You are on page 1of 85

Optical Properties of Solids

Mark Fox
Oxford University Press, 2001
SOLUTIONS TO EXERCISES
These notes contain detailed solutions to the Exercises at the end of each
chapter of the book, for the benet of class instructors. Please note that gures
within the solutions are numbered consecutively from the start of the document
(e.g. Fig. 1) in order to distinguish them from the gures in the book, which
have an additional chapter label (e.g. Fig. 1.1). A similar convention applies to
the labels of tables.
The author would be very grateful if mistakes that are discovered in the solu-
tions would be communicated to him. He is also very appreciative of comments
about the text and/or the Exercises. He may be contacted at the following
address:
Department of Physics and Astronomy
University of Sheeld
Hicks Building
Sheeld, S3 7RH
United Kingdom.
email: mark.fox@shef.ac.uk
c _ Mark Fox 2006
Chapter 1
Introduction
(1.1) Glass is transparent in the visible spectral region and hence we can assume
= = 0. The reectivity is calculated by inserting n = 1.51 and = 0
into eqn 1.26 to obtain R = 0.041. The transmission is calculated from
eqn 1.6 with R = 0.041 and = 0 to obtain T = 0.92.
(1.2) From Table 1.4 we read that the refractive indices of fused silica and
dense int glass are 1.46 and 1.746 respectively. The reectivities are then
calculated from eqn 1.26 to be 0.035 and 0.074 respectively, with = 0
in both cases because the glass is transparent. We thus nd that the
reectivity of dense int glass is larger than that of fused silica by a factor
of 2.1. This is why cutglass products made from dense int glass have a
sparkling appearance.
(1.3) We rst use eqns 1.22 and 1.23 to convert
r
to n, giving n = 3.01 and
= 0.38. We then proceed as in Example 1.2. This gives:
v = c/n = 9.97 10
7
ms
1
,
= 4/ = 9.6 10
6
m
1
,
R = [(n 1)
2
+
2
]/[(n + 1)
2
+
2
] = 25.6 %.
(1.4) The antireection coating prevents losses at the airsemiconductor in-
terface, and 90% of the light is absorbed when exp(l) = 0.1 at the
operating wavelength. With = 1.3 10
5
m
1
at 850 nm, we then nd
l = 1.8 10
5
m = 18 m.
(1.5) We are given n = 3.68 and we can use eqn 1.16 to work out = /4 =
0.083. We then use eqn 1.26 to nd R = 0.328. The transmission coe-
cient is calculated from eqn 1.6 as T = (10.328)
2
exp(1.32) = 0.034.
The optical density is calculated from eqn 1.8 as 0.4341.32 = 1.1. (n.b.
In principle we should take account of multiple reections as in Fig 1.10.
However, we can neglect these eects here because the absorption is very
high: see Exercise 1.9.)
(1.6) 99.8% absorption in 10 m means exp(l) = 0.002, and hence = 0.62 m
1
.
We use eqn 1.16 to nd = /4 = 3.5 10
8
. We thus have n =
1.33 + i 3.5 10
8
. The real and imaginary parts of
r
are found from
eqns 1.20 and 1.21 respectively, and thus we obtain
r
= 1.77+i 9.210
8
.
(1.7) The lter appears yellow and so it must transmit red and green light, but
not blue. The lter must therefore have absorption at blue wavelengths.
(1.8) (i) The beam that has suered no reections from the back surface passes
through the rst interface, then propagates through the material, and
nally passes through the second interface. Its intensity is thus given by:
I
1
= (1 R) e
l
(1 R) .
1
The intensity of the beam that is reected once from the back surface
is found by following its path through the medium as it is reected o
the front and back surfaces and propagates through the medium. The
intensity is thus given by:
I
2
= (1 R) e
l
R e
l
R e
l
(1 R) .
The ratio of the two intensities is thus given by R
2
exp(2l).
(ii) The window is transparent and hence nonabsorbing, implying =
= 0. The reectivity is calculated from eqn 1.26 to be 0.04, and the
ratio is thus R
2
= 1.6 10
3
.
(iii) The intensity is proportional to the square of the eld. (cf. eqn A.40.)
The eld ratio is thus [R
2
exp(2l)]
1/2
= 0.04.
(iv) The eld ratio is important at normal or near-normal incidence be-
cause multiple beam interference can occur for the exiting and reected
beams. The plate would then behave as a FabryPerot etalon. (See optics
texts e.g. Hecht (1998) for further details.) The eects of multiple beam
interference are a nuisance when it comes to extracting reliable values of
the optical constants, but their eects can be assumed to be small in limit
where the reectivity is small (as in this Exercise), and when the sample
is strongly absorbing (as in the next).
(1.9) In Exercise 1.5 we worked out R = 0.328. With = 1.3 10
6
m
1
and
l = 2 10
6
m, the intensity ratio is equal to R
2
exp(2l) = 5.9 10
4
,
while the eld ratio is its square root, namely 0.024.
(1.10) We take the log
10
of eqn 1.6 to obtain (using log
10
x = log
e
x/ log
e
10):
log
10
(T) = 2 log
10
(1 R) +l/ log
e
10 .
We can then substitute l/ log
e
10 from eqn 1.8 to obtain the required
result.
If the medium is transparent at

then we will have that T

= (1
R)
2
, where R is the reectivity at

. We assume that the reectivity


varies only weakly with wavelength. This is a reasonable assumption for
most materials if we choose

sensibly. For example, we would choose

just above the absorption edge we are trying to measure. With this
assumption, the optical density at is then given by
O.D.() = log
10
(T

) + log
10
(T

) .
Measurements of T() and T(

) thus allow the optical density to be


determined. If l is known, the absorption coecient can then be found
from eqn 1.8.
(1.11) The propagation time is equal to l/v = ln/c. The time dierence is
thus l(n
1
n
2
)/c = 27 ns. The pulses at the wavelength with the smaller
refractive index (i.e. 1500 nm) take the shorter time.
(1.12) With = 6.6 10
7

1
m, and = 1.88 10
13
rad/s, we nd
r
=

r
+ 3.97 10
5
i. The imaginary part of
r
is very large, and hence the
2
approximation
2

1
is a good one. In this approximation, we have
(with

i = (1 + i)/

2):
n =
_

r
= (3.97 10
5
)
1/2

i = 445(1 + i) .
By inserting n = = 445 into eqn 1.26, we then obtain R = 99.6 %.
(1.13) We use the same formula for the complex dielectric constant as in the
previous exercise. With = 1.88 10
13
rad/s, we nd
n
2
=
r
=
1
+ 2.94 10
5
i ,
which implies:
n =
_

r
= (2.94 10
5
)
1/2

i = 383(1 + i) .
We then nd from eqn 1.16 that = 4/ = 4.8 10
7
m
1
. Beers law
means that we set exp(l) = 0.5 for a drop in intensity by a factor of 2,
giving l = 1.4 10
8
m = 14 nm.
(1.14) It is apparent from eqn 1.26 that R = 1 when n = 1 and = 0. For zero
reectivity we thus require
r
= (n + i)
2
= 1.
(1.15) (i) We convert wavelengths to photon energies using E = hc/ to obtain
the energy level scheme shown in Fig. 1 of this solutions manual. It is thus
apparent that 0.294 eV of energy is dissipated during each absorption /
emission process.
(ii) When the quantum eciency is 100%, every absorbed photon pro-
duces a luminescent photon. The ratio of the light energy emitted to that
absorbed is then simply given by the ratio of the relevant photon energies.
The emitted power is thus (1.165/1.459) 10 = 8 W, and the dissipated
power is 2 W.
(iii) For a luminescent quantum eciency of 50% the number of photons
emitted drops by a factor of 2 compared to part (ii), and so the light
power emitted falls to 4 W. The remaining 6 W of the absorbed power is
dissipated as heat.
absorption
1.459 eV
emission
1.165 eV
relaxation(0.294 eV)
Figure 1: Energy levels scheme for Exercise 1.15.
(1.16) This is an example of Raman scattering, which is discussed in detail in
Section 10.5. Conservation of energy in the scattering process is satised
when
h
out
= h
in
h
phonon
.
With = c/, we then nd
out
= 521 nm.
3
(1.17) The transmission is given by eqn 1.9, with the wavelength dependence
of the scattering cross section given by eqn 1.10. At 850 nm we have
10% transmission, so that N
s
l = 2.30. The scattering cross-section is
11.1 times larger at 850 nm than at 1550 nm, and so we have N
s
l =
2.30/11.1 = 0.21 at the longer wavelength, implying a transmission of
81 %.
In general, the scattering losses decrease as the wavelength increases, and
hence the propagation losses decrease. Longer wavelengths are therefore
preferable for long range communication systems. At the same time, the
bres start to absorb in the infrared due to phonon absorption. 1550 nm is
the longest practical wavelength for silica bres before phonon absorption
becomes signicant.
(1.18) We again use eqn 1.9 to calculate the transmission, setting exp(N
s
l) =
0.5. This gives N
s
l = 0.69, which implies l = 3.5 m for the given values
on N and
s
.
If the wavelength is reduced by a factor of two, Rayleighs scattering law
(eqn 1.10) implies that
s
increases by a factor of 16. The length required
for the same transmission is thus smaller by a factor of 16: i.e. l =
3.5/16 = 0.22 m.
(1.19) Birefringence is an example of optical anisotropy as discussed in Sec-
tion 1.5.1, and also in Section 2.4. Ice is a uniaxial crystal, and therefore
has preferential axes, making optical anisotropy possible. Water, by con-
trast, is a liquid and has no preferential axes. The optical properties must
therefore be isotropic, making birefringence impossible.
4
Chapter 2
Classical propagation
(2.1) We envisage two displaced masses as shown in Fig. 2. The spring is
extended by a distance (x
1
x
2
) and so the force on the masses are
K
s
(x
1
x
2
). The equations of motion are therefore
m
1
d
2
x
1
dt
2
= K
s
(x
1
x
2
)
and
m
2
d
2
x
2
dt
2
= K
s
(x
2
x
1
) .
Divide the equations by m
1
and m
2
respectively and subtract them to
obtain:
d
2
dt
2
(x
1
x
2
) = K
s
_
1
m
1
+
1
m
2
_
(x
1
x
2
) .
On dening the relative displacement x = x
1
x
2
and introducing the
reduced mass , where 1/ = 1/m
1
+ 1/m
2
, we then have:

d
2
x
dt
2
= K
s
x.
This is the equation of motion of an oscillator of frequency (K
s
/)
1/2
.
m
1
m
2
x
1
x
2
rest
displaced
m
1
m
2
x
1
x
2
rest
displaced
Figure 2: Displacement of two masses as described in Exercise 2.1
(2.2) The solution is simpler if complex exponentials are used. We therefore
write the force as the real part of F
0
e
it
, and look for solutions of the
form x(t) = x
0
e
it
. On substituting into the equation of motion we then
obtain:
m(
2
i +
2
0
)x
0
e
it
= F
0
e
it
,
which implies:
x(t) =
F
0
m
1

2
0

2
i
e
it
.
5
The phase factor comes from the middle term, namely [
2
0

2
i]
1
.
On multiplying top and bottom by the complex conjugate, we nd:
1

2
0

2
i
=
(
2
0

2
) + i
(
2
0

2
)
2
+ ()
2
.
On writing this in the form:
a + ib = re
i
r(cos + i sin ) ,
we then see that the phase factor is given by:
tan =

(
2
0

2
)
.
This implies that the displacement of the oscillator is of the form:
x(t) e
i
e
it
= e
i(t)
,
which shows that the oscillator has a relative phase lag of:
= tan
1
[/(
2
0

2
)] .
(2.3) By applying the Lorentz oscillator model of Section 2.2.1, we realize that
the refractive index will have a frequency dependence as shown in Fig. 2.4,
with
0
corresponding to 500 nm. (i.e.
0
= 3.8 10
15
rad/s.) For
frequencies well above the resonance, we will just have the contribution of
the undoped sapphire crystal:
n

n(
0
) = 1.77 ,
which implies

= (1.77)
2
. The refractive index well below the resonance
can be worked out from eqn 2.19. Using the value of N given in the
question, we nd
st

= 2.23 10
3
. We thus have:
n
st
=

st
= [(1.77)
2
+ 2.23 10
3
]
1/2
.
We thus nd n
st
n

= 6.3 10
4
.
(2.4) We again use the Lorentz oscillator model of Section 2.2. The Exercise
is similar to Example 2.1, because we are dealing with a relatively small
number of absorbers and the overall refractive index will be dominated
by the host crystal. We can therefore assume n = 1.39 throughout the
Exercise. On the other hand, the host crystal is transparent at 405 nm, and
so the absorption will be determined by the impurity atoms. The other
factor we have to include is the low oscillator strength of the transition.
We therefore modify the rst equation in Example 2.1 to:
(
0
) =

2
(
0
)
2n
=
Ne
2
2n
0
m
0
1

0
f ,
where f = 9 10
5
is the oscillator strength. For the absorption line
we have
0
= 2c/405 nm = 4.65 10
15
rad/s, and = = 2 =
5.15 10
14
s
1
. With N = 2 10
26
m
3
and n = 1.39, we then nd
(
0
) = 8.6 10
6
. We nally obtain the absorption at the line centre
(405 nm) from eqn 1.16 as 270 m
1
.
This Exercise is broadly based on the results presented in the paper by
Iverson and Sibley in J. Luminescence 20, 311 (1979).
6
(2.5) The result of this Exercise works in the limit where the contribution of
the particular oscillator to the dielectric constant is relatively small, as in
Example 2.1 and the previous exercise. In this limit we have
2

1
, and
therefore (
0
) =
2
(
0
)/2n, where n =

1
, and
1
(
0
) = 1+. We then
see from eqn 2.16 that
2
(
0
) = Ne
2
/
0
m
0

0
, so that the absorption is
(cf. eqn 1.16):
(
0
) =
4(
0
)

= 4

2
(
0
)
2n
(2c/
0
) =
Ne
2
n
0
m
0
c
.
This shows that it is the linewidth that determines the peak absorption
strength per oscillator. The oscillator strength is, of course, also impor-
tant.
(2.6) The data can be analysed by comparison with Fig. 2.4.
(i) The low frequency refractive index corresponds to

st
. With n(

0
) = 2.43 from the data, we nd
st
= 5.9.
(ii) The resonant frequency is the mid point of the wiggle, i.e. 5.0
10
12
Hz.
(iii) The natural frequency is given by eqn 2.2, which implies K
s
=
2
0
.
The reduced mass is given by eqn 2.1:
1/ = 1/23 + 1/35.5 amu
1
,
which gives = 14 amu = 2.33 10
26
kg. With
0
= 2
0
= 3.1
10
13
rad/s, we nd K
s
= 23 kg s
2
. The restoring force is given by F =
K
s
x, which implies [F[ = 23 N for x equal to unity.
(iv) The oscillator density can be found from eqn 2.19.
st
= 5.9 has been
found in part (i), and

can be read from the graph as

= [n(

0
)]
2
= (1.45)
2
= 2.10. We thus have
st

= 3.8. Using the values of

0
and worked out previously, we then nd N = 3.0 10
28
m
3
.
(v) is equal to the shift between the maximum and minimum in the re-
fractive index in angular frequency units. We can only make a rough esti-
mate of because the data does not follow a simple line shape. The damp-
ing rate depends strongly on the frequency, which is why the resonance
line is asymmetric. By comparison with Fig. 2.4 we nd 110
12
Hz,
and hence = 2 6 10
12
s
1
.
(vi) The result of Exercise 2.5 tells us that = Ne
2
/n
0
c at the line
centre for a weak absorber. This limit does not really apply here, but we
can still use it to get a rough answer. On inserting the values of N, and
found above, and taking n 2, we nd 1 10
6
m
1
.
(2.7) Use = ck/n in eqn 2.25 to obtain:
v
g
=
d
dk
= c
d
dk
(k/n) =
c
n

ck
n
2
dn
dk
.
Then substitute v = c/n to obtain eqn 2.26.
7
(2.8) We consider three separate frequency regions.
(i) <
0
: In this frequency region
r
is real, and increases with frequency.
Since n =

r
, it is apparent that dn/d is positive, so that from eqn 2.26
we see that v
g
< v because k follows . Since
r
> 1, n > 1, and hence
v = c/n < c. Therefore v
g
< c.
(ii)
0
< < (
2
0
+Ne
2
/
0
m
0
)
1/2
: In this frequency region,
r
is negative.
The refractive index is purely imaginary and the wave does not propagate.
This is an example of the reststrahlen eect discussed in Section 10.2.3.
(iii) > (
2
0
+ Ne
2
/
0
m
0
)
1/2
: In this region
r
is positive and increases
with frequency, approaching unity asymptotically. dn/d is therefore pos-
itive, but we cannot use the same line of argument as part (i) because
n < 1 and therefore v > c. We must therefore work out v
g
explicitly using
eqn 2.25. It is easier to take dn/d than dn/dk, and we can use k = n/c
to nd v
g
from:
1
v
g
=
dk
d
=
n
c
+

c
dn
d
.
With n =

r
, we obtain:
dn
d
=
d
d
_
1 +
Ne
2

0
m
0
1

2
0

2
_
1/2
=
1
n
Ne
2

0
m
0

(
2
0

2
)
2
.
Hence
1
v
g
=
n
c
+
Ne
2
nc
0
m
0

2
(
2
0

2
)
2
,
=
1
nc
_
n
2
+
Ne
2

0
m
0

2
(
2
0

2
)
2
_
,
=
1
nc
_

r
+
Ne
2

0
m
0

2
(
2
0

2
)
2
_
,
=
1
nc
_
1 +
Ne
2

0
m
0

2
0
(
2
0

2
)
2
_
.
Hence we nd:
v
g
= nc
_
1 +
Ne
2

0
m
0

2
0
(
2
0

2
)
2
_
1
.
The denominator is greater than unity, and n < 1, so v
g
< c.
(2.9) (i) Consider a dipole p placed at the origin. The electric eld generated
at position vector r is given by:
E(r) =
3(p r)r r
2
p
4
0
r
5
.
The electric eld generated at the origin by a dipole p at position vector
r is therefore given by:
E(r) =
3(p (r))(r) r
2
p
4
0
r
5
=
3(p r)r r
2
p
4
0
r
5
.
8
Consider the ith dipole within the sphere illustrated in Fig. 2.8. We
assume that the dipole is oriented parallel to the z axis so that we can
write p
i
= (0, 0, p
i
). Then, on writing r
i
= (x
i
, y
i
, z
i
) in the formula for
E, we nd that the z component of the eld at the origin from the ith
dipole is:
c
i
=
p
i
(3z
2
i
r
2
i
)
4
0
r
5
i
.
We now sum over the cubic lattice of dipoles within the sphere. By sym-
metry, the x and y components sum to zero, giving a resultant eld along
the z axis of magnitude:
c
sphere
=
1
4
0

i
p
i
3z
2
i
r
2
i
r
5
i
,
as required.
(ii) If all the dipoles have the same magnitude p, then the resultant eld
is given by:
c
sphere
=
p
4
0

i
2z
2
i
x
2
i
y
2
i
r
5
i
.
The x, y and z axes are equivalent for the cubic lattice within the sphere,
and so we must have:

i
x
2
i
r
5
i
=

i
y
2
i
r
5
i
=

i
z
2
i
r
5
i
.
It is thus apparent that

i
2z
2
i
x
2
i
y
2
i
r
5
i
= 0 .
The net eld is therefore zero.
(iii) Consider a hollow sphere of radius a placed within a polarized dielec-
tric medium as illustrated in Fig. 3. (cf. Fig 2.8.) We assume that the
polarization is parallel to the z axis. The surface charge on the sphere must
balance the normal component of the polarization P. With P = (0, 0, P),
the normal component at polar angle is equal P cos , as shown in Fig. 3.
Hence the surface charge density at angle is equal to P cos . The
charge contained in a circular element at angle subtending an incremen-
tal angle d as dened in Fig. 3 is then given by:
dq = dA = P cos (2a sin a d) = 2Pa
2
cos sin d .
The x and y components of the eld generated at the origin by this in-
cremental charge sum to zero by symmetry, leaving just a z component,
with a magnitude given by Coulombs law as:
dc
z
=
dq
4
0
a
2
cos = +
P cos
2
sin d
2
0
.
9
On integrating over , we then obtain:
c
z
=
_

=0
dc
z
=
P
2
0
_

0
cos
2
sin d =
P
3
0
.
Since P is parallel to the z axis, and the x and y components of E are
zero, we therefore have:
E =
P
3
0
,
as required for eqn 2.28.
+
+
+
+
+
+
+
-
-
-
- -
-
-
P
q
dq
a
dq
Pcosq
+
+
+
+
+
+
+
+
+
+
+
+
+
+
-
-
-
- -
-
-
P
q
dq
a
dq
Pcosq
Figure 3: Denition of angles and charge increment as required for Exercise
2.9(iii).
(2.10) If
r
1 is small, the left hand side of the ClausiusMossotti relationship
becomes equal to (
r
1)/3, and we then nd:

r
= 1 +N
a
1 +,
where = N
a
, as in eqn A.4. It is apparent that
r
1 will be small
if either N is small or
a
is small. This means that we either have a low
density of absorbing atoms (as in a gas, for example), or we are working
at frequencies far away from any resonances.
(2.11) We are working with a gas, and we can therefore forget about Clausius-
Mossotti. At s.t.p. we have N
A
(Avogadros constant) molecules in a
volume of 22.4 litres. Hence N = 2.69 10
25
m
3
. We then nd
a
from:

a
= (
r
1)/N = 2.2 10
29
m
3
.
The atomic dipole is worked out from
p =
0

a
c .
The displacement of an electron by 1

A produces a dipole of 1.610


29
Cm.
Hence we require a eld of 0.810
11
V/m. The eld acting on an electron
at a distance r from a proton is given by Coulombs law as:
c =
e
4
0
r
2
.
10
On substituting r = 1

