Physics Notes Complete
Physics Notes Complete
PH-1007 (Physics)
Dr. Gorky Shaw
F ∝ −y
or, F = −ky (1.1)
F = ma (1.2)
Now,
dv d2 y
a= = 2 (1.3)
dt dt
d2 y
m 2 = −ky
dt
d2 y k
=⇒ 2 + y = 0
dt m
2
dy
=⇒ 2 + ω 2 y = 0 (1.4)
dt
k
where ω 2 = is a positive constant.
m
2
1.1 Simple harmonic motion
y = C 1 y 1 + C 2 y2
or, y = C1 eiωt + C2 e−iωt (1.6)
A1 = C1 + C2
and, A2 = i(C1 − C2 ) (1.9)
Thus, from (1.6), (1.8), and (1.9), General solution of (1.4) is,
3
1.1 Simple harmonic motion
0 = A1 + 0 =⇒ A1 = 0 (1.11)
A1 = A sin φ
and, A2 = A cos φ (1.13)
y = A sin(ωt + φ) (1.15)
where the term (ωt + φ) is called the phase of the SHM, and φ is called
the initial phase or phase constant.
(1.6), (1.10), and (1.15) are all valid forms of the general solutions of
(1.4). However, (1.15) is the most convenient form to represent and study
simple harmonic motion.
4
1.1 Simple harmonic motion
d2 y
a = 2 = Aω 2 sin(ωt + φ) (1.20)
dt
= −ω 2 y (1.21)
5
1.1 Simple harmonic motion
1
Kmax = mω 2 A2 at y = 0
2
Kmin = 0 at y = A (1.23)
1
Pmax = mω 2 A2 at y = A
2
Pmin = 0 at y = 0 (1.25)
6
1.2 Damped Harmonic Oscillations
1. Restoring force:
FR ∝ −y
or, FR = −ky (1.27)
d2 y
m = F R + FD
dt2
d2 y dy
or, m 2 = −ky − b
dt dt
d2 y dy
=⇒ m 2 +b + ky = 0
dt dt
d2 y b dy k
=⇒ + + y=0 (1.29)
dt2 m dt m
b k
Let = 2r and = ω02 . Then we get
m m
d2 y dy
2
+ 2r + ω02 y = 0 (1.30)
dt dt
This is the equation of motion of a damped harmonic oscillator.
r is called the damping constant, and ω0 is called the natural frequency of the
7
1.2 Damped Harmonic Oscillations
y = eαt (1.31)
Thus, we get
α2 + 2rα + ω02 = 0
q
=⇒ α = −r ± r2 − ω02 (1.33)
8
1.2 Damped Harmonic Oscillations
y = A1 y1 + A2 y2
√ √
(−r+ r2 −ω02 )t (−r− r2 −ω02 )t
or y = A1 e + A2 e (1.36)
Based on the extent of damping, there are three distinct cases of damped
harmonic motion:
1. r2 < ω02 : underdamped motion.
2. r2 > ω02 : overdamped motion.
3. r2 = ω02 : critically damped motion.
As we will see subsequently, in the different cases, the displacement varies
with time as represented graphically in Figure 1.2.
9
1.2 Damped Harmonic Oscillations
p
where ω = ω02 − r2 is a real number and represent the angular frequency
of damped oscillations. Clearly, ω < ω0 .
Hence, we can rewrite (1.36) as
y = A1 e(−r+iω)t + A2 e(−r−iω)t
y = A1 e−rt eiωt + A2 e−rt e−iωt
= e−rt (A1 eiωt + A2 e−iωt )
= e−rt [A1 (cos ωt + i sin ωt) + A2 (cos ωt − i sin ωt)]
= e−rt [(A1 + A2 ) cos ωt + i(A1 − A2 ) sin ωt] (1.38)
Now, since we are discussing oscillations, which are very real, the displace-
ment y is certainly a real quantity. Hence, both (A1 + A2 ) and i(A1 − A2 )
must be real quantities.
Accordingly, we make the replacement of variables:
A1 + A2 = A0 sin φ
and i(A1 − A2 ) = A0 cos φ (1.39)
Then, from (1.38) and (1.39), we have the general solution for under-
damped harmonic oscillations
10
1.2 Damped Harmonic Oscillations
2π 2π
which represents harmonic oscillations with time period T = =p 2
ω ω0 − r 2
As mentioned earlier, ω0 is the natural frequency (without any damping)
and ω is the damped frequency.
The amplitude of oscillations
A = A0 e−rt (1.41)
Then for the overdamped case, the general solution given by (1.36) can
be rewritten as
11
1.3 Energy decay in underdamped harmonic oscillations
Then, we have
y = A1 e(−r+h)t + A2 e(−r−h)t
= e−rt (A1 eht + A2 e−ht )
= e−rt [A1 (1 + ht + ...) + A2 (1 − ht + ...)]
= e−rt [(A1 + A2 ) + h(A1 − A2 )t] (1.44)
Therefore,
dy
v= = ωA0 e−rt cos(ωt + φ) − rA0 e−rt sin(ωt + φ)
dt
= A0 e−rt [ω cos(ωt + φ) − r sin(ωt + φ)] (1.47)
12
1.4 Parameters associated with underdamped harmonic oscillations
E = E0 e−2rt (1.52)
1
where E0 = mω 2 A20 us the energy at t = 0.
2
Thus, the energy of a damped harmonic oscillator decays exponentially with
time.
Note that while amplitude decay is proportional to e−rt , energy decay is
proportional to e−2rt . This is expected because the energy of oscillation is
proportional to the square of the amplitude, as seen in the above relations.
13
1.4 Parameters associated with underdamped harmonic oscillations
Figure 1.4: Displacement versus time for underdamped oscillations. Note that
successive amplitudes are separated by time intervals of T /2. Suc-
cessive amplitudes of the same sign are separated by time intervals
of T .
tn+1 = tn + T /2 (1.54)
An = A0 e−rtn
and An+1 = A0 e−rtn+1 = A0 e−r(tn +T /2) (1.55)
Therefore,
An A0 e−rtn
= −r(t +T /2)
= erT /2 = constant (1.56)
An+1 A0 e n
rT
loge d = =λ (1.57)
2
is called the logarithmic decrement.
Alternative definition of λ: Instead of considering successive ampli-
tudes (either positive or negative), only positive (or only negative) am-
plitudes may be considered. In this case, successive positive (or negative)
14
1.4 Parameters associated with underdamped harmonic oscillations
2. Relaxation time for amplitude (τA ): it is the time when the amplitude
of oscillations decreases to 1/e times its initial value.
1
If we put t = in the expression for A, then we get
r
A0
A = A0 e−r(1/r) = A0 e−1 = (1.59)
e
1
Thus, τA = is the time in which the amplitude reduces to the initial
r
amplitude, and hence is the relaxation time.
In terms of the relaxation time τA ,
A = A0 e−t/τA (1.60)
E = E0 e−2rt (1.61)
1
Therefore at t = ,
2r
E0
E = E0 e−2r1/2r = E0 e−1 = (1.62)
e
1
Thus, τ = is the relaxation time for energy of oscillations.
2r
In terms of the relaxation time τ ,
E = E0 e−t/τ (1.63)
15
1.5 Forced oscillations
damping.
dE d 1 E
P =− = − (E0 e−t/τ ) = (E0 e−t/τ ) =
dt dt τ τ
E
Therefore, P = (1.64)
τ
Thus, Q = ωτ .
For an ideal oscillator, τ → ∞ (no loss of energy). Therefore, in this
Q → ∞.
1. Restoring force:
FR ∝ −y
or, FR = −ky (1.66)
16
1.5 Forced oscillations
d2 y
m 2 = Fext + FR + FD
dt
d2 y dy
or, m 2 = F0 sin ωt − ky − b
dt dt
d2 y dy
=⇒ m 2 +b + ky = F0 sin ωt
dt dt
d2 y b dy k F0
=⇒ + + y= sin ωt (1.69)
dt2 m dt m m
b k F0
Let = 2r, = ω02 and = f0 . Then we get
m m m
d2 y dy
2
+ 2r + ω02 y = f0 sin ωt (1.70)
dt dt
This is the equation of motion for forced oscillations.
r is the damping constant, and ω0 is natural frequency of the oscillator.
f0 is the driving acceleration.
17
1.5 Forced oscillations
1. Transient solution
It is the solution of the equation obtained by setting the RHS of (1.70)
to zero, i.e.,
d2 y dy
2
+ 2r + ω02 y = 0 (1.71)
dt dt
which is identical to (1.30) and represents damped harmonic motion. As
we have discussed in Section 1.2, the displacement represented by this
equation decays exponentially with time, and hence is not of interest when
considering forced oscillations over relatively long durations of time.
y = A sin(ωt − θ) (1.72)
Now, (1.74) has to be valid for all values of t. Hence, the coefficients of
18
1.5 Forced oscillations
sin(ωt − θ) and cos(ωt − θ) on the LHS and RHS must be separately equal
to each other. That is,
and
The phase difference (θ) between the driving force and the driven system
may be obtained by taking the ratio of (1.75) and (1.76):
2rAω f0 sin θ
2 =
−Aω 2 + ω0 A f0 cos θ
−1 2rAω
=⇒ θ = tan (1.80)
−Aω 2 + ω02 A
19
1.5 Forced oscillations
f0 F0 /m F0
A≈ 2 = = (1.82)
ω0 k/m k
Thus, in this case, the amplitude of oscillations depends only on the restor-
ing force constant k.
20
1.5 Forced oscillations
(−ω02 + ω 2 + 2r2 ) = 0
=⇒ ω 2 = ω02 − 2r2
q
=⇒ ω = ω02 − 2r2 (1.87)
ω ' ω0 (1.89)
21
1.5 Forced oscillations
resonance is flat.
The variation of peak amplitude with damping is schematically repre-
sented in Figure 1.5. The figure also indicates the shift in the resonance
frequency farther away from ω0 with increasing damping, as expected
from (1.88).
Increasing damping also has the effect of shifting the phase difference θ
at resonance from 900 to lower values.
22
1.6 Coupled oscillations
1. mg acting downwards.
We consider the radial (in the direction of the length of string) and tangential
(in the direction of motion of bob) components of the net force F acting on the
pendulum.
The radial component of force is
Fr = mg cos θ − T (1.92)
23
1.6 Coupled oscillations
Now, the displacement x of the pendulum is always along the tangential direc-
tion, as shown in Figure 1.6. Thus, this tangential component of the force acts
along the direction of displacement and opposes it, as indicated in 1.6. So, this
represents the restoring force acting on the pendulum to cause the oscillatory
motion. Therefore,
d2 x
Fθ = m 2 (1.94)
dt
So, from (1.93) and (1.94),
d2 x
m 2 = −mg sin θ (1.95)
dt
(1.95) does not represent simple harmonic motion (SHM). Recall from (1.4)
that for SHM, the restoring force has to be proportional to the displacement,
but in this case, it is not.
However, if we consider small oscillations, then for small θ,
24
1.6 Coupled oscillations
Fθ = −mgθ (1.97)
x
Also, for small θ, θ ' tan θ ' . Therefore,
L
x
θ' (1.98)
L
Therefore, using (1.96) and (1.98) in (1.95), we get
d2 x x
m 2 = −mg (1.99)
dt L
Rearranging, we get
d2 x g
+ x=0 (1.100)
dt2 L
which is the differential equation representing simple harmonic motion of a
simple pendulum for small oscillations.
