[go: up one dir, main page]

0% found this document useful (0 votes)
13 views90 pages

Solution Set For Introduction T - Jcossham

The document discusses the principles of restoring forces and frequencies in various systems, deriving equations of motion using Newton's second law. It covers different scenarios including pendulums, liquid displacements, and gas behaviors, providing formulas for frequency in terms of stiffness and mass. Additionally, it explores harmonic motion and relationships between different frequencies in a system.

Uploaded by

halesito
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
13 views90 pages

Solution Set For Introduction T - Jcossham

The document discusses the principles of restoring forces and frequencies in various systems, deriving equations of motion using Newton's second law. It covers different scenarios including pendulums, liquid displacements, and gas behaviors, providing formulas for frequency in terms of stiffness and mass. Additionally, it explores harmonic motion and relationships between different frequencies in a system.

Uploaded by

halesito
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 90

Chapter 1

1.1
In Figure 1.1(b), the restoring force is given by:
F   mg sin 

By substitution of relation sin   x l into the above equation, we have:

F  mg x l
so the stiffness is given by:
s   F x  mg l
so we have the frequency given by:
2  s m  g l

Since  is a very small angle, i.e.   sin   x l , or x  l , we have the restoring


force given by:
F   mg

Now, the equation of motion using angular displacement  can by derived from
Newton’s second law:
F  mx
i.e.  mg  ml
g
i.e.     0
l
which shows the frequency is given by:
2  g l

In Figure 1.1(c), restoring couple is given by  C , which has relation to moment of


inertia I given by:

 C  I
C
i.e.    0
I
which shows the frequency is given by:
2  C I

In Figure 1.1(d), the restoring force is given by:


F  2T x l

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
so Newton’s second law gives:
F  mx   2Tx l

i.e. x  2Tx lm  0

which shows the frequency is given by:


2T
2 
lm

In Figure 1.1(e), the displacement for liquid with a height of x has a displacement of
x 2 and a mass of Ax , so the stiffness is given by:

G 2 Axg
s   2 Ag
x 2 x

Newton’s second law gives:


 G  mx
i.e.  2 Axg  Alx

2g
i.e. x  x0
l
which show the frequency is given by:
 2  2g l

In Figure 1.1(f), by taking logarithms of equation pV   constant , we have:

ln p   ln V  constant

dp dV
so we have:  0
p V
dV
i.e. dp  p
V
The change of volume is given by dV  Ax , so we have:
Ax
dp  p
V
The gas in the flask neck has a mass of Al , so Newton’s second law gives:

Adp  mx

A2 x
i.e.  p  Alx
V
pA
i.e. x  x0
lV
which show the frequency is given by:
pA
2 
l V

In Figure 1.1 (g), the volume of liquid displaced is Ax , so the restoring force is
 gAx . Then, Newton’s second law gives:

F   gAx  mx
gA
i.e. x  x0
m
which shows the frequency is given by:
 2  gA m

1.2
Write solution x  a cos(t   ) in form: x  a cos  cos t  a sin  sin t and

compare with equation (1.2) we find: A  a cos and B  a sin  . We can also

find, with the same analysis, that the values of A and B for solution
x  a sin(t   ) are given by: A  a sin  and B  a cos , and for solution

x  a cos(t   ) are given by: A  a cos and B  a sin  .

Try solution x  a cos(t   ) in expression x   2 x , we have:

x   2 x  a 2 cos(t   )   2 a cos(t   )  0

Try solution x  a sin(t   ) in expression x   2 x , we have:

x   2 x  a 2 sin(t   )   2 a sin(t   )  0

Try solution x  a cos(t   ) in expression x   2 x , we have:

x   2 x  a 2 cos(t   )   2 a cos(t   )  0

1.3
(a) If the solution x  a sin(t   ) satisfies x  a at t  0 , then, x  a sin   a
i.e.    2 . When the pendulum swings to the position x   a 2 for the first
time after release, the value of t is the minimum solution of equation
a sin(t   2)   a 2 , i.e. t   4 . Similarly, we can find: for x  a 2 ,

t   3 and for x  0 , t   2 .

If the solution x  a cos(t   ) satisfies x  a at t  0 , then, x  a cos   a

i.e.   0 . When the pendulum swings to the position x   a 2 for the first
time after release, the value of t is the minimum solution of equation
a cos t   a 2 , i.e. t   4 . Similarly, we can find: for x  a 2 , t   3

and for x  0 , t   2 .

If the solution x  a sin(t   ) satisfies xa at t0 , then,

x  a sin( )  a i.e.     2 . When the pendulum swings to the position

x  a 2 for the first time after release, the value of t is the minimum

solution of equation a sin(t   2)   a 2 , i.e. t   4 . Similarly, we can

find: for x  a 2 , t   3 and for x  0 , t   2 .

If the solution x  a cos(t   ) satisfies xa at t0 , then,

x  a cos( )  a i.e.   0 . When the pendulum swings to the position

x  a 2 for the first time after release, the value of t is the minimum

solution of equation a cos t   a 2 , i.e. t   4 . Similarly, we can find: for

x  a 2 , t   3 and for x  0 , t   2 .

(b) If the solution x  a sin(t   ) satisfies x  a at t0 , then,

x  a sin   a i.e.     2 . When the pendulum swings to the position

x  a 2 for the first time after release, the value of t is the minimum

solution of equation a sin(t   2)   a 2 , i.e. t  3 4 . Similarly, we can


find: for x  a 2 , t  2 3 and for x  0 , t   2 .

If the solution x  a cos(t   ) satisfies x  a at t0 , then,

x  a cos  a i.e.    . When the pendulum swings to the position

x  a 2 for the first time after release, the value of t is the minimum

solution of equation a cos(t   )   a 2 , i.e. t  3 4 . Similarly, we can

find: for x  a 2 , t  2 3 and for x  0 , t   2 .

If the solution x  a sin(t   ) satisfies x  a at t0 , then,

x  a sin( )  a i.e.    2 . When the pendulum swings to the position

x  a 2 for the first time after release, the value of t is the minimum

solution of equation a sin(t   2)   a 2 , i.e. t  3 4 . Similarly, we can

find: for x  a 2 , t  2 3 and for x  0 , t   2 .

If the solution x  a cos(t   ) satisfies x  a at t0 , then,

x  a cos( )  a i.e.    . When the pendulum swings to the position

x  a 2 for the first time after release, the value of t is the minimum

solution of equation a cos(t   )   a 2 , i.e. t  3 4 . Similarly, we can

find: for x  a 2 , t  2 3 and for x  0 , t   2 .

1.4
The frequency of such a simple harmonic motion is given by:
s e2 (1.6  10 19 ) 2
0     4.5  1016 [rad  s 1 ]
me 4 0 r 3me 4    8.85  10  (0.05  10 )  9.1 10
12 9 3 31

Its radiation generates an electromagnetic wave with a wavelength  given by:


2c 2    3  108
   4.2  10 8 [m]  42[nm]
0 4.5  10 16

Therefore such a radiation is found in X-ray region of electromagnetic spectrum.


1.5
(a) If the mass m is displaced a distance of x from its equilibrium position, either
the upper or the lower string has an extension of x 2 . So, the restoring force of

the mass is given by: F   sx 2 and the stiffness of the system is given by:

s   F x  s 2 . Hence the frequency is given by a2  s m  s 2m .

(b) The frequency of the system is given by: b2  s m


(c) If the mass m is displaced a distance of x from its equilibrium position, the
restoring force of the mass is given by: F   sx  sx  2 sx and the stiffness of
the system is given by: s   F x  2 s . Hence the frequency is given by

c2  s m  2s m .

Therefore, we have the relation: a2 : b2 : c2  s 2m : s m : 2s m  1 : 2 : 4

1.6
At time t  0 , x  x0 gives:

a sin   x0 (1.6.1)

x  v0 gives:

a cos   v0 (1.6.2)

From (1.6.1) and (1.6.2), we have


tan   x0 v0 and a  ( x02  v02  2 )1 2

1.7
In the capacitor of Figure 1.5, the time taken for q changing to q  dq is:

dq
dt 
I

where
I  q  q0 sin t

Charge has a value between q and q  dq twice per oscillation.


2dt
probability   
T
where
2
T  period 


2dt dq 1 dq 1 dq
   
I  q0 sin t 
q  q 2 2
1
T 2
0

So, the same, the probability of the current in the inductance L to lie between I
and I  dI is:
1 dI

I  I 2 2
1
2
0

1.8
Since the displacements of the equally spaced oscillators in y direction is a sine

curve, the phase difference  between two oscillators a distance x apart given is

proportional to the phase difference 2 between two oscillators a distance  apart


by:  2  x  , i.e.   2x  .

1.9
The mass loses contact with the platform when the system is moving downwards and
the acceleration of the platform equals the acceleration of gravity. The acceleration of
a simple harmonic vibration can be written as: a  A 2 sin(t   ) , where A is the

amplitude,  is the angular frequency and  is the initial phase. So we have:

A 2 sin(t   )  g

g
i.e. A
 sin(t   )
2

Therefore, the minimum amplitude, which makes the mass lose contact with the
platform, is given by:
g g 9.8
Amin     0.01[m]
 2
4 f
2 2
4   2  52

1.10
The mass of the element dy is given by: m  mdy l . The velocity of an element

dy of its length is proportional to its distance y from the fixed end of the spring, and

is given by: v  yv l . where v is the velocity of the element at the other end of the

spring, i.e. the velocity of the suspended mass M . Hence we have the kinetic energy
of this element given by:
2
1 1  m  y 
KEdy  mv2   dy  v 
2 2 l  l 
The total kinetic energy of the spring is given by:
2
l 1  m  y 
l mv 2 l 1
KEspring   KEdy dy    dy  v   y dy  6 mv
2 2
0 0 2
l  l  2l 3 0

The total kinetic energy of the system is the sum of kinetic energies of the spring and
the suspended mass, and is given by:
KEtot  mv 2  Mv 2  M  m 3v 2
1 1 1
6 2 2
which shows the system is equivalent to a spring with zero mass with a mass of
M  m 3 suspended at the end. Therefore, the frequency of the oscillation system is

given by:
s
2 
M m 3

1.11
In Figure 1.1(a), the energy is directly given by:
1 1
E  mv 2  sx 2
2 2
The equation of motion is by setting dE dt  0 , i.e.:

d 1 2 1 2
 mx  sx   0
dt  2 2 
s
i.e. x  x0
m
In Figure 1.1(b), the restoring force of the simple pendulum is  mg sin  , then, the

stiffness is given by: s  mg sin  x  mg l . So the energy is given by:


1 2 1 2 1 2 1 mg 2
E mv  sx  mx  x
2 2 2 2 l
The equation of motion is by setting dE dt  0 , i.e.:

d  1 2 1 mg 2 
 mx  x 0
dt  2 2 l 
g
i.e. x  x0
l

In Figure 1.1(c), the displacement is the rotation angle  , the mass is replaced by the
moment of inertia I of the disc and the stiffness by the restoring couple C of the
wire. So the energy is given by:
1 1
E  I 2  C 2
2 2
The equation of motion is by setting dE dt  0 , i.e.:

d  1 2 1 2
 I  C   0
dt  2 2 
C
i.e.    0
I

In Figure 1.1(d), the restoring force is given by:  2 Tx l , then the stiffness is given

by: s  2T l . So the energy is given by:


1 2 1 2 1 2 1 2T 2 1 2 T 2
E mv  sx  mx  x  mx  x
2 2 2 2 l 2 l
The equation of motion is by setting dE dt  0 , i.e.:

d 1 2 T 2
 mx  x   0
dt  2 l 
2T
i.e. x  x0
lm

In Figure 1.1(e), the liquid of a volume of Al is displaced from equilibrium

position by a distance of l 2 , so the stiffness of the system is given by

s  2 gAl l  2 gA . So the energy is given by:

1 2 1 2 1 1 1
E mv  sx  Alx 2  2 gAx 2  Alx 2  gAx 2
2 2 2 2 2
The equation of motion is by setting dE dt  0 , i.e.:

d 1 2
 Alx  gAx   0
2

dt  2 
2g
i.e. x  x0
l

In Figure 1.1(f), the gas of a mass of Al is displaced from equilibrium position by

a distance of x and causes a pressure change of dp   pAx V , then, the stiffness

of the system is given by s   Adp x  pA2 V . So the energy is given by:

1 2 1 2 1 1 pA2 x 2
E mv  sx  Alx 2 
2 2 2 2 V

The equation of motion is by setting dE dt  0 , i.e.:

d 1 1 pA2 x 2 
 Alx 2  0
dt  2 2 V 

pA
i.e. x  x0
lV

In Figure 1.1(g), the restoring force of the hydrometer is  gAx , then the stiffness

of the system is given by s  gAx x  gA . So the energy is given by:


1 2 1 2 1 2 1
E mv  sx  mx  gAx 2
2 2 2 2
The equation of motion is by setting dE dt  0 , i.e.:

d 1 2 1 2
 mx  gAx   0
dt  2 2 
Ag
i.e. x  x0
m

1.12
The displacement of the simple harmonic oscillator is given by:
x  a sin t (1.12.1)
so the velocity is given by:
x  a cos t (1.12.2)
From (1.12.1) and (1.12.2), we can eliminate t and get:

x2 x 2
  sin 2 t  cos 2 t  1 (1.12.3)
a 2
a
2 2

which is an ellipse equation of points ( x, x ) .

