BookSol.dvi Steven H. Simon The Oxford Solid State Basics, Solution Manual University Press (2015)

User Manual:

Open the PDF directly: View PDF PDF.
Page Count: 199

DownloadBookSol.dvi Steven H. Simon - The Oxford Solid State Basics, Solution Manual-Oxford University Press (2015)
Open PDF In BrowserView PDF
The Oxford Solid State Basics
Solutions to Exercises
Steven H. Simon
Oxford University

CLARENDON PRESS
2015

.

OXFORD

iii

These are the solutions to exercises from the Book The Oxford Solid
State Basics by Steven H. Simon, published by Oxford University Press,
2013 edition. Please do everyone a favor and do not circulate these
solutions. Do not post these solutions on your website. Do not put them
on Russian websites. Do not copy them and hand them out to students.
While there is no way for me to enforce these reasonable rules, be assured
that I, being a professor at Hogwarts, am in possession of powerful hexes
which I have used to protect the secrecy of these solutions. Those who
attempt to circulate these solutions unlawfully will activate the hex and
will suffer thirty years of bad luck, including spiders crawling into your
underwear.
Some of these solutions have been tested through use in several years
of courses. Other solutions have not been completely tested. Errors
or ambiguities that are discovered in the exercises will be listed on my
web page. If you think you have found errors in the problems or the
solutions please do let me know, and I will make sure to fix them in the
next version. Doing so will undoubtedly improve your Karma. ,
Steven H Simon
Oxford, United Kingdom
January 2014

Contents
1 About Condensed Matter Physics

1

2 Specific Heat of Solids: Boltzmann, Einstein, and Debye

3

3 Electrons in Metals: Drude Theory

15

4 More Electrons in Metals: Sommerfeld (Free Electron)
Theory
21
5 The Periodic Table

35

6 What Holds Solids Together: Chemical Bonding

39

7 Types of Matter

47

8 One-Dimensional Model of Compressibility, Sound, and
Thermal Expansion
49
9 Vibrations of a One-Dimensional Monatomic Chain

55

10 Vibrations of a One-Dimensional Diatomic Chain

71

11 Tight Binding Chain (Interlude and Preview)

81

12 Crystal Structure

95

13 Reciprocal Lattice, Brillouin Zone, Waves in Crystals

99

14 Wave Scattering by Crystals

111

15 Electrons in a Periodic Potential

125

16 Insulator, Semiconductor, or Metal

135

17 Semiconductor Physics

139

18 Semiconductor Devices

149

19 Magnetic Properties of Atoms: Para- and
Dia-Magnetism

159

vi Contents

20 Spontaneous Magnetic Order: Ferro-, Antiferro-, and
Ferri-Magnetism
167
21 Domains and Hysteresis

175

22 Mean Field Theory

179

23 Magnetism from Interactions: The Hubbard Model

191

About Condensed Matter
Physics
There are no exercises for chapter 1.

1

Specific Heat of Solids:
Boltzmann, Einstein, and
Debye
(2.1) Einstein Solid
(a) Classical Einstein (or “Boltzmann”) Solid:
Consider a three dimensional simple harmonic oscillator with mass m and spring constant k (i.e., the mass
is attracted to the origin with the same spring constant
in all three directions). The Hamiltonian is given in the
usual way by
p2
k
H=
+ x2
2m
2
 Calculate the classical partition function
Z
Z
dp
dx e−βH(p,x)
Z=
(2π~)3

Note: in this problem p and x are three dimensional vectors.
 Using the partition function, calculate the heat capacity 3kB .
 Conclude that if you can consider a solid to consist
of N atoms all in harmonic wells, then the heat capac-

(a)
H=
Z=
Since,

Z

Z

∞

−∞

ity should be 3N kB = 3R, in agreement with the law of
Dulong and Petit.
(b) Quantum Einstein Solid:
Now consider the same Hamiltonian quantum mechanically.
 Calculate the quantum partition function
Z=

2

p
π/a

in three dimensions, we get
h
i3
p
p
Z = 1/(2π~) π/(β/2m) π/(βk/2)) = (~ωβ)−3
p
with ω = k/m. From the partition function
U = −(1/Z)∂Z/∂β = 3/β = 3kB T

e−βEj

where the sum over j is a sum over all eigenstates.
 Explain the relationship with Bose statistics.
 Find an expression for the heat capacity.
 Show that the high temperature limit agrees with
the law of Dulong of Petit.
 Sketch the heat capacity as a function of temperature.
(See also exercise 2.7 for more on the same topic)

dx e−βH(p,x)

dy e−ay =

X
j

p2
k
+ x2
2mZ 2

dp
(2π~)3

2

4 Specific Heat of Solids: Boltzmann, Einstein, and Debye

Thus the heat capacity ∂U/∂T is 3kB .
(b) Quantum mechanically, for a 1d harmonic oscillator, we have
eigenenergies
En = ~ω(n + 1/2)
p
with ω = k/m. The partition function is then
X
Z1d =
e−β~ω(n+1/2)
n≥0

= e−β~ω/2 1/(1 − e−β~ω )
= 1/[2 sinh(β~ω/2)]

The expectation of energy is then
hEi =

1

=

0.75

−(1/Z)∂Z/∂β = (~ω/2) coth(β~ω/2)
1
~ω(nB (β~ω) + )
2

where nB is the boson occupation factor

0.5

nB (x) = 1/(ex − 1)

0.25

(hence again the relationship with free bosons). The high temperature
limit gives nB (x) → 1/(x+ x2 /2) = 1/x− 1/2 so that hEi → kB T . More
generally, we obtain

00

1

2

Fig. 2.1 Heat capacity in the Einstein
model (per atom) in one dimension.
Units are kb on vertical axis and kb T /ω
on horizontal. In three dimensions, the
heat capacity per atom is three times
as large.

C = kB (β~ω)2

eβ~ω
− 1)2

(eβ~ω

In 3D,
En1 ,n2 ,n3 = ~ω[(n1 + 1/2) + (n2 + 1/2) + (n3 + 1/2)]
and
Z3d =

X

e−βEn1 ,n2 ,n3 = [Z1d ]3

n1,n2,n3≥0

and correspondingly

1
hEi = 3~ω(nB (β~ω) + )
2
So the high temperature limit is hEi → 3kB T and the heat capacity
C = ∂hEi/∂T = 3kB . More generally we obtain
C = 3kB (β~ω)2
Plotted this looks like Fig. 2.1.

eβ~ω
(eβ~ω − 1)2

5

(2.2) Debye Theory I
(a)‡ State the assumptions of the Debye model of heat
capacity of a solid.
 Derive the Debye heat capacity as a function of
temperature (you will have to leave the final result in
terms of an integral that cannot be done analytically).
 From the final result, obtain the high and low temperature limits of the heat capacity analytically.
You may find the following integral to be useful
Z ∞
∞
∞ Z ∞
X
X
x3
1
π4
dx x
x3 e−nx = 6
=
=
4
e −1
n
15
0
n=1
n=1 0
By integrating by parts this can also be written as
Z ∞
4π 4
x 4 ex
=
dx x
2
(e − 1)
15 .
0

(b) The following table gives the heat capacity C for
potassium iodide as a function of temperature.
T (K)

C(J K−1 mol−1 )

0.1
1.0
5
8
10
15
20

8.5 × 10−7
8.6 × 10−4
.12
.59
1.1
2.8
6.3

 Discuss, with reference to the Debye theory, and
make an estimate of the Debye temperature.

(a) The key assumption of Debye theory is that the dispersion curve
is linear (ω = vk) up to a cut-off frequency ωDebye determined by the
requirement that the total number of vibrational modes is correct.
For a crystal containing N atoms, the low temperature limiting form
is

3
12N kB π 4 T
C=
(2.1)
5
TD
and the high temperature limit is 3N kB . Here, TD = ~ωDebye /kB .
The full derivation goes as follows. For oscillators with frequency ω(k)
a system has a full energy
Z
E = L3 d3 k(2π)3 ~ω(k)[nB (β~ω(k)) + 1/2]
One includes also a factor of 3 out front to account for the three different
sound modes (two transverse and one longitudinal) and we cut off the
integral at some cutoff frequency ωcutof f . We use the assumption that
ω = v|k| although it is not much harder to consider three different
velocities for the three different modes. We thus obtain
Z ωcutof f
dωg(ω)[nB (β~ω) + 1/2]~ω
E =
0

where




9ω 2
12πω 2
g(ω) = N
=
N
(2π)3 nv 3
ωd3

and we have replaced nL3 = N where n is the density of atoms. Here
ωd3 = 6π 2 nv 3 is the Debye frequency, and ~ωd = kB TDebye defines the
Debye temperature. Note that there is no dependence of g(ω) on the
density n (it cancels). This shows that until the cutoff is imposed, there

6 Specific Heat of Solids: Boltzmann, Einstein, and Debye

is actually no knowledge of the underlying lattice — only the overall
volume and sound velocity.
We should choose the cutoff frequency such that we have the right
number of modes in the system, thus we have
Z ωcutof f
dωg(ω)
3N =
0

performing this integral, we find that the proper value of ωcutof f is
exactly the Debye frequency ωd that we just defined.
The general heat Debye theory heat capacity will then be
Z ωd
kB
eβ~ω
2
C = dhEi/dT =
dωg(ω)(~ω)
(kB T )2 0
(eβ~ω − 1)2
Defining x = ~ω/kB T we obtain
3 Z ~ωd /kB T

T
ex
9
dxx4 x
C = dhEi/dT = N kB
TDebye
(e − 1)2
0
This integral is known as the Debye integral. In the low temperature
limit, we can extend the integral out to infinity whereupon it just gives
the constant 4π 4 /15 recovering the above claimed result Eq. 2.1.
In the high temperature limit, the exponents can be expanded such
that the Debye integral becomes
Z ~ωd /kB T
Z ~ωd /kB T
ex
4
=
dxx2 = (1/3)(~ωd /kB T )3
dxx x
2
(e
−
1)
0
0
which then recovers the law of Dulong-Petit C = 3N kB
(b) Given the heat capacity and the temperature, in the low T limit
we should have (from Eq. 2.1)
TD =



12Rπ 4 T 3
5C

1/3

The table of heat capacity looks like
T (K)
C (J K


−1

12Rπ 4 T 3
5C

mol

1/3

−1

)

(K)

0.1

1.0

5

8

10

15

20

8.5 × 10−7

8.6 × 10−4

1.2 × 10−1

5.9 × 10−1

1.1

2.8

6.3

132

131

127

119

121

132

135

So TDebye is about 130K. The fact that the T 3 fit is not perfect is a
reflection of (a) that Debye theory is just an approximation (in particular
that phonons have a nonlinear spectrum!) and (b) that one needs to be
in the low T limit to obtain perfect T 3 scaling. (Note that at low enough
T , the T 3 scaling does indeed work).

7

(2.3) Debye Theory II
Use the Debye approximation to determine the heat
capacity of a two dimensional solid as a function of temperature.
 State your assumptions.
You will need to leave your answer in terms of an integral that one cannot do analytically.

 At high T , show the heat capacity goes to a constant and find that constant.
 At low T , show that Cv = KT n Find n. Find K in
terms of a definite integral.
If you are brave you can try to evaluate the integral,
but you will need to leave your result in terms of the
Riemann zeta function.

In 2d there should be 2N modes. So high T heat capacity should be
C = 2kb N (Law of Dulong-Petit).
Assume longitudinal and transverse sound velocities are equal.
Z |k|=kDebye 2
2
2(πkDebye
d k
=
2N = 2A
(2π)2
(2π)2
0
with A the area. So
kDebye =

√

4πn

with n = N/A the density. So ΘDebye = ~kDebye c with c the sound
velocity.
Since phonons obey bose statistics we have
Z |k|=kDebye 2
d k
ǫ n (βǫk )
E = 2A
2 k B
(2π)
0
Z |k|=kDebye 2
1
d k
~ck β~ck
= 2A
2
(2π)
e
−1
0
Z |k|=kDebye
1
2π
k dk ~ck β~ck
= 2A
(2π)2 0
e
−1
Z ΘDebye
1
A
dǫ ǫ
ǫ
=
π 0
~c ~c eβǫ − 1
Let z = βǫ = ǫ/(kb T ) and we get
E=

A(kb T )3
π~2 c2

Z

ΘDebye /(kb T )

0

z 2 dz
ez − 1

For large T , Θ/T is small so z is small, so
z 2 dz
=z
ez − 1

so we get

so in this limit
E=

Z

ΘDebye /(kb T )

zdz = (ΘDebye /(kb T ))2 /2

0

A(kb T )Θ2Debye
2π~2 c2

=

2
AkDebye
kb T

2π

=A

(4πN/A)kb T
= 2N kb T
2π

8 Specific Heat of Solids: Boltzmann, Einstein, and Debye

which gives
C = dE/dT = 2N kb
as expected.
For small T , the upper limit of the integral goes to infinity and we
have
Z
A(kb T )3 ∞ z 2 dz
E=
π~2 c2 0 ez − 1
So

Cv = KT 2
where
K=

3Akb3
π~2 c2

To evaluate the integral we have
Z

0

∞

z 2 dz
ez − 1

=
=

Z

0

K=

ΘDebye

Neon
Argon
Krypton
Xenon
Radon
Mercury
Potassium
Rubidium
Cesium

75
92
64
64
64
69
91
56
32

K
K
K
K
K
K
K
K
K

Some Low Debye Temperatures

z 2 dz
ez − 1

−z ∞
z 2 dz X −nz
e
e
0
n=0
∞ Z ∞
∞
X
X
dzz 2 e−nz =
2/n3 = 2ζ(3)

Thus we obtain

Material

0

∞

∞

n=1

(2.4) Debye Theory III
Physicists should be good at making educated guesses:
Guess the element with the highest Debye temperature.

Z

n=1

6Akb3 ζ(3)
π~2 c2

The lowest? You might not guess the ones with the absolutely highest or lowest temperatures, but you should be
able to get close.

Largest Debye temperature should be the one with the highest speed
of sound which is probably the hardest element (ie., highest spring constant) and/or smallest mass. Diamond is the obvious guess (and indeed
it does have the highest Debye temperature). ΘDebye = 2230K. The
lowest is harder to guess. One presumably wants a soft material of some
sort – also possibly a heavy material.
Soft and heavy metals like mercury are good guesses. (in fact mercury
is liquid at room temperature and one has to go to low T to measure
a Debye temperature). Also good guesses are Noble gases where the
spring constant is very low (weak interaction between the atoms). Also
heavy soft group 1 metals are good guesses. Many of these are gas or
liquid at room T and a Debye temeperature can only be measured at
low T .

9

(2.5) Debye Theory IV
From Fig. 2.3 (main text) estimate the Debye temper-

ature of diamond. Why does it not quite match the result
listed in Table 2.2 (main text)?

Extracting the slope from the figure gives C/T 3 ≈ 1.9×10−7J/(mol − K4 )
Then using the formula
12N kB π 4
C=
5



T
TD

3

We obtain
TD ≈ 2200K
The reason that it does not match the Debye temperature given in
the figure caption has to do with the comment in the caption. Debye
theory predicts the heat capacity at all possible temperatures. The
Debye temperature quoted in the text is chosen so as to give a good fit
over the full temperature range. The Debye temperature measured here
is chosen to give a good fit at the lowest temperatures (where Debye
theory can actually be exact).

(2.6) Debye Theory V*
In the text we derived the low temperature Debye heat
capacity assuming that the longitudinal and transverse
sound velocities are the same and also that the sound velocity is independent of the direction the sound wave is
propagating.
(a) Suppose the transverse velocity is vt and the lon-

gitudinal velocity is vl . How does this change the Debye
result? State any assumptions you make.
(b) Instead suppose the velocity is anisotropic. For example, suppose in the x̂, ŷ and ẑ direction, the sound velocity is vx , vy and vz respectively. How does this change
the Debye result?

(a) This is actually quite simple. The derivation of the heat capacity
follows the text (or exercise 2.1). The only difference is in the density
of states. In the isotropic calculation we use


12πω 2
g(ω) = N
(2π)3 nv 3
Recall the origin of these factors. Really we had (See Eq. 2.3 of the
main text)
4πω 2
g(ω) = 3L3
(2π)3 v 3
where the factor of 3 out front is for the three polarizations of the sound
waves. One could just as well have written it as


2
1
1
1
3 4πω
+ 3+ 3
g(ω) = L
(2π)3 v 3
v
v

10 Specific Heat of Solids: Boltzmann, Einstein, and Debye

separating out the three different polarizations. Now, if the three polarizations have three different velocities, we have


2
1
1
1
3 4πω
g(ω) = L
+ 3+ 3
(2π)3 v13
v2
v3
this is true since the density of states of the three different excitation
modes simply add. In an isotropic solid, the two transverse mode have
the same velocity vt and the one longitudinal mode has velocity vl and
we would have


4πω 2 2
1
g(ω) = L3
+
(2π)3 vt3
vl3
The remainder of the derivation is unchanged. Thus, defining v̄ such
that
2
1
3
= 3+ 3
3
v̄
vt
vl

we obtain the low temperature capacity in the usual form

3
12N kB π 4 T
C=
5
TD
where now
(kB TD )3 = 6π 2 n~3 v̄.3
Note that the high frequency cutoff is different for the two types of modes
(but the k cutoff is the same for both modes).
(b) If instead we have three different sound velocities in three different
directions, the situation is more complicated (and here we neglect the
differences between longitudinal and transverse modes). Here we must
make some assumption about the sound velocity in some arbitrary direction. A reasonable guess would be as follows. If you consider a sound
wave in direction k̂ (with k̂ = k/|k| a unit vector), we would have
q
v(k̂) = vx2 k̂x2 + vy2 k̂y2 + vz2 k̂z2
Now, following the usual derivation of Debye theory, we start with


Z
L3
1
hEi = 3
dk
dk
dk
~ω(k)
n
(β~ω(k))
+
x
y
z
B
(2π)3
2 .

And now
ω(k) = v(k̂)|k| =

q
vx2 kx2 + vy2 ky2 + vz2 kz2

Since the system is now not isotropic, we cannot do the usual thing
and convert to spherical polar coordinates directly. Instead, we rescale
the axes first writing (with j = x, y, z)
Kj = kj vj
So that
ω(K) = v(k̂)|k| =

q
Kx2 + Ky2 + Kz2

11

and
hEi =

3

L3
3
(2π) vx vy vz

Z

dKx dKy dKz ~ω(K)



1
nB (β~ω(K)) +
2 .

We can now use spherical symmetry to obtain


Z ∞
4πL3
1
2
hEi = 3
ω dω(~ω) nB (β~ω) +
(2π)3 vx vy vz 0
2 .

(2.2)

The rest of the derivation follows as usual to give the usual expression
for heat capacity

3
12N kB π 4 T
C=
5
TD
where now

(kB TD )3 = 6π 2 n~3 vx vy vz .

(2.7) Diatomic Einstein Solid*
Having studied exercise 2.1, consider now a solid made
up of diatomic molecules. We can (very crudely) model
this as a two particles in three dimensions, connected to
each other with a spring, both in the bottom of a harmonic well.
p2 2
k
k
K
p1 2
+
+ x1 2 + x2 2 + (x1 − x2 )2
H=
2m1
2m2
2
2
2
Here k is the spring constant holding both particles in the
bottom of the well, and K is the spring constant holding
the two particles together. Assume that the two particles
are distinguishable atoms.

(For this problem you may find it useful to transform to
relative and center-of-mass coordinates. If you find this
difficult, for simplicity you may assume that m1 = m2 .)
(a) Analogous to exercise 2.1 above, calculate the classical partition function and show that the heat capacity
is again 3kB per particle (i.e., 6kB total).
(b) Analogous to exercise 2.1 above, calculate the quantum partition function and find an expression for the heat
capacity. Sketch the heat capacity as a function of temperature if K ≫ k.
(c)** How does the result change if the atoms are indistinguishable?

(a) We can write the partition function as
Z
Z
dp1 dp2
Z=
dx1 dx2 e−βH
(2π~)3 (2π~)3
Considering the momentum integrals first, we have
Z

dpe−βp

2

/(2m)

=



2πm
β

3/2

Then the spatial integrals are made simple by transforming
Y

=

y

=

x1 − x2

(x1 + x2 )/2

So the spatial integrals are
Z

dYdye

−β (−ky2 −(k/4+K/2)Y 2 )

=



π
kβ

3/2 

π
β(k/4 + K/2)

3/2

12 Specific Heat of Solids: Boltzmann, Einstein, and Debye

Putting these together we get a partition function
 3/2 
3/2 
3/2 
3/2
π
π
2πm1
2πm2
Z=
∼ β −6
kβ
β(k/4 + K/2)
β
β
The energy is
hEi = −∂ ln Z/∂β = 6/β
So the heat capacity for the two particles is
C = ∂hEi/∂T = 6kB
(b) The case where the two masses are identical is fairly simple. Again,
we construct
Y
y

=
=

x1 − x2
(x1 + x2 )/2

Q

=

(p1 − p2 )/2

and correspondingly

q =

(p1 + p2 )

Note that these two variables are constructed so that [Qj , Yk ] = i~δjk
and [qj , yk ] = i~δjk and all other commutators are zero (in other words,
these are canonical conjugates). The Hamiltonian is now written as


q2
k K
Q2
2
y2
+
+ kY +
+
H=
2(m/2) 2(2m)
4
2
which comprises two independent three-dimensional harmonic oscillators
with frequencies
p
ω1 =
4k/m
p
ω2 =
(k/2 + K)/(2m)

The heat capacity is then (analogous to 2,1)
C = 3kB (β~ω1 )2

eβ~ω2
eβ~ω1
2
+
3k
(β~ω
)
B
2
(eβ~ω1 − 1)2
(eβ~ω2 − 1)2

(2.3)

The case of unequal masses is more tricky. The general method is
similar to the discussion outlined in problem *** below. First, rescale
√
√
ri = pi mi and xi = zi / mi so that the Hamiltonian reads
H=

√
√
k
K
k
r1 2 r2 2
2
z1 2 +
z2 2 + (z1 / m1 − z2 / m2 ) (2.4)
+
+
2
2
2m1
2m2
2

Note that r and z are canonically conjugate just like p and x. The
potential can be viewed as a matrix which we can write as


√
(k + K)/m1 −K/ m1 m2
√
−K/ m1 m2 (k + K)/m2

13

We can diagonalize this matrix to define two new decoupled degrees of
freedom representing two independent harmonic oscillators. The frequencies of these oscillaltors squared are the eigenvalues of the above
Ck_B
matrix.
p
5
(k + K)(m1 + m2 ) + (k + K)2 (m1 − m2 )2 + 4K 2 m1 m2
ω1 =
4
2m1 m2
p
3
(k + K)(m1 + m2 ) − (k + K)2 (m1 − m2 )2 + 4K 2 m1 m2
ω2 =
2
2m1 m2
And the heat capacity is given by formula 2.3 using these two oscillator
frequencies. A plot is given in Fig. 2.2.
(d) If the two atoms are indistinguishable then they must obey either
Bose or Fermi statistics depending on the atom type. The center of
mass degree of freedom (y above) has the same Einstein heat capacity as
calculated before. However, the relative degree of freedom does not. Due
to the statistics, the relative wavefunction must obey Ψ(Y) = ±Ψ(−Y)
with the ± depending on whether we have bosons or fermions. Since
the three dimesional harmonic motion wavefunction can be decomposed
into three one-dimensional wavefunctions Ψ(r) = ψnx (x)ψny (y)ψnz (z),
and ψn is symmetric or antisymmetric depending on whether n is even
or odd, we must have that
nx + ny + nz = even for bosons, odd for fermions
So when we write the partition function for this oscillator, instead of
X
e−β~ω(nx +ny +nz +3/2)
nx ,ny ,nz ≥0

as usual, we instead only include the terms of sum respecting the even/odd
symmetry. This restriction can be handled by writing
X
X
1
(1 ± (−1)nx +ny +nz )
→
2
nx +ny +nz =even/odd

nx ,ny ,nz ≥0

The sum can then be evaluated to give a partition function
"
2 
3 #
3
1
1
Zbose/fermi = e− 2 β~ω 2
±
1 − e−β~ω
1 + e−β~ω
which then can be differentiated to get the heat capacity. I obtained

24e2ω/kb T 1 + 2e2ω/kb T + 5e4ω/kb T
C = kb
2
1 + 2e2ω/kb T − 3e4ω/kb T (kb T )2
for the fermi case and


24e2ω/kb T 5 + 2e2ω/kb T + e4ω/kb T
C = kb
2
−3 + 2e2ω/kb T + e4/kb T (kb T )2

for the bose case. Note that both of these have the correct Dulong-Petit
high temperature limit of 3kb .

1
2

4

6

8

10

T

Fig. 2.2 Heat capacity of two einstein
oscillators. Here ω1 = 1 and ω2 = 10

14 Specific Heat of Solids: Boltzmann, Einstein, and Debye

(2.8) Einstein Versus Debye*
In both the Einstein model and the Debye model the
high temperature heat capacity is of the form
C = N kB (1 − κ/T 2 + . . .)
 For the Einstein model calculate κ in terms of the
Einstein temperature.

 For the Debye model calculate κ in terms of the
Debye temperature.
From your results give an approximate ratio
TEinstein /TDebye . Compare your result to the values for
silver given in Fig. 2.4 (main text). (The ratio you calculate should be close to the ratio stated in the caption
of the Figure. It is not exactly the same though. Why
might it not be?).

Expanding the heat capacity of a single Einstein oscillator
C

=
∼
∼
∼
∼

∼

eβ~ω
(eβ~ω − 1)2
1 + β~ω + (β~ω)2 /2
kb (β~ω 2 )
[β~ω + (β~ω)2 /2 + (β~ω)3 /6]2
1 + β~ω + (β~ω)2 /2
kb
[1 + (β~ω)/2 + (β~ω)2 /6]2
1 + β~ω + (β~ω)2 /2
kb
[1 + (β~ω)/2 + (7/12)(β~ω)2 ]

kb 1 + β~ω + (β~ω)2 /2 [1 − (β~ω)/2 + (5/12)(β~ω)2 ]

kb 1 − (β~ω)2 /12 + . . .
kb (β~ω)2

2
So κ = TEinstein
/12.
We can handle the Debye case by realizing that the heat capacity is
just an integration over Einstein oscillators. So
Z ωD
eβ~ω
dωg(ω)kb (β~ω)2 β~ω
C =
(e
− 1)2
0
Z ωD

dωg(ω) 1 − (β~ω)2 /12 + . . .
∼ kb
0

3
where g(ω) = N 9ω 2 /ωD
. Note that the integration is cut off so that the
integral over 1 gives precisely 3N kb as it should. Thus we obtain

C ∼ 3kB N 1 − (β~ωD )2 /20 + . . .

2
2
2
So κ = TDebye
/20. Setting TEinstein
/12 = TDebye
/20 we would predict

TDebye =

p
5/3 TEinstein ≈ 1.29TEinstein

In the data from Fig 2.4 the ratio is 215/151 ≈ 1.42. The reason this
does not match perfectly with our prediction is mainly because TDebye
and TEinstein are likely fit over the full range of the heat capacities
measured, not just in the high temperature limit. If they were fit parameters for only the high temperature limit, the ratio would come out
as we predicted here.

Electrons in Metals: Drude
Theory
(3.1) Drude Theory of Transport in Metals
(a)‡ Assume a scattering time τ and use Drude theory
to derive an expression for the conductivity of a metal.
(b) Define the resistivity matrix ρ as E = ρj. Use
e ρ for
Drude theory to derive an expression efor the matrix
e
a metal in a magnetic field. (You may assume B parallel
to the ẑ axis. The under-tilde means that the quantity ρ
e
is a matrix.) Invert this matrix to obtain an expression
for the conductivity matrix σ.
e
(c) Define the Hall coefficient.
 Estimate the magnitude of the Hall voltage for a
specimen of sodium in the form of a rod of rectangular
cross-section 5mm by 5mm carrying a current of 1A down
its long axis in a magnetic field of 1T perpendicular to
the long axis. The density of sodium atoms is roughly
1 gram/cm3 , and sodium has atomic mass of roughly 23.
You may assume that there is one free electron per sodium
atom (sodium has valence 1).
 What practical difficulties would there be in measuring the Hall voltage and resistivity of such a specimen.
How might these difficulties be addressed).

(d) What properties of metals does Drude theory not
explain well?
(e)* Consider now an applied AC field E ∼ eiωt which
induces an AC current j ∼ eiωt . Modify the above calculation (in the presence of a magnetic field) to obtain an
expression for the complex AC conductivity matrix σ (ω).
e
For simplicity in this case you may assume that the metal
is very clean, meaning that τ → ∞, and you may assume
that E ⊥ B. You might again find it convenient to assume B parallel to the ẑ axis. (This exercise might look
hard, but if you think about it for a bit, it isn’t really
much harder than what you did above!)
 At what frequency is there a divergence in the conductivity? What does this divergence mean? (When τ is
finite, the divergence is cut off.)
 Explain how could one use this divergence (known
as the cyclotron resonance) to measure the mass of the
electron. (In fact, in real metals, the measured mass of
the electron is generally not equal to the well-known value
me = 9.1095 × 10−31 kg. This is a result of band structure
in metals, which we will explain in Part VI.)

(a) We consider an electron with momentum p at time t and we ask
what momentum it will have at time t + dt. There is a probability dt/τ
that it will scatter to momentum zero. If it does not scatter to momentum zero (with probability 1 − dt/τ ) it simply accelerates as dictated by
its usual equations of motion dp/dt = F Thus


dt
(p(t) + Fdt)
hp(t + dt)i = 1 −
τ

or

dp
p
= F−
dt
τ
where here the force F on the electron is just the Lorentz force
F = −e(E + v × B)

3

(3.1)

16 Drude Theory

In absence of magnetic field
p
dp
= −eE −
dt
τ
In steady state, dp/dt = 0 so we have
mv = p = −eτ E
with m the mass of the electron and v its velocity. If there is a density
n of electrons in the metal, and they are all moving at velocity v then
the electrical current is given by
j = −env =

e2 τ n
E
m

or in other words, the conductivity of the metal is
σ=

e2 τ n
m

(3.2)

(b) In both an electric and a magnetic field
dp
= −e(E + v × B) − p/τ
dt
Again setting this to zero in steady state, and using p = mv and j =
−nev, we obtain an equation for the steady state current
0 = −eE +
or
E=



m
j×B
+
j
n
neτ


1
m
j×B+ 2 j
ne
ne τ

We now define the 3 by 3 resistivity matrix ρ which relates the current
e
vector to the electric field vector
E = ρj
e
We then obtain components of this matrix
ρxx = ρyy = ρzz =

m
ne2 τ

and if we imagine B oriented in the ẑ direction, then
ρxy = −ρyx =

B
ne

Inverting this equation we obtain a conductivity matrix
σzz

=

ne2 τ /m

σxx = σyy

=

ρxx /(ρ2xx + ρ2xy ) = σzz /[1 + (eBτ /m)2 ]

σyx = −σxy

=

ρxy /(ρ2xx + ρ2xy ) = σzz (eBτ /m)/[1 + (eBτ /m)2 ]

17

with all other entries in the σ matrix being zero.
(c) The Hall coefficient is RH = ρyx /B which is −1/ne in Drude
theory. If sodium n= 1 gm /cm3 with atomic mass M = 23, this is then
a density of atoms of
n = NA × n/M = 2.6 × 1028 m−3
hence the same density of electrons assuming one free electron per atom.
The cross section of the rod is L by L with L = 5mm, so the current
density is j = I/L2 . The Hall resistivity is ρxy = B/(ne) so the Hall
voltage is jρxy L. So the total Hall voltage is
V =

IB
= 4.8 × 10−8 Volts
Lne

Some of the problems with making this measurement might be:
• This is a very small voltage: One needs a sensitive voltmeter
• There may be contact resistance: Use a high impedance voltmeter
• Contacts may not be perfectly aligned: Try varying (reversing) the
magnetic field to pick out only the B dependent part (I.e., measure
(V (B) − V (−B))/2).
• Could have heating

Tutors might also use this problem as an opportunity to discuss how
useful lock-in amplifiers are (which most students do not appear to
know).
(d) Drude theory fails to explain why the electrons do not carry heat
capacity of 3/2kB per electron as a classical gas would. This results in
incorrect predictions of, ex, thermoelectric coefficients. Drude theory
also fails to explain why the sign of the Hall effect can be different in
different samples. Drude theory does not explain why we should only
count valence electrons.
(e) Drude theory at finite frequency. We start with the equation of
motion simplified by setting τ → ∞
dp
= −e(E + v × B)
dt
setting E = x̂Ex0 eiωt and B = B ẑ,
mv̇x

=

mv̇y

=

−eEx0 eiωt − evy B
evx B

We can differentiate the first equation to give
mv̈x = −iωeEx0 eiωt − ev̇y B
then plug in the second equation to give
v̈x = −iωe(Ex0 /m)eiωt − (eB/m)2 vx

18 Drude Theory

which is the equation of a driven harmonic oscillator. We use the ansatz
solution vx = vx0 eiωt so we obtain
−ω 2 vx0 = −iω(eEx0 /m) − (eB/m)2 vx0
which we solve
vx0 =
and similarly
vy0 =

−iω(eEx /m)
(eB/m)2 − ω 2

−(eB/m)(eEx /m)
(eB/m)2 − ω 2

with the current being j = −env we obtain
σxx
σyx

= iω(ne2 /m)/[(eB/m)2 − ω 2 ]

= (eB/m)(ne2 /m)/[(eB/m)2 − ω 2 ]

The cyclotron frequency eB/m is the natural oscillation frequency of a
particle of charge −e of mass m in magnetic field B. This divergent
response is easy to detect experimentally as a strong absorbtion of the
ac electric field at a particular frequency. (Then this obviously can be
converted into a measurement of the mass).
The motion in the z-direction is unaffected by the magnetic field in
the z direction so that we have
σzz = ne2 /(iωm)
and off-diagonal terms including z are zero.
Note, the calculation may look a bit nicer if you set v or equivalently
j, and solve for E to obtain the finite frequency resistance matrix, and
then invert last. Lets try doing it that way also. Starting with
dp
= −e(E + v × B)
dt
Writing E = E0 eiωt and j = j0 eiωt and also j = n(−e)v = n(−e)p/m
we then have
iωmj0 /(n(−e)) = −e[E0 + j0 × B/(n(−e))]
or
E0 =



iωm
ne2



j0 − B × j0 /(n(−e))

So assuming B in the ẑ direction, we have a resistivity matrix

 iωm
B/(−ne)
0
ne2
iωm
ρ =  B/(ne)
0 
ne2
iωm
e
0
0
ne2
which we invert to get the same result as above.