1. Equation of motion
Considering only small oscillations, equations of motion of the two masses,
25
1.6 Coupled oscillations
Figure 1.7: Two identical pendulums coupled via a massless spring (Source:
Matt Jarvis, Professor of Astrophysics and Fellow, St Cross Col-
lege).
d2 x x
m 2 = −mg − k(x − y) (1.103)
dt L
d2 y y
and, m 2 = −mg − k(y − x) (1.104)
dt L
or,
d2 x 2 k
+ ω 0 x = − (x − y) (1.105)
dt2 m
and
d2 y 2 k
+ ω 0 y = − (y − x) (1.106)
dt2 m
26
1.6 Coupled oscillations
d2
2
(x + y) + ω02 (x + y) = 0 (1.107)
dt
d2 2 2k
(x − y) + ω 0 (x − y) = − (x − y)
dt2 m
d2
2 2k
or, (x − y) + ω 0 + (x − y) = 0 (1.108)
dt2 m
Replacing
X = x + y in (1.107)
and Y = x − y in (1.108) (1.109)
we get
d2 X
2
+ ω02 X = 0 (1.110)
dt
and
d2 Y
2k
2
+ ω02 + Y =0
dt m
d2 Y
or, 2
+ ω12 Y = 0 (1.111)
dt
(1.110) and
s(1.111) represent SHM with natural angular frequencies ω0
2k
and ω1 = ω02 + , respectively.
m
X = x + y = X0 cos(ω0 t + φ0 ) (1.112)
and
Y = x − y = Y0 cos(ω1 t + φ1 ) (1.113)
1
Note that we can choose to write the general solutions in terms of either sine or cosine functions. Here we
have opted for cosine functions, but the analysis holds equally good with sine functions.
27
1.6 Coupled oscillations
1. In-phase oscillations
In this case x = y, or Y = x − y = 0. Both pendulums follow (1.110)
and oscillate with the angular frequency ω0 . The pendulums are always in
phase. The spring is always unstretched / uncompressed and maintains its
natural length. The relative motion of the pendulums and corresponding
time variation of x and y are schematically represented in Figure 1.8.
Figure 1.8: In-phase coupled oscillations (Sources: Matt Jarvis, Professor of As-
trophysics and Fellow, St Cross College and PPLATO@University
of Reading).
2. Out-of-phase oscillations
In this case x = −y, or X = x + y = 0. Both s pendulums follow (1.111)
2k
and oscillate with the angular frequency ω1 = ω02 + > ω0 . The
m
pendulums are always out of phase. The spring is always stretched /
28
1.6 Coupled oscillations
These two modes are called normal modes of oscillation. ω0 and ω1 are
called normal frequencies. Any other frequency is a linear combination
of ω0 and ω1 .
3. Resonance
In general, from (1.112) and (1.113),
1
x = (X + Y )
2
1
and, y = (X − Y ) (1.114)
2
29
1.6 Coupled oscillations
Consider the special condition where the resultant amplitudes and phases
given by (1.112) and (1.113) are equal. That is,
X0 = Y0 = a (say)
and φ0 = φ1 = 0 (the equal value can be zero without loss of generality)
(1.115)
30
1.6 Coupled oscillations
31
1.6 Coupled oscillations
32
1.7 Waves
1.7 Waves
1.7.1 Introduction
Wave motion is periodic motion in space and time. A disturbance of any
conventional property y of a medium may spread through space with time.
Such a spreading of disturbances in a medium is called a wave. y may be
called the wave function.
Let y represent displacement of particles in the medium. Then wave motion
can result from periodic displacement of particles from their mean positions.
The state of motion of a particle is called phase. Different particles may be
in different phases at a given time.
Wave motion consists of the propagation of phase from point to point in the
medium, distinct from the motion of particles in the medium2 .
Types of waves:
33
1.7 Waves
Figure 1.12: (a) Longitudinal and (b) transverse waves demonstrated using a
slinky.
In the next Sections, we will focus on transverse waves; discuss the important
parameters, and set up the general wave equation and its differential form.
34
1.7 Waves
1
3. Frequency (n or ν): number of wave cycles per second. n = .
T
2π
4. Angular frequency (ω): ω = 2πn = .
T
5. Wavelength (λ): distance travelled by the wave in one cycle, i.e., in
time T , as shown in Figure 1.13.
6. wave number (k): a measure of the number of wave cycles per unit
2π
distance. k = .
λ
λ 1
7. Velocity (v): v = . Since n = , we have
T T
v =n·λ (1.127)
y = a sin ωt (1.128)
35
1.7 Waves
y = a sin(ωt − φ) (1.129)
x=λ
and φ = 2π (1.130)
because over a complete wave cycle, the phase goes through a cycle of 0 to 2π.
Therefore, from (1.130),
2πx
φ= = kx (1.131)
λ
If v is the wave velocity, then
2πv
ω = 2πn = (1.132)
λ
Using these expressions for φ and ω in (1.129), we can write
2πvt 2πx
y = a sin − (1.133)
λ λ
or,
2π
y = a sin (vt − x) (1.134)
λ
2π
y = a sin (vt + x) (1.135)
λ
36
1.7 Waves
and
∂ 2y ω2 ∂ 2y
= 2· 2 (1.141)
∂t2 k ∂x
Now,
ω 2πn
= =n·λ=v (1.142)
k 2π/λ
3
Note that here we switch to partial derivatives from total derivatives. This is because, unlike in case of
oscillations, where the displacement is a function of time (t) only, in case of waves the displacement is a
function of both time (t) and position (x).
37
1.7 Waves
∂ 2y 2
2 ∂ y
=v · 2 (1.143)
∂t2 ∂x
which is the differential equation of motion for a transverse wave travelling with
velocity v.
38
1.7 Waves
Note that, in the above treatment, we have considered that the phase of the
particle at the origin is zero. This does not affect the general applicability of
the wave equation (1.134) or (1.136). However, one may consider some non-
zero initial phase φ0 at the origin, and in this case, the further generalized wave
equation is
y = a sin(ωt − kx − φ0 ) (1.144)
Further note that, instead of (1.134) and (1.136), we can also write the wave
equation as
2π
y = a sin (x − vt) = a sin(kx − ωt) (1.145)
λ
.
Solution:
y = 4 sin[(0.1π)(x − 20t)]
Comparing term-by-term:
(i) amplitude, a = 4 cm.
2π
(ii) = 0.1π =⇒ wavelength λ = 20 cm.
λ
(iii) Initial phase φ0 = 0.
39
1.7 Waves
(iv) v = 20 cm/s.
v 20
(v) Frequency n = = = 1.0 Hz.
λ 20
Solution:
Given, frequency of the radio waves, n = 15×106 Hz, and velocity v = 3×108
m/s.
Therefore, length of the waves (i.e., the wavelength)
v 3 × 108
λ= = = 20 m.
n 15 × 106
********************* The End (of this Unit) *********************
40
Interference of light
PH-1007 (Physics)
Dr. Gorky Shaw
2.1 Introduction
Light is a transverse, electromagnetic wave that is visible to the human eye.
Light is dualistic in nature.
Certain optical phenomena, such as interference, diffraction, and polarization,
are explained by the wave nature of light. Light-matter interaction in processes
of emission and absorption, such as photoelectric effect and Compton effect,
are explained by the corpuscular (particle) nature of light.
2. Superposition principle
When two waves interact, the resulting wave function is the sum of the
two individual wave functions.
2.1 Introduction
Consider two waves with wave functions y1 and y2 , represented by the
equations
y1 = a1 sin(ω1 t − k1 x)
and y2 = a2 sin(ω2 t − k2 x) (2.1)
When these two waves interact, the resultant wave function y is given by
y = y1 + y2
= a1 sin(ω1 t − k1 x) + a2 sin(ω2 t − k2 x) (2.2)
y = y1 + y2 + y3 + ... + yn
Xn
= yi (2.3)
i=1
3. Coherent sources
Two light sources are said to be coherent if they produce waves that have
the same frequency, and a sharply defined phase difference that remains
unchanged with time. The most common example of coherent sources is
Laser.
4. Interference
When two waves produced by coherent sources travel simultaneously in
a medium and superpose each other, the resultant intensity is not dis-
tributed uniformly in space. This modification in intensity is called in-
terference.
2
2.1 Introduction
Figure 2.2: (a) Constructive and (b) destructive interference, a direct conse-
quence of the superposition principle (Source: A-Level Physics).
3
2.2 Analytical treatment of interference by division of wavefront - Young’s
double slit
2. The distance of observation should be large.
For good contrast:
1. Amplitudes of the interfering waves should be nearly equal (and ideally,
exactly equal).
= S2 P − S1 P (2.4)
4
2.2 Analytical treatment of interference by division of wavefront - Young’s
double slit
y1 = a1 sin ωt
and y2 = a2 sin(ωt + δ) (2.6)
y = y1 + y2
= a1 sin ωt + a2 sin(ωt + δ)
= a1 sin ωt + a2 (sin ωt cos δ + cos ωt sin δ)
= (a1 + a2 cos δ) sin ωt + (a2 sin δ) cos ωt (2.7)
5
2.2 Analytical treatment of interference by division of wavefront - Young’s
double slit
a1 + a2 cos δ = R cos θ
and a2 sin δ = R sin θ (2.8)
(2.9) indicates two notable features of the resultant wave: its frequency re-
mains the same (angular frequency ω same as that of the source waves), and its
nature remains unaltered (similar forms of the wave equations (2.6) and (2.9)).
Here, R is the resultant amplitude. Hence the resultant intensity is I = R2 .
Squaring and adding the two equations in (2.8), we get
6
2.2 Analytical treatment of interference by division of wavefront - Young’s
double slit
cos δ = +1
=⇒ δ = 2nπ, n = 0, ±1, ±2, ... (2.11)
That is,
If a1 = a2 = a and I = a2 , then
cos δ = −1
=⇒ δ = (2n + 1)π, n = 0, ±1, ±2, ... (2.16)
7
2.2 Analytical treatment of interference by division of wavefront - Young’s
double slit
That is,
If a1 = a2 = a and I = a2 , then
Imin = 0 (2.20)
Figure 2.5: Variation of resultant intensity (I) with phase difference (δ).
8
2.2 Analytical treatment of interference by division of wavefront - Young’s
double slit
= I1 + I2 (2.21)
9
2.2 Analytical treatment of interference by division of wavefront - Young’s
double slit
(x − d)2
S1 P = D + (2.24)
2D
Therefore,
2xd
S2 P − S1 P = (2.25)
D
For maxima or bright fringes:
1 n
P∞ n k n n(n − 1)...(n − k + 1)
For a real number n, the binomial expansion of (1+x) = k=0 k x , where k = .
k!
Here k is a positive integer.
10
2.2 Analytical treatment of interference by division of wavefront - Young’s
double slit
2xd
= nλ
D
Dλ
=⇒ x = n · (2.26)
2d
For minima or dark fringes:
2xd λ
= (2n + 1)
D 2
(2n + 1) Dλ
=⇒ x = · (2.27)
2 2d
Therefore, spacing between two consecutive bright (or dark) fringes
= xn+1 − xn
Dλ
= (2.28)
2d
which is independent of n.
11
2.3 Numerical problems - Interference by division of wavefront
Exercise 2.1
Two coherent sources with intensity ratio 100:1 produce interference fringes.
Find the ratio between maximum and minimum intensities in the interference
pattern.
Solution:
We know,
Imax (a1 + a2 )2
=
Imin (a1 − a2 )2
Now, given,
I1 a21 100
= 2=
I2 a2 1
a1
=⇒ = 10
a2
=⇒ a1 = 10a2
Therefore,
Exercise 2.2
The ratio between maximum and minimum intensities in a double slit inter-
ference pattern is 36:1. Find the ratio between amplitudes and intensities of
the two interfering waves.
Solution:
12
2.3 Numerical problems - Interference by division of wavefront
Given,
Imax (a1 + a2 )2 36
= =
Imin (a1 − a2 )2 1
=⇒ a1 + a2 = 6(a1 − a2 )
=⇒ (a1 + a2 ) = 6(a1 − a2 )
=⇒ 5a1 = 7a2
I1 a21 49
= 2=
I2 a2 25
Exercise 2.3
In a Young’s double slit experiment with monochromatic light of wavelength
6000 Å, the fringe width is found to be 0.5 mm in the interference pattern on a
screen at a distance 1 m from the slits. Find the separation between the slits.