A, we nd c = 1.4 10
11
V/m. It is not surprising
that these two elds are of similar magnitude because the external eld
must work against the Coulomb forces in the molecule to induce a dipole.
(2.12) (a) If we are far away from resonance frequencies, we can ignore the
damping term, and write eqn 2.24 as:

r
n
2
= 1 +
Ne
2

0
m
0

j
f
j

2
0j

2
.
On substituting = 2c/, this becomes:
n
2
= 1 +
Ne
2

0
m
0
1
(2c)
2

j
f
j

2
j

2
j
,
where
j
= 2c/
0j
. This is of the Sellmeier form if we take:
A
j
= Ne
2
f
j

2
j
/4
2

0
m
0
c
2
.
(b) With the approximations stated in the exercise, we have:
n
2
= 1 +
A
1

2
1
= 1 +A
1
(1
2
1
/
2
)
1
.
With x
2
1
/
2
1, we can expand this to:
n
2
= 1 +A
1
(1 +x +x
2
+ ) ,
which implies:
n = [(1 +A
1
) +A
1
(x +x
2
+ )]
1/2
= (1 +A
1
)
1/2
[1 +A
1
/(1 +A
1
)(x +x
2
+ )]
1/2
= (1 +A
1
)
1/2
[1 + (1/2)A
1
/(1 +A
1
)(x +x
2
) (1/8)(A
1
/(1 +A
1
))
2
(x +x
2
)
2
+ ]
= (1 +A
1
)
1/2
+
A
1
2

1 +A
1
x +
_
A
1
2

1 +A
1

A
2
1
8(1 +A
1
)
3/2
_
x
2
+
On re-substituting for x, we then nd:
n = (1 +A
1
)
1/2
+
A
1
2

1 +A
1

2
1

2
+
_
A
1
2

1 +A
1

A
2
1
8(1 +A
1
)
3/2
_

4
1

4
,
which shows that:
C
1
= (1 +A
1
)
1/2
,
C
2
= A
1

2
1
/2(1 +A
1
)
1/2
,
C
3
= A
1
(4 + 3A
1
)
4
1
/8(1 +A
1
)
3/2
.
Note that the Cauchy formula generally applies to transparent materials
(eg glasses) in the visible spectral region. In this situation, the dispersion
is dominated by the electronic absorption in the ultraviolet. We should
then take
1
as the wavelength of the band gap, and the approximation

2
1
/
2
1 will be reasonable, as we are far away from the band gap
energy.
11
(2.13) (i) On neglecting the
4
term in Cauchys formula, we have
n = C
1
+C
2
/
2
.
On inserting the values of n at 402.6 nm and 706.5 nm and solving, we nd
C
1
= 1.5255 and C
2
= 4824.7 nm
2
, so that we have:
n = 1.5255 + 4824.7/
2
,
where is measured in nm.
(ii) The values are found by substituting into the result found in part (i)
to obtain n = 1.5493 at 450 nm and n = 1.5369 at 650 nm.
(iii) Referring to the angles dened in Fig. 4, we have
sin
in
sin
1
=
sin
out
sin
2
= n,
from Snells law. Furthermore, for a prism with apex angle , we have

1
+
2
= . With
in
= 45

and = 60

, we then nd
out
= 57.17

for
n = 1.5493 (450 nm) and
out
= 55.91

for n = 1.5369 (650 nm). Hence

out
= 1.26

.
q
in
q
out
q
2
q
1
60
q
in
q
out
q
2
q
1
60
Figure 4: Angles required for the solution of Exercise 2.13.
(2.14) The transit time is given by:
=
L
v
g
= L
dk
d
,
which, with k = n/c, becomes:
=
L
c
_
n +
dn
d
_
.
We introduce the vacuum wavelength via = 2c/, and write dn/d =
dn/d d/d, so that we then have:
=
L
c
_
n
dn
d
_
.
The dierence in the transit time for two wavelengths separated by ,
where , is given by:
=
d
d
.
12
On using the result above, we nd
d
d
=
L
c

d
2
n
d
2
,
which implies:
[[ =
L
c

d
2
n
d
2
=
L
c

2
d
2
n
d
2

,
as required.
The timebandwidth product of eqn 2.38 implies that a pulse of light
contains a spread of frequencies and therefore a spread of wavelengths.
In a dispersive medium, the dierent wavelengths will travel at dierent
velocities, and this will cause pulse broadening. With = c/, we have
[[ = (
2
/c), so that:
[[ =
L
c
_

2
d
2
n
d
2
_

c
.
The precise amount of broadening depends on the numerical value of the
timebandwidth product assumed for the pulse. For a 1 ps pulse with
t = 1, we have = 10
12
Hz, and hence [[ = 0.17 ps for the
parameters given in the exercise.
directionof
propagation
e-ray
polarization
vector
indexellipsoid
n
e
n
o
n
o
n
e
n(q )
y
z
q
directionof
propagation
e-ray
polarization
vector
indexellipsoid
n
e
n
o
n
o
n
e
n(q )
y
z
q
Figure 5: Index ellipse for the eray of a wave propagating at an angle to the
optic (z) axis of a uniaxial crystal, as required for Exercise 2.15.
(2.15) It is apparent from eqn 2.46 that
11
=
22
= n
2
o
and
33
= n
2
e
. Hence we
can re-cast the index ellipsoid in the form:
x
2
n
2
o
+
y
2
n
2
o
+
z
2
n
2
e
= 1 .
Owing to the spherical symmetry about the optic (z) axis, we can choose
the axes of the index ellipsoid so that the x axis coincides with the po-
larization vector of the oray as in Fig 2.12(a). The polarization of the
13
eray will then lie in the yz plane, as in Fig. 2.12(b). The projection of
the index ellipsoid onto the plane that contains the direction of propaga-
tion and the polarization vector of the eray thus appears as the ellipse
drawn in Fig. 5. The refractive index n() that we require is the distance
from the origin to the point of the ellipse where the E-vector cuts it. The
co-ordinates of this point are x = 0, y = n() cos and z = n() sin . On
substituting into the equation of the index ellipsoid, we then have:
0 +
n()
2
cos
2

n
2
o
+
n()
2
sin
2

n
2
e
= 1 ,
which implies:
1
n()
2
=
cos
2

n
2
o
+
sin
2

n
2
e
= 1 ,
as required.
opticaxis(z)
E-vector
45
frontsurfaceof
waveplate
y
opticaxis(z)
E-vector
45
frontsurfaceof
waveplate
y
Figure 6: Axes of the wave plate relative to the polarization vector for light
propagating along the x direction, as required for the solution of Exercise 2.16.
(2.16) We assume that the optic axis of the wave plate is along the z axis, as
shown in Fig. 6, and assume that the light is propagating along the x
direction. Light with its polarization vector at 45

to the optic axis will


then have its E-vector as shown in the gure. We resolve the E-vector
into two components of equal amplitude, one along the z axis and the
other along the y axis. The components along the z and y axes experience
refractive indices of n
e
and n
o
respectively. The phase dierence between
the two components at the rear surface of the wave plate is thus =
2Ln/, where L is the thickness, n = n
e
n
o
and is the vacuum
wavelength of the light.
The wave plate operates at a quarter wave plate when = /2. In
this situation, the output is two orthogonally polarized waves of equal
amplitude but with a /2 phase dierence between them, i.e. circularly
polarized light. The condition for this is L = /4n, which is equal to
14 m for the parameters given.
(2.17) The crystal will be isotropic if the medium has high symmetry so that
the x, y and z axes are equivalent. If not, it will be birefringent. For the
14
crystals listed in the Exercise we have:
Crystal Structure x, y, z equivalent ? Birefringent ?
(a) NaCl cubic (fcc) yes no
(b) Diamond cubic yes no
(c) Graphite hexagonal no yes
(d) Wurtzite hexagonal no yes
(e) Zinc blende cubic yes no
(f) Solid argon cubic (fcc) yes no
(g) Sulphur orthorhombic no yes
The two hexagonal crystals are uniaxial, with the optic axis lying along the
direction perpendicular to the hexagons. Sulphur has the lowest symmetry
and is the only biaxial crystal included in the list.
15
Chapter 3
Interband absorption
(3.1) The Bornvon Karmen boundary conditions are satised when:
k
x
L = n
x
L,
k
y
L = n
y
L,
k
z
L = n
z
L,
where n
x
, n
y
and n
z
are integers. The wave vector is therefore of the form:
k = (2/L)(n
x
, n
y
, n
z
) .
The allowed values of k form a grid as shown in Fig. 7. Each allowed
k-state occupies a volume of k-space equal to (2/L)
3
. This implies that
the number of states in a unit volume of k-space is L
3
/(2)
3
. Hence a
unit volume of the material would have 1/(2)
3
states per unit volume of
k-space.
The density of states in k-space is found by calculating the number of
states within the incremental shell between k-vectors of magnitude k and
k + dk, as shown in Fig. 7. This volume of the incremental shell is equal
to 4k
2
dk, and contains 4k
2
dk 1/(2)
3
= k
2
dk/2
2
states, as given in
eqn 3.15.
k
x
k
y
2p/L
k
dk
Figure 7: Grid of allowed values of k permitted by the Bornvon Karmen bound-
ary conditions, as considered in Exercise 3.1. The points of the grid are sepa-
rated from each other by distance 2/L in all three directions, giving a volume
per state of (2/L)
3
. Note that the diagram only shows the x-y plane of k-
space. The incremental shell considered for the derivation of eqn 3.15 is also
shown.
16
(3.2) With E(k) =
2
k
2
/2m

, we have:
dE
dk
=

2
k
m

.
On inserting into eqn 3.1415 and substituting for k, we nd:
g(E) = 2
k
2
/2
2

2
k/m

=
m

2
=
m

2
_
2m

2
_
1/2
=
1
2
2
_
2m

2
_
3/2
E
1/2
,
as required.
(3.3) (i) The parity of a wave function is equal to 1 depending on whether
(r) = (r). Atoms are spherically symmetric, and so measurable
properties such as the probability amplitude must possess inversion sym-
metry about the origin: i.e. [(r)[
2
= [(r)[
2
. This is satised if
(r) = (r). In other words, the wave function must have a de-
nite parity.
(ii) r is an odd function, and so the integral over all space will be zero
unless the product

f

i
is also an odd function. This condition is satised
if the two wave functions have dierent parities (parity selection rule).
Since the wave function parity is equal to (1)
l
, the parity selection rule
implies that l is an odd number.
(iii) In spherical polar co-ordinates (r, , ) we have:
x = r sin cos = r sin (e
i
+ e
i
)/2 ,
y = r sin sin = r sin (e
i
e
i
)/2i ,
z = r cos .
The selection rules on m can be derived by considering the integral over
. For light polarized along the z axis we have:
M
_
2
=0
e
im

1 e
im
d,
since z is independent of . The integral is zero unless m

= m. The
selection rule for z-polarized light is therefore m = 0.
For x or y polarized light we have:
M
_
2
=0
e
im

(e
i
e
i
) e
im
d,
which is zero unless m

= m 1. We thus have the selection rule m =


1 for light linearly polarized along the x or y axis. With circularly
polarized light we have m = +1 for
+
polarization and m = 1 for

polarization.
(3.4) The apparatus required is basically the same as for Fig. 3.13, but with
modications to take account of the fact that the required energy range of
0.30.6 eV corresponds to a wavelength range of 24 m. This is below the
band gap of silicon, and the spectrograph/silicon CCD array arrangement
17
shown in Fig. 3.13 is not appropriate. Instead, a detector with a band gap
smaller than 0.3 eV must be used, e.g. InSb. (See Table 3.2.) As InSb
array detectors are not available, a scanning monochromator with a single
channel detector would normally be used. A thermal source would suce
as the light source.
Another point to consider is that standard glass lenses do not transmit
in this wavelength range, and appropriate infrared lenses would have to
be used, e.g. made from CdSe. (See Fig. 1.4(b).) Also, since the data is
taken at room temperature, no cryostat is needed.
Figure 8 gives a diagram of a typical arrangement that could be used.
InAs sample
scanning
monochromator
computer
:
infraredlenses
whitelight
source
InSb
detector
InAs sample
scanning
monochromator
computer
:
infraredlenses
whitelight
source
InSb
detector
InSb
detector
Figure 8: Apparatus for measuring infrared absorption spectra in the range
24 m, as discussed in Exercise 3.4.
(3.5) The type of band gap can be determined from an analysis of the variation
of the absorption coecient with photon energy. The material is direct
or indirect depending on whether a graph of
2
or
1/2
against is a
straight line. Other factors to consider are that the absorption is much
stronger in a direct-gap material, and that the temperature dependence of
is expected to be dierent. In an indirect gap material, phonon-assisted
absorption mechanisms will freeze out as the temperature is lowered.
(3.6) Plot
2
and
1/2
against as shown in Fig. 9. In the range 2.2
2.7 eV, the graph of
2
is a straight line with an intercept at 2.2 eV. We
thus deduce that GaP has an indirect band gap at 2.2 eV. For > 2.7 eV,
the graph of
1/2
is a straight line with an intercept at 2.75 eV. (Note
the huge dierence in the two axis scales, which is a further indication
of the indirect nature of the transitions below 2.75 eV, and their direct
nature above 2.75 eV.) We thus deduce that GaP has a direct gap at
2.75 eV. Hence we conclude that the conduction band has two minima:
one away from k = 0 at 2.2 eV, and another at k = 0 at 2.75 eV.
(3.7) A wavelength of 1200 nm corresponds to a photon energy of 1.03 eV. This
is above the direct gap of germanium at 0.80 eV, and thus the absorption
will be given by (cf. eqn 3.25):
= C( 0.80)
1/2
.
18
2.2 2.4 2.6 2.8 3.0
0
20
40
60
80
a
2
(
1
0
1
2
m
-
2
)
Energy(eV)
0
200
400
600
800
1000
a
1
/
2
(
m
-
1
/
2
)
2.2 2.4 2.6 2.8 3.0
0
20
40
60
80
a
2
(
1
0
1
2
m
-
2
)
Energy(eV)
0
200
400
600
800
1000
a
1
/
2
(
m
-
1
/
2
)
Figure 9: Analysis of GaP absorption data as required for Exercise 3.6.
The scaling coecient C can be determined from the data in Fig. 3.10:

2
0.5 10
12
m
2
at 0.90 eV implies C 2.2 10
6
m
1
eV
1/2
. With
this value of C, we then nd 1 10
6
m
1
at 1.02 eV.
(3.8) (i) We consider transitions 1 and 2 in Fig. 3.5. The k vectors for the
transitions can be worked out from eqn 3.23. The appropriate parameters
are read from Table C.2 as follows: E
g
= 1.424 eV, m

e
= 0.067m
e
, m

hh
=
0.5m
e
, and m

lh
= 0.08m
e
. For the heavy hole and light hole transitions we
nd from eqn 3.22 that
hh
= 0.059 m
e
and
lh
= 0.036 m
e
respectively.
Hence for = 1.60 eV, we nd k = 5.3 10
8
m
1
for the heavy holes
and k = 4.1 10
8
m
1
for the light holes.
(ii) The air wavelength of the photon is 775 nm. The wavelength inside
the crystal is reduced by a factor n. The photon wave vector inside the
crystal is therefore given by:
k =
2
(/n)
= 3.0 10
7
m
1
.
This is more than an order of magnitude smaller than the electron wave
vector, and hence the approximation in eqn 3.12 is justied.
(iii) Equation 3.24 shows that the joint density of states is proportional
to
3/2
. Hence the ratio of the joint density of states for heavy and light
hole transitions with the same photon energy is equal to:
(
hh
/
lh
)
3/2
= (0.059/0.036)
3/2
= 2.1 .
(iv) It is apparent from Fig.3.5 that the lowest energy (i.e. k = 0) splito
hole transition occurs at = E
g
+ . Reading a value of = 0.34 eV
from Table C.2, we nd = 1.76 eV, which is equivalent to = 704 nm.
(3.9) (i) The lowest conduction band states of silicon are p-like at the point
(i.e k = 0). The spin-orbit splitting is small, and the j = 1/2 and j = 3/2
conduction bands states are degenerate at k = 0, but not for nite k.
19
Electric-dipole transitions from the p-like valence band states are forbidden
to these p-like conduction band states at k = 0. The rst dipole-allowed
transition is to the s-like antibonding state at 4.1 eV. Hence the direct
gap at the point is equal to 4.1 eV.
(ii) The discussion of the atomic character of bands given in Section 3.3.1
only applies at the point where k = 0 and we are considering stationary
states. This means that electric-dipole transitions can be allowed at the
zone edges, even though they are forbidden at k = 0.
(3.10) At low temperatures, phonon absorption is impossible, and the indirect
transition must proceed by phonon emission, with a threshold energy of
E
ind
g
+ . The band structure diagram of germanium given in Fig. 3.9
shows that the indirect gap occurs at the L-point of the Brillouin zone. We
therefore need a phonon with a wave vector equal to the k vector at the
L-point. The energies of these phonons are given in Table 3.1. The lowest
energy energy phonon is the TA phonon with an energy of 0.008 eV. The
absorption threshold would thus occur at E
ind
g
+0.008 eV, i.e. at 0.75 eV.
(3.11) The absorption coecient with a eld applied is given by eqn 3.26. The
absorption decreases exponentially for < E
g
, and this produces an ex-
ponential absorption tail below E
g
. Although there is no clear absorption
edge as for the case at zero eld, a reasonable point to take is when the
absorption has decayed by a factor e
1
from its value at E
g
. This occurs
when
4

2m

e
3[e[c
(E
g
)
3/2
= 1 .
We are looking for the eld at which this condition is satised for =
(E
g
0.01) eV. We thus need to solve:
4

2m

e
3[e[c
(0.01 eV)
3/2
= 1 .
With m

e
= 0.067m
e
, we nd c = 1.8 10
6
V/m.
v
B
r
w
F
v
B
r
w
F
Figure 10: Force acting on a particle moving in a magnetic eld pointing into
the paper, as required for Exercise 3.12. The charge is assumed to be positive.
(3.12) We consider a particle of charge q, mass m and velocity v moving in a
magnetic eld B. The particle experiences the Lorentz force F = qv B
which is at right angles both to the velocity and the eld, as shown in
Fig. 10. This perpendicular force produces circular motion at angular
20
frequency with radius r. We equate the central force with the Lorentz
force to obtain, with v = r:
m
2
r = qrB,
which, with [q[ = e, implies:
= eB/m,
as required.
With a magnetic eld pointing in the z direction, the motion in the x-y
plane is quantized, but the motion in the z direction is free. We have
seen above that, in the classical analysis, the eld causes circular motion.
The quantized motion will therefore correspond to a quantum harmonic
oscillator. These quantized states are called Landau levels. The Bloch
wave functions of eqn 3.78 are therefore modied to the form:

n
(r) u(r)
n
(x, y) e
ikzz
,
where
n
is a harmonic oscillator function, and n is the quantum number
of the Landau level. The selection rule for transitions between the Landau
levels can be deduced by repeating the derivation in Section 3.2 with the
modied wave functions. For x-polarized light, the matrix element is now
given by:
M
_
u

f
(r)
n
(x, y)e
ik

z
z
xu

i
(r)
n
(x, y)e
ik
z
z
d
3
r ,

_
unit cell
u

f
(r) xu

i
(r) d
3
r
n
[
n
) ,
where we assumed k

z
= k
z
as usual in the second line. The electricdipole
matrix element is therefore proportional to the overlap of harmonic oscil-
lator wave functions with dierent values of n. Now harmonic oscillator
functions form an orthonormal set, and so the wave functions of diering
n are orthogonal. Hence the matrix element is zero unless n

= n, i.e.
n = 0.
(3.13) (i) Let z be the free direction. Apply Bornvon Karmen boundary con-
ditions as in Exercise 3.1 to show that k
z
= 2n/L, where n is an integer.
There is therefore one k state in a distance 2/L, so that the density of
states in k-space is 1/2 per unit length of material. For free motion in
the z direction we have E =
2
k
2
/2m for a particle of mass m, which
implies:
dE/dk =
2
k/m =
_
2E/m.
The density of states in energy space is then worked out from eqn 3.14:
g
1D
(E) = 2
g(k)
dE/dk
= 2
1/2

