillation to stop.
In some other instants the damping is so high that we can not possibly notice the response
of the system and to us the system appears un moved whereas we did not realize that the
system did move, but the dynamics of the system were such that it died out in fraction of a
22
3.1. DAMPING 23
second (e.g. a system which oscillates at 100 cycles per second and it requires only 50 cycles
to die out will come to a halt in only half a second).
Now we want to incorporate damping in our system as an oppsosing force which makes the
system come to a stop. We already know one opposing force and that is the spring force whose
magnitude is kx. However, we also know that this term does not help the system to dissipate
energy, rather, this force makes the system vibrate forever. So, the force we are looking for
needs to be dierent from this force. Now spring force is proportional to the diplacement of
the mass from the equilibrium position. So, we need to think in terms of velocity, the next
quantity after displacement. The truth is that one kind of damping forces is proportional to
the velocity of vibration of the oscillating system. There are others which for instance depend
on the friction of the surface on which the system moves. However, we will take up those
models later and will only concentrate on velocity dependent damping forces. Let us rst
produce the mathematical form of the damping force and then elaborate. This form is given
with the help of Eq. (3.1).
f
d
= c
1
x c
2
x
2
(3.1)
3.1.1 Pressure / Drag Damping
When the oscillating object is moving fast, then the opposition in notion is more severe in
terms of the pressure or drag which is proportional to the square of the velocity as shown
in Eq. (3.1). To understand what it feels to be opposed by pressure, take your hand out from
a fast moving car and feel the wind at your hand. Please do it for a short while and with
utter caution, because if you are careless, you might lose your arm. Or, think of a falling drop
of rain. The drop is moving at a high velocity. The wind pressure opposing the fall will be
proportional to the square of the velocity. The term which becomes important in this case is
the second term on the RHS of Eq. (3.1), and the damping force can be given as:
f
d
= c
2
x
2
(3.2)
3.1.2 Viscous Damping
When the oscillating object is moving slow, then the term that takes precedence is the rst
term on the RHS of Eq. (3.1), which represents the viscosity related damping and is pro-
portional to the velocity of the moving object. To appreciate this force, think of an iron bob
slowly drowning in a pool of oil. The upward force felt by the bob will come from the viscosity
of the oil and this force can be given as:
f
d
= c
1
x (3.3)
For all practical purposes and to keep the mathematics simple, we will consider Eq. (3.3) to
model damping in real systems. Since we will only be using one form of damping, therefore,
we will use the notation f
d
= b x for viscous damping force with b acting as the viscous
damping coecient.
24 3.2. A SPRING MASS SYSTEM WITH DAMPING
3.2 A Spring Mass System with Damping
The system in Fig. 3.1 will now be considered to model damping into oscillating systems. A
mass is vertically hung from a spring and a damper as shown in Fig. 3.1. Before attaching the
mass m to the system, the spring and the damper are loosely hanging from the xed support.
The unstretched coordinate of the system is denoted by x which is the vertical coordinate.
It is zero in this unstretched position (x = 0). This position is indicated in Fig. 3.1. In this
state the system is in static equilibrium. Applying Newtons 2
nd
Law on this system yields:

x
F
x
= 0
The mass m is then connected to the system. The force of gravity pulls the mass toward the
earth which results in a deection of the mass. The system is still in static equilibrium,
therefore, the sum of the forces acting on the system is again zero. However, the old coordinate
x is no longer zero. Its value now is x = . So, we dene a new coordinate x whose value is
zero at x = as shown in the gure. Looking at Fig. 3.1 one can immediately recognize the
x = 0
x = x = 0
m
k
f
g
= mg
f
s
= kx f
d
= b x
b
Figure 3.1: A Vertical Spring Mass System with a Damper
equivalence between the two coordinates as shown with Eq. (3.4).
x = x (3.4)
There now are two coordinates x and x with the former representing the position of the mass
from the unstretched spring position and the latter is the coordinate of the mass from the
stretched postion. In the stretched position when the mass is deected by , the balance of
forces in the vertical direction gives:
mg = k (3.5)
3.2. A SPRING MASS SYSTEM WITH DAMPING 25
3.2.1 Equation of Motion from Unstretched Position
The mass in Fig. 3.1 is pulled down and let go. It starts moving away from the direction
of the pull and towards the equilibrium position x = . As of now it is not known whether
due to damping the system will come to a complete stop without crossing the equilibrium
position or the system will oscillate a little before coming to a halt. The behaviour of the mass
under the action of the pull will be discussed later. For now the equation of motion is obtained.
It is now clear that the coordinate at the unstretched position is given by x. From this
positon, writing the balance of forces in the vertical direction produces the following:
m x = b x kx + mg
m x + b x + kx = mg (3.6)
The non zero RHS of Eq. (3.6) implies that when the mass nally cones to rest, its position
will be x = and not x = 0.
3.2.2 Equation of Motion from the Stretched Position
Now the coordinate at the stretched position namely x is considered. Instead of rewriting the
balance of forces, the second equation in Eq. (3.6) is utilized in conjunction with the value of
x in terms of x. From Eq. (3.4) it can be seen that x can be rewritten in terms of x:
x = x + (3.7)
Substituting this value of x in Eq. (3.6) results in the following:
m x + b x + kx = mg
m