Solution:
Dλ 100 × 6 × 10−7
2d = = = 0.0012 cm
β 0.05
Exercise 2.4
Two coherent sources placed 0.2 mm apart produce an interference pattern
observed on a screen 1 m away. With a certain monochromatic light source, the
fourth bright fringe is situated at a distance 10.0 mm from the central fringe.
Find the wavelength of light used.
13
2.3 Numerical problems - Interference by division of wavefront
Solution:
Exercise 2.5
In an interference pattern, 12th order maximum is observed for λ = 6000 Å.
What order maximum is visible with light of wavelength 4800 Å?
Solution:
We know, the path difference between source waves for the nth order maxi-
mum ∝ nλ.
Given, maxima (of different orders) for two wavelengths occur at the same
position, or
n2 λ2 = n1 λ1
n1 λ1 12 × 6000
or, n2 = = = 15
λ2 4800
14
2.4 Interference due to thin films
15
2.5 Interference due to parallel-sided thin films: reflected rays
But,
and
16
2.5 Interference due to parallel-sided thin films: reflected rays
Therefore,
In this condition, the thin film will appear dark in reflected light.
Note that the choice of + and − signs, as well as values of n (0, 1, 2... or 1, 2, 3...)
in the conditions (2.36) and (2.37) are not arbitrary. These are carefully chosen
17
2.6 Interference due to thin films - further points
so that the value of t, the thickness of the film, is always positive. Negative
value of t would not make physical sense.
18
2.6 Interference due to thin films - further points
Figure 2.8: Colours in thin films (a) Soap bubbles, (b) Oil spill, and (c) Bottom
of a cooking pot.
For a wedge-shaped thin film with wedge angle θ, as shown in Figure 2.9, the
effective path difference at thickness t is given by
λ
peff = 2µt cos (r + θ) ± (2.40)
2
Accordingly, in reflected light, for maximum intensity,
λ
2µt cos (r + θ) = (2n − 1) , n = 1, 2, 3... (2.41)
2
and for minimum intensity,
For such a film, the thickness varies across the film. Hence, alternate locations
will demonstrate constructive and destructive interference, according to the
thickness of the film at that location. This is the basis of Newton’s rings
formation, which is discussed next.
19
2.7 Newton’s rings
20
2.7 Newton’s rings
terference between light waves reflected from the upper and lower surfaces of
the wedge-shaped air film. The thickness of the air film is zero at the point of
contact and gradually increases outwards. Due to the shape of the plano-convex
lens, the locus of points where the thickness of the air film is constant is a circle,
with the point of contact as its centre.
21
2.7 Newton’s rings
the lens. Let rn be the radius, and Dn = 2rn the diameter, of the Newton’s ring
corresponding to a point P where the film thickness is t.
We use the property of a circle2 :
rn2 = PN2 = ON × NE
= t × (2R − t)
= 2Rt − t2 (2.48)
rn2 = 2Rt
r2
=⇒ 2t = n (2.49)
R
2
Verify this yourself!
22
2.7 Newton’s rings
Therefore, from (2.45) and (2.49), for a bright ring (constructive interference),
rn2 λ
µ = (2n − 1) , n = 1, 2, 3...
R 2
λR
=⇒ rn2 = (2n − 1) (2.50)
2µ
rn2
µ = nλ, n = 0, 1, 2, 3...
R
nλR
=⇒ rn2 = (2.52)
µ
23
2.7 Newton’s rings
Therefore,
2 2
Dm+p − Dm = 4pλR (2.57)
That is,
2 2
Dm+p − Dm
λ= (2.58)
4pR
In the experimental setup shown in Figure 2.10, the cross-wire of the eye-
piece of the microscope is focused on any dark ring (say 24th, 16th etc.), and
reading of the microscope is noted. Similar readings are noted by focusing the
cross-wire after every few dark rings. Diameters of different rings are calculated,
and a graph is plotted between the square of the diameter (Dn2 ) and the ring
number n, which is a straight line, as shown in Figure . Its slope gives the value
2 2
Dm+p − Dm
of . If R is known, the wavelength λ of the monochromatic light
p
can then be determined from 2.58.
24
2.8 Numerical problems - Interference by division of amplitude
If a liquid is put in between the plano-convex lens and the glass plate, then for
the liquid film,
2 2 4pλR
[Dm+p − Dm ]liquid = (2.60)
µ
Therefore, from (2.59) and (2.60),
2 2
[Dm+p − Dm ]air
µ= 2 2]
(2.61)
[Dm+p − Dm liquid
Hence, refractive index (µ) of the liquid can be determined by measuring the
diameters of Newton’s rings in the liquid and in air.
Exercise 2.6
A soap film (µ = 1.33) seen by sodium light (λ = 5893 Å) by normal reflection
appears dark. Find the minimum thickness of the film.
Solution:
For the film to appear dark, i.e., for destructive interference, we have
2µt cos r = nλ
Clearly, the minimum possible value of the thickness of the film, t, is obtained
for n = 1.
For normal incidence, r = 0, cos r = 1
Therefore,
nλ 1 × 5893
t= = = 2215 Å
2µ 2 × 1.33
25
2.8 Numerical problems - Interference by division of amplitude
Exercise 2.7
Newton’s rings are observed in reflected light of wavelength 6250 Å. Diameter
of the 10th dark ring is 0.5 cm. Find (i) the radius of curvature of the lens (ii)
thickness of the air film.
Solution:
Dn2 = 4nλR
Dn2 (0.5)2
R= =
4nλ 4 × 10 × 6.25 × 10−5
0.25
=
250 × 10−5
= 100 cm
= 1m
2t = nλ
nλ
=⇒ t =
2
10 × 6.25 × 10−5
=
2
= 31.25 × 10−5
= 3.125 µm
Exercise 2.8
In a Newton’s rings experiment, the diameter of the 15th ring is found to
be 0.590 cm and that of the 5th ring 0.336 cm. If radius of curvature of the
plano-convex lens is 100 cm, calculate the wavelength of light used.
Solution:
26
2.8 Numerical problems - Interference by division of amplitude
Exercise 2.9
In a Newton’s rings formation in a liquid film with light of wavelength 6000
Å, diameter of the sixth bright ring is 3.1 mm and radius of curvature of the
curved surface is 1 m. Calculate the refractive index of the liquid.
Solution:
2 × (12 − 1) × 6 × 10−5 × 1
µ=
0.31
2 × 11 × 6 × 10−5
=
0.31
≈ 1.374
27
2.9 Michelson interferometer
Figure 2.13: Michelson interferometer (a) Typical experimental setup (b) Pla-
nar view showing the propagation of light in the set up (Source:
Physics LibreTexts).
The observer sees the interference pattern due to beams reflected from the
mirror M1 and a virtual image M02 of the mirror M2 formed by M, which act
as coherent sources. The arrangement is optically equivalent to a thin air film
between M1 and M02 .
The schematics in Figure 2.13 only a single reference light beam emergent
from the light source S. However, it is an extended source (rather than a point
source), which means there are light beams are emergent from various points
on the source and are incident on M at different angles. With monochromatic
light, the resultant of interference between all these beams from the extended
source is a circular pattern of alternating dark and bright fringes (similar in
appearance to Newton’s rings) when the two mirrors M1 and M02 are parallel
28
2.9 Michelson interferometer
29
Diffraction
PH-1007 (Physics)
Dr. Gorky Shaw
3.1 Introduction
According to geometrical optics, if a plane wave is incident on a long narrow
slit, as shown in Figure 3.1, the region PQ is illuminated and the rest of the
region forms the geometrical shadow.
However, if the slit is made very narrow (comparable to the wavelength of
light), then light bends into the geometrical shadow region. This is known as
diffraction.
1. Fresnel diffraction: The source, slit and screen are at finite distance
from each other. The incident wavefront is spherical or cylindrical.
2
3.2 Fraunhofer diffraction at a single slit
3
3.2 Fraunhofer diffraction at a single slit
πe sin θ
Let = α. Then
λ
a sin(nd/2)
R=
sin(d/2)
a sin α
=
sin(α/n)
a sin α
' (for small α/n)
α/n
na sin α
= (3.6)
α
As n → ∞, a → 0. But the product na remains finite.
Let na = A. Then
A sin α
R= (3.7)
α
Therefore, resultant intensity at P,
A2 sin2 α
2
I=R = (3.8)
α2
sin α = 0 (3.10)
4
3.2 Fraunhofer diffraction at a single slit
or
α cos α − sin α = 0
or, α − tan α = 0 (3.11)
sin α = 0 (but α 6= 0)
=⇒ α = nπ (n = ±1, ±2, ±3...but n 6= 0)
πe sin θ
=⇒ = nπ (3.12)
λ
That is,
e sin θ = nλ (3.13)
πe sin θ/λ = 0
=⇒ sin θ = 0
=⇒ θ = 0 (3.15)
That is the same direction as that of the incident light. This is known as the
principal maximum or central maximum.
5
3.2 Fraunhofer diffraction at a single slit
α − tan α = 0 (3.16)
πe sin θ π
' (2n + 1) , n = ±1, ±2, ±3... (3.18)
λ 2
or,
λ
e sin θ ' (2n + 1) , n = ±1, ±2, ±3... (3.19)
2
6
3.2 Fraunhofer diffraction at a single slit
etc.
7
3.3 Plane Diffraction Grating
0
where A = Resultant amplitude from each slit
A sin α
= (3.26)
α
That is,
A sin α sin N β
R= · (3.27)
α sin β
8
3.3 Plane Diffraction Grating
The first factor gives a diffraction pattern due to a single slit. The second
factor gives the interference pattern due to N slits.
9
3.3 Plane Diffraction Grating
sin N β = 0 (3.29)
or
3.3.3 Minima
When sin N β = 0 (but sin β 6= 0), we have the minimum intensity I = Imin = 0.
sin N β
When sin β = 0, the term is indeterminate (and corresponds to principal
sin β
maxima, as discussed in Section 3.3.4 below.
Therefore, the condition for minimum intensity, given by (3.29) is
π
=⇒ N (e + d) sin θ = mπ
λ
=⇒ N (e + d) sin θ = mλ (3.32)
where m takes all integral values except 0, ±N, ±2N , ±3N , ..., ±nN (where
n is an integer). This is because these values of m give sin β = 0, corresponding
to principal maxima, as discussed in Section 3.3.4. Thus, there are (N − 1)
minima between successive principal maxima.
10
3.3 Plane Diffraction Grating
A2 sin2 α
Ip = 2
· N2 (3.35)
α
These maxima are most intense and are called principal maxima. These
are obtained in the directions given by the condition (3.33):
β = nπ
π
=⇒ (e + d) sin θ = nπ
λ
=⇒ (e + d) sin θ = nλ (3.36)
where n = 0, ±1, ±2, ±3... correspond to the nth order principal maxima.
11
3.3 Plane Diffraction Grating
sin N β
Figure 3.5: Triangle to determine .
sin β
sin2 N β N 2 tan2 β
=
sin2 β (1 + N 2 tan2 β) sin2 β
N2
=
(cot2 β + N 2 ) sin2 β
N2
=
cos2 β + N 2 sin2 β
N2
= 2 2 (using sin2 β + cos2 β = 1) (3.38)
1 + (N − 1) sin β
Therefore, intensity of the secondary maxima,
0 A2 sin2 α N2
I = ·
α2 1 + (N 2 − 1) sin2 β
0 Ip
=⇒ I = (3.39)
1 + (N 2 − 1) sin2 β
12
3.3 Plane Diffraction Grating
sin2 α sin2 N β
Variation of the and terms with the direction θ, and the
α2 sin2 β
resultant grating spectrum (intensity curve), are shown in Figure 3.6.