_
2E/m
= (2m/Eh
2
)
1/2
.
We thus see that g
1D
(E) E
1/2
.
(ii) The threshold energy will be equal to the band gap E
g
of the semi-
conductor as for a 3-D material. Fermis golden rule indicates that the
21
0 2 4 6 8 10
0
2
4
6
A
b
s
o
r
p
t
i
o
n
(
a
.
u
.
)
Energyrelativeto E
g
in arb.units
(ii) (iii)
0 1 2 3
0
2
4
6
A
b
s
o
r
p
t
i
o
n
(
a
.
u
.
)
Energyinunitsof hw
L
relativeto E
g
0 2 4 6 8 10
0
2
4
6
A
b
s
o
r
p
t
i
o
n
(
a
.
u
.
)
Energyrelativeto E
g
in arb.units
0 2 4 6 8 10
0
2
4
6
A
b
s
o
r
p
t
i
o
n
(
a
.
u
.
)
Energyrelativeto E
g
in arb.units
(ii) (iii)
0 1 2 3
0
2
4
6
A
b
s
o
r
p
t
i
o
n
(
a
.
u
.
)
Energyinunitsof hw
L
relativeto E
g
0 1 2 3
0
2
4
6
A
b
s
o
r
p
t
i
o
n
(
a
.
u
.
)
Energyinunitsof hw
L
relativeto E
g
Figure 11: (ii) Absorption of a one-dimensional semiconductor as discussed in
Exercise 3.13. (iii) Absorption for a system with quantized landau levels in
two directions and free motion in the third.
L
is the Landau level angular
frequency.
absorption is proportional to the density of states. Since g
1D
(E) E
1/2
,
we therefore expect ( E
g
)
1/2
, for > E
g
. See Fig. 11(ii).
(iii) The magnetic eld quantizes the motion in two dimensions to give
Landau levels, leaving the particle free to move in the third dimension.
Optical transitions are possible between Landau levels with the same value
of n. (See Exercise 3.12.) These will occur at energies given by (c.f. eqn
3.32):
E
n
= E
g
+ (n + 1/2)
L
,
where

L
= eB/m

e
+eB/m

h
= eB/.
Each Landau level transition has a 1-D density of states due to the free
motion parallel to the eld. Hence for each Landau level we expect:
( E
n
)
1/2
.
The total absorption is found by adding the absorption for each Landau
level transition together, as shown in Fig. 11(iii).
When comparing to the experimental data in Fig. 3.7, we expect () to
diverge each time the frequency crosses the threshold for a new value of
n. These divergences are broadened by scattering. We therefore see dips
in the transmission at each value of that satises eqn 3.32.
(iv) Minima in the transmission occur at 0.807 eV, 0.823 eV, 0.832 eV and
0.844 eV, with an average separation of 0.012 eV. We equate this separation
energy to eB/, and hence nd = 0.035m
e
. If m

h
m

e
, we will have
= m

e
, and hence we deduce m

e
0.035m
e
. The lowest energy transition
occurs at E
g
+ (1/2)eB/, which implies E
g
= 0.80 eV.
The values we have deduced refer to the point of the Brillouin zone.
(See Fig. 3.9.) The indirect transitions for > 0.66 eV are too weak to
be observed compared to the direct transitions above 0.80 eV.
(3.14) The responsivity is calculated using eqn 3.38. The device will be most
ecient if it has 100% quantum eciency, and so we set = 1, giving:
Responsivity
max
=
e

(1 e
l
) .
22
On inserting the values given in the Exercise, we nd responsivities of
0.46 A/W at 1.55 m and 1.05 A/W at 1.30 m.
(3.15) (i) The p-i-n diode structure is described in Appendix D. The p- and
n-regions are good conductors, whereas the i-region is depleted of free
carriers and therefore acts like an insulator. We thus have two parallel
conducting sheets separated by a dielectric medium, as in a parallel-plate
capacitor.
(ii) The capacitance of a parallel-plate capacitor of area A, permittivity

0
and plate separation d is given by:
C =
A
r

0
d
.
In applying this formula to a p-n junction, we should use the depletion
region thickness for d. In the case of a p-i-n diode, we assume that the
depletion lengths in the highly doped p- and n-regions are much smaller
than the i-region thickness, so that can set d = l
i
. We then nd C = 10 pF
for a silicon p-i-n diode with A = 10
6
m
2
,
r
= 11.9 and l
i
= 10
5
m.
(iii) The electric eld for an applied bias of 10 V can be calculated from
eqn D.3 as:
c =
1.1 (10)
10
5
= 1.1 10
6
V/m.
The electron and hole velocities can be calculated from the eld and the
respective mobilities:
v = c .
This gives v = 1.7 10
5
m/s for the electrons and v = 5 10
4
m/s for
the holes. The drift time is nally calculated from t = l
i
/v, which gives
60 ps for the electrons and 200 ps for the holes.
Note that velocity saturation eects have been neglected here. The linear
relationship between the velocity and eld breaks down at high elds,
and the velocity approaches a limiting velocity called the saturation drift
velocity. The eld in this example is quite large, and the transit times will
actually be slightly longer than those calculated from the mobility due to
the saturation of the velocity.
(iv) With R = 50 and C = 10 pF, we nd RC = 500 ps. To obtain the
same transit time, we need a velocity of
v = l
i
/t = 10
5
/5 10
10
= 2 10
4
m/s .
This velocity occurs for an electric eld of v/
e
= 1.3 10
5
V/m. We
nally nd the voltage from eqn D.3:
c = 1.3 10
5
=
[1.1 V [
10
5
,
which gives V = 0.2 V. We therefore need to apply a reverse bias of
0.2 V.
The point about this last part of the question is to make the students
think about the factors that limit the response time of the photodetector.
23
In most situations, the time constant will be capacitance-limited because
the transit time is much shorter than the RC time constant. It is only in
small-area low-capacitance devices that we need to worry about the drift
transit time.
24
Chapter 4
Excitons
(4.1) The Hamiltonian for the hydrogen atom has three terms corresponding to
the kinetic energies of the proton and electron, and the Coulomb attraction
between them. On writing the position vectors of the electron and proton
as r
1
and r
2
, the Hamiltonian thus takes the form:

H =

2
2m
1

2
1


2
2m
2

2
2

e
2
4
0
[r
1
r
2
[
,
where m
1
= m
e
, m
2
= m
p
, and

2
i
=

2
x
2
i
+

2
y
2
i
+

2
z
2
i
.
We introduce the relative co-ordinate r and the centre of mass co-ordinate
R according to:
r = r
1
r
2
R =
m
1
r
1
+m
2
r
2
m
1
+m
2
.
Now

x
1
=
x
x
1

x
+
X
x
1

X
=

x
+
m
1
m
1
+m
2

x
2
=
x
x
2

x
+
X
x
2

X
=

x
+
m
2
m
1
+m
2

X
,
which implies

x
2
1
=

2

x
2
+
2m
1
m
1
+m
2

xX
+
_
m
1
m
1
+m
2
_
2

X
2

x
2
1
=

2

x
2

2m
2
m
1
+m
2

xX
+
_
m
2
m
1
+m
2
_
2

X
2
.
It is then apparent that:
1
m
1

x
2
1
+
1
m
2

x
2
2
=
_
1
m
1
+
1
m
2
_

x
2
+
1
m
1
+m
2

X
2
.
Similar results may be derived for the other co-ordinates, so that we have:
1
m
1

2
1
+
1
m
2

2
2
=
_
1
m
1
+
1
m
2
_

2
r
+
1
m
1
+m
2

2
R
25
On introducing the total mass M and reduced mass m according to:
M = m
1
+m
2
,
1
m
=
1
m
1
+
1
m
2
,
and substituting into the Hamiltonian, we then nd

H =

2
2M

2
R


2
2m

2
r

e
2
4
0
[r[
.
The three terms in the Hamiltonian now represent respectively:
the kinetic energy of the whole atom,
the kinetic energy due to relative motion of the two particles,
the Coulomb attraction, which depends only on the relative co-ordinate.
The Hamiltonian thus breaks down into two terms:

H =

H
whole atom
+

H
relative
,
where:

H
whole atom
=

2
2M

2
R

H
relative
=

2
2m

2
r

e
2
4
0
[r[
.
These two terms correspond respectively to the motions of:
a free particle of mass M moving with the centre of mass co-ordinate,
a particle of mass m experiencing the Coulomb force and moving
relative to a stationary origin.
The Hamiltonian is thus separable into the free kinetic energy of the atom
as a whole and the bound motion of the electron relative to the nucleus.
For the latter case, we describe the motion by using the reduced mass m,
rather than the individual electron mass. The reduced mass correction
does not make much dierence for hydrogen itself, where m
2
m
1
and
hence m m
1
, but it is very important for excitons, where the electron
and hole masses are typically of the same order of magnitude.
(4.2) (i) The Hamiltonian describes the relative motion of the electron and hole
as they experience their mutual Coulomb attraction within the semicon-
ductor. The kinetic energy of the exciton as a whole is not included. As
explained in Exercise 4.1, the appropriate mass is the reduced electron-
hole mass , and the co-ordinate r is the position of the electron relative
to the hole. The rst term represents the kinetic energy, and the second is
the Coulomb potential. The inclusion of
r
in the Coulomb terms accounts
for the relative permittivity of the semiconductor.
(ii) We substitute into the Schrodinger equation with the
2
operator
written in spherical polar co-ordinates:

2
=
1
r
2

r
_
r
2

r
_
+
1
r
2
sin

_
sin

_
+
1
r
2
sin
2

2
.
26
Since does not depend on or , the Schrodinger equation becomes:

2
2
1
r
2

r
_
r
2

r
_

e
2
4
0

r
r
= E .
On substituting = C exp(r/a
0
), we obtain:
_


2
2a
2
0
+

2
a
0
r

e
2
4
0

r
r
_
= E .
The wave function is therefore a solution if
+

2
a
0
r

e
2
4
0

r
r
= 0 ,
which implies
a
0
=
4
0

2
e
2

r
m
0

a
H
,
where a
H
is the hydrogen Bohr radius. With this value of a
0
, we then
nd:
E =

2
2a
2
0
=
e
4
8
2
r

2
0
h
2


m
0
1

2
r
R
H
,
where R
H
is the hydrogen Rydberg energy.
The normalization constant is found by solving:
_

r=0
_

=0
_
2
=0

r
2
sin drdd = 1 .
This gives:
4C
2
_

r=0
r
2
e
2r/a0
dr = 4C
2

a
3
0
4
= 1 ,
which implies:
C =
_
1
a
3
0
_
1/2
.
In the language of atomic physics, the wave function considered here is in
fact a 1s state.
(4.3) The radial probability density P(r) is proportional to r
2
[R(r)[
2
, where
R(r) is the radial part of the wave function. For the 1s wave function of
Exercise 4.2 we then have:
P(r) r
2
e
2r/a0
.
On dierentiating, we nd that this peaks at a
0
. The expectation value
of r is found from
r) =
_

r=0
_

=0
_
2
=0

r r
2
sin drdd,
= 4
_
1
a
3
0
__

r=0
r
3
e
2r/a
0
dr ,
= (3/2)a
0
.
The peak of the probability density and the expectation value thus dier
by a factor of 3/2.
27
(4.4) The purpose of this exercise is to familiarize the student with the varia-
tional method and demonstrate that it works. This will be useful to us
later for obtaining an estimate of the wave function and binding energy
of the excitons in quantum wells. (See Exercise 6.9.)
(i) We require a spherically symmetric wave function with a functional
forms that makes the probability density peak at some nite radius and
then decay to zero for large values of r. The given wave function satises
these criteria. It is actually a correctly normalized 1s-like atomic wave
function, but with a variable radius parameter .
(ii) We substitute into the Hamiltonian with the
2
written in spherical
polar co-ordinates, as in Exercise 4.2. Since again depends only on r,
this gives:

H =

2
2
1
r
2

r
_
r
2

r
_

e
2
4
0

r
r
.
On evaluating the derivatives, we nd:

H =
_


2
2
2
+
_


e
2
4
0

r
_
1
r
_ _
1

3
_
1/2
e
r/
.
The expectation value is then given by:
E) =
_

r=0
_

=0
_
2
=0


H r
2
sin drdd
= 4
_
1

3
__

r=0
_


2
2
2
r
2
+
_


e
2
4
0

r
_
r
_
e
2r/
dr
=

2
2
2

e
2
4
r

.
(iii) On dierentiating E) with respect to , we nd a minimum when =
4
0

2
/e
2
. The value of E) at this minimum is E) = e
4
/8h
2

2
0

2
r
.
(iv) The value of that minimizes E) is equal to a
0
, and the minimum
value of E) is the same energy as that found in Exercise 4.2.
The variational method gives exactly the right energy and wave function
here because our guess wave function had exactly the right functional
form. In other situations, this will not be the case, and the energy and
wave functions obtained by the variational method will only be an approx-
imation to the exact ones. The accuracy of the results will depend on how
good a guess we make for the functional form of the trial wave function.
(4.5) (i) The electron performs circular motion around the nucleus with quan-
tized angular momentum equal to n. The orbits are stable, and photons
are only emitted or absorbed when the electron jumps between orbits.
(ii) The central force for the circular motion is provided by the Coulomb
attraction, and the electron velocity v must therefore satisfy:
v
2
r
=
e
2
4
0

r
r
2
,
28
where r is the radius of the orbital and is the reduced mass. (See
Exercise 4.2 for a discussion of why it is appropriate to use the reduced
mass here.) The Bohr assumption implies that the angular momentum is
quantized:
L = vr = n .
On eliminating v from these two equations we nd:
r =
4
0

2
e
2
n
2

m
0


r
n
2
a
H
,
where a
H
= 4
0

2
/m
0
e
2
is the hydrogen Bohr radius. The energy is
found from:
E =
1
2
v
2

e
2
4
0

r
r
.
On solving for v and substituting, we nd:
E =
e
4
8h
2

2
0

2
r
n
2


m
0
1

2
r
R
H
n
2
,
where R
H
= m
0
e
4
/8h
2

2
0
is the hydrogen Rydberg energy. These are the
same results as in eqns 4.1 and 4.2.
(iii) E is identical to the exact solution of the hydrogen Schrodinger equa-
tion.
(iv) The radius for n = 1 corresponds to the peak in the radial probability
density for the ground state 1s wave function. For higher atomic shells, the
Bohr radius corresponds to the peak radial density of the wave function
with the highest value of the orbital quantum number l, namely l = n 1.
(4.6) The binding energies and radii can be calculated from eqns 4.1 and 4.2
respectively. The reduced mass is calculated from eqn 3.22 to be =
0.179m
0
, and with
r
= 7.9 we then nd R
X
= 39.1 meV and a
X
= 2.3 nm.
Hence we obtain E(1) = 39.1 meV, E(2) = 9.8 meV, r
1
= 2.3 nm and
r
2
= 9.3 nm.
We expect the excitons to be stable if E(n) > k
B
T. At room temperature
k
B
T = 25 meV, so that we would expect the n = 1 exciton to be stable,
but not the n = 2 exciton.
(4.7) The reduced mass is calculated from eqn 3.22 to be 0.056m
0
, and hence we
calculate R
X
= 4.9 meV from eqn 4.1 using
r
= 12.4. Equation 4.4 gives
the wavelengths of the n = 1 and n = 2 excitonic transitions as 873.7 nm
and 871.4 nm respectively. Hence = 2.3 nm.
(4.8) We assume that the exciton has a Lorentzian line with a centre energy of
1.5149 eV and a full width at half maximum of 0.6 meV. We then expect
the absorption and refractive index to follow a frequency dependence as
in Fig. 2.5. This implies that the maximum in the refractive index would
occur at
0
/2, i.e. [1.5149 (0.6/2)] = 1.5146 eV.
The peak value of the refractive index can be worked out from eqns 2.17
21. At line centre, we have from eqns 1.16 and 1.22:
=
4

=
4
2
2n
,
29
which implies
2
(
0
) = 1.37 if we assume that n 3.5 (i.e. the excitonic
contribution to the refractive index is relatively small.) From eqn 2.21 we
have

2
(
0
) = 1.37 = (
st

= (
st

)
1.5149
0.6
,
implying (
st

) = 5.4 10
4
. With
st
= (3.5)
2
= 12.25, we can
then nd the dielectric constant at (
0
/2) using eqns 2.2021. This
gives (
0
/2) = 12.93 + 0.685i. We nally nd n from eqn 1.22 to be
3.60. This justies the approximation n 3.5 in the calculation of
2
(
0
)
above.
(4.9) The n = 1 and n = 2 excitons have energies of E
g
R
X
and E
g
R
X
/4
respectively. Hence the n = 1 2 transition occurs at a photon energy
of 3R
X
/4. We calculate R
X
= 4.2 meV from eqn 4.1, and hence con-
clude that the transition energy is 3.15 eV. This occurs at a wavelength of
394 m.
(4.10) The magnitude of the electric eld between an electron and hole sepa-
rated by a distance of r in a medium of relative permittivity
r
is given
by Coulombs law as:
c =
e
4
0

r
r
2
.
In the Bohr model we have (see e.g. Exercise 4.5):
[E(n)[ =
e
4
8(
0

r
hn)
2
and
r
n
=
4
0

2
n
2
e
2
.
It is thus apparent that for n = 1 we have:
2[E(1)[
er
1
=
2R
X
ea
X
=
e
4
0

r
_
e
2

r
h
2
_
2
=
e
4
0

r
r
2
.
Hence c = 2R
X
/ea
X
.
(4.11) We use eqn 3.22 to nd = 0.028m
0
, and then use eqns 4.12 with

r
= 16 to calculate E(1) = 1.5 meV and r
1
= 31 nm. Then, using the
result of the previous exercise, we nd that the internal eld in the exciton
has a magnitude of 9.7 10
4
V/m. We expect the excitons to be ionized
whenever the applied eld exceeds this value. The voltage at which this
occurs can be worked out from eqn 4.5. With V
bi
= 0.74 V and l
i
=
2 10
6
m, we nd c = 9.7 10
4
V/m for V
0
= +0.55 V. Hence we need
to apply a forward voltage of 0.55 V.
(4.12) The cyclotron energy is given by eqn 4.6 and the exciton Rydberg by
eqn 4.1. The condition
c
= R
X
can thus be written:
eB

=
R
H
m
0

2
r
,
30
which, on solving for B, gives:
B =

2
R
H
m
0

2
r
e
.
On inserting the numbers for GaAs we nd B = 1.8 T.
(4.13) The magnetic eld is related to the vector potential by
B = A.
Thus for A = (B
0
/2)(y, x, 0), we obtain:
B =
B
0
2
_
_

z
_
_

_
_
y
x
0
_
_
=
_
_
0
0
B
0
_
_
.
In the analysis following eqn B.17, we neglected the term in A
2
because
the magnetic vector potential of a light wave is small. However, we are
now considering the interaction between an exciton and a strong magnetic
eld, and it is precisely the term in A
2
that gives rise to the diamagnetic
shift. It is then apparent from eqn B.17 that the diamagnetic perturbation
for a vector potential of A = (B
0
/2)(y, x, 0) is given by:

=
e
2
A
2
2m
0
=
e
2
B
2
0
8m
0
(x
2
+y
2
) .
The diamagnetic energy shift is calculated from
E = [

[) =
e
2
B
2
0
8m
0
[(x
2
+y
2
)[) .
In the case of an exciton, the wave functions are spherically symmetric so
that:
x
2
) = y
2
) = z
2
) = r
2
)/3 = r
2
n
/3 .
The total shift is obtained by summing the energy shifts of the electrons
and holes to obtain:
E =
e
2
B
2
0
8m

e
2r
2
n
3
+
e
2
B
2
0
8m

h
2r
2
n
3
=
e
2
B
2
0
r
2
n
12
.
(4.14) It is shown in Example 4.1 that the radius of the ground state exciton in
GaAs is 13 nm. Hence for = 0.05m
0
we calculate E = +4.9 10
5
eV
at B = 1 T. The wavelength shift can be calculated from:
=
d
dE
E =
hc
E
2
E = 0.026 nm.
(4.15) The given eective masses imply = 0.17m
0
from eqn 3.22, so that we
can calculate r
1
= 3.1 nm and r
2
= 12.3 nm from eqn 4.2 using
r
= 10.
The Mott densities are estimated from eqn 4.8. Hence we obtain N
Mott
=
8.1 10
24
m
3
and 1.3 10
23
m
3
for the n = 1 and n = 2 excitons
respectively.
31
(4.16) The classical theory of equipartition of energy states that we have a
thermal energy of k
B
T/2 per degree of freedom. A free particle has three
degrees of freedom corresponding to the three velocity components. Hence
the total thermal kinetic energy is given by:
p
2
2m
=
3
2
k
B
T .
This implies a thermal de Broglie wavelength of:

deB
=
h
p
=
h

3mk
B
T
.
The particle density at the BoseEinstein condensation temperature is
given by eqn 4.9. On setting N = 1/r
3
, where r is the average inter-
particle distance, we nd
1
r
= (2.612)
1/3

