You are on page 1of 48

Foreword

Assalamualaikum warahmatullahi wabarakatuh,

Bismillahirrahmanirrahim,

Alhamdulillah, praise and gratitude to Allah swt, The cherisher and sustainer of the
worlds; God who has been giving His blessing and mercy to the writers to complete this
paper entitled "Harmonic Oscillator."

Writers also say thanks to all of the parties who already helped the writer in compled
thas paper. Certainly in writing it, this paper has many flaws, and therefore the writers beg
criticism and suggestions from other parties for the perfection of the writing of this paper. I
welcome any and look forward to criticism and suggestions from various parties.

This paper is made so that we can know about the Harmonic Oscillator. Hopefully
this paper useful for the readers, If there are many mistake in the writing of this paper the
writers apologize.

Wassalam.

Pekanbaru, 01 November 2012

Writers
Content

FOREWORD

CONTENT

HARMONIC OSCILLATORS

1. Introduction
2. Linear and non linear oscillations
3. Linear harmonic oscillators
4. Damped Harmonic oscillators
5. Quality factor
6. Forced harmonic oscillators (driven oscillators)
7. Amplitude resonance
8. Energy resonance
9. Rate of energy dissipation

CONCLUSION

REFERENCES
Harmonic Oscillators

3.1 INTRODUCTION

Consider a system in static or dynamic stable equilibrium. When such a system is


displaced slightly from its equilibrium position, the resulting oscillatory motion is called
harmonic motion. Such motions occur frequently in natme and are investigated, both from a
practical as well as a theoretical point of view, in physics and engineering. A few examples of
such motions are elastic springs, bending beams, pendula, vibrating strings, resonance of air
cavities, and the motion of charges in certain electrical circuits and cavities.

To start, we shall study the motion of a linear harmonic oscillator (motion resulting
from a small displacement of a system from its equilibrium) in one dimension. Unavoidable
inclusion of friction in such motion leads to the investigation of a damped harmonic
oscillator. To maitain oscillatory motion in the presence of friction, some external force must
be applied. Such an oscillating system is called a forced or driven oscillator.

When the displacement of the system from equilibrium is large, the system is no
longer linear. Such oscillating systems are called nonlinear. We divide our study into two
parts. This chapter is mainly devoted to the study of a linear system including damped and
forced harmonic oscillators. The study of nonlinear oscillations, the electrical equivalent of
mechanical oscillators and multidimensional oscillators, will be investigated in Chapter 4. It
may be pointed out that. in general, oscillations of systems occurring in nature are nonlinear.
But their approximation to linear systems allows us to use strong analytical techniques
developed for this purpose.

3.2 LINEAR AND NONLINEAR OSCILLATIONS

Consider a particle of mass tn moving in an arbitrary conservative force field for


which the potential energy V(x) of the particle as a function of its displacement is represented
by a heavy curve, as shown in Fig. 3.1. For a conservative force field, the total energy E of
the particle is

E=K + V = constant (3.1)

If is the velocity of the particle,

(3.2)

which when solved for yields


(3.3)

If E = E0, as shown in Fig. 3.1, then Eo - V(x) = 0 and , = 0; that is, the particle
stays at rest in a stable equilibrium at x --- x0. Let us consider the case in which the particle
energy E1 is slightly greater than E0. For x < xj and x > x2, will be imaginary; hence the
particle cannot exist in these regions. Thus a particle with energy E1 is constrained to move in
a potential well (or valley) between x1 and x2. The particle moving to the right is reflected
back when it reaches .r2, and when traveling to the left it is reflected again at x1. The points x1
and x2 are called turning points, and the velocity of the particle at these points is zero. These
points are obtained by solving Et - V(x) = 0. In between these points, the velocity of rn
changes continuously depending on the value of V(x). Hence a particle in a potential well
moves back and forth and oscillates between x1 and x2 when its energy is greater than E0.

The position x(t.) of a particle moving in potential well can be found by integrating

Eq. (3.3); that is,

(3.4)

Figure 3.1 A particle of mass m and energy E is moving in an arbitrary potential energy function V(x) shown by
the solid heavy curve. The dotted curve is the parabolic potential approximation of the arbitrary potential.

while the time period T of one complete oscillation is given by


(3.5)

Equations (3.4) and (3.5) cannot be solved unless we know the form of the potemial
function V(x). The motion of the particle can be limited to the region in the neighborhood of
x0, and for small displacements, such as these, it is possible to approximate the arbitrary
potential function V(x) by a parabolic potential shown by the dotted curve in Fig. 3.1. As we
shall show later, this potential may be written as V(x) = 1/2k(x - x0) 2, where k is a constant,
thereby enabling us to solve Eqs. (3.4) and (3.5).

Suppose a particle is oscillating aborn a point of stable equilibrium x0, where the
minimum potential is V(xo) at x = Xo. Let us expand the potential function V(x) in a Taylor
series about the point x0.