If both (3.40) and (3.41) are satisfied for a given θ, a particular maximum of
order n will be missing in the grating spectrum.
13
3.4 Dispersive power and resolving power of a grating
dθ n
=⇒ = (3.44)
dλ (e + d) cos θ
14
3.4 Dispersive power and resolving power of a grating
15
3.4 Dispersive power and resolving power of a grating
The first minimum of λ adjacent to the nth principal maximum, along the
direction (θn + dθn ), will correspond to m = nN + 1, for which we get
If the nth principal maximum of λ + dλ, occurs in the same direction (see
Figure 3.9), then
Therefore,
λ
= nN
dλ
λ N (e + d) sin θn
or, = (3.50)
dλ λ
16
3.5 Numerical problems - Diffraction
Exercise 3.1
Light of wavelength 5000 Å is incident normally on a plane transmission grat-
ing. Find the difference in deviations in the first and third order spectra. No.
of lines per cm on the grating surface is 6000.
Solution:
(e + d) sin θ1 = n1 λ
n1 λ 1 × 5 × 10−5
=⇒ sin θ1 = = = 0.3
e+d 1/6000
(e + d) sin θ2 = n2 λ
n2 λ 3 × 5 × 10−5
=⇒ sin θ2 = = = 0.9
e+d 1/6000
Exercise 3.2
Light of wavelength 5000 Å falls normally on a plane transmission grating
having 15000 lines in 3 cm. Find the angle of diffraction for maximum intensity
17
3.5 Numerical problems - Diffraction
Solution:
nλ 1 × 5 × 10−5
sin θ = = = 0.25
e+d 1/5000
Exercise 3.3
A single slit is illuminated by light composed of two wavelengths λ1 and λ2 . It
is observed that the first diffraction minimum for λ1 coincides with the second
diffraction minimum for λ2 . What is the relation between λ1 and λ2 ?
Solution:
e sin θ = mλ
e sin θ = λ1 = 2λ2
=⇒ λ1 = 2λ2
Exercise 3.4
What is the minimum number of lines required in grating to just resolve the
lines of wavelength 5890 Å and 5896 Å in the second order?
Solution:
18
3.5 Numerical problems - Diffraction
Exercise 3.5
How many orders will be observed by a grating with 4000 lines/cm, if it is
illuminated by light of wavelength in the range 5000-7500 Å?
Solution:
For principal maxima in a grating diffraction pattern,
nλ = (e + d) sin θ
19
Laser
PH-1007 (Physics)
Dr. Gorky Shaw
4.1 Introduction
LASER stands for Light Amplification by Stimulated Emission of Radiation.
In the classical view, the energy of an electron orbiting an atomic nucleus
is larger for orbits further from the nucleus of an atom. However, quantum
mechanical effects force electrons to take on discrete positions in orbitals. Thus,
electrons are found in specific energy levels of an atom, and can transition
between energy levels via suitable absorption or emission of energy. An idealized
atom with two electron energy levels and one electron is shown in Figure 4.1.
These processes are more conveniently represented in energy level diagrams
as shown in Figure 4.2.
Figure 4.2: Schematic energy level diagrams showing the absorption, sponta-
neous emission and stimulated emission processes in an idealized
atom with a single electron and two available energy levels.
2
4.1 Introduction
2. An excited state is an upper energy level with very short lifetime. That
is, electrons reaching these energy levels from the ground state by absorp-
tion of energy (such as from incident photons) almost instantaneously
drop back down to a lower energy level with emission of energy (mostly
in the form of radiation).
N1 ∝ e−E1 /kT
N2 ∝ e−E2 /kT
N3 ∝ e−E3 /kT ...etc. (4.1)
3
4.1 Introduction
5. Pumping is the method of raising the atoms from lower energy levels to
higher ones, in order to achieve population inversion.
Some pumping techniques are:
Optical pumping: Optical energy in the form of photons is used to
excite the atoms in the medium. An external light source (like xenon flash
lamp) is used to achieve population inversion the laser medium. Optical
pumping is used in solid-state lasers such as ruby lasers, discussed in
Section 4.6.
Electrical pumping: High-voltage electric discharge acts as the pump
source and direct electron excitation occurs through electric discharge.
Electrical pumping is used in gas lasers (e.g., Argon ion lasers).
Inelastic collision between atoms: In this case also, high-voltage
electric discharge acts as a pump source, but a combination of two types
of gases, with comparable excited states, is used. This method is used in
the Helium–Neon (He-Ne) laser, and described in detail in Section 4.7.
Direct conversion: In this method, electrical energy applied to semi-
conductors and LEDs is directly converted into light energy due to recom-
bination of electrons and holes. This method of pumping is used in semi-
conductor laser such as GaAs (gallium arsenide), AlGaAs (Aluminium
gallium arsenide), GaN (gallium nitride) etc.
Chemical reaction: An atom or a molecule produced to be in the
excited state at the time of production through some chemical reaction
can be used for pumping. It is used in chemical lasers such as hydrogen
fluoride (HF) laser.
4
4.2 Three-level Laser
Consider a system with three energy levels E1 < E2 < E3 , as shown in Figure
4.4. E1 is the ground state. E3 is called the pump level. E2 is a metastable state
with lifetime much longer than that of E3 . In normal conditions, populations
in these levels are N1 N2 > N3 .
When supplied with light of energy hν = E3 − E1 , the atoms in the lower
energy state E1 are excited to the pump level E3 . This process of supplying
energy is called pumping. Out of these excited atoms, some reach the ground
state by spontaneous emission. But many of them undergo fast spontaneous
non-radiative transition to the metastable state E2 instead.
Due to the longer lifetime of the level E2 , spontaneous transitions from E2
to E1 do not occur often, leading to accumulation of atoms in the level E2 .
Eventually, we have a situation where N2 > N1 > N3 . Thus in the three-level
system, population inversion is achieved between the energy levels E1 and E2 .
A photon spontaneously emitted when an atom transitions from E2 to E1
interacts with more atoms in the level E2 and causes stimulated emission of
more photons. Photons emitted during these stimulated emission processes
again interact with atoms in the level E2 . This triggers an avalanche effect.
This way, a large number of photons are emitted and laser beam is obtained.
5
4.3 Four-level Laser
6
4.4 Components of a laser system
The pump source provides energy to the laser medium to attain population
inversion.
The laser medium or active medium is a medium which, when excited,
attains the state of population inversion and causes light amplification. Laser
medium can be solid, liquid, or gas.
Optical resonator: The laser medium is surrounded by two parallel mir-
rors which provide feedback of light. One mirror is perfectly reflective (high
reflector) while the other one is partially reflective (output coupler). Together
they are known as optical resonator or optical cavity or resonating cavity. The
optical coupler allows some light to leave the optical cavity to produce the laser
output beam. Photons in the laser medium bounce back and forth between the
two mirrors, causing an avalanche of stimulated emissions to achieve optical
gain.
Thus, amplified light is produced by stimulated emission and hence will travel
to large distances without spreading out in space, after having escaped through
the output coupler.
7
4.5 Types of Laser
8
4.6 Ruby Laser
The light or photons emitted due to stimulated emission will escape through
the partially reflecting mirror or output coupler to produce laser light.
The ruby laser works in pulsed mode due to the high pump energy required
to operate a three-level laser. At room temperature, a ruby laser will only emit
short bursts of laser light, each laser pulse occurring after a flash of the pumping
light.
9
4.7 He-Ne laser
It would be better to have a laser that emits light continuously. Such a laser
is called a continuous wave (CW) laser. Four-level lasers are suitable to work
in CW mode.
10
why He emission not possible 4.7 He-Ne laser
of the helium atoms gain enough energy and jump into the excited states or
metastable states. The metastable state electrons of the helium atoms cannot
return to ground state by spontaneous emission. However, they can return to
ground state by transferring their energy to the lower energy state electrons
of the neon atoms. The metastable states of neon have the longer lifetime.
Therefore, a large number of neon atoms will remain in the metastable states
and hence population inversion is achieved.Thus, helium atoms help neon atoms
in achieving population inversion.
Figure 4.10 shows the energy level diagram of Helium-Neon laser, with the
possible transitions. The amount of Helium in the tube is about 6 times the
amount of Neon. Thus Helium atoms have more chance to receive energy from
the accelerated electrons, and transfer into the excited energy levels E3 and E5 .
Neon atoms have two excited energy levels (E3 and E5 ) which are very close
to the excited energy levels of the Helium atom. The excited Helium atoms
transfer their excitation energy to the Neon atoms by collisions by a process
called resonance excitation.
The metastable state electrons (E3 and E5 ) of the neon atoms will sponta-
neously fall into the next lower energy states (E2 and E4 ) by releasing photons
corresponding to red light and infrared radiation. The electrons continue on to
the ground state E1 through radiative and nonradiative transitions.
The photons emitted from the neon atoms will move back and forth between
two mirrors until they stimulate other excited electrons of the neon atoms and
causes them to emit light. Thus, optical gain is achieved.
The light or photons emitted due to stimulated emission will escape through
the partially reflecting mirror or output coupler to produce laser light.
Energy from the He-Ne laser is emitted at wavelengths which correspond to
11
4.8 Applications of Laser
E5 − E4 ⇒ λ1 = 3.391 µm (near-infrared)
E5 − E2 ⇒ λ2 = 0.632 µm (visible light, red color)
E3 − E2 ⇒ λ3 = 1.152 µm (mid-infrared)
3. Surgery
6. Thermonuclear reactions
7. Holography
8. Isotope separation.
12
Quantum Mechanics - I
PH-1007 (Physics)
Dr. Gorky Shaw
5.1 Introduction
As discussed in earlier chapters, light is a transverse, electromagnetic wave that
is visible to the human eye. Light is dualistic in nature. That is, certain optical
phenomena, such as interference, diffraction, and polarization, are explained by
the wave nature of light. Light-matter interaction in processes of emission and
absorption, such as photoelectric effect and Compton effect, are explained by
the corpuscular (particle) nature of light.
Development of quantum mechanics began in early 20th century with path-
breaking ideas of Max Planck and Niels Bohr. One of the major driving forces
towards this development was the failure of classical physics to satisfactorily
explain this wave-particle duality of light.
In particular, some of the phenomena which could not be explained by clas-
sical theories are:
1. Blackbody radiation
2. Stability of the atom
3. Photoelectric effect
4. Compton effect
5. Specific heat of solids at low temperatures
E = σT 4 (5.1)
2
5.2 Blackbody radiation
Figure 5.3: Radiation curves for stars at different temperatures (Source: Hy-
perPhysics).
3
5.2 Blackbody radiation
temperature T is given by
a
Eλ dλ = dλ (5.3)
λ5 eb/λT
where a and b are constants.
According to Rayleigh-Jeans law, based on classical physical arguments and
empirical facts, the energy density in the wavelength range λ to (λ + dλ) at
temperature T is given by
8πkT
Eλ dλ = dλ (5.4)
λ4
where k is the Boltzmann constant.
As schematically represented in Figure 5.4, both of these distribution laws fail
to satisfactorily the experimental observations. Wien’s law fails to accurately
fit the data for long wavelengths. Rayleigh–Jeans law fails to describe the short
wavelength spectrum of blackbody radiation, which is also known as ultravi-
olet catastrophe as the law begins to diverge with empirical observations as
these frequencies reach the ultraviolet region of the electromagnetic spectrum.
4
5.2 Blackbody radiation
According Planck’s radiation law, the energy density in the wavelength range
λ to (λ + dλ) at temperature T is given by
8πhc
Eλ dλ = dλ (5.6)
λ5 ehc/λkT − 1
1
In essence: every Macroscopic phenomenon which appears to be continuous is microscopically quantized.