19

(3.2) Scattering Times
The following table gives electrical resistivities ρ, densities n, and atomic weights w for the metals silver and
lithium:

Ag
Li

ρ (Ωm)

n (g/cm3 )

w

1.59 × 108
9.28 × 108

10.5
0.53

107.8
6.94

 Given that both Ag and Li are monovalent (i.e.,
have one free electron per atom), calculate the Drude
scattering times for electrons in these two metals.

In the kinetic theory of gas, one can estimate the scattering time using the equation
τ =

where n is the gas density, hvi is the average velocity (see
Eq. 3.4 main text), and σ is the cross-section of the gas
molecule—which is roughly πd2 with d the molecule diameter. For a nitrogen molecule at room temperature,
we can use d = .37nm.
 Calculate the scattering time for nitrogen gas at
room temperature and compare your result to the Drude
scattering times for electrons in Ag and Li metals.

Note: the table should read 10−8 not 108 !
We use σ = ρ−1 = N e2 τ /m with m the free electron mass and where
N here is the electron density which we calculate by
N =n

Avagadro Number
106
mol-weight in grams/cm3

Solving for τ we get
τAg

=

τLi

=

3.8 × 10−14 sec

8.3 × 10−15 sec

The second part should say room temperature and pressure. The
weight of a Nitrogen molecule is about 28 times that of a proton (two
nitrogen atoms of atomic weight 14). So the velocity at 300 K is
s
8kB T
≈ 475m/sec
hvi =
π28mp
uncoincidentally being close to the speed of sound in air. The density
is given by n = P/RT with R the gas constant. At P = 105 pascals
and T = 300 K, this gives .025 mol/m3 . (This should be the usual 22.4
moles per liter that people remember, but we used 300 K instead of 273
and we approximated the pressure). Multiplying by Avagadro’s number
give the density that we should use in the equation
τ=

1
nhviσ

1
≈ 2 × 10−10 sec
nhviσ

So electrons scatter much much much more often — this is not surprising
considering how much higher their density is than that of the nitrogen
gas.

20 Drude Theory

(3.3) Ionic Conduction and Two Carrier Types
In certain materials, particularly at higher temperature, positive ions can move throughout the sample in
response to applied electric fields, resulting in what is
known as ionic conduction. Since this conduction is typically poor, it is mainly observable in materials where
there are no free electrons that would transport current. However, occasionally it can occur that a material
has both electrical conduction and ionic conduction of

roughly the same magnitude—such materials are known
as mixed ion–electron conductors.
Suppose free electrons have density ne and scattering
time τe (and have the usual electron mass me and charge
−e). Suppose that the free ions have density ni , scattering time τi , mass mi and charge +e. Using Drude theory,
(a) Calculate the electrical resistivity.
(b) Calculate the thermal conductivity.
(c)* Calculate the Hall resistivity.

If we fix the electric field, both species respond to the electric field independently. So the total conductivity is the sum of the two independent
conductivities


ne τe
ni τi
2
σ = σe + σi = e
+
me
mi
And thus

ρ=

e2



1
ne τe
me

+

ni τi
mi



The thermal conductivity is similar – both pieces add


2
ni τi
4kB
T ne τe
+
κ = κe + κi =
π
me
mi
Note that the Weidemann-Franz law continues to hold here in the ratio
of σ to κ.
The Hall resistivity is more complicated. To simplify, if we apply
magentic field in the z direction, we need only keep track of conductivity
in the x, y plane (i.e., we can think of this as a two dimensional problem).
For a single species, we have (See exercise 3.1)


r
BR
ρ=
−BR
r
where r = m/(nq 2 τ ) and R = q/n with q the charge on the charge
carrier. We define tensors ρe and ρi for the two separate species in
terms of rj = mj /(nj qj2 τj ) and Rj = qj /nj with j = e or i. The
conductivity tensors are σj = ρ−1
j and then the total conductivity tensor
is σ = σe + σi . Finally this is inverted to give the tensor ρtotal = σ −1 .
There is a lot of algebra involved in this. I obtained
ρxx

=

ρxy

=

B 2 (re Ri2 + ri Re2 ) + ri re (re + ri )
B 2 (Re + Ri )2 + (re + ri )2

B B 2 Re Ri (Re + Ri ) + Ri re2 + Re ri2
B 2 (Re + Ri )2 + (re + ri )2

More Electrons in Metals:
Sommerfeld (Free
Electron) Theory

(4.1) Fermi Surface in the Free Electron (Sommerfeld) Theory of Metals
(a)‡ Explain what is meant by the Fermi energy, Fermi
temperature and the Fermi surface of a metal.
(b)‡ Obtain an expression for the Fermi wavevector and
the Fermi energy for a gas of electrons (in 3D).
 Show that the density of states at the Fermi sur-

face, dN/dEF can be written as 3N/2EF .
(c) Estimate the value of EF for sodium [The density
of sodium atoms is roughly 1 gram/cm3 , and sodium has
atomic mass of roughly 23. You may assume that there
is one free electron per sodium atom (sodium has valence
one)]
(d) Now consider a two-dimensional Fermi gas. Obtain an expression for the density of states at the Fermi
surface.

(a.i) Fermi Energy EF is chemical potential at T = 0. Note, if there is
a filled band the chemical potential is mid-gap, and this differs from the
conventional intuition that it is the highest filled state at zero temperature. Note that some books define fermi energy to be chemical potential
as a function of temperature. This is annoying — why define a new
quantity if it is just another name for the old quantity?!
(a.ii) Fermi temperature TF = EF /kb with kb being Boltzmann’s constant.
(a.iii) Fermi surface is the surface in momentum space separating the
filled and unfilled states at zero temperature. (This is ill-defined for the
case of a filled band – but we don’t do band theory until later in the
course). Or the manifold of states having energy EF . Note that it need
not be a sphere, for example, if the effective mass (defined later!) is
anisotropic you get an ellipsoid instead.
(b)
Z
2V 4πkF3
dk
=
N = 2V
3
(2π)3 3
k 2
n(E) =

1/2 − (E − µ)/(4kb T ) otherwise

0123

In this approximation, one can calculate that the energy is given by
E(T ) = Constant + D

Z

124536789

2/kb T

−2/kb T

dEE(1/2 − E/(4kb T ))

= Constant + D(kb T )2 (4/3)
which results in C = (8/3)kb (kb T )D(EF ).
An exact calculation (See Ashcroft+Mermin, or exercise 4.9.b below)
of this result is
C = (π 2 /3)kb (kb T )D(EF ) = N (π 2 /2)kb (kb T )/EF

(4.2)

This calculation is an exercise given as an additional problem.
Note that as discussed above, for a free fermi gas D(EF ) = 3N/2EF .
Thus up to constants of order one we have
C ∼ kb N (kb T /EF )
which is very small since T ≪ TF .
(b) There are several ways that the electrons can respond to the magnetic field. First, we assume that the magnetic field couples only to
the spins of the electron (we ignore orbital effects). The Hamiltonian
(neglecting the Lorentz force of the magnetic field) becomes
H=

p2
+ gµB B · σ
2m

where g = 2 is the g-factor of the electron B is the externally applied
magnetic field and σ is the spin of the electron which takes eigenvalues
±1/2. Here µB ≈ .67(K/T )/kB the conventional Bohr magneton. Thus
in the magnetic field the energy of an electron with spin up or down

Fig. 4.1 The Fermi function (green)
and a simple approximation to the
fermi function (blue).

26 Sommerfeld Theory

(with up meaning it points the same way as the applied field)
ǫ(k, ↑) =
ǫ(k, ↓) =

~2 |k|2
+ µb |B|
2m
~2 |k|2
− µb |B|
2m

The spin magnetization of the system in the direction of the applied
magnetic field will then be
M =−

1 dE
= −([# up spins] − [# down spins]) µB /V
V dB

(4.3)

So when the magnetic field is applied, it is lower energy for the spins to
be pointing down, so more of them will point down.
Let us now calculate the Pauli paramagnetism of the free electron
gas at T = 0. With zero magnetic field applied, both the spin up and
spin down states are filled up to the Fermi energy (i.e, to the Fermi
wavevector). Near the Fermi level the density of states per unit volume
for spin up electrons is g(EF )/2 and similarly the density of states per
unit volume for spin down electrons is g(EF )/2. When B is applied, the
spin ups will be more costly by an energy µB B. Thus, (assuming that the
chemical potential does not change) we will have (g(EF )/2)µB B fewer
spin ups electrons per unit volume. Similarly, the spin downs will be less
costly by the same amount, so we will have (g(EF )/2)µB B more spin
downs per unit volume. Note that the total number of electrons in the
system did not change, so our assumption that the chemical potential did
not change is correct. (Recall that chemical potential is always adjusted
so it gives the right total number of electrons in the system). Thus,
using Eq. 4.3 the moment per unit volume is given by
M = g(EF )µ2B B
and hence the magnetic susceptibility χ = ∂M/∂H is given (at T = 0
by)
χP auli = µ0 µ2B g(EF )
(c) For a classical monatomic gas, the specific heat is given by the
equipartition law C = 3kB N which is larger than the result above by
roughly a factor of EF /kb T which could be a factor of 100 or more.
Similarly, for a single isolated spin 1/2 we can calculate the partition
function (this calculation was done in stat mech class last year).
Z = eβgµB B/2 + e−βgµB B/2
The expectation of the moment (per spin) is then
m = −d log Z/d(Bβ) = (gµB /2) tanh(βgµB B/2)
For small B this is
m = (gµB /2)2 (B/kb T )

27

thus the total susceptibility for N spins (recall susceptibility is measured
per unit volume) is
dM/dH = µ0 dM/dB = N µ0 (gµB /2)2 (1/kb T )
which, for any T ≪ TF is much larger than the Pauli susceptibility
calculated above (by a factor of approximately TF /T ).
Other properties that differ from the classical prediction include: Thermopower, Peltier Coefficinent, Average Electron Velocity, Compressibility, ...
(d) The T 3 term is clearly from Debye phonon specific heat. The T linear term is the specific heat of free electrons. Using the above formula
Eq. 4.2 yields
EF = (π 2 /2)R(T /Clinear ) =≈ 2 × 104 KkB ≈ 1.7eV
The real value is roughly 2.1 eV.
Another method would be to use the density of potassium and assume
the valence is 1. If you do this, you get something much closer to the
right answer.

(4.4) Another Review of Free Electron Theory
 What is the free electron model of a metal.
 Define Fermi energy and Fermi temperature.
 Why do metals held at room temperature feel cold
to the touch even though their Fermi temperatures are
much higher than room temperature?
(a) A d-dimensional sample with volume Ld contains
N electrons and can be described as a free electron model.
Show that the Fermi energy is given by
EF =

~2
(N ad )2/d
2mL2

Find the numerical values of ad for d = 1, 2, and 3.
(b) Show also that the density of states at the Fermi

energy is given by
g(EF ) =

Nd
2Ld EF

 Assuming the free electron model is applicable, estimate the Fermi energy and Fermi temperature of a
one-dimensional organic conductor which has unit cell of
length 0.8 nm, where each unit cell contributes one mobile
electron.
(c) Consider relativistic electrons where E = c|p|. Calculate the Fermi energy as a function of the density for
electrons in d = 1, 2, 3 and calculate the density of states
at the Fermi energy in each case.

The free electron model of a metal describes electrons in a metal as a
noninteracting gas of fermions at some fixed density (usually chosen to
be v electrons per unit cell of the metal where v is the valence).
Fermi Energy EF is chemical potential at T = 0. Note, if there is a
filled band the chemical potential is mid-gap, and this differs from the
conventional intuition that it is the highest filled state at zero temperature. Note that some books define fermi energy to be chemical potential
as a function of temperature. This is annoying — why define a new
quantity if it is just another name for the old quantity?!
Fermi temperature TF = EF /kb with kb being Boltzmann’s constant.
Due to Pauli exclusion, a metal can have a very high Fermi tempera-

28 Sommerfeld Theory

ture (high chemical potential) even if the material is at zero temperature
— i.e., if the material is in its ground state. When you touch a material
and it feels hot, this is because heat has flowed from the material to you.
If the material is in its ground state (despite having a high fermi temperature) it cannot lower its own energy and therefore cannot transfer heat
to you. Note: Having a high chemical potential DOES mean that the
material might have a tendency to transfer electrons to another body
with fewer electrons (although this might create a charge imbalance that
then prevents further flow of electrons). We discuss physics of this type
in Chapter 18.
(a) In any number of dimensions we can write
Z
dd k
d
N = 2L
d
|k| 1 term at
all!
(b) In the more general case where g(E) is not a constant, one must use
the so-called Sommerfeld expansion. This is quite a bit more complex.
First we will quote the key formula and use it to derive our result for
the heat capacity, then we will go back to derive Sommerfeld’s formula.
Defining the Fermi function
nF (ǫ) =

1
eβ(ǫ−µ)

+1

The Sommerfeld formula is
Z ∞
Z µ
π
H(ǫ)nF (ǫ)dǫ =
H(ǫ)dǫ + (kb T )2 H ′ (µ) + O(T 4 )
6
∞
∞

(4.4)

The intuition behind this formula is that at low temperature the fermi
function is a step function (given by the first term). The finite slope of
the fermi function where it is almost a step creates the small correction
term which is determined by H ′ . Note that as we determinmed in 4.6
and above, if the argument H is constant, then up to exponentially small
terms, there is no correction term at any order in T . One could carry
this expansion to higher order and pick up terms related to H ′′ etc as
well.
Expecting that the chemical potential will remain very close to the
Fermi energy at low temperatures, we can conclude
Z ∞
Z EF
π
H(ǫ)dǫ + (µ − EF )H(EF ) + (kb T )2 H ′ (EF ) + . . .
H(ǫ)nF (ǫ)dǫ =
6
∞
∞
Using this equation to write an expression for the density
Z ∞
N
g(ǫ)nF (ǫ)dǫ
(4.5)
=
V
0
Z EF
π
dǫg(ǫ) + (µ − EF )g(EF ) + (kb T )2 g ′ (µ) + . . .
=
6
∞
n
o
π
N (T = 0)
+ (µ − EF )g(EF ) + (kb T )2 g ′ (EF )
(4.6)
=
V
6

33

Since we are fixing the density N , the term in brackets must remain
zero for all temperatures (accurate to order T 2 ). This confirms that
µ = EF + O(T 2 ) and confirms our suspicion that it should be very close
to EF at low temperature. Note also that for g ′ = 0, such as in 2d,
µ = EF independent of temperature.
We can then write a similar expression for the energy density
Z ∞
E
=
ǫg(ǫ)nF (ǫ)dǫ
V
0
Z EF
π
ǫg(ǫ)dǫ + EF (µ − EF )g(EF ) + (kb T )2 (ǫg(ǫ))′ǫ=EF + . . .
=
6
∞
n
o
π
E(T = 0)
+ EF (µ − EF )g(EF ) + (kb T )2 g ′ (EF )
=
V
6
π2
2
+ (kb T ) g(EF ) + . . .
(4.7)
6
Note that in the final equation the term in brackets is the same as the
term in brackets from Eq. 4.6 which we have set to zero in order to keep
the density constant as a function of temperature. Thus we obtain
E
E(T = 0) π 2
=
+
(kb T )2 g(EF ) + . . .
V
V
6
Which we differentiate to obtain the heat capacity
C/V =

π2
kb (kb T )g(EF )
3

as claimed.
Finally we return to prove the Sommerfeld formula 4.4. The quantity
we would like to evaluate is
Z ∞
I = H(ǫ)nF (ǫ)dǫ
∞

Let us define a function
K(ǫ) =

Z

ǫ

dǫ′ H(ǫ′ )

−∞

So that, integrating by parts, we have
Z ∞
−∂nF (ǫ)
I=
K(ǫ)
dǫ
∂ǫ
∞
there are no boundary terms (to accuracy of O(e−βµ ) because ∂nF /∂ǫ
decays very rapidly away from the chemical potential. Note now that
the function ∂nF /∂ǫ is a symmetric (even) function around the chemical
potential. Thus let if expand K in a taylor series around the chemical
potential only the even terms will have a nonzero contribution, thus we
have

Z ∞
−∂nF (ǫ)
1
2 ′′
dǫ
(4.8)
I=
K(µ) + (ǫ − µ) K (µ) + . . .
2
∂ǫ
∞

34 Sommerfeld Theory

Z

∞

∞

−∂nF (ǫ)
=1
∂ǫ

the first term in the expansion of Eq. 4.8 is just
Z µ
K(µ) =
H(ǫ)dǫ
−∞

as required. The next term gives a prefactor of K ′′ (µ) = H ′ (µ) and
requires that we evaluate the integral
Z
Z ∞
1 ∞ x2 eβx
1
2 −∂nF (ǫ)
= (kb T )2 π 2 /6
(ǫ − µ)
dǫ =
∂ǫ
2 ∞ (eβx + 1)2
∞ 2
using the given integral. Thus we have obtained the first two terms of
Eq. 4.8. By followin a similar procedure one can evaluate higher terms
in the expansion and in particular we will find that the next term in the
expansion must be proportional to T 4 .

5

The Periodic Table

(5.1) Madelung’s Rule
 Use Madelung’s rule to deduce the atomic shell filling configuration of the element tungsten (symbol W)
which has atomic number 74.
 Element 118 has recently been discovered, and is

expected to be a noble gas, i.e., is in group VIII. (No
real chemistry tests have been performed on the element yet, as the nucleus decays very quickly.) Assuming
that Madelung’s rule continues to hold, what should the
atomic number be for the next noble gas after this one?

Angular momentum l orbitals (l = 0 is called s, l = 1 is called p, etc)
contain up to 2(2l + 1) electrons. Madulung’s rule fills orbitals according
to the diagram Fig. 5.1: from lowest n + l to highest, and for cases of
the same n + l, fill the lower n first. So we have
Tungsten atomic number 74:
1s2 2s2 2p6 3s2 3p6 4s2 3d1 0 4p6 5s2 4d10 5p6 6s2 4f14 5d4 .
Note that the “exponents” add to 74. Or equivalently we write
[Xe] 4f14 5d4 .
Note that the noble gases occur whenever a p-shell has just filled.
Element 118 has a filled 7p shell. Madelung’s rule tells us that we then
have to fill 8s, 5g, 6f, 7d and finally 8p. This brings us to element
168. For entertainment sake (and you can try to prove this) note that
the sequence of nobel gas element numbers 2, 10, 18, 36, 54, 86, 118,
168 has successive differences which are twice the perfect squares each
occurring twice. (10-2)/2 = 4 , (18-10)/2 = 4, (36-18)/2 = 9, (54-36)/2
= 9, (86-54)/2 = 16, (118-86)/2=16, (168-118)/2 = 25, and so forth.

(5.2) Effective Nuclear Charge and Ionization
Energy
(a) Let us approximate an electron in the nth shell
(i.e., principal quantum number n) of an atom as being
like an electron in the nth shell of a hydrogen atom with
an effective nuclear charge Z. Use your knowledge of
the hydrogen atom to calculate the ionization energy of
this electron (i.e., the energy required to pull the electron
away from the atom) as a function of Z and n.
(b) Consider the two approximations discussed in the
text for estimating the effective nuclear charge:

01
21
31
41
51
61


27
37
47
57
67







38
48 49
58 59 5
68 69 6 6

Fig. 5.1 Ordering of filling orbitals in
atoms (Madelung’s rule).

• (Approximation a)
Z = Znuc − Ninside
• (Approximation b)
Z = Znuc − Ninside − (Nsame − 1)/2
where Znuc is the actual nuclear charge (or atomic number), Ninside is the number of electrons in shells inside of n
(i.e., electrons with principal quantum numbers n′ < n),
and Nsame is the total number of electrons in the nth
principal shell (including the electron we are trying to
remove from the atom, hence the −1).

36 The Periodic Table
actual ionization energies (you will have to look these up
on a table).
Your results should be qualitatively quite good. If you
try this for higher atomic numbers, the simple approximations begin to break down. Why is this?

 Explain the reasoning behind these two approximations.
 Use these approximations to calculate the ionization energies for the atoms with atomic number 1 through
21. Make a plot of your results and compare them to the

Neglecting fine structure, the energy of an electron in the nth shell of
hydrogen is
−Ry
En =
n2
where Ry=13.6 eV is the Rydberg constant. For a hydrogenic atom
with nuclear charge Z, the Coulomb interaction is Z times as strong
as in hydrogen, resulting in binding energy which is Z 2 as strong. To
see this in detail, one can solve the Schroedinger equation in detail.
However, without doing this one can get it by a scaling argument as
well. The bound state is a balancing of the kinetic with the potential
energy. So roughly one should be able to estimate the binding energy
by setting these equal to each other. Setting the the length scale to a,
(i.e, define a to be the effective Bohr radius) we have
KE =

Ze2
~2
= PE =
2
ma
4πǫ0 a

solving for a obtains a ∼ 1/Z and plugging back into P E or KE we
determine that the kinetic energy should scale as Z 2 .
Ionization Energies
250
experiment
"approx1.txt"
"approx2.txt"
200

150

100

Fig. 5.2 Ionization Energy (eV) as a
function of atomic number. Exact compared to the two proposed approximations

50

0
0

5

10

15

20

25

The two approximations are plotted here with the exact ionization
energies. Qualitatively they are OK, but quantitatively not so good I
guess.
One can do a bit better (See Fig. 5.3) by assuming that p-shells are
“outside” of s-shells. In other words, a single electron in a p-shell sees
a charge of Z = 1 since the entire s-shell is inside of it. In this case one
gets the following figure (in the two apperoximations discussed above).
In fact, this is getting to be pretty decent. Notice with this second
approximation one obtains a dip in the ionization energy for filled sshells (such as atomic numbers 4 and 12) which is seen in the experiment,
although is weaker in reality.

37

Ionization Energies
140
experiment
"approx3.txt"
"approx4.txt"
120

100

80

Fig. 5.3 Ionization Energy (eV) as
a function of atomic number. Exact
compared to the two proposed approximations with the modification that pshells are declared to be outside of sshells.

60

40

20

0
0

5

10

15

20

25

Once you get to the transition metals, the d-shells really are not very
easily described as being inside or outside of anything. And often when
transition metals ionize, they lose their s-electrons.

(5.3) Exceptions to Madelung’s Rule
Although Madelung’s rule for the filling of electronic
shells holds extremely well, there are a number of exceptions to the rule. Here are a few of them:
Cu = [Ar] 4s1 3d10
Pd = [Kr] 5s0 4d10
Ag = [Kr] 5s1 4d10
Au = [Xe] 6s1 4f 14 5d10

elements followed Madelung’s rule and the Aufbau principle?
 Explain how the statement “3d is inside of 4s”
might help justify this exception in copper.

 What should the electron configurations be if these

Madelung’s rule incorrectly predicts:
Cu = [Ar] 4s2 3d9
Pd = [Kr] 5s24d8
Ag = [Kr] 5s2 4d9
Au = [Xe] 6s2 4f 14 5d9
For copper, the fact that 3d is inside 4s makes the 4s electron less well
bound than you might otherwise expect. Thus the d electrons can fill
preferentially over the s in some cases.

(5.4) Mendeleev’s Nobel Prize
Imagine writing a letter to the Nobel committee nominating Mendeleev, the creator of the periodic table, for
a Nobel Prize. Explain why the periodic table is so important. Remember that the periodic table (1869) was

devised many years before the structure of the hydrogen
atom was understood. (If you do not already have some
background in chemistry, you may want to read the next
chapter before attempting this exercise.)

38 The Periodic Table

Dear Nobel Committee,
Do I have to smack you upside the head? Do
the right thing and give the prize to Mendeleev for
God sake!
Sincerely,
Professor Steven H. Simon

What Holds Solids
Together: Chemical
Bonding
(6.1) Chemical Bonding
(a) Qualitatively describe five different types of chemical bonds and why they occur.
 Describe which combinations of what types of
atoms are expected to form which types of bonds (make
reference to location on the periodic table).
 Describe some of the qualitative properties of materials that have these types of bonds.

6
(Yes, you can just copy the table out of the chapter
summary, but the point of this exercise is to learn the
information in the table!)
(b) Describe qualitatively the phenomenon of van der
Waals forces. Explain why the force is attractive and proportional to 1/R7 where R is the distance between two
atoms.

(a) Just look at the table in the Chapter Summary of chapter 6.
(b) van der Waals forces are from correlated dipole flucuations. If
the electron is a given fixed position, there is a dipole moment p = er
where r is the vector from the electron to the proton. With the electron
“orbiting” (i.e, in an eigenstate), the average dipole moment is zero.
However, if an electric field is applied to the atom, the atom will develop
a polarization (i.e., it will be more likely for the electron to be found on
one side of the nucleus than on the other). We write
p = χE
with χ the polarizability. .
Now, let us suppose we have two such atoms, separated a distance r
in the x̂ direction. Suppose one atom momentarily has a dipole moment
p1 (for definiteness, suppose this dipole moment is in the ẑ direction).
Then the second atom will feel an electric field
p1
E=
4πǫ0 r3
in the negative ẑ direction. The second atom then, due to its polarizability, develops a dipole moment p2 = χE whi ch in turn is attracted
to the first atom. The potential energy between these two dipoles is
U=

−p1 χE
−|p1 |2 χ
−|p1 ||p2 |
=
=
4πǫ0 r3
(4πǫ0 )r3
(4πǫ0 r3 )2

40 Chemical Bonding

corresponding to a force which is attractive and proportional to 1/r7 .
Note that while for a single isolated atom hpi = 0 the result is proportional instead to h|p|2 i ∼ h|r|2 i ∼ with r the position of an electron, is
nonzero. This calculation is done more carefully in problem 6.6 below.

(6.2) Covalent Bonding in Detail*
(a) Linear Combination of Atomic Orbitals:
In Section 6.2.2 we considered two atoms each with a
single atomic orbital. We called the orbital |1i around nucleus 1 and |2i around nucleus 2. More generally we may
consider any set of wavefunctions |ni for n = 1, . . . , N .
For simplicity, let us assume this basis is orthonormal
hn|mi = δn,m (More generally, one cannot assume that
the basis set of orbitals is orthonormal. In Exercise 6.5
we properly consider a non-orthonormal basis.)
Let us write a trial wavefunction for our ground state
as
X
φn |ni.
|Ψi =
n

This is known as a linear combination of atomic orbitals,
LCAO, or tight binding (it is used heavily in numerical
simulation of molecules).
We would like to find the lowest-energy wavefunction
we can construct in this form, i.e., the best approximation to the actual ground-state wavefunction. (The more
states we use in our basis, generally, the more accurate
our results will be.) We claim that the ground state is
given by the solution of the effective Schroedinger equation
Hφ = E φ
(6.1)

where φ is the vector of N coefficients φn , and H is the
N by N matrix
Hn,m = hn|H|mi
with H the Hamiltonian of the full system we are considering. To prove this, let us construct the energy
E=

hψ|H|ψi
hψ|ψi

 Show that minimizing this energy with respect to
each φn gives the same eigenvalue equation, Eq. 6.1.
(Caution: φn is generally complex! If you are not comfortable with complex differentiation, write everything in
terms of real and imaginary parts of each φn .) Similarly,

the second eigenvalue of the effective Schroedinger equation will be an approximation to the first excited state of
the system.
(b) Two-orbital covalent bond
Let us return to the case where there are only two orbitals in our basis. This pertains to a case where we have
two identical nuclei and a single electron which will be
shared between them to form a covalent bond. We write
the full Hamiltonian as
H=

p2
+ V (r − R1 ) + V (r − R2 ) = K + V1 + V2
2m

where V is the Coulomb interaction between the electron
and the nucleus, R1 is the position of the first nucleus
and R2 is the position of the second nucleus. Let ǫ be the
energy of the atomic orbital around one nucleus in the
absence of the other. In other words
(K + V1 )|1i

=

ǫ|1i

(K + V2 )|2i

=

ǫ|2i

Define also the cross-energy element
Vcross = h1|V2 |1i = h2|V1 |2i
and the hopping matrix element
t = −h1|V2 |2i = −h1|V1 |2i
These are not typos!
 Why can we write Vcross and t equivalently using
either one of the expressions given on the right-hand side?
 Show that the eigenvalues of our Schroedinger
equation Eq. 6.1 are given by
E = ǫ + Vcross ± |t|
 Argue (perhaps using Gauss’s law) that Vcross
should roughly cancel the repulsion between nuclei, so
that, in the lower eigenstate the total energy is indeed
lower when the atoms are closer together.
 This approximation must fail when the atoms get
sufficiently close. Why?

41

(a) writing
|ψi =

X

hψ|H|ψi
E=
=
hψ|ψi

n

φn |ni

P

φ∗n Hnm φm
P
2
n |φn |

n,m

(6.2)

We can extremize by differentiating with respect to φ∗n . Note: When
working with complex quantities we can simplify life by treating φn and
φ∗n as independent variables. We thus have
P

P
∗
∂E
φ
n,m φn Hnm φm
m Hnm φm
P
P
P n 2
0 =
=
−
2
2
∂φ∗n
|φ
|
|φ
|
p
n
p
n
p |φp |
X
0 =
Hnm φm − Eφn
m

where we have used Eq. 6.2 to identify E.
(b) The 2 by 2 matrix given by H in this basis is


ǫ + Vcross
t
t∗
ǫ + Vcross
which has eigenvalues E = ǫ + Vcross ± |t|. Note that t can always be
taken as real by simply making a gauge transform on the single particle
wavefunction (i.e., redefining the phase of one of the wavefunctions to
absorb the phase of t).
Here Vcross is the potential felt by the electron on atom 2 due to the
nucleus of atom 1. Since the charge distribution of the electron on atom
2 is roughly spherical we can use Gauss’s law to calculate its interaction
energy with the nucleus of atom 1. If the nucleus of atom 1 is outside
of this spherical distribution of charge of the electron, Gauss’s law tells
us that we can treat the entire spherical distribution of charge as if it is
all at the center of the sphere. In this way, the charge of the electron on
atom 2 exactly cancels the charge of the nucleus of atom 2.
When the two nuclei get close, this argument no longer works, as the
nucleus is then inside much of the distribution of the electron charge. If
the electron charge distribution remains spherical, the nucleus of atom
2 will only see electron charge that is at smaller distances (inside) to
the center of this spherical distribution. When the nuclei are very close
together, the nucleus does not see the electron charge at all, and only
sees the other nucleus. Also, when the atoms get close together, the
assumption of orthogonality of grounds tate orbitals on different sites
breaks down. (One has to choose whether you want orthogonal orbitals
or orbitals that are eigenstates of the single atom Hamiltonian.)

42 Chemical Bonding
(6.3) LCAO and the Ionic–Covalent Crossover
For Exercise 6.2.b consider now the case where the
atomic orbitals |1i and |2i have unequal energies ǫ0,1 and
ǫ0,2 . As the difference in these two energies increases
show that the bonding orbital becomes more localized on

the lower-energy atom. For simplicity you may use the
orthogonality assumption h1|2i = 0. Explain how this
calculation can be used to describe a crossover between
covalent and ionic bonding.

Here we have instead, the two by two hamiltonian matrix


ǫ1 t
t∗ ǫ 2
where we have now absorbed Vcross into the values of ǫi . The lower
energy eigenstate is
o
p
1n
(ǫ1 + ǫ2 ) + (ǫ1 − ǫ2 )2 + 4t2
Eground =
2

with normalized eigenvector

Probability
1

(X, 2t)
ψ= √
4t2 + X 2

0.8
0.6
0.4

with

0.2

-7.5

-5

-2.5

2.5

5

7.5

X = E2 − E1 +

HE2 - E1Lt

Fig. 6.1 Probability (squared amplitude) of ground state wavefunction being on site 1 (solid) or site 2 (dashed)
as a function of E2 − E1 .

p
(E2 − E1 )2 + 4t2

When E2 − E1 ≫ t then X ≫ t and all of the wavefunction ends
up on the first atom (i.e., the one with the lower energy). Similarly, if
E1 − E2 ≫ t then conversely all of the weight of the wavefunction ends
up on the second atom. In Fig. 6.1 it is shown how the weight of the
wavefunction moves from towards the lower energy atom as a function
of energy.
When the energies on the two sites are equal, one has an equal sharing of the wavefunction in the ground state (as in the prior problem).
However, as the energy difference is increased, the ground state moves
more towards the lower energy site, until the bond is completely ”ionic“
meaning that the electron is completely transferred from one atom to
the other.

(6.4) Ionic Bond Energy Budget
The ionization energy of a sodium atom is about 5.14
eV. The electron affinity of a chlorine atom is about 3.62
eV. When a single sodium atom bonds with a single chlorine atom, the bond length is roughly 0.236 nm. Assuming that the cohesive energy is purely Coulomb en-

ergy, calculate the total energy released when a sodium
atom and a chlorine atom come together to form a NaCl
molecule. Compare your result to the experimental value
of 4.26 eV. Qualitatively account for the sign of your error.

(c) The cohesive energy is (with d the bond distance)
Ecoh =

e2
= 6.10eV
4πǫ0 d

43

Thus the total bonding energy is

✻
V (x)

E = −5.14eV + 3.62eV + 6.10eV = 4.58eV
which is slightly larger than the experimentally measured bonding energy. The reason for the discrepancy is that there must be a repulsive
force in addition to the coulomb attractive force which reduces the magnitude of the cohesive (binding) energy.
In the Fig. 6.2 the lower curve is the pure Coulomb energy. The upper
curve includes a short range repulsion. The repulsion must be there, or
there would be no minimum in the curve!

(6.5) LCAO Done Right*
(a)* In Exercise 6.2 we introduced the method of linear combination of atomic orbitals. In that exercise we
assumed that our basis of orbitals is orthonormal. In this
exercise we will relax this assumption.
Consider now many orbitals on each atom (and potentially many atoms). Let us write
|ψi =

N
X
i=1

φi |ii

for an arbitrary number N of orbitals. Let us write the
N by N overlap matrix S whose elements are
Si,j = hi|ji
In this case do not assume that S is diagonal.

Hφ = ESφ

(6.3)

with the same notation for H and φ as in Exercise 6.2.
This equation is known as a “generalized eigenvalue problem” because of the S on the right-hand side.
(b)** Let us now return to the situation with only two
atoms and only one orbital on each atom but such that
h1|2i = S1,2 6= 0. Without loss of generality we may assume hi|ii = 1 and S1,2 is real. If the atomic orbitals
are s-orbitals then we may assume also that t is real and
positive (why?).
Use Eq. 6.3 to derive the eigenenergies of the system.

