Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 19

CHAPTER 1: OSCILLATIONS

Hour: 1

1. INTRODUCTION TO OSCILLATION AND VIBRATION


There is lot of importance of vibration and oscillation in our practical world from the shock absorbers
in vehicles and machineries or cantilever beams in houses, oscillatory electric and magnetic field to a
simple swing in the garden. Both oscillation and vibration are seen in our daily life and sometimes
they are beneficial and at times become detrimental. In simple words, the motion in which an object
executes a to and fro motion about an equilibrium position is called an oscillatory motion but
vibrations can be in multiple directions. In oscillation the body undergoes a maximum displacement
with fixed time period of oscillation. Vibration can be referred to the physical change brought about
due to movement of the body. Oscillation can be seen in the swinging of a pendulum clock and
vibration in the plucking of guitar string.

In oscillations the objects subjected to a restoring force when displaced from its equilibrium position
performing to and fro or motion. Such an object moves back and forth along path, repeating
over and over a series of motions. Such a motion of constant frequency is called periodic
motion or harmonic motion and the objects performing such type of motion are called harmonic
oscillators. There are three main types of simple harmonic motion:
(a) Free oscillations – simple harmonic motion with a constant amplitude and period and no external
influences.
(b) Damped oscillations – simple harmonic motion but with a decreasing amplitude and varying
period due to external or internal damping forces.
(c) Forced oscillations – simple harmonic motion but driven externally

(a) Free Oscillations of Simple Harmonic Motion:


Let us consider an object of mass, m attached to massless spring of spring constant, k fixed to
a rigid support is displaced from its equilibrium position by distance A along the x-axis as
shown in Fig 1.1.

Fig 1.1 Simple Harmonic Motion of an object.


Since the restoring force is directly proportional to the displacement and opposite in direction, we can
write it as,
F  x
1.1
F  kx

where, k is the spring constant, x is the displacement from the equilibrium position. The motion
executed by the object is because of this restoring force. Hence, using Newton’s second law of the
motion the equation of motion for the object can be written as follows.

max   kx 1.2
d 2x
m   kx
dt 2

d 2x k
 x0 1.3
dt 2 m

d 2x
2
 02 x  0 1.4
dt

Equation (1.4) is the differential equation for simple harmonic motion, which can be written as,

d 2x
2
 02 x 1.5
dt

k
Where  0 
2
is a constant dependent on the system parameters, k and m.
m

To solve equation (1.4 ) we consider a tentative solution for x in such a manner, which do not alters
after differentiating. Let the tentative solution for x as follows.

x  Ce t 1.6

Where, C and  are some arbitrary constants to be determined from the initial conditions of
oscillations. Now differentiating equation (1.6 ) twice w.r.t. x, we get the following equations.

dx
 Ce t
dt

d 2x
dt 2

 C 2 e t   2 Cet 
d 2x
Or, 2
  2x 1.7
dt

Now comparing equations (1.5) and (1.7) we can write

 2 x  02 x

 2

  02 x  0 1.8
Now in equation (1.8), x =0 is a trivial solution, i.e. the object is in the equilibrium position. For all
other positions, x≠0, hence

 2

 02  0

This means   i0 , which implies there can be two solutions of x for each value of from
equation (1.6). The most general solution can be written as a linear combination of these two solutions
as given below.

xt   C1e  i0t  C2 e i0t 1.9

Where, C1 and C2 are some arbitrary constants.


CHAPTER 1: OSCILLATIONS
Hour: 2

Using De Moivre’s theorem we can rewrite equation (1.9) as


xt   C1 cos 0t  i sin 0t   C2 cos 0t  i sin 0t 
xt   C1  C2 cos 0t  iC1  C2 sin 0t 1.10
Now we use the following substitutions
C1  C2   Asin  1.11
iC1  C2   A cos  1.12
Where, A and  are constant s to be determined from the initial condition of the system. Substituting
equations (1.11) and (1.12) in equation (1.10) we will get,
xt   Asin  cos 0t  A cos  sin 0t
xt   Asin 0t    1.13
The variation of x(t), vx(t) and ax(t) is shown in Fig. 1.1. given below.