The energy of the simple harmonic oscillator is given by:


1 1
E  mx 2  sx 2 (1.12.4)
2 2
Write (1.12.3) in form x 2   2 (a 2  x 2 ) and substitute into (1.12.4), then we have:
1 2 1 2 1 1
E mx  sx  m 2 (a 2  x 2 )  sx 2
2 2 2 2
Noting that the frequency  is given by:  2  s m , we have:
1 1 1
E s (a 2  x 2 )  sx 2  sa 2
2 2 2
which is a constant value.

1.13
The equations of the two simple harmonic oscillations can be written as:
y1  a sin(t   ) and y2  a sin(t     )

The resulting superposition amplitude is given by:


R  y1  y2  a[sin(t   )  sin(t     )]  2a sin(t     2) cos( 2)

and the intensity is given by:


I  R 2  4a 2 cos 2 ( 2) sin 2 (t     2)

i.e. I  4a 2 cos 2 ( 2)

Noting that sin 2 (t     2) varies between 0 and 1, we have:

0  I  4a 2 cos 2 ( 2)

1.14
1
2   40 2
LC
103  106
(a)  2   103   31.2 Hz
4  250  100
106  106 1010
(b)  2    25  107  250  106   16  103 Hz
40  25  4 40

1.15
By elimination of t from equation x  a sin t and y  b cos t , we have:

x2 y 2
 1
a 2 b2
which shows the particle follows an elliptical path. The energy at any position of x ,
y on the ellipse is given by:

1 2 1 2 1 2 1 2
E mx  sx  my  sy
2 2 2 2
1 1 1 1
 ma 2 2 cos 2 t  ma 2 2 sin 2 t  mb 2 2 sin 2 t  mb 2 2 cos 2 t
2 2 2 2
1 1
 ma 2 2  mb 2 2
2 2
1
 m 2 (a 2  b 2 )
2
The value of the energy shows it is a constant and equal to the sum of the separate
1
energies of the simple harmonic vibrations in x direction given by m 2 a 2 and in
2
1
y direction given by m 2b 2 .
2
At any position of x , y on the ellipse, the expression of m( xy  yx ) can be

written as:
m( xy  yx )  m(ab sin 2 t  ab cos 2 t )  abm (sin 2 t  cos 2 t )  abm
which is a constant.

1.16
The period of rapid oscillation is:
2 4 2
T1   
(1  2 ) / 2 1  2 
The period of slow frequency envelope is:
2 4
T2  
(2  1 ) / 2 
The number of rapid oscillations between two consecutive zeros of the slow
frequency envelope is:
T2 / 2 

T1 
Chapter 2

2.1
Eq. 2.3 is x  ( A  Bt )e 0t so x  A at t  0 gives x  Ae 0t . x  0 at t  0 gives
 0 Ae 0t  Be 0t  0 .  B  0 A ,  solution is x  A(1  0t )e 0t .

2.2
r r  force 
Log dec. :     0.2 ,  r  1 
2m  10  0.5  velocity 
s r2 
( 2 ) 2  40  2  40 2    
2 
 m 4 m 
 r2 
 s  40m 2  200  0.52  50 Nm -1  neglecting 
 4m 2 

2.3
Critical damping gives:
r s
 0 
2m m
where sx  mg and x is the spring extension.
1
g 9.81
 r  2m  10  10(100) 2
x 0 .1
r  100 kgs -1

2.4
Use the relation between current and charge, I  q , and the voltage equation:
q C  IR  0
we have the equation:
Rq  q C  0
solve the above equation, we get:
q  C1e  t RC
where C1 is arbitrary in value. Use initial condition, we get C1  q0 ,
i.e. q  q0e  t RC
which shows the relaxation time of the process is RC s.

2.5
(a) 02 -  2  10-6 02  r 2 4m 2 => 0 m r  500
The condition also shows    0 , so:
Q   m r  0 m r  500
2
Use    , we have:

r r  
 
  
2m m  Q 500
(b) The stiffness of the system is given by:
s  02 m  1012  1010  100[ Nm 1 ]

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
and the resistive constant is given by:
 m 106  1010
r 0   2  10 7 [ N  sm 1 ]
Q 500
(c) At t  0 and maximum displacement, x  0 , energy is given by:
1 1 1 2 1
E  mx 2  sx 2  sxmax   100  10 4  5  10 3[ J ]
2 2 2 2
1
Time for energy to decay to e of initial value is given by:
m 1010
t   0.5[ms]
r 2  10 7
(d) Use definition of Q factor:
E
Q  2
 E
where, E is energy stored in system, and  E is energy lost per cycle, so energy
loss in the first cycle,  E1 , is given by:
E 5  103
 E1  E  2  2   2  10 5 [ J ]
Q 500

2.6
The frequency of a damped simple harmonic oscillation is given by:
r2 r2 r2
 2  02  2 =>  2  02  =>      -  
4m 2    0 
0
4m 4m 2
m
Use     and Q  0 we find fractional change in the resonant frequency is
r
given by:
    0
0

0

r2
8m 0
2 2
 8Q 2 1

2.7
The damped frequency is given by:
  
2
r2
 2  02   02   
 2 
2
4m
r  r 1
   Angle    
2m 20 2m0 2Q
(Note: system approaches centre twice per cycle.)

2.8
Equation of motion:
e2 x
mx  n
0
ne 2
 e 
me 0
2.9
The relaxation time:
t 4
   5.77 sec s
log( E0 / E ) log e 2
Q  0  2  330  5.77  1.2  104
E 2
   5.3  10  4
E Q
Chapter 3

3.1
In the resonant LCR series circuit shown in Figure 3.3, the maximum potential across
the capacitor is given by:
VC max 02

V0 
 2 R 2 / L2   2  02 
2

1/ 2

which has a maximum value at the minimum of the term:

 2 R 2 / L2   2  02 
2

that is when
d
d

 2 R 2 / L2   2  02   0
2

i.e. when
2 R 2
 4  2  02   0
L2
which can be simplified as:
R2
 2  02 
2L2

By substituting 02  LC 


1
and Q0  0 L / R into the above equation, we have:

1 1 1
 R 2

2 R  2
1 
2 2
   02     02  02
2 
  0 1 
2 2 

2 
 2L   20 L   2Q0 

3.2
The maximum potential across the inductance occurs when the maximum current
I 0 max flowing through the inductance impedance Z L  L , so:

LV0  2V0
VL max  I 0 maxL  1/ 2
 1/ 2
L  2 R 2 2  2 R 2 2 2
   2  02    L2    0  
2

 L 2
  
which has a maximum value at the minimum of the term:

 2 R 2 / L2   2  02 
2

i.e. when
d
d

 2 R 2 / L2   2  02   0
2

Using the same derivation in question 3.1, we have:

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
1
 1 2
  0 1  2 
 2Q0 

3.3
1
(a)  2   398 Hz
LC
V 15
(b)   0.2 amps
R 75

3.4
The low frequency limit of the bandwidth of the resonance absorption curve 1

satisfies the equation:


1m  s 1   r
which defines the phase angle given by:
1m  s 1
tan 1   1
r
The high frequency limit of the bandwidth of the resonance absorption curve 2

satisfies the equation:


2 m  s 2  r
which defines the phase angle given by:
2 m  s  2
tan 2  1
r

3.5
(a) 0  100  2 sec 1

(b)   10  2 sec 1

(c) Q  100 / 10  10

rt r 10
(d)   e 1 at   1 cycle
m mv 10

3.6
Acceleration amplitude x  F0 / Z m . Velocity resonance frequency is 0 where

02  s / m i.e. m  s /  , so:


0 F0 0 F0 0 F0 F0
x    
r 
2 1/ 2
sm 
1/ 2
0 m m
At very high frequencies:
F0 F0 Z 1
x   and the limit of m as    
Zm Zm
1  m

3.7
(a) F0 / s ; (b) F0 / r ; (c) F0 / m .

3.8
Write the equation of an undamped simple harmonic oscillator driven by a force of
frequency  in the vector form, and use F0eit to represent its imaginary part

F0 sin t , we have:

mx  sx  F0eit (3.8.1)

We try the steady state solution x  Aeit and the velocity is given by:

x  iAeit  ix
so that:
x  i 2 2 x   2 x
and equation (3.8.1) becomes:
( A 2 m  As )eit  F0eit
which is true for all t when
 A 2 m  As  F0
F0
i.e. A
s   2m
F0
i.e. x eit
s  m2

The value of x is the imaginary part of vector x , given by:


F0
x sin t
s   2m
F0 sin t s
i.e. x where 02 
m(02   2 ) m
Hence, the amplitude of x is given by:
F0
A
m(   2 )
2
0

and its behaviour as a function of frequency is shown in the following graph:

By solving the equation:


mx  sx  0
we can easily find the transient term of the equation of the motion of an undamped
simple harmonic oscillator driven by a force of frequency  is given by:
x  C cos 0t  D sin 0t

where, 0  s m , C and D are constant. Finally, we have the general solution


for the displacement given by the sum of steady term and transient term:
F0 sin t
x  C cos 0t  D sin 0t (3.8.2)
m(02   2 )

3.9
The undamped oscillatory equation for a bound electron is given by:
x  02 x  ( eE0 m) cos t (3.9.1)
Try solution x  A cos t in equation (3.9.1), we have:
( 2  02 ) A cos t  ( eE0 m) cos t
which is true for all t provided:
( 2  02 ) A   eE0 m

eE0
i.e. A
m(02   2 )
So, we find a solution to equation (3.9.1) given by:
eE0
x cos t (3.9.2)
m(02   2 )
For an electron number density n , the induced polarizability per unit volume of a
medium is given by:
nex
e   (3.9.3)
0E

By substitution of (3.9.2) and E  E0 cos t into (3.9.3), we have

nex ne 2
e   
 0 E  0 m(02   2 )

3.10
The forced mechanical oscillator equation is given by:
mx  rx  sx  F0 cos t

which can be written as:


mx  rx  m02 x  F0 cos t (3.10.1)

where, 0  s m . Its solution can be written as:

F0 r FX
x sin t  0 2m cos t (3.10.2)
Z m2
Z m

 m  s  
where, X m  m  s  , Z m  r 2  (m  s  ) 2 ,   tan 1  
 r 
By taking the displacement x as the component represented by curve (a) in Figure
3.11, i.e. by taking the second term of equation (3.10.2) as the expression of x , we
have:
F0 X m F0 m(02   2 )
x cos  t  cos t (3.10.3)
Z m2 m 2 (02   2 ) 2   2 r 2
The damped oscillatory electron equation can be written as:
mx  rx  m02 x  eE0 cos t (3.10.4)
Comparing (3.10.1) with (3.10.4), we can immediately find the displacement x for a
damped oscillatory electron by substituting F0  eE0 into (3.10.3), i.e.:

eE0 m(02   2 )
x cos t (3.10.5)
m 2 (02   2 ) 2   2 r 2

By substitution of (3.10.5) into (3.9.3), we can find the expression of  for a

damped oscillatory electron is given by:


nex ne 2 m(02   2 )
  