(3.6)

We limit our discussion to small displacements in symmetrical potentials. The term


X(xo) is a constan and can be dropped without affecting the results. Also, since xo is a point
of minimum, for stable equilibrium in a symmetrical potential, the odd terms must be zero.
[Note that if the expression resulting from the exposion of F(x) were used the even terms
would be zero.] Therefore,

(3.7a)

while

(3.7b)

Define

(3.8)

(3.9)
(3.10)

Then the potential function may be written as

(3.11)

Let us assume that the origin is located at the equilibrium point so that x0 = 0 and x' = x,
and by neglecting the higher-order terms in Eq. (3.11), we get

(3.12)

Furthermore, since the motion of the particle is in a conservative force field, using the
definition

and substituting for V(x) from Eq. (3.12), we may write

(3,13)

Linear Oscillations

In the first approximation, we can neglect all terms except the first in Eqs. (3.12) and
(3.13) so that

(3.14)

(3.15)

where

(3,16)

since (d2V/dx2)0 is always positive, k will be positive also. Hence a force F(x) = -kx is
always directed toward the center and proportional to x. Such a force is called a linear
restoring force.
the potential corresponding to such a force is parabolic as given by Eq. (3.14) and
shown by the dotted curves in Figs. 3.1 and 3.2 for different values of k. The corresponding
linear forces are shown by the dotted lines.

Physical systems involving springs, pendula, and elastic deformation are described by
Eqs. (3.14) and (3.15) and are said to obey Hooke's law. This is true only if the displacements
are small and we remain within elastic limits, as shown in Fig. 3.2(a). Moreover, the results
obtained are still approximate. We shall spend most of the time discussing linear oscillations
resulting from approximate linear systems. k has been given several names, but is usually
called the spring constant or stifflyess constant. k is defined as the force per unit length with
units of Newtons per meter (N/m). I/k is called the compliance of the spring.

Nonlinear Oscillations

If the displacement of the system from stable equilibrium is not small (or if we
definitely want higher order improvements in the linear approximation), we cannot drop the
second term in Eqs. (3.12) and (3.13). Thus, according to Eq. (3.13), the force is no longer
linear because of the presence of a x3 term, while the potential is no longer parabolic because
of the presence of ax4 term. Different forms of forces and potentials are illustrated in Fig. 3.2
for systems with large displacements (hence no longer linear).

Let us further consider Eq. (3.13) for a nonlinear system; that is,

(3.13)

We must remember that is a very small quantity as compared to k, but its magnitude
and sign affect the linear term - kx, hence the resulting force F(x). If < 0, the magnitude of
the force F(x) will be less than the linear force kx alone and the system is said to be soft. On
the other hand, if > 0, the magnitude of the force F(x) is greater than the linear force kx
alone and the system is said to be hard. The forces and potentials of such systems are shown
in Fig. 3.2.
Figure 3.2 The plots of F(x) versus x and V(x) versus x for a variety of systems. The magnitude and sign
of determines whether the system is hard or soft. For soft systems, < 0, while for hard systems, > 0.

3.3 LINEAR HARMONIC OSCILLATOR

Consider the prototype of a linear or simple harmonic oscillator shown in Fig. 3.3. It
consists of a mass m tied to a spring having a force constant k. The spring-mass system
oscillates in one dimension along the X-axis on a horizontal frictionless surface. The system
obeys Hooke's law;

Figure 3.3 Prototype of a linear


harmonic oscillator showing the
maximum and minimum values of x, v,
a and F.

x=-A x=0 x= +A

v=0 v=Vmaks v=0


a =+amax a=0 a= -amax

F=+kA F=0 F= -kA

Hence the system is linear. Measuring the displacement x from the equilibrium
position, the potential energy V(x) is

(3.17)

While the restoring force F(x) is

(3.18)

From Newton's second law, F(x) = rn(d2x/dt2); therefore,

(3.19)

Or (3.20)

Where (3.21)

is a constant and is called the free natural angular frequency (or free oscillation
frequency) of the system.

Our aim is to solve Eq. (3.20) for x(t). Before we do this, it must be pointed out that an
equation of this form is frequently encountered both in physics and engineering; hence its
solution must be thoroughly investigated. Equation (3.20) is a second-order, linear,
homogeneous differential equation. The highest derivative that occurs in a differential
equation is called its order, while a differential equation is linear if it does not contain terms
higher than the first degree in the dependent variable [x is the dependent variable in Eq.
(3.20)] and its derivative. Also, Eqs (3.20) is homogeneous because it does not contain terms
other than the dependent variable and its derivatives. Thus the most general form of a
differential equation of the nth order, linear and inhomogeneous, is

(3.22)

If b(t) = 0, the equation is homogeneous. The coefficients Cn Cn-1,..., C1 are constants


that may or may not be independent of time, but we assume them to be independent of time.

We shall be dealing with second-order differential equations. We summarize next some


properties of such equations, which will be helpful.
1. The general solution of any second-order differential equation depends on only
two arbitrary constants. Suppose we choose C1 and C2 to be the arbitrary
constants; then

x = x(t; C1, C2)

C1 and C2 are arbitrary because any values of C1 and C2 will satisfy a second-
order differential equation.
2. If x1(t) is any solution of a linear homogeneous differential equation, then
Cx1(t) is also a solution, where C is an arbitrary constant.
3. If x1(t) and x2(t) are solutions of a linear homogeneous differential equation,
then x1(t) + x2(t) or any other linear combination C1x1(t) + C2x2(t) is also a
solution.