(6.4)

We can extremize by differentiating with respect to φ∗n to give
! P
P
P
∗
Snm φm
Hnm φm
∂E
n,m φn Hnm φm
m
P n,m ∗
− P
= P
0 =
∗
∗
∗
∂φn
n,m φn Snm φm
n,m φn Snm φm
n,m φn Snm φm
X
X
0 =
Hnm φm − E
Snm φm
m

Fig. 6.2 Probability (squared amplitude) of ground state wavefunction being on site 1 (solid) or site 2 (dashed)
as a function of E2 − E1 .

Using a similar method as in Exercise 6.2, derive the
new “Schroedinger equation”

(a) This is very similar to 6.2.
P
∗
hψ|H|ψi
n,m φn Hnm φm
= P
E=
∗
hψ|ψi
n,m φn Snm φm

✲x

m

where we have used Eq. 6.4 to identify E.
(b) An s-orbital can be taken to be manifestly postive everywhere (no
nodes), so overlaps Sij must be real and positive. Here the easiest thing

44 Chemical Bonding

to do is to just apply S −1 to both sides of the equation to give the
eigenvalue problem
SHφ = Eφ
Here, let us write the Hamiltonian as

ǫ
H=
t∗

t
ǫ



where we have absorbed Vcross into ǫ, and


1 S
S=
S 1
with S = S12 . The matrix we want to diagonalize is then


1
ǫ − St t − ǫS
−1
S H=
t − ǫS ǫ − St
1 − S2
The eigenvalues are easily seen to be
E=

(6.6) Van der Waals Bonding in Detail*
(a) Here we will do a much more precise calculation
of the van der Waals force between two hydrogen atoms.
First, let the positions of the two nuclei be separated by
a vector R, and let the vector from nucleus 1 to electron
1 be r1 and let the vector from nucleus 2 to electron 2 be
r2 as shown in the following figure.
-

✒

■ r1

R

+

✲

-

r2

+
Let us now write the Hamiltonian for both atoms
(assuming fixed positions of nuclei, i.e., using Born–
Oppenheimer approximation) as
H

=

H0

=

H1

=
−

H0 + H1
p1 2
p2 2
e2
e2
+
−
−
2m
2m
4πǫ0 |r1 |
4πǫ0 |r2 |
e2
e2
+
4πǫ0 |R|
4πǫ0 |R − r1 + r2 |
e2
e2
−
4πǫ0 |R − r1 |
4πǫ0 |R + r2 |

Here H0 is the Hamiltonian for two non-interacting hydrogen atoms, and H1 is the interaction between the

1
([ǫ − St] ± |t − ǫS|)
1 − S2

atoms.
Without loss of generality, let us assume that R is in
the x̂ direction. Show that for large R and small ri , the
interaction Hamiltonian can be written as
H1 =

e2
(z1 z2 + y1 y2 − 2x1 x2 ) + O(1/R4 )
4πǫ0 |R|3

where xi , yi , zi are the components of ri . Show that this
is just the interaction between two dipoles.
(b) Perturbation Theory:
The eigenvalues of H0 can be given as the eigenvalues of the two atoms separately. Recall that the
eigenstates of hydrogen are written in the usual notation as |n, l, mi and have energies En = −Ry/n2 with
Ry = me4 /(32π 2 ǫ20 ~2 ) = e2 /(8πǫ0 a0 ) the Rydberg (here
l ≥ 0, |m| ≤ l and n ≥ l + 1). Thus the eigenstates
of H0 are written as |n1 , ll , m1 ; n2 , l2 , m2 i with energies
En1 ,n2 = −Ry(1/n21 + 1/n22 ). The ground state of H0 is
|1, 0, 0; 1, 0, 0i.
 Perturbing H0 with the interaction H1 , show that
to first order in H1 there is no change in the ground-state
energy. Thus conclude that the leading correction to the
ground-state energy is proportional to 1/R6 (and hence
the force is proportional to 1/R7 ).
 Recalling second-order perturbation theory show

45
that we have a correction to the total energy given by
δE =
X

n1 , n2
l1 , l2
m1 , m2

| < 1, 0, 0; 1, 0, 0| H1 |n1 , ll , m1 ; n2 , l2 , m2 i|2
E0,0 − En1 ,n2

 Show that the force must be attractive.
(c)*Bounding the binding energy:
First, show that the numerator in this expression is
zero if either n1 = 1 or n2 = 1. Thus the smallest En1 ,n2
that appears in the denominator is E2,2 . If we replace
En1 ,n2 in the denominator with E2,2 then the |δE| we
calculate will be greater than than the |δE| in the exact
calculation. On the other hand, if we replace En1 ,n2 by

0, then the |δE| will always be less than the δE of the
exact calculation.
 Make these replacements, and perform the remaining sum by identifying a complete set. Derive the bound
6e2 a50
8e2 a50
≤ |δE| ≤
6
4πǫ0 R
4πǫ0 R6
You will need the matrix element for a hydrogen atom
h1, 0, 0|x2 |1, 0, 0i = a20
where a0 = 4πǫ0 ~2 /(me2 ) is the Bohr radius. (This last
identity is easy to derive if you remember that the groundstate wavefunction of a hydrogen atom is proportional to
e−r/2a0 .)

(a) In fact, this is more or less the definition of dipole interaction! Let
us start by deriving
1
|R − a|

1
q
2a·R
a2
R 1 + R2 + R
2




a·R
1 a2
3 (a · R)2
1
(6.5)
1−
+
.
.
.
+
−
+
R
R2
2 R2
2 R4

=

=

Applying this to all the terms in
H1

=
−

e2
e2
+
4πǫ0 |R| 4πǫ0 |R − r1 + r2 |
e2
e2
−
4πǫ0 |R − r1 | 4πǫ0 |R + r2 |

We discover that the first two leading orders completely cancel, thus
leaving us with only the contribution of the terms coming from the
square bracketed terms of Eq. 6.5. We then obtain
 2

r1 + r22 − |r1 − r2 |2
3(R · (r2 − r1 ))2 − (R · r1 )2 − (R · r2 )2 )
e2
+
H1 =
4πǫ0 R
2R2
R4
with R in the x̂ direction, this simplifies to
H1

=



e2
r1 · r2 + 3 (x1 − x2 )2 − x21 − x22
3
4πǫ0 R

which then simplifies to the desired answer.

(b) The expectation of x or y or z in the ground state of the hydrogen
atom |100i is zero due to the fact that the state is spherically symmetric.
As a result taking the expectation of H1 in the ground state gives zero.
Thus the leading correction to the energy occurs at second order in H1

46 Chemical Bonding

and is proportional to 1/R6 (assuming it does not vanish, which we will
show next).
Given then the form of the 2nd order perturbation theory, note that
every term has an overall negative sign (numerator is positive, denominator is negative). Thus the interaction is of the form −C/R6 for some
positive constant C (to be calculated), and hence the force is attractive.
(c) As just above, since x or y or z has zero expectation in the ground
state, this means H1 has zero matrix element unless n1 > 1 and n2 > 1
(for n = 1 we must have l = m = 0).
First let us consider the quantity
X
I=
| < 1, 0, 0; 1, 0, 0| H1 |n1 , ll , m1 ; n2 , l2 , m2 i|2
(6.6)
n1 , n2
l1 , l2
m1 , m2

X

=

n1 , n2
l1 , l2
m1 , m2

h1, 0, 0; 1, 0, 0| H1 |n1 , ll , m1 ; n2 , l2 , m2 i

× hn1 , ll , m1 ; n2 , l2 , m2 | H1 |1, 0, 0; 1, 0, 0i

(6.7)

(6.8)

We notice the complete set in the middle here, so we do the sum over
the set to obtain
I

=
=

h1, 0, 0; 1, 0, 0| H12 |1, 0, 0; 1, 0, 0i
 2 2

e
hy12 ihy22 i + hz12 ihz22 i + 4hx21 ihx22 i
4πǫ0 R

with all expectations being in the ground state of the respective hydrogen
atom (each bracket gives a20 ). Thus we obtain
2

e2
I=
6a40
4πǫ0 R3
Now, the upper bound is defined by setting En1 ,n2 = E2,2 whereas
the lower bound is defined by setting En1 ,n2 = 0 so we obtain
I
I
< |δE| <
E0,0
E0,0 − E2,2
Here using the excitation spectrum of the hydrogen atom, we have
E0,0 =-2Ry and E2,2 = −Ry/2 so E0,0 − E2,2 = −(3/2)Ry, and using the fact that the Rydberg is e2 /(8πǫ0 a0 ) we have
W = I/Ry =
so our inequality is

12e2 a50
4πǫ0R6

W
2W
< |δE| <
2
3
which is the required result.

Types of Matter
There are no exercises for chapter 7.

7

One-Dimensional Model of
Compressibility, Sound,
and Thermal Expansion

(8.1) Potentials Between Atoms
As a model of thermal expansion, we study the distance
between two nearest-neighbor atoms in an anharmonic
potential that looks roughly like this

✻
V (x)

kB T

x0

✻ ✲

x

where x is the distance between the two neighboring
atoms. This potential can be expanded around its minimum as
κ3
κ
(x − x0 )3 + . . .
(8.1)
V (x) = (x − x0 )2 −
2
3!
where the minimum is at position x0 and κ3 > 0. For

small energies, we can truncate the series at the cubic
term. (Note that we are defining the energy at the bottom of the well to be zero here.)
A very accurate approximate form for interatomic potentials (particularly for inert atoms such as helium or
argon) is given by the so-called Lennard-Jones potential
 

σ 12  σ 6
V (x) = 4ǫ
+ǫ
(8.2)
−
x
x
where ǫ and σ are constants that depend on the particular
atoms we are considering.
 What is the meaning of the exponent 6 in the second term of this expression (i.e., why is the exponent
necessarily chosen to be 6).
 By expanding Eq. 8.2 around its minimum, and
comparing to Eq. 8.1, calculate the values of the coefficients x0 , κ, and κ3 for the Lennard-Jones potential in
terms of the constants ǫ and σ. We will need these results
in Exercise 8.3.

The exponent 6 determines the long range behavior of the potential
and is fixed by the form of the Van der Waals interaction.
By setting dV /dx = 0 we find the minimum at x0 = 21/6 σ.
Taking second and third derivatives at this position

and

8

d2 V
= 36 × 22/3 ǫ/σ 2 = κ ≈ 57ǫ/σ 2
dx2 x=x0

(8.3)

√
d3 V
= −756 2 ǫ/σ 3 = −κ3 ≈ 1069ǫ/σ 3
dx3 x=x0

(8.4)

50 Compressibility, Sound, and Thermal Expansion

(8.2) Classical Model of Thermal Expansion
(i) In classical statistical mechanics, we write the expectation of x as
R
dx x e−βV (x)
hxiβ = R
dx e−βV (x)

Although one cannot generally do such integrals for arbitrary potential V (x) as in Eq. 8.1, one can expand the
exponentials as


βκ
2
βκ3
e−βV (x) = e− 2 (x−x0 ) 1 +
(x − x0 )3 + . . .
6
and let limits of integration go to ±∞.

 Why is this expansion of the exponent and the extension of the limits of integration allowed?
 Use this expansion to derive hxiβ to lowest order in
κ3 , and hence show that the coefficient of thermal expansion is
1 dL
1 dhxiβ
1 kB κ3
α=
≈
=
L dT
x0 dT
x0 2κ2
with kB Boltzmann’s constant.
 In what temperature range is the above expansion
valid?
 While this model of thermal expansion in a solid is
valid if there are only two atoms, why is it invalid for the
case of a many-atom chain? (Although actually it is not
so bad as an approximation!)

R
dx x e−βV (x)
hxiβ = R
dx e−βV (x)

with
e

−βV (x)

=e

2
− βκ
2 (x−x0 )



βκ3
(x − x0 )3 + . . .
1+
6



Redefine y = (x − x0 ) so we have
hxiβ

Using

i
h
βκ 2
dy (y + x0 ) e− 2 y 1 + βκ6 3 y 3 + . . .
h
i
=
R
βκ 2
dye− 2 y 1 + βκ6 3 y 3 + . . .
βκ 2
βκ3 R
dy y 4 e− 2 y
6
+ ...
= x0 +
R
βκ 2
dy e− 2 y
R

Z

2

dxe−ax =

p
π/a

as a generating function, we have
Z
Z
p
p
2
2
dxx4 e−ax = (d/da)2 dxe−ax = (d/da)2 π/a = (3/4) π/a5

gives

hxiT = x0 +
Thus

κ3 (kb T )
+ ...
2κ2

1 κ3 kb
1 dhxiT
=
x0 dT
x0 2κ2

In order for this calculation to be valid, since we have treated the cubic
term perturbatively, this term actually must be small compared to the
leading term.

51
2
Roughly, if the leading term is most important, we have
p κ(x − x0 ) ∼
kb T which means that the typical deviation is |x − x0 | ∼ kb T /κ. Then
in order to have the leading term be larger than the cubic term, we have

κ|x − x0 |2 ≫ κ3 |x − x0 |3
or
kb T ≫ κ3 (kb T /κ)3/2
or equivalently
kb T ≪ κ3 /κ23
For a many atom chain, one must solve for the normal modes of the
chain. Then at finite temperature one should “occupy” the phonons
thermally and then calculate the effect of the nonlinear terms in this
state. What we have done in this problem is more or less the thermal
expansion of an Boltzmann model of a solid.

(8.3) Properties of Solid Argon
For argon, the Lennard-Jones constants ǫ and σ from
Eq. 8.2 are given by ǫ = 10meV and σ = .34nm. You will
need to use some of the results from Exercise 8.1.
(a) Sound
Given that the atomic weight of argon is 39.9, estimate
the sound wave velocity in solid argon. The actual value
of the longitudinal velocity is about 1600 m/sec.
(b) Thermal Expansion

Using the results of Exercise 8.2, estimate the thermal
expansion coefficient α of argon. Note: You can do this
part even if you couldn’t completely figure out Exercise
8.2!
The actual thermal expansion coefficient of argon is
approximately α = 2 × 10−3 /K at about 80K. However,
at lower temperature α drops quickly. In the next exercise will use a more sophisticated quantum model to
understand why this is so.

Sound: From the text in one dimension we have
q
v = κx20 /m

where x0 is the neighbor distance. From our above problem on LennardJones, we have
κx20 = 72ǫ
and the mass m = .0399kg/NA . Which gives
v = 1320m/s
Thermal expansion. Plugging in our above expression for x0 , κ and
κ3 for the Lennard Jones case, we obtain
α=

7kb
1 κ3 kb
=
≈ .0012/K
2
x0 2κ
48ǫ

52 Compressibility, Sound, and Thermal Expansion

(8.4) Quantum Model of Thermal Expansion
(a) In quantum mechanics we write a Hamiltonian
H = H0 + δV
where

κ
p2
+ (x − x0 )2
(8.5)
2m
2
is the Hamiltonian for the free Harmonic oscillator, and
δV is the perturbation (see Eq. 8.1)
H0 =

δV = −

κ3
(x − x0 )3 + . . .
6

where we will throw out quartic and higher terms.
 What value of m should be used in Eq. 8.5?
Using perturbation theory it can be shown that, to
lowest order in κ3 the following equation holds
hn|x|ni = x0 + En κ3 /(2κ2 )

(8.6)

where |ni is the eigenstate of the Harmonic oscillator
whose energy is
En = ~ω(n +

1
) + O(κ3 )
2

n≥0

p
with ω = κ/m. In (c) we will prove Eq. 8.6. For now,
take it as given.
 Note that even when the oscillator is in its ground
state, the expectation of x deviates from x0 . Physically
why is this?

(b)* Use Eq. 8.6 to calculate the quantum expectation
of x at any temperature. We write
P
−βEn
n hn|x|nie
P −βE
hxiβ =
n
ne

 Derive the coefficient of thermal expansion.
 Examine the high temperature limit and show that
it matches that of Exercise 8.2.
 In what range of temperatures is our perturbation
expansion valid?
 In light of the current quantum calculation, when
is the classical calculation from Exercise 8.2 valid?
 Why does the thermal expansion coefficient drop
at low temperature?
(c)** Prove Eq. 8.6 by using lowest-order perturbation
theory.
Hint: It is easiest to perform this calculation by using
raising and lowering (ladder) operators. Recall that one
can define operators a and a† such that [a, a† ] = 1 and
√
a† |ni0 =
n + 1|n + 1i0
√
n|n − 1i0 .
a|ni0 =
Note that these are kets and operators for the unperturbed Hamiltonian H0 . In terms of these operators, we
have the operator x − x0 given by
r
~
(a + a† ).
x − x0 =
2mω

(a) For a system of two atoms, one should use the ”reduced” mass
µ = m1 m2 /(m1 + m2 ) which for two identical atoms is m/2.
Quantum mechanically the oscillator does not sit in its exact minimum, rather it has “quantum fluctuations” around this minimum . Since
the well is asymmetric around the minimum, the expectation value of
the position is slightly greater than x0 even in the ground state.
(b)
hxiβ

=

P

hn|x|nie−βEn
nP
−βEn
ne

=

P

n hn|(x0

+ En κ3 /(2κ2 ))|nie−βEn
P −βE
n
ne

hEiβ κ3
2κ2
where hEiβ is the energy expectation of a harmonic oscillator of frequency ω at temperature β = 1/(kb T ). As derived above when we
discussed Einstein model, this expectation is
=

x0 +

1
~ω
hEiβ = (nB (β~ω) + )~ω =
coth(β~ω/2)
2
2

53

with nB the boson occupation factor. Thus we obtain
hxiβ = x0 + (κ3 ~ω/(4κ2 )) coth(β~ω/2)
The coefficient of thermal expansion is then
α=

κ3 dhEi
1 dhxiβ
=
x0 dT
2x0 κ2 dT

where C = dhEi/dT is exactly the specific heat of a harmonic oscillator
(recall einstein model). Thus we obtain
α=

κ3 C
κ3
eβ~ω
2
=
k
(β~ω)
b
2x0 κ2
2x0 κ2
(eβ~ω − 1)2

We can see that the thermal expansion drops as modes ”freeze” out,
entirely analogous to the specific heat dropping at low temperature.
In the high T limit, the specific heat C goes to kb , so we obtain
α = (κ3 kb /(2x0 κ2 ))
in agreement with the classical result.
For the classical calculation to be valid, we must have kb T ≫ ~ω so
that quantum mechanics can be replaced by classical mechanics.
(c) First let us consider the unperturbed harmonic oscillator. Recall
the raising and lowering operators
r
mω
a =
[(x − x0 ) + (i/mω)p]
2~
r
mω
a† =
[(x − x0 ) − (i/mω)p]
2~
(8.7)
So that the position is given by
p
x − x0 = ~/(2mω)(a + a† )

Write the eigenstates |ni0 to mean the eigenstates of the unperturbed
oscillator with energy En = ~ω(n + 1/2). In lowest order perturbation
theory, the perturbed eigenket is given by

3/2 X
† 3
−κ3
~
0 hm|(a + a ) |ni0
|ni = |ni0 +
|mi0
6
2mω
En − Em
m
So that we have

−2κ3
hn|x − x0 |ni =
6



~
2mω

2 X
m

0 hn|a

+ a† |mi0 0 hm|(a + a† )3 |ni0
En − Em

There are two nonzero terms of the sum, the one where n = m + 1 which
gives us a denominator ~ω and a numerator
†
†
†
†
0 hn|a |n − 1i0 0 hn − 1|a aa + aa a + aaa |ni0
√ 
√ √
√
=
n n(n − 1) + n n + (n + 1) n = 3n2

54 Compressibility, Sound, and Thermal Expansion

and also the term with n = m − 1 which gives a denominator −~ω and
a numerator

=

+ 1i0 0 hn + 1|a† a† a + a† aa† + aa† a† |ni0
√
√
√
√

n + 1 n n + 1 + (n + 1) n + 1 + (n + 2) n + 1 = 3(n + 1)2

0 hn|a|n

Thus we get

2κ3
hn|x − x0 |ni =
6



~
2mω

2

1
(6n + 3) = En κ3 /(2κ2 )
~ω

Vibrations of a
One-Dimensional
Monatomic Chain
(9.1) Classical Normal Modes to Quantum
Eigenstates
In Section 9.3 we stated without proof that a classical
normal mode becomes a quantum eigenstate. Here we
prove this fact for a simple diatomic molecule in a potential well (see Exercise 2.7 for a more difficult case, and see
also Exercise 9.7 where this principle is proven in more
generally).
Consider two particles, each of mass m in one dimension, connected by a spring (K), at the bottom of a potential well (with spring constant k). We write the potential
energy as
U=

K
k 2
(x1 + x22 ) + (x1 − x2 )2
2
2

 Write the classical equations of motion.
 Transform into relative xrel = (x1 − x2 ) and center
of mass xcm = (x1 + x2 )/2 coordinates.
(a) Show that in these transformed coordinates, the
system decouples, thus showing that the two normal
modes have frequencies
p
ωcm =
k/m
p
(k + 2K)/m.
ωrel =

9
Note that since there are two initial degrees of freedom,
there are two normal modes.
Now consider the quantum-mechanical version of the
same problem. The Hamiltonian is
H=

 Again transform into relative and center of mass coordinates.
Define the corresponding momenta prel = (p1 − p2 )/2
and pcm = (p1 + p2 ).
(b) Show that [pα , xγ ] = −i~δα,γ where α and γ take
the values cm or rel.
(c) In terms of these new coordinates show that the
Hamiltonian decouples into two independent harmonic
oscillators with the same eigenfrequencies ωcm and ωrel .
Conclude that the spectrum of this system is
Enrel ,ncm = ~ωrel (nrel +

=

mx¨2

=

−kx1 − K(x1 − x2 )

−kx2 − K(x2 − x1 )

Taking the sum and difference of these two equations gives
mẍcm
mẍrel

=
=

−kxcm
−(k + 2K)xrel

1
1
) + ~ωcm (ncm + )
2
2

where ncm and nrel are non-negative integers.
(d) At temperature T what is the expectation of the
energy of this system?

(a) The equations of motion are
mx¨1

p21
p2
+ 2 + U (x1 , x2 )
2m
2m

56 Vibrations of a One-Dimensional Monatomic Chain

which we identify as harmonic oscillators of frequencies
p
ωcm =
k/m
p
(k + 2K)/m
ωrel =

(b) We have [pi , xj ] = −i~δij . Defining prel = (p1 − p2 )/2 and pcm =
p1 + p2 and xrel = x1 − x2 and xcm = (x1 + x2 )/2 we obtain
1
([p1 , x1 ] − [p1 , x2 ] − [p2 , x1 ] + [p2 , x2 ])
2
1
(−i~)(1 − 0 − 0 + 1) = −i~
2
1
([p1 , x1 ] + [p1 , x2 ] − [p2 , x1 ] − [p2 , x2 ])
4
1
(−i~)(1 + 0 − 0 − 1) = 0
4
([p1 , x1 ] − [p1 , x2 ] + [p2 , x1 ] − [p2 , x2 ])
(−i~)(1 + 0 − 0 − 1) = 0
1
([p1 , x1 ] + [p1 , x2 ] + [p2 , x1 ] + [p2 , x2 ])
2
1
(−i~)(1 + 0 + 0 + 1) = −i~
2

[prel , xrel ] =
=
[prel , xcm ] =
=
[pcm , xrel ] =
=
[pcm , xcm ] =
=

One could alternately dmonstrate that prel = −i~∂/∂xrel etc using jacobians to confirm these commutations.
(c) We have
p1

=

pcm /2 + prel

p2
x1

=
=

pcm /2 − prel
xcm + xrel /2

x2

=

xcm − xrel /2

So the Hamiltonian becomes
H

=
=

k
K
1 2
(p + p22 ) + (x21 + x22 ) + (x1 − x2 )2
2m 1
2
2
1 2
k
K
(p /2 + 2p2rel ) + (2x2cm + x2rel /2) + x2rel
2m cm
2
2

which decouples into (note the total mass is 2m and the reduced mass
is m/2)
H1

=

H2

=

p2cm
2k 2
+
x
2(2m)
2 cm
p2rel
(k/2 + K) 2
+
xrel
2(m/2)
2

which are two independent harmonic oscilators with frequencies
p
k/m
ωcm =
p
ωrel =
(k + 2K)/m

57

such that the spectrum can be written as
1
1
Enrel ,ncm = ~ωrel (nrel + ) + ~ωcm (ncm + )
2
2
At temperature T , the expectation of the energy of this system will
correspondingly be
hEi = ~ωrel (nB (β~ωrel ) + 1/2) + ~ωcm (nB (β~ωcm ) + 1/2)
where nB (x) = 1/(ex − 1) is the usual Bose factor.
The purpose of this exercise is not just to do another quantum mechanics problem. It is here to point out that coupled deg rees of freedom
act just like a single simple harmonic oscillator once the degrees of freedom are “rediagonalized”. This is important motivation for treating
phonons (coupled modes of springs) as individual harmonic oscillators.

(9.2) Normal Modes of a One-Dimensional
Monatomic Chain
(a)‡ Explain what is meant by “normal mode” and by
“phonon”.
 Explain briefly why phonons obey Bose statistics.
(b)‡ Derive the dispersion relation for the longitudinal
oscillations of a one-dimensional mass-and-spring crystal
with N identical atoms of mass m, lattice spacing a, and
spring constant κ (motion of the masses is restricted to
be in one dimension).
(c)‡ Show that the mode with wavevector k has the
same pattern of mass displacements as the mode with
wavevector k + 2π/a. Hence show that the dispersion
relation is periodic in reciprocal space (k-space).
 How many different normal modes are there.
(d)‡ Derive the phase and group velocities and sketch
them as a function of k.
 What is the sound velocity?
 Show that the sound velocity is also given by vs =

√
1/ βρ where ρ is the chain density and β is the compressibility.
(e) Find the expression for g(ω), the density of states
of modes per angular frequency.
 Sketch g(ω).
(f) Write an expression for the heat capacity of this
one-dimensional chain. You will inevitably have an integral that you cannot do analytically.
(g)* However, you can expand exponentials for high
temperature to obtain a high-temperature approximation. It should be obvious that the high-temperature
limit should give heat capacity C/N = kB (the law of
Dulong–Petit in one dimension). By expanding to next
non-trivial order, show that
C/N = kB (1 − A/T 2 + . . .)
where

(a) A normal mode is a periodic collective motion where all particles
move at the same frequency. A phonon is a quantum of vibration.
[I do not like the definition ”a quantum of vibrational energy”. The
vibration does carry energy, but it carries momentum as well, so why
specify energy only?]
Each classical normal mode of vibration corresponds to a quantum
mode of vibration which can be excited multiple times. A single mode
may be occupied by a single phonon, or it may be occupied with multiple phonons corresponding to a larger amplitude oscillation. The fact

A=

~2 κ
2
6mkB
.

58 Vibrations of a One-Dimensional Monatomic Chain

that the same state may be multiply occupied by phonons means that
phonons must be bosons.
(b) The equation of motion for the nth particle along the chain is given
by
mẍn = κ(xn+1 − xn ) + κ(xn−1 − xn ) = κ(xn+1 + xn−1 − 2xn )
note that na is the equilibrium position of the nth particle. Using the
ansatz
xn = Aeiωt−ikna
we obtain
−ω 2 meiωt−ikna

ω2m =

or
ω=

= κeiωt (eik(n+1)a + eik(n−1)a − 2eikn )

= 2κ(cos(ka) − 1)

p
p
(2κ/m)(cos(ka) − 1) = 2 κ/m | sin(ka/2)|

ω=2

κ

√m

ω

Fig. 9.1 Dispersion relation for
vibrations of the one-dimensional
monatomic harmonic chain.
The
dispersion is periodic in k → k + 2π/a.

k=−π/a

0

k=+π/a

(c)
e−i(k+2π/a)na = e−i(k+2π/a)na = e−ikna
If you assume periodic boundary conditions, then k = 2πm/L but k is
identified with k + 2π/a so that there are therefore exactly N = L/a
different normal modes.
(d)
vphase = ω(k)/k = 2
and
vgroup = dω(k)/dk =

p
κ/m | sin(ka/2)|/k

q
p
κ/m a cos(|k|a/2) = (a/2)ω0 1 − ω 2 /ω02

59

5676589 84  
3

5676589 84  
3




01

02

03

3

2

1 45676589 3

01

02

03

3

0

0

03

03

2

1 45676589 3

Fig. 9.2 The monatomic harmonic chain. Right: Phase velocity. Left: Group
velocity. Note velocities are signed quantities, to the left of the origin, the velocity
should have negative sign.

p
where ω0 = 2 κ/m. Note that the phase velocity is not periodic in the
Brillouin zone! One can understand this if you think carefully about
aliasing of waves. The phase velocity is the velocity at which the peaks
of waves move. However, the waves are only defined at the position of
the masses along the chain. We write cos(kna) for the positions of the
masses at some time, but this only defines the value of the wave for
integer n. For integer n, we have k is the same as k + 2π/a. However,
the ”peak” of this function may be between the integer values of n.
However, when we make n non-integer, then k is no longer the same as
k + 2π/n.
For sketches see figure 9.2
The sound velocity is the velocity at small k. This is
p
v = a κ/m
. The density of the chain is ρ = m/a and the compressibility is β =
−(1/L)dL/dF = 1/(κa). Thus we obtain v −2 = ρβ
(e) Note first that

(ω(k)/2)2 + (vgroup (k)/a)2 = κ/m

(9.1)

Density of states is uniform in k. If there are N sites in the system,
there are N modes total. The density of states in k is therefore dN/dk =
N a/(2π) = L/(2π) where L is the length of the system.
Thus we have
g(ω) =
=
=

dN/dω = (dN/dk)(dk/dω) =
N
p
2π κ/m cos(|k|a/2)
2N
p
2π (κ/m) − (ω(k)/2)2

where we have used Eq. 9.1.

Na
2πvgroup

(9.2)

60 Vibrations of a One-Dimensional Monatomic Chain

856789

5   6

412
4
312
3
012
012

3

312

4 56789

5 6

Fig. 9.3 The one dimensional harmonic chain. Density of states g(ω). Note that the
p
DOS diverges at ω = 2 k/m where the group velocity goes to zero.

The additional factor of 2 that appears up top is to account for the
fact that for each value of ω > 0 there are actually two values of k with
that ω. (Note if you integrate over frequency you correctly get back N
degrees of freedom).
(f) The energy stored in the chain is given by
Z
U = dωg(ω)~ω(nB (ω) + 1/2)
so the heat capacity is C = ∂U/∂T . Note that we can drop the +1/2
since it has no derivative.
(g) To recover the law of Dulong-Petit, one takes the high temperature
limit of nB (ω) = kB T /~ω so th at we have
Z
Z
∂
C=
dωg(ω)(kB T ) = kB dωg(ω) = kB N
∂T
To go further, we use the high temperature expansion (expanding
1/(ex − 1) for small x)
nB (ω) + 1/2 =

1 ~ω
kB T
+
+ ...
~ω
12 kB T

So that we now have
1
∂U
= kB N − 2
C=
∂T
T

Z



1 ~ω
dω~ωg(ω)
12 kB



So that the coefficient A defined in the problem has the values
Z
~2
dωω 2 g(ω)
A=
2
12N kB
Inserting our expression for g(ω) we obtain
Z ωmax
ω2
~2
dω p
A=
2
12πkB 0
(κ/m) − (ω/2)2

61

p
Defining x = (ω/2) m/κ we obtain
Z
~2 8κ 1
x2
A=
dx √
2
12πkB m 0
1 − x2

The integral is evaluated to give π/4 (make the substitution x = sin θ).
Thus we obtain
~2 κ
A= 2
6kB m
as required.

(9.3) More Vibrations
Consider a one-dimensional spring and mass model of
a crystal. Generalize this model to include springs not
only between neighbors but also between second nearest
neighbors. Let the spring constant between neighbors be

called κ1 and the spring constant between second neighbors be called κ2 . Let the mass of each atom be m.
(a) Calculate the dispersion curve ω(k) for this model.
(b) Determine the sound wave velocity. Show the group
velocity vanishes at the Brillouin zone boundary.

(a) Use the same approach
mẍn

=
=

κ1 (xn+1 − xn ) + κ1 (xn−1 − xn ) + κ2 (xn+2 − xn ) + κ1 (xn−2 − xn )
κ1 (xn+1 + xn−1 − 2xn ) + κ2 (xn+2 + xn−2 − 2xn )

Using the same ansatz
xn = Aeiωt−ikna
we obtain
−mω 2 = 2κ1 (cos(ka) − 1) + 2κ2 (cos(2ka) − 1)
so

(9.3)

r

2κ1
2κ2
(cos(ka) − 1) +
(cos(2ka) − 1)
m
m
(b) To obtain the sound velocity, expand for small k to obtain
ω=

ω

=

r

2κ1 (ka)2
2κ2 (2ka)2
+
=
m 2
m
2

a

r

κ1 + 4κ2
m

!

k

Thus the sound velocity is
vs = a

r

κ1 + 4κ2
m

The easiest way to examine ∂ω/∂k at the zone boundary is to differentiate Eq. 9.3 to given
mω∂ω/∂k = −2aκ1 sin(ka) − 4aκ2 sin(2ka)
At the zone boundary k = π/2 both terms on the right hand side are
zero, hence we have zero group velocity.

62 Vibrations of a One-Dimensional Monatomic Chain

(9.4) Decaying Waves
In the dispersion curve of the harmonic chain (Eq. 9.3),
there is a maximum possible frequency of oscillation
ωmax . If a vibration with frequency ω > ωmax is forced
upon the chain (say by a driving force) the “wave” will
not propagate along the chain, but rather will decay as

one moves away from the point where the oscillation is imposed (this is sometimes known as an “evanescent” wave).
With ω > ωmax solve Eq. 9.3 for a complex k to determine
the decay length of this evanescent wave. What happens
to this length as ω → ωmax ?

The usual form is
Ω = ω/ωmax = ± sin(ka/2)
with ± chosen so Ω is positive. Thus
ka = ±2 sin−1 Ω
When Ω = 1 then ka = π. When Ω > 1 then ka becomes complex
with a real part of π. The complex part gets smaller and vanishes as
Ω approaches 1 from above. The complex part of k gives the inverse
length scale of the wave’s decay (and it can decay left-going or rightgoing depending on the ± sign). So as Ω approachs 1 from above, the
length scale of decay grows longer and longer until at 1 it extends over
the entire system and is a nondecaying wave.
See also the solution of 9.6 below, which is quite similar.
To check these properties of the arcsin, consider
y = sin(x) =

z − z −1
2i

where z = eix . Multiplying by z we obtain the quadratic
z 2 − 2izy − 1 = 0
Solving this equation gives
z = iy ±

p
1 − y2

Note that for y > 1 we have z = eix along the positive imaginary axis,
which means x = π/2 + real. For y = 1 we have x = π/2 whereas for
y < 1 we obtain a complex valued z with unti magnitude, thus a real x.
If one is concerned that this solution might not work, one can always
go back to the equations of motion
mẍn = κ(xn+1 − xn ) + κ(xn−1 − xn ) = κ(xn+1 + xn−1 − 2xn )
Using the ansatz
xn = Aeiωt−ikna
we obtain
−ω 2 meiωt−ikna
ω2m =

= κeiωt (eik(n+1)a + eik(n−1)a − 2eikn )
= 2κ(cos(ka) − 1)

63

Note that nowhere did we specify that k is real (!) so we can happily
extend into the complex plane. Using double angle formula
ω2 =

4κ
sin2 (ka/2)
m

The only thing we would need to be careful about is choosing the sign
of the square root.