2mk
B
T
h
.
Hence:
r

deB
=
1
(2.612)
1/3
_
3
2
= 0.50 .
(4.17) BoseEinstein condensation refers to the quantization of the kinetic en-
ergy due to free translational motion. In applying this to excitons, we need
to consider the centre of mass motion of the whole exciton, which behaves
as a composite boson. The appropriate mass to use to calculate the con-
densation temperature is therefore the total mass of the exciton, namely
(m

e
+m

h
) = 1.7m
0
. On substituting into eqn 4.9 with N = 1 10
24
m
3
and solving for T
c
, we nd T
c
= 17.2 K.
(4.18) The excitonic radii calculated from eqn 4.2 are r
1
= 0.85 nm and r
2
=
3.4 nm. For the n = 1, the radius is comparable to the unit cell size a and
thus the Wannier model is invalid. On the other hand, the n = 2 exciton
satises the condition r a, and the Wannier model is valid.
32
Chapter 5
Luminescence
(5.1) Indirect transitions involve the absorption or emission of a phonon to con-
serve momentum. This means that they have a low transition probability,
and hence low quantum eciency. See Section 5.5.2.
(5.2) This occurs because the electrons can relax very rapidly to the bottom
of the conduction band by emission of phonons, and similarly for holes.
The relaxation processes occur on a much faster timescale (ps) than
the radiative recombination (ns). Hence all the carriers have relaxed
before emission occurs, and it makes no dierence where they were initially
injected. See Section 5.3.1.
(5.3) In the 2p 1s transition, the upper 2p level has n = 2, l = 1 and m = 1,
0 or +1, while the lower 1s level has n = 1, l = 0 and m = 0. The Einstein
A coecient is therefore given from eqn B.29 as:
A =
e
2

3
3
0
c
3
1
3

m=1,0,1
[2p, m[r[1s)[
2
,
where the factor of 1/3 accounts for the triple degeneracy of the 2p state.
(It is not necessary to consider the spin degeneracies here because they
cancel out.) We write:
r = x

i +y

j +z

k,
so that:
[r)[
2
= [x)[
2
+[y)[
2
+[z)[
2
.
Now atoms are spherically symmetric, and so that it must be the case
that:
[x)[
2
= [y)[
2
= [z)[
2
.
We therefore only have to evaluate one of these, and we chose z) because
the mathematics is easier. Written explicitly, with z = r cos , we have:
z) =
_

r=0
_

=0
_
2
=0

2p
r cos
1s
r
2
sin drdd.
This has to be evaluated for each of the three possible m values of the 2p
state. However, since z has no dependence on , and the dependence of
the wave functions is determined only by m, the integral is only non-zero
for the m = 0 level of the 2p state. Hence we only have to evaluate one
integral. On inserting the explicit forms of the wave functions, we then
have:
z) =
_

r=0
R

21
rR
10
r
2
dr
_

=0
_
2
=0
Y

1,0
cos Y
0,0
sin dd,
=
1

6a
4
H
_

r=0
r
4
exp(3r/2a
H
) dr

3
4
_

=0
_
2
=0
cos
2
sin dd.
33
This gives:
z) =
24

6
_
2
3
_
5
a
H

3
= 3.94 10
11
m.
Then, with [r)[
2
= 3[z)[
2
, we obtain:
A =
e
2

3
3
0
c
3
1
3
3[z)[
2
=
e
2

3
3
0
c
3
[z)[
2
.
The 2p 1s transition in hydrogen has an energy of (3/4)R
H
= 10.2 eV,
and therefore = 1.55 10
16
rad/s. Hence we obtain A
2p1s
= 6.27
10
8
s
1
. We then see from eqn 5.2 that the radiative lifetime
R
= 1/A =
1.6 ns. This concurs with the experimental value.
(5.4) It follows from eqn 5.4 that:
1

=
1

R
+
1

NR
,
where 1/
NR
is the non-radiative recombination rate. Hence the excited
state lifetime can be shorter than the radiative lifetime.
R
would normally
be independent of T, but the non-radiative recombination rate 1/
NR
gen-
erally increases with T due to increased non-radiative recombination by
phonon emission. Hence decreases with T. The quantum eciency can
calculated from eqn 5.5 to be equal to /
R
. At 300 K we nd
R
= 79 %,
while at 350 K we nd
R
= 56 %.
(5.5) Semiconductors emit light at their band gap energy. We are therefore
looking for a semiconductor with E
g
= 2.3 eV. Inspection of Table C.3
suggest two possibilities: ZnTe or GaP. The latter has an indirect band
gap and would therefore not be very inecient. Hence the most likely
candidate material is ZnTe. Note, however, that we could also make an
emitter at this wavelength by using an alloy, for example: Ga
x
In
1x
N.
(See Fig. 5.11 and Exercise 5.15.) Linear extrapolation between the GaN
and InN band gaps would suggest that a composition with x 0.26 would
emit at 540 nm.
(5.6) (i) Consider an incremental beam slice at a position z within the absorbing
material as shown in Fig. 12. Let A be the area of the slice, dz its thickness,
I the incoming intensity (power per unit area), and I the loss of intensity
due to absorption within the slice. From Beers law (eqn 1.3) we have:
I = Idz .
We assume that each absorbed photon generates an electron-hole pair.
The number of electron-hole pairs generated within the slice per unit time
is therefore equal to AI/h. Hence the carrier generation rate G per unit
volume is given by:
G =
AI/h
Adz
=
AIdz/h
Adz
=
I
h
.
34
(ii) The rate equation for the carrier density N is
dN
dt
= G
N

=
I
h

N

,
where the rst term accounts for carrier generation and the second for
carrier recombination. In steady-state conditions we must have that
dN
dt
= 0 .
Hence:
N =
I
h
.
(iii) We calculate the laser intensity from:
I =
P
A
=
1 mW
(50 m)
2
= 1.3 10
5
Wm
2
.
The sample is antireection coated, and so this is also the intensity inside
the sample. We can then calculate the carrier density using the result
from part (ii) with the values of and given in the exercise. Hence we
nd N = 6.6 10
20
m
3
for h = 2.41 eV. Note that this is the carrier
density at the front of the sample. The intensity will decay exponentially
(cf. eqn 1.4) and the carrier density will follow a similar exponential decay.
Incoming
photonnumber
N(z)
Outgoing
photonnumber
N-dN
A
dz
Incoming
intensity
I(z)
Outgoing
intensity
I - dI
z
Incoming
photonnumber
N(z)
Outgoing
photonnumber
N-dN
A
dz
Incoming
intensity
I(z)
Outgoing
intensity
I - dI
z
Figure 12: An incremental slice of a laser beam at a position z within an
absorbing material. A is the area of the slice and dz its thickness. In Exercise 5.6
we consider a continuous laser beam with an intensity I(z), whereas in Exercise
5.7 we consider a pulse with a photon number of N(z).
(5.7) (i) The argument proceeds along similar lines as for the previous exercise.
Consider again an incremental beam slice of area A and thickness dz, as
shown in Fig. 12. Since we are now dealing with a pulse rather than a
continuous beam, we need to consider the photon number rather than the
intensity. Let N be the incoming photon number, and N the number of
photons absorbed within the slice. From Beers law (eqn 1.3) we have:
N = Ndz .
We assume that each absorbed photon generates an electron-hole pair.
Furthermore, we assume that the pulse is so short that no recombina-
tion takes place while the material is being excited. Hence the number
35
of electron-hole pairs generated is just equal to the number of photons
absorbed. The carrier density generated is then given by:
N =
N
Adz
=
Ndz
Adz
=
N
A
.
Now the number of photons N in the pulse is just equal to E/h, where
E is the pulse energy. Hence we obtain:
N =
E
Ah
.
On inserting the relevant numbers from the exercise, we nd N = 1.9
10
24
m
3
.
(ii) The pulse excites the carrier density calculated in part (i) at time t = 0.
The carriers then recombine and the carrier density decreases according
to
N(t) = N
0
exp(t/) ,
where is the total decay rate including both radiative and non-radiative
recombination. (See eqn 5.4.) With = (1/
R
+ 1/
NR
)
1
= 0.89 ns, we
nd N(t) = N
0
/2 at t = 0.62 ns.
(iii) The quantum eciency is calculated from eqn 5.5 as = 89%. Each
laser pulse contains E/h photons. We are told that the crystal is thick,
and so we can assume that all the laser photons are absorbed in the crystal.
The number of photons re-emitted by luminescence is therefore E/h =
3.5 10
10
photons.
hn
E
g
E
k
k =0
E
e
E
h
hn
E
g
E
k
k =0
E
e
E
h
Figure 13: Denition of energies as required for Exercise 5.8.
(5.8) The emission rate is proportional to the probability that the upper level
is occupied and that the lower level is empty. These probabilities can
be calculated from the FermiDirac functions for the electrons and holes,
with f
e
as the electron occupancy of the upper level and f
h
as the hole
occupancy of the lower level (i.e. the probability that the lower level is
empty.) Hence the emission probability is proportional to f
e
f
h
.
36
In the classical limit, the occupancy factors follow Boltzmann statistics
with:
f
e,h
exp(E
e,h
/k
B
T) ,
where E
e
and E
h
are the electron and hole energies within the conduction
or valence band, respectively. (See Fig. 13.) Hence:
f
e
f
h
exp(E
e
/k
B
T) exp(E
h
/k
B
T) = exp[(E
e
+E
h
)/k
B
T] .
Now it is apparent from Fig. 13 that
h = E
g
+E
e
+E
h
,
and hence that (E
e
+E
h
) = h E
g
. We therefore conclude that:
I(h) f
e
f
h
exp[(h E
g
)/k
B
T] .
(5.9) The number of electrons in the conduction band is given by eqn 5.6. In the
classical limit, we have (E E
F
) k
B
T, so that the electron occupancy
factor is given by:
f
e
(E) =
1
exp[(E E
F
)/k
B
T] + 1
exp[(E
F
E)/k
B
T] .
On using the density of states given in eqn 5.7, we then obtain:
N
e
=
1
2
2
_
2m

2
_
3/2
exp
_
E
F
k
B
T
__

E
g
(E E
g
)
1/2
exp
_
E
k
B
T
_
dE .
On introducing the variable x = (E E
g
)/k
B
T we then obtain:
N
e
=
1
2
2
_
2m

e
k
B
T

2
_
3/2
exp
_
(E
F
E
g
)
k
B
T
__

0
x
1/2
exp(x) dx.
The nal result is obtained by setting (E
F
E
g
) E
c
F
, where E
c
F
is the
electron Fermi energy measured relative to the bottom of the conduction
band.
For the case of GaAs at 300 K, we can insert the eective mass and
temperature, and use the denite integral given in the Exercise, to obtain:
N
e
= exp
_
E
c
F
k
B
T
_
(4.4 10
23
) m
3
.
In part (a) we then nd E
c
F
= 0.216 eV 8.4k
B
T, whereas in part (b)
we nd E
c
F
= +0.021 eV +0.83k
B
T. The approximations are therefore
valid for part (a) because we have (E E
F
) k
B
T for all states in the
conduction band, but not for part (b), where the Fermi level comes out
above the conduction band minimum.
The aim of this exercise is to get the students to think about the conditions
under which the use Boltzmann statistics is valid.
37
(5.10) At T = 0 we have f(E) = 1 for E < E
F
and f(E) = 0 for E > E
F
.
We can therefore cut o the integral at the relevant Fermi energy, and
replace the Fermi function by unity up to this energy. With the energies
measured relative to the band edge, we then nd:
N
e
=
1
2
2
_
2m

2
_
3/2
_
E
c
F
0
E
1/2
dE ,
for the electrons and similarly for the holes.
On evaluating the integral we nd:
N
e
=
1
2
2
_
2m

2
_
3/2

2
3
(E
c
F
)
3/2
,
and similarly for the holes. Equation 5.13 then follows by simple re-
arrangement.
(5.11) We use eqn 5.13 to evaluate the Fermi energies. In part (a) we nd
E
c
F
= 0.36 meV for the electrons, and E
v
F
= 0.073 meV for the holes. In
part (b) we nd E
c
F
= 36 meV for the electrons, and E
v
F
= 7.3 meV for the
holes.
For the conditions of degeneracy to apply, we need E
F
k
B
T. In part
(a), the electrons will be degenerate for T 4.2 K, while the holes will
be degenerate for T 0.9 K. In part (b), the electrons will be degenerate
for T 420 K, while the holes will be degenerate for T 85 K. This
shows that it is much easier to obtain degenerate statistics for the lighter
electrons than for the holes.
(5.12) The Fermi wave vector k
F
is, in general, related to the Fermi energy E
F
by:
E
F
=

2
k
2
F
2m
.
With E
F
given by eqn 5.13, we then nd:
k
F
= (3
2
N)
1/3
,
where N = N
e
or N
h
as appropriate. In a photoluminescence experiment,
we excite equal numbers of electrons and holes, so that N
e
= N
h
. Hence
the Fermi wave vectors of the electrons and holes are identical. The fact
that the mass does not appear in the formula for the Fermi wave vector
is a consequence of the fact that the density of states in k-space does not
depend on the mass.
(5.13) (i) The solid angle subtended by an object of area dA at a distance r is
given by (See Fig 14(a)):
d =
dA
4r
2
.
The lens will be positioned at a distance equal to its focal length from the
sample, and so we set r = 100 mm in this case. We then nd:
d =
(25 mm/2)
2
4(100 mm)
2
= 0.049 .
38
dA
r
dW
(a)
q
inside
q
outside
medium
n
air
n =1
S
(b)
dA
r
dW
(a)
q
inside
q
outside
medium
n
air
n =1
S
(b)
Figure 14: (a) Denition of solid angle d for an object of area dA at a distance
of r, as required for Exercise 5.13. (b) Emission of a ray at angle
inside
from a
source S embedded within a medium of refractive index n.
(n.b. The answer given in some of the printed versions of the book is
incorrect.)
(ii) Consider a ray emitted by a source embedded within a medium of
refractive index n as illustrated in Fig. 14. The ray is refracted at the
surface according to Snells law, with:
sin
outside
sin
inside
= n.
The light is being collected by a lens of radius 12.5 mm positioned 100 mm
from the surface, and therefore only those rays that satisfy
outside

12.5/100 will be collected (assuming small angles). By Snells law we
deduce that only those rays within a cone with
inside
0.125/n will be
collected.
The source emits uniformly over all 4 steradians within the medium.
The fraction F of the photons emitted that can be collected by the lens is
worked out by considering a circle of radius r
inside
placed at a distance
of r from the source:
F =
dA
4r
2
=
r
2
(
inside
)
2
4r
2
=
(
inside
)
2
4

(
outside
)
2
4n
2
.
Finally, we have to consider that some of the photons will be reected
at the surface. The reectivity is calculated from eqn 1.26 to be 21% for
n = 2.7. Hence the nal fraction collected is 0.79F = 4.2 10
4
.
(iii) The semiconductor absorbs photons of energy 2.41 eV and emits lu-
minescent photons at the band gap energy of 1.61 eV with a probability
equal to
R
. Hence the power emitted is equal (1.61/2.41)
R
times the
power absorbed. The incoming laser will be partially reected at the
surface, and so the maximum power that can be absorbed is equal to
(1 R) P
incident
= 0.79 1 mW. Hence the maximum luminescent
power is equal to:
P
lum
=
1.61
2.41

R
0.79 1 mW = 0.53
R
mW.
39
(iv) We obtain the luminescent power collected by the lens by multiplying
the total luminescent power by the collected fraction:
P
collected
= 0.53
R
4.2 10
4
mW = 0.22
R
W.
(5.14) (i) The electron Fermi energy is calculated from eqn 5.13 to be 0.14 eV.
(ii) The hole Fermi energy is calculated from:
N
h
=
_
E
v
F
0
[g
hh
(E) +g
lh
(E)] dE
=
_
E
v
F
0
1
2
2
_
2

2
_
3/2
(m
3/2
hh
+m
3/2
lh
) E
1/2
dE
=
1
3
2
_
2

2
_
3/2
(m
3/2
hh
+m
3/2
lh
) (E
v
F
)
3/2
.
Hence for N
h
= 2 10
24
m
3
we obtain E
v
F
= 0.012 eV.
(iii) The carriers will be degenerate if E
F
> k
B
T. At 180 K we have
k
B
T = 0.018 eV, and so the electrons are degenerate, but not the holes.
(iv) When both the electrons and holes are degenerate, we expect emission
from the band edge up to the Fermi energies as illustrated in Fig. 5.7. We
then expect emission from the band gap to E
g
+ E
c
F
+ E
v
F
. At T = 0
we would have a sharp cut-o at this energy, but at nite T the edge is
broadened over k
B
T. The luminescence would then be expected to fall
to 50 % of its peak value at E
g
+ E
c
F
+ E
v
F
. On inserting the calculated
Fermi energies, and the value of the band gap, we expect the 50% point
at around 0.95 eV. This can be compared to the experimental value of
0.94 eV. The agreement between the experiment and model is thus
good.
(v) The luminescence spectrum at 250 ps given in Fig. 5.8 is consistent
with a value of the electron Fermi energy of 0.035 eV. This implies from
eqn 5.13 that the carrier density is N
e
3 10
23
m
3
. The electrons are
still degenerate at this density for T = 55 K. The lifetime is estimated
from:
N(t) = N
0
exp(t/) ,
which implies N(250)/N(24) = exp(226/), where is the lifetime in
ps. On inserting the two values, we nd 0.13 ns.
(5.15) We follow Example 5.1(i). We require a band gap equivalent to 500 nm,
namely 2.48 eV. Linear extrapolation between the band gaps of GaN and
InN implies that an alloy with x = 0.39 would have the required band
gap.
(5.16) (i) The reectivity is calculated from eqn 1.26 to be 31% for n = 3.5.
(ii) The cavity mode frequency is given by eqn 5.15, and implies a mode
spacing of = c/2nl. With n = 3.5 and l = 1 mm, we nd =
4.3 10
10
Hz.
(iii) The threshold gain can be calculated from eqn 5.19. We are told to
ignore background absorption and scattering, and so we set
b
= 0. The
40
coated facet has a reectivity of 95%, while the other facet will just have
the natural reectivity of 31%. Hence we nd:

th
=
1
2 10
3
m
ln(0.95 0.31) = 610 m
1
.
(5.17) (i) The maximum possible power would be obtained if one photon is
emitted for each electron that ows through the device. The power is
then equal to the electron ow rate multiplied by the photon energy:
P
max
= h
_
100 mA
e
_
= 150 mW.
(ii) The electrical power input is equal to IV = 0.1 1.9 = 0.19 W =
190 mW. With a power output of 50 mW, the power conversion eciency
is therefore equal to (50/190) = 26 %.
(iii) The power output will vary as shown in Fig. 5.15(a). The slope
eciency is calculated from eqn 5.21 as (50/(100 35) = 0.77 mW/mA =
0.77 W/A. The dierential quantum eciency is calculated from eqn 5.21
to be 51 %.
41
Chapter 6
Semiconductor quantum
wells
(6.1) Substitute into eqn 6.3 with x = 10
6
m and m = 0.1m
0
to obtain
T 0.01 K.
(6.2) The energy dierence between the n = 1 and n = 2 levels is worked out
from eqn 6.13 to be 3
2

2
/2m

d
2
. On setting this energy dierence to be
equal to k
B
T/2, we then derive the required result, namely
d =

3
2

2
m

k
B
T
.
On substituting into this formula with T = 300 K, we nd d = 9.3 nm for
m

= m
0
, and d = 30 nm for m

= 0.1m
0
.
On comparing with eqn 6.4, it is apparent that d =

3
2
x. This shows
that the two criteria used to determine when quantum size eects are
important give the same answer, apart from a numerical factor of 5.
The diering numerical factor is to be expected, given the approximate
nature of the criteria.
k
x
k
y
(2p/L)
2
k
dk
k
x
k
y
(2p/L)
2
k
dk
Figure 15: Grid of allowed k states in a two-dimensional material of area L
2
,
as required for the solution of Exercise 6.3.
(6.3) The x and y components of the k vector must each satisfy the criterion
exp(ikL) = 1, which implies k = integer 2/L. The possible values of
the k vector therefore form a regular grid in k-space as shown in Fig. 15,
42
with a grid-spacing of 2/L. The area per k-state is (2/L)
2
, and the
density of states for an area L
2
is therefore L
2
/(2)
2
. This implies that
the density of states in k-space for a unit area of crystal is 1/(2)
2
.
It is also apparent from Fig. 15 that the area of k-space enclosed by the
increment k k +dk is 2kdk. If we call the number of k-states enclosed
by this area g
L
(k)dk, we then nd:
g
L
(k)dk =
2kdk
(2/L)
2
=
L
2
2
kdk .
Hence for a unit area of crystal we have:
g(k)dk =
k
2
dk .
The density of states in energy space can be found from eqns 3.1314. (The
factor or two accounts for the fact that spin 1/2 particles have two spin
states for each k-state.) With E(k) =
2
k
2
/2m, we have dE/dk =
2
k/m,
and hence
g(E) =
2g(k)
dE/dk
=
2k/2

2
k/m
=
m

2
,
as required.
(6.4) The depth of the potential well enters eqn 6.26 via , which is dened in
eqn 6.27. The function on the right hand side of eqn 6.26 decreases from
+ at x = 0 to zero at x =