Let us now go back to Eq. (3.20) and try to find its solutions. To start, we may write it
as

Now, multiplying both sides by

and integrating, we get

where C is a constant. When thus

which, after separating the variables, may be written as

Integrating this equation, we get

where is a constant, called the initial phase or phase constant. We may write this
equation as
(3.24)

Thus the solution of Eq. (3.20), which is a second-order differential equation, is given
by Eq. (3.24) and contains two arbitrary constants A and to be determined from the initial
conditions. Equation (3.24) is a solution of a linear oscillator or harmonic oscillator. The
graph of x versus t is shown in Fig. 3.4. x is called the displacement,' the maximum
displacement is called the amplitude of the oscillator and is equal to A. The quantity to is
called the angular frequency and is given by Eq. (3.21). Also, where vo is called
the frequency of the oscillator. the time period To of the oscillator is the time required to
complete one oscillation. Thus, in time t = To in Eq. (3.24), t increases by 2 that is,

or (3.25a)

or (3.25b)

The expressions for velocity and acceleration may be obtained by differentiating Eq.
(3.24); that is,

(3.26a)

(3.26b)

The plots of x, v, and a versus t are shown in Fig. 3.4.

The solution given by Eq. (3.24) may be written in a different form as follows:

Substituting

(3.27a)

And (3.27b)

we get

(3.28)
which is another form of the solution for a second-order differential equation. Squaring
and adding Eqs. (3.27a) and (3,27b), we get

(3.29a)

and dividing Eq. (3.27b) by (3.27a), we get

(3,29b)

Equations (3.27) and (3.29) give the relations between the constants A, , and B, C.
We may differentiate Eqs. (3.28) to get expressions for velocity and acceleration.

Equation (3.20) or any other second-order differential equation with constant


coefficients may be solved by a trial solution of an exponential form, as explained next. Let
the trial solution be

(3.30)

where A is a constant to be determined. Substituting Eq. (3.30) into (3.20),

(3.31)

This is called the characteristic, initial, or auxiliary equation. Thus

(3.32)

where Thus, with two roots . the general solution is,


using A+ and A- as two constants,

(3.33)

All three solutions, Eqs. (3.24), (3.28), and (3.33), are equivalent, and any one can be
derived from the other two. Each contains two constants. Equation (3.33) may be reduced to
the other two by using the Euler formulas

(3.34)

It may be well to remember that in a solution given by Eq, (3.33), the constants A+ and
A- are complex quantities and sometime inconvenient to use. Still another way of writing this
general solution is (with A and qb as two constants)
(3.35)

where both the real and the imaginary parts of this equation are solutions of a general
differential equation.

Energy of the Simple Harmonic Oscillator

For a simple harmonic oscillator, the displacement is

while the velocity is

and the maximum value of the velocity vo is

Hence the kinetic energy K of oscillator is

(3.36)

Where Ko is the kinetic maximun energy given by

(3.37)

The potential energy of the system is equal to the work done by the applied force F a=-
F=-(-kx)=kx in displacing the system from x=0 to x=x. Thus

(3.38)

Subtituting for x

(3.39)

where Vo is the maximum potential energy when x = A; that is,

(3.40)
Thus the total energy E, which is always constant whenever there is a conservative
force field, is

(3.41)

This equation can be solved for x(0 and provides more information about the problem
under consideration, as we shall show next. From Eq. (3.41),

(3.42a)

or (3.42b)

we get the solution for x to be

(3.43a)

or (3.43b)

where 1 and 2 are constants, while the amplitude A is given by

(3.44)

This relation tells us that x can vary between +A and -A, that is, between

This has to be true because only then will x be real as given by Eqs. (3.42). The value
of x then must lie between two limits that are determined by the energy E and the spring
constant k.

To find the average values of V and K over one complete time period, we use the
following general expression for the average value of quantityf (t):

(3.45)

That is,
(3.46)

and similarly

(3.47)

That is,

(3.48)

If, instead of time averages, we calculate space averages over one complete time
period, we get (see Problem 3.1)

(3.49)

and

(3.50)

3.4 DAMPED HARMONIC OSCILLATOR

Theoretically, a linear or a simple harmonic oscillator once set into motion will
continue oscillating forever. Such oscillations are called free oscillations. In practice,
however, in any physical situation there are dissipative or damping forces, and the oscillating
system will lose energy with time. Thus the oscillating system is damped and eventually
comes to rest. The differential equation for a linear oscillator given by Eq. (3.20) must be
modified to include the effect of damping.