(9.5) Reflection at an Interface*
Consider a harmonic chain of equally spaced identical
masses of mass m where left of the n = 0 mass the spring
constant is κL but right of the n = 0 mass, the spring
constant is κR , as shown in this figure.
n=0

a

m

κL

m

κL

m

κL

κR

m

from the left, where it can be either transmitted with
amplitude T or reflected with amplitude R. Using the
following ansatz form

T eiωt−ikL na
n≥0
δxn =
iωt−ikR na
Ie
+ Reiωt+ikR na n < 0
derive T /I and R/I given ω, κL , κR and m.

m

κR

I
R

m

κR

m

T

A wave with amplitude I is incident on this interface

Oops sorry about the typo, it should read

T eiωt−ikR na
δxn =
iωt−ikL na
Ie
+ Reiωt+ikL na

n≥0
n<0

On both the left and the right, we must have the usual relations
between kL,R and ω
ω=2
or

q
κL,R /m| sin(kL,R a/2)|

kL,R =

2
sin−1
a


 q
ω
m/κL,R
2

with these values of k, the ansatz form will then satisfy the equations of
motion except near the boundary.
Near the boundary, we have equations of motion for δx−1 and δx0
which involve δx coordinates on both sides of the junction. We write
¨ −1
m δx
¨0
m δx

=
=

κL (δx−2 + δx0 − 2δx−1 )
κL (δx−1 − δx0 ) + κR (δx1 − δx0 )

64 Vibrations of a One-Dimensional Monatomic Chain

Plugging in the ansatz form of the wavefunction and doing a bit of
algebra gives us
(−mω 2 + 2κL ) IeikL a + Re−ikL a



(−mω 2 + κL + κR )T


κL IeikL 2a + Re−ikL 2a + T

κL IeikL a + Re−ikL a + κR T e−ikR a

=
=

(9.4)

The first equation simplifies quite a bit by rewriting the dispersion relation as
−mω 2 = 2κL (cos(kL a) − 1)
So that
−mω 2 + 2κL = κL eikL a + e−ikL a
and the first equation becomes simply



(9.5)

I +R=T
which is just extending the wavefunction from the left hand side until
it hits position zero and matching it to the value of the wave from the
right at this position.
Plugging this into the second equation of Eq. 9.4 and solving to obtain
T /I =

κL 2i sin(kL a)
−mω 2 + κL (1 − e−ikL a ) + κR (1 − e−ikR a )

In the denominator we can simplify further applying 9.5 to give us
−mω 2 /2 + κL (1 − e−ikL a ) = iκL sin(kL a)
and similarly for the right hand wave so that we get
T /I =

(9.6) Impurity Phonon Mode*
Consider a harmonic chain where all spring constants
have the same value κ and masses have value m, except
for the mass at position n = 0 which instead has value
M < m as shown in this figure:
n=0
a

m

κ

m

κ

m

κ

M

κ

m

κ

m

κ

m

2
1+

κR sin(kR a)
κL sin(kL a)

Along with traveling wave solutions, there can be a
standing wave normal mode localized near the impurity.
Use an ansatz of the form
δxn = Aeiωt−q|n|a
with q real to solve for the frequency of this impurity
mode. Consider your result in the context of Exercise
9.4.

65

Another error here: q needs to have both real and imaginary parts.
This problem is fairly easy once you have solved 9.4. For 0 < ω <
ωmax the only solution to our equations of motion (except at the impurity site) is for the usual waves
−mω 2 = 2κ(cos(ka) − 1)

(9.6)

or equivalently
ω = ωmax | sin(ka/2)|
or
ka = 2 sin−1 (ω/ωmax)
p
with ω > ωmax = 2 κ/m we have decaying solutions of the form given
by the usual where (as shown in Exercise 9.4)
ka = π ± iqa
with q here real. Then on the left we choose + and on the right we
choose − so that the wave decays away from the impurity. So we have

Aeiωt+iπn+qan n ≤ 0
δxn =
Aeiωt−iπn−qan n ≥ 0
with q > 0 this solves all of the equations of motion except for the one
at n = 0
¨0
M δx

= κ(δx−1 + δx1 − 2δx0 )

Plugging in our ansatz, we obtain the three equations
−M ω 2

=

κ(−e−qa − e−qa − 2)

Taking this equation with to Eq. 9.6 and eliminating ω 2 we obtain
M (cosh(qa) + 1) = m(e−qa + 1)
Now, if M = m the only place the two sides are equal is at q = 0, which
means there is no bound state. Further, since the left hand side grows
with increasing q and the right shrinks with q it is clear that there can
be no solution for M > m. However, for M < m there is one crossing
of the two curves, which finds the q appropriate for the bound state. In
fact, one can go a bit further analytically, writing z = eqa we have
(M/m)((z + z −1 )/2 + 1) = z −1 + 1
which is a quadratic equation in z, which we can solve to given
z = eqa =

2m
−1
M

66 Vibrations of a One-Dimensional Monatomic Chain

Or




2m
qa = log
−1
M
which, since we require q > 0 has a solution only for M < m. Finally
to obtain the frequency, one just has to plub back into our dispersion
relation to give
−mω 2 = 2κ(cos(π + iqa) − 1)
which gives a real frequency greater than ωmax

(9.7) General Proof That Normal Modes Become Quantum Eigenstates∗
This proof generalizes the argument given in Exercise
9.1. Consider a set of N particles a = 1, . . . N with masses
ma interacting via a potential
1X
U=
xa Va,b xb
2 a,b
where xa is the deviation of the position of particle a from
its equilibrium position and V can be taken (without loss
of generality) to be a symmetric matrix. (Here we consider a situation in 1d, however, we will see that to go
to 3d we just need to keep track of three times as many
coordinates.)
√
(i) Defining ya = ma xa , show that the classical equations of motion may be written as
X
Sa,b yb
ÿa = −

(ii) Recall the orthogonality relations for eigenvectors
of hermitian matrices
X (m) ∗ (n)
[sa ] [sa ] = δm,n
(9.7)
a

X

1
1
Va,b √
ma
mb
Thus show that the solutions are
Sa,b = √

ya(m) = e−iωm t s(m)
a
where ωm is the mth eigenvalue of the matrix S with
(m)
corresponding eigenvector sa . These are the N normal
modes of the system.

]

=

δa,b.

(9.8)

m

Since S is symmetric as well as hermitian, the eigenvectors can be taken to be real. Construct the transformed
coordinates
X (m) √
sa x a m a
Y (m) =
(9.9)
a

P

(m)

=

X

√
s(m)
a p a / ma

(9.10)

a

show that these coordinates have canonical commutations
[P (m) , Y (n) ] = −i~δn,m

b

where

(m)

∗
[s(m)
a ] [sb

(9.11)

and show that in terms of these new coordinates the
Hamiltonian is rewritten as
X  1 (m) 2 1 2 (m) 2 
H=
[P
] + ωm [Y
]
(9.12)
2
2
.
m
Conclude that the quantum eigenfrequencies of the system are also ωm . (Can you derive this result from the
prior two equations?)

2
Yet another typo, in part (i) it should say −ωm
is the mth eigenvaluer.
(i) First given the expression for U , differentiate with resepect to xa
to obtain the force on particle a
X
∂U
Fa = −
=−
Va,b xb
∂xa
b

The equation of motion is then

ma x¨a = −

X
b

Va,b xb

67

√
Now defining xa = ya / ma we obtain
y¨a = −

X
b

X
1
1
Sa,b yb
√ Va,b √ yb = −
ma
ma
b

−iωt

as required. Using the ansatz ya = e
sa we have
X
ω 2 sa =
Sa,b sa

(9.13)

b

which is solved by sa being the eigenvector of S and ω 2 its eigenvalue.
(ii) Given
X
√
Y (m) =
s(m)
a xa ma
a

P (n)

X

=

√
(n)
s b p b / mb

b

we calculate
[P (n) , Y (m) ] =

X
a,b

p
(n)
sb s(m)
a [pb , xa ] ma /mb

Using the canonical commutations
[pb , xa ] = −i~δab
we have
[P (n) , Y (m) ] = −i~

X
a

(n)
s(m)
a sa = −i~δnm

where we have used the orthogonality of eigenstates.
Next let us look at the terms of the proposed Hamiltonian
X1
XX 1
√
(m)
[P (m) ]2 =
s(m)
a s b p a p b / ma mb
2
2
m
m a,b
X 1
=
p2a
2m
a
a
where we have used the orthogonality of s to collapse the m sum and
generate δab .
Now examining the second term of the proposed Hamiltonian
XX 1
X1
√
2
2 (m)
ωm
[Y (m) ]2 =
s(m)
a ωm sb xa xb ma mb
2
2
m
m
a,b

But here we can use Eq. 9.13 to replace ω 2 with the matrix S giving us
X1
XX 1
√
2
(m)
ωm
[Y (m) ]2 =
s(m)
a Sb,c sc xa xb ma mb
2
2
m
m a,b,c
X1
X1
√
=
Sb,a xa xb ma mb =
xa Va,b xb
2
2
a,b

a,b

68 Vibrations of a One-Dimensional Monatomic Chain

where we have used the orthogonality of s and taken the sum over m to
give a δac .
Thus the given Hamiltonian is equivalent to the original Hamitlonian
X 1
X1
H=
p2a +
xa Va,b xb
2ma
2
a
a,b

The new form of the Hamiltonian in terms of P and Y has each m
coordinate completely decoupled. Further the P and Y satisfy canonical
commutations for momenta and positon so each m is simply a harmonic
oscillator with frequency ωm .
If one wanted to derive the spectrum from the Hamiltonian, one could
follow the usual procedure of writing
√ 
√
1
a† = √
P/ ω + i ωY
2~

to rewrite the Hamiltonian for each decoupled mode as
H = ~ω(a† a + 1/2)

(9.8) Phonons in 2d*

Consider a mass and spring model of a two-dimensional
triangular lattice as shown in the figure (assume the lattice is extended infinitely in all directions). Assume that
identical masses m are attached to each of their six neighbors by equal springs of equal length and spring constant κ. Calculate the dispersion curve ω(k). The twodimensional structure is more difficult to handle than the
one-dimensional examples given in this chapter. In Chapters 12 and 13 we study crystals in two and three dimensions, and it might be useful to read those chapters first
and then return to try this exercise again.

Ok, this one is pretty hard if you have never seen such a thing before.
Probably it should have two stars.
First of all, one assumes that the crystal starts in equilibrium with
the springs unstretched. Forces on the springs will be proportional to
the amount of stretching. Note however, that if one of the masses is
displaced in a direction perpendicular to one of its attaching springs, to
linear order, the spring is not stretched at all (it is only rotated)! So in
other words, we need only keep track of the stretching in the direction
parallel to the spring.
Let us let the edge have unit length for simplicity of notation. If
we let the position of a sites in equilibrium be called rn and let the
displacements from this equilibrium be called urn . Now let us define

69

three unit vectors ai pointing
in the three directions
of the lattice edges
√
√
a1 = (1, 0), a2 = (1/2, 3/2), a2 = (−1/2, 3/2). The total potential
energy of the system can then be written as
U=

3
2

κ XX 
urn − urn +aj · aj
2 r j=1

6

4

n

where here we have included each spring once in the sum and we have
dotted the displacement with its direction so as to only count stretching
of the spring and not rotation. The force on a mass, and hence using
newton’s law we have
α
müα
rn = Frn =

−∂U
∂uα
rn

where α is a basis direction (say, x or y). Carefully taking the derivative
we get



3
X β
X
 [ur +a + uβr −a − 2uβr ]aβj 

aα
müα
j
rn = −κ
n
n
n
p
j
j=1

2

0

-2

-4

-6
-6

-4

-2

0

2

4

6

Fig. 9.4 Contour plot of the transverse
mode frequency of the triangular lattice
phonons in kx , ky space.

6

β

Using a wave ansatz

α iωt−ik·rn
uα
rn = U e

We obtain the two by two eigenvalue problem
X
ω2U α =
Dαβ (k)U β
β

4

2

0

-2

-4

-6
-6

-4

-2

0

2

4

6

where the so-called dynamical matrix is given by
3

κ X α β ik·aj
a a (e
+ e−ik·aj − 2)
Dαβ (k) =
m j=1 j j

Fig. 9.5 Contour plot of the longitudinal mode frequency of the triangular
lattice phonons in kx , ky space.

Note that this matrix is symmetric in α and β. Plugging in the values
of aj and using some trig identities we gets
√ 
√ 





κ
3
4 0
1
1
− 3
2 k · a2
2 k · a3
2 k · a1
√
√
sin (
sin (
D=
sin (
)+
)+
)
0 0
3 3
− 3
3
m
2
2
2
For each k there will be two eigenvalues which correspond to the longitudinal and transverse phonon modes. The eigenvalues of this becomes


q
2κ
S1 + S2 + S3 ± S12 + S22 + S32 − S1 S2 − S1 S3 − S2 S3
ω2 =
m

2
1.5
1

where

k · ai
)
2
These spectra are shown in figures 9.4 - 9.6. Note that the spectrum
is periodic in the Brillouin zone – which has the shape of a hexagon.
Si = sin2 (

0.5
-3

-2

-1

1

2

3

Fig. 9.6 Cut along k = (kx, 0) of
the spectrum of the triangular lattice
phonons.

Vibrations of a
One-Dimensional Diatomic
Chain

(10.1) Normal modes of a One-Dimensional Diatomic Chain
(a) What is the difference between an acoustic mode
and an optical mode.
 Describe how particles move in each case.
(b) Derive the dispersion relation for the longitudinal oscillations of a one-dimensional diatomic mass-andspring crystal where the unit cell is of length a and each
unit cell contains one atom of mass m1 and one atom of
mass m2 connected together by springs with spring constant κ, as shown in the figure (all springs are the same,
and motion of particles is in one dimension only).
a

m1

κ

m2

κ

10

(c) Determine the frequencies of the acoustic and optical modes at k = 0 as well as at the Brillouin zone
boundary.
 Describe the motion of the masses in each case (see
margin note 4 of this chapter!).
 Determine the sound velocity and show that the
group velocity is zero at the zone boundary.
p  Show that the sound velocity is also given by vs =
β −1 /ρ where ρ is the chain density and β is the compressibility.
(d) Sketch the dispersion in both reduced and extended
zone scheme.
 If there are N unit cells, how many different normal
modes are there?
 How many branches of excitations are there? I.e.,
in reduced zone scheme, how many modes are there there
at each k?
(e) What happens when m1 = m2 ?

The following figure depicts a long wavelength acoustic wave: All
atoms in the unit cell move in-phase with a slow spatial modulation.
Acoustic waves ω ∼ k for small k.
a

acoustic
κ
m1

κ
m2

72 Vibrations of a One-Dimensional Diatomic Chain

The following depicts a long wavelength optical wave: The two different types of atoms move out of phase, with a slow spatial modulation.
(In general a long wavelength optical mode is any long wavelength mode
where not all atoms in the unit cell are moving in phase). Note that the
amplitude of motion of the different atoms in the cells is generally not
the same. Optical modes have ω nonzero as k → 0.
a

optical
κ
m1

κ
m2

(b) Let xn be the position of the nth particle of mass m1 and yn be
the position of the nth particle of mass m2 . We can assume that the
equilibrium position of xn is given by na and the equilibrium position
of yn is given by na + d.
We write the equations of motion for the deviations from these equilibrium positions δxn and δyn .
¨n
m1 δx
¨
m2 δy

=

−κ(δxn − δyn−1 ) − κ(δxn − δyn )

=

n

−κ(δyn − δxn ) − κ(δyn − δxn+1 )

Writing the ansätze
δxn
δyn

= Ax eikan−iωt
= Ay eikan−iωt

we obtain the equations
−m1 ω 2 Ax eikna

−m2 ω 2 Ay eikna

= −2κAx eikna + κAy (eikna + eik(n−1)a )

= −2κAy eikna + κAx (eikna + eik(n+1)a )

or
ω 2 Ax
2

ω Ay

= 2(κ/m1 )Ax − (κ/m1 )(1 + e−ika )Ay
= 2(κ/m2 )Ay − (κ/m2 )(1 + e

ika

)Ax

(10.1)
(10.2)

which is an eigenvalue problem from ω 2 . Thus we need to find the roots
of the determinant
2(κ/m1 ) − ω 2
−(κ/m1 )(1 + e−ika )
ika
−(κ/m2 )(1 + e )
2(κ/m2 ) − ω 2

73

which gives the equation
0 =
0 =


κ2
ω 4 − ω 2 (2κ(1/m1 + 1/m2 )) +
4 − (1 + eika )(1 + e−ika )
m1 m2


2
κ
2(m
+
m
)κ
1
2
+
(2 − 2 cos(ka)))
ω4 − ω2
m1 m2
m1 m2

with the solution (skipping a few steps)


q
κ
2
2
2
ω =
m1 + m2 ± m1 + m2 + 2m1 m2 cos(ka)
m1 m2


q
κ
m1 + m2 ± (m1 + m2 )2 − 4m1 m2 sin2 (ka/2)
=
m1 m2
(c) At k = 0, cos(ka) = 1, the acoustic mode has zero energy, whereas
the optical mode has energy
s
2κ(m1 + m2 )
ω=
m1 m2
At the zone boundary cos(ka) = −1, so the two modes have energy
r
r
2κm1
2κm2
ω=
and
m1 m2
m1 m2
the greater of which is the optical mode, the lesser being the acoustic
mode.
To find the motions corresponding to these modes we need to plug
our frequencies back into Eqs. 10.1 and 10.2 to find the relation between
Ax and Ay . For the acoustic mode at k = 0 we obtain Ax = Ay which
means the two masses move in phae with the same amplitude. For the
optical mode at k = 0 we have Ax = −(m2 /m1 )Ay meaning that the
two different masses move in opposite directions with the heavier mass
moving with lower amplitude. At the zone boundary the two modes correspond to one of the masses staying still and the other mass moving.
For example, for the lower frequency mode, the higher mass particles
move and the lower mass particle stays fixed. Since we are at the zone
boundary, every other higher mass particle moves in the opposite direction (thus compressing symmetrically around the fixed particle. An
example of a zone boundary mode is shown in the following figure

a

zone boundary
κ
m1

κ
m2

74 Vibrations of a One-Dimensional Diatomic Chain

To find the sound velocity, expand the cos around k = 0, one obtains
the acoustic mode velocity ω = vk with
r
κ
v=a
2(m1 + m2 )
We check that v −2 = ρβ. The density of the chain is ρ = (m1 + m2 )/a,
the compressibility of two springs in series is κ/2 so the compressibility
of the chain is β = −(1/L)dL/dF = 2/(κa).
Near the zone boundary, since the group velocity is dω/dk and since
dω/d cos(ka) is nonsingular, the group velocity must be zero by using
the chain rule since d cos(ka)/dk = a sin(ka) = 0 at the zone boundary
(k = π/a).

5676589 8 85 84

1
2
3
01

02

03

3

2

1

45676589



5676589 8 85 84

1
2
3
01

02

03

3

2

1 45676589



Fig. 10.1 Diatomic Chain. Top Reduced Zone Scheme. Bottom Extended Zone
Scheme. Both pictures use m1 /m2 = .4.

If there are N unit cells, therefore 2N atoms, there are 2N modes.
There are 2 modes per k in the reduced zone scheme, therefore two
branches.

75

(e) When m1 = m2 the unit cell is now of size a/2 so the Brillouin
zone is doubled in size. In this limit, the gap at the Brillouin zone
boundary vanishes and the two branches become the single branch of
the monatomic chain (this is most easily described in extended zone
scheme).

(10.2) Decaying Waves
Consider the alternating diatomic chain dispersion as
discussed in the text Eq. 10.6 and shown in Fig. 10.6.
For frequencies above ω+ (k = 0) there are no propagating wave modes, and similarly for frequencies between
ω− (k = π/a) and ω+ (k = π/a) there are no propagating

wave modes. As in Exercise 9.4, if this chain is driven at
a frequency ω for which there are no propagating wave
modes, then there will be a decaying, or evanescent, wave
instead. By solving 10.6 for a complex k, find the length
scale of this decaying wave.

As in problem 9.4, we simply want to analytically extend k to complex
numbers. From the text Eq. 10.6 we can rearrange to obtain



1 
(mω 2 − κ1 − κ2 )2 − κ21 − κ22
ka = cos−1
2κ1 κ2

The argument on the right hand size is greater than 1 for ω larger than
the q = 0 optical mode freuquency, whereas for ω between the zone
boundary acoustic and optical model frequencies, the right hand side is
less than −1. The arccos of a number greater than one is pure imaginary
and grows from 0 as the argument increases from unity. Whereas the
arcos of a number less than -1 is π + imaginary with the imaginary part
growing from zero as the argument decreases from -1. The length scale
of decay is always given by L = a/q with q the imaginary part of k.

(10.3) General Diatomic Chain*
Consider a general diatomic chain as shown in Fig. 10.1
with two different masses m1 and m2 as well as two different spring constants κ1 and κ2 and lattice constant a.

(a) Calculate the dispersion relation for this system.
(b) Calculate
the acoustic mode velocity and compare
p
it to vs = β −1 /ρ where ρ is the chain density and β is
the compressibility.

(a) This is the same approach as the prior cases, just a bit more
algebra to keep track of
¨n
m1 δx
¨
m2 δy

n

= −κ1 (δxn − δyn−1 ) − κ2 (δxn − δyn )
= −κ2 (δyn − δxn ) − κ1 (δyn − δxn+1 )

Writing the ansätze
δxn

= Ax eikan−iωt

δyn

= Ay eikan−iωt

76 Vibrations of a One-Dimensional Diatomic Chain

we obtain the equations
−m1 ω 2 Ax eikna
2

−m2 ω Ay e

ikna

=
=

−(κ1 + κ2 )eikna Ax + (κ2 eikna + κ1 eik(n−1)a )Ay

−(κ1 + κ2 )eikna Ay + (κ2 eikna + κ1 eik(n+1)a )Ax

or
ω 2 Ax
ω 2 Ay






κ1 + κ2
κ1
κ2
Ax +
Ay
+ e−ika
m1
m
m1

 1


κ1
κ2
κ1 + κ2
Ay +
Ax
+ eika
=
m2
m2
m2
=

which is an eigenvalue equation for ω 2 . The secular equation is then
0 = (κ1 + κ2 − m1 ω 2 )(κ1 + κ2 − m2 ω 2 ) − κ21 − κ22 − 2κ1 κ2 cos(ka)
or
ω4 −

(κ1 + κ2 )(m1 + m2 ) 2 4κ1 κ2
ω −
sin2 (ka/2) = 0
m1 m2
m1 m2

with the solution (skipping a few steps)
ω2

=

(κ1 + κ2 )(m1 + m2 )
2m1 m2
s
1 (κ1 + κ2 )2 (m1 + m2 )2
16κ1 κ2
±
−
sin2 (ka/2)
2
2
(m1 m2 )
m1 m2

(b) Simply taylor expanding the dispersion gives us
ω=

r

or
vs =

κ1 κ2
ak
(κ1 + κ2 )(m1 + m2 )

s

a2 κ1 κ2
(κ1 + κ2 )(m1 + m2 )

Now the hydrodynamic approach. The density of the chain is ρ =
(m1 +m2 )/a, the spring constant of the two springs in series is κ1 κ2 /(κ1 κ2 )
so the compressibility of the chain
p is β = −(1/L)dL/dF = (κ1 κ2 )/(κ1 +
κ2 )/. Plugging this into vs = β −1 /ρ gives the same result.

77

(10.4) Second Neighbor Diatomic Chain* Consider the diatomic chain from Exercise 10.1. In addition
to the spring constant κ between neighboring masses, suppose that there is also a next nearest-neighbor coupling

with spring constant κ′ connecting equivalent masses in
adjacent unit cells. Determine the dispersion relation for
this system. What happens if κ′ ≫ κ?

We write the equations of motion
¨n
m1 δx
¨
m2 δy
n

=
=

−κ(δxn − δyn−1 ) − κ(δxn − δyn ) − κ′ (δxn − δxn−1 ) − κ′ (δxn − δxn+1 )
−κ(δyn − δxn ) − κ(δyn − δxn+1 ) − κ′ (δyn − δyn−1 ) − κ′ (δyn − δyn+1 )

Writing the usual ansätze
δxn

= Ax eikan−iωt

δyn

= Ay eikan−iωt

we obtain the equations
−m1 ω 2 Ax eikna

=

−m2 ω 2 Ay eikna

−2(κ + κ′ )Ax eikna + κAy (eikna + eik(n−1)a ) + κ′ Ax (eik(n+1)a + eik(n−1)a )

=

−2(κ + κ′ )Ay eikna + κAx (eikna + eik(n+1)a ) + κ′ Ay (eik(n+1)a + eik(n−1)a )

=
=

2 [(κ + κ′ ) − κ′ cos(ka)] Ax − κ(1 + e−ika )Ay
2 [(κ + κ′ ) − κ′ cos(ka)] Ay − κ(1 + eika )Ax

or
m1 ω 2 Ax
m2 ω 2 Ay

which is an eigenvalue problem for ω 2 . Using some trig identies and
some algebra the secular equation can be reduced to


0 = m1 m2 ω 4 − (m1 + m2 ) 2κ + 4κ′ sin2 (ka/2) ω 2 + [4κ2 + 16κκ′ ] sin2 (ka/2) + 16κ′2 sin4 (ka/2)

with the solution
ω2

=

=


1
(m1 + m2 )(2κ + 4κ′ sin2 (ka/2))
2m1 m2

q
± (m1 + m2 )2 (2κ + 4κ′ sin2 (ka/2))2 − 4m1 m2 ([4κ2 + 16κκ′ ] sin2 (ka/2) + 16κ′2 sin4 (ka/2))

1 
(m1 + m2 )(κ + 2κ′ sin2 (ka/2))
m1 m2

q

± 4m1 m2 κ2 cos2 (ka/2) + 4(m1 − m2 )2 κ′2 sin4 (ka/2) + κκ′ sin2 (ka/2)

Note that when κ2 = 0 we recover the diatomic chain from Exercise
10.1. When κ′ ≫ κ we get two decoupled monatomic chains (as should
be expected!).

78 Vibrations of a One-Dimensional Diatomic Chain

(10.5) Triatomic Chain*
Consider a mass-and-spring model with three different
masses and three different springs per unit cell as shown
in this diagram.
a

m3

κ3

m1

κ1

m2

κ2

m3

κ3

m1

κ1

m2

As usual, assume that the masses move only in one
dimension.
(a) At k = 0 how many optical modes are there? Cal-

culate the energies of these modes. Hint: You will get a
cubic equation. However, you already know one of the
roots since it is the energy of the acoustic mode at k = 0
(b)* If all the masses are the same and κ1 = κ2 determine the frequencies of all three modes at the zone
boundary k = π/a. You will have a cubic equation, but
you should be able to guess one root which corresponds
to a particularly simple normal mode.
(c)* If all three spring constants are the same, and
m1 = m2 determine the frequencies of all three modes at
the zone boundary k = π/a. Again you should be able to
guess one of the roots.

With three particles in the unit cell, in one dimesion there is one
acoustic mode and two optical modes.
Declaring the unit cell to be one set of m1 , m2 , m3 with positions
x, y, z our equations of motion are
¨n
m1 δx
¨
m2 δy
n
¨n
m3 δz

= −κ3 (δxn − δzn−1 ) − κ1 (δxn − δyn )

= −κ1 (δyn − δxn ) − κ2 (δyn − δzn )
= −κ2 (δzn − δyn ) − κ3 (δzn − δxn+1 )

Plugging in the usual wave ansatz gives us the secular determinant
0=

κ 3 + κ 1 − m1 ω 2
κ1
κ3 eika

κ1
κ 1 + κ 2 − m2 ω 2
κ2

κ3 e−ika
κ2
κ 2 + κ 3 − m3 ω 2

which multiplied out gives the secular polynomial
0 =

m1 m2 m3 ω 6
− [κ1 (m1 + m2 )m3 + κ3 m2 (m1 + m3 ) + κ2 m1 (m2 + m3 )] ω 4
+ [(κ2 κ3 + κ1 κ2 + κ3 κ1 )(m1 + m2 + m3 )] ω 2
− [κ1 κ2 κ3 (2 − 2 cos(ka))]

(10.3)

At k = 0, the final term vanishes and the secular polynomial clearly
has ω = 0 a root. This we could have guessed in advance since we know
the acoustic mode should come down to zero frequency at k = 0. Once
we remove the obvious factor of ω 2 the remaining polynomial is just
quadratic and we can obtain the two roots in the obvious way.
(b) Setting all the masses the same and κ1 = κ2 , and setting ka = π
we obtain


0 = m3 ω 6 − 2m2 (2κ1 + κ3 ) ω 4




+ (2κ1 κ3 + κ21 )(3m) ω 2 − κ21 κ3 (4)
(10.4)

79

Here we have a difficult cubic. However, with some intuition we realize
that there should be a mode at the zone boundary where the κ3 spring
does not compress at all – the two masses connected to κ3 moving in
unison, but since we are at the zone boundary every other κ3 spring
moves in the opposite direction. The two masses moving in unison have
mass 2m and have a restoring force of 2κ1 from the two springs on the
two sides. Thus we expect a frequency of ω 2 = κ1 /m. Indeed, this
frequncy solves Eq. 10.4. We can thus factor out this root to obtain a
quadratic which we then solve to get the two other roots at the zone
boundary
p
3κ1 + 2κ3 ± 9κ21 − 4κ1 κ3 + 4κ23
2
ω =
2m
(c) Setting all of the spring constants the same and m1 = m2 and
setting ka = π we obtain
0

=

m21 m3 ω 6 − [2κm1 (2m3 + m1 )] ω 4


+ (2m1 + m3 )(3κ2 ) ω 2 − 4κ3

(10.5)

Very similar reasoning suggests a mode at this wavevector with frequency
ω 2 = κ/m1 (which is indeed a solution) which we can then factor out to
give the two additional zone boundary modes
!
p
2 − 4m m + 4m2
3m
+
2m
±
9m
3
1
1
3
3
1
ω2 = κ
2m1 m3

Tight Binding Chain
(Interlude and Preview)

(11.1) Monatomic Tight Binding Chain
Consider a one-dimensional tight binding model of electrons hopping between atoms. Let the distance between
atoms be called a, and here let us label the atomic orbital on atom n as |ni for n = 1 . . . N (you may assume
periodic boundary conditions, and you may assume orthonormality of orbitals, i.e., hn|mi = δnm ). Suppose
there is an on-site energy ǫ and a hopping matrix element
−t. In other words, suppose hn|H|mi = ǫ for n = m and
hn|H|mi = −t for n = m ± 1.
 Derive and sketch the dispersion curve for electrons.
(Hint: Use the effective Schroedinger equations of Exer-

11

cise 6.2a. The resulting equation should look very similar
to that of Exercise 9.2.)
 How many different eigenstates are there in this
system?
 What is the effective mass of the electron near the
bottom of this band?
 What is the density of states?
 If each atom is monovalent (it donates a single electron) what is the density of states at the Fermi surface?
 Give an approximation of the heat capacity of the
system (see Exercise 4.3).
 What is the heat capacity if each atom is divalent?

Write a general wavefunction as
X
Ψ=
φn |ni
n

The schroedinger equation (See problem 6.2a) is
Eφn = ǫφn − t(φn+1 + φn−1 )
write the ansatz
φn = eikna
and we obtain
E(k) = ǫ − 2t cos(ka)
(I won’t sketch this because it is rather trivial). Note however, for t > 0
the minimum occurs at k = 0 whereas for t < 0 the minimum is at the
zone boundary.
If there are N sites in the system, the total length of the system (assumed periodic) is N a. Thus k must be quantized k = 2πm/(N a). Thus
there are exactly N different values of k within the Brillouin zone. This
is to be expected. If we started with a Hilbert space of N dimensions
(described by states |ni), then we expect that there should be exactly

82 Tight Binding Chain (Interlude and Preview)

N eigenstates. (And one gets twice as many eigenstates if you include
spin).
Expanding around the minimum one obtains
E = const + |t|a2 k 2
which we set equal to a free particle dispersion
~2 k 2 /(2m∗ ) = |t|a2 k 2
yielding an effective mass
m∗ = ~2 /(2|t|a2 )
The density of states in k space is uniform as usual dN/dk = L/(2π) =
N a/(2π) (I have not included spin here yet) so the density of states in
energy
dN/dE = (dN/dk)(dk/dE)
Now using

we obtain

q
dE/dk = 2ta sin(ka) = 2|t|a 1 − ((E − ǫ)/(2t))2
dN/dE =

N
2π

2∗2
q
2
2|t| 1 − ((E − ǫ)/(2t))

and, as in exercise 9.2 there is an extra factor of 2 being that one has
two possible k states for each energy E and still another factor of 2
to include the 2 spin states. (Note if you integrate over frequency you
correctly get back 2N degrees of freedom, now with the factor of 2 for
spin included).
If each atom is monovalent, the band is exactly half filled (since two
electrons can go into each orbital). Thus the density of states is
dN/dE =

N 1
π |t|

The heat capacity is given in terms of the density of states as
2
CV = γ̃kB
g(EF )V T

with γ̃ = π 2 /6 (the exact number is from a more complicated calculation
1
that the students are not responsible for). Here g(EF )V = N
π |t| .
If there were two electrons per atom, then the band would be completely full, and there would be no freedom in the system at all. The
heat capacity would be zero. The spin susceptibility would be zero too.

83

(11.2) Diatomic Tight Binding Chain
We now generalize the calculation of the previous exercise to a one-dimensional diatomic solid which might look
as follows:
−A − B − A − B − A − B−
Suppose that the onsite energy of type A is different from
the onsite energy of type B. I.e, hn|H|ni is ǫA for n being
on a site of type A and is ǫB for n being on a site of type
B. (All hopping matrix elements −t are still identical to
each other.)
 Calculate the new dispersion relation. (This is extremely similar to Exercise 10.1. If you are stuck, try

studying that exercise again.)
 Sketch this dispersion relation in both the reduced
and extended zone schemes.
 What happens if ǫA = ǫB ?
 What happens in the “atomic” limit when t becomes very small.
 What is the effective mass of an electron near the
bottom of the lower band?
 If each atom (of either type) is monovalent, is the
system a metal or an insulator?
 *Given the results of this exercise, explain why LiF
(which has very ionic bonds) is an extremely good insulator.