. (See, for example, Fig. 6.4, for the case


with = 13.2.) The function on the RHS of eqn 6.26 will therefore always
cross the tan x function between 0 and /2, no matter how small is.
0.0 0.4 0.8 1.2
0
2
4
6
8
10
y
x
y =tan(x)
y =0.82(33.5-x
2
)
1/2
/x
x =1.30
0.0 0.4 0.8 1.2
0
2
4
6
8
10
y
x
y =tan(x)
y =0.82(33.5-x
2
)
1/2
/x
x =1.30
Figure 16: Graphical solution required for Exercise 6.5.
(6.5) We follow the method of Example 6.1. We have
_
m

w
m

b
_
1/2
=
_
0.34
0.5
_
1/2
= 0.82 ,
and
=
0.34m
0
(10
8
)
2
0.15 eV
2
2
= 33.5 .
43
Hence we must solve:
tan x = 0.82
_
33.5 x
2
x
_
.
The two functions are plotted in Fig. 16, which shows that the solution is
x = 1.30. The energy is then found from
E =
2
2
x
2
m

w
d
2
= 7.6 meV,
where d = 10 nm. The equivalent energy for an innite well is calculated
from eqn 6.13 to be 11 meV.
(6.6) (i) The wave functions of an innite potential well form a complete or-
thonormal basis, with
_
+

n
dz =
n,n
,
where
n,n
is the Kronecker delta function:

n,n
= 1 if n = n

= 0 if n ,= n

.
It is apparent from eqn 6.11 that the wave functions depend only on n and
are therefore identical for electrons and holes with the same n. Hence the
orthonormality condition applies irrespective of whether
n
and
n
are
electron or hole wave functions, or a mixture. Hence:
M
n,n
= 1 if n = n

= 0 if n ,= n

.
This result can also be derived by explicit (and somewhat tedious) substi-
tution of the wave functions from eqn 6.13 into the formula for M
n,n
.
(ii) This result follows from parity arguments. The wave functions of
a nite well have well-dened parities as a consequence of the inversion
symmetry about the centre of the well. Wave functions with odd n have
even parity, while those with even n have odd parity. Hence the product

en

hn
has even parity if n n

is even and has odd parity for odd n n

.
The integral of an odd function from + is zero, and so the
overlap integral is zero if n is equal to an odd number.
n Electron Heavy hole Light hole
1 225 30 188
2 898 120 750
Table 1: Connement energies in meV calculated for an innite quantum well
of width 5 nm, as required for Exercise 6.7.
(6.7) The connement energies of the electrons, heavy holes and light holes
calculated from eqn 6.13 are given in Table 1. With an innite well, we
only need consider n = 0 transitions. The threshold photon energies for
these transitions are given by eqn 6.39. Two transitions fall in the range
1.4 2.0 eV:
44
Heavy hole 1 electron 1 at 1.679 eV,
Light hole 1 electron 1 at 1.837 eV.
For each transition we expect a step at the threshold energy as in Fig. 6.8.
In 2-D materials the joint density of states is proportional to the electron-
hole reduced mass (see eqn 6.41), and hence the relative height of the
heavy-hole and light-hole transition steps is in proportion to their reduced
masses, that is 0.059 : 0.036. Hence the nal spectrum would appear as
in Fig. 17.
1.4 1.6 1.8 2.0
0.0
0.1
A
b
s
o
r
p
t
i
o
n
(
a
r
b
.
u
n
i
t
s
)
Photonenergy(eV)
hh1 e1
lh1 e1
1.679 eV
1.837 eV
1.4 1.6 1.8 2.0
0.0
0.1
A
b
s
o
r
p
t
i
o
n
(
a
r
b
.
u
n
i
t
s
)
Photonenergy(eV)
hh1 e1
lh1 e1
1.679 eV
1.837 eV
Figure 17: Absorption spectrum for Exercise 6.7
(6.8) (i) In nite wells the connement energies are reduced compared to those
of an innite well of the same width. Hence the transition energies would
be lower. Furthermore, transitions that are forbidden in innite wells
become weakly allowed, such as the hh3 e1 transition. This transition
would fall within the observed energy range.
(ii) Peaks would appear below the steps in the absorption spectrum due
to exciton absorption. The dierence in energy between the peak and the
continuum absorption is the exciton binding energy.
(6.9) (i) To prove normalization, we must show that
_ _

dA = 1. For the
given wave function we have:
_

r=0
_
2
=0

rdrd =
_
2

2
__

r=0
_
2
=0
exp(2r/) rdrd ,
=
_
2

2
_
2
_

r=0
exp(2r/) rdr ,
= 1 ,
as required.
(ii) We rst compute the eect of

H on , using the fact that / = 0:

H =

2
2r
d
dr
_
r
d
dr
_

e
2
4
0

r
r
,
=

2
2
2
+
_

2
2
2

e
2
4
0

r
_

r
.
45
We can then calculate E)
var
from:
E)
var
=

2
2
2
_

r=0
_
2
=0

rdrd +
_

2
2
2

e
2
4
0

r
__

r=0
_
2
=0

drd ,
=

2
2
2
1 +
_

2
2
2

e
2
4
0

r
_

,
= +

2
2
2

e
2
2
0

.
(iii) On dierentiating E)
var
with respect to , we nd that E)
var
achieves
its minimum value of E
min
= e
4
/8(
0

r
)
2
for = 2
2

r
/e
2
. The
minimum energy is four times larger than the bulk exciton binding energy
found in Exercise 4.4.
(iv) In part (iii) we found
min
= 2
2

r
/e
2
. This can be written in
terms of the bulk exciton radius a
X
dened in eqn 4.2 as:

min
= a
X
/2 .
Hence the radius of the exciton in 2-D is half the radius of the bulk.
(6.10) At d = we have bulk GaAs, while at d = 0 we have bulk AlGaAs.
For intermediate values of d, we have a GaAs quantum well exciton with
an enhanced binding energy. In an ideal 2-D system we would expect
four times the binding energy of bulk GaAs (i.e. 16 meV), but in realistic
systems, the enhancement might be smaller due to the imperfect quantum
connement of the nite-height AlGaAs barriers. Thus as d is reduced
from , the binding energy increases from 4 meV, going through a peak,
and then dropping to 6 meV as d 0. The height of the peak might
typically be around 10 meV.
1.58 1.60 1.62 1.64
0.0
1.0
Energy(eV)
P
L
E
i
n
t
e
n
s
i
t
y
hh1 e1
exciton
lh1 e1
exciton
hh1 e1bandedge
lh1 e1
bandedge
1.58 1.60 1.62 1.64
0.0
1.0
Energy(eV)
P
L
E
i
n
t
e
n
s
i
t
y
hh1 e1
exciton
lh1 e1
exciton
hh1 e1bandedge
lh1 e1
bandedge
Figure 18: Interpretation of the data in Fig. 6.17 as required for Exercise 6.11.
(6.11) (i) See Section 5.3.4.
(ii) The interpretation of the principal features in the data is shown in
Fig. 18. For both heavy and light hole transitions, we expect to observe a
peak due to excitonic absorption followed by a step at the band edge. The
two strong peaks observed in the data correspond to the excitons for the
46
hh1 e1 and lh1 e1 transitions, while the at absorption bands above
both excitons correspond to the interband transitions. These interband
absorption bands are at because the density of states is independent of
the energy in 2-D systems (see eqn 6.41.)
(iii) The hh1 e1 interband continuum starts at 1.592 eV. In the innite
well model, the transition energy is given by eqn 6.42 with n = 1, which,
on using E
g
= 1.519 eV, m

e
= 0.067m
0
and m

hh
= 0.5m
0
, implies d =
9.3 nm. The real well width would be smaller, because the innite well
model overestimates the connement energy.
(iv) The binding energies can be read from Fig. 6.17 as the energy gap
between the exciton peak and the appropriate band edge. With the tran-
sitions identied as in Fig. 18, we nd binding energies of 11 meV for the
heavy holes and 12 meV for light holes. For a perfect 2-D system we would
expect 4R
X
for GaAs, i.e. 16.8 meV. The experimental values are lower
because a real quantum well is not a perfect 2-D system.
(6.12) (i) The interaction between an electric dipole and an external electric
eld is of the form:
H

= p E .
With the eld aligned along the z axis, this reduces to H

= p
z
c
z
. If the
position vector of the electron relative to the origin is r, then the electron
dipole moment is er, with z component p
z
= ez. Hence H

= +ezc
z
.
Note that the choice of origin does not matter in the nal result, because
we only calculate the shift of the electron relative to its original position.
(ii) With H

= +ezc
z
, we have:
E
(1)
= ec
z
_
+

zdz .
Since z is an odd function, and

[[
2
is an even one, the integral is
zero.
(iii) With H

= +ezc
z
, the second-order energy shift is given by:
E
(2)
= e
2
c
2
z

n=2
[1[z[n)[
2
E
1
E
n
.
Since the wave functions have parity (1)
n+1
, and z is an odd parity
operator, all the terms with odd n are zero. Hence we have:
E
(2)
= e
2
c
2
z
[1[z[2)[
2
E
1
E
2
+e
2
c
2
z
[1[z[4)[
2
E
1
E
4
+ .
Now the terms with n 4 are much smaller than the term with n = 2.
(This can be veried by working through the integrals, but it is fairly
obvious given the larger denominator.) Hence we only need to consider
the rst term:
E
(2)
= e
2
c
2
z
[1[z[2)[
2
E
1
E
2
.
47
On substituting the energies for an innite well from eqn 6.13, this be-
comes:
E
(2)
=
2e
2
c
2
z
m

d
2
3
2

2
[1[z[2)[
2
.
Now, on redening the origin so that the quantum well runs from z = 0
to z = +d, we have:
1[z[2) =
_
+

1
z
2
dz ,
=
2
d
_
d
0
sin(z/d) z sin(2z/d) dz ,
=
16d
9
2
.
Hence we nd
E
(2)
=
2e
2
c
2
z
m

d
2
3
2

16d
9
2

2
= 24
_
2
3
_
6
e
2
c
2
z
m

d
4

2
,
as required.
(6.13) (i) From the result in the previous Exercise, we would expect E
c
2
z
. We work out the eld strength from eqn 6.44, which implies c
z
=
11.5 MV/m at 10 V and c
z
= 6.5 MV/m at 5 V. For small shifts we have
E, and so we nd:
(5 V) =
_
6.5
11.5
_
2
(10 V) = 0.32 10.5 nm = 3.4 nm.
(ii) The wavelength red shift of 10.5 nm corresponds to an energy shift of
18 meV. This energy shift is related to the net electron-hole displacement
z) by:
E = p
z
c
z
= ez)c
z
.
With c
z
= 11.5 MV/m, we thus obtain z) = 1.6 nm.
Sample A Sample B
c
z
(MV/m) E
calc
E
exp
E
calc
E
exp
3 1.5 1 15 6
6 5.9 5 62 27
9 13 13 139 54
Table 2: Comparison of the calculated and measured Stark shifts (in meV) for
the two samples discussed in Exercise 6.13.
(6.14) The experimental data is taken from Polland et al., Phys. Rev. Lett. 55,
2610 (1985). We analyse it by using the energy shift calculated by second-
order perturbation theory in Exercise 6.12(iii). Since we are considering
an electron-hole transition, we must add together the Stark shifts of the
electrons and holes, giving:
E
(2)
= 24
_
2
3
_
6
e
2
c
2
z
d
4

2
(m

e
+m

hh
) .
48
The calculated shifts using m

e
= 0.067m
0
and m

hh
= 0.5m
0
are compared
to the experimental ones in Table 2. It is apparent that the model works
well for sample A, but not for sample B. The model breaks down when the
size of the Stark shift becomes comparable to the energy splitting of the
unperturbed hh1 and hh2 levels. This is essentially the same criterion as
for the transition from the quadratic to the linear Stark eect in atomic
physics. In sample B, we are in this regime at all the elds quoted.
(6.15) At c
z
= 0 the quantum well is symmetric about the centre of the well.
The electron and hole states therefore have a denite parity with respect
to inversion about z = 0. The parity is (1)
(n+1)
, and the electronhole
overlap given by eqn 6.36 is zero if n is odd: see Exercise 6.6(ii). At
nite c
z
, the inversion symmetry is broken, and the states no longer have
a denite parity. Therefore, the selection rule based on parity no longer
holds.
(6.16) With the connement energy E d
2
, we have dE/dd 2/d
3
, and
hence we expect:
E
E
= 2
d
d
.
A 5% change in d is thus expected to give E/E = 10%. With
E = 0.1 eV, we then expect a full-width broadening of 0.02 eV. This is
comparable to the linewidth observed in the 10 K data shown in Fig. 6.13.
A 5% variation in d corresponds more or less to a uctuation of one
atomic layer. Such monolayer uctuations are unavoidable in the crystal
growth. The linewidth at room temperature is further broadened by the
thermal spread of the carriers in the bands.
(6.17) We assume innite barriers and use eqn 6.42 to calculate the transition
energy with n = 1. With = 0.80 eV, E
g
= 0.75 eV, and = 0.038m
0
,
we nd d = 14 nm. In reality, the quantum well would have to be narrower
to compensate for the imperfect connement of the barriers.
(6.18) (i) z is an odd function with respect to inversion about z = 0. The
integral from to + will therefore be zero unless

n
is also an
odd function, which requires that the wave functions must have dierent
parities. Since the wave functions have parities of (1)
n+1
, the condition
is satised if n is an even number and n

odd, and vice versa. Hence n


must be equal to an odd number.
(ii) The strength of the intersubband transitions is proportional to the
square of the matrix elements. These matrix elements can be evaluated
by substituting the wave functions from eqn 6.11. On redening the origin
so that the quantum well runs from z = 0 to z = +d, we nd:
1[z[2) =
2
d
_
d
0
sin(z/d) z sin(2z/d) dz =
16
9
2
d ,
and
1[z[4) =
2
d
_
d
0
sin(z/d) z sin(4z/d) dz =
4
45
2
d .
49
Hence the 1 4 transition is weaker than the 1 2 transition by a factor
[(4/45)/(16/9)]
2
= [1/20]
2
= 2.5 10
3
.
It is apparent from Fig. 6.14 that the wavelength of the 1 2 transition
is given by
hc

= E
2
E
1
=
3
2

2
2m

e
d
2
,
where we used the innite well energies of eqn 6.13 in the second equality.
On inserting m

e
= 0.067m
0
and d = 20 nm, we nd hc/ = 0.042 eV, and
hence = 29 m.
q
q
semiconductor,refractiveindex n
air
n =1
E
E
z
quantumwell
q
q
semiconductor,refractiveindex n
air
n =1
E
E
z
quantumwell
Figure 19: Refraction of light entering a semiconductor containing a quantum
well, as discussed in Exercise 6.19.
(6.19) Consider a ray incident at angle to the normal as shown in Fig. 19.
The ray will be refracted according to Snells law, with
sin
sin

= n,
where

is the angle inside the crystal. For intersubband transitions, we


are interested in the z component of the electric eld of the light at the
quantum well, namely:
c
z
= c

sin

= c

sin /n.
If I
0
is the incident intensity, and there are no intensity losses at the
surface, then the intensity in the z component at the quantum well is
given by
I
z
= I
0
(sin /n)
2
,
since the intensity is proportional to c
2
. Hence the fraction of the power
of the beam in the z polarization at the quantum well is (sin /n)
2
. This
fraction has a maximum value of 1/n
2
for = 90

. Therefore even if we
completely absorb all the z polarized light by intersubband transitions,
and we use glancing incidence, we can only remove a fraction of 1/n
2
of
the power in the incident beam. This fractional absorption is equal to 9%
if n = 3.3.
(6.20) For a cubic dot, the energies of the quantized levels are given by eqn 6.47
with d
x
= d
y
= d
z
= d, implying:
E(n
x
, n
y
, n
z
) =

2

2
2m

d
2
(n
2
x
+ n
2
y
+ n
2
z
) ,
50
n
x
n
y
n
x
(n
2
x
+ n
2
y
+ n
2
z
)
1 1 1 3
2 1 1 6
2 2 1 9
3 1 1 11
2 2 2 12
3 2 1 14
3 2 2 17
Table 3: Quantum numbers of the energy levels for a cubic quantum dots in
order of increasing energy, as discussed in Exercise 6.20.
where n
x
, n
y
and n
z
are integers with a minimum value of 1. As demon-
strated by Table 3, the quantized levels occur at energies of 3, 6, 9, 11,
12, 14, 17,. . . in units of h
2
/8m

d
2
.
51
Chapter 7
Free electrons
(7.1) The method for determining the electron Fermi energy was considered in
Exercise 5.10, and the formula for E
F
is given in eqn 5.13:
E
F
=

2
2m
(3
2
N)
2/3
,
which implies:
N =
1
3
2
_
2mE
F

2
_
3/2
.
On substituting this form of N into the formula for the plasma frequency
in eqn 7.6, we then obtain
E
3
F
=
9
2
0

2
8m
_

p
e
_
4
.
(7.2) We expect 100% reectivity below the plasma frequency and transmission
at higher frequencies. (See Fig. 7.1.) Hence we can set
p
/2 = 3 MHz
in this example. On substituting into eqn 7.6 and solving for N, we nd
N 10
11
m
3
. (n.b. The electron density calculated here is just a typical
one. The value of N varies somewhat due to atmospheric conditions.)
(7.3) The formula for the skin depth is given in eqn 7.20. On inserting the
value of for salt water we then nd 0.5 m for /2 = 200 kHz. This
shows that electromagnetic waves of this frequency penetrate less than 1 m
from the surface of the sea. To obtain a skin depth of 30 m as required
for communication with a submarine submerged at this depth, we require
/2 = 70 Hz. Even lower frequencies are required for deeper depths. The
data transmission rate is very low at these small carrier frequencies.
(7.4) From Table 7.1 we read N = 0.91 10
28
m
3
for cesium, while the trans-
mission edge at 440 nm implies
p
= 440 nm, and hence:

p
= 2c/
p
= 4.3 10
15
rad/s .
On substituting into eqn 7.6 and solving for m, we nd m

e
= 1.4
10
30
kg 1.6m
0
.
(7.5) This Exercise closely follows Example 7.3. We read
p
= 1.3610
16
rad/s
from Table 7.1, and work out = 4.0 10
14
s from eqn 7.14 using the
value of N from Table 7.1 and the value of
0
given in the exercise. At
500 nm we have = 3.77 10
15
rad/s, and the relative permittivity at
this wavelength is then given by eqns 7.1617 as:

r
= 12.0 + 0.086i .
We nally use eqns 1.2223 to calculate n = 0.012 +3.5i, and substitute
n = 0.012 and = 3.5 into eqn 1.26 to obtain R = 99.6 %.
52
(7.6) We rst use eqn 7.14 to nd = 1.2 10
13
s, taking N = 5.9 10
28
m
3
from Table 7.1, and then proceed as in Example 7.3 to nd the extinction
coecient . With
p
= 1.37 10
16
rad/s (cf. Table 7.1), and =
2c/ = 1.88 10
15
rad/s, we nd from eqns 7.1617 that
r
= 52.1 +
0.235i, and hence = 7.22 (see eqn 1.23). This value of implies, through
eqn 1.16, an absorption coecient = 9.1 10
7
m
1
. The transmission
is given by Beers law (see eqn 1.4) as exp(z) = 0.16 for z = 20 nm.
(7.7) Based on the plasma frequency of gold given in Table 7.1, we expect high
reectivity above 140 nm. The low reectivity up to 600 nm is caused by
interband transitions as illustrated schematically in Fig 7.4. The energy
gap between the d-bands and the Fermi energy can be read from the
data as the energy equivalent of 520 nm, i.e.: 2.4 eV. It is apparent
from the reectivity spectrum that gold reects red, orange and yellow
light stronger than green and blue. This accounts for its characteristic
yellowish colour.
(7.8) It is apparent from eqn 1.26 that R = 0 when n = 1, and hence
r
= 1.
The relative permittivity of a doped semiconductor is given by eqn 7.22,
and light damping implies that we take = 0. Hence we shall have zero
reectivity when: (see eqn 7.24 for
p
):
1 =
opt

Ne
2
m

2
=
opt
_
1

2
p

2
_
.
Equation 7.25 is then derived by solving this formula for
2
.
N (10
24
m
3
) (R = 0) (m) (R = 0) (10
13
rad/s) m

e
/m
0
0.35 33 5.7 0.020
0.62 27 7.0 0.028
1.2 22 8.6 0.036
2.8 16 12 0.044
4.0 14 13 0.048
Table 4: Eective masses of n-type InSb calculated from the data in Fig 7.7 as
required for Exercise 7.9.
(7.9) The easiest way to determine the eective mass from the spectra is to read
the wavelength at which R = 0 from the data. We then use eqns 7.2425
to write:
m

=
Ne
2

0
(
opt
1)
2
,
where is the angular frequency corresponding to R = 0. The values
of the eective masses found in this way with
opt
= 15.6 are given in
Table 4.
It is apparent from Table 4 that m

e
increases strongly with N, rising from
0.020m
0
at 3.5 10
23
m
3
to 0.048m
0
at 4 10
24
m
3
. This increase in
the eective mass with carrier density is caused by non-parabolicity in
53
the conduction band of InSb. The eective mass approximation assumes
that the bands are parabolic in shape as in Fig 3.5. However, a glance at
the band structure of a real III-V semiconductor (eg GaAs, see Fig 3.4)
indicates that this approximation only holds near k = 0. To get a good
t to the energy bands for larger values of k (but still below the maxima),
we have to re-write eqn 3.17 as:
E
c
(k) = E
g
+

2
k
2
2m

(k)
,
where m

(k) = m

e
for small k, but then increases at larger values of
k. As we dope the semiconductor with more electrons, the Fermi energy
increases (see eqn 5.13) and we are probing states of the conduction band
with larger values of E and k. It is therefore to be expected that the
eective mass should increase with the doping density.
(7.10) The free carrier absorption coecient in the limit 1 is given by
eqn 7.28. At 10 m we have = 1.9 10
14
rad/s, and on inserting the
relevant values into eqn 7.28 we nd = 1.0 ps. This implies 200,
so that the approximations used to derive eqn 7.28 are justied.
(7.11) The laser beam generates free carriers in the conduction and valence
bands which can then induce free carrier absorption. The carrier density
generated by a continuous laser beam is given by (see Exercise 5.6):
N =
I