Once again we consider a mass m tied to a spring, as shown in Fig. 3.5, as a prototype
and restrict its motion to one dimension. As the mass moves in a fluid, air or liquid, the
frictional force is the viscous force that produces the damping. As long as the speed of the
mass is small so as not to cause turbulence, the frictional force or damping force Fa may be
assumed to be proportional to the velocity. That is,

(3.51)

where b must be a positive constant. The net force Fet due to forces acting on mass m
as shown in Fig. 3.5 is
(3.52)

Using Newton's second law and substituting Fnet = m in Eq. (3.52), we get

(3.53)

which is a second-order differential equation for a damped harmonic oscillator. To


solve this equation, we divide both sides by rn and substitute

(3.54a)

(3.54b)

And to obtain

(3.55)

As before, let us try an exponential solution of the form

Figure 3.5 Forces acting on a prototype of a damped harmonic oscillator.

and substituting in Eq.(3.55), we get

Since 0, we must have

(3.56)

This auxiliary equation has the roots

(3.57a)

(3.57b)
The following three cases of this solution are of special interest and will be discussed in
some details

In case a Underdumped Oscillations, , for this case, it is convenien to make


substitution:

(3.59)

Thus the exponentials inside the parentheses in Eq. (3.58) are imaginary, and we may
write this equation as

(3.60)

which is a solution of an underdamped oscillator. Using the relation


, we may write Eq. (3.60) as

Substituting i(A1 - A2) = B and (A1 + A2) = C, we obtain an alternative solution:

(3.61)

This may still be written in a slightly different form by making the following
substitutions in Eq. (3.61).

Thus we obtain

(3.62)

Of the three solutions given by Eqs. (3.60), (3.61), and (3.62), we shall concentrate on
Eq. (3.62). It may be pointed out that the constants A1 and2 in Eq. (3.60) are complex
quantities, while B and C in Eq. (3.61) and A and in Eq. (3.62) are all real quantifies. The
solution given by Eq. (3.62) indicates that for a damped oscillator the motion is osciliatory,
but the amplitude of the oscillations decays exponentially, as shown in Fig. 3.6. The natural
angular frequency, oh, or the frequency of the damped oscillator is always less than the free
oscillation frequency mo. The natural frequency is not a frequency in the true sense of the
word because the oscillator never passes through the same point twice with the same velocity;
that is, the motion is not periodic. But if is very small, then (as shown later), and
we can call the "frequency." If T is small, we can expand Eq. (3.59) (using the binomial
expansion) as

(3.63a)

(3.63b)

(3.63c)

According to Eq. (3.62) the case for = 0 is shown in Fig. 3.6. Equation (3.62) states
(and this is demonstrated in Fig. 3.6) that the maximum amplitude of the oscillations
decreases exponentially with time because of the factor and lies between the two curves
given by

(3.64)

where Aet (t) is the envelope that limits the displacement of the oscillations. For
comparison the x0 graph represents the oscillations for a free oscillator, that is, for = 0. The
graphs with different dampings are shown in Fig. 3.6. Plots x4 and x5 represent the envelope
of the damped motion that (due to the presence of the cosine term) touches the
envelope at that is, at times w , where n is an integer.
The period of the damped oscillation is . Since , that is, the damped
frequency is smaller than the free

For undamped and underdamped oscillators,

Below is a graph of x versus t for the following degrees of freedom:


xO Undamped,

x1 Lightly damped,

x2 Moderately damped,

x3 Heavily damped,

x4 and x5 are two envelopes for

plot x1 showing that the amplitude


0f the oscillations decays

exponentially.

a. What equations (other than cosine and sine) can be used to obtain the same graph?
b. How do the graphs of the four functions differ in their time periods, amplitude, and
frequencies? Explain.
frequency, the period T1 of the damped oscillation is longer than the free period To. This is
clear
from the comparison of the values of the plots in Fig. 3.6. It is worth noting that the points
of
the curve that touch the envelope are Tj/2 (= /) apart, but the maxima and the minima of
the
curve, even though separated by T1/2, do not coincide with the points of maxima and minima
of
the undamped motion curve.
Figure 3.6 shows the plots of x(t) versus t for different degrees of damping. The ratio /
determines the essential feature of these plots. If /1, the amplitude envelope Ae(t) given
by Eq. (3.64) changes very slowly with time, while the cosine term in x(t) makes several zero
crossings. Such a system is said to be lightly damped. On the other hand, if / >> 1, the
sys-
tem is said to be heavily damped because Ae(t) will decrease very rapidly and goes to zero,
while
the cosine term makes only a few zero crossing. In either case, the ratio of the two successive
maxima is given by

where t=tm is the time when the first maximum occurs and t2 =tm + T1 is the time when the

next maximum occurs, T1 being the time period of the damped oscillation. The quantity exp

(T) is called the decrement of motion, while its logarithm, T, is called the logarithmic
decre-

ment, ; that is,

Case (b) Critically Damped, o2 = 2 For this case, the two roots 1 and 2 given by Eqs.
(3.57) are equal, that is,

and the general solution given by Eq. (3.58) takes the form

where (A1 + A2) = B1 = constant. This is not a general solution because it contains only one

constant. We can show that in such cases, if e- is a solution,


is also a solution. Substituting in the differential equation

we get

Since o = the equation is satisfied, and te- is also a solution. Thus, for a critically damped
case, the general solution is a linear combination of e- and te- ; that is,

where B1 and B2 are constants to be determined by the initial conditions.