The unit cell a is now the distance from an A atom to the next A
th
atom. Let φA
site of
n be the amplitude of the wavefunction on the n
B
th
type A and φn be the amplitude on the n site of type B.
The Schroedinger equation becomes
EφA
n

=

EφB
n

=

B
B
ǫA φA
n − t(φn + φn−1 )

A
A
ǫB φB
n − t(φn + φn+1 )

Using the ansätz
φA
n

=

Aeikna

φB
n

=

Beikna

gives
EA =
EB =

ǫA A − t(1 + e−ika )B
ǫB B − t(1 + eika )A

again giving a two by two eigenvalue problem, where we solve for the
roots of the determinant
ǫA − E
−t(1 + e−ika )
ika
−t(1 + e )
ǫB − E
yielding the secular equation
0 = E 2 − E(ǫA + ǫB ) + (ǫA ǫB − t2 (2 + 2 cos(ka)))
with the solutions
E± (k) =


p
1
ǫA + ǫB ± (ǫA − ǫB )2 + 4t2 (2 + 2 cos(ka))
2

Note that in the limit that ǫA = ǫB we recover
E = ǫ ± 2|t| cos(ka/2)

84 Tight Binding Chain (Interlude and Preview)

5676589 8

2

3

01

02

03

3

2

1 45676589



2

1 45676589



5676589 8

2

3

01

02

03

3

Fig. 11.1 Diatomic Tight Binding Problem in 1D. Left: Reduced Zone scheme.
Right: Extended Zone scheme. In both pictures we have chosen ǫA = 2.56 and
ǫB = 1.56

which matches the solution of part (a) above, but for the change in the
definition of the size of the unit cell. (See the figure for sketches of the
dispersion)
In the limit of small t the bands become very flat at energies ǫA and
ǫB .
Expanding around the minimum gives
2t2 (ka)2
E = constant + p
(ǫA − ǫB )2 + 16t2

which we set equal to ~2 k 2 /(2m∗ ) to yield
p
~2 (ǫA − ǫB )2 + 16t2
∗
m =
4t2 a2
If each atom is monovalent, there are now two electrons per unit cell,
and this fills exactly the lower band (leaving the upper band empty).
The system is therefore an insulator.
If ǫA = ǫB this becomes a monotonic chain, the gap closes and it
becomes a metal.
For LiF we can expect a much lower energy for electrons on F than
on Li (F high high electron affinity, Li has low ionization energy). So

85

we can set ǫA ≪ ǫB . What happens in this limit is that the band
are extremely far apart – thus a very good insulator. Further, if we
look a the eigenvectors in the lower band, we will find that they are
almost completely on the lower energy atoms. Thus the free electron is
transferred almost completely from the higher to the lower energy atom.

(11.3) Tight Binding Chain Done Right
Let us reconsider the one-dimensional tight binding
model as in Exercise 11.1. Again we assume an on-site energy ǫ and a hopping matrix element −t. In other words,
suppose hn|H|mi = ǫ for n = m and hn|H|mi = −t for
n = m±1. However, now, let us no longer assume that orbitals are orthonormal. Instead, let us assume hn|mi = A
for n = m and hn|mi = B for n = m + 1 with hn|mi = 0
for |n − m| > 1.

 Why is this last assumption (the |n − m| > 1 case)
reasonable?
Treating the possible non-orthogonality of orbitals here
is very similar to what we did in Exercise 6.5. Go back
and look at that exercise.
 Use the effective Schroedinger equation from Exercise 6.5 to derive the dispersion relation for this onedimensional tight binding chain.

One can assume that orbitals far apart have no overlap since the tails
of wavefunctions decay exponentially.
The generalized Schoedinger equation is
X
X
Hnm φm = E
Snm φm
m

m

Hnm

=

Snm

=

ǫ0 δnm − t(δn,m+1 + δn,m−1 )

where here

Aδnm + Bδn,m+1 + B ∗ δn,m−1

(note we did not promise that B is real!). Plugging in our wave ansatz
φm = eimka into the Schroedinger equation, we obtain
E(A + Beik + B ∗ e−ik ) = ǫ0 − 2t cos(ka)
or
E=

ǫ0 − 2t cos(ka)
A + Beik + B ∗ e−ik

(11.4) Two Orbitals per Atom
(a) Consider an atom with two orbitals, A and B havB
ing eigenenergies ǫA
atomic and ǫatomic . Now suppose we
make a one-dimensional chain of such atoms and let us
assume that these orbitals remain orthogonal. We imagine hopping amplitudes tAA which allows an electron on
orbital A of a given atom to hop to orbital A on the neighboring atom. Similarly we imagine a hopping amplitude

tBB that allows an electron on orbital B of a given atom
to hop to orbital B on the neighboring atom. (We assume that V0 , the energy shift of the atomic orbital due
to neighboring atoms, is zero).
 Calculate and sketch the dispersion of the two resulting bands.
 If the atom is diatomic, derive a condition on the
B
quantities ǫA
atomic − ǫatomic , as well as tAA and tBB which

86 Tight Binding Chain (Interlude and Preview)
determines whether the system is a metal or an insulator.
(b)* Now suppose that there is in addition a hopping
term tAB which allows an electron on one atom in orbital

5

A to hop to orbital B on the neighboring atom (and vice
versa). What is the dispersion relation now?

(a) The two orbitals are completely decoupled from each other. So we
get two independent dispersions (we assume t’s real here)

4
3
2
1
-3

-2

-1

E = ǫA − 2tAA cos(ka)

(11.1)

E = ǫB − 2tBB cos(ka)

(11.2)

and
1

2

3

-1
-2

Fig. 11.2 Dispersion in the two oribtal tight binding model (Eqs. 11.1 and
11.2). Here, tAA = 1, tBB = 2, ǫA = 0,
ǫB = 4. Horizontal axis is ka.

A sketch is given in Fig. 11.2.
It should say ”’if the atom is divalent” not diatomic (doh!). If the
atom is divalent, then it will be a metal if the bands overlap and an
insulator (or semiconductor) if the bands do not overlap. A band with
the dispersion as in Eqs. 11.1 has a band which extends from ǫA − 2|tAA |
to ǫA + 2|tA A| (and similar for the other band – this holds even if t’s are
complex). In order for the bands to not overlap we must have either
ǫA + 2|tAA | < ǫB − 2|tBB |
or
ǫB + 2|tBB | < ǫA − 2|tAA |

If neither of these is satisfied the bands must overlap and we have a
metal. Note if we are tricky we can summarize the two above conditions
as one condition which, if satisfied, tells us we have an insulator
|ǫA − ǫB | > 2 (|tAA | − |tBB |)
(b) When there is hopping tAB between an A orbital on one site and a
B orbital on a neighboring site, we have a matrix schroedinger equation
EφA
n
EφB
n

A
∗
A
B
∗
B
= ǫA φA
n − tAA φn+1 − tAA φn−1 − tAB φn+1 − tAB φn−1

B
∗
B
A
∗
A
= ǫB φB
n − tBB φn+1 − tBB φn−1 − tAB φn+1 − tAB φn−1

Using the usual ansätz
φA
n

=

Aeikna

φB
n

=

Beikna

we obtain
EA

=

EB

=



ǫA − tAA eika − t∗AA e−ika A + −tAB eika − t∗AB e−ika B


ǫB − tBB eika − t∗BB e−ika B + −tAB eika − t∗AB e−ika A

which is an eigenvalue equation for E. Defining
TAA
TBB
TAB

=
=
=

tAA eika + t∗AA e−ika = 2Re[tAA eika ]
ika

tBB e + t∗BB e−ika
tAB eika + t∗AB e−ika

ika

= 2Re[tBB e ]
= 2Re[tAB eika ]

(11.3)
(11.4)
(11.5)

87

We can write the secular equation as
0

= E 2 + E (TAA + TBB − ǫA − ǫB )

2
+ (ǫA − TAA ) (ǫB − TBB ) − TAB

(11.6)

which we solve as a quadratic equation to give


q
1
2
2
ǫA + ǫB − TAA − TBB ± (ǫA − ǫB − TAA + TBB ) − 4TAB
E=
2

(11.5) Electronic Impurity State*
Consider the one-dimensional tight binding Hamiltonian given in Eq. 11.4. Now consider the situation where
one of the atoms in the chain (atom n = 0) is an impurity
such that it has an atomic orbital energy which differs by
∆ from all the other atomic orbital energies. In this case
the Hamiltonian becomes
Hn,m = ǫ0 δn,m − t(δn+1,m + δn−1,m ) + ∆δn,m δn,0 .
(a) Using an ansatz
φn = Ae−qa|n|

with q real, and a the lattice constant, show that there
is a localized eigenstate for any negative ∆, and find the
eigenstate energy. This exercise is very similar to Exercise
9.6.
(b) Consider instead a continuum one-dimensional
Hamiltonian with a delta-function potential
H=−

Similarly show that there is a localized eigenstate for any
negative ∆ and find its energy. Compare your result to
that of part (a).

(a) Consider first without the impurity. Propose exponentially decaying or growing solutions
φn = Ae±qa|n|
plugging these in we obtain
E = ǫ0 − 2t cosh(qa)

~2 2
∂ + (a∆)δ(x).
2m∗ x

(11.7)

these give solutions with q real for (ǫ0 − E)/2t > 1, or in other words,
for energies below the bottom of the band. These are analogous to the
evanescent waves discussed in Exercises 9.4 and 10.2.
Now, we can patch together two of these evanescent waves at the
impurity. Examining the schroedinger equation at position zero, we
have
Eφ0 = (ǫ0 + ∆)φ0 − t(φ1 + φ−1 )
which gives us

E − ǫ0 − ∆ = 2te−qa

Then plugging in the value of E in terms of q from Eq. 11.7, we obtain
−∆ = 2 sinh(qa)
recalling that ∆ was negative we have a solution for any value of ∆ < 0
with
qa = sinh−1 (|∆|/(2t))

88 Tight Binding Chain (Interlude and Preview)

Plugging back into Eq. 11.7 we obtain the bound state energy
p
E = ǫ0 − 2t [∆/(2t)]2 + 1

and note that for ∆ = 0 this gives us the energy at the bottom of the
band.
(b) For the continuum problem we have similarly a decaying waves of
the form e−qx with energies
E=

−~2 q 2
2m∗

Then we need to patch together solutions at zero, we have
∂ψ− − ∂ψ+ = 2q = (a|∆|)(2m∗ )/~2
or
q = (a|∆|m∗ )/~2
with energy
E = −a2 m∗ ∆2 /(2~2 )
To compare to the above tight binding problem, we use the value of
the effective mass m∗ = ~2 /(2ta2 ) to obtain
q = |∆|/(2t)
which matches our above result for small ∆. Correspondingly the energy
is
E = −∆2 /(4t)
which also matches the above result for small ∆ so long as we choose
the bottom of the band to be called zero energy.

(11.6) Reflection from an Impurity*
Consider the tight binding Hamiltonian from the previous exercise representing a single impurity in a chain.
Here the intent is to see how this impurity scatters a
plane wave incoming from the left with unit amplitude
(this is somewhat similar to Exercise 9.5). Use an ansatz

wavefunction
φn =



T e−ikna
e−ikna + Re+ikna

n≥0
n<0

to determine the transmission T and reflection R as a
function of k.

Use the schroedinger equation at position n = −1
Eφ−1 = ǫ0 φ−1 − t(φ0 + φ−2 )
and at position zero
Eφ0 = (ǫ0 + ∆)φ0 − t(φ1 + φ−1 )

89

and we plug in the ansatz form. The former equation simply tells us
that we must have 1 + R = T (To see this recall that the relationship
between E and k is fixed by the plane wave away from positon zero.
So this first equation tells us that position φ0 must just be the usual
continuation of this wave-front with no changes.) The second equation
then gives us
E T = (ǫ0 + ∆)T − t(T eika + eika + Re−ika )
Using 1 + R = T as well as using E = ǫ0 + 2t cos(ka) this becomes
T =

1
∆
2it sin(ka)

+1

Note that when ∆ = 0, we have complete transmission T = 1 and
R = 0. With R = T − 1, we have |R|2 + |T |2 = 1 which indicates current
conservation.

(11.7) Transport in One Dimension*
(a) Consider the one-dimensional tight binding chain
discussed in this chapter at (or near) zero temperature.
Suppose the right end of this chain is attached to a reservoir at chemical potential µR and the left end of the
chain is attached to a reservoir at chemical potential µL
and let us assume µL > µR . The particles moving towards the left will be filled up to chemical potential µR ,
whereas the particles moving towards the right will be
filled up to chemical potential µL , as shown in the bottom of Fig. 11.4, and also diagrammed schematically in
the following figure
µL
µL

µR
µR

(i) Argue that the total current of all the particles moving to the right is
Z ∞
dk
v(k)nF (β(E(k) − µL ))
jR =
π
0
with v(k) = (1/~)dǫ(k)/dk the group velocity and nF the
Fermi occupation factor; and an analogous equation holds
for left moving current.
(ii) Calculate the conductance G of this wire, defined
as
Jtotal = GV
where Jtotal = jL − jR and eV = µL − µR , and show
G = 2e2 /h with h Planck’s constant. This “quantum” of

conductance is routinely measured in disorder free onedimensional electronic systems.
(iii) In the context of Exercise 11.2, imagine that an impurity is placed in this chain between the two reservoirs
to create some backscattering. Argue that the conductance is reduced to G = 2e2 |T |2 /h. This is known as the
Landauer formula and is a pillar of nano-scale electronics.
(b) Now suppose that the chemical potentials at both
reservoirs are the same, but the temperatures are TL and
TR respectively.
(i) Argue that the heat current j q of all the particles
moving to the right is
Z ∞
dk
q
jR =
v(k) (E(k) − µ) nF (βL (E(k) − µ))
π
0
and a similar equation holds for left-moving heat current.
(ii) Define the thermal conductance K to be
J q = K(TL − TR )
q
q
where J q = jL
− jR
and TL − TR is assumed to be small.
Derive that the thermal conductance can be rewritten as
Z
−2 ∞
∂
K=
nF (β(E − µ)).
dE(E − µ)2
hT −∞
∂E

Evaluating this expression, confirm the Wiedemann–
Franz ratio for clean one-dimensional systems
2
π 2 kB
K
=
2
TG
3e

(Note that this is a relationship between conductance and
thermal conductance rather than between conductivity

90 Tight Binding Chain (Interlude and Preview)
and thermal conductivity.) In evaluating the above integral you will want to use
Z ∞
∂
1
π2
dx x2
=−
x +1
∂x
e
3 .
−∞
If you are very adventurous, you can prove this nasty

J

identity using the techniques analogous to those mentioned in footnote 20 of Chapter 2, as well as the fact that
the Riemann zeta function takes the value ζ(2) = π 2 /6
which you can prove analogous to the appendix of that
chapter.

(a)(i) Genereally the velocity of a wavepacket is the group velocity
v = dω/dk. Here, ω = Ek /~ To find the total right going current, one
integrates over all possible k, the occupancy of that k times the group
velocity, to obtain the current. There should be an additional factor of
−e out front to turn this into an electrical current. The integral over
k usually has 2π downstairs, but the 2 is cancelled by a factor of 2 out
front for two spins. The integral over k also usually comes with a factor
of volume (or length in this case) out front. To see why this is missing we
have to think carefully about the definition of current – which counts the
number of particles crossing a given point in some unit of time. If one
has a particle in a k eigenstate, it is delocalized over the entire system.
Thus the ”probability“ that with velocity v it crosses that given point
in a unit of time is given by v/L.
(ii) We can write the total current as
Z ∞

Z −∞
dE(k)
dE(k)
e
nF (β(Ek − µR ))dk −
nF (β(Ek − µL ))dk
= −
π 0 dk
dk
Z ∞

Z ∞ 0
e
dEnF (β(E − µR )) −
dEnF (β(E − µL ))
= −
~π 0
0
e
[µL − µR ]
= −
~π
where in the last step we have assumed that µ ≫ kB T so that the
occupancy at E = 0 is unity — in which case the integral of the Fermi
function simply gives the chemical potential (one can see this by starting
at T = 0 where the Fermi function is just a step function. As we raise
the temperature, the change in the fermi function is symmetric around
the chemical potential, so even though the fermi function gets smaller
below µ and larger above µ the integral is unchanged).
Plugging in the expression for voltage
J =−

e2
V
π~

or G = 2e2 /h.
(iii) If the amplitude of transmission through the impurity is T then
the probability of transmission is |T |2 . The resulting current is then
reduced by the probability that each particle is transmitted. Thus we
obtain
G = 2e2 |T |2 /h

Note that as we discovered in exercise 11.6, T is a function of wavevector
and hence energy. Once should use k near kF to calculate T . The

91

reason for this is that all of the net current is coming from near the
Fermi surface. Well below the fermi surface the currents of left and
right movers exactly cancel.
(b) (i) The argument here is quite similar. Here each electron is
moving with velocity v, but is carrying energy E − µ. Here energy is
measured with respect to the chemical potential – this is an appropriate
definition since even at T = 0 the addition of an electron carries the
chemical potential worth of energy. The difference from this T = 0
energy is the excess heat added.
(ii) We take the difference of right moving and left moving energy
currents (the sign gives current moving from left to right)
Z ∞

Z −∞
−1
dE(k)
dE(k)
Jq =
(Ek − µ)nF (βR (Ek − µ))dk −
(Ek − µ)nF (βL (Ek − µL ))dk
π
dk
dk
0
Z0 ∞

Z ∞
−1
=
dE(E − µ)nF (βR (E − µ)) −
dE(E − µ)nF (βL (E − µ))
~π 0
0
Z ∞
−1
dnF (β(E − µ))
=
(βR − βL )
dE(E − µ)
~π
dβ
0
(11.8)
where we have made the approximation that βR − βL is small to
replace the finite difference with a derivative with respect to β. This
then becomes.
−1
J =
(βR − βL )
~πβ
q

Z

∞
0

dE(E − µ)2

dnF (β(E − µ))
dE

then using
βR − βL ≈

TL − TR
kB T 2

gives the desired result. Note that the formula also extends the lower
limit of integration to −∞. This is allowed since dnF /dE is strongly
peaked around µ.
Finally scaling out some factors of β we obtain
−2T
K=
h

Z

∞

dxx2

−∞

2
1
2π 2 kB
T
d
=
dx ex + 1
3h

Thus dividing by the expresion for G we obtain
2
π 2 kB
T
K
=
2
GT
3he

which is the Weideman-Franz law.
For completeness, we evaluate the integral. First, note that the inte-

92 Tight Binding Chain (Interlude and Preview)

grand is symmetric around zero so we can write
Z ∞
Z ∞
Z ∞
1
1
e−x
2 d
2 d
I =
dxx
=
2
dxx
=
−4
dxx
dx ex + 1
dx ex + 1
1 + e−x
−∞
0
0
Z ∞
∞
∞
X
X (−1)n
= 4
dxx
(−ex )n = 4
n2
0
n=1
" ∞ n=1
#
∞
X
X 1
1
2
= −4
= 4(1 − )ζ(2) = 2ζ(2)
−
2
2
2
n
n
4
n=1,evens
n=1
The method of calculating ζ(2) proceeds similar to the appendix of
Chapter 1. First, consider the function x in the range [−π, π] as a fourier
series. It can be writen as
x=

∞
X

an sin(nx)

n=1

with
an =

1
π

The calculate

Z

π

−π

Z

x sin(nx) = −2(−1)n /n

π

dx x2 = 2π 3 /3

−π

but also caculate the same quantity in terms of its fourier transforms
(using percival’s theorem)
Z

π

dx x2 =

−π

Z

π

dx =
−π

∞
X

n=1

|an |2 sin2 (nx) =

∞
X
4
π
2
n
n=1

setting these two expressions equal to each other gives
∞
X
π2
1
=
n2
6
n=1

(11.8) Peierls Distortion*
Consider a chain made up of all the same type of atom,
but in such a way that the spacing between atoms alternates as long-short-long-short as follows
−A = A − A = A − A = A−
In a tight binding model, the shorter bonds (marked with
=) will have hopping matrix element tshort = t(1 + ǫ)
whereas the longer bonds (marked with −) have hopping
matrix element tlong = t(1 − ǫ). Calculate the tightbinding energy spectrum of this chain. (The onsite energy ǫ is the same on every atom). Expand your result to

linear order in ǫ. Suppose the lower band is filled and the
upper band is empty (what is the valence of each atom
in this case?). Calculate the total ground-state energy of
the filled lower band, and show it decreases linearly with
increasing ǫ.
Now consider a chain of equally spaced identical A
atoms connected together with identical springs with
spring constant κ.
Show that making a distortion
whereby alternating springs are shorter/longer by δx
costs energy proportional to (δx)2 . Conclude that for a
chain with the valence as in the first part of this problem,
a distortion of this sort will occur spontaneously. This is
known as a Peierls distortion.

93

OK, I messed up this problem a bit, so bear with me.
First we need to evalulate the spectrum. Consider the unit cell to be
a single unit like this A = A−. Call the left side the A site and the
right site the B site. Set the onsite energy epsilon0 to zero for simplicity
and assume t is real (without loss of generality). We then have the
tight-binding schroedinger equation
EφA
n

=

EφB
n

=

B
−t(1 + ǫ)φB
n − t(1 − ǫ)φn−1
A
−t(1 + ǫ)φA
n − t(1 − ǫ)φn+1

Using the usual ansatz this gives us
EA =
EB =

−t(1 + ǫ)B − t(1 − ǫ)Be−ika
−t(1 + ǫ)A − t(1 − ǫ)Aeika

which is an eigenvalue problem with solutions
p
E(k) = ± |2t[cos(ka/2) + iǫ sin(ka/2)]| = ±|2t| ǫ2 + (1 − ǫ2 ) cos2 (ka/2)

To find the total energy of a filled lower band, we need
Z
dk
E(k)
Etot = 2L
2π

which is some horrid ellipic integral. Ideally, we just evaluate this horrid
integral and we are done, but this is very difficult analytically (we could
do it numerically though!).
However, all we need to do is to make an estimate of how this changes
to lowest order in ǫ. Near the zone boundary, since the cosine term is
small the effect of epsilon is most pronounced where the energy reduction
in the lower band is −|2tǫ|. As we move away from the zone boundary
the cosine term becomes more important. Define κa = π − ka to be the
distance to the zone boundary (focusing for now on the zone boundary
at ka = π).
For small κ we see that the two term in the square root become roughly
equal when
κa ≈ ǫ
And when κa ≫ ǫ the cosine term dominates. Let us then split the
integration roughly at this intermediate point κa ≈ ǫ. For smaller κ, the
energy reduction can be approximated as −|2tǫ| for each value of κ and
integrating then a range of κ that is ǫ large we get an energy reduction
of ∼ L|t|ǫ2 , which we are not interested in. For the region further from
the zone boundary, we have roughly
Z cutoff
hp
i
δE ≈ |2tL|
dκ
ǫ2 + (κa)2 /4 − κa/2
κa=ǫ
Z cutoff
ǫ2
dκ ≈ |2tL|ǫ2 log(ǫ/cutoff)
≈ |2tL|
κ
κa=ǫ

94 Tight Binding Chain (Interlude and Preview)

here the cutoff is some arbitrary momentum much further from the zone
boundary (it will not matter where we choose the cutoff!). Thus we see
that the energy saving is proportional to ǫ2 log ǫ.
(ii) The energy of stretching springs is always quadratic in the stretching (hookes’ law), hence proportional to ǫ2 . For small ǫ the electronic
energy saving ǫ2 log ǫ always wins, so the system always distorts spontaneously!

(11.9) Tight Binding in 2d*
Consider a rectangular lattice in two dimensions as
shown in the figure. Now imagine a tight binding model
where there is one orbital at each lattice site, and where
the hopping matrix element is hn|H|mi = t1 if sites n and
m are neighbors in the horizontal direction and is = t2 if
n and m are neighbors in the vertical direction. Calculate
the dispersion relation for this tight binding model. What
does the dispersion relation look like near the bottom of
the band? (The two-dimensional structure is more difficult to handle than the one-dimensional examples given
in this chapter. In Chapters 12 and 13 we study crystals in two and three dimensions, and it might be useful
to read those chapters first and then return to try this

exercise again.)

t2
t1

This is a lot easier than it looks! Let us label the sites (m,n). Let
the wavefunctions on the sites be φn,m accordingly. The Schroedinger
equation is then
Eφn,m = ǫ0 φn,m − t1 (φn+1,m + φn−1,m ) − t2 (φn,m+1 + φn,m−1 )
Using an ansatz
φn,m = eikx nax +ky may
where ax and ay are the lattice distances in the two directions.
and assuming both t1 and t2 real
E = ǫ0 − 2t1 cos(kx ax ) − 2t2 cos(ky ay )
Assuming both t1 and t2 positive, the minimum occurs at kx = ky = 0.
Near this minimum, we can expand to gets
E ≈ (ǫ0 − 2t1 − 2t2 ) +


1
t1 (kx ax )2 + t2 (ky ay )2
2

which has equal-E contours which are ellipses. Thus near the bottom of
the band we have an ellipsoidal bowl.

12

Crystal Structure

(12.1) Crystal Structure of NaCl Consider the
NaCl crystal structure shown in Fig. 12.21. If the lattice
constant is a = 0.563 nm, what is the distance from a

sodium atom to the nearest chlorine? What is the distance from a sodium atom to the nearest other sodium
atom?

Super easy:
√ The Na-Cl distance is a/2 = .2815 nm whereas the Na-Na
distance is a 2/2 = .398 nm.

(12.2) Neighbors in the Face-Centered Lattice.
(a) Show that each lattice point in an fcc lattice has
twelve nearest neighbors, each the same distance from the
initial point. What is this distance if the conventional
unit cell has lattice constant a?
(b)∗ Now stretch the side lengths of the fcc lattice

such that you obtain a face-centered orthorhombic lattice
where the conventional unit cell has sides of length a, b,
and c which are all different. What are the distances to
these twelve neighboring points now? How many nearest
neighbors are there?

(a) Given the primitive lattice vectors, one can define the fcc lattice
as being vectors of the form a2 (n, m, l) where n, m, l are either all even
or two odd and one even. Among this sets of possible lattice points, the
closest ones to [0, 0, 0] are of the form a2 [1, 1, 0]. Here the two entries that
are 1 could be either ±1 which gives four possibilities. Further the 0 can
be in one of three places. Thus we have √
12 possibilities. The distance
from [0, 0, 0] to any of these points is a 2/2 (analogous to previous
problem).
(b) In units of the three different (unequal) conventional unit cell
lattice constants, the 12 points are still of the form 12 [1, 1, 0] and permutations. The distances to these 12 points are then
[±1, ±1, 0] →
[±1, 0, ±1] →
[0, ±1, ±1] →

1p 2
a + b2
2
1p 2
d=
a + c2
2
1p 2
b + c2
d=
2
d=

There are four nearest neighbors corresponding to the smaller two of
a, b, c.

96 Crystal Structure

(12.3) Crystal Structure
The diagram of Fig. 12.22 shows a plan view of a
structure of cubic ZnS (zincblende) looking down the z
axis. The numbers attached to some atoms represent the
heights of the atoms above the z = 0 plane expressed as a
fraction of the cube edge a. Unlabeled atoms are at z = 0
and z = a.
(a) What is the Bravais lattice type?
(b) Describe the basis.
(c) Given that a = 0.541 nm, calculate the nearestneighbor Zn–Zn, Zn–S, and S–S distances.

1
2

3
4

1
4

a

1
2

1
2

3
4

1
4

1
2

a
Zn=

S=

Fig. 12.23 Plan view of conventional unit cell of
zincblende.

(a) The lattice is FCC (otherwise known as ”cubic F”)
(b) The basis can be described as Zn at position [0,0,0] and S at
position [ 14 , 14 , 34 ] (or equivalently, but less standard notation [ 14 , 14 , −1
4 ]
all in units of the lattice constant.
√
(c) (just a bit of geometry:)pnearest neighbor Zn-Zn is a/ 2 = 0.383
2
2
2
nm; nearest neighbor
√ Zn-S is a 1/4 + 1/4 + 1/4 = 0.234 nm; nearest
neighbor S-S is a/ 2 = 0.383 nm;

(12.4) Packing Fractions
Consider a lattice with a sphere at each lattice point.
Choose the radius of the spheres to be such that neighboring spheres just touch (see for example, Fig. 12.8. The
packing fraction is the fraction of the volume of all of
space which is enclosed by the union of all the spheres

(i.e., the ratio of the volume of the spheres to the total
volume).
(a) Calculate the packing fraction for a simple cubic
lattice.
(b) Calculate the packing fraction for a bcc lattice.
(c) Calculate the packing fraction for an fcc lattice.

(a) The volume of a conventional unit cell is Vcell = a3 . Each cell corresponds to a single sphere. The radius of this sphere is a/2 so its volume
is Vsphere = 4π/3(a/2)3 . Thus the packing fraction is Vsphere /Vcell =
π/6 ≈ .52.

(b) The volume of a conventional unit cell is Vcell = a3 . Each conven√
tional unit cell contains two lattice points which are a distance a 3/2

97

√
apart from each other. Thus
√ the 3radius of the each sphere is a 3/4 so
has volume Vsphere√= 4π(a 3/4) /3. Thus we have a packing fraction
2Vsphere /Vcell = π 3/8 ≈ .68
(c) The volume of a conventional unit cell is Vcell = a3 . Each conven√
tional unit cell contains four lattice points which are a distance√a 2/2
apart from each other. Thus
√ the radius of the each sphere is a 2/4 so
has volume Vsphere =√4π(a 2/4)3 /3. Thus we have a packing fraction
4Vsphere /Vcell = π/(3 2) ≈ .74

(12.5) Fluorine Beta Phase
Fluorine can crystalize into a so-called beta-phase at
temperatures between 45 and 55 Kelvin. Fig. 12.24 shows
the cubic conventional unit cell for beta phase fluorine in

1
2

1
4

and

1
2

1
2

3
4

1
2

three-dimensional form along with a plan view.
 How many atoms are in this conventional unit cell?
 What is the lattice and the basis for this crystal?

Fig. 12.23

A conventional unit cell for fluorine beta
phase. All atoms in the picture are fluorine. Lines are
drawn for clarity Top: Three-dimensional view. Bottom:
Plan view. Unlabeled atoms are at height 0 and 1 in units
of the lattice constant.
1
4

and

3
4

1
2

There are 8 atoms in the conventional unit cell. 1 in the corner (8
times 1/8). Each atom on the face counts 1/2. Then there is one in the
center.
The lattice is simple cubic. The basis is
[0, 0, 0], [0, 1/2, 1/4], [0, 1/2, 3/4], [1/4, 0, 1/2], [3/4, 0, 1/2], [1/2, 1/4, 0], [1/2, 3/4, 0], [1/2, 1/2, 1/2]

Reciprocal Lattice,
Brillouin Zone, Waves in
Crystals

(13.1) Reciprocal Lattice
Show that the reciprocal lattice of a fcc (face-centered
cubic) lattice is a bcc (body-centered cubic) lattice. Correspondingly, show that the reciprocal lattice of a bcc
lattice is an fcc lattice. If an fcc lattice has conventional
unit cell with lattice constant a, what is the lattice con-

13

stant for the conventional unit cell of the reciprocal bcc
lattice?
Consider now an orthorhombic face-centered lattice
with conventional lattice constants a1 , a2 , a3 . What it
the reciprocal lattice now?

Brief Solution
Say we have an fcc lattice in real space. This can be written as cubic
with a basis R1 = [0, 0, 0], and R2 = [1/2, 1/2, 0] and R3 = [1/2, 0, 1/2]
and R4 = [0, 1/2, 1/2] in units of the lattice constant a. In reciprocal
space, we propose a basis for the reciprocal lattice which is S1 = (0, 0, 0)
and S2 = (1/2, 1/2, 1/2) in units of the reciprocal cubic lattice constant
4π/a. For these to be reciprocal, we must therefore have
ei2πRi ·Sj = 1
for all i and j which is easy to check is true. Further, we can show that
given the fcc basis there is no third point that can be added to the bcc
basis which would still have the same property. Similarly, given the bcc
basis there is no fifth point we can add to the fcc basis.
For the orthorhombic face centered system one would similarly obtain
a reciprocal lattice which is an orthorhombic body centered systems with
basis vectors 4π/ai .
Detailed Solution
The more straightforward way to find the reciprocal lattice of a direct
lattice is by construction. Let us start with the direct lattice of the BCC

100 Reciprocal Lattice, Brillouin Zone, Waves in Crystals

lattice which has primitive lattice vectors
a1

=

[1, 0, 0]a

a2
a3

=
=

[0, 1, 0]a
[1/2, 1/2, 1/2]a

we then construct the primitive lattice vectors of the reciprocal lattice
via
2πbf aj × ak
bi =
a1 · a2 × a3
with i, j, k cyclic. This gives
b1

=

b2

=

b3

=

4π
a
4π
(0, 1/2, −1/2)
a
4π
(0, 0, 1)
a
(1/2, 0, −1/2)

we can then reassemble these vectors to give the standard primitive
lattice vectors for FCC
b′1
b′2
b′3

4π
a
4π
= b2 + b3 =
(0, 1/2, 1/2)
a
4π
= b1 + b2 + b3 = (1/2, 1/2, 0)
a
= b1 + b3 =

(1/2, 0, 1/2)

which shows us tht the fcc lattice has lattice constant 4π/a. NOTE
about making this transformation from b into b′ . We are guaranteed
that that b are primitive lattice vectors. We assemble them together
with integer coefficients to make b′ . However we must also be able to
reassemble the b′ with integer coefficients to get back the b in order for
this to be an allowed transformation from one set of primitive lattice
vectors to another.
Slick Solution
Once one has read chapter 14 on scattering and selection rules, we
simply note that the selection rules on miller indices tell us the form
of the reciprocal lattice. For example, if the direct lattice is BCC, the
selection rule is that h + k + l is even. This means that either all h, k, l
are even or two are odd and one is even. Thus if we write


k
l
4π h
x̂ + ŷ + ẑ
G(hkl) =
a 2
2
2
we must have either all h/2, k/2, l/2 integer or two half-odd integer and
one integer. This is precisely the condition we would use to define an
FCC lattice. The prefactor 4π/a then gives us the lattice constant.