,
where I is the intensity, is the absorption coecient, is the carrier
lifetime, and is the photon energy. On inserting the appropriate values
as given in the Exercise, we nd N = 1.9 10
22
m
3
. This is the density
of electrons in the conduction band and holes in the valence band, both of
which cause free carrier absorption. The free carrier absorption coecient
at 10.6 m can be calculated separately for the electrons and holes using
eqn 7.28. With = 1.8 10
14
rad/s, we obtain = 130 m
1
for the
electrons and = 70 m
1
for the holes. Hence the total free carrier
absorption coecient is 200 m
1
.
Here are a few points to note about this exercise:
Students should be careful not to confuse the two dierent usages of
. In the calculation of the carrier density, represents the carrier
lifetime, whereas in eqn 7.28 it represents the momentum scattering
time.
Intervalence band absorption has been neglected here. However, with
only 10
22
m
3
holes, it is probable that intervalence band absorp-
tion will be insignicant. (See the next exercise.)
This exercise is actually an example of a nonlinear optical eect: the
light beam at 633 nm induces changes in the optical properties. The
mechanism is that the beam creates carriers which then alter the
optical properties. These types of nonlinear eects are called free
carrier nonlinearities for obvious reasons. See Chapter 11 for more
information on nonlinear optics.
54
k
E
F
D
E

min

max

max

min
max
,
min
k
F
lh
k
F
hh
k
E
F
D
E

min

max

max

min
max
,
min
k
F
lh
k
F
lh
k
F
hh
k
F
hh
Figure 20: Intervalence band transitions as required for the solution of Exercise
7.12. The Fermi energy is not drawn to scale: with the parameters given in the
Exercise, the Fermi energy is just above the split-o band. The labels (1), (2)
and (3) refer to the lhhh, solh, and sohh transitions as in Fig. 7.9.
(7.12) (i) The holes are distributed between the heavy and light hole bands,
and so we can nd the Fermi energy from (see eqn 3.16):
N =
_
EF
0
(g
hh
+g
lh
) dE
=
2
3/2
2
2

3
(m
3/2
hh
+m
3/2
lh
)
_
E
F
0
E
1/2
dE .
With m

hh
= 0.5m
0
and m

lh
= 0.08m
0
we then nd E
F
= 0.032 eV for
N = 1 10
25
m
3
. Note that this is smaller than the spin-orbit energy
and so our neglect of the occupancy of the split-o band is justied. The
wave vector at the Fermi energy is worked out from:
E
F
=

2
k
2
F
2m
.
This give k
hh
F
= 6.5 10
8
m
1
and k
lh
F
= 2.6 10
8
m
1
for the heavy and
light holes respectively. (See Fig. 20.)
(ii) The energies of the three types of intervalence band transitions indi-
cated in Fig. 7.9 can be calculated by using eqns 3.1820:
(1) lh hh : = [E
lh
(k) E
hh
(k)[ =

2
2
_
1
m

lh

1
m

hh
_
k
2
,
(2) so lh : = [E
so
(k) E
lh
(k)[ = +

2
2
_
1
m

so

1
m

lh
_
k
2
,
(3) so hh : = [E
so
(k) E
hh
(k)[ = +

2
2
_
1
m

so

1
m

hh
_
k
2
,
55
where k is the wave vector at which the transition occurs. Transitions can
only take place from occupied states to empty ones. The upper and lower
limits of the transition energies are therefore set by the Fermi wave vectors
as indicated in Fig. 20. The energy limits calculated using the Fermi wave
vectors from part (i) are therefore as follows:
1. lhhh transitions: The lower and upper limits are set by k
lh
F
and
k
hh
F
respectively. On inserting the into (1) above, we nd a transition
range of 0.03 0.17 eV.
2. solh transitions: Since the light hole eective mass is smaller than
the split-o mass, the lower and upper limits correspond to the tran-
sitions at k
lh
F
and k = 0 respectively. On inserting into (2) above we
nd a range 0.32 0.34 eV.
3. sohh transitions: The lower limit is set by the transition at k = 0,
and the upper limit by the one at k
hh
F
, giving a range of 0.34
0.42 eV.
(7.13) (i) The energies of the 2p
0
, 3p
0
and 4p
0
transitions can be read from
Fig. 7.11 as 34.0, 40.1, and 42.4 meV respectively. These energies t well
to the formula:
h = R

_
1
1
n
2
_
,
with R

= 45.2 meV. This is consistent with eqn 7.30 if we set m

e
=
0.85m
0
for the donor levels.
(ii) The energies of the 2p

, 3p

, 4p

and 5p

transitions can be read


from Fig. 7.11 as 39.2, 42.4, 43.3 and 44.1 meV respectively. The np

transitions correspond to transitions from np hydrogenic states to the 1s


state, with energies given by:
h = [E
1s
E
np
[ .
The 1s energy is just equal to the value of R

found in the part (i), and


so we can identify R

0
= R

= 45.2 meV. The value of R

is then found
by tting the transition energies. A good t is found with R

25 meV.
(7.14) These are donor level transitions as sketched in Fig. 7.10(a). Their tran-
sition energies are given by eqn 7.30. It is apparent that they depend only
on the eective mass and relative permittivity of the host crystal, and
not on any of the properties of the dopant atom. The reason why this is
so is that the donor levels are formed by the interaction between a band
electron and an ionized impurity. The donor atoms are chosen so that
they have a single excess electron, and so they are all singly ionized. The
donor level energies are therefore hydrogenic with the relevant mass and
permittivity of the band electron.
On comparing with eqn 7.30, we see that we must have R

= (m

e
/m
0

2
r
)
R
H
. On setting this equal to 2.1 meV we nd m

e
= 0.036m
0
for
r
= 15.2.
(7.15) Doping with acceptors creates a p-type semiconductor with a series of
acceptor levels just above the valence band as indicated in Fig. 21. The
energy of the acceptor levels relative to the top of the valence band will
56
conductionband
valenceband
acceptor
levels
E
g
conductionband
valenceband
acceptor
levels
conductionband
valenceband
acceptor
levels
E
g
Figure 21: Acceptor to conduction band transitions in a p-type semiconductor
as discussed in Exercise 7.15. Note that this gure is not drawn to scale: the
acceptor energies are much smaller than the band gap.
be given by eqn 7.29 with m

e
replaced by m

h
. Electrons from the valence
band can be thermally excited to ll the acceptor levels, giving rise to
the possibility of acceptor to conduction band transitions as indicated in
Fig. 21. These transitions occur at a photon energy of E
g
E
A
. The ab-
sorption edge of a p-type semiconductor will therefore decrease on doping
from E
g
to E
g
E
A
, where E
A
is the energy of n = 1 acceptor level. A
wavelength shift from 5.26 m to 5.44 m corresponds to an energy shift
of 8 meV, and hence we deduce E
A
8 meV.
Students might get confused about the stipulation of 10 K for the temper-
ature in this exercise. At low temperatures we might expect the acceptors
to be frozen out, with holes in the acceptor levels rather than in the
valence band. The point is that the acceptors here are rather shallow and
a signicant fraction are ionized even at 10 K. It would not be possible to
observe the acceptor to conduction band transitions at higher tempera-
tures due to the thermal broadening of the edge. The exercise is in fact
based on real experimental data. See Figure 2 in Johnson, E.J. and Fan,
H.Y. (1965): Impurity and Exciton Eects on the Infrared Absorption
Edges of III-V Compounds, Phys. Rev. 139, A1991A2001.
(7.16) The peak is caused by Raman scattering from plasmon modes. The
energy of the scattered photons is given by eqn 7.34. In this exercise, we
are clearly considering the case of plasmon emission where the sign is
appropriate. We thus deduce:

p
=
in

out
= (2.410 2.321) eV = 0.089 eV.
On substituting into eqn 7.24 with
opt
= n
2
and the given value of m

, we
nd N = 4.2 10
24
m
3
. Note that we would also expect Raman signals
from optical phonons, but these would occur at smaller energy shifts with
300 cm
1
. (See Fig. 10.11.)
(7.17) We set

p
=
_
Ne
2

opt

0
m

e
_
1/2
=
LO
,
57
with
opt
= n
2
and solve for N. A wave number of 297 cm
1
implies
= 2c = 5.6 10
13
rad/s, and hence we nd N = 7.2 10
23
m
3
. A
mixed plasmonphonon mode is formed at this doping density because the
two longitudinal excitations couple strongly together if their frequencies
are the same.
58
Chapter 8
Molecular materials
(8.1) This is a standard example covered in all elementary quantum mechanics
texts. On substituting
1
into the Schrodinger equation, we obtain:


2
2m
_
x
2
a
4

1
a
2
_

1
+
1
2
m
2
x
2

1
= E
1
,
which can be re-arranged to give:
_
1
2
m
2


2
2ma
4
_
x
2

1
+

2
2ma
2

1
= E
1
.

1
is therefore a solution if we set a = (/m)
1/2
to eliminate the rst
term. We can then deduce:
E
1
=

2
2ma
2
=
1
2
.
By a similar method we can show that
2
and
3
are solutions with
E
2
= (3/2) and E
3
= (5/2) respectively, and with a again equal to
(/m)
1/2
.
(8.2) The energy levels for a particle of mass m in an innite potential well of
width d are given by eqn 6.13 as:
E
n
=
h
2
n
2
8md
2
.
Selection rules permit transitions with n equal to an odd number. (See
the discussion of intersubband transitions in Section 6.7.) The lowest
energy transition occurs for n = 1 2, with h = 3h
2
/8md
2
. On setting
m = m
0
for the -electron, and h = 2.5 eV, we nd d = 6.7 10
10
m.
This corresponds to about seven carboncarbon bonds.
(8.3) The ratio of the number of molecules with one vibrational quantum excited
compared to those with none is given by Boltzmanns law as exp(h/k
B
T),
where is the frequency of the vibration. The calculated ratios for the
three modes at 300 K are:
= 2 10
13
Hz: 4 10
2
,
= 4 10
13
Hz: 1.6 10
3
,
= 7 10
13
Hz: 1.4 10
5
.
The point of this exercise is to make the student appreciate that it is a
good approximation to assume that all the molecules are in the vibrational
ground state at room temperature.
59
(8.4) The energy of isolated hydrogen atoms is just given by the standard Ry-
dberg formula with E
n
= R
H
/n
2
. The energy required to promote one
of the atoms from a 1s to a 2p state is therefore (3/4)R
H
= 10.2 eV. This
is smaller than the equivalent transition in the H
2
molecule. The reason
is that the energy of a diatomic molecule is given by:
E
molecule
= E
atom1
+E
atom2
E
binding
,
where E
atomi
is the energy of the isolated atoms, and E
binding
is the
binding energy of the molecule. Now the ground state of the molecule
is more strongly bound than the excited state, and the transition includes
the dierence of the two binding energies. Hence in the molecule the
transition energy is:
h = (3/4)R
H
+ E
binding
.
We can then account for the molecular transition energy of 11.3 eV if we
assume that the dierence of the binding energies of the 1s1s and 1s2p
molecular congurations is 1.1eV.
0.5 1.0 1.5 2.0
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
U
(
r
)
r
0.5 1.0 1.5 2.0
-0.2
0.0
0.2
0.4
0.6
0.8
1.0
U
(
r
)
r
Figure 22: LennardJones potential for A = B = 1, as considered in Exercise
8.5.
(8.5) (i) The attractive r
6
term is caused by the van der Waals interaction,
which is the main attractive force between neutral molecules. This is a
dipole-dipole interaction. A uctuating dipole p
1
on molecule 1 generates
an electric eld of strength c
1
p
1
/r
3
at molecule 2. This induces a
dipole of magnitude p
2
c
1
p
1
/r
3
on molecule 2, which then generates
a eld c
2
p
2
/r
3
p
1
/r
6
at molecule 1. The interaction energy is then
p
1
c
2
(p
1
)
2
/r
6
. Although the time average of p
1
will be zero, the time
average of (p
1
)
2
is not. Hence the dipole-dipole mechanism generates an
attractive potential r
6
.
(ii) The Lennard-Jones potential is plotted for the case A = B = 1 in
Fig. 22. The potential is attractive at large r, but the repulsive term
dominates for small r. This gives a minimum at a well-dened value of r.
The position of the minimum r
0
can be calculated by setting:
dU
dr
=
12A
r
13
+
6B
r
7
= 0 ,
at r = r
0
, which implies r
0
= (2A/B)
1/6
.
60
(iii) The Taylor expansion for U(r) near r
0
is:
U(r) = U(r
0
) +
_
dU
dr
_
(r r
0
) +
1
2
_
d
2
U
dr
2
_
(r r
0
)
2
+ ,
where the derivatives are evaluated at r = r
0
. If r
0
is the position of the
minimum, then the rst derivative will be zero, and U(r) reduces to:
U(r) = U(r
0
) +
1
2
_
d
2
U
dr
2
_
(r r
0
)
2
+ .
Now
d
2
U
dr
2
= +
156A
r
14

42B
r
8
,
which, with r = r
0
= (2A/B)
1/6
, gives:
d
2
U
dr
2
=
18B
2
A
_
B
2A
_
1/3
.
Thus:
U(r) = U(r
0
) +
18B
2
A
_
B
2A
_
1/3
(r r
0
)
2
+ ,
U(r
0
) +
1
2

2
(r r
0
)
2
.
Hence
2
= (18B
2
/A)(B/2A)
1/3
.
0 2 4 6
4.6
4.8
5.0
5.2
5.4
T
r
a
n
s
i
t
i
o
n
e
n
e
r
g
y
(
e
V
)
Quantumnumber n
0 2 4 6
4.6
4.8
5.0
5.2
5.4
T
r
a
n
s
i
t
i
o
n
e
n
e
r
g
y
(
e
V
)
Quantumnumber n
Figure 23: Analysis of transition energies in Exercise 8.6.
(8.6) The absorption spectrum will consist of a series of vibronic lines with
energies given by eqn 8.3. The longest wavelength will correspond to
the transition with n = 0, and the others to increasing values of n. The
transition energies are plotted against this assignment of n in Fig. 23. The
good linear t conrms the assignment. The tting parameters that come
out of the analysis are
0
= 4.65 eV and
2
= 0.113 eV. We thus identify
the energy of the S
1
state as 4.65 eV, and nd /2 = 2.7 10
13
Hz.
61
E
Q
Q
0
S
0
S
1
S
2
5
.
7
e
V
7
.
4
e
V
n =5
n =6
0.13 eV
0.11 eV
E
Q
Q
0
S
0
S
1
S
2
5
.
7
e
V
7
.
4
e
V
n =5
n =6
0.13 eV
0.11 eV
Figure 24: Conguration diagram deduced from the data of Fig. 8.9 as consid-
ered in Exercise 8.7.
(8.7) The absorption spectrum shows a progression of vibronic lines obeying
eqn 8.3 with at least two excited electronic states. The rst progression
starts at 5.7 eV, and has a vibrational splitting of 0.11 eV. The intensity
of the lines peaks for n = 6. The second progression starts at 7.3 eV, has
a vibrational splitting of 0.13 eV, and has maximum intensity for n = 5.
We thus deduce that we have two excited states:
S
1
at energy 5.7 eV with = 0.11 eV,
S
2
at energy 7.3 eV with = 0.13 eV.
The line with the maximum intensity tells us about the relative values of
Q
0
for the states. The intensity peaks when the overlap of the vibronic
states is largest. The wave function of the ground state S
0
peaks at Q
0
,
while for the excited states the wave function gradually peaks more and
more at the classical turning points. The potential minima of S
1
and S
2
therefore occur so that the edge of their sixth and fth vibronic levels align
respectively with Q
0
for S
0
. We thus obtain the schematic conguration
diagram given in Fig. 24.
(8.8) The spin-orbit interaction introduces a coupling mechanism between the
spin and orbital angular momenta S and L. It is therefore possible to have
an interaction between dierent S states via their spin-orbit interaction
with a common L state. This interaction produces a small mixing of the
wave functions, so that triplet states contain a small admixture of singlet
character. This small singlet admixture gives a nite probability for a
triplet-to-singlet transition, which would otherwise be totally forbidden if
the spin states were pure.
Spin-orbit coupling is now routinely used to increase the intensity of phos-
phorescence in organic LEDs. A heavy metal (eg platinum) is introduced
into the molecule, and this increases the spin-orbit interaction, because the
spin-orbit coupling generally scales as Z
2
, where Z is the atomic number.
62
The use of heavy-metal dopants strongly enhances the triplet-to-singlet
transition rate, and hence the overall light emission eciency. This is
especially important in electrically-pumped devices, which are otherwise
limited to a maximum eciency of 25%. (See Exercises 8.15 and 8.16.)
(8.9) The weak nature of the emission and the long radiative lifetime indicates
that we are dealing with phosphorescence from a triplet state. The energy
level scheme for pyrromethene 567 would be qualitatively similar to that
of anthracene shown in Fig. 8.12. From the spectra shown in Fig. 8.10 we
deduce that the S
1
level has an energy of 2.3 eV, while the wavelength of
the phosphorescence indicates that the triplet lies at an energy of 1.6 eV.
In the case of optical excitation, the phosphorescence could be caused by
intersystem crossing from excited singlet states.
0 2 4
3.2
3.6
4.0
4.4
E
n
e
r
g
y
(
e
V
)
Vibronic number
crystal
solution
0 2 4
3.2
3.6
4.0
4.4
E
n
e
r
g
y
(
e
V
)
Vibronic number
0 2 4
3.2
3.6
4.0
4.4
E
n
e
r
g
y
(
e
V
)
Vibronic number
crystal
solution
crystal
solution
Figure 25: Analysis of the vibronic peaks of anthracene shown in Fig. 8.13 as
considered in Exercise 8.10.
(8.10) The assignment of the vibronic peaks of the solution is given in Fig. 8.13.
The energies of the vibronic peaks are plotted in Fig. 25 and a good
straight line is obtained. The linear t according to eqn 8.3 gives the
vibrational energy as 0.16 eV.
In the case of the crystal, we have rst to identify the various peaks in the
absorption spectrum. A strong vibronic progression with energies of 3.13,
3.30, 3.46 and 3.61 eV is observed in the data. These can be identied as
the 00, 01, 02, and 03 transitions, and a linear t according to eqn 8.3
gives the vibrational energy as 0.18 eV. (See Fig. 25.)
Two other features can also be identied in the absorption spectrum of
the crystal at 3.18 and 3.35 eV. These have the same splitting as the other
progression and therefore involve similar types of vibrations. The most
probable cause is a splitting of the electronic states in the lower symme-
try of the crystal, eg by the Davydov eect. (See Pope and Swenberg,
Electronic processes in organic crystals and polymers, 2nd edn, Oxford
University Press, 1999, Section I.D.5, pp 5966.)
(8.11) When the mirror symmetry rule works, we expect the emission spectrum
to be a mirror image of the absorption spectrum about the 00 transition.
(See, for example, Figs 8.10 or 8.17.) We thus expect a broad vibronic
63
band extending downwards from the 0-0 transition at 3.13 eV. The width
of the band will be about 1 eV. A series of vibronic peaks will occur with
energies given by h (3.13 n), with 0.18 eV. We would thus
expect peaks at 3.13, 2.95, 2.77, 2.59 eV .
(8.12) The S
1
absorption band has a 00 transition at 1.9 eV and extends to
2.8 eV. The emission band would thus have a 00 transition at 1.9 eV
and extend down to about 1.0 eV. The 01 vibronic peaks occurs at 2.1 eV
in the absorption spectrum, which implies a vibrational energy of 0.2 eV,
and hence a 01 transition in emission at around 1.7 eV.
(8.13) The dierence between the absorption and photoconductivity edges is
caused by excitonic eects. The absorption edge corresponds to the cre-
ation of tightly-bound (Frenkel) excitons. Since excitons are neutral par-
ticles, they do not contribute to the photoconductivity. The photoconduc-
tivity edge therefore corresponds to the band edge where free electrons and
holes are rst created. The dierence in the two edges gives the exciton
binding energy, which works out to be 1.1 eV.
It is important to realize that this is a dierent situation to that encoun-
tered for weakly-bound (Wannier) excitons. Weakly-bound excitons can
be easily ionized to produce free electrons and holes, and hence produce
a photocurrent. (See, for example, Figs 4.5 and 6.12.)
(8.14) The dominant vibrational frequency can be deduced by analysing the vi-
bronic progression of either the absorption or emission spectra in Fig. 8.17
according to eqns 8.3 or 8.5 as appropriate. This gives 0.17 eV, im-
plying 2.6 10
14
rad/s. With = 2.98 10
15
rad/s, we then nd