 0 E  0 [m 2 (02   2 ) 2   2 r 2 ]
So we have:
ne 2 m(02   2 )
r  1   1 cos t
 0 [m 2 (02   2 ) 2   2 r 2 ]

3.11
Considering an electron in an atom as a lightly damped simple harmonic oscillator,
we know its resonance absorption bandwidth is given by:
r
  (3.11.1)
m
On the other hand, the relation between frequency and wavelength of light is given
by:
c
f  (3.11.2)

where, c is speed of light in vacuum. From (3.11.2) we find at frequency resonance:
c
f   
20

where 0 is the wavelength at frequency resonance. Then, the relation between

angular frequency bandwidth  and the width of spectral line  is given by:
2c
  2 f   (3.11.3)
20
From (3.11.1) and (3.11.3) we have:
20 r r 
   0  0
2cm 0 m Q
So the width of the spectral line from such an atom is given by:
0 0.6  10 6
    1.2  10 14 [m]
Q 5  10 7

3.12

The transient term of a forced oscillator decay with e  rt 2 m to e - k at time t , i.e.:

 rt 2m   k

so, we have the resistance of the system given by:


r  2mk t
(3.12.1)
For small damping, we have
  0  s m (3.12.2)
We also have steady state displacement given by:
x  x0 sin(t   )

where the maximum displacement is:


F0
x0 
 r  (m  s m) 2
2

(3.12.3)
By substitution of (3.12.1) and (3.12.2) into (3.12.3), we can find the average rate of
growth of the oscillations given by:
x0 F0

t 2km0
Chapter 4

4.1
The kinetic energy of the system is the sum of the separate kinetic energy of the two
masses, i.e.:
1 1 1 1 1  1 1
Ek  mx 2  my 2  m  ( x  y ) 2  ( x  y ) 2   mX 2  mY 2
2 2 2 2 2  4 4
The potential energy of the system is the sum of the separate potential energy of the
two masses, i.e.:
1 mg 2 1 1 mg 2 1
Ep  x  s( y  x)2  y  s( x  y)2
2 l 2 2 l 2
1 mg 2
 ( x  y 2 )  s( x  y)2
2 l
1 mg  1 1 
  ( x  y)2  ( x  y)2   s( x  y)2
2 l 2 2 
1 mg 2  1 mg 
 X   s Y 2
4 l 2 l 
Comparing the expression of Ek and E p with the definition of E X and EY given
by (4.3a) and (4.3b), we have:
1 mg 1 mg
a  m, b  , c  m , and d  s
2 4l 4 2l
Noting that:
12 12
m m
X q    ( x  y)    X
2 2
12 12
m m
and Yq    ( x  y )    Y
2 c
1 2 1 2
m m
i.e. X    X q and Y    Yq
2 2
we have the kinetic energy of the system given by:
2 2
1  2   1  2  1 1
E k  aX  cY  m 
2 2
X q   m Yq   X q2  Yq2
4  m  4  m  2 2
and
2 2
mg  2   mg  2  1 g 2  g 2s  2
E p  bX  dY 
2 2
 Xq    s  Yq   X q    Yq
4l  m   2l  m  2 l l m
which are the expressions given by (4.4a) and (4.4b)

4.2
s
2  gives s  46 N  m -1
m

4.3
x  2a , y  0 :

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
-X + Y

≡ +

-2a y=0 -a -a -a a

x  0 , y  2 a :
-X - Y

≡ +

x=0 -2a -a -a -a a

4.4
For mass m1 , Newton’s second law gives:
m1x1  sx
For mass m2 , Newton’s second law gives:
m2 x2   sx
Provided x is the extension of the spring and l is the natural length of the spring,
we have:
x2  x1  l  x
By elimination of x1 and x2 , we have:
s s
 x x  x
m2 m1
m  m2
i.e. x  1 sx  0
m1m2
which shows the system oscillate at a frequency:
s
2 

where,
m1m2

m1  m2
For a sodium chloride molecule the interatomic force constant s is given by:
(2 ) 2 mNa mCl 4 2  (1.14  1013 ) 2  (23  35)  (1.67  10 27 ) 2
s   2   27
 120[ Nm 1 ]
mNa  mCl (23  35)  1.67  10

4.5
If the upper mass oscillate with a displacement of x and the lower mass oscillate
with a displacement of y , the equations of motion of the two masses are given by
Newton’s second law as:
mx  s ( y  x )  sx
my  s ( x  y )
i.e.
mx  s ( x  y )  sx  0
my  s ( x  y )  0
Suppose the system starts from rest and oscillates in only one of its normal modes of
frequency  , we may assume the solutions:
x  Aeit
y  Beit
where A and B are the displacement amplitude of x and y at frequency  .
Using these solutions, the equations of motion become:
[m 2 A  s( A  B)  sA]eit  0
[m 2 B  s( A  B)]eit  0
We may, by dividing through by me it , rewrite the above equations in matrix form as:

2 s m   2  s m   A
    0 (4.5.1)
 s m s m   2   B
which has a non-zero solution if and only if the determinant of the matrix vanishes;
that is, if
(2 s m   2 )( s m   2 )  s 2 m 2  0
i.e.  4  (3s m)  2  s 2 m 2  0
s
i.e.  2  (3  5 )
2m
In the slower mode,   (3  5 ) s 2 m . By substitution of the value of frequency
2

into equation (4.5.1), we have:


A s s  m 2 5 1
  
B 2 s  m 2
s 2
which is the ratio of the amplitude of the upper mass to that of the lower mass.
Similarly, in the slower mode,  2  (3  5 ) s 2m . By substitution of the value of
frequency into equation (4.5.1), we have:
A s s  m 2 5 1
  
B 2 s  m 2
s 2

4.6
By adding up the two equations of motion, we have:
m1x  m2 y  (m1 x  m2 y)( g l )
By multiplying the equation by 1 (m1  m2 ) on both sides, we have:
m1 x  m2 y g m1 x  m2 y

m1  m2 l m1  m2
m1x  m2 y g m1 x  m2 y
i.e.  0
m1  m2 l m1  m2
which can be written as:
X  12 X  0 (4.6.1)
where,
m1 x  m2 y
X  and 12  g l
m1  m2
On the other hand, the equations of motion can be written as:
g s
x   x  ( x  y )
l m1
g s
y  y ( x  y)
l m2
By subtracting the above equations, we have:

g  s s 
x  y   ( x  y )    ( x  y )
l  m1 m2 
g  1 1 
i.e. x  y    s    ( x  y )  0
l  m1 m2  
which can be written as:
Y  22Y  0 (4.6.2)
where,
g  1 1 
Y  x y and 22   s  
l  m1 m2 
Equations (4.6.1) and (4.6.2) take the form of linear differential equations with
constant coefficients and each equation contains only one dependant variable,
therefore X and Y are normal coordinates and their normal frequencies are given
by 1 and 2 respectively.

4.7
Suppose the displacement of mass M is x , the displacement of mass m is y ,
and the tension of the spring is T . Equations of motion give:
Mx  kx  F0 cos t  T (4.7.1)
my  T (4.7.2)
s ( y  x)  T (4.7.3)
Eliminating T , we have:
Mx  kx  F0 cos t  s ( y  x)
so for x  0 at all times, we have
F0 cos t  sy  0
that is
F
y   0 cos t
s
Equation (4.7.2) and (4.7.3) now give:
my  sy  0
with   s m , so M is stationary at  2  s m .
2

This value of  satisfies all equations of motion for x  0 including


T   F0 cos t
4.8
Noting the relation: V  q C , the voltage equations can be written as:
q1 q2 dI
 L a
C C dt
q2 q3 dI
 L b
C C dt
so we have:
q1  q 2  LCIa
q  q  LCI
2 3 b
i.e.
q1  q2  LCIa
q  q  LCI
2 3 b

By substitution of q1   I a , q 2  I a  I b and q3  I b into the above equations, we


have:
 I a  I a  I b  LCIa
a I  I  I  LCI
b b b
i.e.
LCIa  2 I a  I b  0
LCIb  I a  2 I b  0
By adding up and subtracting the above equations, we have:
LC ( Ia  Ib )  I a  I b  0
LC ( I  I )  3( I  I )  0
a b a b
Supposing the solutions to the above normal modes equations are given by:
I a  I b  A cos t
I a  I b  B cos t
so we have:
( A 2 LC  A) cos t  0
( B 2 LC  3B ) cos t  0
which are true for all t when
1
2  and B0
LC
or
3
2  and A0
LC
which show that the normal modes of oscillation are given by:
1
I a  I b at 12 
LC
and
3
I a   I b at 22 
LC

4.9
From the given equations, we have the relation between I1 and I 2 given by:
iM
I2  I1
Z 2  iLs
so:
  2M 
E  iL p I1  iMI 2   iL p   I1
 Z 2  iLs 
i.e.
E  2M 2
 iL p 
I1 Z 2  iLs
which shows that E I1 , the impedance of the whole system seen by the generator, is
the sum of the primary impedance, iLp , and a ‘reflected impedance’ from the
secondary circuit of  2 M 2 Z s , where Z s  Z 2  iLs .

4.10
Problem 4.9 shows the impedance seen by the generator Z is given by:
 2M 2
Z  i L p 
Z 2  iLs
Noting that M  L p Ls and L p Ls  n 2p ns2 , the impedance can be written as:

iL p Z 2   2 L p Ls   2 M 2 i L p Z 2   2 M 2   2 M 2 i L p Z 2
Z  
Z 2  iLs Z 2  iLs Z 2  iLs
so we have:
1 Z 2  iLs 1 1 1 1
     2
Z i L p Z 2 i L p L p iL p n p
Z2 Z2
Ls ns2
which shows the impedance Z is equivalent to the primary impedance iLp
connected in parallel with an impedance (n p ns ) 2 Z 2 .

4.11
Suppose a generator with the internal impedance of Z1 is connected with a load with
an impedance of Z 2 via an ideal transformer with a primary inductance of L p and
the ratio of the number of primary and secondary transformer coil turns given by
n p ns , and the whole circuit oscillate at a frequency of  . From the analysis in
Problem 4.9, the impedance of the load is given by:
1 1 1
  2
Z L i L p n p
Z2
ns2
At the maximum output power: Z L  Z1 , i.e.:
1 1 1 1
  2 
Z L iL p n p Z1
2
Z2
ns
which is the relation used for matching a load to a generator.
4.12
By substitution of j  1 and n  3 into equation (4.12), we have:
   2
12  202 1  cos   202 1    (2  2 )0
2

 4   2
By substitution of j  2 and n  3 into equation (4.15), we have:
 2 
12  202 1  cos   202
 4 
By substitution of j  3 and n  3 into equation (4.12), we have:
 3   2
12  202 1  cos
  2 2
0 1    (2  2 )0
2

 4  2 
In equation (4.11), we have A0  A4  0 when n  3 , and noting that 02  T ma ,
equation (4.11) gives:
 2 
when r  1 :  A0   2  2  A1  A2  0

 0 
 2 
i.e.  2  2  A1  A2  0 (4.12.1)
 0 
 2 
when r  2 :  A1   2  2  A2  A3  0
 (4.12.2)
 0 
 2 
when r  3 :  A2   2  2  A3  A4  0
 0 
 2 
i.e.  A2   2  2  A3  0 (4.12.3)
  0 

Write the above equations in matrix format, we have:


 2   2 02 1 0  A1 
  
 1 2   2 02  1  A2   0
 0 1 2   2 02  A3 

which has non zero solutions provided the determinant of the matrix is zero, i.e.:
(2   2 02 )3  2(2   2 02 )  0
The solutions to the above equations are given by:
12  (2  2 )02 ,` 22  202 , and 12  (2  2 )02

4.13
By substitution of 12  (2  2 )02 into equation (4.12.1), we have:
2 A1  A2  0 i.e. A1 : A2  1 : 2
By substitution of 12  (2  2 )02 into equation (4.17.3), we have:
 A2  2 A3  0 i.e. A2 : A3  2 : 1
Hence, when 12  (2  2 )02 , the relative displacements are given by:
A1 : A2 : A3  1 : 2 : 1
By substitution of 22  202 into equation (4.12.1), we have:
A2  0
By substitution of 2  20 into equation (4.12.2), we have:
2 2

 A1  A3  0 i.e. A1 : A3  1 : 1
Hence, when 22  202 , the relative displacements are given by:
A1 : A2 : A3  1 : 0 : 1
By substitution of 22  (2  2 )02 into equation (4.12.1), we have:
 2 A1  A2  0 i.e. A1 : A2  1 :  2
By substitution of 12  (2  2 )02 into equation (4.17.3), we have:
 A2  2 A3  0 i.e. A2 : A3   2 : 1
Hence, when 12  (2  2 )02 , the relative displacements are given by:
A1 : A2 : A3  1 :  2 : 1
The relative displacements of the three masses at different normal frequencies are
shown below:

As we can see from the above figures that tighter coupling corresponds to higher
frequency.