101

Similarly in reverse, if the direct space lattice is FCC, the selection
rule is h, k, l all even or all odd. This means (h/2, k/2, l/2) all integer or
all half-odd integer. This is precisely the condition for defining a BCC
lattice.
The moral of this story is that scattering occurs when (h, k, l) represents a reciprocal lattice vector. The selection rules tell you when this
is so.

(13.2) Lattice Planes
Consider the crystal shown in Exercise 12.3. Copy this

figure and indicate the [210] direction and the (210) family of lattice planes.

[210]

y

x

(210) planes
NOTE: As drawn here this is a family of planes (it is a family of
lattice planes for the corresponding simple cubic). For the FCC to be
a family of lattice planes, it must have half the plane spacing so that it
is called (420) and it cuts though every lattice point. So the question is
not correctly written as it stands.

(13.3) Directions and Spacings of Crystal
Planes
 ‡Explain briefly what is meant by the terms “crystal planes” and “Miller indices”.
 Show that the general direction [hkl] in a cubic
crystal is normal to the planes with Miller indices (hkl).
 Is the same true in general for an orthorhombic
crystal?

 Show that the spacing d of the (hkl) set of planes
in a cubic crystal with lattice parameter a is
d= √

a
h2 + k2 + l2

 What is the generalization of this formula for an
orthorhombic crystal?

102 Reciprocal Lattice, Brillouin Zone, Waves in Crystals

A crystal plane is a plane which intersects at least three non-colinear
(and therefore an infinite number of) points of the lattice.
Miller indices are a set of three integers which specify a set of parallel
planes (or equivalently specify a vector in reciprocal space). If the axes
of a lattice are mutually orthogonal , the Miller indices (hkl) specifies
the reciprocal lattice vector (2π)(hx̂/ax +kŷ/ay +lẑ/az where ax , ay and
az are the length of the three basis vectors. In terms of lattice planes,
one can determine the lattice vectors by picking any given plane from
the family, and finding its intersection points with the three coordinate
axes. These intersection points (kx , 0, 0) and (0, ky , 0) and (0, 0, kz ) have
the property that they are in the ratios
h:k:l=

1
1
1
:
:
kx ky kz

The actual values of h : k : l are then the smallest integer values with
these ratios. By convention note that negative numbers are denoted as
an integer with a bar on top, ex (1, 1, 1̄) to denote (1, 1, −1).
Let the basis vectors of the lattice be a, b and c, assumed to be orthogonal. Let the lengths of these three lattice vectors be a, b, c respectively.
The plane (hkl) can be defined as the plane connecting the points a/h,
b/k and c/l. To construct a vector normal to this plane, take any two
(noncolinear) vectors in this plane and take their cross product
n

=
=

(a/h − b/k) × (a/h − c/l)


k
l
abc h
a
+
b
+
c
hkl a2
b2
c2

This is only parallel to the vector [hkl] in the case of the cubic crystal
a = b = c.
Much more succinctly one could also note that for orthogonal axes,
bi = 2π/ai , and the family of lattice planes (hkl) is normal to the corresponding reciprocal lattice vector.
The unit normal in this direction is


k
l
1
h
a + 2b + 2c
n̂ = p
b
c
(h/a)2 + (k/b)2 + (l/c)2 a2

To find the inter-planar spacing, take any vector that connects two adjacent planes and find the component in the direction of the normal to
the plane. For example, the vector a/h connects two adjacent planes,
and we can resolve it parallel to n by taking the dot product with the
unit normal vector n̂. Thus we obtain
1
dhkl = n̂ · a/h = p
(13.1)
2
(h/a) + (k/b)2 + (l/c)2

with the case a = b = c appropriate for cubic crystals.
Note: In cases where the axes are not orthogonal, this formula does
not work.

103

(13.4) ‡Reciprocal Lattice
(a) Define the term Reciprocal Lattice.
(b) Show that if a lattice in 3d has primitive lattice
vectors a1 , a2 and a3 then primitive lattice vectors for
the reciprocal lattice can be taken as
a2 × a3
b1 = 2π
(13.2)
a1 · (a2 × a3 )
a3 × a1
b2 = 2π
(13.3)
a1 · (a2 × a3 )
a1 × a2
b3 = 2π
(13.4)
a1 · (a2 × a3 )

What is the proper formula in 2d?
(c) Define tetragonal and orthorhombic lattices. For an
orthorhombic lattice, show that |bj | = 2π/|aj |. Hence,
show that the length of the reciprocal lattice vector
G = hb1 + kb2 + lb3 is equal to 2π/d, where d is the
spacing of the (hkl) planes (see question 13)

(a) Given a lattice of points (in 3d it would look like R[uvw] = ua1 +
va2 + wa3 with [uvw] integers), the reciprocal lattice is defined by the
set of points in k space such that
eik·R = 1
for all points R in the lattice . Note that this set of points forms a lattice
of values of k.
(b) If we take
b1
b2
b3

a2 × a3
a1 · (a2 × a3 )
a3 × a1
= 2π
a1 · (a2 × a3 )
a1 × a2
= 2π
a1 · (a2 × a3 )
= 2π

(13.5)

Then we can show the key formula
bi · aj = 2πδij .
If these are our basis vectors for the reciprocal lattice, we then have a
general reciprocal lattice point given by
Ghkl = hb1 + kb2 + lb3

(13.6)

with h, k, l integers. It is trivial to then see that
eiK·R = 1

(13.7)

for any lattice vector R and any reciprocal lattice vector K. However,
this does not quite prove the desired statement. What it proves is that
vector K of the form of Eq. 13.6 are in the reciprocal lattice. We need
to show that there are no other vectors in the reciprocal lattice as well
so that b1 , b2 , and b3 can indeed be taken as the basis.

104 Reciprocal Lattice, Brillouin Zone, Waves in Crystals

To show this, consider an arbitrary vector K of the form of Eq. 13.6
but do not require that h, k, l are integers. Given an arbitrary Ruvw of
the real space lattice, in order that Eq. 13.7 is satisfied, we must have
uk + vh + wl = integer
for this to be true for arbitrary u, v, w which are integers, we can conclude
that k, h, l are integers.
We can derive the 2d analogous formula by setting a3 = ẑ the unit
vector normal to the plane, more conveniently written as
a2 × ẑ
b1 = 2π
ẑ · (a1 × a2 )
ẑ × a1
b2 = 2π
ẑ · (a1 × a2 )

(c) A tetragonal lattice is a lattice (in 3d) where all three basis vectors
are normal to each other, and two of them are the same length but the
third is a different length. An orthorhombic lattice is a lattice (in 3d)
where all three basis vectors are normal to each other and all three have
different lengths. (Note: crystals may have orthorhombic or tetragonal
symmetry even if lattice constants in the three directions are all equal.
The symmetry of a crystal has to do with whether it looks the same
under various types of rotations. An orthorhombic crystal does not look
the same under the three 90 degree rotations – this may be the case even
if all three lattice constants are the same).
For an orthorhombic crystal, without loss of generality, let us write
a1 = a1 x̂ and a2 = a2 ŷ and a3 = a3 ẑ. We then just use the above
formula Eq. 13.5 to obtain
a2 a3
ŷ × ẑ = 2πx̂/a1
b1 = 2π
a1 a2 a3
a2 a3
b2 = 2π
ẑ × x̂ = 2πŷ/a2
a1 a2 a3
a2 a3
b3 = 2π
x̂ × ŷ = 2πẑ/a3
a1 a2 a3
So |bi | = 2π/|ai |. Thus the length of G = hb1 + kb2 + lb3 is given by
p
p
|Ghkl | = h2 |b1 |2 + k 2 |b2 |2 + l2 |b3 |2 = (2πh/a1 )2 + (2πk/a2 )2 + (2πl/a3 )2 = 2π/dhk
as given in Eq. 13.1.

(13.5) More Reciprocal Lattice
A two-dimensional rectangular crystal has a unit cell
with sides a1 = 0.468 nm and a2 = 0.342 nm.
(a) Draw to scale a diagram of the reciprocal lattice.

 Label the reciprocal lattice points for indices in the
range 0 ≤ h ≤ 3 and 0 ≤ k ≤ 3.
(b) Draw the first and second Brillouin zones using the
Wigner–Seitz construction.

The reciprocal lattice vectors are b1 = 2πx̂/|a1 |, and b2 = 2πŷ/|a2 |.
Their magnitudes are
|b1 | = 13.4nm−1

|b2 | = 18.4nm−1

105

(a) A diagram of the reciprocal lattice is given in the figure.
12
32
42
012

032

042

42

32

12

042
032
012

Fig. 13.1 Figures for problem 13. First Two Brillouin Zones. Red is the 1st Zone.
Blue is the 2nd zone.

To find the Brillouin zones, one first constructs perpendicular bisectors
between the origin and any given lattice point (shown as dotted lines on
the plot). Then starting at the origin, the region one can get to without
crossing a dotted line is the first zone. Crossing only one dotted line
gets one to the second zone. etc.

(13.6) Brillouin Zones
(a) Consider a cubic lattice with lattice constant a.
Describe the first Brillouin zone. Given an arbitrary
wavevector k, write an expression for an equivalent
wavevector within the first Brillouin zone (there are several possible expressions you can write).

(b) Consider a triangular lattice in two dimensions
(primitive lattice vectors given by Eqs. 12.3). Find the
first Brillouin zone. Given an arbitrary wavevector k (in
two dimensions), write an expression for an equivalent
wavevector within the first Brillouin zone (again there
are several possible expressions you can write).

(a) The reciprocal lattice of a cubic lattice with lattice constant a
is a cubic lattice with lattice constant 2π/a. The first Brillouin zone
is a cube centered around the origin in reciprocal space (0,0,0), with
side length 2π/a. It spans kx , ky , kz ∈ [−π/a, π/a]. Given an arbitrary
k = (kx , ky , kz ) we can write the equivalent wavevector as (kx′ , ky′ , kz )′
where
kj′ = kj − [[(kj + π/a)/(2π/a)]] ∗ 2π/a

(13.8)

with j = x, y, z and [[ ]] is the floor function – meaning the greatest
integer less than its argument. The point of Eq. 13.8 is that it takes k in
units of the reciprocal lattice constant (2π/a) and returns a number between -1/2 and +1/2 of the reciprocal lattice constant by appropriately

106 Reciprocal Lattice, Brillouin Zone, Waves in Crystals

adding or subtracting integer units of 2π/a.
(b) Start with the primitive lattice vectors
a1
a2

=
=

ax̂
√
(a/2)x̂ + (a 3/2)ŷ

We need to find vectors b1 , b2 such that ai · bj = 2πδi,j . One way to
do this is to use the usual 3d formula
a2 × a3
a1 · (a2 × a3 )
a3 × a1
b2 = 2π
a1 · (a2 × a3 )
a1 × a2
b3 = 2π
a1 · (a2 × a3 )
√
then assigning a3 = ẑ. The denominator is a2 3/2. Thus we have
√
√
√
b1 = 2π[(a 3/2)x̂ − (a/2)ŷ]/(a2 3/2) = (2π/a)[x̂ − ( 3/3)ŷ]
√
√
b2 = 2πŷa/(a2 3/2) = (2π/a)ŷ(2 3/3)
b1

= 2π

These form the primitive lattice vectors of a triangular lattice in k-space.
The first Brillouin zone is the Wigner-Seitz cell of the reciprocal lattice
— which is a hexagon centered around zero wavevector (see fig 12.6 of
the book).
Two write a formula that translates any k into the first Brillouin zone,
we first write any k point in terms of the primitive lattice vectors
k = α1 b1 + α2 b2
We are going to need to solve for αi in terms of k. To do this we write
k · b1

k · b2

= α1 |b1 |2 + α2 b1 · b2
= α1 b1 · b2 + α2 |b2 |2

and we will want to solve for the αs. In this particular case, things are
quite easy since |b1 |2 = |b2 |2 = −2b1 · b2 = (2π/a)2 (4/3). So this
becomes
βk · b1
βk · b2

= α1 − α2 /2
= −α1 /2 + α2

where we have defined
β=

3 a2
4 (2π)2

Solving this system of equations we obtain
α1
α2

=
=

(β/3)(2k · b1 + k · b2 )
(β/3)(k · b1 + 2k · b2 )

107

Now to obtain α within the first brillouin zone we need to shift both
α’s by integers until they lie in the range [−1/2, 1/2]. To do this we
write
α′j = αj − [[αj + 1/2]]

again with [[ ]] being the floor function. Once we have α′ ’s we construct
k′ = α′1 b1 + α′2 b2
which is now within the first Brillouin zone.

(13.7) Number of States in the Brillouin Zone
A specimen in the form of a cube of side L has a primitive cubic lattice whose mutually orthogonal fundamental
translation vectors (primitive lattice vectors) have length
a. Show that the number of different allowed k-states

within the first Brillouin zone equals the number of primitive unit cells forming the specimen. (One may assume
periodic boundary conditions, although it is worth thinking about whether this still holds for hard-wall boundary
conditions as well.)

We are given a cubic lattice with lattice constant a and an overall size
L on a size. The number of atoms in this sample are N = (L/a)3 . It
is simplest to assume periodic boundary conditions. Since eikx (x+L) =
eikx x we must have kx = 2πn/L for n an integer. Thus (considering all
three directions) there is one eigenstate per volume (2π/L)3 in k-space.
The first Brillouin zone extends from −π/a ≤ kx , ky , kz ≤ π/a. This
has volume in k-space of (2π/a)3 . Dividing this by the volume occupied
by one eigenstate gives a total of (L/a)3 states in the first Brillouin zone.
This agreement is not coincidental. If one thinks in a tight-binding or
atomic orbital picture, there should be exactly one orbital per atom per
band when hopping is turned off. Once hopping is turned back on, each
atomic orbital spreads into a band that fills the Brillouin zone, but the
total number of states stays constant.
One can also consider the same problem with hard wall boundary
conditions, although it is less convenient. In this case the eigenstates
are not of the form eikx but are rather sin(kx) where k > 0, so one is
not really talking about pure plane waves. In this case the analogue of
the Brillouin zone goes from 0 ≤ kx , ky , kz ≤ π/a, but the density of
k states is doubled, k = π/L in each direction, so the total number of
states remains the same as using periodic boundary conditions.

108 Reciprocal Lattice, Brillouin Zone, Waves in Crystals

(13.8) Calculating Dispersions in d > 1*
(a) In Exercises 9.8 and 11.9 we discussed dispersion
relations of systems in two dimensions (if you have not
already solved those exercises, you should do so now).
 In Exercise 11.9 describe the Brillouin zone (you
may assume perpendicular lattice vectors with length a1
and a2 ). Show that the tight-binding dispersion is periodic in the Brillouin zone. Show that the dispersion curve
is always flat crossing a zone boundary.

 In Exercise 9.8, describe the Brillouin zone. Show
that the phonon dispersion is periodic in the Brillouin
zone. Show that the dispersion curve is always flat crossing a zone boundary.
(b) Consider a tight binding model on a threedimensional fcc lattice where there are hopping matrix
elements −t from each site to each of the nearest-neighbor
sites. Determine the energy spectrum E(k) of this model.
Show that near k = 0 the dispersion is parabolic.

(a) [First see solutions for 9.8 and 11.9]
In Exercise 11.9, for a rectangular lattice with lattice vectors of length
a1 and a2 the reciprocal lattice is rectangular with lattice vectors of
length b1 = (2π)/a1 and b2 = (2π)/a2 . The Brillouin zone is hence a
rectangle extending between kx ∈ [−π/a1 , π/a1 ] and ky ∈ [−π/a2 , π/a2 ].
Given the derived dispersion
E = ǫ0 − 2t1 cos(kx a1 ) − 2t2 cos(ky a2 )
at the zone boundary, the cos is at its maximum, and its derivative is
zero. Thus the dispersion is always flat approaching the zone boundary.
In Exercise 9.8 we have a triangular lattice, whose reciprocal lattice is
also triangular and has a Wigner-Seitz cell which is hexagonal (see exercise 13.6.b also). Note also the figures shown in the solution of 9.8 which
clearly show the hexagonal periodicity of the dispersion curve. Note that
the dispersion must be symmetric as k approaches the zone boundary.
Let the zone boundary wavevector be bi /2 + k⊥ where k⊥ · bi = 0.
Imagine approaching the zone boundary via k = αbi /2 + k⊥ and taking
α to one. Note that the dispersion must be symmetric around α = 1
since (by reflection symmetry of the problem) the frequency at k is the
same as the frequency at k′ = −αbi /2 + k⊥ . However, one can translate
this by a reciprocal lattice vector to get (2 − α)bi /2 + k⊥ which is the
same point reflected around the zone boundary. Thus, so long as the
dispersion does not have a cusp at the zone boundary, it must have zero
derivative.
Recall from Eq. 9.8 that the derived freuquency spectrum is


q
2κ
2
2
2
2
S1 + S2 + S3 ± S1 + S2 + S3 − S1 S2 − S1 S3 − S2 S3
ω =
m
where

k · ai
)
2
where ai are three independent vectors along the lattice. In terms of
primitive lattice vectors here we can take a1 and a2 to be primitive
lattice vector, and then a3 = a2 − a1 .
Si = sin2 (

109

The frequency is an analytic function in k so long as the argument of
the square-root is nonzero. By rewriting the argument as

1
(S1 − S2)2 + (S2 − S3)2 + (S3 − S1)2
2

it is clear the argument must be nonzero except at k = 0. Thus we can
conclude that the dispersion is analytic and symmetric around the zone
boundary and therefore must have zero slope. This may seem like a bit
of a cheat, but it is perfectly rigorous.
For those who would prefer to see this statment proven a bit more
directly, we take the following approach. Let us start with the primitive
lattice vectors
a1
a2

= ax̂

√
= (a/2)x̂ + (a 3/2)ŷ

Recall that the reciprocal lattice vectors (see 13.6.b) are given by
√
√
√
b1 = 2π[(a 3/2)x̂ − (a/2)ŷ]/(a2 3/2) = (2π/a)[x̂ − ( 3/3)ŷ]
√
√
b2 = 2πŷa/(a2 3/2) = (2π/a)ŷ(2 3/3)
For simplicity let us imagine approachig the zone boundary
k = αb2 /2 + k⊥
(the other zone boundaries will be the same by symmetry). Here the
direction orthogonal to b2 is b1 + b2 /2. So we can write k⊥ = β(b1 +
b2 /2) for some value of β. We then have (using ai · bj = 2πδij for
i, j ∈ 1, 2)
k · a1

k · a2

k · a3

=

(β + α)π

=

β2π

=

(β − α)π

Now let us examine the terms in the dispersion. In particular we are
concerned with the dependence of the dispersion on α near α = 1. Again
all we need to do is to show that the dispersion is symmetric and analytic
around α = 1. Here S2 is independent of α so we are not concerned with
that piece. Using sin2 (x/2) = (1 − cos(x))/2 we have that
S1

=

S2

=

[1 − cos((β + α)π)]/2

[1 − cos((β − α)π)]/2

The cos is periodic in α → α ± 2, and reflection of α around 1 (i.e.,
α → 2 − α) turns S1 into S2 , hence leaves the freqeuncy unchanged.
Thus we have the dispersion symmetric and analytic around the zone
boundary, thus having zero slope.
(b) For an fcc lattice, if the conventional unit cell has lattice constant
a, the vectors to the 12 nearest neighbors of a lattice point are given by

110 Reciprocal Lattice, Brillouin Zone, Waves in Crystals

(see exercise 12.2)
[±1, ±1, 0]a/2

[±1, 0, ±1]a/2
[0, ±1, ±1]a/2

Thus we have the schroedinger equation (setting the onsite energy ǫ0 to
zero for simplicity)
X
Eψ(r) = −t
ψ(r + u)
u

where the sum is over u being these 12 vectors. Using the usual ansatz
ψ(r) = Aeik·r we obtain
i
h
X
E = −t
eik·[α,β,0]a/2 + eik·[α,0,β]a/2 + eik·[0,α,β]a/2
α=±1,β=±1

or
E = −4t [cos(kx a/2) cos(ky a/2) + cos(kx a/2) cos(kz a/2) + cos(ky a/2) cos(kz a/2)]
Expanding around k = 0 to second order we obtain


E = −12t + t kx2 + ky2 + kz2 (a2 ) + . . .
which is parabolic as claimed.

Wave Scattering by
Crystals

(14.1) Reciprocal Lattice and X-ray Scattering
Consider the lattice described in Exercise 13.5 (a twodimensional rectangular crystal having a unit cell with
sides a1 = 0.468 nm and a2 = 0.342 nm). A collimated
beam of monochromatic X-rays with wavelength 0.166 nm
is used to examine the crystal.
(a) Draw to scale a diagram of the reciprocal lattice.

14
(b) Calculate the magnitude of the wavevectors k and
k′ of the incident and reflected X-ray beams, and hence
construct on your drawing the “scattering triangle” corresponding to the Laue condition ∆k = G for diffraction
from the (210) planes (the scattering triangle includes k,
k′ and ∆k).

A diagram of the reciprocal lattice is given in the figure. The scatter-

42
32
12

012

12

32

42

012
Fig. 14.1 A picture of the reciprocal Lattice with the scattering triangle for the
(210) reciprocal wavevector.

ing triangle is the triangle such that |k| = |k′ | and k + G = |k′ | (Note
for fixed G there are two such possible triangles).

112 Wave Scattering by Crystals

(14.2) ‡ X-ray scattering II
BaTiO3 has a primitive cubic lattice and a basis with
atoms having fractional coordinates
Ba
Ti
O

[0,0,0]
[ 12 , 21 , 12 ]
[ 12 , 21 , 0],

[ 12 , 0, 12 ],

[0, 12 , 21 ]

 Sketch the unit cell.
 Show that the X-ray structure factor for the (00l)

Bragg reflections is given by
h
i
S(hkl) = fBa + (−1)l fT i + 1 + 2(−1)l fO

where fBa is the atomic form factor for Ba, etc.
 Calculate the ratio I(002) /I(001) , where I(hkl) is the
intensity of the X-ray diffraction from the (hkl) planes.
You may assume that the atomic form factor is proportional to atomic number (Z), and neglect its dependence
on the scattering vector. (ZBa = 56, ZTi = 22, ZO = 8.)

01
23
4
Fig. 14.2 Unit cell of BaTiO3

The X-ray structure factor is given by
X
X
S(hkl) =
fd eik(hkl) ·Rd =
fd e2πi(hxd +kyd +lzd )
d

d

where Rd = (xd , yd , zd ) are the positions of atom d in the unit cell, and
fd it the corresponding form factor (which we take to be proportional
to Zd ). If we are interested in (00l) we set h = k = 0 and obtain
S(00l) = fBa + (−1)l fT i + [1 + 2(−1)l ]fO
The Bragg peak intensity is proportional to the square of the structure
factor (times a multiplicity factor, but the multiplicity for all (00l) are
the same!), thus we obtain
2

(fBa + fT i + 3fO )
I002
≈ 15.4
=
I001
(fB a − fT i − fO )2
In reality the form factor depends on the scattering vector, and the
variation is different from each atom, so this is just an approximation.

113

(14.3) ‡ X-ray scattering and Systematic Absences
(a) Explain what is meant by “Lattice Constant” for a
cubic crystal structure.
(b) Explain why X-ray diffraction may be observed in
first order from the (110) planes of a crystal with a bodycentered cubic lattice, but not from the (110) planes of a
crystal with a face-centered cubic lattice.
 Derive the general selection rules for which planes
are observed in bcc and fcc lattices.
(c) Show that these selection rules hold independent of
what atoms are in the primitive unit cell, so long as the
lattice is bcc or fcc respectively.
(d) A collimated beam of monochromatic X-rays of

wavelength 0.162 nm is incident upon a powdered sample
of the cubic metal palladium. Peaks in the scattered Xray pattern are observed at angles of 42.3◦ , 49.2◦ , 72.2◦ ,
87.4◦ , and 92.3◦ from the direction of the incident beam.
 Identify the lattice type.
 Calculate the lattice constant and the nearestneighbor distance.
 If you assume there is only a single atom in the basis does this distance agree with the known data that the
density of palladium is 12023 kg m−3 ? (Atomic mass of
palladium = 106.4.)
(e) How could you improve the precision with which
the lattice constant is determined. (For one suggestion,
see Exercise 14.10.)

(a) For a cubic lattice, the lattice constant is the distance between
one lattice point and the nearest neighbor lattice point.
(b) The (110) planes of a body-centred cubic lattice contain all the
lattice points, whereas the (110) planes of a crystal with a face-centred
cubic lattice contain only half the lattice points. The remaining fcc
lattice points lie on a set of planes half-way in between the (110) planes,
and so X-rays reflected from these planes interfere destructively with
X-rays reflected from the (110) planes.
Let us see this more analytically now:
We view both the bcc and fcc lattices as being a cubic lattice with a
basis. For a cubic lattice, we must have khkl = (2π/a)(hx̂ + kŷ + lẑ) (or
in other words we consider reciprocal lattice vectors (h, k, l).)
Generally we sum over lattice points Ri in a unit cell to get the structure factor
X
eiRi ·k(hkl)
S(hkl) ∼
Ri

The bcc lattice is a cubic lattice with a basis (0,0,0) and (1/2,1/2,1/2).
So we obtain
bcc
S(hkl)
∼ 1 + eik(hkl) ·[(1/2)x̂+(1/2)ŷ+(1/2)ẑ] = 1 + (−1)h+k+l

This is only nonzero when h + k + l = even.
The fcc lattice is a cubic lattice with with basis (0,0,0) and (1/2,1/2,0)
and (1/2,0,1/2) and (0,1/2,1/2). Analogously we obtain
f cc
S(hkl)
∼ 1 + (−1)h+k + (−1)h+l + (−1)k+l

This is only nonzero when h, k, l are either all even or all odd.
Note that these selection rules hold for an orthorhombic lattice as
well as for cubic (nowhere did we use the lattice constants in the three

114 Wave Scattering by Crystals

directions!).
(c) This is the general principle that you multiply a lattice by a basis.
Suppose we have a Bravais lattice (either fcc or bcc) and then we have
a basis of m atoms which are at positions rj (j = 1, . . . m) with respect
to the lattice points. These atoms may have scattering form factors fj
as well.
As above, we treat the Bravais lattice as a cubic lattice with a basis
Ri (i=1,2 for bcc and i = 1 . . . 4 for fcc). So positions of all of the atoms
(2m or 4m of them) in the full cubic unit cell can be written as
Rij = Ri + rj
Writing the structure factor
S(hkl) =

X

fj e

ik(hkl) Rij

i,j

=

X

fj

j

"
X
i

e

ik(hkl) Ri

#

eik(hkl) rj

(14.1)

The factor inside the brackets is exactly the structure factor for the Bravais lattice as calculated above. So if the fcc structure factor vanishes,
then the structure will also vanish for any lattice having an fcc lattice
with any basis.
(d) Note that the angle given is 2θ (the full deflection angle)! We are
given 2θ then we use Bragg’s law λ = 2d sin θ to get the distance between
lattice planes. We then examine the ratio’s between these distances.
√
d = λ/(2 sin θ) (dmax /d)2 aestimate = d h2 + k 2 + l2
2θ
42.3◦
49.2◦
72.2◦
87.4◦
92.3◦

.224
.195
.137
.117
.112

nm
nm
nm
nm
nm

1
1.33
2.66
3.67
4.00

.389
.389
.389
.389
.389

nm
nm
nm
nm
nm

we see that the values of 1/d2 are in ratios 3:4:8:11:12. Recall (see
problem 13.3) that for a cubic lattice (a2 /d2 ) = h2 + k 2 + l2 . So the
observed peaks are of the types (1,1,1), (2,0,0), (2,2,0), (1,1,3), and
(2,2,2). These are all even or all odd, which is characteristic of the fcc
lattice. All measured scattering points give the same estimate of the
lattice constant a to three digits (which is the same number of digits
that the measurement contains... so no higher accuracy estimate can be
made from this data).
.1064 kg
m3
= 1.4682 × 10−29 m3 /atom
12023 kg 6.022 × 1023 atoms
Then for an fcc there are 4 atoms per unit cell, so the volume of the unit
cell is 5.872 × 10−29 m3 . The cube root of this gives 3.887 nm which is
in good agreement. An equivalent calculation is to calculate the density

115

based on the lattice constant

3
4 ∗ .1064 kg
1
= 12006 kg/m3
6.022 × 1023 atoms .389 × 10−9 m
in good agreement.

(e) to improve the precision of the measurement one would need to obtain more “digits” of resolution in the measurement of the angle. There
are several things that will help this. From Bragg’s law d ∼ 1/ sin θ so

1 ∂d
= cot θ
d ∂θ
So if there is an error in measurement of θ one ends up with a fractional
error in d given by
δ log d = (cot θ)(δθ)

As a result, error is minimized if one can work as close as possible to
scattering angle of 90 degrees (See exercise 14.10).
In addition some experimental issues can be listed. A higher brightness source will frequently reduce the noise and make the signal easier to
analyze. Better columnation will make measurements of angle more precise. At least as important is the calibration of the device that measures
the angles! Defining the wavelength λ more precisely becomes necessary
at some point. Temperature control is necessary since lattice constants
do change as a function of temperature due to thermal expansion.

(14.4) ‡ Neutron Scattering
(a) X-ray diffraction from sodium hydride (NaH) established that the Na atoms are arranged on a face-centered
cubic lattice.
 Why is it difficult to locate the positions of the H
atoms using X-rays?
The H atoms were thought to be displaced from the Na
atoms either by [ 14 , 41 , 41 ] or by [ 21 , 12 , 21 ], to form the ZnS
(zincblende) structure or NaCl (sodium chloride) structure, respectively. To distinguish these models a neutron
powder diffraction measurement was performed. The intensity of the Bragg peak indexed as (111) was found to
be much larger than the intensity of the peak indexed as
(200).

 Write down expressions for the structure factors
S(hkl) for neutron diffraction assuming NaH has
(i) the sodium chloride (NaCl) structure
(ii) the zinc blende (ZnS) structure.
 Hence, deduce which of the two structure models is correct for NaH. (Nuclear scattering length of Na
= 0.363 × 105 nm; nuclear scattering length of H =
−0.374 × 105 nm.)
(b) How does one produce monochromatic neutrons for
use in neutron diffraction experiments?
 What are the main differences between neutrons
and X-rays?
 Explain why (inelastic) neutron scattering is well
suited for observing phonons, but X-rays are not.

[This problem is based on a classic experiment by Shull et al, who
later won a Nobel prize for his work on neutron scattering]
Note, the scattering length should be in units of 10−5 nm not 105 nm.
It doesn’t change the problem much though.
It is difficult to see H atoms with X-rays since the form factor (amplitude) of scattering is proportional to the charge of the nucleus (the
atomic number), and the charge of H is very small.

116 Wave Scattering by Crystals

From Eq. 14.1 above we have
S(hkl) =

X
i


!
X
bj eik(hkl) rj 
eik(hkl) Ri 
j

where j are the elements of the basis and i are the 4-elements of the fcc
lattice basis when fcc is viewed as a cubic lattice with a basis. Here I have
replaced fj by bj which is the conventional notation for the scattering
length (which is analogous to the form factor in X-rays) in neutron
scattering. Note that the first factor in brackets is the scattering from
the fcc lattice which always gives (See problem 14)
f cc
Shkl
= (1 + eiπ(h+k) + eiπ(k+l) + eiπ(h+l) )

For the case of the NaCl structure, we have a basis of Na at [0,0,0]
and H at [1/2,1/2,1/2]. So we obtain
S(hkl) = (1 + eiπ(h+k) + eiπ(k+l) + eiπ(h+l) )(bN a + bH eiπ(h+k+l) )
For the case of the ZnS structure we have a basis of Na at [0,0,0] and
H at [1/4,1/4,1/4] so we obtain
S(hkl) = (1 + eiπ(h+k) + eiπ(k+l) + eiπ(h+l) )(bN a + bH ei(π/2)(h+k+l) )
The powder scattering intensity is proportional to the structure factor
squared times a multiplicity factor that counts how many symmetry related vectors give the same scattering pattern. The multiplicity of (111)
is 8 (since each factor of 1 could have been 1̄), whereas the multiplicity
of (200) is 6 since the 2 could be of either sign, and any one of the 3
entries could contain the 2.
Thus we obtain in the NaCl case a ratio of scattering intensities
I(111)
(bN a − bH )2 8
=
= 6000
I(200)
(bN a + bH )2 6
whereas in the ZnS case we have
I(111)
(b2 + b2H ) 8
= Na
= 0.67
I(200)
(bN a − bH )2 6
Thus we conclude that NaH has NaCl structure.
(b) Two typical schemes for producing monochromatic neutrons. (1)
time of flight (a ”chopper”) can select a particular velocity (2) Bragg
scattering of a given angle off of a crystal can select a particular wavelength.
The main things to know about the differences between X-rays and
neutrons are:
X-rays are scattered by electrons. The electromagnetic interaction is
relatively strong, and so X-rays typically penetrate only a few microns
into a sample. The intensity of scattering varies with the atomic form

117

factor f (∆k), which approximately scales with the number of electrons
in the atom.
Neutrons, on the other hand, are electrically neutral, and so interact
weakly with matter. They penetrate typically a few cm into materials. Neutrons may scatter from nuclei (via the strong nuclear force)
or unpaired electrons (via the magnetic dipole-dipole interaction). The
diffraction theory is the same as for X-rays, except that for scattering
from nuclei the atomic form factor f (∆k) is replaced by the nuclear scattering length b, which is independent of ∆k. (Note, we typically make
the approximation that f is independent of ∆k, to make calculations
easier, but it is not really such a good approximation. In the neutron
case, it really is a good approximation because the scattering is essentially off of a point nucleus). The scattering length b varies irregularly
from nucleus to nucleus so that the scattering from light elements is similar in strength to that from heavy elements. Neutrons can thus easily
see small atoms, and can easily distinguish atoms with similar atomic
number.
Magnetic neutron diffraction can also be used to determine magnetic
structures (in which case there is a form factor similar to that for Xrays).
The reason inelastic scattering of neutrons is more useful than X-rays
for studying phonons is because the X-ray velocity (velocity of light)
is huge. This makes it very difficult to conserve both momentum and
energy in any inelastic process involving light since a very small momentum change corresponds to a huge energy change. Although with
modern tools measurement of phonons using X-rays is now possible.

(14.5) And More X-ray Scattering
A sample of aluminum powder is put in an Debye–
Scherrer X-ray diffraction device. The incident X-ray radiation is from Cu–Ka X-ray transition (this just means
that the wavelength is λ = .154 nm). The following scattering angles were observed:

19.48◦ 22.64◦ 33.00◦ 39.68◦ 41.83◦ 50.35◦ 57.05◦ 59.42◦
Given also that the atomic weight of Al is 27, and the
density is 2.7 g/cm3 , use this information to calculate
Avagadro′ s number. How far off are you? What causes
the error?