Fig 1.1

Equation (1.13) is the required solution of S.H.M. Since, 1  sin   1, hence  A  xt   A . i.e. A
2
is the maximum displacement of the object. Again, if an increment of is given to t i.e.
0
2
t  t in equation 1.13, which will give
0

  2  
xt   Asin 0  t      A sin 0t    . It means the object arrives at the same position after
    
2
a time interval of, T  , which is the time period of motion. Again at t=0, for small values of
0
the object will not be in its equilibrium position but have some initial small displacement, which
may be on either side of the equilibrium position depending upon the value of is called the initial
phase of motion.
Inferences about the solution:
1. Solution is periodic in nature.
2. Displacement has sinusoidal (cosine) variation w.r.t.
3. It has constant amplitude throughout time.
4. It may have initial phase.
5. It oscillates with fixed time period.

Energy in Simple Harmonic Motions:

For a mechanical system executing simple harmonic motion the total energy is the sum total of
kinetic energy and potential energy possessed by the object at any position and at any instant of time.

E  K.E.  P.E.
E is the total energy, K.E. is the kinetic energy and P.E. is the potential energy.
Now,
2
1 1  dx 
K .E.  mvx2  m  . 1.14
2 2  dt 
Now differentiating 1.13 w.r.t. t, and substituting it in the above equation we will get
1

K .E.  mA202 cos 2 02 t  
2
 1.15

Similarly, P.E. is given by the amount work done by the restoring force and is given by the following
equation.
1
P.E.   F .dx     kxdx  kx2  C  .
2
Where C  is the constant of integration and can be set to zero, since x=0 in equilibrium position.
Hence from the above equation we can write,
1
P.E.  kx2 1.16
2
P.E.  kA2 sin 2 0 t   
1
1.17
2
Hence the total energy is given by
mA202 cos 2 0t     kA2 sin 2 0t    .
1 1
T .E. 
2 2
Since, m 0  k , we can rewrite the above equation as
2

2
 
T .E.  kA2 cos 2 0 t     sin 2 0 t     kA2  m02 A 2
1 1
2
1
2
1.18
The total energy in S.H.M. is always constant. However the K.E. and P.E. varies w.r.t. time and
position.
From equations (1.14) and (1.16), we see that K.E. and P.E. has parabolic variation w.r.t. position.
The variations of the energies w.r.t. position are given in fig. 1.2.

Fig 1.2
But the variations of energies w.r.t. time are oscillatory with periodicity of T/2. The variations w.r.t.
time is shown in fig. 1.3.

The average energies for a complete cycle can be evaluated as by averaging over one time period as
given below.

T T

0 2 mA 0 cos 0t   dt  cos  t   dt


1 2 2 2 2
0
1 1 1 1 1
K .E.  T
 mA202 0
 mA202  mA202  kA2
2 T 2 2 4 4
 dt
0
1.19

T T

0 2 kA sin 0t   dt  sin  t   dt


1 2 2 2
0
1 1 1 1 1
P.E.  T
 kA2 0
 kA2  kA2  m02 A 2
2 T 2 2 4 4
 dt
0
1.20
Average total energy,
1 1 1 1
T .E.  K .E.  P.E.  kA2  kA2  kA2  m02 A 2  T .E. 1.21
4 4 2 2
Inferences:
1. Average K.E. and P.E. energy over a cycle remains constant and are equal.
2. Average T.E. is equal to the T.E.
CHAPTER 1: OSCILLATIONS
Hour: 3

(b) Damped Harmonic Motion

In all practical cases of applications there is no true simple harmonic oscillation in mechanical
systems. For example, if we observe the oscillation of a simple pendulum made by hanging a bob with
a light thread and made to oscillate then it comes to rest after some time. There is resistance offered
by the medium say air to the free oscillation of the pendulum and ultimately brining the pendulum to
rest. Even we plan to oscillate the pendulum in an evacuated chamber there will be friction at the
point of pivot (hinge) which will slow down the oscillation. Again consider a situation when such a
system where a spring-block system is made to oscillate in a thick medium like mustard oil or
glycerine then the system may come to rest without executing a single oscillation. Such resistive force
opposes the free oscillation of the system and damps the amplitude is called damping force. The
damping force is also opposite to the displacement and has the same direction as the restoring force.
The magnitude of the damping force depends upon the velocity of the object.