Figure 3.7 represents three cases of interest (for a critically damped oscillator) resulting from
the solution given by Eq. (3.68). If we differentiate x with respect to t and equate to zero, we
get the positions of maxima in the plot of x versus t. Thus dx/dt = 0 gives

If

the curves for x versus t have maxima at t = 0 and t > 0, as in Fig. 3.7.

If

the curve does not have a maximum for t > 0, as demonstrated in Fig. 3.7.
Critical damping plays a very important role in the design of such instruments as
galvanometers, hydraulic springs, and pointer reading meters. It is desired that the system
attain an equilibrium position rapidly and smoothly in the presence of frictional damping.

Case (c) Overdamped, o2 2 : If the damping increases such that o2 2, then the two
roots 1 and 2 are real. If we represent

the general solution given by Eq. (3.58) takes the form


Note that 2 is no longer a frequency because the motion is no longer oscillatory. The
exponents are real, and both terms on the right decay exponentially, one faster than the other.
As shown in
Fig. 3.8, for the case x(0) 0, (0) 0, the displacement goes to zero asymptotically, but not
as rapidly as in the case of a critically damped system. For the case when x(0) > 0 or < 0,
Fig. 3.8, shows how x(t) varies with time. For x(0) > 0, x(t) reaches maximum for t > 0. For
x(0) < 0, but small, x(t) has no maximum for t > 0. For x(0) 0, but sufficiently large, x(t)
has a maximum for t > 0, as shown.

For a critically damped oscillator (not oscillatory)

Plots are for the same values of B 1 and B2 but

different values of ?. For the present situation,


B2/B 1 = 0.5, while 7 = b/2m. The graphs for
three different values of ? and b are below.

(a) What is the effect of increasing the


value of 7- Explain

(b) What causes the changes in the values


of B I and B27 Explain

(c) How will a change in the value of b


and m change the values of B and ?

(d) What is the significance of the following


maximum values in the three cases?
(e) How does the graph of a critically damped oscillator differ from that of an undamped
oscillator in terms of frequency of oscillations and amplitude?

For an overdamped oscillator


(not oscillatory)

Below is the graph of x versus t


for three different initial conditions.

(a) How do different values of the


coefficiems B, C, and D affect
different plots?

(b) What causes the change in the

x alue of to and how does it affect


the graph?

(c) Explain why the two values of


X1 and x3 have opposite signs.
max(xl) = 6.382 max(x3) = 2
min(xl) =0.21 min(x3) =-1.683

Consider a damped harmonic oscillator and graph its motion using the following data and the
initial conditions, where m = mass, k = spring constant, and b = damping constam

Solution
Approach using Eq. (3.63):
(a) Underdamped oscillator with b = 0.1

A and are determined


from initial conditions.

the alternative approach


below.
(b) Critically damped oscillator with b = 1.0

(c) Overdamped oscillator with b = 5.0

Alternate approach
using Eqs. (3.60), (3.61), and (3.62):
(A) Underdamped oscillator with = 1.0

Evaluate the constants


By using the initial
conditions and Eq. (3.61).

(B) Critically damped oscillator with ? = 10

(C) Overdamped Oscillator with ? = 50

It is clear from these two graphs that both approaches give the same results. The third
treatment mentioned in Section 3.4 also will yield the same results.
Energy Considerations

The total energy E(O of a damped harmonic system at any time t is given by

where E(0) is the total energy at time t = 0 and Wf s is the work done by friction in the time
interval 0 to t. Assuming the dissipative frictional forcef = - bx = - bv, we can calculate Wf as
follows:

Thus the rate of energy loss by friction may be written as

which is negative and represents the rate at which energy is being dissipated into heat. Since
Wf < 0, Et , continuously decreases with time and may be calculated in the following manner:

From Eq. (3.62),

Let us assume that the system is lightly damped so that /1 1, and neglecting the second
Term on the right in the preceding expression for , we can substitute for x and in Eq.
(3.73):

Since we assumed light damping, we may write 2 02= k/m; hence this equation takes the
form

while the initial energy of the system is obtained by substituting t = 0 in Eq. (3.74); that is,

Thus
That is, the energy decreases (or decays) exponentially at a much faster rate (e-2y) than the
rate at which the amplitude decreases or decays (e-y).
The time in which E decreases to 1/e ( = 0.368) of its initial value is called the
characteristic time or decay constant and may be evaluated by substituting E(t) = Eo/e and t =
in Eq. (3.76):

If is very small, -->, and if is very large, --> 0.


Also, using Eq. (3.76), we may write the logarithmic derivative of E as

(1/E)(dE/dt) represents the fractional rate of decrease in energy. Since the rate of energy loss
is proportional to the square of the velocity [Eq. (3.72b)], the loss in energy is not uniform.
dE/dt will be maximum when is maximum (near the equilibrium), and it will drop to zero
when is zero near maximum amplitude. The plots of E and dE/dt are shown in Fig. 3.9.