4.14
By expansion of the expression of  2j , we have:
2T  j  2T  ( j n  1) 2 ( j  n  1) 4 ( j n  1) 6 
 2j  1  cos      
ma  n  1  ma  2! 4! 6! 
If n  1 and j  n , j n  1 has a very small value, so the high order terms of
the above equation can be neglected, so the above equation become:
2T  ( j n  1) 2  T  j 
2

 2j  
 ma  n  1 
ma  2!   
j T
i.e. j 
n  1 ma
which can be written as:
j T
j 
l 
where,   m a and l  (n  1)a
4.15
L
d
I r1  I r   Vr 1  Vr   Vr  Vr1   Vr1  2Vr  Vr 1 
dt
qr
 Vr
c
dq d
 r  C Vr
dt dt
and
d 2 qr d2
V  I r 1  I r 
d
2
 C 2 r
dt dt dt
2
dV
 2r 
1
Vr 1  2Vr  Vr 1 
dt LC

4.16
(a)
2 y
By substitution of y into , we have:
t 2
 2 y  2 it ikx
 2 (e e )   2ei (t kx )
t 2
t
 y
2
By substitution of y into , we have:
x 2
 2 y  2 it ikx
 2 (e e )  k 2ei (t kx )
x 2
x
If   ck , we have:
2 y 2 2 y
c  ( 2  c 2 k 2 )ei (t kx )  0
t 2
x 2

i.e.
2 y 2  y
2
 c
t 2 x 2

(b)
  y     y 
   iy  ky   
x  t  x x t  x  t
Chapter 5

5.1
If y  f1 (ct  x ) , the expression for y at a time t  t and a position x  x ,

where t  x c , is given by:

yt t , x x  f1[c(t  t )  ( x  x)]


 f1[c(t  x c)  ( x  x)]
 f1[ct  x  x  x]
 f1[ct  x]  yt , x
i.e. the wave profile remains unchanged.
If y  f 2 (ct  x ) , the expression for y at a time t  t and a position x  x ,

where t   x c , is given by:

yt  t , x  x  f1[c(t  t )  ( x  x)]
 f1[c(t  x c)  ( x  x)]
 f1[ct  x  x  x]
 f1[ct  x]  yt , x
i.e. the wave profile also remains unchanged.

5.2
The pulse shape before reflection is given by the graph below:

The pulse shapes after of a length of l of the pulse being reflected are shown
below:
(a) l  l 4

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
(b) l  l 2

(c) l  3l 4

(d) l  l
5.3
1
 T 2 1
1/ 2
 12 
Power: P   2 A2    .0340  9  10 4 .0152 
1
  240Watts
2  2  .03 

P   2  2 requires 4 times the power  960 Watts

1 1
P  A2  amplitude requires of the power  60 Watts
2 4

5.4
  c and L   / 2  fundamenta l   2L

T  c 2   ( 2 L ) 2  3 2  0.7  220   161N


1 .7 2

10
This is equivalent to supporting a mass  T / g  16.4 kg .
There are 4 strings. Very old instruments were made to withstand much less stress
from gut strings and the introduction of later strings damaged and in some cases
destroyed many valuable instruments.

5.5
The boundary condition yi  yr  yt gives:

A1ei (t kx )  B1ei (t kx )  A2ei (t kx )

At x  0 , this equation gives:


A1  B1  A2 (5.5.1)

 
The boundary condition Ma  T yt  T ( yi  yr ) gives:
x x
Ma  ikTA2 e i (t kx )  ikTA1e i (t kx )  ikTB1e i (t kx )

At x  0 , a  yt  yi  yr , so the above equation becomes:

T T T
  2 MA2  i A2  i A1  i B1
c c c
T T  T
i.e. i A1  i B1    M  i  A2
c c  c

Noting that T c  c , the above equation becomes:

icA1  icB1   M  ic A2 (5.5.2)

By substitution of (5.5.1) into (5.5.2), we have:


icA1  icB1   M  ic ( A1  B1 )

i.e.
B1  iq

A1 1  iq

where q  M 2  c

By substitution of the above equation into (5.5.1), we have:


iq
A1  A1  A2
1  iq

i.e.
A2 1

A1 1  iq

5.6
Suppose the wave equation is given by: y  sin(t  kx) . The maximum value of

transverse harmonic force Fmax is given by:

 y    TA
Fmax  T    T  A sin(t  kx)  TAk 
 x  max  x  max c

i.e.
T Fmax 0.3 0.3
  
c A 0.1 2  5 
Noting that c  T c , the rate of energy transfer along the string is given by:

c 2 A2 1 T 2 2 1 0.3 3
P   A    (2  5) 2  0.12  [W ]
2 2c 2  20
so the velocity of the wave c is given by:
2P 2  3 20 30
c   [ms 1 ]
 A
2 2
0.01 (2  5)  0.1
2 2

5.7
(a) 2nd harmonic = 880 Hz. 3rd harmonic = 1320 Hz.
(b) each octave adds a harmonic so n  440  16000 n  36
(c) reducing the length of the string increases the frequency,
L2  1
  L2  26.9 or 5.1 cm from the end of the string.
L1  2

5.8
The impedance of the anti-reflection coating Z coat should have a relation to the

impedance of air Z air and the impedance of the lens Z lens given by:

1
Z coat  Z air Z lens 
nair nlens

So the reflective index of the coating is given by:


1
ncoat   nair nlens  1.5  1.22
Z coat

and the thickness of the coating d should be a quarter of light wavelength in the
coating, i.e.
 5.5  10 7
d   1.12  10 7 [m]
4ncoat 4  1.22

5.9
y
By substitution of equation (5.4) into , we have:
x
y n t
 ( An cos nt  Bn sin nt ) cos n
x c c
so:

2 y n2 n t n2
  ( A cos  t  B sin  t ) sin   y
x 2
n n n n
c2 c c2
n
Noting that k  , we have:
c

2 y n2 n2
 k 2
y   y  y0
x 2 c2 c2

5.10
By substitution of the expression of ( yn2 ) max into the integral, we have:
1 l 1 l x
n2  ( yn2 ) max dx  n2 ( An2  Bn2 )  sin 2 n dx
2 0 2 0 c
1 l 1  cos(2n x c)
 n2 ( An2  Bn2 )  dx
2 0 2
1  c 2 l 
 n2 ( An2  Bn2 ) l  sin n 
4  2n c 

nc 2 l
Noting that n  , i.e. sin n  sin 2n  0 , the above equation becomes:
l c
1 l 1
n2  ( yn2 ) max dx  ln2 ( An2  Bn2 )
2 0 4
which gives the expected result.

5.11
Expend the expression of y ( x, t ) , we have:
y ( x, t )  A cos(t  kx)  rA cos(t  kx)
 A cos t cos kx  A sin t sin kx  rA cos t cos kx  rA sin t sin kx
 A(1  r ) cos t cos kx  A(1  r ) sin t sin kx
which is the superposition of standing waves.
Chapter 6

6.1
The wave group has a modulation envelope of:

  k 
A  A0 cos t x
 2 2 

where   1  2 is the frequency difference and k  k1  k2 is the wave

number difference. At a certain time t , the distance between two successive zeros of
the modulation envelope x satisfies:
k
x  
2
Noting that k  2  , for a small value of   , we have: k  (2 2 ) , so

the above equation becomes:


2
x  
22
i.e.

 x 

which shows that the number of wavelengths  contained between two successive
zeros of the modulating envelop is   

6.2
The expression for group velocity is given by:
d d dv
vg   (kv)  v  k
dk dk dk
By substitution of the expression of v into the above equation, we have:
sin( ka 2) d  sin( ka 2) 
vg  c k c ka 2 
ka 2 dk 
sin( ka 2) (ka 4) cos(ka 2)  (a 2) sin( ka 2)
2
c  ck
ka 2 (ka 2) 2
sin( ka 2) ka sin( ka 2)
c  c cos  c
ka 2 2 ka 2
ka
 c cos
2
At long wavelength, i.e. k  0 , the limiting value of group velocity is the phase
velocity c .

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
6.3
Noting that the group velocity of light in gas is given on page 124 as:
   r 
Vg  v1  
 2 r  
we have:
   r     r 
Vg  r  v1   r  v  r  
 2 r    2  
 B     B 2 
 v  A  2  D2    A  2  D 
   2    
 B    2B 
 v  A  2  D2     3  2 D 
   2  
 B   B 
 v  A  2  D2     2  D2 
     
 v( A  2 D2 )

6.4

 
2
c2
The relation  r  2  1   e  gives:
v 

 2c 2
2
  2  e2
v

By substitution of v   k , the above equation becomes:

 2  e2  c 2 k 2 (6.4.1)

As   e , we have:

 
2
c2
1  e  1

2
v
i.e. v  c , which means the phase velocity exceeds that of light c .
From equation (6.4.1), we have:
d ( 2 )  d (e2  c 2 k 2 )

i.e. 2d  2kc 2 dk


which shows the group velocity vg is given by:
d k c2 c
vg   c2   cc
dk  v v
i.e. the group velocity is always less than c .

6.5
From equation (6.4.1), we know that only electromagnetic waves of   e can

propagate through the electron plasma media.


For an electron number density ne ~ 1020 , the electron plasma frequency is given by:

ne 10 20
e  e  1.6  10 19   5.65  1011[rad  s 1 ]
me 0 31
9.1 10  8.8  10 12

Now consider the wavelength of the wave in the media given by:
v 2v 2v 2c 2  3  108
      3  10 3[m]
f  e e 5.65  1011

which shows the wavelength has an upper limit of 3 103 m .

6.6

The dispersion relation  2 c 2  k 2  m 2c 2  2 gives

d ( 2 c 2 )  d (k 2  m 2c 2  2 )

2
i.e. d  2kdk
c2
 d
i.e.  c2
k dk
Noting that the group velocity is d dk and the particle (phase) velocity is  k ,

the above equation shows their product is c 2 .

6.7
h h 6.63  10 34
p     1.2  10 9 m
 p 5.4  10 25

6.8
The series in the problem is that at the bottom of page 125. The frequency
components can be expressed as:
sin(  t 2)
R  na cos  t
  t 2

which is a symmetric function to the average frequency 0 . It shows that at

2
t  , R  0 ,  t    2

In k space, we may write the series as:
y ( k )  a cos k1 x  a cos( k1  k ) x    a cos[ k1  (n  1)k ]x

As an analogy to the above analysis, we may replace  by k and t by x , and


2
R is zero at x  , i.e. kx  2
k

6.9
The frequency of infrared absorption of NaCl is given by:
 1 1 
  2  15   
2T 1 1
    27
  27 
 3.608  1013[rad  s 1 ]
a  mNa mCl   23  1.66  10 35  1.66  10 

The corresponding wavelength is given by:

2c 2  3  108
   52[ m]
 3.608  1013
which is close to the experimental value: 61m
The frequency of infrared absorption of KCl is given by:
 1 
  2  15   
2T 1 1 1
   27
 27 
 3.13  1013[rad  s 1 ]
a  mK mCl   39  1.66  10 35  1.66  10 

The corresponding wavelength is given by:

2c 2  3  108
   60[ m]
 3.13  1013
which is close to the experimental value: 71m

6.10

c 3  108
frequency     .75  1018 Hz
 4  10 10

1 1 1
L2   
 2  1.26  5.8  10  .75  1018
344  1018
1
L  m
18  109
1
 L  5  10 11 m i.e.  of  (blue)
10
Chapter 7

7.1
The intensity of sound wave can be written as:
1
I   0 c 2 2
2
where  is the displacement amplitude of an air molecule, so we have:

1 2I 1 2 10
    6.9 105[m]
2 0c 2  500 1.29  330

7.2
The expression of displacement amplitude is given by Problem 7.1, i.e.:

1 2  10 10 I 0 1 2  10  10 10  10 2
    10 10 [m]
2 0c 2  500 1.29  330

7.3
The audio output is the product of sound intensity and the cross section area of the
room, i.e.:
P  IA  100 I 0 A  100  10 2  3  3  10[W ]

7.4
The intensity of sound wave can be written as:
I  P 2  0c

where P is acoustic pressure,  0 is air density, and c is sound velocity, so we

have:
P  I 0 c

so the ratio of the pressure amplitude in water and in air, at the same sound intensity,
are given by:
pwater I (  0 c ) water (  0 c ) water 1.45  10 6
    60
pair I (  0 c ) air (  0 c) air 400

and at the same pressure amplitudes, we have:


I water (  0 c ) air 400
   3  10 4
I air (  0 c ) water 1.45  10 6

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
7.5
If  is the displacement of a section of a stretched spring by a disturbance, which

travels along it in the x direction, the force at that section is given by: F  Y ,
x
where Y is young’s modulus.
The relation between Y and s , the stiffness of the spring, is found by considering
the force required to increase the length L of the spring slowly by a small amount
l  L , the force F being the same at all points of the spring in equilibrium. Thus
 l Y 
 and F   l
x L L
If l  x in the stretched spring, we have:

Y 
F  sx    x and Y  sL .
L
If the spring has mass m per unit length, the equation of motion of a section of
length dx is given by:

 2 F  2
m dx  dx  Y dx
t 2 x x 2

 2 Y  2 sL  2
or  
t 2 m x 2 m x 2

sL
a wave equation with a phase velocity
m

7.6
The Poissons ratio   0.25 gives:

 0.25
2(    )

i.e. 
So the ratio of the longitudinal wave velocity to the transverse wave velocity is given
by:
vl   2   2
   3
vt  
In the text, the longitudinal wave velocity of the earth is 8kms 1 and the transverse
wave velocity is 4.45kms 1 , so we have:
  2 8

 4.45
i.e.   1.23
so the Poissons ratio for the earth is given by:
 1.23
   0.276
2(   ) 2  (1.23   )

7.7
At a plane steel water interface, the energy ratio of reflected wave is given by:
2 2
I r  Z steel  Z water   3.9  10 7  1.43  10 6 
      86%
6 
I i  Z steel  Z water   3.9  10  1.43  10 
7

At a plane steel water interface, the energy ratio of transmitted wave is given by:
It 4 Z ice Z water 4  3.49  106  1.43  106
   82.3%
I i ( Z ice  Z water ) 2 (3.49  106  1.43  106 ) 2

7.8
Solution follow directly from the coefficients at page 153.
r
Closed end is zero displacement with  1 (node).
i

r
Open end:  1 (antinode,  is a max)
i

pr
Pressure: closed end:  1 . Pressure doubles at antinode
pi

pr
Open end :  1 (out of phase – cancels to give zero pressure, i.e. node)
pi

7.9

(a) The boundary condition  0 at x  0 gives:
x
( Ak sin kx  Bk cos kx) sin t x0  0

i.e. B  0 , so we have:   A cos kx sin t


The boundary condition  0 at x  L gives:
x
 kA sin kx sin t xl  0
i.e. kA sin kL sin t  0
2 2l
which is true for all t if kl  n , i.e. l  n or  
 n
The first three harmonics are shown below:


(b) The boundary condition  0 at x  0 gives:
x
( Ak sin kx  Bk cos kx) sin t x0  0

i.e. B  0 , so we have:   A cos kx sin t

The boundary condition   0 at x  L gives:

A sin kx sin t xl  0

i.e. A cos kl sin t  0

 1 2  1 4l
which is true for all t if kl   n   , i.e. l   n   or  
 2   2 2n  1

The first three harmonics are shown below:


0

7.10
The differentiation of equation  2   e2  3aTk 2 gives:

d
2  6aTk
dk
 d
i.e.  3aT
k dk

where a represents Boltzmann constant,  k is the phase velocity, d dk is the

group velocity.

7.11
(a) Since h   , i.e. kh  1 , we have: tanh kh  1 , therefore:

 g Tk  g Tk g Tk gT
v 2     tanh kh   2  2
k   k  k  
i.e. the velocity has a minimum value given by:
4 gT
v4 

g Tk g T
when  , i.e. k 2  or c  2
k  T g
(b) If T is negligible, we have:
g
v2  tanh kh
k
and when   c , k  0 , and for a shallow liquid, h  0 . Noting that when

hk  0 , tanh kh  kh , we have:

g g
v tanh kh  kh  gh
k k
(c) For a deep liquid, h   i.e. tanh kh  1 , the phase velocity is given by:

g g g
v 2p  tanh kh  i.e. v p 
k k k
and the group velocity is given by:
kdv p 1 g g 1 g 1 g
vg  v p   vp  k 3   
dk 2 k k 2 k 2 k
(d) For the case of short ripples dominated by surface tension in a deep liquid, i.e.
g Tk
 and h   , we have:
k 

Tk Tk Tk
v 2p  lim tanh kh  i.e. v p 
h   
and the group velocity is given by:
kdv p Tk k T 3 Tk 3
vg  v p      v
dk  2 k 2  2 p

7.12
Before the source passes by the observer, the source has a velocity of u , the
frequency noted by the observer is given by:
c
1  
cu
After the source passes by the observer, the source has a velocity of  u , the
frequency noted by the observer is given by:
c
2 

cu
So the change of frequency noted by the observer is given by:

 c c  2cu
   2  1      2
c u c u  (c  u 2 )

7.13
By superimposing a velocity of  v on the system, the observer becomes stationary
and the source has a velocity of u  v and the wave has a velocity of c  v . So the
frequency registered by the observer is given by:
cv cv
    
c  v  (u  v) c  u

7.14
Suppose the aircraft is flying at a speed of u , and the signal is being transmitted from
the aircraft at a frequency of  and registered at the distant point at a frequency of
  . Then, the Doppler Effect gives:
c
  
c u
Now, let the distant point be the source, reflecting a frequency of   and the flying
aircraft be the receiver, registering a frequency of   . By superimposing a velocity of
 u on the flying aircraft, the distant point and signal waves, we bring the aircraft to
rest; the distant point now has a velocity of  u and signal waves a velocity of
 c  u . Then, the Doppler Effect gives:
c u cu cu
       
 c  u  ( u ) c c u

which gives:

    15  103
u c c  3  108  750[ms 1 ]
   2   2  3  109

i.e. the aircraft is flying at a speed of 750 m s

7.15
Problem 5.24 shows the Doppler Effect in the format of wavelength is given by:
cu
  
c
where u is the velocity of gas atom. So we have:
u
       
c
i.e.

 2  10 12
u      c  3  108  1 103[ms 1 ]
 6  10 7
The thermal energy of sodium gas is given by:
1 3
mNau 2  kT
2 2
where k  1.38  10 23[ JK 1 ] is Boltzmann’s constant, so the gas temperature is given

by:

mNau 2 23  1.66  1027  1000 2


T   900[ K ]
3k 3  1.38  10 23

7.16
A point source radiates spherical waves equally in all directions.

 vc 
v    : Observer is at rest with a moving source.
 c  u 

 c  v 
v    : Source at rest with a moving observer.
 c 

 c  v 
v    : Source and observer both moving.
 c  u 
Chapter 8

8.1
The integral of magnetic energy over the last quarter wavelength is given by:
2
0 1 0 1  2V0  0 V02 1  cos 4x  L0V02
 4 2   4 2  Z 0
    4 Z 02  
2
L0 I dx L0 cos kx  dx 2 L0 dx
 2 4 Z 02
The integral of electric energy over the last quarter wavelength is given by:

2 1  cos 4x  C0V02


1 1
 
0 0 0
 4 2 0   4 2 0 0   4 0 0  
2 2
C V dx C 2V sin kx dx 2C V dx
2 4

Noting that Z 0  L0 C0 , we have:

0 1 L0V02 C0V02 0 1
 4 2 0   
2
L I dx 2
C0V 2 dx
4Z 0 4  4 2

8.2
The maximum of the magnetic energy is given by:

1  1  2V  
2
2 2L V 2
( Em ) max   L0 I    L0  0
cos kx    0 2 0   2C0V02
2  max  2  Z 0   max Z0

The maximum of the electric energy is given by:

1  1 2
( Ee ) max   C0V 2    C0 2V0  sin kx    2C0V02
2  max  2  max
The instantaneous value of the two energies over the last quarter wavelength is given
by:
2
1  2V  1
( Em  Ee )i  L0  0  cos kx   C0 (2V02 sin kx) 2
2  Z0  2
 2C0V02 cos 2 kx  2C0V02 sin 2 kx
 2C0V02
So we have:
( Em ) max  ( Ee ) max  ( Em  Ee )i  2C0V02

8.3
For a real transmission line with a propagation constant  , the forward current wave

I x at position x is given by:

I x   I 0 e  x  Ae x

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
where I 0  A is the forward current wave at position x  0 . So the forward voltage

wave at position x is given by:


Vx  Z 0 I x  Z 0 Ae x

The backward current wave I x at position x is given by:

I x   I 0  e  x  Be  x

where I 0   B is the backward current wave at position x  0 . So the backward

voltage wave at position x is given by:


Vx    Z 0 I x    Z 0 Be  x

Therefore the impedance seen from position x is given by:

Vx  Vx Z 0 Ae x  Z 0 Be  x Ae x  Be  x


Zx    Z
I x  I x Ae x  Be x Ae x  Be x
0

If the line has a length l and is terminated by a load Z L , the value of Z L is given

by:
VL Vl   Vl  Ae l  Be  l
ZL    Z0
I L Il  Il Ae l  Be l

8.4
The impedance of the line at x  0 is given by:
 Ax  Bex  A B
Z i   Z 0 x   Z0
x 
 Ae  Be  x0 A B

Noting that:

Ae l  Bel
Z L  Z0
Ae l  Bel
we have:
( Z 0  Z L ) Ae l  ( Z 0  Z L ) Be l

i.e.
A ( Z 0  Z L ) 2l
 e
B (Z0  Z L )

so we have:
A B 1 Z (el  e l )  Z L (el  e l ) Z sinh l  Z L cosh l
Zi  Z0  Z 0 0 l l l l
 Z0 0
A B 1 Z 0 ( e  e )  Z L (e  e ) Z 0 cosh l  Z L sinh l

8.5
If the transmission line of Problem 8.4 is short-circuited, i.e. Z L  0 , The expression

of input impedance in Problem 8.4 gives:


Z 0 sinh l
Z sc  Z 0  Z 0 tanh l
Z 0 cosh l

If the transmission line of Problem 8.4 is open-circuited, i.e. Z L   , The expression

of input impedance in Problem 8.4 gives:


Z L cosh l
Z oc  Z 0  Z 0 coth l
Z L sinh l

By taking the product of these two impedances we have:


Z sc Z oc  Z 02 , i.e. Z 0  Z sc Z oc

which shows the characteristic impedance of the line can be obtained by measuring
the impedances of short-circuited line and open-circuited line separately and then
taking the square root of the product of the two values.

8.6
The forward and reflected voltage waves at the end of the line are given by:
Vl   Vl   V0 e  ikl

where V0 is the forward voltage at the beginning of the line. So the reflected voltage

wave at the beginning of the line is given by:


V0  Vl  e  ikl  V0 e  i 2 kl

The forward and reflected current waves at the end of the line are given by:
I l   Vl  Z 0  V0 e ikl Z 0  I 0 e  ikl
I l    Vl  Z 0  Vl  Z 0  I 0 e ikl

where I 0 is the forward current at the beginning of the line. So the reflected current

wave at the beginning of the line is given by:


I 0  I l  e  ikl  I 0 e  i 2 kl
Therefore the input impedance of the line is given by:
V0  V0 V0 (1  e i 2 kl ) V0 (eikl  e  ikl ) sin kl L 2l
Zi     iZ 0  i 0 tan
i 2 kl
I 0  I 0 I 0 (1  e ) I 0 (e  e )
ikl ikl
cos kl C0 

The variation of the ratio Z i L0 C0 with l is shown in the figure below:

… …

8.7
Analysis in Problem 8.6 shows the impedance of a short-circuited loss-free line has an
impedance given by:
2l
Z i  iZ 0 tan

so, if the length of the line is a quarter of one wavelength, we have:
2  
Z i  iZ 0 tan  iZ 0 tan  
 4 2
If the this line is bridged across another transmission line, due to the infinite
impedance, the transmission of fundamental wavelength  will not be affected.
However for the second harmonic wavelength  2 , the impedance of the bridge line

is given by:
2 
Z i  iZ 0 tan  iZ 0 tan   0
 24

which shows the bridge line short circuits the second harmonic waves.

8.8
1
For Z 0 to act as a high pass filter with zero attenuation, the frequency  2  ,
2 LC
where Z 0  L C .

The exact physical length of Z 0 is determined by  . Choosing the frequency 1

determines k1  2 1 .