5
5.3 Quantization and Bohr model of the atom
For small λ, Planck’s reduces to Wien’s law. For small values of λ, ehc/λkT >> 1.
Therefore, in this limit, (5.6) may be rewritten as
8πhc
Eλ dλ ' dλ
λ5 ehc/λkT
a
= dλ (5.8)
λ5 eb/λT
which is Wien’s law (5.3), with a = 8πhc and b = hc/k.
En − En−1 = hν (5.9)
6
5.4 Photoelectric effect
5.4.2 Observations
1. Photoelectric effect is an instantaneous process.
2. Photoelectric current (number of electrons ejected) is proportional to the
intensity of radiation.
3. A minimum frequency (ν0 ) of light is needed to cause photoelectric emis-
sion from a metal surface. The corresponding minimum energy required
for photoelectric emission is known as the work function (φ0 = hν0 ) of
the metal.
4. Kinetic energy of photoelectrons is proportional to the frequency of the
incident radiation.
7
5.5 Particle nature of Radiation
from the atom. If this quantum of energy exceeds this minimum energy needed
for the electron to escape from metal surface, i.e., the work function φ0 , then
the electron is emitted from the metal and the excess energy (hν − φ0 ) appears
1
as its kinetic energy mv 2 . Hence the equation for photoelectric emission is
2
1 2
mv = hν − φ0 (5.10)
2
2. Each photon has an energy hν and speed c (in vacuum), the speed of
light.
8
5.5 Particle nature of Radiation
4. Photons are electrically neutral and are not deflected by electric or mag-
netic field.
2 m0 c2
E = mc = r (5.11)
v2
1− 2
c
where m0 is the rest mass of the particle.
Since a photon moves with speed of light, we have v = c. If m0 c2 = 0, then
from (5.11), E would be infinite. But energy E of the photon is finite (= hν).
Hence, the photon must have zero rest mass (i.e., m0 = 0), and E cannot be
0
determined from this expression, which takes the indeterminate form.
0
We rather use the relativistic energy-momentum relationship
E 2 = p2 c2 + m20 c4 (5.12)
E = pc (5.13)
9
5.6 De Broglie hypothesis of matter waves
h
p=
λ
h
or, λ = (5.14)
p
De Broglie proposed that the role of matter waves in guiding the motion of
material particles is analogous to that of electromagnetic waves in guiding the
motion of photons.
10
5.6 De Broglie hypothesis of matter waves
Hence,
h h h
λ= =√ =√ (5.18)
p 2mK 2mE
p2
K= = qV (5.19)
2m
Hence, in this case
h h h
λ= =√ =√ (5.20)
p 2mK 2mqV
11
5.6 De Broglie hypothesis of matter waves
and
r
v2
h h 1 − c2
λ= = (5.24)
p m0 v
Figure 5.7: Electron orbits as standing waves around the nucleus (Source:CK-12
Foundation).
12
5.6 De Broglie hypothesis of matter waves
is quantized.
vp = νλ (5.27)
For a simple harmonic wave propagating along the +ve x-direction, given by
where a is the amplitude and (ωt − kx) is the phase of the wave, the phase
velocity
dx
vp = (5.29)
dt
is given by
2πν ω
vp = νλ = = (5.30)
2π/λ k
13
5.6 De Broglie hypothesis of matter waves
In terms of the reduced Planck’s constant, the energy and momentum of the
particle associated with the wave are given by
h
E = hν = × 2πν = ~ω
2π
h h 2π
and, p = = · = ~k (5.33)
λ 2π λ
Accordingly, the phase velocity can also be expressed as the ratio of the energy
and momentum of the associated particle:
ω ~ω E
vp = = = (5.34)
k ~k p
1 p2
E = mv 2 = (p = mv) (5.35)
2 2m
Therefore,
E p2 p mv v
vp = = = = = (5.36)
p 2mp 2m 2v 2
which indicates the wave travels at half the speed of the associated par-
ticle.
14
5.6 De Broglie hypothesis of matter waves
Therefore,
E mc2 c2
vp = = = (5.38)
p mv v
Now, for material particles, v < c always. Hence, we have vp > c.
Hence, from (5.37) and (5.38), we note that for both non-relativistic and rela-
tivistic particles, the phase velocity of matter wave is different from the veloc-
ity of the associated particles. This apparent anomaly was resolved by Erwin
Schrödinger by introducing the wave packet representation of matter waves,
i.e., considering that a packet (or group) of waves, rather than a single wave, is
associated with a particle in motion.
15
5.6 De Broglie hypothesis of matter waves
simple harmonic waves with the same amplitude but with a difference of dω
and dk in angular frequency and wave number, respectively, given by2
y1 = a cos(ωt − kx)
and y2 = a cos[(ω + dω)t − (k + dk)x] (5.39)
2
we could as well use sin instead of cos, the end result would be the same.
16
5.6 De Broglie hypothesis of matter waves
Reordering,
dω dk
y = 2a cos t − x cos(ωt − kx) (5.42)
2 2
where
dω dk
A = 2a cos t− x (5.44)
2 2
It represents the envelope of the wave group. Group velocity, the velocity of
propagation of the envelop, and hence of the modulated wave group, is given
by
dω
vg = (5.45)
dk
dω d(~ω) dE
vg = = = (5.46)
dk d(~k) dp
1. Non-relativistic particle with velocity v:
2
d p p
vg = = =v (5.47)
dp 2m m
17
5.6 De Broglie hypothesis of matter waves
From (5.47) and (5.48), for both non-relativistic and relativistic particles, we
have
vg = v (5.49)
That is, group velocity is always equal to particle velocity. Which mean the de
Broglie wave group always travels along with the particles, hence resolving the
anomaly associated withe the phase velocity and particle velocity.
18
5.7 Heisenberg’s uncertainty principle
Or,
dvp
vg = vp − λ (5.54)
dλ
dvp
In a normal dispersive medium3 , > 0. Therefore, vg < vp .
dλ
If vp is independent of λ (i.e., in a non-dispersive medium, such as vacuum),
then vg = vp .
∆p · ∆x = h (5.55)
y1 = a cos(ωt − kx)
and y2 = a cos[(ω + ∆ω)t − (k + ∆k)x] (5.57)
3
A dispersive medium is a medium in which waves of different wavelengths travel at different velocities.
19
5.7 Heisenberg’s uncertainty principle
is given by
∆ω ∆k
y = 2a cos t− x cos(ωt − kx) (5.58)
2 2
The particle could be found in any of the regions of width λm /2, as shown in
Figure 5.11. Hence, the uncertainty in particle position is
λm 2π
∆x = = (5.60)
2 ∆k
Now since de Broglie wavelength associated with the particle with momentum
4
Note the change in notation.
20
5.7 Heisenberg’s uncertainty principle
p is
h
λ= (5.61)
p
we have
2π 2πp
k= = (5.62)
λ h
and
2π
∆k = ∆p (5.63)
h
Therefore, from (5.60) and (5.63),
2πh h
∆x = =
2π∆p ∆p
=⇒ ∆x · ∆p = h (5.64)
∆x · ∆p ≥ h (5.65)
21
5.7 Heisenberg’s uncertainty principle
Now, the momentum p must at least be of the same order as ∆p and corre-
sponding velocity should at least be of the order of the uncertainty in velocity
∆p 0.527 × 10−20
∆v = = −31
= 0.0579 × 1011 ≈ 5.8 × 109 > c (5.70)
m 9.1 × 10
which is not possible.
Further, energy is required in order to confine a particle in a given volume.
Energy of an electron inside the nucleus would be at least
q
E = p2 c2 + m20 c4 (5.71)
22
5.7 Heisenberg’s uncertainty principle
Now,
and
Therefore,
E ≥ 10 MeV (5.74)
23
Quantum Mechanics - II
PH-1007 (Physics)
Dr. Gorky Shaw
2. The wave function associated with a particle describes its position and
state. If we know the wave function ψ(x, y, z, t) at all points in space and
all instants of time, then the particle behavior can be completely specified.
1
Given a complex number z = a + ib, a, b ∈ R, it can be shown that its complex
√ conjugate is z ∗ = a − ib.
The normal length of z is a positive (and thus real) number number, |z| = a + b2 . It can be verified that
2
2
|z| = zz ∗ .
6.1 The wave function
Note: From this point onward, our discussion will be strictly limited to
one-dimension, unless otherwise specified.
10. Integral version: The probability Pab (t) of finding a particle in the interval
x ∈ [a, b] is,
Z b Z b
2
Pab (t) = |ψ(x, t)| dx = ρ(x, t)dx (6.1)
a a
2
6.2 Schrödinger equation
E = hν = ~ω
h
and, p = = ~k (6.7)
λ
The energy and momentum are related as
p2
E= + V (x, t) (6.8)
2m
where V (x, t) is the potential energy of the particle.
3
6.2 Schrödinger equation
~2 k 2
~ω = +V (6.9)
2m
Multiplying (6.9) by ψ, we get
~2 k 2
~ωψ = ψ+Vψ (6.10)
2m
Now, from (6.5),
~ ∂ψ ∂ψ
~ωψ = − = i~ (6.11)
i ∂t ∂t
From (6.6),
~2 k 2 ~2 ∂ 2 ψ
ψ=− (6.12)
2m 2m ∂x2
Therefore, using (6.11) and (6.12) in (6.10), we get
∂ψ ~2 ∂ 2 ψ
i~ =− +Vψ (6.13)
∂t 2m ∂x2
This is the time-dependent Schrödinger equation (TDSE).
~2 d2 ψ
Eψ = − +Vψ (6.15)
2m dx2
Rearranging (6.15), we get
4
6.2 Schrödinger equation
d2 ψ 2m
+ 2 (E − V )ψ = 0 (6.16)
dx2 ~
This is the time-independent Schrödinger equation (TISE).
ψ(x, t) = Aei(kx−ωt)
= Aeipx/~−Et/~
= Aeipx/~ e−iEt/~
= φ(x)e−iEt/~ (6.17)
where
is a function of x only.
Using (6.18) in (6.13), we get
∂ h i ~2 ∂ 2 h i
−iEt/~ −iEt/~
i~ φ(x)e = − + V (x) φ(x)e
∂t 2m ∂x2
2 2
h
iE −iEt/~ ~ ∂ −iEt/~
i
=⇒ i~ − φ(x)e = − + V (x) φ(x)e
~ 2m ∂x2
(6.19)
~2 ∂ 2
Eφ(x) = − + V (x) φ(x) (6.20)
2m ∂x2
Since there is no longer any time dependence in (6.20), we replace the partial
derivative with total derivative. Hence, we get
5
6.3 Observables, operators, eigenfunctions, eigenvalues, expectation values
~2 d2
Eφ(x) = − + V (x) φ(x)
2m dx2
~2 d2
or simply, Eφ = − +V φ (6.21)
2m dx2
d2 φ 2m
+ 2 (E − V )φ = 0 (6.22)
dx2 ~
This is the time-independent Schrödinger equation (TISE).
6
6.3 Observables, operators, eigenfunctions, eigenvalues, expectation values
Mathematically,
For example,
d 2
(x ) = 2x (6.24)
dx
Order of operation is important. For example,
d d d
[xf (x)] = f (x) + x f (x) 6= x f (x) (6.25)
dx dx dx
That is, operators are not necessarily commutative.
In quantum mechanics, an operator applied to a wave function gives the
corresponding observable quantity, multiplied by the wave function.