Fig 1.4

Q. Why is the damping force directly proportional to the velocity of the object?

Hence the different forces acting on the object are


1. The restoring force, -kx, as explained in the previous section.
2. Damping force, -bvx, where b is the damping constant.
Since the motion of the object is governed by Newton’s 2nd law, we can write the following equation
of motion.

max  kx  bv x
d 2x dx
m 2  kx  b 1.22
dt dt
d 2 x b dx k
  x0 1.
dt 2 m dt m
The above equation (1.23) is the equation of motion for damped harmonic motion. We rewrite the
b k
above equation by using 2 s  , and 02  .
m m
d 2x dx
2
 2s  02 x  0 . 1.24
dt dt
We consider a tentative solution for equation (1.24) like that of S.H.M.
Let,
xt   Ce t 1.25
where, C and are arbitrary constants to be determined from the initial conditions. Now using the
first and second derivative of equation (1.25) i.e. equations w.r.t. t in equation (1.24) we will get
 2Cet  2sCet  02 x  0

Or,   2s  0 x  0
2 2
 1.26
We have used equation (1.25) in the above equation. Again, x=0 is a trivial solution. Hence for
nontrivial solutions we will have
 
 2  2s  02  0 , which is a quadratic equation in .
  s  s 2  02 1.27
This means equation (1.24) has two solutions for the two roots of . The most generalized solution
can be written as linear combinations of these two solutions, which will give us
  s  s 2  2  t  s  s 2  2  t
xt   C1e 
 C2 e  
0 0
1.28
Where, C1 and C2 are arbitrary constants. Considering that the object to be at rest at t=0, we can set
C1=-C2. Hence, equation (1.28) can be rewritten as,
  s 2 02  t  s 2 02  t 
xt   C1e  e 
 st 
e  
 1.29
 
The value of C1 can be again be determined from the initial value of velocity at t=0. Here, we shall
limit our self to study the behaviour of the solution only. Now from equation (1.29) we observe that
the displacement at any instant of time depends upon the value of s and  0 . There may be three
different solutions depending upon s  0 , s  0 and s   0 . We shall discuss each of the conditions
separately as given below.
Case-I: Over damped or Heavily damped oscillation for s  0 .

We consider   s  0  0 but lesser than s. Then equation (1.29) can be written as


2 2

xt   2C1e st


e  t
 e  t 
 Ae st sinh( t ) 1.30
2
Initially x(t) increases w.r.t. t for smaller values but it starts decreasing exponentially as time
increases. We will not observe any oscillations and the amplitude decreases exponentially and comes
to rest without completing any cycle. Such type of damping is called over damped or heavily damped.
This type of damping is useful if we want a system to come to rest without any when displaced from
its equilibrium position or released from its maximum displacement. For example if a spring block
system is made to oscillate in a highly viscous medium like glycerine, honey, we may expect it to
observe heavy damping. Some other examples like that of dead beat galvanometer, tuned dampers
used in car engines or skyscrapers which damp the vibrations or oscillations heavily. A very common
example of damping is a series RLC circuit. For R 2  4 L , it behaves as an over damped circuit.
C
The solution given by equation (1.30) and the charge variation in the RLC circuit is shown in Fig 1.6

Fig 1.5
Fig 1.6

Case-II: Critical damping for s  0 .