3.5 QUALITY FACTOR

The quality factor Q , or simply Q value, is a frequently used term in mechanical oscillatory
systems, as well as electrical oscillatory systems. Q is a dimensionless quantity and
represents the degree of damping of an oscillator. The quality factor is defined as 2 times the
ratio of the energy stored to the average energy loss per period. Thus

If P is defined as the power loss or the rate at which the energy is dissipated, and the time
period of oscillation T 1= 2/1, we can write the denominator as PT1 = P2/1 , and hence
Eq. (3.79) as

But lt 1/1 is the time of motion for 1 radian. Thus

As should be clear, for the lightly damped oscillator, Q will be very large, while for a heavily
damped oscillator, it will be very small. We can calculate the Q value of the lightly damped
oscillator as follows, The energy of the oscillator and the rate at which it loses energy are
given by [Eqs. (3.76)
and (3.78)]

Below are the graphs of energy


E and the rate of energy loss

ET = dE/dt versus time t for a


damped oscillator

For clarity, instead of

graphing ET, we have graphed


ET.5

a). Why is E positive

and ET negative?
b). Is the rate of loss of

energy ET in phase with


or out of phase with
energy E? What is the
significance of this?
And

Thus the energy dissipated in time At will be

If At is the time for 1 radian of oscillation, At = 1/1 ; hence

For light damping, 1 0 ; hence

If is small, Q will be large, and vice versa. Ordinary mechanical systems, such as
loudspeakers and rubber bands, are heavily damped and may have Q values from 5 to 100.
On the other hand, systems such as tuning forks and violin strings may have a Q value as
high as 1000. A typical microwave cavity resonator has a Q value of about 104. Systems with
extremely light damping are excited atoms (Q 107), excited nuclei (Q 1012), and gas lasers
(Q 1014).

3.6 Forced Harmonic Oscillator (Driven Oscillator)

A free oscillator will oscillate forever. But, in reality, every system has some damping
present (the energy is dissipated, say in the form of heat) and the system will eventually stop
oscillating. To maintain the oscillations, energy from an external source must be supplied at a
rate equal to the energy dissipated by the oscillator in the damping medium. Such motion in
which energy is supplied externally is called forced oscillations or driven oscillations, while
the system is called a forced oscillator or a driven oscillator. If the system is acted on by a
driving force Fd then the net force, Fnet , acting on the system is given by

Where

and from Newton's second law Fnet = m. Equation (3.85) cannot be solved unless we know
the form of the applied force Fd. Since we have been limiting our discussion to linear
oscillators, it is easier if we assume that the driving force has a sinusoidal form given by

We have good reasons to assume this form for the driving force. First, many actual situations
involve just such a force, as, for example, the response of a bound electron when
electromagnetic waves are incident on it, that is, in the scattering of light from bound
electrons. Second, any periodic function of time can be represented as a sum of several
harmonics (or sinusoidal) terms. Using the techniques of Fourier series, one can solve for the
motion of the system under any periodic driving force (as discussed in Chapter 4.).
We may combine these equations and write the following equation that describes the motion
of a driven harmonic oscillator:

This is an inhomogeneous, second-order, linear differential equation. The solution of Eq.


(3.87) is given by the sum of two parts according to the following theorem:

If xi(t) is a particular solution of an inhomogeneous differential equation and the


complementary function x(t) is the solution of the corresponding homogeneous
equation [that is, Eq. (3.87) with the right side equal to zero], then x(t) = xi(t) +
xh(t) is also a solution of the inhomogeneous differential equation.

Thus the general solution of Eq. (3.87) is

where xh is the solution of the homogeneous equation

From Section 3.4, the general solution of this homogeneous equation is given by any one of
the three forms, Eq. (3.60), (3.61), or (3.62); hence

Since the oscillations of a damped oscillator eventually decay to zero, the xh part of the
solution is called the transient term. After a certain time, the xh part of the solution is of no
consequence; hence, for a steady-state solution we must concentrate on finding the particular
solution xi(t).

According to Eq. (3.87), the applied force varies sinusoidally, so we expect the
resulting qeady-state solution xi(t) to vary sinusoidally. A solution of the form x = A cos t
would have been perfectly acceptable if the left side of the equation did not have an term.
To take care of this situation, we must have a solution of the form

Let us assume a solution of the form

To calculate A and , we substitute for xi in Eq. (3.87), and after setting 0 = O, we get

Rearranging,

For this to hold for all values of t, the coefficients of the cos t and sin t terms on each
sidsast be separately equal. That is,

From Eq. (3.93), we obtain an expression for the phase angle to be

Using the usual notation k/m = 02 and = b/2m, we get

From whicw we obtain

And

If we substitute these in Eq. (3.92), we get


Thus a particular solution of the inhomogeneous equation is
f0
m
xi (t) = cos(t
2
(20 2 ) +4y2 2

). 0.00

Where = tan1
2y
0.01
20 2

[A slightly different procedure for obtaining the general solution is more convenient when
the driving force Fd given by Eq. is written in exponential form as

Fd =
i(t0 )
F0 e 0.02

and we can obtain the same results.]