For example, the Hamiltonian operator Ĥ
Ĥψ = Eψ (6.26)
x̂ψ = xψ
ŷψ = yψ
ẑψ = zψ (6.27)
7
6.3 Observables, operators, eigenfunctions, eigenvalues, expectation values
(a) Time-dependent:
∂
Ê(x) = i~ (6.30)
∂t
~2 ∂ 2
Ĥ(x) = − + V (x) (6.31)
2m ∂x2
Eφ = Ĥφ
or, Ĥψ = Eψ (6.32)
~2 ∂ 2
K̂(x) = − (x-component)
2m ∂x2
~2 2
K̂ = − ∇ (3-D) (6.33)
2m
Âψ = aψ (6.34)
The possible values of the constant a are called eigenvalues of the observable
associated with the operator Â. The corresponding state of the system is known
as the eigenstate of the associated observable.
d2
For example, Let ψ(x) = sin 2x and  = 2 .
dx
d2 d
Âψ = (sin 2x) = (2 cos 2x) = −4 sin 2x = −4ψ (6.35)
dx2 dx
2
Note that the wave function symbol is changed from φ to ψ in (6.32). This of course does not change the
result.
8
6.3 Observables, operators, eigenfunctions, eigenvalues, expectation values
d2
Hence, sin 2x is an eigenfunction and −4 an eigenvalue of the operator .
dx2
d
On the other hand, Let ψ(x) = sin 2x and  = .
dx
d
Âψ = (sin 2x) = 2 cos 2x 6= aψ (6.36)
dx
d
Hence sin 2x is not an eigenfunction of the operator .
dx
Z
hAi = ψ ∗ Âψdx (6.38)
where the integral is over all available space and hAi may vary in time.
9
6.4 Particle in an one-dimensional potential well of infinite height
That is,
hAi = a (6.41)
d2 ψ 2m
+ 2 (E − V )ψ = 0 (6.43)
dx2 ~
Now inside the box, V = 0. Therefore, Schrödinger equation of the particle
10
6.4 Particle in an one-dimensional potential well of infinite height
d2 ψ 2mE
+ 2 ψ=0 (6.44)
dx2 ~
Now,
p2 ~2 k 2 ~2 k 2
E= +V = +0=
2m 2m 2m
2mE
=⇒ 2
= k2 (6.45)
~
Using (6.45) in (6.44), we get
d2 ψ
2
+ k2ψ = 0 (6.46)
dx
General solution of (6.46) can be written as
ψ = 0 at x = 0 (6.48)
and
ψ = 0 at x = L (6.49)
0 = A sin 0 + B cos 0
=⇒ 0 = 0 + B
=⇒ B = 0 (6.50)
Therefore,
11
6.4 Particle in an one-dimensional potential well of infinite height
0 = A sin(kL)
=⇒ kL = nπ , n = 0, 1, 2, ... (6.52)
(the other possibility from (6.52), i.e., A = 0 is discarded because that would
imply ψ = 0 everywhere, which means the particle would not exist at all).
Further, n = 0 =⇒ k = 0, which once again means ψ = 0 for all x, and the
particle would not exist. Hence n = 0 is also discarded. Therefore, we have
kL = nπ , n = 1, 2, ...
nπ
=⇒ k = (6.53)
L
Therefore, we have
nπx
ψn (x) = A sin (6.54)
L
where we have added the suffix n to ψ to indicate dependence of ψn on n. Now,
the wave function must be normalized. That is,
Z L
|ψ(x, t)|2 dx = 1
0
Z L
2 nπx
2
=⇒ A sin dx = 1
L
2 Z L
0
A 2nπx
=⇒ 1 − cos dx = 1
2 0 L
A2
L 2nπx
=⇒ x− sin =1
2 2nπ L
A2
L 2nπL
=⇒ L− sin =1
2 2nπ L
A2 L
1
=⇒ 1− sin (2nπ) = 1
2 2nπ
A2 L
=⇒ =1
2 r
2
=⇒ A = (6.55)
L
Therefore, we have
12
6.4 Particle in an one-dimensional potential well of infinite height
r
2 nπx
ψn (x) = sin (6.56)
L L
which is the wave function associated with a particle enclosed in an infinitely
deep potential well.
The wave function given by (6.56) can be written in terms of the wavelength
2L
λn = :
n
r
2 2πx
ψn (x) = sin (6.57)
L λn
~2 k 2
En =
2m
~2 n2 π 2 n2 π 2 ~2
= · 2 = , n = 1, 2, 3... (6.58)
2m L 2mL2
n2 π 2 ~2 n2 h2
En = = , n = 1, 2, 3... (6.59)
2mL2 8mL2
Hence, we have
En ∝ n2
and, En = n2 E1 (6.60)
Note that the lowest energy level is not E = 0. Which means a particle in a
box has a zero-point energy 3 .
The variation of wave function (ψn ), probability density (|ψn |2 ), and energy
levels (En ) of the particle are shown in Figure 6.2.
A significant feature of the states of the particle in a box is the occurrence of
nodes. These are points, other than the two end-points (which are fixed by the
3
The zero-point energy is fundamentally related to the Heisenberg uncertainty principle.
13
6.5 Quantum tunneling of particles through potential barriers
Figure 6.2: (a) Wave function, (b) Probability density, and (c) Energy levels of
a particle in an one-dimensional potential well of infinite height.
4
Also, easy to remember: n gives the number of crests that the wave function has inside the box.
14
6.5 Quantum tunneling of particles through potential barriers
d2 ψI 2mE
Region I: + 2 ψI = 0 (6.62)
dx2 ~
d2 ψII 2m
Region II: + 2 (E − U0 )ψII = 0 (6.63)
dx2 ~
d2 ψIII 2mE
Region III: + 2 ψIII = 0 (6.64)
dx2 ~
Evidently, ψI and ψIII will be sinusoidal functions and ψII will be an exponential
function 5 , as qualitatively represented in Figure 6.3(b).
5
This should be obvious from Section 6.4 and earlier discussion on damped oscillations (solution for overdamped
oscillations. Try to figure it out!)
15
6.5 Quantum tunneling of particles through potential barriers
3. Tunnel diodes - Electrons tunnel through a potential barrier (at the con-
tact point of two different semiconductors) much larger than their kinetic
energies.
16
6.5 Quantum tunneling of particles through potential barriers
close to conducting surface with a voltage bias, electrons are able to tunnel
between the tip and the sample, generating an electric current.
By using piezoelectric regulators, the tip-sample height can be adjusted
to keep the tunneling current constant. Thus, path of the tip along the
sample surface can be used to obtain a gray-scale image representing the
topography of the sample surface.
17
Vector Algebra
PH-1007 (Physics)
Dr. Gorky Shaw
~ and B
The scalar product (or dot product) of two vectors A ~ is defined by
~·B
A ~ = AB cos ϕ (6.1)
6.2 Vector products
where A, B are magnitudes (lengths) of the vectors A, ~ B, ~ respectively, and ϕ
is the angle between the vectors, as shown in Figure 6.1(a). The dot product
between of vectors is based on the projection of one vector in the direction of
the other, as illustrated in Figures 6.1(b) and 6.1(c). The result of the dot
product itself is a scalar, hence the name scalar product.
The scalar product of two orthogonal vectors (ϕ = 900 ) vanishes. In this case
~·B
A ~ = AB cos 900 = 0 (6.2)
The scalar product of a vector with itself gives the square of its magnitude:
~·A
A ~ = A2 cos 00 = A2 (6.3)
~ and B
The cross product of two vectors A ~ is defined by
~ =A
C ~×B
~ = AB sin ϕ n̂ (6.4)
2
6.3 Vector algebra - component form
either one of them. The direction of the vector product is determined by the
direction of the angle ϕ, as indicated in Figures 6.2(a) and 6.2(b).
As evident from Figure 6.2,
~ × B)
(A ~ = −(B
~ × A)
~ (6.5)
~·B
A ~ = AB sin 00 = 0 (6.7)
(6.7) implies that the cross product of a vector with itself vanishes.
3
6.3 Vector algebra - component form
î · ĵ = ĵ · k̂ = k̂ · î = 0 (6.8)
î · î = ĵ · ĵ = k̂ · k̂ = 1 (6.9)
~ = Ax î + Ay ĵ + Az k̂
A (6.10)
where
~ · î, Ay = A
Ax = A ~ · ĵ, Az = A
~ · k̂ (6.11)
3. To calculate the dot product, mulitply like components, and then add:
~·B
A ~ = Ax Bx + Ay By + Az Bz (6.14)
This also follows from the properties of the unit vectors given by (6.8)
and 6.9.
In particular,
~·A
A ~ = A2x + A2y + A2z = A2 (6.15)
4. To calculate the cross product, form the determinant whose first row is
4
6.4 Special vectors
î ĵ k̂
~ ~
A × B = Ax Ay Az
Bx By Bz
= (Ay Bz − Az By )î + (Az Bx − Ax Bz )ĵ + (Ax By − Ay Bx )k̂ (6.16)
~r = xî + y ĵ + z k̂ (6.17)
5
6.4 Special vectors
where
∂f ∂f ∂f
∇f = î + ĵ + k̂ (6.24)
∂x ∂y ∂z
is a vector quantity called the gradient of f , with
∂ ∂ ∂
∇= î + ĵ + k̂ (6.25)
∂x ∂y ∂z
is maximum for ϕ = 0, i.e., when ∇f and d~l are in the same direction.
Thus, the gradient points in the direction of maximum increase of the function
f . The magnitude |∇f | gives the slope (rate of increase) along this direction.
6
6.5 Operations of the ∇ operator
6.5.1 Divergence
~ ∂ ∂ ∂
∇·F = î + ĵ + k̂ · Fx î + Fy ĵ + Fz k̂
∂x ∂y ∂z
∂Fx ∂Fy ∂Fz
= + + (6.28)
∂x ∂y ∂z
which is a scalar quantity.
7
6.5 Operations of the ∇ operator
6.5.2 Curl
î ĵ k̂
∂ ∂ ∂
∇ × F~ =
∂x ∂y ∂z
F x Fy Fz
∂Fz ∂Fy ∂Fx ∂Fz ∂Fy ∂Fx
= − î + − ĵ + − k̂ (6.29)
∂y ∂z ∂z ∂x ∂x ∂y
Curl of ~r,
î ĵ k̂
∂ ∂ ∂
∇ × ~r = =0
∂x ∂y ∂z
x y z
Hence, ~r is irrotational.
~ = −y î + xĵ.
Example 2: Find the divergence and curl of the function A
8
6.5 Operations of the ∇ operator
Solution:
Divergence of A,~
∇·A ~ = ∂(−y) + ∂x + ∂(0) = 0
∂x ∂y ∂z
Hence, ~r is solenoidal.
~
Curl of A,
î ĵ k̂
−y x 0
Tip: Try to plot and check what the function above looks like, and if the
obtained values of divergence and curl make sense.
∂ 2f ∂ 2f ∂ 2f
∇ · (∇f ) = 2
+ 2 + 2 = ∇2 f
∂x ∂y ∂z
9
6.6 Integral calculus
where d~l is the infinitesimal displacement vector along a path from the point ~a
to the point ~b.
For a closed loop, i.e., ~b = ~a, the line integral is represented by
I ~b
F~ · d~l (6.31)
~a
For example, work done by a force F~ in moving an object from the point the
R ~b
point a to the point b is given by W = ~a F~ · d~l.
d~a = dxdy k̂ or
= dydz î or
= dzdxĵ (6.33)
For a closed surface (e.g., a balloon, a hollow sphere), the surface integral is
represented by
I
F~ · d~a (6.34)
S
10
6.6 Integral calculus
amount per unit time passing through the surface, and is called flux of the
vector F~ .
dV = dxdydz (6.36)
For a vector F~ ,
Z Z Z Z
F~ dV = Fx dV î + Fy dV ĵ + Fz dV k̂ (6.37)
V V V V
Corollaries
R ~b
1. The integral ~a ∇f · d~l is path-independent (as it is dependent only on
the values of the function at the endpoints of the path, as evident from
(6.39)).