This is a special case of over damped motion. We shall consider the case of s   0 . Under such

situation, say   s  0  1. From equation (1.28) we can expand the exponential terms under
2 2

the square root and retain up to the first order as   1 . Hence, we can write,
  s 2 02 t  s 2 02  t 
xt   e st  C1e   C2 e  

 
xt   e C1e  t  C2 e  t 
 st

xt   e  st C1 1  t   C2 1  t   e  st C1  C2   C1  C2 t 


1.31
x(t )  e st P  Qt 
Where, C1  C 2  P and C1  C 2   Q , which can be determined from the initial conditions.
Equation (1.31) is product of an exponentially decaying term and an algebraic term. Initially when t is
very small the amplitude grows linearly but soon starts damping as the exponential term starts
dominating as time increases. Hence, with time the object comes to rest without completing a single
oscillation. The amplitude decreases exponentially as in the previous case. The object reaches its
equilibrium in the minimum time without completing an oscillation. Such type of damping is called
critical damping and the displacement w.r.t. time is shown in Fig. 1.7.

Fig. 1.7

The condition for critical damping can easily be found out from the following condition.
k
s 2   02  1.32
m
b
Using, 2 s  in the above equation we will get,
m
b2 k
2

4m m 1.33
 b  2 mk

Equation (1.33) gives the value of the damping constant for which we can achieve critical damping
for a system. A very useful practical application of critical damping is pointer type galvanometer.
Another common application of over damping is automatic door closer. A series RLC circuit for
R 2  4L , behaves as a critically damped circuit.
C

Case-III: Under damped for s   0

Under such situation the term s 2  02 becomes a complex number. This can be written as
s 2  02  i 02  s 2  i  1.34
Using equation (1.24) we can rewrite equation (1.28) as
xt   C1e si  t  C2 e si  t
 xt   e  st C1 cos  t  i sin  t   C2 cos  t  i sin  t 
 xt   e  st C1  C2 cos  t  iC1  C2 sin  t 
 xt   e  st A sin  cos  t  A cos  sin  t  1.35
Where, A sin   C1  C2  and A cos   C1  C2  1.36
Now equation (1.35) can be written as
xt   Ae st sin  t    1.37
Using the value of   we can rewrite equation (1.37) as

xt   Ae st sin 02  s 2 t    1.38
The solution as given above is now a product of an exponentially damping term and an oscillatory
sinusoidal function. This tells us that the amplitude decreases exponentially but it completes number
of oscillations before coming to rest. The angular frequency of oscillation also decreases compared to
free oscillations. Such types of oscillations are called underdamped oscillations. For example, the
oscillation of a spring-block system on a table attached to a fixed support, plucking of the guitar
string, tuning fork, etc. A series RLC circuit for R
2
 4L , behaves as an under damped circuit.
C

Fig. 1.8
A comparison of the oscillations for all the three damping cases along with the free oscillations is
shown in the figure below.

Fig. 1.9
CHAPTER 1: OSCILLATIONS
Hour: 4

Time Constant or Decay Time: In all damped harmonic oscillations the amplitude decays with time.
The time in which the amplitude decays to 1/eth of its initial amplitude is known as the decay time or
time constant. If A is the initial displacement at t=0 and τ is the time in which the amplitude falls to
1/e of A then it can be found as follows.
A
 Ae  s
e
b
A 
  Ae m 2
e
b

1
e e 2m

b
 1
2m
2m
  1.39
b
(c) Forced Oscillations
We have learnt that in damped harmonic oscillations the amplitude decreases exponentially, hence its
energy also dissipates with time. For example if we pull a block attached to a spring from its
equilibrium and allow it to oscillate then it will come to rest after some time. Instead if we move our
hand up and down with rhythm the block will keep on oscillating without coming to rest as shown in
the figure below.