For a high frequency load Z L and a loss- free line, we have, for the input

impedance:
 Z cos kl  iZ 0 sin kl 
Z in  Z 0  L 
 Z 0 cos kl  iZ L sin kl 
For n even, we have:
2 n1
cos k1l  cos  cos n  1
1 2
For n odd, we have:
2 n1
cos k1l  cos  cos n  1
1 2

The sine terms are zero.


So Z in  Z L for n odd or even, and the high frequency circuits, input and load, are

uniquely matched at 1 when the circuits are tuned to 1 .

8.9
The phase shift per section  should satisfy:

Z1 i L  2 LC
cos   1   1  1
2Z 2 2 i C 2

i.e.

 2 LC
1  cos  
2
i.e.

  2 LC
2 sin 2

2 2
For a small  ,   sin  , so the above equation becomes:
   2 LC
2

2  
2 2
i.e.

   LC   v  k

where the phase velocity is given by v  1 LC and is independent of the

frequency.

8.10

k  L0C 0 L0 L0
  Q
2 R0 C0 R0

8.11
R0 G0
Suppose   K , where K is constant, the characteristic impedance of a
L0 C0

lossless line is given by:


R0  iL0 KL0  iL0 L0
Z0   
G0  iC0 KC0  iC0 C0

which is a real value.

8.12
Try solution    m ex in wave equation:

 2 8 2 m
 2 ( E  V )  0
x 2 h
we have:

8 2 m
2  (V  E )
h2

For E  V (inside the potential well), the value of  is given by:

2
 in  i
2 m( E  V )
h
So the  has a standing wave expression given by:
2 2
i m (V  E ) x i m (V  E ) x
  Ae h
 Be h
where A , B are constants.
For E  V (outside the potential well), the value of  is given by:

2
 out   2m(V  E )
h
So the expression of  is given by:
2 2
m (V  E ) x  m (V  E ) x
  Ae h  Be h

where A , B are constants. i.e. the x dependence of  is e  x , where

2
  2m(V  E )
h
Chapter 9

9.1
Write the expressions of E x and H y as:

2
E x  E0 sin (v t  z )
E E
2
H y  H 0 sin (v t  z )
H H

where E and H are the wavelengths of electric and magnetic waves respectively,

and vE and vH are the velocities of electric and magnetic waves respectively.

By substitution of the these expressions into equation (9.1a), we have:


2 2 2 2
 vH H 0 cos (v H t  z )   E0 cos (v E t  z )
H H E E

vH H 0 2 E0 2
i.e.  cos (v H t  z )  cos (v E t  z )
H H E E
which is true for all t and z , provided:
E0
v H  vE 
H 0

and  H  E

so, at any t and z , we have:


E0
E   H  tz
H 0

Therefore E x and H y have the same wavelength and phase.

9.2
Energy Force  Distance Force
   pressure
Volume L3 Area

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
Currents in W into page. Field lines at A cancel. Those at C force wires together.
Reverse current in one wire. Field lines at A in same direction, force wires apart.
Fig Q.9.2.a

Motion

Field lines at C in same direction as those from current in wire –


in opposite direction at A . Motion to the right
Fig Q.9.2.b

9.3
The volume of a thin shell of thickness dr is given by: 4r 2 dr , so the electrostatic
energy over the spherical volume from radius a to infinity is given by:
 1
a 2  0 E (4r )dr , which equals mc , i.e.:
2 2 2

 1
a 2  0 E (4r )dr  mc
2 2 2

By substitution of E  e 4 0 r 2 into the above equation, we have:

 1 e2
a
0
2 (4 0 r )
2 2
(4r 2 )dr  mc 2

e2  1
i.e.
8 0 
a r 2
dr  mc 2

e2
i.e.  mc 2
8 0 a

Then, the value of radius a is given by:


e2 (1.6  10 19 ) 2
a   1.41 10 15 [m]
8 0 mc 2 8  8.8  10 12  9.1 10 31  (3  108 ) 2

Another approach to the problem yields the value:

a  2.82  10 15 [m]


9.4
The magnitude of Poynting vector on the surface of the wire can be calculated by
deriving the electric and magnetic fields respectively.
The vector of magnetic field on the surface of the cylindrical wire points towards the
azimuthal direction, and its magnitude is given by Ampere’s Law:
I
H  H e  e
2r
where r is the radius of the wire’s cross circular section, and I is the current in the
wire.
Ohm’s Law, J  E , shows the vector of electric field on the surface of the
cylindrical wire points towards the current’s direction, and its magnitude equals the
voltage drop per unit length, i.e.:
V IR
E  Eze z  e z  e z
l l
where, l is the length of the wire, and the V is the voltage drop along the whole
length of the wire and is given by Ohm’s Law: V  IR , where R is the resistance of
the wire.
Hence, the Poynting vector on the surface of the wire points towards the axis of the
wire is given by:
S  E  H  E z e z  H  e   E z H e r

which shows the Poynting vector on the surface of the wire points towards the axis of
the wire, which corresponds to the flow of energy into the wire from surrounding
space. The product of its magnitude and the surface area of the wire is given by:
IR I
S  2rl  Ez H  2rl  2rl  I 2 R
l 2r
which is the rate of generation of heat in the wire.

9.5
For plane polarized electromagnetic wave ( Ex , H y ) in free space, we have the

relation:
Ex 0

Hy 0
Its Poynting vector is given by:
Ex 0 2 1
S  Ex H y  Ex  Ex   0 E x2  c 0 E x2
0  0 0 0 0

where c 1 0 0 is the velocity of light.


Noting that:
2
1 1    1
 0 E x2   0  0 H y   0 H y2
2 2  0  2
we have:

1 1 
S  Ex H y  c  0 Ex2  0 H y2   c 0 E x2
2 2 
Since the intensity in such a wave is given by:
1
I  S av  c 0 E 2  c 0 Emax
2

2
we have:
1
S  3  108  8.8 1012  Emax
2
 1.327  103 Emax
2

2
2 12 2
Emax  S  S 1 2  27.45S 1 2 [Vm 1 ]
c 0 3  10  8.8  10
8 12

0 2 12 2
H max  Emax  S  S 1 2  7.3  10 2 S 1 2 [ Am 1 ]
0 c0 3  10  4  10
8 -7

9.6
The average intensity of the beam and is given by:
Power Energy 0.3
I    1.53  108[Wm  2 ]
area area  pulse duration   (2.5  10 3 ) 2  10 4

Using the result in Problem 8.6, the root mean square value of the electric field in the
wave is given by:

I 1.53  108
E2    2.4  105[Vm 1 ]
c 0 3  10  8.8  10
8 12

9.7
Using the result of Problem 9.6, the amplitude of the electric field at the earth’s
surface is given by:

E0  27.45S 1 2  27.45  1350  1010[Vm 1 ]

and the amplitude of the associated magnetic field in the wave is given by:

H 0  7.3  10 2  1350  2.7[ Am 1 ]

The radiation pressure of the sunlight upon the earth equals the sum of the electric
field energy density and the magnetic field energy density, i.e.
1 1
prad   0 E02  0 H 02   0 E02  8.8 1012 10102  8.98 106 [ Pa]
2 2

9.8
The total radiant energy loss per second of the sun is given by:
Eloss  S  4r 2  1350  4  (15  1010 ) 2  3.82  1026 [ J ]

which is associated with a mass of:

3.82  10 26
m  Eloss c 2   4.2  109 [kg ]
(3  108 ) 2

9.9
At a point 10km from the station, the Poynting vector is given by:

P 105
S   1.6  10 4 [W m 2 ]
2r 2
2    (10  10 )
3 2

Using the result in Problem 8.6, the amplitude of electric field is given by:

E0  27.45  S 1 2  27.45  1.6 104  0.346[V m]


The amplitude of magnetic field is given by:

H 0  7.3 102 S 1 2  7.3  102  1.6  104  9.2 104 [ A m]

9.10
Analysis in page 195 and 196 shows, in a conducting medium, the wave number of
electromagnetic wave is given by:


k
2
where  is angular frequency of the electromagnetic wave,  and  are the
permeability and conductivity of the conducting medium.
Differentiation of the above equation gives:

1 2  1 
dk  d  d
2  2 2 2

d 2 2 
i.e. 2 2 2
dk   k
which shows, when a group of electromagnetic waves of nearly equal frequencies
propagates in a conducting medium, where the group velocity and the phase velocity
can be treated as fixed values, the group velocity, vg  d dk , is twice the wave
velocity, v p   k .

9.11
  0.1
(a)    36 109  720  100
 2 r  0 2  50 10  50
3

which shows, at a frequency of 50kHz , the media is a conductor


  0.1
(b)    36 109  3.6 103  102
 2 r  0 2 10  10  50
4 6

which shows, at a frequency of 104 MHz , the media is a dielectric.

9.12
The Atlantic Ocean is a conductor when:
 
  100
 2 r  0

 4.3
i.e.    36 109  10[ MHz]
2  100   r  0 2 100  81

Therefore the longest wavelength that could propagate under water is given by:
v c
   10  106
max  r max

c 3  108
i.e. max    3[m]
 r  10  106 81  10  10 6

9.13
Analysis in page 205 and 206 shows that when an electromagnetic wave is reflected
normally from a conducting surface its reflection coefficient I r is given by:

2 0
I r  1 2

Noting that  r  1 , the fractional loss of energy is given by:

 2 0  8 0 8   r 8


1  I r  1  1  2 
  
     
9.14
Noting that the relation between refractive index n of a dielectric and its impedance
Z0
Z d is given by: n  , where Z 0 is the impedance in free space, so, when light
Zd

travelling in free space is normally incident on the surface of a dielectric, the reflected
intensity is given by:
2 2 2
E   Zd  Z0   1  Z0 Zd   1  n 
2

I r   r          
 Ei   Zd  Z0   1  Z0 Zd   1  n 
and the transmitted intensity is given by:
2 2
Z E 2 Z  2Z d  Z  
2

  n
2 2  4n
I t  0 t2  0    0   
Z d Ei Zd  Zd  Z0  Zd  1  Z0 Zd  1 n  (1  n) 2

9.15
If the dielectric is a glass (nglass  1.5) , we have:

2
 1  nglass   1  1.5  2
    4%
 1  n   1  1.5 
I r _ glass
 glass 
4nglass 4  1.5
I t _ glass    96%
(1  nglass ) 2
(1  1.5) 2

Problem 9.12 shows water is a conductor up to a frequency of 10MHz, i.e. water is a


dielectric at a frequency of 100MHz and has a refractive index of:

nwater   r  81  9

So, the reflectivity is given by:


2
 1  nwater   1  9 
2

I r _ water        64%
 1  nwater   1  9 
and transmittivity given by:
4nwater 49
I t _ water    36%
(1  nwater ) 2
(1  9) 2

9.16
The loss of intensity is given by:
I loss  1  I t1 I t 2

where I t1 is the transmittivity from air to glass and I t 2 is the transmittivity from
glass to air. Following the discussion in problem 9.14, we have:
2 2 2
Z i Et2 Z d  2Z 0  Z  2  1 2  4n
It 2      d        I t1
Z t Ei2 Z 0  Z 0  Z d  Z0  1  Zd Z0  n 11 n  (1  n) 2
So we have:
I loss  1  I t21  1  0.962  7.84%

9.17
Noting that c  1 0 0   2 , the radiating power can be written as:

dE q 2 4 x02
P 
dt 12 0c 3
 2 x02
  2q 2
12 0c  c 2

4 2 2 x02
 0 0 I 02
12 0 
2 2

1 2 0  x0  2
2

    I
2 3 0    0

2 0  x0 
2 2
x 
i.e. R    787 0  []
3 0    
By substitution of given parameters, the wavelength is given by:

3  108
   600[m]  x0  30[m]
c 5  105
So the radiation resistance and the radiated power are given by:
2 2
x   30 
R  787   0   787     1.97[]
  600 
1 2 1
P RI 0  1.97  202  400[W ]
2 2
Chapter 10

10.1

v 2  S / p  2 / 0.1  20  4.5 m  s -1

10.2

2 z 2 z
Substituting the expression of z into  , we have:
x 2 y 2

2 z 2 z
  (k12  k22 ) Aei[t ( k1xk2 y )]  (k12  k22 ) z
x 2 y 2

Noting that k 2   2 c 2  k12  k 22 , we have:

2 z 2 z 2
   z
x 2 y 2 c2

1 2 z
Substituting the expression of z into , we have:
c 2 t 2

1 2 z  2 [t ( k1xk2 y )] 2
  Ae   z
c 2 t 2 c2 c2
So we have:

2 z 2 z 1 2 z
 
x 2 y 2 c 2 t 2

10.3
Boundary condition z  0 at y  0 gives:

A1ei (t k1x )  A2ei (t k1x )  0 i.e. A1   A2

so the expression of z can be written as:

z  A1{ei[t ( k1xk2 y )]  ei[t ( k1xk2 y )]}  A1[ei (t k1x ) (e ik2 y  eik2 y )]  2iA1 sin( k 2 y )ei (t k1x )

Therefore, the real part of z is given by:


zreal  2 A1 sin k 2 y sin(t  k1 x)

Using the above expression, boundary condition z  0 at y  b gives:

z  2iA1 sin k2bei (t k1x )  0

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
n
which is true for any t and x , provided: sin k2b  0 , i.e. k 2  .
b

10.4
Using boundary condition z  0 at x  0 in the displacement equation gives:
( A1  A4 )ei (t k2 y )  ( A2  A3 )ei (t k2 y )  0
which is true for any t and y if:

A1   A4 and A2   A3

so we have:
z  A1{ei[t ( k1xk2 y )]  ei[t (  k1xk2 y )]}  A2{ei[t ( k1xk2 y )]  ei[t (  k1xk2 y )]
 2 A1i sin k1 xei (t k2 y )  2 A2i sin k1 xei (t k2 y ) (10.4.1)
 2i sin k1 x[ A1ei (t k2 y )  A2ei (t k2 y ) ]
Using boundary condition z  0 at y  0 in equation (10.4.1) gives:

 2i sin k1 x( A1  A2 )eit  0

which is true for any t and x if:


A1  A2

Therefore, equation (10.4.1) becomes:

z  4 A1 sin k1 x sin k 2 yeit

and the real part of z is given by:


zreal  4 A1 sin k1 x sin k2 y cos t (10.4.2)

Using boundary condition z  0 at x  a in equation (10.4.2) gives:


n
sin k1a  0 , i.e. k1  1 , where n1  1,2,3, .
a
Using boundary condition z  0 at y  b in equation (10.4.2) gives:
n
sin k 2b  0 , i.e. k2  2 , where n2  1,2,3, .
b

10.5
As an analogy to discussion in text page 212, electric field Ez between these two

planes is the superposition of the incident and reflected waves, which can be written
as:
i[( k x x  k y y ) t ] i [(  k x x  k y y ) t ]
E z  E1e  E2 e

where k x  k cos  and k y  k sin 


Boundary condition Ez  0 at x  0 gives:

i ( k y y t )
( E1  E2 )e 0

which is true for any t and y if E1   E2  E0 , so we have:

i[( k x x  k y y )t ] i[(  k x x k y y )t ] i ( k y y t )


Ez  E0e  E0e  E0 (eikx x  e ikx x )e

Using the above equation, boundary condition Ez  0 at x  a gives:

i ( k y y t )
Ez  E0 (eikxa  eikxa )e 0

i ( k y y t )
i.e. sin k x ae 0

which is true for any t and y if sin k x a  0 , i.e. k x  n a .

1 1
By substitution of the expression of c and  g into  , we have:

2
c 2g

2
1 1  kx   k y 
2
k x2  k y2 k2   
2
1
 2           2
c g  2   2 
2
( 2 ) 2
( 2 ) 2
 2c  0

10.6
sin my sin nz sin 
Ex  A t  k x x 
a b cos 

10.7

2  m2 n2  2
k  2  k  2    2  2 
2
x
2 2

c c a b 
1
m n  2 2 2
   c  2  2 
a b 
that is when m  n  1

10.8
Rearrange the dispersion relation:

2
k2  2
 k x2
c
gives:
 2  k 2  k x2 c 2

Applying partial differentiation of k x on the above equation gives:

 2 
k x
 
k x
k 2  k x2 c 2

i.e.

2  2c 2 k x
k x

i.e.
 
 c2
k x k x

10.9
As shown in page 190, the energy per unit volume between the parallel conducting
1
plates is given by: E02 . The cross section of the electromagnetic wave travel
2
1
through has an area of ab , and the electromagnetic wave travels at a velocity: .


So the power transmitted is given by:


1 2 1 1 
P E 0  ab   abE 02
2  2 

10.10

2me  18.2  10 31 kg.  p  18.2  1.6  10 50 


1/ 2
 5.4  10 25

dB  h / p  1.22  10 9 m

10.11

Expanding e h / kT  1 for small h gives 1  h / kT  1 in the denominator, that

is kT / h in the numerator.

10.12
The most sensitive wavelength to the human eye can be given by substituting the
sun’s temperature T  6000[ K ] into equation ch m  5kT , i.e.:

ch 3 108  6.63 10 34


m    4.7  10 7 [m]
5kT 5 1.38  10 23  6000
which is in the green region of the visible spectrum.

10.13
Substituting the tungsten’s temperature T  2000[ K ] into equation ch m  5kT ,

i.e.:

ch 3 108  6.63 10 34


m   23
 14 10 7 [m]
5kT 5 1.38 10  2000
which is well into infrared.
Chapter 11

11.1
The wave form in the upper figure has an average value of zero and is an odd function
of time, so its Fourier series has a constant of zero and only sine terms. Since the
wave form is constant over its half period, the Fourier coefficient bn will be zero if

n is even, i.e. there are only odd harmonics and the harmonics range from 1,3,5 to
infinity.
The wave form in the lower figure has a positive average value and is a even function
of time, so its Fourier series has a constant of positive value and only cosine terms.
Since  T  1 2 , there are both odd and even harmonics. The harmonics range from

1,2,3 to infinity.

11.2
Such a periodic waveform should satisfy: f ( x)   f ( x  T 2) , where T is the

period of the waveform. Its Fourier coefficient of cosine terms can be written as:
2 T 2nx
an 
T 0
f ( x) cos
T
dx

2 T2 2nx T 2nx 
   f ( x) cos dx   f ( x) cos dx 
T 0 T T 2 T 
2 T2 2nx T 2nx 
   f ( x) cos dx    f ( x  T 2) cos d ( x  T 2) 
T 0 T T 2 T 
If n is even, we have
2n( x  T 2)  2nx  2nx
cos  cos  n   cos
T  T  T

Hence, by substituting into an and using u  x  T 2 , we have:

2 T2 2nx T 2 2nu 
an  
T0 f ( x) cos
T
dx   f (u ) cos
0 T
du   0

Similarly, the coefficient of sine terms is given by:
2 T 2nx
bn 
T 0 f ( x) sin
T
dx

2 T2 2nx T 2nx 
   f ( x) sin dx   f ( x) sin dx 
T0 T T 2 T 
2 T2 2nx T 2nx 
   f ( x) sin dx    f ( x  T 2) sin d ( x  T 2) 
T0 T T 2 T 

If n is even, using u  x  T 2 , we have:

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
2 T2 2nx T 2 2nu 
bn  
T0 f ( x) sin
T
dx   f (u ) sin
0 T
du   0

Therefore, if n is even, the Fourier coefficients of both cosine and sine terms are
zero, i.e. there are no even order frequency components.

11.3
A phase advance of  / 2 in the series of Figure 11.1 gives:
4h    1   1   1   
f ( x)   sin x    sin 3 x    sin 5 x    sin 7 x   
   2 3  2 5  2 7  2 
4h  1 1 1 
  cos x  cos 3 x  cos 5 x  cos 7 x 
  3 5 7 
which is the series in Figure 11.2

11.4
The constant term of the Fourier series is given by:
1 1 2 1  h
2
a0 
2 0
ydx 
2 0
h sin xdx 

The Fourier coefficient of cosine term is given by:
1 2 h 
an   y cos nxdx   sin x cos nxdx
 0  0

when n  1 , we have:
h  h 

 
a1  sin x cos xdx  sin 2 xdx  0
0 2 0

when n  1 , we have:
h 
an 
  0
sin x cos nxdx

h 

2 
0
sin(1  n) x  sin(1  n) xdx

h  1 1 
  cos(1  n) x  cos(1  n) x 
2 1  n 1 n 0
which gives:
h 2 h 2 h 2
a2   , a3  0 , a4   , a5  0 , a6   ,…
 1 3  35  57
The Fourier coefficient of sine term is given by:
1 2 h 
bn   y sin nxdx   sin x sin nxdx
 0  0

when n  1 , we have:
h  h  h
 
b1  sin x sin xdx  sin 2 xdx 
0 0 2
when n  1 , we have:
h 


bn  sin x sin nxdx
0

h 

2 
0
cos(1  n) x  cos(1  n) xdx

h  1 1 
 1  n sin(1  n) x  1  n sin(1  n) x 
2 0

0
Overall, the Fourier series is given by:
 
1
y a0   an cos nx   bn sin nx
2 1 1

h  2 2 2 
 1  sin x  cos 2 x  cos 4 x  cos 6 x 
  1 2 1 3 35 57 

11.5
Such a wave form is a even function with a period of  . Hence, there are only
constant term and cosine terms.
The constant term is given by:
1 1  2h
a0   h sin xdx 
2  0 
which doubles the constant shown in Problem 11.4
The coefficient of cosine term is given by:
4h  2 2nx
an 
  0
sin x cos

dx

4h  2

  0
sin x cos 2nxdx

2h  2

  0
[sin(1  2n) x  sin(1  2n) x]dx
 2
2h  1 1 
  cos(1  2n) x  cos(1  2n) x 
 1  2n 1  2n 0
which gives:
h 2 h 2 h 2
a1   , a2   , a3   ,…
 1 3  35  5 7
Therefore the Fourier series is given by:
1 
2nx
y a0   an cos
2 1 
h 2 2 2 
 1  cos 2 x  cos 4 x  cos 6 x 
  1 3 35 57 
h
Compared with Problem 11.4, the modulating ripple of the first harmonic sin x
2
disappears.
11.6
Change every t in the half wave series to t    and add the resulting series to the
half-wave series.

11.7
V q CV V 2t
I    
R t 2t V RC

11.8

Total energy  2Td 2 / l . Equation 11.3 and 11.4 give

16d 2T  1 1  1.87Τd 2 1.87


E1  E 2  E3   1    i.e.  93.5%
 2 l  32 52  l 2

11.9
f (x) is even function in the interval   , so its Fourier series has a constant term

given by:
1 1  1  2
2
a0 
2  
f ( x)dx 
2 

x 2 dx 
3
The coefficient of cosine term is given by:
2  2nx

an  f ( x) cos
dx
0 
2  2  2
  x 2 cos nxdx 
n 0
x d sin nx
 0
2  2  4 
x sin nx   sin nxdx 2   2  xd cos nx


n  0 0  n  0

 2  x cos nx 0   cos nxdx   2 cos n  (1) n 2
4  4 4
n  0 
 n n
Therefore the Fourier series is given by:
1 
2nx 1 2  4
f ( x)  a0   an cos     (1) n 2 cos nx
2 1 2 3 1 n

11.10
The square wave function of unit height f (x) has a constant value of 1 over its first

half period [0,  ] , so we have:

f ( 2)  1
By substitution into its Fourier series, we have:
4  1 3 1 5 
f ( 2)   sin  sin  sin  1
 2 3 2 5 2 
i.e.
1 1 1 
1   
3 5 7 4

11.11
It is obvious that the pulse train satisfies f (t )  f (t ) , i.e. it is an even function. The

cosine coefficients of its Fourier series is given by:



4  2nx 4 T 2nx 2 2
an   cos dx   sin  sin n
T 0 T T 2n T 0 n T

11.12
2 2
As  becomes very small, sin n  n , so we have:
T T
2 2 2 2 4
an  sin n   n 
n T n T T
We can see as   0 , an  0 , which shows as the energy representation in time

domain  0 , the energy representation in frequency domain  0 as well.

11.13
The constant term of the Fourier series is given by:
1 1 T2 1 1  1 1
a0   dt   dt 
2 T T 2 2 T 2
 T
The coefficient of cosine term is given by:

4  1 2nt 4 1 T 2nt 1 2n
an  
T 0 2
cos
T
  
T 2 2n
sin 
T 0 n
sin
T
As   0 , we have:
2n
1 1 2n 2
an  
sin  
n T n T T
Now we have the Fourier series given by:
1 
2nt 1 2  2nt
f (t )  a0   an cos    cos
2 n1 T T T n1 T
Chapter 12

12.1

Fig.A.12.1

As shown in Fig.A.12.1, the air gap thickness t is given by:


r2 r2
t  t2  t1  
2 R2 2 R1
Noting that there is a  rad of phase shift upon the reflection at the lower surface of
the air gap, the thickness of air gap at dark rings should satisfy:
2t  n
r2 r2
i.e.   n
R2 R1

which yields the radius rn of the nth dark ring given by:

R1 R2 n
rn2 
R1  R2

12.2
The ring system shrinks towards the centre because the radius of each ring is reduced

by n ( r 2 of the ring  nt where t is the film thickness). So the new radius of the
2.6
40th ring is r / n   2 cms .
1.3

12.3

(a) Angular separation  5  10 4 radians to give linear fringe separation 0.5mm.