Note: The angles listed are θ not 2θ here. If you use 2θ by mistake

118 Wave Scattering by Crystals

you will discover that the ratios of d values do not make sense!
√
θ
d = λ/(2 sin θ) (dmax /d)2 aestimate = d h2 + k 2 + l2
19.48◦
22.64◦
33.00◦
39.68◦
41.83◦
50.35◦
57.05◦
59.42◦

.2309
.2000
.1414
.1206
.1155
.1000
.0918
.0894

nm
nm
nm
nm
nm
nm
nm
nm

1
1.33
2.67
3.66
4.00
5.33
6.33
6.66

.3999
.4001
.3999
.4000
.4000
.4000
.4000
.4000

nm
nm
nm
nm
nm
nm
nm
nm

From the third column we see that we have an fcc lattice. From the
final column we conclude the conventional lattice constant is about .4000
nm. There are four atoms per unit cell, so we would predict an atomic
density of 4/(4nm)3 = 6.28 × 1028m−3 . On the other hand, with atomic
weight of 27 and density 2.7 g/cm3 we predict an atomic density of
NA × 105 m−3 with NA = 6.022 × 1023 being Avagadro’s number. (Or
in other words, our slighly inaccurate prediction of Avagadro’s number
is 6.28 × 1023 ). So we are off by abour 4%. Where does the error come
from? There are a few possible places. First, the atomic mass is not
quite 27, it is actually about 26.982 (depending a bit on isotopic abundances, but this is typical). But this is an error of much less than an
percent. Probably the biggest possible error is that the density changes
as a function of temperature. If the density is measured at one temperature, but the lattice constant is measured at another temperature,
then there can be a substantial disagreement. In fact the linear thermal
expansion coefficient of aluminum is about (22 × 10( − 6))K−1 which
would give a density different between T = 0 and room temperature of
a few percent (and indeed, the measurements could even be done above
room temperature!). In the literature X-ray measurements of the lattice
constant of alumium give numbers of about .404 nm at room temperature (which would match the stated density much more closely). Thus I
suspect the data given above was measured at low temperature, hence
the disagreement.

(14.6) More Neutron Scattering
The conventional unit cell dimension for a particular
bcc solid is .24nm. Two orders of diffraction are observed.

What is the minimum energy of the neutrons? At what
temperature would such neutrons be dominant if the distribution is Maxwell–Boltzmann.

For bcc, the lowest two order diffraction peaks are the (110) and the
(200). The corresponding plane spacings are
p
d = a/ h2 + k 2 + l2

119

√
Or d = a/ 2 and d = a/2. Using Bragg’s law,
d = λ/(2 sin θ)
the wavelength must be at least as small as 2d in order to see a particular
peak. Or in other words λ ≤ a = .24nm. The corresponding energy of
neutrons is
~2 k 2
~2 (2π)2
E=
=
= 2.28 × 10−21 J
2mn
2mn λ2
and
E/kB = 165K

(14.7) Lattice and Basis
Prove that the structure factor for any crystal (described with a lattice and a basis) is the product of the

structure factor for the lattice times the structure factor
for the basis (i.e., prove Eq. 14.14).

We have the definition of the structure factor
X
S(hkl) =
fj e2πi(hxj +kyj +lzj )
atoms j in (conventional) unit cell

When we have a lattice and a basis, all of the positons rj of atoms j in
the unit cell can be written as the sum of a lattice point Ra and a basis
vector uα . Thus we can write
rj = Ra + uα
and the sum over j becomes a sum over a and α. Thus we have
X
S(hkl) =
fα e2πi(h(Rx,a +ux,α )+k(Ry,a +uy,α )+l(Rz,a +uz,α ))
α,a

Note that the atom type depends only on the basis vector not on the
lattice point. We split this sum into two
X
X
S(hkl) =
fα e2πi(hux,α )+kuy,α +luz,α )
e2πi(hRx,a +kRy,a +lRz,a )
α

a

which is simply the product of basis and lattice structure factors
basis lattice
S(hlk)
S(hkl) = S(hkl)

120 Wave Scattering by Crystals

(14.8) Cuprous Oxide and Fluorine Beta
(a) The compound Cu2 O has a cubic conventional unit
cell with the basis:
O
Cu

[000] ; [ 12 , 21 , 12 ]
[ 14 , 41 , 14 ] ; [ 14 , 34 , 34 ] ; [ 43 , 14 , 34 ] ; [ 43 , 34 , 41 ]

type? Show that certain diffraction peaks depend only
on the Cu form factor fCu and other reflections depend
only on the O form factor fO .
(b) Consider fluorine beta phase as described in exercise 12.5. Calculate the structure factor for this crystal.
What are the selection rules?

Sketch the conventional unit cell. What is the lattice

(a) the conventional unit cell of Cu2 O is shown in Figure 14.3.
The structure factor is
S(hkl)

= fO (1 + eiπ(h+k+l) )
h π
i
π
π
π
+fCu ei 2 (h+k+l) + ei 2 (h+3k+3l) + ei 2 (3h+k+3l) + ei 2 (3h+3k+l)

Note that sqaure bracketed expression can be simplified to
i
h
π
ei 2 (h+k+l) 1 + eiπ(k+l) + eiπ(h+l) + eiπ(k+l)

Fig. 14.3 Conventional unit cell of
Cu2 O. The darker atoms at the corners
and the center are O and the lighter colored atoms are Cu.

Note that the coefficient of f0 vanishes unless h + k + l is even. Indeed,
this tells us that the oxygens alone form a bcc lattice. The coefficient of
the fCu vanish unless h, k, l are all even or all odd, which tells us that
the Cu alone form an fcc lattice (perhaps less obvious!). Thus peaks
such as (110) depend only on fO wheras peaks such as (111) depend
only on fCu .
(b) From 12.5 the basis of Flourine beta is
[0, 0, 0], [0, 1/2, 1/4], [0, 1/2, 3/4], [1/4, 0, 1/2],
[3/4, 0, 1/2], [1/2, 1/4, 0], [1/2, 3/4, 0], [1/2, 1/2, 1/2]
The structure factor is then
1 + eiπk+iπl/2 + eiπk+iπ3l/2 + eiπl+iπh/2 + eiπl+iπ3h/2
+eiπh+iπk/2 + eiπh+iπ3k/2 + eiπ(h+k+l)
which can be simplified to
1 + (−1)h+k+l + (−1)k [il + i−l ] + (−1)l [ih + i−h ] + (−1)h [ik + i−k ]
Note that the first two terms cancel
give 2. Let us examine the last three

 0
2
ik + i−k =

−2

if h + k + l is odd and otherwise
terms. Note that
l = odd
l = 0 mod 4
l = 2 mod 4

121

Thus we obtain the following rules for (hkl) where the structure factor
does not vanish
one even two odd if the even one is 2 mod 4
one odd two even if the two even indices are not the same mod 4
all even with either one or all three indices being 2 mod 4

(14.9) Form Factors
(a) Assume that the scattering potential can be written
as the sum over the contributions of the scattering from
each of the atoms in the system. Write the positions of
the atoms in terms of a lattice plus a basis so that
X
Vα (x − R − yα )
V (x) =
R,α

where R are lattice points, α indexes the particles in the
basis and yα is the position of atom α in the basis. Now
use the definition of the structure factor Eq. 14.5 and de-

(a) Start with 14.5
Z
XZ
iG·x
dx e
V (x) =
S(G) =
unitcell

R,α

unitcell

rive an expression of the form of Eq. 14.8 and hence derive
expression 14.9 for the form factor. (Hint: Use the fact
that an integral over all space can be decomposed into a
sum over integrals of individual unit cells.)
(b) Given the equation for the form factor you just derived (Eq. 14.9), assume the scattering potential from an
atom is constant inside a radius a and is zero outside that
radius. Derive Eq. 14.10.
(c)* Use your knowledge of the wavefunction of an electron in a hydrogen atom to calculate the X-ray form factor
of hydrogen.

dx eiG·x Vα (x − R − yα )

we then using eiG·R = 1 we have
XZ
S(G) =
dx eiG·(x−R) Vα (x − R − yα )
R,α

unitcell

Now we note that the integral dx over a unit cell summed over all
possible unit cells indexed by R is equivalent to a single integral over all
of space, so we have
XZ
XZ
iG·x
S(G) =
dx e
Vα (x − yα ) =
dx eiG·(x+yα ) Vα (x)
α

=

X

α

e

iG·yα

fα (G)

α

with

fα (G) =
as required.

Z

dx eiG·x Vα (x)

(b) Assuming the potential is V0 inside a radius a and zero outside we
have
Z radius=a
Z
Z a
iG·x
2
fα (G) = V0
dx e
= 2πV0
r dr sin θdθeirGcosθ
0

0

122 Wave Scattering by Crystals

where θ is the angle between x and G. So we have
Z a
Z 1
Z
2
irGz
fα (G) = 2πV0
r dr
dze
= 2πV0
0

−1

a

r2 dr

0

2 sin(rG)
rG

sin(x) − x cos(x)
= 4πV0 a3
x3

where x = Ga. Now comparing this to 14.10, it looks identical except for
the prefactor. Note however, that the charge of atom is spread out over
a volume 4πa3 /3. So setting V0 4πa3 /3 = Z makes the two equations
match.
(c) The normalized wavefunction for an electron in the ground state
of a hydrogen atom is
2 −r/a0
1
ψ=√
e
3/2
4π a0
with a0 the Bohr radius. The scattering potential is proportional to the
electron density |ψ|2 . Let us call the constant of proportionality K. We
then have (using similar calculation as above)
Z
Z ∞
2 sin(rG)
|ψ(r)|2
f (G) = K dx eiG·x |ψ(r)|2 = K2π
r2 dr
rG
0
So we want

Z
2K ∞ 2 2 sin(rG) −2r/a0
e
r dr
a30 0
rG
The integration is not too difficult and gives the result
f (G) =

f (G) =

(14.10) Error Analysis
Imagine you are trying to measure the lattice constant
a of some crystal using X-rays. Suppose a diffraction peak
is observed at a scattering angle of 2θ. However, suppose

16K
(a20 G2

2

+ 4)

that the value of θ is measured only within some uncertainty δθ. What is the fractional error δa/a in the resulting measurement of the lattice constant? How might this
error be reduced? Why could it not be reduced to zero?

See 14.3.e.
(e) From Bragg’s law d ∼ 1/ sin θ so
1 ∂d
= cot θ
d ∂θ
So if there is an error in measurement of θ one ends up with a fractional
error in d given by
δ log d = (cot θ)(δθ)
As a result, error is minimized if one can work as close as possible to
scattering angle of 90 degrees. Although at 90 degrees one could have

123

this expression be exactly zero. However, this does not mean the error
is actually zero. One has only made the error in measurement zero to
lowest order. One also has to worry about higher derivative term
1 ∂2d
= 1 6= 0
d ∂θ2
at 90 degrees.

Electrons in a Periodic
Potential

(15.1) ‡Nearly Free Electron Model
Consider an electron in a weak periodic potential in one
dimension V (x) = V (x + a). Write the periodic potential
as
X iGx
e VG
V (x) =
G

where the sum is over the reciprocal lattice G = 2πn/a,
and VG∗ = V−G assures that the potential V (x) is real.
(a) Explain why for k near to a Brillouin zone boundary (such as k near π/a) the electron wavefunction should
be taken to be
ψ = Aeikx + Bei(k+G)x

(15.1)

where G is a reciprocal lattice vector such that |k| is close
to |k + G|.
(b) For an electron of mass m with k exactly at a zone
boundary, use the above form of the wavefunction to show

15

that the eigenenergies at this wavevector are
E=

where G is chosen so |k| = |k + G|.
 Give a qualitative explanation of why these two
states are separated in energy by 2|VG |.
 Give a sketch (don’t do a full calculation) of the
energy as a function of k in both the extended and the
reduced zone schemes.
(c) *Now consider k close to, but not exactly at, the
zone boundary. Give an expression for the energy E(k)
correct to order (δk)2 where δk is the wavevector difference from k to the zone boundary wavevector.
 Calculate the effective mass of an electron at this
wavevector.

(a) A periodic lattice can only scatter a wave by a reciprocal lattice
vector (Bragg diffraction). In the nearly free electron picture, the scattering perturbation is weak, so that we can treat the scattered wave in
perturbation theory. In this case, there is an energy denominator which
suppresses mixing of k-vectors which have greatly different unperturbed
energies. Thus, the only mixing that can occur is between two states
with similar energies that are separated by a reciprocal lattice vector.
Degenerate perturbation theory tells us that we should first diagonalize
within the degenerate space spanned by only these two eigenstates.
(b)We have our (variational) trial wavefunction given by
|ψi = A|ki + B|k + Gi
or equivalently

~2 k2
+ V0 ± |VG |
2m

(15.2)

√
ψ = (Aeikx + Bei(k+G)x )/ L

To maintain normalization we can insist that |A|2 + |B|2 = 1. Taking k
and k + G both on a Brillouin zone boundary we have k = nπ/a and k +

126 Electrons in a Periodic Potential

G = −nπ/a, where here we have chosen the nth zone boundary, and we
must have G = −2nπ/a the reciprocal lattice vector. The Hamiltonian
H in question is the usual Kinetic term plus V (x).
Approach 1: Diagonalize H within the 2d degenerate space
hk|H|ki = ~2 (nπ/a)2 /(2m) + V0

hk + G|H|k + Gi
hk|H|k + Gi
hk + G|H|ki

= ~2 (nπ/a)2 /(2m) + V0
= V2nπ/a

= V−2nπ/a

Diagonalizing this two by two matrix

 2
~ (π/a)2 /(2m) + V0
V2nπ/a
V−2nπ/a
~2 (π/a)2 /(2m) + V0
gives eigenstates
E=

~2 (nπ/a)2
+ V0 ± |V2nπ/a |
2m

∗
where we have used VG = V−G
. The eigenstates are correspondingly

|ψ± i = |ki + |k + Gi
which are (proportional to) the functions sin(2nπx/a) and cos(2nπx/a).
Interpretation: If we have considered only the V2nπ/a and V−2nπ/a
Fourier modes of the potential then we have V = 2V2nπ/a cos(2nπr/a).
Assuming V2nπ/a > 0, then the higher energy state is the ψ = cos(2nπr/a)
which puts the maximum amplitude of the wavefunction exactly at the
maxima of the potential. Similarly, the lower energy wavefunction is the
sin(2nπr/a) which has the minimum amplitude of the wavefunction at
the maximum of the potential. In the case of V2nπ/a < 0 the sin is the
higher energy wavefunction.
Approach 2: Variational.
If we simply calculate the expectation value of H in the trial state
given by Eq. 15.2 we obtain
hψ|H|ψi

∗
+ B ∗ A V2nπ/a
= ~2 (nπ/a)2 /(2m) + V0 + A∗ B V2nπ/a

Using the variational principle, the eigenstate is the trial wavefunction
which minimizes the total energy while preserving the normalization.
One way to do this is to write A = cos(θ) and B = eiχ sin(θ) which is
the most general form we can write while still preserving |A|2 + |B|2 = 1
(we can arbitrarily choose A to be real, since that only introduces an
irrelevant overall phase). In terms of these parameters we have
hψ|H|ψi = ~2 (nπ/a)2 /2m + V0 + 2Re[V2nπ/a eiχ sin(θ) cos(θ)]
for V2nπ/a > 0 this is minimized for χ = π and θ = π/4 (or equivalently χ = 0 and θ = 3π/4). It gives minimum energy states for
ψ = sin(2nπr/a) as above.

127

See figure 15.1 for a sketch of the bands in the extended zone scheme.

2nd Zone

2nd Zone

1st Brillouin Zone

Extended
Zone
Scheme
−2π/a

−π/a

π/a

0

2π/a

Reduced
Zone
Scheme
−2π/a

−π/a

π/a

0

2π/a

Fig. 15.1 Diagram of the dispersion in a nearly free electron model. Top: extended
zone scheme. Bottom: reduced zone scheme. Note that gaps open up at all zone
boundaries

(c) This calculation is entirely analogous to that above, only here we
need to consider k not on the zone boundary. Letting k = nπ/a + δk
and k + G = −nπ/a + δk we have
hk|H|ki =
hk + G|H|k + Gi =
hk|H|k + Gi

=

hk + G|H|ki =

~2 (δk + nπ/a)2 /(2m) + V0
~2 (δk − nπ/a)2 /(2m) + V0
V2nπ/a

V−2nπ/a

which we now need to diagonalize. We obtain
s
2
 2
~2 [(δk)2 + (nπ/a)2 ]
~ 2(δk)nπ/a
+ |V2nπ/a |2
+ V0 ±
E± =
2m
2m
expanding the square-root we obtain
E± =

~2 (nπ/a)2
~2 (δk)2
+ V0 ± |V2nπ/a | +
2m
2m



~2 (nπ/a)2
1±
m|V2nπ/a |

which is a quadratic correction as we move away from the Brillouin
zone. Note that for this expansion to remain valid we must have the
bracketed term in the square root two equations up small compared to
the |V2nπ/a |2 term.
The effective mass is then obtained by setting


1
~2 (nπ/a)2
1
1
±
=
2m∗
2m
m|V2nπ/a |

128 Electrons in a Periodic Potential

or equivalently
m∗ =

m
1±

~2 (nπ/a)2
m|V2nπ/a |

with the + being for the upper band.

(15.2) Periodic Functions
Consider a lattice of points {R} and a function ρ(x)
which has the periodicity of the lattice ρ(x) = ρ(x + R).

Show that ρ can be written as
X
ρG eiG·x
ρ(x) =
G

where the sum is over points G in the reciprocal lattice.

Generally we can always write ρ(x) in terms of its Fourier transform
Z
dk
ρ(x) = V
ρk eik·x
(2π)3
Now by periodicity we know that ρ(x) = ρ(x + R) for any lattice vector
R, so let us take a sum over all N lattice vectors in the system
Z
dk
V X
1 X
ρk eik·(x+R)
ρ(x + R) =
ρ(x) =
N
N
(2π)3
R
ZR
X
V
dk
ik·x
=
ρk e
eik·R
N
(2π)3
R

The sum gives

X
R

eik·R =

(2π)3 X 3
δ (k − G)
v
G

where the sum is over all G which are reciprocal lattice vectors and v is
the volume of the unit cell. Letting the delta function act and cancelling
some factors, we then directly obtain
X
ρ(x) =
eiG·x ρG
G

(15.3) Tight Binding Bloch Wavefunctions
Analogous to the wavefunction introduced in Chapter
11, consider a tight-binding wave ansatz of the form
X ik·R
e
|Ri
|ψi =
R

where the sum is over the points R of a lattice, and |Ri
is the ground-state wavefunction of an electron bound to

a nucleus on site R. In real space this ansatz can be
expressed as
X ik·R
ψ(r) =
e
ϕ(r − R).
R

Show that this wavefunction is of the form required by
Bloch’s theorem (i.e., show it is a modified plane wave).

129

Start by writing the function ϕ in its fourier representation
Z
dq iq·r
ϕ(r) = V
e ϕq
(2π)3
so that
ψ(r) =

X

eik·R V

R

The sum over R gives
X
R

ei(k−q)·R =

Z

dq iq·(r−R)
e
ϕq
(2π)3

(2π)3 X 3
δ (k − q − G)
v
G

and we allow the delta function to act, giving
V X −iG·r
e
ϕk−G
ψ(r) = eik·r
v
G

The sum over G is a function periodic in r → r + R (since eiG·R = 1)
hence this is in Bloch form.

(15.4) *Nearly Free Electrons in Two Dimensions Consider the nearly free electron model for a square
lattice with lattice constant a. Suppose the periodic potential is given by
V (x, y)

=

2V10 [cos(2πx/a) + cos(2πy/a)]

+

4V11 [cos(2πx/a) cos(2πy/a)]

(a) Use the nearly free electron model to find the energies
of states at wavevector G = (π/a, 0).
(b) Calculate the energies of the states at wavevector G = (π/a, π/a). (Hint: You should write down a 4
by 4 secular determinant, which looks difficult, but actually factors nicely. Make use of adding together rows or
columns of the determinant before trying to evaluate it!)

Note, we should not call the points (π/a, 0) and (π/a, π/a) as G since
they are not reciprocal lattice vectors!
(a) The first part of this problem is no different from the one dimensional problem posed in 15.1! The fourier components V10 couple
(π/a, 0) and (−π/a, 0). The energies of the states are
E = ~2 (π/a)2 /(2m) ± |V10 |
(b) The second part of the problem is more complicated. Here, all four
points (±π/a, ±π/a) all have the same energy and are coupled together
by the scattering potential. Therefore we must treat all of these states
in degenerate perturbation theory. Let us label the points
|1i = (+π/a, +π/a)
|2i = (+π/a, −π/a)
|3i = (−π/a, −π/a)
|4i = (−π/a, +π/a)

130 Electrons in a Periodic Potential

Writing a general wavefunction within this space as
tonian matrix within this reduced Hilbert space is


ǫ
V10 V11 V10
 V10
ǫ
V10 V11 


 V11 V10
ǫ
V10 
V10 V11 V10
ǫ

P

i

φi |ii the Hamil-

where ǫ = ~2 (π/a)2 /m. To find the eigenenergies we would thus like to
solve the determinant equation

0=

ǫ−E
V10
V11
V10

V10
ǫ−E
V10
V11

V11
V10
ǫ−E
V10

V10
V11
V10
ǫ−E

Adding and subtracting rows and columns leaves the determinant unchanged. So we can subtract row 3 from row 1 and subtract row 4 from
row 2. Then add column 3 to column 1 and add column 4 to column 2.
The result is

0=

0
0
V11 + ǫ − E
2V10

0
0
2V10
V11 + ǫ − E

V11 − ǫ + E
0
ǫ−E
V10

0
V11 − ǫ + E
V10
ǫ−E

Because of all the zeros, method of minors can then evaluate the determinant easily to given


2
0 = (V11 − ǫ + E)2 (V11 + ǫ − E)2 − 4V10

Thus giving solutions

(15.5) Decaying Waves
As we saw in this chapter, in one dimension, a periodic potential opens a band gap such that there are no
plane-wave eigenstates between energies ǫ0 (G/2) − |VG |
and ǫ0 (G/2) + |VG | with G a reciprocal lattice vector.
However, at these forbidden energies, decaying (evanes-

E

=

ǫ + V11

(two solutions)

E

=

ǫ − V11 ± 2|V10 |

cent) waves still exist. Assume the form
ψ(x) = eikx−κx
with 0 < κ ≪ k and κ real. Find κ as a function of energy
for k = G/2. For what range of VG and E is your result
valid?

Here, we can think of the wavevector k as taking an imaginary part
(i.e., absorb κ into k)
k = −G/2 + iκ

131

so that
ǫ0 (k)

=

ǫ0 (k + G) =

~2 k 2
2m

=

~2 (k+G)2
2m

=

~2 ((G/2)2 − iGκ − κ2 )
= ǫR − iǫI
2m
~2 ((G/2)2 + iGκ − κ2 )
= ǫR − iǫI
2m

where we have defined
ǫR =

~2 ((G/2)2 − κ2 )
2m
~2 (Gκ)
ǫI =
2m

As worked out in the text (Eq. 15.8), the characteristic equation is
0 =
=
=

(ǫ0 (k) − E) (ǫ0 (k + G) − E) − |VG |2
(ǫR − iǫI − E)(ǫR + iǫI − E) − |VG |2
(ǫR − E)2 + ǫ2I − |VG |2

Thus, we have
E = ǫR ±

(15.3)

q
|VG |2 − ǫ2I

Note here that for |ǫI | > |VG | the energy becomes imaginary and the
solution is not valid. Thus we must have
|VG | ≥ |~2 Gκ/(2m)|
In Eq. 15.3 we have a quadratic equation for κ2 (there are no lone factors
of κ), which can be solved to then give only one (possibly) positive
solution
r
~2 (G/2)2
~2 κ2
~2 (G/2)2
=−
− E + 4E
+ |VG |2
2m
2m
2m
In order for this solution to be valid, we must have the right hand side
be positive. We can write this condition as
 2
2
~ (G/2)2
~2 (G/2)2
+E
+ |VG |2
≤ 4E
2m
2m
Or equivalently

~2 (G/2)2
− E ≤ |VG |
2m

or in other words, that the energy is inside the gap!

(15.6) Kronig–Penney Model*
Consider electrons of mass m in a so-called “delta-

function comb” potential in one dimension
V (x) = aU

X
n

δ(x − na)

132 Electrons in a Periodic Potential
(a) Argue using the Schroedinger equation that inbetween delta functions, an eigenstate of energy E is always of a plane wave form eiqE x with
√
qE = 2mE/~.
Using Bloch’s theorem conclude that we can write an
eigenstate with energy E as
ψ(x) = eikx uE (x)
where uE (x) is a periodic function defined as
uE (x) = A sin(qE x) + B cos(qE x)

0 ay . In this case, imagine that the hoppings have
a value −tx in the x-direction and a value −ty in the ydirection, with ty > tx . (Why does this inequality match
ax > ay ?)
 Write an expression for the dispersion of the electronic states ǫ(k).
 Suppose again that the atoms are monovalent, what
is the shape of the Fermi surface now?

(a) The dispersion of the tight binding model is given by
ǫ(kx , ky ) = −2t (cos(kx a) + cos(ky a))
A contour plot of this energy is given in the left of Fig. 16.1. If we are
considering a monovalent unit cell, then the Brillouin zone is half filled.
This then gives a Fermi surface in the shape of a diamond, as shown in
the right of Fig. 16.1.
3

3

2

2

1

1

0

0

-1

-1

-2

-2

-3

-3
-3

-2

-1

0

1

2

3

-3

-2

-1

0

1

2

Fig. 16.1 Left: Dispersion of a 2D tight binding model on a square lattice. Right:
the Fermi surface for one electron per unit cell.

(b) If ax > ay one expects the hopping magnitude to be smaller in
the x direction since the atoms are further apart (although this is not
holy, as the orbitals, such as px orbitals, may not be isotropic). We then
expect a dispersion of the form
ǫ(kx , ky ) = −2tx cos(kx ax ) − 2ty cos(ky ay )

3

137

As an example, let us choose ty = 2tx but ax = ay = a for simplicity.
A contour plot of the energy is given in the left of Fig. 16.2. If we are
considering a monovalent unit cell, then the Brillouin zone is half filled.
This then gives a Fermi surface in the shape of ... well, i’m not sure
what to call it. But it is shown in the right of Fig. 16.2.
3

3

2

2

1

1

0

0

-1

-1

-2

-2

-3

-3
-3

-2

-1

0

1

2

3

-3

-2

-1

0

1

2

3

Fig. 16.2 Left: Dispersion of a 2D tight binding model on a square lattice with
anisotropic hoppings. Right: the corresponding Fermi surface for one electron per
unit cell.

(16.3) More Fermi Surface Shapes*
Consider a divalent atom, such as Ca or Sr, that forms
an fcc lattice (with a single atom basis). In the absence

of a periodic potential, would the Fermi surface touch the
Brillouin zone boundary? What fraction of the states in
the first Brillouin zone remain empty?

This is some nasty geometry. First, recall that the reciprocal lattice
of an FCC lattice with lattice constant a is a BCC lattice with lattice
constant 4π/a (See excercise 13.1). The BCC lattice has two lattice
points per conventional unit cell, so the primitive unit cell (or the Brillouin zone in this case) has volume 21 (4π/a)3 in k-space. Now since we
have a divalent unit cell in real space, we should have enough electrons
to fill exactly the volume of the Brillouin zone. Thus for electrons in the
absence of a periodic potential we have
 3
4πkF3
1 4π
≈ 992/a3
=
3
2 a
or
kF = (24π 2 )1/3 /a ≈ 6.187/a
We would like to know if this hits the Brillouin zone boundary or not.
For a BCC lattice, the nearest neighbor of the point [0,0,0] is the point

138 Insulator, Semiconductor, or Metal

[1/2,1/2,1/2] in units of the lattice
√constant. The perpendicular bisector
to this point is thus a distance 3/4. For our reciprocal lattice with
lattice constant 4π/a, the distance to this perpendicular bisector (i.e.,
to the Brillouin zone boundary) is then
√
dk = π 3/a ≈ 5.441/a
which is less than the Fermi wavevector, thus telling us that the Fermi
surface hits the Brillouin zone boundary (it is obvious that it must hit
the Brillouin zone boundary since the volume of the fermi surface must
equal the volume of the Brillouin zone and they are not the same shape!).
The Fermi surface thus goes into the 2nd Brillouin zone as a spherical
cap of radius 6.187/a where the Brillouin zone boundary is of radius
5.441/a. Note that the center of this spherical cap is in the L direction
(in the language of Brillouin zones, see fig 13.6 of the book). We should
check that the spherical cap remains on the L-face. We can check this
by noting that the angle subtended by this spherical cap is only θ =
cos−1 (5.441/6.187) ≈ 28.425 degrees, which is much smaller than the
angle to say the K point.
The height of the spherical cap is h = 6.187/a − 5.441/a = 0.745/a.
A well known geometric formula gives us that the volume of a cap is
V =

πh2
(3r − h) = 10.4/a3
3

However, note that we have 8 such spherical caps in all of the 8 equivalent
L directions, thus giving a total volume of
V = 83/a3
in the 2nd Brillouin zone, compared to the total volume of the fermi
surface which is 992/a3 . Thus the caps in the second Brillouin zone
account for roughly 8% of the filled states.

Semiconductor Physics

(17.1) Holes
(a) In semiconductor physics, what is meant by a hole
and why is it useful?
(b) An electron near the top of the valence band in a
semiconductor has energy
E = −10−37 |k|2

where E is in Joules and k is in m−1 . An electron is removed from a state k = 2 × 108 m−1 x̂, where x̂ is the unit

17

vector in the x-direction. For a hole, calculate (and give
the sign of!)
(i) the effective mass
(ii) the energy
(iii) the momentum
(iv) the velocity.
 If there is a density p = 105 m−3 of such holes all
having almost exactly this same momentum, calculate the
current density and its sign.

(a) A hole is the absence of an electron in an otherwise filled valence
band. This is useful since instead of describing the dynamics of all the
(many) electrons in the band, it is equivalent to describe the dynamics
of just the (few) holes.
(b) Effective mass ~2 k 2 /(2m∗) = (10−37 Joule · meter2 )k 2 . So m∗ =
5 × 10−32 kg or .05 the mass of the electron. This mass is positive in the
usual convention. The energy is E = (10−37 Joule·meter2 ))k 2 = 4×10−21
J, or about 0.025 eV. This energy is positive (it takes energy to ”push”
the hole down into the fermi sea, like pushing a balloon under water).
Getting the momentum and velocity right are tricky. First, note that
the velocity of an eigenstate is the same whether or not the state is filled
with an electron. It is always true that the velocity of an electron in a
state is ∇k Ek /~ where Ek is the electron energy. Thus the hole velocity
here is negative v = −~k/m∗ = −3.8 × 105 m/s (i.e the velocity is in
the negative x̂) direction.
For momentum, since a filled band carries no (crystal) momentum,
and for electrons crystal momentum is always ~k, the removal of an
electron leaves the band with net momentum −~k which we assign as
the momentum of the hole. Thus we obtain hole momentum −~k =
−2.1 × 10−26 kg-m/s which is also in the negative x̂ direction. (this
matches well to the intuition that p = mv with a positive effective mass
for holes). With p the density of such holes, the total current density is
pev = −6 × 10−9 Amp/m2 also in the negative x̂ direction (noting that
the charge of the hole is positive).
Note that it is typical to define the wavevector of a hole to be negative

140 Semiconductor Physics

of the wavevector of the missing electron.

(17.2) Law of Mass Action and Doping of Semiconductors
(a) Assume that the band-gap energy Eg is much
greater than the temperature kB T . Show that in a pure
semiconductor at a fixed T , the product of the number of
electrons (n) and the number of holes (p) depends only on
the density of states in the conduction band and the density of states in the valence band (through their effective
masses), and on the band-gap energy.
 Derive expressions for n for p andR for the product
∞
np. You may need to use the integral 0 dx x1/2 e−x =
√
π/2.

germanium at room temperature.
(c) The graph in Fig. 17.1 shows the relationship between charge-carrier concentration for a certain n-doped
semiconductor.
 Estimate the band gap for the semiconductor and
the concentration of donor ions.
 Describe in detail an experimental method by
which these data could have been measured, and suggest
possible sources of experimental error.

(b) The band gaps of silicon and germanium are 1.1
eV and 0.75 eV respectively. You may assume the effective masses for silicon and germanium are isotropic,
roughly the same, and are roughly .5 of the bare electron
mass for both electrons and holes. (Actually the effective
masses are not quite the same, and furthermore the effective masses are both rather anisotropic, but we are just
making a rough estimates here.)
 Estimate the conduction electron concentration for
intrinsic (undoped) silicon at room temperature.
 Make a rough estimate of the maximum concentration of ionized impurities that will still allow for this
“intrinsic” behavior.
 Estimate the conduction electron concentration for

Fig. 17.1 Figure for Exercise 17.