11
6.6 Integral calculus
12
Electromagnetic theory
Part I: Maxwell’s equations
PH-1007 (Physics)
Dr. Gorky Shaw
(8.1) and (8.4) are Gauss’s law in electrostatics in integral form and differen-
tial form, respectively.
(8.5) and (8.6) are Gauss’s law in magnetostatics in integral form and differ-
ential form, respectively.
Unlike (8.1) and (8.4), the RHS of (8.5) and (8.6) are zero. This indicates,
unlike electric charges, magnetic charges or magnetic monopoles do not exist.
d
Z Z ~
∂B
ε=− ~ · d~a = −
B · d~a (8.9)
dt ∂t
(In the last step of (8.9), the total derivative is replaced by the partial derivative
because in general B ~ is a function of several variables, and not just t).
Also,
I Z
ε= E ~ · d~l = (∇ × E) ~ · d~a (8.10)
2
8.1 Fundamental laws of electromagnetism
with the integral taken over any surface bounded by the loop. And
I Z
B~ · d~l = (∇ × B)
~ · d~a (8.15)
~ = µ0 J~
=⇒ ∇ × B (8.16)
(8.13) and (8.16) are Ampere’s law in integral form and differential form,
respectively.
3
8.1 Fundamental laws of electromagnetism
4
8.1 Fundamental laws of electromagnetism
~ = ∇ · (µ0 J)
∇ · (∇ × B) ~
~ = µ0 (∇ · J)
or, ∇ · (∇ × B) ~ (8.22)
∇ · J~ = 0 (8.24)
Ampere’s law in the form of ∇ · J~ = 0 given by (8.24) is valid only when the
charge density ρ is static and does not change with time.
which is consistent.
5
8.2 Maxwell’s equations of electromagnetism
~ does not change with time, we get back the original Ampere’s law
When E
~ = µ0 J~
∇×B (8.30)
(8.28) and (8.29) are modified Ampere’s law in differential form and integral
form, respectively.
6
8.2 Maxwell’s equations of electromagnetism
form are:
~ = ρ
∇·E (Gauss’s law of electrostatics (8.4)) (8.31)
0
~ = 0 (Gauss’s law of magnetostatics (8.6))
∇·B (8.32)
∂B~
~
∇×E =− (Faraday’s law of electromagnetic induction (8.12)) (8.33)
∂t
~
~ = µ0 J~ + µ0 0 ∂ E (Modified Ampere’s law (8.28))
∇×B (8.34)
∂t
7
8.2 Maxwell’s equations of electromagnetism
are:
I
~ · d~a = 0
E (8.43)
I
~ · d~a = 0
B (8.44)
!
I Z ~
∂B
~ · d~l = −
E · d~a (8.45)
∂t
!
I Z ~
∂E
~ · d~l =
B µ0 0 · d~a (8.46)
∂t
~ and H
8.2.3 Maxwell’s equations in terms of D ~
~ is the electric displacement vector D
Associated with the electric field vector E ~
given by
~ = 0 E
D ~ (8.47)
8
8.2 Maxwell’s equations of electromagnetism
9
Electromagnetic theory
Part II: Electromagnetic waves
PH-1007 (Physics)
Dr. Gorky Shaw
~
~ · (∇ × E)
B ~ · ∂B
~ = −B (9.2)
∂t
Maxwell’s fourth equation is:
~
~ = µ0 J~ + µ0 0 ∂ E
∇×B (9.3)
∂t
~ we get
Taking dot product of (9.3) with E,
~
~ · (∇ × B)
E ~ = µ0 (E
~ · J) ~ · ∂E
~ + µ0 0 E (9.4)
∂t
Using (9.2) and (9.4), and the vector identity
~ × B)
∇ · (E ~ =B
~ · (∇ × E)
~ −E
~ · (∇ × B)
~ (9.5)
we get
! !
~
∂B ~
~ × B)
∇ · (E ~ = ~
−B · − µ0 (E ~ · J) ~ · ∂E
~ + µ0 0 E
∂t ∂t
!
~ ~
~ × B)
∇ · (E ~ · ∂ B + µ0 0 E
~ =− B ~ · ∂ E − µ0 (E ~ · J)
~ (9.6)
∂t ∂t
9.1 Poynting theorem: electromagnetic energy
Now,
~
~ · ∂ B = Bx ∂Bx + By ∂By + Bz ∂Bz
B
∂t ∂t ∂t ∂t
1∂ 2
= (Bx + By2 + Bz2 )
2 ∂t
1∂ 2
= (B ) (9.7)
2 ∂t
Similarly,
~
~ · ∂ E = 1 ∂ (E 2 )
E (9.8)
∂t 2 ∂t
Using (9.7) and (9.8) in (9.6), we get
∇ · (E ~ = − 1 ∂ (B 2 + µ0 0 E 2 ) − µ0 (E
~ × B) ~ · J)
~ (9.9)
2 ∂t
Rearranging,
2
~ · J~ = − 1 ∂ 2 B 1 ~ × B)
~
E 0 E + − ∇ · (E (9.10)
2 ∂t µ0 µ0
3. The second term on the RHS represents the rate at which energy is trans-
2
9.2 Electromagnetic waves in vacuum
ported out of the volume V , across its boundary surface, by the electro-
magnetic fields.
The energy per unit time, per unit area, transported by the fields is called
~
the Poynting vector, S:
~ = 1 (E
S ~ × B)
~ (9.12)
µ0
Vacuum can be treated as a free space with no charges and currents, i.e., ρ = 0
and J~ = 0.
We have Maxwell’s third equation:
~
~ = − ∂B
∇×E (9.13)
∂t
Applying the curl to the LHS of (9.13), we get
~ = ∇(∇ · E)
∇ × (∇ × E) ~ − ∇2 E
~
~
= 0 − ∇2 E
~
= −∇2 E (9.14)
3
9.2 Electromagnetic waves in vacuum
~
∂ 2E
2~
∇ E = µ0 0 2 (9.16)
∂t
Similarly, it can be shown that
~
∂ 2B
2~
∇ B = µ0 0 2 (9.17)
∂t
These are wave equations for the electric field and magnetic field components,
respectively, of electromagnetic waves (EM waves) in free space.
~ and B
In general, each Cartesian component of the E ~ fields satisfies the
three-dimensional wave equation
1 ∂ 2y
∇2 y = (9.18)
v 2 ∂t2
where velocity of the wave is given by
1
v=√ (9.19)
µ0 0
Now,
µ0
= 10−7 N/A2 and
4π
1
= 9 × 109 Nm2 /C2 (9.20)
4π0
Therefore,
10−7 1
µ0 0 = = (9.21)
9 × 109 9 × 1016
Hence,
1
v=√ = 3 × 108 m/s = c (the speed of light). (9.22)
µ0 0
That is, electromagnetic waves propagate in free space with the speed of light.
4
9.2 Electromagnetic waves in vacuum
~
1 ∂ 2E
2~
∇E= 2 2 (9.23)
c ∂t
and
~
1 ∂ 2B
~ =
∇2 B (9.24)
c2 ∂t2
~ r, t) = E~0 ei(~k·~r−ωt)
E(~
~ r, t) = B
and B(~ ~0 ei(~k·~r−ωt) (9.25)
where E~0 and B~0 are the complex amplitudes. The physical fields are their
real parts. And, ω = ck, where ~k is the propagation vector:
~k = kn̂ = 2π n̂
λ
2πν ω
= n̂ = n̂ (9.26)
c c
where n̂ is the unit vector along the direction of propagation of the EM waves.
Applying the divergence to the E ~ field given by (9.25), we get
~ ~ i(~k·~r−ωt)
∇ · E = ∇ · E0 e
∂ ∂ ∂
= î + ĵ + k̂ · (E0x î + E0y ĵ + E0z k̂)ei(kx x+ky y+kz z−ωt)
∂x ∂y ∂z
∂
i(kx x+ky y+kz z−ωt)
∂ i(kx x+ky y+kz z−ωt)
= E0x e + E0y e
∂x ∂y
∂ i(kx x+ky y+kz z−ωt)
+ E0z e (9.27)
∂z
5
9.2 Electromagnetic waves in vacuum
~ = i(~k · E)
∇·E ~ =0
i.e., ~k · E
~ =0 (9.30)
~ = 0 in free space.
since ∇ · E
Similarly, it can be shown that
~k · B
~ =0 (9.31)
~
∂B
− ~0 ei(~k·~r−ωt)
= iω B
∂t
~
= iω B (9.34)
6
9.2 Electromagnetic waves in vacuum
i(~k × E)
~ = iω B
~
=⇒ ~k × E~ = ωB~ (9.35)
~ and B
(9.35) and (9.37) imply that E ~ are also mutually perpendicular to
each other.
kE = ωB
E ω
=⇒ = =c (9.38)
B k
Using B = µ0 H, we get
r
E µ0
Z0 = = µ0 c = = 376.72 Ω (9.39)
H 0
~ = 1 (E
S ~ × B)
~
µ0
1 ~
= [E × (~k × E)]
~ (9.40)
µ0 ω
7
9.2 Electromagnetic waves in vacuum
we get
~ = 1 [(E
S ~ ~k − (E
~ · E) ~ · ~k)E]
~
µ0 ω
E2 ~ ~ · ~k = 0)
= k (since E
µ0 ω
E2
= n̂ (using (9.26)) (9.42)
µ0 c
~
Over a complete cycle of an EM wave, average value of S:
~ = hE 2 i
hSi n̂ (9.43)
µ0 c
For a sinusoidal wave function given by
~ r, t) = E~0 ei(~k·~r−ωt)
E(~ (9.44)
we have
2 E02 2
hE i = = Erms (9.45)
2
where
E0
Erms = √ (9.46)
2
is the RMS value of the electric field.
Therefore,
2
~ Erms
hSi = n̂ (9.47)
µ0 c
8
9.3 Wave equation in a charge-free non-conducting medium
B2
1 2
U= 0 E + (9.48)
2 µ0
hU i = 0 hE 2 i = 0 Erms
2
(9.51)
Hence,
~
∂ 2E
~ = µ
∇2 E (9.54)
∂t2
9
9.4 Wave equations in terms of scalar and vector potentials
and
~
∂ 2B
2~
∇ B = µ 2 (9.55)
∂t
These are, however, valid only if and µ are independent of position and time.
Speed of EM wave in the medium,
1
v=√ (9.56)
µ
Since we always have µ > µ0 and > 0 , (9.56) indicates that we always have
v < c.
∇ · (∇ × F~ ) = 0 (9.58)
~ = 0, then corresponding to G,
Conversely, if ∇ · G ~ we can always define a vector
potential F~ such that G~ = ∇ × F~ .
10
9.4 Wave equations in terms of scalar and vector potentials
~ =∇×A
B ~ (9.61)
~ (A
(9.61), (9.62), and (9.63) imply that A, ~ + C),
~ and (A
~ + ∇f ) all give the
~
same B.
So, the magnetic vector potential is not unique, but arbitrary to the extent
of addition of a constant vector or the gradient of a scalar.
This arbitrariness allows one to choose a convenient value of A~ for mathe-
matical simplicity.
Since ∇ × (∇f ) for any scalar function f , we can define a scalar function φ
such that
!
~
∇× E ~ + ∂ A = −∇ × (∇φ) = 0
∂t
~
∂A
~
or, E + = −∇φ (9.65)
∂t
11
9.4 Wave equations in terms of scalar and vector potentials
~ is defined
φ is called the scalar potential, and from (9.65), the electric field E
in terms of the scalar potential (and the vector potential) as
~
~ = −∇φ − ∂ A
E (9.66)
∂t
~
∂A
For a time-independent field, = 0, so we get
∂t
~ = −∇φ
E (9.67)
A ~ − ∇f and φ → φ + ∂f
~→A (9.68)
∂t
and
~ → ∇ × (A
B ~ − ∇f )
=∇×A ~+0
~
=B (9.70)
12
9.4 Wave equations in terms of scalar and vector potentials
~ and B
That is, the E ~ fields remain unchanged under the transformations of φ
~ given by (9.68)
and A
Transformations of the potentials that leave the fields invariant are called
Gauge transformations.