(b) Approximately  /   500 / 40  25 fringes


(c) Angle subtended by source to Young’s slits must be << fringe angular separation

i.e.  5  10 4 radians giving say 2  10 2 mm .

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
12.4
As shown in page 274 of the text, the intensity distribution of the interference pattern
is given by:

I  4a 2 cos 2
2
where  is the phase difference between the two waves transmitted from the two
radio masts to a point P and is given by:
2 2
  kf sin   f sin    400  sin   4 sin 
 3  10 1500  103
8

so we have:
4 sin 
I  4a 2 cos2  2 I 0 [1  cos(4 sin  )]
2
where I 0  a 2 represents the radiated intensity of each mast.

The intensity distribution is shown in the polar diagram below:

Fig.A.12.4

12.5
(a)
Analysis is the same as Problem 12.3 except:
2 
   0  kf sin      sin      sin 
 2
Hence, the intensity distribution is given by:
      sin   2   sin  
I  4a 2 cos 2    4a 2 cos 2    4 I s sin  
2  2   2 

where I s  a 2 is the intensity of each source.


The polar diagram for I versus  is shown below:

Fig.A.12.5(a)

(b)
In this case, the phase difference is given by:
 2     sin 
   0  kf sin     sin  
2  4 2
Hence, the intensity distribution is given by:
    sin    
I  4a 2 cos 2  4a 2 cos 2  4 I s cos 2 (1  sin  )
2 4  4 

where I s  a 2 is the intensity of each source.


The polar diagram for I versus  is shown below:
Fig.A.12.5(b)

12.6
Finesse   R /(1  R )  29.8   FSR  29.8  m  8.94  10 10 m .
The number of beams is the Finesse  29.8  30 .

12.7
(a)

Fig.A.12.7(a)

Fig.A.12.7(a) shows elements of a vertical column and a horizontal row of radiators in


a rectangular lattice with unit square cells of side d . Rays leave each lattice point at
an angle  to reach a distant point P . If P is simultaneously the location of the
mth spectral order of interference from the column radiation and the nth spectral
order of interference from the row radiation, we have from pages 281/2 the relations:
d sin   m and d cos   n
Thus
sin  m
 tan  
cos n
where m and n are integers.

(b)

Fig.A.12.7(b)

Waves scattered elastically (without change of  ) by successive planes separated by


a distance d in a crystal reinforce to give maxima on reflection when the path
difference 2d sin   n . In Fig.A.12.7(b), the path difference ABC between the
incident and the reflected rays  2d sin  .

12.8
Using the Principal Maximum condition:
f sin   n

at    , we have: f  n , which shows the minimum separation of equal
2
sources is given by: f   .

When N  4 , the intensity distribution as a function of  is given by:


sin 2 (4 sin  )
I  Is
sin 2 ( sin  )
The N  1  3 points of zero intensity occur when:
  3
f sin   , ,
4 2 4
1 1 3
i.e. sin   , ,
4 2 4
The position of the N  2  2 points of secondary intensity maxima should occur
between the zero intensity points and should satisfy:
dI
0
d
dI d  sin 2 (4 sin  ) 
 0
d d  sin 2 ( sin  ) 
i.e. Is

i.e. 6 cos 2 ( sin  )  1  0

1  1 
which yields: sin   arccos  
  6

i.e. secondary intensity maxima occur when   21.5 and   39.3 .


The angular distribution of the intensity is shown below:

12.9
The angular width of the central maximum  is the angular difference between +1
and -1 order zero intensity position and should satisfy:
2 2  0.21
sin     1.875  10  3 or   6
Nf 32  7
The angular separation between successive principal maxima  is given by:
 0.21
sin     0.03 or   1 42
f 7

12.10
2
Total phase of main beam remains zero.   0   / 3   f sin 

 sin   1 / 18  0.055 or   3 09'
Chapter 13

13.1
(a) and (b) are both solved by   1.22 / d .
1
(a)   1.22  .21 / 76  radians  0.2 
300
(b)   1.22  550  10 9 / 2.4  2.8  10 7 rads

13.2


90   
90 
120 60 120 60 

150  30  150  30 

180  0  180  0

210  330  210  330 

240  300  240  300 


 
270 270

Fig.A.13.2

The above polar diagrams show the traces of the tip of the intensity of diffracted light
I for monochromatic light normally incident on a single slit when the ratio of slit
width to the wavelength d  changes from 1 to 4. It is evidently shown that the

polar diagram becomes concentrated along the direction   0 as d  becomes

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
larger.

13.3
It is evident that   0 satisfies the condition:   tan  .
By substitution of   3 2   into the condition:   tan  we have:

3 2    tan(3 2   )

i.e. 3 2    cot 

i.e. (3 2   ) sin   cos 

when  is small, we have:

2
(3 2   )  1 
2
The solution to the above equation is given by:   0.7 .
Using the similar analysis for   5 2   and   7 2   , we can find

  0.041 and   0.029 respectively. Therefore the real solutions for  are
  0,1.43 ,2.459 ,3.471 , etc .

13.4
If only interference effects are considered the intensity of this grating is given by:
sin 2 3
I  I0
sin 2 

The intensity of the principal maximum is given by: I max  9I 0 when   0 .

The  for the secondary maximum should satisfy:

d  sin 2  
   0
d  sin 
2

i.e. sin 2   1

i.e.  sec_ max  (2n  1) , where n is integer
2
Hence, at the secondary maximum:
sin 2 3 sec_ max 1
I sec_ max  I 0  I0  I max
sin  sec_ max
2
9
13.5
Suppose a monochromatic light incident on a grating, the phase change d required

to move the diffracted light from the principal maximum to the first minimum is given
by:
 f sin   f f  
d  d   d (sin  )   
     Nf N
Since N is a very large number, we have:

d 
0
N
Then, suppose a non-monochromatic light, i.e.  is not constant, incident on the
same grating, the phase change d required to move the diffracted light from the

principal maximum to the first minimum should be the same value as given above, so
we have:
 f sin   f 1
d  d   d (sin  )  f sin d  
    
f f sin  
 cosd  d  0
 2
N
which gives:
d  (nN cot  ) 1

13.6
(a)
The derivative of the equation:
f sin   n

gives:
f cosd  nd

when  is a small angle we have:


d n

d f
When the diffracted light from the grating is projected by a lens of focal length F
on the screen, the relation between linear spacing on the screen l and the diffraction
angle  is given by:
l  F
Its derivative over  gives:
dl d nF
F 
d d f
(b)
Using the result given above, the change in linear separation per unit increase in
spectral order is given by:
dl Fd 2  (5.2  10 7  5  10 7 )
  6
 2  10 2 [m]
n f 2  10

13.7
(a)
Using the resolving power equation:

 nN
d
we have:
 (5.89  10 7  5.896  10 7 ) 2
N   328
nd 3  (5.896  10  7  5.89  10  7 )
(b)
Using the resolving power equation:

 nN
d
we have:
 6.5  107
d    2.4  10 12 [m]
nN 3  9  10 4

13.8
When the objects O and O are just resolved at I and I  the principal
maximum of O and the first minimum of O are located at I . Rayleigh’s criterion
thus defines the path difference:
OBI  OAI  OB  OA  1.22 ( BI  AI )

Also OB  OA giving
(OB  OB)  (OA  OA)  1.22

Fig.A.13.8 shows OA parallel to OA and OB parallel to OB , so:


OA  OA  OC  OO sin i  s sin i
and
OB  OB  OC   OO sin i  s sin i
We therefore write:
1.22 1.22
s or s  if i  45
2 sin i 2 sin i
Fig.A.13.8

13.9
The darkest point is point 3 on Figure 13.23 and 13.24 where the second half zone
almost cancels the first bright spot. Using the relation A  Rn2  nr0 where n is
the number of zones. We take n  2 so that we have a dark central spot surrounded
by a bright ring. Any other pattern reduces the effect of the dark spot. Thus

Rn2  .25  10 3   2  5  10 7 r0


2

 r0  6.25 cm

13.10
The first zone between positions 1 and 2 contributes 4I 0 . Halfway round the 2nd zone

between positions 2 and 3 thus intensity has been reduced by 2I 0 to 2I 0 . The phase

difference at position 2 lags  radians behind that of position 1. Halfway between


positions 2 and 3 thus phase difference has been increased by  / 2 radians and the
phase difference between the first and last vectors is 3 / 2 radians.
Chapter 14

14.1
For  0  30 , we have:

 1 30 
T  T0 1  sin 2   1.017T0
 4 2 

T  T0
i.e.  1.7%  2%
T0

For 0  90 , we have:

 1 90 
T  T0 1  sin 2   1.125T0
 4 2 

T  T0
i.e.  12.5%
T0

14.2
Multiplying the equation of motion by 2 dx dt and integrating with respect to t

gives:
2
 dx  x
m   A  2 f ( x)dx
 dt  0

dx
where A is the constant of integration. The velocity is zero at the maximum
dt
x0
displacement x  x0 , giving A  2  f ( x)dx .
0

2
 dx  x0 x
i.e. m   2  f ( x)dx  2 f ( x)dx  2 F ( x0 )  2 F ( x)
 dt  0 0

dx 2
i.e.  [ F ( x0 )  F ( x)]
dt m
Upon integration of the above equation, we have:
m dx
t
2 F ( x0 )  F ( x)

If x  0 at time t  0 and  0 is the period of oscillation, then x  x0 at t   0 4 ,

so we have:

Solution Set for Introduction to Vibrations and Waves, First Edition. H. J. Pain and Patricia Rankin.
© 2015 John Wiley & Sons, Ltd. Published 2015 by John Wiley & Sons, Ltd.
Companion website: http://booksupport.wiley.com
m x0 dx
0  4 
2 0 F ( x0 )  F ( x)

14.3
By substitution of the solution into x :

 n2 n n2 n 
x    an cos   bn sin  
n 1  9 3 9 3 

Since s3  s1 , we have s ( x )  s1 x , so:

 
 n2  n  n2  n 
x  s( x)   an  s1   cos   bn  s1   sin    F0 cos t
n1   9 3  9 3 

   n2  n  n2  n 
i.e.  an  s1   cos   bn  s1   sin    F0 cos 
n1   9 3  9  3 
i.e.
The above equation is true only if bn  0 and the even numbered cosine terms are

zero. By neglecting the zero terms, we have:


a3 ( s1  1) cos   a9 ( s1  9) cos 3    F0 cos 

i.e. a3 ( s1  1) cos   a9 ( s1  9)(4 cos3   3 cos  )    F0 cos 

i.e. [a3 ( s1  1)  3a9 ( s1  9)] cos  4a9 ( s1  9) cos3     F0 cos

As we can see, only a3 and a9 are the main coefficients in the solution, i.e. the

fundamental frequency term and its third harmonic term are the significant terms in
the solution.

14.4
Since V  V0 at r  r0 , by expanding V at r0 , we have:

 dV   d 2V 
V  V0    ( r  r0 )   2  (r  r0 ) 2  
 dr  r0  dr  r0
Noting that:
 dV   r 6 r 12 
   12V0  07  013   0
 dr  r0  r0 r0 
 d 2V   13r12 7r 6  V
 2   12V0  140  80   72 02
 dr  r0  r0 r0  r0

We have:
72V0
V  V0  2
(r  r0 ) 2  
r0

The expression of potential energy for harmonic oscillation is given by:


1 72V
V  V0  sx 2 , hence s  2 0 , and the oscillation frequency is given by:
2 r0

s 72V0
2  
m mr02

14.5
Relative velocity 12% of 330 m/sec  40 m/sec. Time to travel 50 m = 1:25 secs 
extra time  0.26 secs. Irrespective of the velocities of the crest and the trough in the
pulse of Figure 14.5a, its leading edge, its tail and its midpoint all have the velocity
c0 so the pulse retains a constant length.

You might also like