Fig. 1.10
So, we can say that when an external periodic force is applied to a damped harmonic oscillator, then
the amplitude of oscillation can be maintained at constant amplitude without damping. The energy
loss due to dissipation is balanced by the energy input from the external force. A common example is
a swing where someone gives a push to the swing at regular intervals. Such types of oscillations are
called forced oscillations. The periodic external force is called the driving force. The different forces
that will be acting on the system are:
1. The restoring force, -kx, as explained in the previous section.
2. Damping force, -bvx, where b is the damping constant.
3. Periodic external force or driving force, F0 sin t , where F0 and  are the magnitude and
angular frequency of the driving force.
It is to be noted that the driving force is in opposite direction to that of the damping and restoring
force. Using Newton’s second law of motion we can write the equation of motion as

max  kx  bv x  F0 sin t


d 2x dx
m 2  kx  b  F0 sin t
dt dt
1.40

d 2 x b dx k F
 2   x  0 sin t 1.41
dt m dt m m

d 2x dx
 2  2s  02 x  F sin t 1.42
dt dt
b 2 k F
Where, 2s  , 0  and F  0 1.43
m m m
After some initial adjustment the oscillation achieves constant amplitude and oscillates like SHM.
Hence, we can consider a tentative solution similar to that of the SHM for equation (1.43).

Let x t  A sin t     1.44
Where A is the amplitude in the steady state condition. Now differentiating equation (1.44) w.r.t.
time.

 A cost   
dx
dt
d 2x
2
  2 A sin t   
dt
Substituting the above equations back in eqn. (1.44) and rearranging the RHS, we will get.
  2 A sin t     2sA cos t     02 A sin t     F sin t     
 
 02   2 A sin t     2sA cos t     F sin t    cos   cos t   sin  
1.45
Now comparing the coefficients of sin t    and cost    right and side and left side of
equation (1.45) we can write,

A 02   2  F cos   1.46
2sA  F sin  1.47

 
Squaring and adding equation (1.46) and (1.47) we will get
2
A2 02   2  4s 2 2  F 2 cos 2   sin 2   F 2  
F2
A 
 
2
2
0 2
  2  4s 2 2
Hence the amplitude of oscillation is given by
F
 
A 1.48
2
0  
2 2
 4s  2 2

The value of δ can be evaluated from equations (1.46) and (1.47), which will give us
2s
tan  
 2  2
0 
 2s 
   tan 1  2 2 
 0      1.49

From equations (1.48) and (1.49) we see that the amplitude and phase in force oscillations depends
mainly on the (i) driver’s frequency (ii) natural frequency of the oscillator and (iii) damping constant
for a given magnitude of the driving frequency. For a particular system the amplitude can be varied by
varying the frequency of the driver’s frequency. There are three possibilities depending upon the
value of the driver’s frequency.
Case-I: 0   , Very small driving frequency. The amplitude of oscillation will be
F F F
     4s  
A  
2 2   2  2 2 02
2
2 2 2
0
02 1  2   4s 4 
 0  0 

F0 F
A  0 1.50
m0 2
k
This means the amplitude depends upon the spring constant and the magnitude of the external force.
 
 
The phase,   tan
1  2s   tan 1  2s   tan 1 0  0
 2 2   2  1.51
 0 
0 1  2  
  0  
This means that the driving force and displacement are in phase.

Case-II: 0   , Driving frequency is same as the natural frequency of the system. This is also
called the resonant frequency. The amplitude of oscillation will be
F0
F F F m  F0
 
A   
4s  
1.52
2
0  
2 2
 4s  2 2 2 2 2s b  b
m
The amplitude depends upon the damping constant, magnitude of the external force and frequency of
the external force. In case of s  0 then A   , i.e. amplitude tends to infinity.

2s   2s  
  tan 1   tan 1   tan 1   
The phase,
 0   
2

2 
 0   2
1.53


This means the displacement lags behind the force by a phase of .
2
Case-III: 0   , Driving frequency is larger than the natural frequency of the system. The
amplitude of oscillation will be
F F F F0
 
A    1.54
2
0  
2 2
 4s  2 2   2
 2  0
 1
 4s 2 
2

  2 
2 m 2

  4
  

The amplitude depends upon the mass of the oscillator, magnitude of the external force and frequency
of the external force.
 