Various gauge conditions can be applied on the potentials depending on the
requirement and convenience.
Two important and frequently used Gauge conditions are: the Lorenz gauge
condition given by
~+ 1 ∂φ
∇·A =0 (9.71)
c2 ∂t
or
~ + µ0 0 ∂φ = 0
∇·A (9.72)
∂t
and the Coulomb gauge condition given by
~=0
∇·A (9.73)
~
9.4.5 Wave equations in terms of φ and A
In free space, we have
~ =0
∇·E (9.74)
~ in terms of φ and A,
Expressing E ~ we get
!
~
∂A
∇· −∇φ − =0
∂t
∂ ~ =0
=⇒ ∇2 φ + (∇ · A) (9.75)
∂t
Using the Lorenz gauge condition (9.71) in (9.75), we get
∂ 1 ∂φ
∇2 φ + − 2 =0
∂t c ∂t
2 1 ∂ 2φ
=⇒ ∇ φ − 2 2 = 0 (9.76)
c ∂t
13
9.4 Wave equations in terms of scalar and vector potentials
(9.76) gives
1 ∂ 2φ
∇2 φ =
c2 ∂t2
2 ∂ 2φ
or, ∇ φ = µ0 0 2 (9.77)
∂t
which is the wave equation in terms of the scalar potential.
From the Lorenz condition (9.72), the first term of (9.78) is zero. Therefore,
we get
~
∂ 2A
~ = µ0 0
∇2 A
∂t2
1 ∂ 2A~
2~
or, ∇ A = 2 2 (9.79)
c ∂t
which is the wave equation in terms of the magnetic vector potential.
14
Electromagnetic theory
Part II: Electromagnetic waves
PH-1007 (Physics)
Dr. Gorky Shaw
Vacuum can be treated as a free space with no charges and currents, i.e., ρ = 0
and J~ = 0.
We have Maxwell’s third equation:
~
~ = − ∂B
∇×E (9.1)
∂t
Applying the curl to the LHS of (9.1), we get
~ = ∇(∇ · E)
∇ × (∇ × E) ~ − ∇2 E
~
~
= 0 − ∇2 E
~
= −∇2 E (9.2)
~
∂ 2E
~ = µ0 0
∇2 E (9.4)
∂t2
9.1 Electromagnetic waves in vacuum
~ and B
In general, each Cartesian component of the E ~ fields satisfies the
three-dimensional wave equation
21 ∂ 2y
∇y= 2 2 (9.6)
v ∂t
where velocity of the wave is given by
1
v=√ (9.7)
µ0 0
Now,
µ0
= 10−7 N/A2 and
4π
1
= 9 × 109 Nm2 /C2 (9.8)
4π0
Therefore,
10−7 1
µ0 0 = = (9.9)
9 × 109 9 × 1016
Hence,
1
v=√ = 3 × 108 m/s = c (the speed of light). (9.10)
µ0 0
That is, electromagnetic waves propagate in free space with the speed of light.
Thus, we can rewrite the wave equations (9.4) and (9.5) as
~
1 ∂ 2E
2~
∇E= 2 2 (9.11)
c ∂t
2
9.1 Electromagnetic waves in vacuum
and
~
1 ∂ 2B
2~
∇B= 2 2 (9.12)
c ∂t
~ r, t) = E~0 ei(~k·~r−ωt)
E(~
~ r, t) = B
and B(~ ~0 ei(~k·~r−ωt) (9.13)
where E~0 and B~0 are the complex amplitudes. The physical fields are their
real parts. And, ω = ck, where ~k is the propagation vector:
~k = kn̂ = 2π n̂
λ
2πν ω
= n̂ = n̂ (9.14)
c c
where n̂ is the unit vector along the direction of propagation of the EM waves.
Applying the divergence to the E ~ field given by (9.13), we get
~ ~ i(~k·~r−ωt)
∇ · E = ∇ · E0 e
∂ ∂ ∂
= î + ĵ + k̂ · (E0x î + E0y ĵ + E0z k̂)ei(kx x+ky y+kz z−ωt)
∂x ∂y ∂z
∂
i(kx x+ky y+kz z−ωt)
∂ i(kx x+ky y+kz z−ωt)
= E0x e + E0y e
∂x ∂y
∂ i(kx x+ky y+kz z−ωt)
+ E0z e (9.15)
∂z
The first term of RHS of (9.15)
3
9.1 Electromagnetic waves in vacuum
~ = i(~k · E)
∇·E ~ =0
i.e., ~k · E
~ =0 (9.18)
~ = 0 in free space.
since ∇ · E
Similarly, it can be shown that
~k · B
~ =0 (9.19)
~
∂B
− ~0 ei(~k·~r−ωt)
= iω B
∂t
~
= iω B (9.22)
i(~k × E)
~ = iω B
~
=⇒ ~k × E~ = ωB~ (9.23)
4
9.1 Electromagnetic waves in vacuum
~ and B
(9.23) and (9.25) imply that E ~ are also mutually perpendicular to
each other.
kE = ωB
E ω
=⇒ = =c (9.26)
B k
Using B = µ0 H, we get
r
E µ0
Z0 = = µ0 c = = 376.72 Ω (9.27)
H 0
5
Magnetic properties of matter
PH-1007 (Physics)
Dr. Gorky Shaw
10.1 Introduction
All magnetic phenomena are due to electric charges in motion. Motion of
electrons - orbital motion about the nucleus and spin motion about their own
axes, produce their own magnetic field.
Electrons possess an orbital magnetic moment as well as a spin magnetic
moment, which are randomly oriented in normal state. Hence, the resultant
magnetic moment of an atom is zero.
The moments tend to align in a certain way in response to an external mag-
netic field, and the material is said to be magnetized.
2. Magnetic flux density (B): the number of lines of force crossing per
unit area of the medium. Unit: Wb/m2 or Tesla.
B = µ0 (H + M ) (10.1)
M ∝H
or, M = χH (10.2)
10.2 Classification of magnetic materials
where χ is called the magnetic susceptibility of the material. Materials
that obey (10.2) are called linear media.
5. Relative permeability (µr ): From (10.1),
B∝H
or, B = µH (10.3)
B0 = µ0 H (10.4)
B = µ0 (H + M )
=⇒ µH = µ0 (H + χH)
µ
=⇒ =1+χ
µ0
=⇒ µr =1+χ (10.6)
2
10.2 Classification of magnetic materials
Figure 10.1: Magnetic field lines in the vicinity of (a) Diamagnetic, (b) param-
agnetic, and (c) ferromagnetic materials.
3
10.3 Magnetic moment of an atom
4
10.3 Magnetic moment of an atom
5
10.4 Paramagnetic susceptibility
µ0 nµ2m
χ= (10.13)
3kT
where µm is the atomic magnetic dipole moment.
6
10.5 Ferromagnetism
10.5 Ferromagnetism
10.5.1 Domain theory
Pierre-Ernest Weiss proposed a hypothetical concept of ferromagnetic domains.
Some of the assumptions in the domains concept are as follows (illustrated in
Figure 10.4):
3. In bulk materials, domains are randomly oriented and cancel out each
other.
5. Also, domains rotate and reorient in the direction of the applied field.
6. After the removal of the external field, domains may not completely revert
to the original random distribution.
7
10.5 Ferromagnetism
Heff = H + λM (10.17)
µ0 nµ2m
χ= (10.18)
3k(T − Tc )
where
µ0 nµ2m λ
Tc = = Cλ (10.19)
3k
is called the Curie temperature. Curie temperature is the temperature above
which ferromagnetic materials lose their permanent magnetization and exhibit
paramagnetic behaviour. So, ferromagnetic susceptibility
C
χ= (10.20)
(T − Tc )
This is known as Curie-Weiss law of ferromagnetism.
8
10.5 Ferromagnetism
9
10.5 Ferromagnetism
positive and negative values. A typical hysteresis loop is shown in Figure 10.6
and described below.
As H is varied through (0 → +Hm → −Hm → +Hm ), the B − H curve first
follows the path from 0 to a, and then traces the closed loop abcdef a.
Subsequent field cycling through +Hm → −Hm → +Hm retraces the same
closed loop.
Increasing the applied field beyond ±Hm leaves the value of B fixed at the
saturation value of ±Bm (points a and d on the curve).
The non-zero value of B (given by ±Br (points b and e, respectively)) when
the applied field is reduced to zero from ±Hm is called residual magnetism or
remanence or retentivity.
The value of H at which B eventually goes to zero (points c and f for the
(+Hm → −Hm ) and (−Hm → +Hm ) cycles, respectively) is called the coercive
field, coercive force, or Coercivity, Hc .
Materials with large coercivity are called hard ferromagnets. Those with small
coercivity are called soft ferromagnets.
Hard ferromagnets with high retentivity are useful for construction of per-
manent magnets. Examples: Steel, AlNiCo, NdFeB, SmCo.
Soft ferromagnets are useful for making cores of electromagnets. Examples:
Soft iron, Permalloy.
10
10.5 Ferromagnetism
1. high retentivity
11
Mechanical properties of matter
PH-1007 (Physics)
Dr. Gorky Shaw
Figure 11.2: Stress-strain curves for brittle, ductile, and plastic materials.
2
11.2 Elastic moduli
Hooke’s law : Within the elastic limit, stress is proportional to the corre-
sponding strain. i.e.,
stress ∝ strain
or, stress = E × strain
stress
or, E = (11.2)
strain
where the material-specific constant of proportionality E is known as elastic
modulus or modulus of elasticity and has the same unit as stress, i.e., N/m2
(SI) or dyne/cm2 (CGS).
3
11.2 Elastic moduli
shear stress
Shear modulus =
Shear strain
Ft /A Ft
In Figure 11.4, η = = (11.4)
δt/l A·γ
4
11.2 Elastic moduli
The three types of stresses and strains are not independent of each other. For
example, a longitudinal stress producing longitudinal elongation also produces
lateral compression, so that the volume of the material is preserved (as shown
in Figure 11.6).
Poisson’s ratio (σ) is a measure of the Poisson effect that describes the ex-
pansion or contraction of a material in directions perpendicular to the direction
of loading. It is the ratio between longitudinal strain (denoted by α) and lateral
strain (denote by β).
Mathematically,
lateral strain β
Poisson’s ratio = − =− (11.6)
longitudinal strain α
In Figure 11.6,
δw δd
β=− =−
w d
δl
and, α = (11.7)
l
Therefore, Poisson’s ratio
β δw/w δd/d
σ=− = = (11.8)
α δl/l δl/l
1
Poisson’s ratio has no unit and has values lying between + and −1.
2
5
11.3 Relation between elastic constants
1
The numerator of (11.9) is just F because area of the cube is unity.
2
The minus sign due to lateral contraction is absorbed in β itself.
6
11.3 Relation between elastic constants
ez = 0 (11.17)
because along Z-direction, the compressive stress along X-axis produces lateral
extension, and the extensive stress along Y-axis produces equal lateral com-
pression.
It can be shown that a force F of compression along X-axis and an equal
force of extension along Y-axis are equivalent to a shear stress of magnitude F
at 450 to X and Y axes. This shear stress produces a shearing strain of angle θ
in the XY-plane, such that
7
11.3 Relation between elastic constants
Y = 2η(1 + σ) (11.20)
Therefore,
3K − 2η
σ= (11.22)
6K + 2η
9Kη
Y = (11.24)
3K + η
8
11.3 Relation between elastic constants
Therefore, we have,
1
−1 < σ < (11.27)
2