Raman
= () = 2.7210
15
rad/s, which is equivalent to a wavelength
of 693 nm.
Two points could be made here:
The molecule will have other vibrational modes, and these will give
additional Raman lines.
Not all vibrational modes are Raman-active, (see Section 10.5.2) and
it is not immediately obvious that the 0.17 eV mode responsible for
the vibronic spectra will show up in the Raman spectrum. In fact,
these modes are observed in the experimental Raman spectra, but
it requires a careful analysis by group theory to demonstrate this
theoretically.
(8.15) Optical excitation creates only singlets because the ground state is a
singlet and optical transitions do not change the spin. With only singlet
states excited, the recombination of the electrons and holes is optically
allowed for all the carriers.
On the other hand, with electrical injection there is no control of the
relative spin of the electrons and holes. The spins can be either parallel or
anti-parallel, and this gives rise to four possible total spin wave functions,
as indicated in Table 5. Three of these are triplets and only one is a
singlet state. The relative number of triplet and singlet excitons created
by electrical injection is therefore in the ratio 3:1, which implies that only
64
Wave function S
z
S State

h
+1 1 triplet
(
e

h
+
e

h
)/

2 0 1 triplet
(
e

h

e

h
)/

2 0 0 singlet

h
1 1 triplet
Table 5: Possible arrangements of relative electron-hole spins as discussed in
Exercise 8.15.
25% of the excitons are in singlet states. The remaining 75% are in triplet
states with very low emission probabilities. Hence the emission is expected
to be weaker than that for optical excitation by a factor of four.
The creation of triplets in electrically-driven organic LEDs is a serious
issue that limits their eciency. One way to enhance the eciency is
to increase the spin-orbit interaction to encourage inter-system crossing.
This is typically done by including a heavy metal atom in the molecule.
(See Exercise 8.8.)
(8.16) (i) As we have seen in Exercise 8.15, we expect that 75% of the excitons
created will be in triplet states with very low emission probabilities. Hence
the maximum quantum eciency that we can expect corresponds to the
number of singlet excitons that we create, namely 25%.
(ii) The number of electrons and holes owing into a device carrying a
current i is equal to i/e. The quantum eciency is dened as the ratio of
photons out to electrons in, and so the number of photons emitted will be
equal to i/e. The power emitted is then equal to h i/e, and we have:
P = 2.25 eV 25%10 mA/e = 5.6 mW.
(iii) The electrical power consumed by the device is equal to iV = 50 mW.
The power conversion eciency is thus equal to 5.6/50 = 11 %. The
eciency of a real device would be much lower, mainly due to the diculty
of collecting the photons, which are emitted in all directions (ie over 4
solid angle). Only a small fraction of these would be collected by the
optics. This latter point is exacerbated by the high refractive index of the
molecular material, which tends to limit the eective collection eciency
even further. (See Exercise 5.13.)
65
Chapter 9
Luminescence centres
(9.1) The solution for a one-dimensional potential well with innite barriers
is given in Section 6.3.2. In a cube the motion is quantized in all three
dimensions, and the energies for the x, y and z directions just add together.
For each direction, we have (cf eqn 6.13):
E =

2

2
2m

(2a)
2
n
2
where n is the quantum number. The quantum numbers for the three
degrees of freedom are independent of each other, and we derive eqn 9.4
by adding the quantized energies for the x, y and z directions together.
Note that the similar case of a quantum dot was considered previously in
Section 6.9.
(9.2) Equation 9.5 predicts E = 0.28/a
2
. The experimental energies are lower
because a real F-centre is not a rigid cubic box, and hence it would be
more appropriate to use a nite rather than innite potential well model.
As discussed in Section 6.3.3, the quantization energies of nite potential
wells are smaller than those of innite wells of the same dimension.
(9.3) We can set a = 0.33 nm and calculate h = 2.6 eV from eqn 9.5. Alterna-
tively, we can just read h 2 eV from Fig. 9.5.
2a
- +
+ -
- +
2a
+
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
+
- +
+ -
- +
-
+
-
+
-
+
F
2
+
2b
b
b
(a) (b)
2a
- +
+ -
- +
2a
+
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
+
- +
+ -
- +
-
+
-
+
-
+
F
2
+
2a
- +
+ -
- +
- +
+ -
- +
2a
+
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
- +
- +
+ -
+
- +
+ -
- +
- +
+ -
- +
-
+
-
-
+
-
+
-
+
+
-
+
F
2
+
F
2
+
2b
b
b
2b
b
b
(a) (b)
Figure 26: (a) An F
+
2
centre in an alkali halide with cell size 2a, as considered
in Exercise 9.4. The colour centre is modelled as a rectangular box with square
cross-section as shown in part (b).
(9.4) The energy of the electron will be given by eqn 6.47 with d
x
= d
y
= b and
66
d
z
= 2b. We therefore have:
E =

2

2
2m
0
_
n
2
x
b
2
+
n
2
y
b
2
+
n
2
z
4b
2
_
.
We can apply this model to the F
+
2
centre by taking the appropriate value
of b. The rectangle equivalent to the F
+
2
centre is sketched in Fig. 26. If the
cubic unit cell size is 2a, then the square end of the box has dimension

2a,
and the longer dimension is 2

2a. We therefore need to take b =

2a for
an F
+
2
centre. The lowest energy transition is the n
z
= 1 2 transition,
which has an energy of h = 3h
2
/32m
0
b
2
. With b =

2a, this gives
h = 3h
2
/64m
0
a
2
, which is half the value given in eqn 9.5.
The experimental absorption peak of KF:F
+
2
occurs at 1.1 eV (see Fig. 9.6),
while that of KF:F occurs at 2.9 eV (see Fig. 9.5.) The experimental ra-
tio is thus 0.4, which is close to the predicted value of 0.5. This is
remarkably good agreement considering the simplicity of the model.
(9.5) (a) We use average values for the atomic numbers of the relevant series:
i.e. we take n = 3 and Z = 25 for the 3d series, and n = 4 and Z = 64 for
the 4f series. We thus obtain a crude estimate of the ratio of their radii
from:
r
3d
r
4f

(3
2
/25)
(4
2
/64)
1.4 .
This shows that the 3d series is expected to have a larger radius by a
factor of 1.4.
(b) The 3d transition metal ions have lost the 4s electrons, and so the 3d
electrons are the outermost orbitals. By contrast, the 4f orbitals in the
rare earths are inside the lled 5s and 5p orbitals. The 5s and 5p orbitals
have n = 5 and therefore have larger radii than the 4f shell due to the n
2
dependence of r.
(9.6) (i) The three p orbitals are dumb-bell shaped as shown in Fig. 27(a) for the
case of the p
z
orbital. In an octahedral lattice, the x, y and z directions
are all equivalent. This implies that the p
x
, p
y
and p
z
orbitals must all
experience the same interaction energy with the crystal. This is apparent
from Fig. 27(b), which shows that the distance from the electron cloud to
the ions is the same for the the p
z
, p
x
and p
y
orbitals. Hence they will
experience identical Coulomb interactions.
(ii) In a uniaxial crystal, the octahedral symmetry is lost and the z direc-
tion is now dierent. This means that the p
z
orbitals are closer to the ions
than the p
x
or p
y
orbitals. (See Fig. 27(c).) The Coulomb interactions
between the electron cloud will now be dierent for the p
z
orbital and the
other two, and so its energy will be dierent. On the other hand, the x
and y directions remain equivalent (see lower half of Fig. 27(c)), and so
the p
x
and p
y
orbitals are still degenerate. Hence the triplet p state splits
into a singlet and a doublet.
(iii) If the nearest neighbour ions are negative, the p
z
electrons will ex-
perience a stronger repulsive interaction with the lattice because of the
smaller distance to the ion. Hence the p
z
states will have a larger energy
67
z
x, y
y
x
z
x
z
x, y
(a) (b) (c) (d)
smaller
distance
z
x, y
z
x, y
y
x
y
x
z
x
z
x
z
x, y
z
x, y
(a) (b) (c) (d)
smaller
distance
Figure 27: Discussion of p orbitals as required for Exercise 9.6. (a) A p
z
orbital.
(b) p
z
, p
x
and p
y
orbitals in an octahedral lattice, as seen in the x-z and x-y
planes. (c) p
z
, p
x
and p
y
orbitals in a uniaxially-distorted lattice, as seen in the
x-z and x-y planes. (d) Splitting of the p orbitals in the distorted lattice, with
negative nearest neighbours.
than the p
x
and p
y
states. This will give a splitting as shown in Fig. 27(d),
with the singlet at higher energy.
(9.7) The 1064 nm line in Nd:YAG corresponds to a transition from the 11 502 cm
1
level as shown in Fig. 9.8(b). The relative populations of the 11 502 cm
1
and 11 414 cm
1
levels of the
4
F
3/2
term are proportional to the Boltz-
mann factor:
N(11502 cm
1
)
N(11414 cm
1
)
= exp
_

E
k
B
T
_
,
where E = 88 cm
1
. This ratio equals 0.19 at 77 K and 0.66 at 300 K.
The spontaneous emission rate increases in proportion to these factors,
and therefore the relative intensity of the 1064 nm line increases with T.
(9.8) The transition rates between the two levels are governed by the Einstein
coecients, the populations of the levels, and the light energy density.
(See Section B.1 in Appendix B.) Three types of transitions are possible,
namely spontaneous emission, stimulated emission, and absorption. For
gain, we need that the stimulated emission rate should exceed the absorp-
tion rate. (Spontaneous emission is negligible at high light intensities.)
The condition for this to occur is (see eqns B.5 and B.6):
B
21
N
2
u() > B
12
N
1
u() ,
which implies:
N
2
N
1
>
B
12
B
21
.
On substituting from eqn B.10, we derive the condition for net gain:
N
2
N
1
>
g
2
g
1
,
68
where g
2
and g
1
are the degeneracies of the two levels. This condition is
called population inversion.
0
1
2
PUMP
LASEREMISSION
694.3nm
rapiddecay
groundstate
PUMPINGBANDS
0
1
2
PUMP
LASEREMISSION
694.3nm
rapiddecay
groundstate
PUMPINGBANDS
Figure 28: Three-level laser scheme for ruby, as required for Exercise 9.9.
(9.9) Ruby is a three-level laser, with a level scheme as shown in Fig. 28. The
lower laser level (level 0) is the ground state, and the upper level is an
excited state (level 2). Ruby has strong absorption bands in the green/blue
spectral regions (see Fig. 1.7), and these are used as intermediate pumping
bands (level 1) to produce population inversion after rapid decay to the
upper lasing level.
When the pump is turned o, all the atoms will be in the ground state,
so that:
N
0
= N ,
N
2
= 0 .
where N is the total number of atoms. When the pump is turned on, N
atoms will be pumped to the upper laser level, so that:
N
0
= N N .
N
2
= N .
For population inversion, we require N
2
> N
0
, which implies N > N/2.
Population inversion is therefore only achieved when more than 50% of
the atoms are pumped to the excited state.
In this particular example, we have N
2
= 0.6N at t = 0, and we have seen
above that the laser will stop oscillating when N
2
= 0.5N. The number of
atoms that make stimulated radiative transitions (and hence the number
of photons emitted) during the laser pulse is therefore 0.1N. This gives a
pulse energy of 0.1N h, where h is the laser photon energy, namely
1.79 eV. On inserting the appropriate numbers we nd:
E
pulse
= 0.1 (10
25
10
6
) 1.79 eV = 0.3 J .
Please note: students might have diculty with the exercise if they have
not already done a separate laser physics course: the information provided
69
in the text is insucient for them to be able to answer it properly. I shall
endeavour to correct this in the second edition, should there ever be one.
(9.10) (i) The optical intensity I(t) is proportional to [c(t)[
2
, and hence if we
have Gaussian intensity pulse as dened in the exercise, we have a time-
varying electric eld of the form (note the extra factor of two):
c(t) = c
0
exp(t
2
/2
2
) e
i0t
,
where
0
is the centre angular frequency. The spectrum of the pulse is
found by rst taking the Fourier transform of c(t):
c() =
1

2
_
+

c(t) e
it
dt ,
=
c
0

2
_
+

exp(t
2
/2
2
) e
i(0)t
dt ,
= c
0
exp[
2
(
0
)
2
/2] ,
where we used the standard integral:
F() =
_
+

exp(t
2
/
2
)e
it
dt =

exp(
2

2
/4) ,
in the last line. The nal result is obtained by using:
I() [c()[
2
exp[
2
(
0
)
2
] .
(ii) The full width at half maximum (FWHM) of the pulse in the time
domain is found by nding the times for which I(t) = I
0
/2, i.e. by solving:
exp(t
2
/
2
) = 0.5 .
This gives t =

ln 2 , so that t = 2

ln 2 . The FWHM of the pulse


in the frequency domain is likewise found by solving:
exp[
2
(
0
)
2
] = 0.5 ,
which gives (
0
) =

ln 2/, and hence = 2

ln 2/. We then
obtain:
t = t/2 = 2 ln 2/ = 0.441 .
(9.11) The crystal-eld shifts of the energy levels depend on the local environ-
ment of the ion. (e.g. small perturbations to the atomic positions aect
the energy levels of the ion through the alterations to the local electric
elds the ion experiences.) There will be much larger inhomogeneity in
the local environment in a glass than in a crystal, due to the lack of long-
range order. Hence we expect much larger inhomogeneous broadening of
the crystal-eld split transitions in a glass than in a crystal.
If we assume a Gaussian pulse, we expect a time-bandwidth product of
0.441, and hence t
min
= 60 fs. Other pulse shapes would give comparable
minimum pulse durations.
70
(9.12) The subscript g stands for gerade and implies even parity. A gg transi-
tion therefore involves no parity change and is forbidden for electric-dipole
transitions. (See Section B.3.) The transition must therefore take place
by a higher-order process, e.g. the magnetic-dipole or electric-quadrupole
interaction. These have much lower probabilities than electric-dipole pro-
cesses, and hence give long excited state lifetimes in the microsecond or
millisecond range. Since the lifetime is long, the radiation would be clas-
sied as phosphorescence rather than uorescence. (See Section B.3.)
(9.13) The excited state lifetime is determined by both radiative and non-
radiative processes. It follows from eqn 5.4 that:
1

=
1

R
+
1

NR
.
The radiative lifetime is governed by atomic transition probabilities and
is not expected to vary signicantly with the temperature. On the other
hand, the non-radiative transition rate is governed by phonon-assisted
processes and is expected to increase strongly with T. On substituting

R
= 1.8 ms into the equation above, we nd
NR
= 6.3 ms at 77 K
and 0.062 ms at 300 K. This implies, through eqn 5.5, that the radiative
quantum eciency is 78% at 77 K and 3% at 300 K. The radiative eciency
is too low at 300 K to allow lasing.
(9.14) The level scheme for Ti:sapphire is shown in Fig. 9.13. If we assume that
the radiative quantum eciency is 100% (i.e. one laser photon emitted
for each pump photon absorbed), then the ratio of the output power to
the input power would just be proportional to the ratio of the respective
photon energies, which implies:
P
out
=
h
out
h
in
P
in
=
1/800
1/514
P
in
= 3.2 W.
In this case, the remaining 1.8 W of power produces phonons (i.e. heat)
in the crystal. Note that a substantial amount of heat is generated in
the crystal even for the ideal case of 100% quantum eciency due to the
dierence in the photon energies of the argon and Ti:sapphire lasers.
In reality, a Ti:sapphire laser typically gives about 1 W for a pump power
of 5 W at 500 nm. This reduction of the power from the ideal value is
caused by a number of factors:
less than unity quantum eciency due to signicant non-radiative
decay;
1
imperfect absorption of the pump laser in the crystal;
optical losses within the cavity.
1
The operating temperature of the laser crystal will be above room temperature due to
the heat generated within it, and this further increases the non-radiative decay rate. Cooling
of the laser rod is therefore essential.
71
Chapter 10
Phonons
(10.1) The phonon modes of purely covalent crystals do not give rise to infrared
absorption because the atoms are neutral and do not interact with the
electric eld of the light wave. Of the ve materials listed, germanium
and argon are elemental, and must therefore have neutral atoms with non-
polar bonds, and hence no infrared absorption. The other three, namely
ice, ZnSe and SiC, are polar, and would therefore have some infrared-active
phonons.
(10.2) It is apparent from eqn 1.26 that R = 0 when n = 1, and hence
r
= 1.
For an undamped oscillator,
r
() is given by eqn 10.16. We thus solve:

r
() =

+ (
st

)

2
TO
(
2
TO

2
)
= 1 ,
for . Rearrangement gives:

2
=
_

st
1

1
_

2
TO
,
leading to the result quoted in the exercise.
(10.3) The exercise follows example 10.1(i). We calculate
LO
= 20 THz from
the LST relationship, and hence that the reststrahlen band runs from 9.2
to 20 THz, i.e. 15 m to 33 m.
(10.4) The exercise closely follows example 10.1(ii). The relative permittivity
is given by eqn 10.10 as:

r
() = 10 +
210
100
2
i

/2
,
where is measured in THz and

= /10
12
. We calculate
LO
= 11 THz
from the LST relationship, so that the reststrahlen band runs from 10 to
11 THz. We thus need to evaluate
r
at = 10.5 THz.
(a) With = 10
11
s
1
, we nd
r
= 10.48+0.336i at 10.5 THz, and hence
that n = 0.0519 and = 3.238 from eqns 1.2223. Then from eqn 1.26 we
nd R = 0.98.
(b) With = 10
12
s
1
, we nd
r
= 9.958+0.252i and n = 0.509+3.196i
at 10.5 THz, and hence that R = 0.84.
(10.5) (i) We identify the reststrahlen band from the region of high reectivity
from 3032 m. (See Fig. 29(a).) On equating the upper and lower wave-
length limits with
TO
and
LO
respectively, we nd
TO
9.5 THz and

LO
10 THz.
72
16 20 24 28 32 36 40
0.0
0.2
0.4
0.6
0.8
1.0
R
e
f
l
e
c
t
i
v
i
t
y
Wavelength( mm)
AlSb
n
TO
n
LO
R()
R(0)
(a)
0.0 0.4 0.8 1.2 1.6
0.84
0.88
0.92
0.96
1.00
R
e
f
l
e
c
t
i
v
i
t
y
g (10
12
s
-1
)
(b)
16 20 24 28 32 36 40
0.0
0.2
0.4
0.6
0.8
1.0
R
e
f
l
e
c
t
i
v
i
t
y
Wavelength( mm)
AlSb
n
TO
n
LO
R()
R(0)
(a)
16 20 24 28 32 36 40
0.0
0.2
0.4
0.6
0.8
1.0
R
e
f
l
e
c
t
i
v
i
t
y
Wavelength( mm)
AlSb
n
TO
n
LO
R()
R(0)
(a)
0.0 0.4 0.8 1.2 1.6
0.84
0.88
0.92
0.96
1.00
R
e
f
l
e
c
t
i
v
i
t
y
g (10
12
s
-1
)
(b)
0.0 0.4 0.8 1.2 1.6
0.84
0.88
0.92
0.96
1.00
R
e
f
l
e
c
t
i
v
i
t
y
g (10
12
s
-1
)
(b)
Figure 29: (a) Interpretation of the data in Fig. 10.14 as required for Exercise
10.5(i) and (ii). (b) Calculated reectivity versus damping constant, as required
for Exercise 10.5(iii).
(ii) The high and low frequency permittivities can be deduced from the
asymptotic reectivities. (See Fig. 29(a).) The low frequency limit gives

st
from (see eqn 1.26, with n =

r
):
R(0) =
_

st
1

st
+ 1
_
2
.
At 0, there is no absorption, and so
r
will be real. On reading
R 0.30 R(0) at long wavelengths, we deduce
st
12. On similarly
equating the short wavelength limit of R, namely 26%, with R(), we
deduce

9.5.
(iii) The peak reectivity is about 90%, and is limited by , which in
turn is determined by the lifetime of the TO phonons. The middle of the
reststrahlen band occurs at 9.75 THz. We thus need to evaluate
r
from
(see eqn 10.10, with measured in THz and

= /10
12
):

r
() =

+ (
st

)

2
TO
(
2
TO

2
) i/2
= 9.5 +
225
90
2
i

/2
,
at = 9.75 THz. We split this into the real and imaginary parts, compute
n and from eqns 1.2223, and R from eqn 1.26. The reectivity calcu-
lated in this way is plotted as a function of in Fig. 29(b). It is apparent
that we have R = 0.9 for = 8.6 10
11
s
1
. This implies, from = 1/,
that = 12 ps.
The values of
LO
and
TO
found in part (i) can be compared to the
LyddaneSachsTeller relationship, which predicts
LO
/
TO
= 1.11. The
experimental ratio is slightly smaller. The values given here are only
approximate, and depend on how exactly they are extracted from the
data. The departure from LST is therefore not very signicant.
(10.6) The exercise closely follows example 10.1(iii). We rst use eqn 10.18
to nd
r
at
TO
, which gives
r
= 10 + 132i for = 10
12
s
1
and

r
= 10 + 1320i for = 10
11
s
1
. We then use eqn 1.23 to nd and
73
eqn 1.16 to nd .
(a) For = 10
12
s
1
we nd = 7.8 and = 3.4 10
6
m
1
.
(b) For = 10
11
s
1
we nd = 26 and = 1.1 10
7
m
1
.
Note that the peak absorption increases for the smaller value of the damp-
ing, as normal for a damped oscillator.
(10.7) The peak reectivity is governed by the damping constant . (See, for
example, Fig. 29(b) above.) As the temperature increases, we expect
to increase, and hence R to decrease, due to the increased probability
of anharmonic decay processes of the type illustrated in Fig. 10.13. The
reason why anharmonic phonon decay increases with T is that phonons are
bosons, and the probability for phonon emission increases as the thermal
population of the nal state increases. (n.b. This contrasts with fermions,
for which the transition probabilities decrease with increasing occupancy
of the nal state.)
(10.8) With negligible damping, we can use eqn 10.16 to calculate
r
= 21.5 at
8 THz. We then substitute this value of
r
into eqn 10.19 to compute the
wave vector. This gives:
q =

c
=

21.5 2 8 10
12
3 10
8
= 7.8 10
5
m
1
.
(10.9) (i) The condition for cyclotron resonance is given in eqn 10.24. In a polar
material, the mass that is measured is the polaron mass m