 
1  2s   tan 1  2s   tan 1  0  
The phase,   tan
 2 2    
 
  2 
0
 1
    
This means the displacement lags behind the force by a phase of  .

Resonance:

This is a special case as mentioned in case-II. The condition when the frequency of the external
periodic force matches with the natural frequency of the oscillator, then the amplitude of oscillation is
maximized. Under such situation there is maximum power transfer takes place from the driver to the
oscillator. This phenomenon is called resonance. A very important application of resonance is seen in
electronic circuits. For example a driven LCR circuit as shown in figure 1.11 with an a.c. source
1
exhibits resonance when the frequency of the a.c. signal is equal to . If damping becomes
2𝜋√𝐿𝐶
weaker than under resonance the amplitude becomes larger.

The variation of the amplitude for different damping is shown in the figure 1.12 below.

Fig 1.11
CHAPTER 1: OSCILLATIONS
Hour: 5

TWO COUPLED OSCILLATORS / NORMAL MODES

Overview and Motivation: This topic will help us to understand the some physics of the mechanism
of excitation of a mechanical wave in a medium. We will not yet observe waves, but this step is
important in its own right. The step is the coupling together of two oscillators via a spring that is
attached to both oscillating objects.

Two Coupled Oscillators:


Real systems are never isolated. From a more practical point of view it is important to study the
effects of their coupling with the neighbouring systems and environments.
As a step in this direction we consider systems of coupled oscillators, specifically point like massive
objects coupled by stiff springs (constant spring constant). We first consider a system of two similar
masses A and B. This can be generalize to n oscillators and finally consider the continuous limit n→∞
to understand the excitation and propagation of waves.
Consider two identical linear, simple harmonic uncoupled (independent) oscillators consisting of
equal point masses m, and connected to a rigid support by a spring in Fig 1.

Each is connected to a spring of constant k0, as shown. The other ends of these springs are fixed. Now
the two oscillators are connected by another spring of constant k as shown in Fig. 2(a). When the
whole system oscillates as a conjugate system, then it is called a coupled oscillation and their equation
of motion cannot be treated independently. The oscillation of such a two coupled oscillator is shown
in Fig. 2 (b). In this case we shall consider the oscillations along the horizontal direction, hence it is
called as longitudinal oscillations.

Let the equations of motion for the free oscillators before they are coupled be given by
mx1  F1  k 0 x1
mx2  F2  k 0 x2

where, x1, x2 are both positive in the same sense to the right. Now couple the two masses together by a
symmetric spring of constant k (i.e. force   kx ) with a natural length L. Assume all springs are
initially at their natural lengths and ignore friction. Consider the oscillations along the horizontal line
(x-axis). Now a force is applied to pull the left mass a distance x1 to the right and push the right mass
a distance x2 to the right. The coupling spring is therefore compressed by x1 and stretched by x2.
The resultant forces acting on the left mass are: a force  k 0 x1 towards the left due to the stretching of
the left spring; a force  kx1 to the left due to the compression of the coupling spring and a force
 kx2 to the right due to the stretching of the coupling spring. The resultant total force on the left mass
is therefore
F1  k0 x1  kx1  kx2 (2)
Likewise, the total force on the right mass is
F2  k0 x2  kx1  kx2 (3)
Thus the equations of motion for the two coupled masses are given by
mx1  F1  k0 x1  k x1  x2  (4)
mx2  F2  k 0 x2  k x1  x2  (5)
Important feature of coupled oscillators: the equation of motion for each oscillator includes
variables belonging to the neighboring oscillators, i.e. they are coupled ODEs. This tells us at once
that in general they do not undergo pure SHM.
To solve coupled ODEs, we need to find new variables such that each equation involves only one
dependent (new) variable. The method which we will follow here is the use of a new set of
coordinates called normal coordinates in which the equations will appear to be uncoupled oscillators.
Adding and subtracting Eqs. (4) and (5), we find
mx1  x2   k 0 x1  x2 
mx1  x2   k0  2k x1  x2 
This suggests choosing new coordinates
X  x1  x2 (6)
Y  x1  x2 (7)
to rewrite the above equations as
mX   k 0 X
mY  k  2k Y
0
Or, equivalently, as
k0
X  02 X , 02  (8)
m
k  2k
Y  12Y , 12  0 (9)
m
Eqs. (8) and (9) are both in a form describing simnple harmonic motion.