x + b

x + k ( x + ) = mg
m

x + b

x + k x + k = mg
m

x + b

x + k x = 0 (3.8)
Eq. (3.8) has been simplied by considering that mg = k and also the fact that the rst and
the second time derivatives of are zero and therefore, with m x and b x, the deection does
not appear.
With two coordinates it will be a hassle to keep track of them when both the coordinates
represent the motion of the same object. A more preferable approach is to keep the same
coordinate x for both the cases and only concentrate on the concept. The concept is explained
again as follows:
When the equation of motion is derived from the unstretched spring position, the RHS
has the gravitational force mg.
m x + b x + kx = mg
When the equation of motion is derived from the stretched spring position, the RHS
does not have the gravitational force term.
m x + b x + kx = 0
26 3.3. DAMPED OSCILLATIONS SOLUTION
3.3 Damped Oscillations Solution
Now the distinction between the unstretched and stretched postions is clear, the latter
(Eq. (3.8)) which is also the simpler one can be divided by m and rewritten to produce the
following:
x + x +
o
2
x = 0 (3.9)
It should be clear that using the stretched version of the equation, the force of gravity term mg
has already been canceled out. is dened as the normalized damping coecient whose value
is given as b/m.
o
2
is the square of the natural frequency given by k/m. It is worth noting
that the dimensions of and
o
are the same. Table (3.1) highlights important parameters
in Eq. (3.9) and their interpretations.
Description Symbol Equals
Normalized Damping Coecient b/m
Natural Frequency
o
_
k/m
Table 3.1: Eq. (3.9) Parameters and their Equivalence.
There are many ways of solving Eq. (3.9). One of those consists of considering the solution
function x(t) to be complex valued. However, since the system lives in the physical world,
any transformation that is done to convert the real world displacement coordinate x into a
complex numbered coordinate has to be reverted back. A simple way of expressing such a
transformation is given as shown below:
x(t) = (z(t)) (3.10)
z(t) is a complex valued displacement coordinate. Eq. (3.9) can therefore be rewritten as:
z + z +
o
2
z = 0 (3.11)
Next the complex valued solution z(t) is further expressed as:
z(t) = Ae
iat
(3.12)
Substituting Eq. (3.12) into Eq. (3.11) results in the following:
_
a
2
+ ia +
o
2
_
Ae
iat
= 0 (3.13)
The only non trivial solution is obtained by putting the terms in brackets on the LHS equal to
zero. However, care must be taken in separating the real and imaginary parts as the quantity
in the bracket is complex valued. Doing just the same gives rise to:
a
2
+
o
2
= 0
a = 0 (3.14)
It is impossible to solve Eq. (3.16) because the set of equations given by Eq. (3.16) suggests
two values for the arbitrary constant a and a cannot have two contradictory values at the
3.3. DAMPED OSCILLATIONS SOLUTION 27
same time. There is only one way out of this confusion. That is, a has to be complex itself
and that shall resolve this problem. In other words a is considered to be:
a = + i (3.15)
Rewriting the bracketed expression in Eq. (3.13) in terms of Eq. (3.15), the following is
achieved:
( + i)
2
+ i ( + i) +
o
2
= 0

2
+
2
i2 + i +
o
2
= 0 (3.16)
It is apparent that Eq. (3.16) has complex numbered parameters. In order to solve this
equation, the real and imaginary parts have to be written separately. Writing these parts
separately means both parts would identically equal to zero. These parts are given below:
i (2 ) = 0

2
+
2
+
o
2
= 0 (3.17)
Solving Eq. (3.17) is pretty straight forward. Doing that results in the following:
=

2
(3.18)

2
=
o
2
+

2
4


2
2
=
_

o
2


2
4
(3.19)
has two roots, one with plus sign and the other with minus sign. The desired function z
will, therefore, also have two possible solutions, one according to the plus root and the other
according to the minus root. The solution to z is thus given as:
z(t) = A
1
e
i(
1
+i)t
+A
2
e
i(
2
+i)t
(3.20)
The above solution may be written in simplied form as:
z(t) = A
1
e
t
e
i
1
t
+A
2
e
t
e
i
2
t
(3.21)
3.3.1 Solution Analysis
At no stage should the reader ever forget that z(t) is a complex valued solution to Eq. (3.8).
The presence of complex exponential also arms this. In addition to this, there is no re-
striction on the two arbitrary constants A
1
and A
2
to be real valued. These two constants
could also be complex valued. However, oscillatory systems under study live in real world
and this fact leads to the question: Is a complex valued solution acceptable? The answer is
obviously no and a simple solution to this inconvinience is to only consider the real part of
the solution while discarding the imaginary part. This could mathematically be represented
by reproducing Eq. (3.10) as shown below:
x(t) = (z(t)) (3.22)
28 3.4. THE GREAT BATTLE: DAMPING VS. OSCILLATION
3.3.2 Roots of Dierential Equation
The complex roots of the dierential equation Eq. (3.11) in terms of the constants and
are given in Eq. (3.19) and Eq. (3.18) respectively. In general the behaviour of the system
depends on the values of these constants. In particular the value of decides if the system
behaves like an oscillatory system or not. The value of determines the sharpness of the
exponentially decaying envelope. Examples of such envelopes are given in Fig. 3.2. Observing
the gure makes evident that larger the value of the quicker the curve goes to zero.
t
f(t) = e
t
= 0.5
= 1
= 3
Figure 3.2: Examples of Exponentially Decaying Envelopes for Dierent Values of
The other part of complex roots appears in the form of . This constant has two possible
values, one postive and the other negative as shown in Eq. (3.19). But there really is more to
than meets the eye. In fact this constant may also assume values based on the discriminant
present under the square root. The value of this discriminant decides the way in which the
system would oscillate. This is dealt with in details in the next section.
3.4 The Great Battle: Damping vs. Oscillation

You might also like