. We thus
obtain:
m

=
eB
2c
= 0.097 m
0
.
(ii) The rigid lattice mass m

can be calculated from eqn 10.25. On


inserting the relevant values into eqn 10.20, we nd
ep
0.33 for CdTe.
Then from eqn 10.25 we nd that m

= 0.097 m
0
implies m

= 0.092 m
0
.
(10.10) The Raman spectra for a number of III-V crystals are shown in Fig. 10.11.
In each case we observe two peaks: one for the TO phonons and the other
for the LO phonons. These two phonon modes have dierent frequencies
because III-V compounds have polar bonds with partially charged atoms.
They therefore interact with light, and obey the LST relationship. The
situation in diamond is dierent because it is a purely covalent crystal,
with neutral atoms that do not interact with the light. The LST analysis
does not apply, and the optical phonons are degenerate at q = 0. The
Raman spectrum therefore has only one peak for the optical phonons.
(10.11) Silicon, like diamond in the previous exercise, is covalent, and its LO
and TO phonons are degenerate at q = 0. The two peaks correspond to
the Stokes and anti-Stokes lines from these degenerate optical phonons.
The line at 501.2 nm is shifted up in frequency compared to the laser and
is thus the anti-Stokes line, while that at 528.6 nm is the Stokes line. The
phonon frequency can be worked out from eqn 10.28, which gives, for the
case of the Stokes line:
/2 =
c
514.5 nm

c
528.6 nm
= 15.5 THz .
74
The relative intensities of the lines are given by eqn 10.30:
I(501.2 nm)
I(528.6 nm)
= exp
_
h 15.5 10
12
k
B
T
_
= 0.08 .
invert invert
Figure 30: Inversion of a TO mode in an ionic crystal, as considered in Exercise
10.12. The inversion centre is circled.
(10.12) NaCl has an inversion centre and so the mutual exclusion rule applies. It
is apparent from Fig. 30 that the TO mode has odd parity under inversion.
The TO mode is therefore IR active but not Raman active.
Crystal Raman line 1 Raman line 2 TO phonon energy LO phonon energy
cm
1
cm
1
meV meV
GaAs 262 286 32.5 35.5
InP 299 341 37.1 42.3
AlSb 312 332 38.7 41.1
GaP 364 403 45.1 50.0
Table 6: Raman shifts deduced from the data in Fig. 10.11, as considered in
Exercise 10.13.
(10.13) The energies of the phonon modes can be deduced directly from the
Raman spectra by applying eqn 10.28, with the + sign as appropriate for
a Stokes shift. This shows that the Raman shift is exactly equal to the
phonon frequency. For each crystal, two lines are observed. The lower
frequency line comes from the TO phonons, while the higher frequency
line originates from the LO phonons. The Raman shifts in cm
1
from
Fig. 10.11 are given in Table 6, together with the energies deduced ac-
cording to:
E (meV) = 0.124 Raman shift (cm
1
) .
On comparing the frequencies of the TO and LO phonons of GaAs in
Table 6 with those deduced from the infrared reectivity data in Fig. 10.5,
we see that there is a small shift of a few wave numbers between the two
sets of data. This is caused by the slight decrease of the optical phonon
frequencies between 4 K and 300 K.
(10.14) Equation 10.29 implies that momentum is conserved during the Ra-
man scattering process so that the vectors form a triangle as depicted in
Fig. 31(a). In the case of inelastic scattering by acoustic phonons, the
frequency shift of the photon is very small because . This implies
that the magnitude of the photon wave vector hardly changes, so that we
can approximate:
[k
1
[ = [k
2
[ k =
n
c
.
75
k
1
k
2
q
q/2
k
1
k
2
q
(a) (b)
k
1
k
2
q
q/2
k
1
k
2
q
(a) (b)
Figure 31: (a) Conservation of momentum during Raman scattering by a phonon
of wave vector q, as as considered in Exercise 10.14. (b) Back-scattering geom-
etry.
It is then apparent from Fig. 31(a) that:
q
2
= k sin

2
=
n
c
sin

2
.
On writing q = /v
s
as appropriate for acoustic phonons, we derive the
result in the exercise. Equation 10.33 then follows by writing:
= [
2

1
[ = .
In back-scattering geometry, we have = 180

, so that q = 2k. (See


Fig. 31(b) with [k
1
[ = [k
2
[ = k.) It then follows from eqn 10.33 with
sin(/2) = 1 that:
v
s
=
c
2n
=

2n
.
On inserting the data given in the exercise, we nd v
s
= 813 m s
1
.
(10.15) (i) The negative term in r
1
is the attractive potential due to the
Coulomb interaction between the ions. The Madelung constant ac-
counts for the summation of the contributions of the positive and nega-
tive ions over the whole crystal. The positive term in r
12
represents the
short range repulsive force due to the Pauli exclusion principle when the
electron wave functions overlap.
(ii) The graph of U(r) is qualitatively similar to that for the Lennard-Jones
potential, being attractive for large r, repulsive for small r, and with a
minimum at some intermediate value of r, labelled r
0
. (cf. Fig. 22.) The
value of r
0
is found by dierentiating U(r):
dU
dr
=
12
r
13
+
e
2
4
0
r
2
,
and nding the value of r for which dU/dr = 0, namely:
12
r
13
0
=
e
2
4
0
r
2
0
,
which implies r
11
0
= 12 4
0
/e
2
.
(iii) The Taylor series for U(r) expanded about r
0
is:
U(r) = U(r
0
)+
dU
dr r=r0
(rr
0
)+
1
2
d
2
U
dr
2
r=r0
(rr
0
)
2
+
1
6
d
3
U
dr
3
r=r0
(rr
0
)
3
+ .
76
Now U(r) is a minimum at r
0
, and the rst derivative is therefore zero,
so that:
U(r) = U(r
0
) +
1
2
d
2
U
dr
2
r=r0
(r r
0
)
2
+
1
6
d
3
U
dr
3
r=r0
(r r
0
)
3
+ .
We can reconcile this with eqn 10.34 by taking x = r r
0
and U(x) =
U(r) U(r
0
). It is then apparent that:
C
2
=
1
2
d
2
U
dr
2
r=r
0
,
C
3
=
1
6
d
3
U
dr
3
r=r0
.
At r = r
0
we have
d
3
U
dr
3
=
2184
r
15
0
+
6e
2
4
0
r
4
0
,
=
_
6e
2
4
0

2184
r
11
0
_
1
r
4
0
,
=
176e
2
4
0
r
4
0
where we used the result of part (ii) to derive the last line. Hence:
C
3
= 22e
2
/3
0
r
4
0
.
(10.16) If the Raman spectrum is lifetime broadened, we shall have a Lorentzian
line shape with:
t =
1
2
.
Hence with t = , and = c, we have:
=
1
2c
.
On inserting the data given in the exercise, we nd = 6 ps. This value
agrees with the lifetime measured by time-resolved Raman scattering. See:
von der Linde et al., Phys. Rev. Lett. 44, 1505 (1980).
77
Chapter 11
Nonlinear optics
(11.1) In the Bohr model for hydrogen, the radius of the electron in the nth
quantum level is given by:
r
n
=
4
0

2
me
2
n
2
Z
=
n
2
Z
a
H
.
The magnitude of the electric eld is given by the standard Coulomb
formula:
c =
Ze
4
0
r
2
,
which, on inserting r
n
from the Bohr formula, gives:
c =
Ze
4
0
r
2
n
=
e
4
0
a
2
H
Z
3
n
4
.
For the outer 3s and 3p electrons in atomic silicon, use n = 3 and an
eective nuclear charge Z 4.
2
We then obtain a value of c = 5
10
11
Vm
1
. The eld for the conduction electrons in crystalline silicon
would, of course, be dierent due to the high relative permittivity and the
low eective mass.
(11.2) We can relate the optical intensity to the electric eld by using eqn A.40.
(a) The optical intensity is found from:
I =
P
A
=
E
pulse
/
pulse
r
2
=
1 10
8
(2.5 10
3
)
2
= 5.1 10
12
W/m
2
.
Hence with n = 1 for air, we nd from eqn A.40 that c = 6.210
7
V m
1
.
(b) The optical intensity is found from:
I =
P
A
=
10
3
20 10
12
= 5 10
7
W/m
2
.
Then with n = 1.45 as appropriate for the bre, we nd from eqn A.40
that c = 1.6 10
5
Vm
1
.
(11.3) With no external eld applied, the gas is isotropic and therefore possesses
inversion symmetry. Hence
(2)
= 0, and no frequency doubling will occur.
With the electric eld applied, the gas is no longer isotropic as the electron
clouds of the atoms will be distorted along the axis dened by the eld.
This means that inversion symmetry no longer holds, so that
(2)
,= 0 and
frequency doubling can, in principle, occur. However, it would give a very
weak signal due to the low density of atoms.
2
Z
e
is the dierence between the nuclear charge and the total number of inner shell
electrons that screen the valence electrons from the nucleus. i.e. Z
e
= 14 10, where 10 is
the total number of electrons in the 1s, 2s and 2p shells.
78
(11.4) The second-order nonlinear susceptibility is zero if the material has an
inversion centre. We must therefore consider the microscopic structure to
see if the material has inversion symmetry or not.
(a) NaCl is a face-centred cubic crystal with inversion symmetry:
(2)
= 0.
(b) GaAs has the zinc-blende structure, which is similar to the diamond
structure except that the bonds are asymmetric. It does not possess in-
version symmetry and therefore
(2)
,= 0.
(c) Water is a liquid and is therefore isotropic; hence inversion symmetry
applies and
(2)
= 0.
(d) Glass is amorphous and has no preferred axes; inversion symmetry
applies and
(2)
= 0.
(e) Crystalline quartz is a uniaxial crystal with the trigonal 3m structure.
It does not possess inversion symmetry, so that
(2)
,= 0.
(f) ZnS has the hexagonal wurtzite structure (6mm), without inversion
symmetry. Hence
(2)
,= 0.
(11.5) (i) Consider the absorption and stimulated emission transitions as indi-
cated in Fig. 11.2. (Spontaneous emission can be neglected if u

is suf-
ciently large.) The absorption and stimulated emission rates are equal
to N
1
B
12
u

and N
2
B
21
u

respectively (see eqns B.56.). If the levels are


non-degenerate, then eqn B.12 tells us that B
12
= B
21
. At t = 0, all the
atoms are in level 1, and there is net absorption, which increases N
2
and
decreases N
1
. As the atoms are pumped to level 2, the stimulated emis-
sion rate becomes increasingly signicant. Eventually, we reach a stage
where N
1
= N
2
= N
0
/2, and the stimulated emission and absorption rates
are identical. There is therefore no net absorption or emission, and N
2
cannot increase further. The maximum value of N
2
that can be achieved
is therefore N
0
/2.
3
(ii) If we only have two levels and we neglect spontaneous emission, then
the rate equations for N
1
and N
2
are:
dN
1
dt
= B
12
N
1
u

+B
21
N
2
u

,
dN
2
dt
= +B
12
N
1
u

B
21
N
2
u

.
On setting B
12
= B
21
as appropriate for non-degenerate levels, and sub-
tracting, we nd:
d
dt
(N
1
N
2
) =
dN
dt
= 2B
12
u

N ,
where N = N
1
N
2
. Integration yields:
N(t) = N(0) exp(2B
12
u

t) ,
which, with N(0) = N
0
, gives the required result.
The result quanties the way the laser pumps atoms from level 1 to level
2, and hence reduces the net absorption. The equation implies that the
3
This shows that it is not possible to achieve population inversion (i.e. N
2
> N
1
) in a
two-level system: three or more levels are required. This is why lasers, in which population
inversion is essential, are always classied as either three or four-level systems.
79
populations will eventually equalize no matter how weak the laser beam
is. This unphysical result arises from neglecting spontaneous emission and
transitions to other levels that occur in real atoms.
directionofpropagation
x
y
E
q
directionofpropagation
x
y
E
q
Figure 32: Propagation and polarization vectors for the light wave considered
in Exercise 11.6.
(11.6) With the beam propagating in the x-y plane and with its polarization in
the same plane, the light must be linearly polarized as shown in Fig. 32.
The z-component of the electric eld is therefore zero. The nonlinear
polarization, for the given nonlinear optical tensor, is found from eqn 11.46
to be:
_
_
_
P
(2)
x
P
(2)
y
P
(2)
z
_
_
_ =
_
_
0 0 0 d
14
0 0
0 0 0 0 d
25
0
0 0 0 0 0 d
36
_
_
_
_
_
_
_
_
_
_
c
x
c
x
c
y
c
y
c
z
c
z
2c
y
c
z
2c
z
c
x
2c
x
c
y
_
_
_
_
_
_
_
_
=
_
_
2d
14
c
y
c
z
2d
25
c
z
c
x
2d
14
c
y
c
z
_
_
.
On setting c
z
= 0, we obtain:
_
_
_
P
(2)
x
P
(2)
y
P
(2)
z
_
_
_ =
_
_
0
0
2d
14
c
y
c
z
_
_
.
Only P
(2)
z
is non-zero, and therefore the second harmonic beam must be
polarized along z.
We assume that the direction of propagation makes an angle with respect
to the x axis as shown in Fig. 32 and that the light has an electric eld
magnitude of c
0
. It will then be the case that c
x
= c
0
cos and c
y
=
c
0
sin , and hence that:
P
(2)
z
= 2d
36
c
2
0
cos sin = d
36
c
2
0
sin 2 .
This is maximized when 2 = 90

: i.e. = 45

.
(11.7) We consider an electro-optic crystal with axes as dened in Fig. 33. The
voltage is applied so as to produce an electric eld of magnitude c
z
along
z axis.
(i) If the light propagates along the z axis, the polarization vector will
lie in the x-y plane, or equivalently, in the x

-y

plane. We resolve the


80
y
z
x
y
x
L
V
y
z
x
y
x
L
V
Figure 33: Experimental geometry of the electro-optic crystal considered in
Exercise 11.7.
light polarization vector into its components along the x

and y

axes.
The phase shift induced by a refractive index change n in a medium of
length L is, in general, given by:
=
2

nL.
On applying this to the two components along the x

and y

axes, we then
have:

x
=
2L

n
x
=
Ln
3
0
r
41

c
z

y
=
2L

n
y
=
Ln
3
0
r
41

c
z
where n
x
and n
y
are the eld-induced refractive index changes along
the x

and y

axes as given in the exercise. The phase dierence is


thus given by (with c
z
= V/L):
=
x

y
=
2Ln
3
0
r
41
c
z

=
2n
3
0
r
41
V

.
(n.b. The formula given in the text is wrong by a factor of two.)
(ii) On setting = , we nd:
V

=

2n
3
0
r
41
.
With the appropriate gures for CdTe given in the exercise, we obtain
V

= 44 kV.
(11.8) This exercise closely follows Example 11.2, and the phase-matching angle
is found by substituting the appropriate refractive indices into eqn 11.56.
With the data given in the exercise, this gives:
1
(1.506)
2
=
sin
2

(1.490)
2
+
cos
2

(1.534)
2
.
On using cos
2
= 1 sin
2
and re-arranging, we nd sin
2
= 0.626 and
hence = 52.3

.
81
(11.9) In a third-order nonlinear medium, the change in the relative permittivity
is given by (see eqn 11.62):

r
=
1
+ i
2
=
(3)
c
2
.
We therefore have
2
= Im(
(3)
)c
2
and hence that
2
Im(
(3)
)I
because I c
2
. It follows from eqns 1.16 and 1.21 that
2
.
Hence, in a medium with Im(
(3)
) ,= 0, we expect
Im(
(3)
)I .
A saturable absorber has an absorption coecient which obeys eqn 11.40.
In the limit of small intensities, this is of the form:
(I) =
0

0
I/I
s
=
0
,
where =
2
I and
2
=
0
/I
s
. We thus have an intensity dependence
exactly as described above, and we therefore conclude that the saturable
absorber must have Im(
(3)
) ,= 0.
(11.10) We can choose our axes as we please in an isotropic medium. Therefore
choose z as the direction of propagation, and x as the polarization vector,
so that the electric eld is given by:
E = (c
x
, 0, 0) .
The third-order nonlinear polarization is given by eqn 11.12, and, with
c
y
= c
z
= 0, the nonlinear polarization is of the form:
_
_
_
P
(3)
x
P
(3)
y
P
(3)
z
_
_
_ =
0
c
3
x
_
_

xxxx

yxxx

zxxx
_
_
.
However, we see from Table 11.3 that
yxxx
=
zxxx
= 0. Hence we nd
P =
0

xxxx
c
3
x
(1, 0, 0) ,
which means that P is parallel to E.
(11.11) This exercise is very similar to Example 11.3. Form eqn 11.70, we
require that:

nonlinear
=
2

n
2
Il = ,
which implies:
I =

2n
2
l
=
1.55 10
6
2 3 10
20
10
= 2.6 10
12
Wm
2
.
The optical power to produce this intensity is given by:
P = IA = 2.6 10
12
(2.5 10
6
)
2
= 50 W.
82
(11.12) It is apparent from Fig. 11.8 that the presence of electrons causes the
states in the conduction band up to E
c
F
to be lled up, and likewise for
the holes in the valence band. The absorption between E
g
and (E
g
+E
c
F
+
E
v
F
) will therefore be blocked, and the new absorption edge will occur at
(E
g
+E
c
F
+E
v
F
). The shift in the absorption edge is therefore (E
c
F
+E
v
F
).
The Fermi energy in the conduction band can be calculated from eqn 5.13.
On inserting m

e
= 0.067m
0
, we nd E
c
F
= 0.054 eV. In the case of the
valence band, we must consider the occupancy of both the heavy and light
hole bands. On using the result of Exercise 5.14(ii), we have:
N
h
=
1
3
2
_
2

2
_
3/2
(m
3/2
hh
+m
3/2
lh
) (E
v
F
)
3/2
,
which gives E
v
F
= 0.007 eV for m

hh
= 0.5m
0
and m

lh
= 0.08m
0
. Hence the
absorption edge will shift to higher energy by 0.054 + 0.007 = 0.061 eV.
4
(11.13) If we treat the exciton as a classical oscillator, we can use the results
derived in Chapter 2. If we assume that the contribution of the exciton to
the refractive index is small compared to the non-resonant value, (which
is indeed the case, as we shall show below,) then we expect a refractive
index variation as in Fig. 2.5. The refractive index will thus have local
maxima and minima just below and above the centre of the absorption
line. It is this extra contribution that we are considering in this exercise.
The magnitude of the excitonic contribution can be calculated by following
the method of Example 2.1. In part (i) of Example 2.1 it is shown that

max
=
Ne
2
2n
0
m
0

0
,
where
max
is the extinction coecient at the line centre, while in part (iii)
it is shown that:
n
max
=
_

+
Ne
2
2
0
m
0

0
_
1/2
.
It then follows, with

= n
2
, that:
n
max
= n
_
1 +

max
n
_
1/2
,
where n
max
is the maximum value of the refractive index. If we assume,
as is demonstrated below, that
max
n, we then nd:
n
max
= n +
max
/2 .
Now we know from eqn 1.16 and the data given in the exercise that

max
=

max
4
=
847 10
9
8 10
5
4
= 0.054 .
4
In a real experiment, the behaviour would be more complicated due to many-body eects
such as band gap renormalization. This causes a shift of E
g
N
1/3
, which reduces the blue
shift of the absorption edge.
83
We thus deduce that there is a local maximum in the refractive index
just below the absorption line with (n
max
n) = 0.027. If the exciton
absorption is saturated, this local maximum will disappear. Hence the
maximum change in the refractive index is 0.027.
(11.14) We see from eqn 4.4 that the n = 1 exciton absorption line will occur
at an energy of
h = E
g
R
X
.
For InP we have:
= (1/m

e
+ 1/m

h
)
1
0.06m
0
,
for m

e
= 0.077m
0
and m

h
0.3m
0
(i.e. a mean of m

hh
and m

lh
). Hence
we see from eqn 4.1 that R
X
= (0.06/12.5
2
)R
H
= 5.2 meV. The exciton
energy is thus 1.34 eV at low temperatures. (We consider low temperatures
here because the exciton would be ionized at room temperature.)
The saturation density for excitons is given by the Mott density of eqn 4.8.
We nd from eqn 4.2 that r = 11 nm for the n = 1 exciton, and hence
N
Mott
1.8 10
23
m
3
. The saturation intensity I
s
is the optical in-
tensity required to produce this carrier density. By using the result of
Exercise 5.6(ii), namely:
N =
I
h
,
we nd
I
s
=
h

N
Mott
=
1.34 eV
10
6
10
9
(1.8 10
23
) 4 10
7
Wm
2
.
84

You might also like