Using the solution given by Eq together with the homogeneous solution xh , given

by Eq. 0.00, we get the general solution

x = xh + xi
f0
m
= Ah ey cos(i t + h ) + cos(t ).
2
(20 2 ) +4y2 2

0.03

As required, this solution contains two arbitrary constants (of integration) Ah , and 0 , while
is not a constant and is given by Eq. (0.01 ). The first part of the solution oscillates with a
natural frequency 0 . Because of the damping, the oscillations die out for large values of
time, that is, for t >>1/y The homogeneous solution xh is called the transient soluion, while
the particular solution xi is the steady-state solution. The general solution x will be
independent of the influence of the initial conditions except in the beginning when the
transient term is still contributing. Figure 3.10 illustrates this for two special cases: (a) for
< 1 that is, the driving frequency is less than the natural frequency; (b) for > 1 , that
is, the driving frequency is greater than the natural frequency. For both cases, the plots of the
homogeneous solution xh versus t as well as plots of the particular solution xi versus t are
shown. The resultant of these two, that is, the plots of x = xh + xi versus t are also shown. As
is clear from these plots, the transient solution xh is effective only in the beginning and
decays to zero as time passes, while the steadystate solution remains constant with time. Thus
the transient solution effects the general solution only in the beginning. Furthermore, if <
1 , the transient term xh causes distortion of the resulting sinusoidal waveform as shown in
Fig. 3.10(a). On the other hand, if > 1 , the transient term xh , instead of causing
distortion, has the effect of modulating the oscillations due to the force function as shown in
Fig. 3.10(b). Of course, in both cases, after the transient term has died out, the oscillations are
governed by the force function. In addition to the relative values of and 1 , , initial
conditions will also affect the detailed motion, but only in the beginning. It is important to
note that the transient terms play an important role in electrical circuits. In designing such
circuits, it is necessary to avoid peak voltages and currents when initially the circuits are
closed.

Since for t 1/y , x x1 we shall concentrate on the discussion of the steady-state solution,
that is, the particular solution xi given by Eqs. (0.00) and (0.01). This solution is independent
of the initial conditions.

Below are two special cases of the influence of the transient term on the steady-state solution.

1. When the driving frequency (=10) is greater than the natural frequency 1 (2.828),
this leads to distortion, as illustrated.
n
N = 30 , n = 0. . , t n =
5
A = 1 , = 1, = 0, 0 = 3

M = 1, = 10, F0 = 50
1
2.
i = tan (2 2 ) 1 = (|20 2 |)2
0

i = 0,21 1 = 2.828
0

= . . cos(1. + ) =
1 . cos(. + )
1 2
|(20 2 )2 +4.2 .2 |

= +
(b) When the driving frequency (=3) is less than natural frequency 1(9.95), this leads to
modulation, as illustrated.
n
N = 30 , n = 0. . , t n =
10
A = 1 , = 1, = 0, 0 = 10

M = 1, = 3, F0 = 50
1
2.
i = tan (2 2 ) 1 = (|20 2 |)2
0

i = 0.066 1 = 9.95
0

= . . cos(1. + ) =
1 . cos(. + )
1 2
2
|(0 2 )2 +4.2 .2 |

= +

If the driving frequency is equal to the natural frequency, what effect it will have on the
amplitude?

Example 0.00
0
Consider a damped oscillator, for which = 4 . Acted on by a driving force =
0 cos . Find the general solution().

Solution

The second-order differential equation describing the driven oscillator is

+ + = 0 cos
Or + 2 + 20 =
0
cos ()

0
Where = 4 and
1
12 = (20 2 )2
0,970 .. ()

The transient (or the homogeneous) solution

() =
(1 cos 1 + 2 cos 1 ).. ()

takes the form

() =
(1 cos 0,97 + 2 cos 0,97) ()

Let us assume the particular solution for the applied force = 0 cos to be

() =
1 cos + 2 sin . ()

() =
1 sin + 2 cos .. ()

() = 2 1 cos 2 2 sin ..
()

Substituting these three equations in Eq. (i) and rearranging gives

0
(1 sin + 2 cos ) cos + (1 sin + 2 cos ) = cos

Equating the coefficients gives

(20 2 )1 + 22 = 0

21 + (20 2 )2 = 0

Solving these equations for B and B 2 in terms of F 0 gives

f0 (20 2 )
1 = 2
m[(20 2 ) +4y2 2 ]

2F
2 = 2
m[(20 2 ) +4y2 2 ]

Hence the general solution is given by


() = () + ()

= (1 cos 0,97 + 2 cos 0,97)+1 cos + 2 sin

where 1and 2are given by Eqs. (xi) and (xii), while 1 and 2 are to be evaluated using the
initial conditions in Eq. (xiii) by the usual procedure (see Example 3.1).

EXERCISE 3.3 Complete the example for the driving force = 0 sin with initial
conditions t = 0,

u 0) = 0, and (0) = 0, that is, calculate 1 and 2 . Also graph , , and = + .