Thus we find that the motion of the coupled system can be described in terms of two coordinates X
and Y each of which has an equation which describes simple harmonic motion. These two motions are
completely uncoupled and independent although the moving parts of the system are not. One motion
can exist without the other.
If Y = 0, x1 = x2 at all times, so that the motion is completely described by Eq. (8). The angular
k
frequency given by 0  which is same as that of the either of the oscillator in isolation. The
m
effect of the spring constant is absent. This is because both the oscillators are always in phase and the
spring is at its natural length throughout the motion as shown in Fig. 3(a). If X = 0, then x1 = -x2 at all
times, so that the motion is completely described by Eq. (9). The angular frequency in this case is
k 0  2k
1  , which is greater than  0 . The oscillators are oscillating harmonically at the same
m
frequency 1 and are always out of phase as shown in Fig. 3(b). The spring is either compressed or
extended during the motion and coupling is effective

Normal Coordinates and Normal Modes


The significance of choosing X and Y to describe the motion of a coupled system is that these
parameters give a very simple illustration of what we call the normal coordinates. The normal
coordinates of a coupled system are the parameters in terms of which the equations of motion of the
system can be written as a set of linear differential equations with constant coefficients in which each
equation contains only one dependent variable. In the above example X  x1  x2 and
Y  x1  x2 are the normal coordinates and the equations that contain only one dependent variable
are Eqs. (8) and (9).
The simple harmonic motion associated with each normal coordinate is called a normal mode of the
coupled system. Each normal mode has its own characteristic frequency called the normal mode

frequency. The ratio of the displacements 


 x1  of the moving parts of the system remains constant

 x2 
for a normal mode and is called the shape or configuration of that mode. In the above example, there
are two normal modes, one associated with normal mode X and the other with Y. The shape of the first
mode is shown in Fig. 3(a). Both the masses are displaced, say to the right by the same amount so that
x1
 1. When they are released, each mass will execute SHM at the same frequency and will
x2
x1
always stay in phase and
x2 will remain constant equal to unity throughout the motion. This is
called the “in phase” mode of the system. The shape of the second normal mode is shown in Fig.
x1  1. When
3(b). The masses are displaced in opposite directions by the same amount so that
x2
k 0  2k
they are released, each mass will execute SHM at the same frequency 1  and will
m
x1
always be out of phase by 180o and
x2 will remain constant equal to -1 throughout the motion.
Both the masses will pass through their respective equilibrium positions simultaneously, except that
the amplitude of one is opposite in sign. Hence, in a given normal mode, both the moving parts
execute SHMs at the same frequency and the same phase constant.
In conclusion one can say that although the general motion of a coupled system is not periodic, yet if
the system is started in just the right way, the moving parts of the system will execute SHM. For a
system having two degrees of freedom, there are exactly two possible ways if starting the system so
that it will oscillate in one of the two possible normal modes. With which frequency the system will
oscillate depends on how we start the system in motion. In the case of two identical SHOs, if the
displacements are related as x1  x2 , then and only then, the system will oscillate with SHM at
frequency  0 , the ratio between the displacements remaining the same throughout the motion.
Similarly if the displacements are related as x1   x2 , then and only then, the system will oscillate
with SHM at frequency 1 , the ratio between the displacements remaining the same throughout the
motion. If the motion is started in any other way, there will be no permanent ratio between the
displacements of the two oscillators and the motion will not even be periodic. As mentioned earlier
these two special ways of motion of system are called its normal modes of vibration.

Reference: The Physics of Waves and Oscillations by N. K. Bajaj, Tata Mc Graw Hill.

You might also like