3.7 AMPLITUDE RESONANCE

The amplitude A and the phase angle 4' of steady-state motion according to Eqs. (3.98) and
f0
m
A= cos(t )
2
(20 2 ) +4y2 2

................................................................................................ 0.04
2y
= tan1 2 2....................................................................................................................
0
0.05

For a fixed value of 0 , the variations in A and with the driving frequency for different
values of are shown in Fig. 3.11. As illustrated, the behavior of these quantities strongly

depends on the ratio As stated earlier , represents the phase difference between the
0
driving force F and the resulting motion x; that is, it represents a delay between the action and
the response. As shown in Fig. 3.11, this phase lag, which is , = 0 when = 0, increases to
( = /2 for = 0 and reaches , = as --> ; that is, at very high frequencies the
oscillations of the system are 1800 out of phase with the driving force. It is interesting to note
that as --> 0 the phase change occurs more and more rapidly, and in the extreme case when
= 0, the phase changes suddenly from 0 to , at = 0 From Fig. 3.11, it is clear that,
depending on the values of there is a certain driving frequency at which the amplitude A
has a maximum value. The frequency at which the amplitude is maximum is called the
amplitude resonance frequency for. This frequency , may be calculated from Eq, (3.103) by
setting

] =
=0

0......................................................................................................................................0.06

Upon solving the resulting equation, we get


= r = (20
1
2 2 )2 .............................................................................................................0.07

which states that as the damping coefficient decreases the resonance frequency increases, and

in the limit as --> 0, ->0 , decreases the natural frequency of a free oscillator. In the
case

of extremely small damping, we can express the fight side of Eq. by using the binomial

theorem; that is,


1 22
= r (2 +)
20

2
Or = r ......................................................................0.07
0

Equations (0.06) and (0.07) for a driven oscillator may be compared with the case of a
damped

oscillator discussed previously; that is,


1
= (20 2 )2

a) The graph below shows amplitude resonance in the variation of amplitude A versus
frequency ratio 0 for different values of ?
b) The graph below shows the phase angle variation in the phase angle versus frequency

ratio 0
What is the effect of using values of ? greater than 1 in (a) and less than 1 in (b).

and, for small .

2
1 = 0
0

while for a free oscillator20 = km. Thus 1 lies to the right of r , while 0 is still farther
away

from 1 , as shown in Fig. 3.12 where 2 /A20 =1 at = r .

Thus the maximum amplitude A =A1 that occurs - r may be obtained from

Eqs. (3.103) and (3.106) to be

0

=
2
2(0 2 )

In the case of small damping, we assume that --> 0; hence


0 0
Ar 2
It is clear that if b is small or --> 0, the amplitude Ao becomes very large. For undamped
systems, b -- 0 and hence A 0 = ; but there are hardly any systems that are undamped.The
graph below shows the relative positions of the resonance frequency for, the natural
frequency 1, and the natural free frequency 0. Q(o3) are the positions of the resonance
amplitude for frequencies , O, and 1. The graphs are the ratio of the square of the
amplitudes versus n r

) are different values of Q for different frequencies. x i and are ploted against x i - r. Note
the small differences in the sitions of the resonance frequencies.
How do you explain the gradual change in Q for different values of ?

The fact that the amplitude of the oscillations of the system is very large at the
resonance frequency has both desirable and undesirable effects. In the case of electrical
circuits that are sed in tuning radios and in certain types of musical instruments such as organ
pipes, it is de- sirable to have a large response for a small driving force. On the other hand, it
is very undesir-able to have a very large amplitude in mechanical systems, such as in the
springs of an automobile or in the spring mounting of an electric motor. There the aim is to
minimize the transmission of vibrations.

ENERGY RESONANCE

In most practical situations involving oscillating systems in nature, the quantity


observed experimentally is energy and not amplitude. Also, the total energy of an oscillating
system is pro- portional to the square of the amplitude near resonance; hence we should look
for the variation of A 2 versus o. Once again we assume that o is fixed. For steady-state
motion, the amplitude A is constant, and we may write

Let us now calculate time averages of K(t), U(t), and E(t) in the case when A changes with
a,

and remembering that for an average over one period. Substituting for A from Eq. (3.103)
into Eq. (3.112), we get
Below is the graph of the resonance curve (or Lorentzian), that is, the graph of L(o) versus 0)
for different values of ?.

(a) What causes the change in the values of L?

(b) What determines the width of the energy resonance curve?

(c) What does the following analysis indicate about the location of the resonance frequency
on the oyaxis?
RATE OF ENERGY DISSIPATION
CONCLUTION

In classical mechanics, a harmonic oscillator is a system that, when displaced from its
equilibrium position, experiences a restoring force, F, proportional to the displacement, x:

where k is a positive constant.

If F is the only force acting on the system, the system is called a simple harmonic oscillator, and
it undergoes simple harmonic motion: sinusoidal oscillations about the equilibrium point, with a
constant amplitude and a constant frequency (which does not depend on the amplitude).

For a conservative force field, the total energy E of the particle is

E=K + V = constant

Let us further consider Eq. (3.13) for a nonlinear system; that is,

Thus the most general form of a differential equation of the nth order, linear and inhomogeneous,
is

The total energy E(O of a damped harmonic system at any time t is given by

the time period T of one complete oscillation is given by

The amplitude A and the phase angle 4' of steady-state motion according to Eqs. (3.98) and

f0
m
A=
2
(0 )2 + 4y 2 2
2
REFERENCES

Arya, P Atam.1990.Introduction to Classical Mechanics Second Edition.New Jersey:


Prentice Hall.

You might also like