(a) The density of states per unit volume of free electron with dispersion E = ~2 k 2 /(2m) is given by (including spin)
g(E) =

√
m3/2
2E 3 2
~ π

So if the dispersion near the valence band edge and conduction band
edges are
Ee (k)

= Ec + ~2 k 2 /(2me )

Eh

= Ev − ~2 k 2 /(2mh )

we obtain density of states for conduction electrons and valence holes
given by
ge (E > Ec )
gh (E < Ev )

3/2
p
me
2(E − Ec ) 3 2
~ π
3/2
p
m
2(Ev − E) 3h 2
=
~ π

=

141

At fixed chemical potential µ and temperature β the number density of
electrons in the conduction band is
Z ∞
ge (E)nF (β(E − µ))dE
n=
Ec

where nF (x) = 1/(ex + 1) is the Fermi occupation factor. Assuming
that µ is well below the conduction band (by at least energy kb T ), then
x = β(E − µ) is very positive and it is acceptable to replace nF (x) by
the Boltzmann factor e−x , thus we obtain.
Z ∞
ge (E)e−β(E−µ) dE
n=
Ec

Similarly, the number of holes in valence band is given by
Z Ev
gh (E)(1 − nF (β(E − µ))dE
p=
−∞

Assuming that µ is well above the valence band (by at least energy kb T )
then x = β(E − µ) is very negative and we can replace the fermi factor
1 − nF (x) by ex resulting in
Z Ev
gh (E)eβ(E−µ) dE
p=
−∞

It is then clear immediately, that when we multiply np the variable µ
completely vanishes.
A bit more manipulation obtains
√ 3/2 Z ∞
p
eβµ 2me
(E − Ec ) e−βE dE
n=
3
2
~ π
Ec
redefining variables y = E − Ec and performing the integral, one obtains

3/2
1 2me kb T
n=
e−β(Ec −µ)
(17.1)
4
π~2
and similarly
p=

1
4

Obtaining
np =

1
16



2mh kb T
π~2

3/2

e−β(µ−Ev )

 √
3
2 mh me kb T
e−β(Ec −Ev )
π~2

(17.2)

which depends only on the band gap, T and the effective masses.
(b) In the undoped case n = p, and we are assuming mh = me = m/2
as well so we have

3/2
1 2(m/2)kb T
e−β(Eg /2)
n=
4
π~2

142 Semiconductor Physics

with Eg the gap value. Plugging in numbers gives
n = 5.26 × 1015 m−3
for Silicon and
n = 4.54 × 1018 m−3
for Germanium. When the doping level gets to on the order of the
expected intrinsic level, then you no longer have intrinsic behavior.
(c) The concentration of donor atoms is simply the saturation concentration at low T (which I estimate from the figure to be about 2 × 1019
m−3 ). Note that at very low temperature one could get carrier freeze-out
where the density drops again.
To estimate the gap, we need to measure the slope of the curve at
high temperature.
Extracting a slope
log n/(m3 ) = −1500K/T + Constant.
Or equivalently
n ∼ e−1500K/T
We then set 1500K = Eg /(2kb ) and obtain a band gap of about .26 eV.
It might be useful to also mention to the students that Si and Ge
both have valley degeneracies (i.e, multiple equi-energy minima in the
conduction band). These may add an additional factor to the law of
mass action.
The concentration is most likely measured by a Hall effect measurement. Several possible sources of error can occur here. First, when there
are both electrons and holes present, then you measure some (nontrivial) combination of the Hall resistivities weighted by their concentrations
and by their mobilities (in a very nontrivial way).
RH =

Re ρ2e + Rh ρ2h
(ρe + ρh )2

(See exercise 17.9). One only gets an accurate absolute measurement of
the electron concentration to the extent that the electron resistivity is
much lower than the hole resistivity.
There are other more obvious sources of experimental error such as
heating when one runs current through a sample to measure it – thus it
requires measuring small voltages accurately. To measure hall resistivity,
without mixing in longitudinal resisitivity, one needs to align contacts
exactly parallel to each other in a hall bar. One has to also make sure
that the voltages being measured are due to the sample and not the
contacts/wires/amplifiers etc.

143
(17.3) Chemical Potential
(a) Show that the chemical potential in an intrinsic
semiconductor lies in the middle of the gap at low temperature.
(b) Explain how the chemical potential varies with temperature if the semiconductor is doped with (i) donors (ii)
acceptors.

(c) A direct-gap semiconductor is doped to produce a
density of 1023 electrons/m3 . Calculate the hole density
at room temperature given that the gap is 1.0 eV, and
the effective mass of carriers in the conduction and valence band are 0.25 and 0.4 electron masses respectively.
Hint: use the result of Exercise 17.2.a.

(a) In an intrinsic semiconductor n = p so we can set
n
=1
p
Referrring back to the previous problem we can insert the expressions
Eq. 17.1 and Eq. 17.2 for n and p respectively. Almost all of the nasty
prefactors cancel and we obtains
3/2

me e−β(Ec −µ)
3/2

mh e−β(µ−Ev )

=1

We can solve this trivially to obtain
µ=

3
Ec + Ev
+ (kB T ) log(mh /me )
2
4

So at low temperature the chemical potential lies mid-gap. (Incidentally,
this is why it is never a good idea to say that the fermi energy is the
energy of the highest filled state. There may be a very large difference
between the highest filled state and the chemical potential!).
(b) Let us assume for a moment we are well above the freezeout temperature, so the doping can be thought of as going directly into the
conduction band. For simplicity let us assume me = mh , so the intrinsic
behavior is then simply that the chemical potential is fixed as a function of temperature. If the doner density is much higher than nint then
n ≈ ndopant (i.e., the thermally excited electrons are irrelevant compared
to those that are there from doping). Then looking at Eq. 17.1 fixing n
to be ndopant and solving for µ we have


ndopant
µ ≈ Ec − kB T log 
 
1 2me k T 3/2
b

4

π~2

Similarly for acceptor impurities



µ ≈ Ev + kB T log 

1
4


pacceptor 

2m k T 3/2
h b

π~2

So at low tempearture, the chemical potential is essentially right at the
conduction (donor) or valence (acceptor) band and moves towards midgap as the temperature is increased. When the intrinsic density exceeds

144 Semiconductor Physics

the dopant density, then one expects to have µ given by the intrinsic
expression from part a, which is roughly to have the chemical potential
mid-gap with a small slope dependent on the ratio of masses.
Now the story is a bit more complicated if one wants to think about
the very low temperature regime where there is carrier freezeout. At
zero temperature all of the electrons are bound to their dopant nuclei
and one can think of this as being a filled “impurity band” playing the
role of a filled valence band, and the nearby conduction band is empty.
As is usually the case, at zero temperature the chemical potential is
midway between the top of the impurity band and the bottom of the
conduction band. As the temperature is increased, since the real valence
band is very far away compared to the temperature so we can ignore it.
It is then only a matter of figuring out how the chemical potential moves
between the filled impurity band and the empty conduction band. Since
the density of states in the conduction band is larger than the those
in the sparse impurity band, as the temperature is raised, the chemical
potential moves up towards the conduction band.
(c) We first calculate the undoped intrinsic carrier concentration using
the law of mass action with n = p = nintrinsic . At T =293 Kelvin, I
obtained
nintrinsic = 1016 m−3
Then since ndopant ≫ nintrinsic we can set p = n2intrinsic /n (from the
law of mass action) to obtain
p = 109 m−3

(17.4) Energy Density
Show that the energy density of electrons in the valence
band of a semiconductor is
3
(ǫc + kB T ) n
2

where n is the density of these electrons and ǫc is the
energy of the bottom of the conduction band.

In short this is just a matter of realizing that the electrons in the
conduction band are essentially classical (are activated with Boltzmann
factors not fermi factors), so classical stat mech applies and one can apply the equipartition theorem. One obtains ǫc for each particle excited,
then an extra 32 kB T for the translational degrees of freedom as usual in
equipartition theorem.
One can, of course, do the calculation more rigorously writing the
total energy density as (compare problem 17.2)
Z ∞
Ege (E)nF (β(E − µ))dE
E/V =
Ec

where nF (x) = 1/(ex + 1) is the Fermi occupation factor. As in 17.2 it

145

is acceptable to replace nF (x) by the Boltzmann factor to get
Z ∞
Ege (E)e−β(E−µ) dE
E/V =
Ec

The same manipulation obtains
√ 3/2 Z ∞
p
eβ(µ−Ec ) 2me
E/V =
[(E − Ec ) + Ec ] (E − Ec ) e−β(E−Ec ) dE
3
2
~ π
Ec
Note the second term in brackets, the Ec term (compare 17.2) gives the
same integral as the above calcualtion of n in 17.2 so we obtain nEc .
To evaluate the remaining term we redefine variables y = E − Ec and
performing the integral, whixh we recognize as being precisely ∂/∂β of
the prior integral from 17.2. Since the prior integral was proportional to
T ( 3/2) we obtain (3/2)(kB T ) times the prior integral and thus a total
of (3/2)(kB T )n proving the result.

(17.5) Semiconductors
Describe experiments to determine the following properties of a semiconductor sample: (i) sign of the majority

carrier (ii) carrier concentration (assume that one carrier
type is dominant) (iii) band gap (iv) effective mass (v)
mobility of the majority carrier.

(i,ii) Sign of majority carrier and carrier concentration (assuming there
is only one type of carrier) are both easily measured with Hall effect. (iii)
band gap may be measured optically. Or by carrier concentration (essentially conductance) as a function of temperature. (iv) Effective mass
is measured with cyclotron resonance (v) mobility is easily measured via
resistivity once concentration of carriers is known.

(17.6) More Semiconductors
Outline the absorption properties of a semiconductor
and how these are related to the band gap. Explain the
significance of the distinction between a direct and an

indirect semiconductor. What region of the optical spectrum would be interesting to study for a typical semiconducting crystal?

Optical absorbtion can occur when a photon is absorbed while exciting
an electron out of the valence band into the conduction band. This
requires a minimum of the gap energy (Small amounts of absorbtion can
occur below the gap for impure semiconductors if there are impurity or
defect states within the gap – also very weak nonlinear processes can
allow multiple photons to be absorbed while exciting a single electron).
In the absorbtion process energy and momentum must both be conserved. Since photons carry very little momentum given a certain energy

146 Semiconductor Physics

(being that c is very large) one should think of this absorbtion as not
tranferring any momentum to the system. This means that direct gap
absorbtion (where the momentum of the electron does not change) is
highly favored over indirect gap absorbtion. Indirect gap absorbtion
can occur, but it must be assisted by a phonon or some other process
that can account for the necessary momentum.
Semiconductor gaps tend to be in the optical, or infra-red range (somewhere from 400 nm to 3 micron, or roughly 3 eV to .5 eV). Only a very
few wide gap semiconductors reach the optical blue range and UV.

(17.7) Yet More Semiconductors
Outline a model with which you could estimate the energy of electron states introduced by donor atoms into an

n-type semiconductor. Write down an expression for this
energy, explaining why the energy levels are very close to
the conduction band edge.

One can consider a simple hydrogenic schroedinger equation with an
attractive proton being the ionized donor and the single electron. The
main differences are that the mass of the electron is replaced by the
band electron mass, and ǫ0 is multiplied by the dielectric constant ǫr of
the semiconductor. As a result, the Rydberg becomes replaced by an
effective Rydberg
R∗ = R0 (m∗ /m)(1/ǫ2r )
This gives us a hydrogenic binding energy that can be extremely small
for typical semiconductors, hence the bound states remain very close the
conduction band edge.

(17.8) Maximum Conductivity*
Suppose holes in a particular semiconductor have mobility µh and electrons in this semiconductor have mobility µe . The total conductivity of the semiconductor will
be
σ = e (n µe + p µh )
with n and p the densities of electrons in the conduction

band and holes in the valence band. Show that, independent of doping, the maximum conductivity that can be
achieved is
√
σ = 2e nintrinic µe µh
with nintrinsic the intrinsic carrier density.
value of n − p is this conductivity achieved?

For what

Actually this is easy. Using law of mass action np = n2intrinsic . Thus
we write

nintrinsic 
σ = e n µe +
µh
n
Now set dσ/dn = 0 to maximize and solve to obtain
p
n = nintrinsic µh/µe

Which correspondingly results in

p = nintrinsic

p
µe/µh

147

plugging into the original expression for σ with a tiny bit of algebra we
obtain
√
σ = 2e nintrinic µe µh
as required, and we also obtain
n − p = nintrinsic


p
p
µh/µe − µe/µh

(17.9) Hall Effect with Both n- and p-Dopants*
Suppose a semiconductor has a density p of holes in
the valence band with mobility µh and a density n of

electrons in the conduction band with mobility µn . Use
Drude theory to calculate the Hall resistivity of this sample.

See also exercise 3.3c. For a single species, we have (See exercise 3.1)


r
BR
ρ=
−BR
r
where r = nµ and R = q/n with q the charge on the charge carrier and
n the carrier density. We define tensors ρe and ρh for the two species
with re = nµn and rh = pµh and Re = e/n and Rh = −e/p. The
conductivity tensors are σj = ρ−1
j and then the total conductivity tensor
is σ = σe + σi . Finally this is inverted to give the tensor ρtotal = σ −1 .
There is a lot of algebra involved in this. I obtained
ρxx

=

ρxy

=

B 2 (re Rh2 + rh Re2 ) + rh re (re + rh )
B 2 (Re + Rh )2 + (re + rh )2

B B 2 Re Rh (Re + Rh ) + Rh re2 + Re rh2
B 2 (Re + Rh )2 + (re + rh )2

Semiconductor Devices

(18.1) Semiconductor Quantum Well
(a) A quantum well is formed from a layer of GaAs
of thickness L nm, surrounded by layers of Ga1−x Alx As
(see Fig. 18.2). You may assume that the band gap of the
Ga1−x Alx As is substantially larger than that of GaAs.
The electron effective mass in GaAs is 0.068 me whereas
the hole effective mass is 0.45 me with me the mass of

the electron.
 Sketch the shape of the potential for the electrons
and holes.
 What approximate value of L is required if the band
gap of the quantum well is to be 0.1 eV larger than that
of GaAs bulk material?
(b) *What might this structure be useful for?

(a) This is a particle in a box problem. Both electron and holes are
particles in a box of length L. Thus the lowest lying electron in the well
is
 π  2 ~2
Ee = Ec +
L 2m∗e
with Ec the bulk conduction band minimum. Similarly the highest lying
hole state in the well is
 π  2 ~2
Eh = Ev −
L 2m∗h
Thus the difference in energy is
Ee − Eh = Ebulkgap +

18



1
~2  π 2 1
−
2 L
m∗e
m∗h

Setting Ee − Eh − Ebulkgap to 1eV and solving for L gives 8 nm.
(b) This type of quantum well device is useful to precisely design a
band gap for example for a laser where one wants to fix the emission
wavelength. If one puts donor impurities outside of the well (on both
sides, say) the donated electrons can reduce their energies by falling
into the well, but the ionized dopants remain behind. This is known as
modulation doping. It is useful because one can obtain extremely high
mobility electrons within the quantum well since there are no ionized
dopants in the well to scatter off of. One uses these structures heavily
for fundamental physics studies of clean (unperturbed) electrons.

150 Semiconductor Devices
(18.2) Density of States for Quantum Wells
(a) Consider a quantum well as described in the previous exercise. Calculate the density of states for electrons
and holes in the quantum well. Hint: It is a 2D electron
gas, but don’t forget that there are several particle-in-abox states.

(b) Consider a so-called “quantum wire” which is a
one-dimensional wire of GaAs embedded in surrounding
AlGaAs. (You can consider the wire cross-section to be a
square with side 30nm.) Describe the density of states for
electrons or holes within the quantum wire. Why might
this quantum wire make a very good laser?

(a) For a 2D electron gas for electrons with mass m, we quickly calcualte the density of states.
Z kF
dk
N = 2A
(2π)2
0
with A the area and the factor of 2 out front for spin and k a two
dimensional vector. This can be converted to
n = N/A =

kF2
2π

When the energy of an electron is given by
E=

~2 k2
2m

we have k 2 = 2mE/~2 and we then have a density of states per unit
volume of
m
dn
=
g=
dE
π~2
This is the correct answer for any E ≥ 0 and the density of states is zero
for any E < 0. We then write more precisely that
g(E) =

m
Θ(E)
π~2

where Θ is the step function which has value 1 for nonnegative argument
and value 0 for negative argument.
In our quantum well we must make a few minor changes. First of
all, we should use the effective mass rather than the actual mass of
the electron. Secondly the energy of the particle in the quantum well
also includes its particle-in-a-box energy for its motion transverse to the
2D quantum well. Thus for an electron in the conduction band of the
quantum well we have
E = Ec +

~2 k2
~2 π 2 a2
+
2m∗e L2
2m∗e

where Ec is the bulk conduction band bottom, and a = 1, 2, 3, . . . is
the transverse quantum number of the particle in the well. Fixing the
transverse quantum number a, the density of states would be


m∗e
~2 π 2 a2
g(E) =
Θ E − Ec −
π~2
2m∗e L2

151

Now accounting for the fact that there may be many transverse modes
we have


m∗e X
~2 π 2 a2
g(E) =
Θ E − Ec −
π~2 a>0
2m∗e L2

and this expression remains true up to an energy where the transverse
modes spill out of the box.
Analogously the density of holes in the valuence band in the quantum
well is


~2 π 2 a2
m∗ X
−
E
Θ Ev −
g(E) = h2
π~ a>0
2m∗h L2

(b) First we determine the density of states for a one dimensional
electron gas.
Z kF
dk
N = 2L
−kF 2π

with L the Length of the system and the factor of 2 out front for spin.
This can be converted to
n = N/L =
using k =

√

2kF
π

2mE/~ we then obtain a density of states
√
2m −1/2
dn
=
E
Θ(E)
g(E) =
dE
π~

Now in the seminconductor quantum wire we must consider the transverse modes. In general the energy of an electron in the wire is then given
by
~2 π 2 a21 + a22
~2 k 2
E = Ec +
+
2m∗e L2
2m∗e
where a1 and a2 are the mode indices (integers greater than zero) in the
two transverse directions. Adding the density of states associated with
all of these modes we obtain
−1/2


√ ∗ X 
~2 π 2 a21 + a22
~2 π 2 a21 + a22
2me
E − Ec −
Θ
E
−
E
−
g(E) =
c
π~ a ,a >0
2m∗e L2
2m∗e L2
1

2

and similarly in the valence band
p ∗

−1/2


2mh X
~2 π 2 a21 + a22
~2 π 2 a21 + a22
Ev −
−
E
−
E
g(E) =
Θ
E
−
v
π~ a ,a >0
2m∗h L2
2m∗h L2
1

2

(18.3) p-n Junction*
Explain the origin of the depletion layer in an abrupt
p-n junction and discuss how the junction causes rectifca-

tion to occur. Stating your assumptions, show that the
total width w of the depletion layer of a p-n junction is:
w = wn + w p

152 Semiconductor Devices
where



1/2
2ǫr ǫ0 NA φ0
wn =
eND (NA + ND )
and a similar expression for wp Here ǫr is the relative
permittivity and NA and ND are the acceptor and donor
densities per unit volume, while φ0 is the difference in
potential across the p-n junction with no applied voltage.
You will have to use Poisson’s equation to calculate the

form of φ given the presence of the ion charges in the
depletion region.
 Calculate the total depletion charge and infer how
this changes when an additional voltage V is applied.
 What is the differential capacitance of the diode
and why might it be useful to use a diode as a capacitor
in an electronic circuit?

Let us set the position x to be zero, where we have an n doped region
to the left (negative x) and a p doped region to the right (positive x).
Let the depletion widths be wn and wp respectively, and the doping
densitities be ND and NA respectively. Within the depletion widths
there is a net charge built up (see Fig 18.4 of the book for example).
We must solve the Poisson equation ∂x2 φ = ρ/(ǫ0 ǫr ) where ρ is the local
charge density (which is constant in each region). Setting φ(x = 0) = 0
for simplicity, we immediately obtain
φ(x)

=

φ(x)

=

−eND 2
x + CD x
2ǫ0 ǫr
eNA 2
x + CA x
2ǫ0 ǫr

x<0
x>0

where CD and CD are constants to be fixed here. We have additional
boundary conditions that the electric field must go to zero at the edge
of the depletion region, so we have ∂x φ(x = −wn ) = ∂x (x = wa ) = 0.
This fixes the constants so that we now have
φ(x)

=

φ(x)

=


−eND 2
x + 2wn x
2ǫ0 ǫr

eNA 2
x − 2wa x
2ǫ0 ǫr

x<0
x>0

The total potential drop across both regions is then
φ0 = φ(−wn ) − φ(wa ) =


e
ND wn2 + NA wa2
2ǫ0 ǫr

(18.1)

We also note that the total charge in the two depletion regions must sum
to zero (since the depletion occurs by annihilation of opposite charges)
so we have
wa NA = wn ND
(18.2)
Plugging Eq. 18.2 into Eq. 18.1 and solving for wn yields the desired
expression.

1/2
2ǫr ǫ0 NA φ0
wn =
eND (NA + ND )
and similarly

wa =



2ǫr ǫ0 ND φ0
eNA (NA + ND )

1/2

153

When an additional voltage is added, it simply shifts φ0 to φ0 + eV .
The total depletion charge per unit cross sectional area is then
q = wn ND = wa NA =



2ǫr ǫ0 ND NA (φ0 + eV )
e(NA + ND )

1/2

The differential capacitance per unit cross sectional area is
∂q
C=
=
∂V



ǫr ǫ0 ND NA
2e(NA + ND )

1/2

(φ0 + eV )−1/2

This provides a useful circuit element as it allows one to control a capacitance by applying a DC voltage.

(18.4) Single Heterojunction*
Consider an abrupt junction between an n-doped semiconductor with minimum conduction band energy ǫc1 and
an undoped semiconductor with minimum conduction

band energy ǫc2 where ǫc1 < ǫc2 . Describe qualitatively
how this structure might result in a two-dimensional electron gas at the interface between the two semiconductors.
Sketch the electrostatic potential as a function of position.

It should say, ǫc1 > ǫc2 .
In this problem, nothing interesting happens in the valence band — so
we only draw the conduction band. The situation is quite similar to the
p-n junction. Before the two semiconductors are brought together, there
are electrons in the conduction band on the left (or in dopant orbitals
slightly below the conduction band as shown in the figure), but empty
states on the right at lower energy because the conduction band energy
is lower. This situation is shown in figure 18.1

Fig. 18.1 Before electrons are allowed to flow between
the two seminconductors, there are electrons on the left
at higher energy than empty states on the right. The
blue lines indicate the conduction band minima. The
dotted line is the fermi energy (slighly below the conduction band assuming that electrons are bound to dopants).

As in the p-n junction, electrons want to flow between the two semiconductors in order to lower their energies, but in so doing charge builds
up (a postive charge is left on the left and negative charged electrons
accumulate on the right. When equilibrium is established the electrochemical potential is as shown in figure 18.2. The electrons are back

154 Semiconductor Devices

attracted to the positive charges they left behind, thus accumulating in
a roughly triangular well that forms near the interface. If the confinement is sufficiently strong, this becomes a particle in a box problem and
the electrons in the well may become restricted to a single transverse
wavefunction — thus becoming a strictly two dimensional electron gas.

Fig. 18.2 Electrochemical potential of a semiconductor
heterostructure. There is net postive charge on the left
where electrons have left their donor atoms behind and
net negative charge on the right where electrons have
accumulated without positively charged nuclei.

(18.5) Diode Circuit
Design a circuit using diodes (and any other simple
circuit elements you need) to convert an AC (alternating

current) signal into a DC (direct current) signal.
 *Can you use this device to design a radio reciever?

For the purpose of this problem we will make the crude assumption
that a diode (arrow in a circuit diagram) is an ideal circuit element that
allows current flow one way but no current flow the other way.

Fig. 18.3 A half-wave rectifier.

The circuit shown in Fig. 18.3 is known as “half-wave rectifier”. If the
source voltage is sinusoidal, V = V0 sin(ωt), the voltage across the load

155

resistor is then
V = V0 sin(ωt)Θ(sin(ωt))
where Θ is the Heaviside step function. I.e., the current goes to zero
instead of going negative. Now instead of giving a true DC output, this
gives an output that is nonnegative. In order to smooth the nonnegative
(but fluctuating) voltage into a smooth DC voltage one inserts a capacitor into the circuit to act as a “battery” as shown in Fig. 18.4. This
effectively damps the high frequency components of the signal leaving
only the low frequency DC component. One should choose a capacitor
such that Rload C ≫ 1/ω in which case the non-dc component of the
resultant signal is of magnitude VAC ≈ V0 /(ωRC) .

Fig. 18.4 A half-wave rectifier with a smoothing capacitor.

One can do a bit better in circuit design by using more diodes to
construct a “Full wave rectifier” as shown in Figure 18.5.

Fig. 18.5 A full-wave rectifier.

In this case, in the absence of the smoothing capacitor, the voltage

156 Semiconductor Devices

across the load is given by
V = |V0 sin(ωt)|
a smoothing capacitor can be used as above. This scheme has the advantage that the resulting DC voltage ends up being of higher magnitude
than that of the half-wave, and the fluctuating DC component ends up
being smaller.
 In order to design a radio, we must first think about how radios
encode information. The simplest encoding system is AM or amplitude
modulation. In this case a signal A(t) (for simplicity let us assume the
signal is everywhere positive) is multiplied by a high frequency carrier
wave sin(ωt) and the product A(t) sin(ωt) is transmitted and then received by an antenna. To convert the received signal back into A(t) one
simply puts it through a rectifier. If one wants to do a slightly better job, one wants to recieve only signals where the carrier frequency is
very close to some given frequency ω0 . To do this, one needs to build
a resonant (LC) circuit which will respond only to frequencies near its
resonance. A sample 1-diode radio circuit (often known as a “crystal”
radio after the fact that the diode was often made of a small crystal)
is shown in Fig. 18.6. In this figure the right hand side (D1,C3,E1) is
just a half-wave rectifier as discussed above. Here E1 is the “earphone”
or output load of the circuit. The middle of the circuit L2,C2 is the
resonant LC circuit, which is inductively coupled to the incoming signal
from the antenna. Note that the capacitors are made tunable here so
that the resonance frequencies can be modified. Many other designs are
possible as well.

Fig. 18.6 A radio circuit. Here E1 is the “earphone”
or output load of the circuit. The incoming antenna is
on the far left. A very crude radio could be made by
with only the rectifier (D1,C3) attached directly to an
incoming antenna and an outgoing earphone E1.

157

(18.6) CMOS Circuit*
Design a circuit made of one n-MOSFET and one pMOSFET (and some voltage sources etc.) which can act
as a latch—meaning that it is stable in two possible states

and can act a single bit memory (i.e., when it is turned
on it stays on by itself, and when it is turned off it stays
off by itself).

A simple (and very rough) CMOS latch circuit (also known as an
SCR or silicon-controlled rectified) is shown in Fig. 18.7. When the
switch (S1) is closed no current flows. The reason for this is that there
is no voltage on either of the two gates, so both transistors are “off”
meaning they prevent current flow (and therefore prevent voltage from
being transmitted). However, when a voltage is momentarily applied
to the ON input, then this activates the P-MOSFET, allowing current
to flow to the gate of the N-MOSFET, which then activates it, allowing
current to flow to the gate of the P-MOSFET even when the ON voltage
is removed. Thus the circuit is latched in the “On” state. To put
the circuit back in the “Off” state, one must disconnect the switch S1
momentarily.

Fig. 18.7 A CMOS latch circuit.

Magnetic Properties of
Atoms: Para- and
Dia-Magnetism

(19.1) ‡ Atomic Physics and Magnetism
(a) Explain qualitatively why some atoms are paramagnetic and others are diamagnetic with reference to the
electronic structure of these materials.
(b) Use Hund’s rules and the Aufbau principle to de-

19

termine L, S, and J for the following isolated atoms:
(i) Sulfur (S) atomic number = 16
(ii) Vanadium (V), atomic number = 23
(iii) Zirconium (Zr), atomic number = 40
(iv) Dysprosium (Dy), atomic number = 66

(a) To a first approximation, paramagnetic atoms have net moment
J 6= 0 which can be re-aligned, whereas the typical diamagnetic atoms
have no moment.
If an atom has completely filled shells (say, the noble gases), then J =
L = S = 0, and the atom is diamagnetic due to Larmor Diamagnetism.
If an atom has a net moment J 6= 0, from unfilled shells, then it is
paramagnetic. This is almost all other atoms.
However (more advanced answer): it is also possible to have J = 0
while having L and S nonzero. In this case, the atom can either be paramagnetic or diamagnetic. Both para and dia terms are weak and either
one can win in this case. This can occur for atoms that are one electron
short of half-filled shells. (Known as Van Vleck Paramagnetism).
In metals one can have Pauli paramagnetism associated with re-orientation
of the spins of the conduction electron. One can also have Landau diamagnetism (beyond the scope of the course) which is the diamagnetic
response of the orbital motion of the conduction electrons. Pauli paramagnetism in a metal is much weaker than Curie paramagnetism and
can be roughly the same size as diamagnetic effects.
(b) Shells are filled in the order
1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, 5s, 4d . . .
with s,p,d shells containing 2,6,10 electrons respectively.
(i) 16 S: [Ne]3s2 3p4
(ii) 23 V: [Ar]4s2 3d3
(iii) 40 Zr: [Kr]5s24d2

160 Magnetic Properties of Atoms: Para- and Dia-Magnetism

(iv) 54 Xe: [Xe] (All filled shells)
(v) 66 Dy: [Xe]6s2 4f 10
Where filled shell configurations [Ne] contains 10 electrons, [Ar] contains 18, [Kr] contains 36 and [Xe] contains 54.
Thus
(i) Sulfer has 4 electrons in a p-shell, these fill lz = −1, 0, 1 with spin
up and lz = 1 with spin down. Thus L = 1 and S = 1, and since the
shell is more than half full J = L + S = 2.
(ii) Vanadium has 3 electrons in a d-shell, these fill lz = 2, 1, 0 with
spin up giving L = 3 and S = 3/2. Since the shell is less than half filled
J = L − S = 3/2.
(iii) Zirconium has 2 electrons in a d-shell, these fill lz = 2, 1 with
spin up giving L = 3 and S = 1. Since the shell is less than half filled
J = L − S = 2.
(iv) Xenon is a noble gas, meaning all shells are filled, so J = L =
S = 0.
(v) Dysprosium has 10 electrons in an f-shell, these fill all the spin up
state (7 of them) and lz = 3, 2, 1 for spin down giving L = 6 and S = 2.
Since the shell is more than half filled J = L + S = 8.
Note that none of these atoms violate the Madelung rule which dictates the filling order or violates Hund’s rules when the atoms are isolated. (Violations do sometimes occur but these atoms work as they are
supposed to).

(19.2) More Atomic Physics
(a) In solid erbium (atomic number=68), one electron
from each atom forms a delocalized band so each Er atom
has eleven f electrons on it. Calculate the Landé g-factor
for the eleven electrons (the localized moment) on the Er
atom.

(b) In solid europium (atomic number =63), one electron from each atom forms a delocalized band so each Eu
atom has seven f electrons. Calculate the Landé g-factor
for the seven electrons (the localized moment) on the Eu
atom.

Er typically is in a +3 state. 2 of those are from the core s-orbitals.
1 is from the f orbital.
(a) For 11 electrons in an f-shell, using Hund’s first rule we obtain
S = 3/2 and Hund’s second rule we have L = 6. Since the shell is more
than half filled, J is given by the sum J = 6 + 3/2 = 15/2. Using the
formula for the Lande g factor (with g = 2) we have


1
1
S(S + 1) − L(L + 1)
g̃ = (g + 1) + (g − 1)
= 6/5
2
2
J(J + 1)
This is almost exactly right.
(b) The counting here is messed up. There are 7 electrons before
losing one. For 7 electrons in an f-shell, L = 0 and S = 7/2 and J = S.
is purely spin, so g̃ = g = 2 (it also comes out of the above formula as

161

well). In fact, we should have 6 electrons when one is lost, and we get
L = 3 = S and J = 0. However, This is a van-vleck ion, so in fact the
result is not what is predicted here.

(19.3) Hund’s Rules*
Suppose an atomic shell of an atom has angular momentum l (l = 0 means an s-shell, l = 1 means a p-shell
etc, with an l shell having 2l + 1 orbital states and two

spin states per orbital.). Suppose this shell is filled with
n electrons. Derive a general formula for S, L, and J as
a function of l and n based on Hund’s rules.

The shell has 2l +1 orbital states and 2 spin states per orbital. Hund’s
first rule tells us that
 n
0 ≤ n ≤ 2l + 1
 2
S(n, l) =
 4l+2−n
2l + 1 ≤ n ≤ 4l + 2
2

The second rule tells us that for n ≤ 2l + 1 we have
L(n, l) =

x=l
X

x

x=l−n+1

We can do this sum to obtain
L(n, l) =

1
n(2l + 1 − n)
2

For 2l + 1 ≤ n ≤ 4l + 2 we can consider only the electrons in addition
to the L = 0 half-filled shell, so L(n, l) = L(n − (2l + 1), l) So we have
in all,
 1
0 ≤ n ≤ 2l + 1
 2 n(2l + 1 − n)
L(n, l) =
 1
2l + 1 ≤ n ≤ 4l + 2
2 (n − 2l − 1)(4l + 2 − n)

And thus we have (using Hund’s third rule) J = L − S for less than half
filled and J = L + S for more than half filled so
 1
0 ≤ n < 2l + 1
 2 n(2l − n)
J(n, l) =
 1
2l + 1 ≤ n ≤ 4l + 2
2 (n − 2l)(4l + 2 − n)
‡ (19.4) Para and Diamagnetism
Manganese (Mn, atomic number=25) forms an atomic
vapor at 2000K with vapor pressure 105 Pa. You can
consider this vapor to be an ideal gas.

(a) Determine L, S, and J for an isolated manganese
atom. Determine the paramagnetic contribution to the
(Curie) susceptibility of this gas at 2000K.
(b) In addition to the Curie susceptibility, the man-

162 Magnetic Properties of Atoms: Para- and Dia-Magnetism
ganese atom will also have some diamagnetic susceptibility due to its filled core orbitals. Determine the Larmor
diamagnetism of the gas at 2000K. You may assume the

atomic radius of an Mn atom is one Ångstrom.
Make sure you know the derivations of all the formulas
you use!

Using the ideal gas law, the density of Mn atoms is n = P/(kB T ). We
will use this below.
(a) Atomic Mn has orbital configuration 25 Mn: [Ar]4s2 3d5 , meaning
a half filled d-shell. This then has L = 0 and S = 5/2. Since this is
purely spin moment the g-factor is 2.
Now we need to determine the paramagnetic susceptibility of a spin
S = 5/2. Let us write the partition function
Z=

5/2
X

e−βgµB Bm

m=−5/2

Since we will be concerned with small B, it is probably simplest to
expand the partition function directly
Z

≈

5/2
X

m=−5/2

= 6+



1
1 − (βgµB Bm) + (βgµB Bm)2 + . . .
2



35
(βgµB B)2
4

Calculating the moment
moment = −∂(kB T log Z)/∂B =

35
β(gµB )2 B
12

Yielding a susceptibility of the gas
χ = ∂M/∂H = nµ0

35
35
β(gµB )2 = [P/(kB T )2 ] µ0 µ2B = 1.6 × 10−7
12
3

(b) Diamagnetism is understood from the schroedinger equation and
expanding for small magnetic field
H=

(p + eA)2
p2
1
e2 A2
=
+
(p · A + Ap) +
2m
2m 2m
2m

The middle term is a L · B, and has zero expectation if there is a filled
shell such that hLi = 0. The final term is the source of diamagnetism.
One can write in circular gauge
A=

1
r×B
2

and the final term ends up being
e2
e2
|B|2 (r⊥ )2 =
|B|2 r2
8m
6m

163

where we have used the spherical symmetry of the atom so that x2 +y 2 =
(2/3)r2 . We thus have the susceptibility
ρe2 2
hr i
6m
where here ρ is the density of electrons. Since there are 20 core electrons
per atom, we have ρ = 20n with n the gas density (using 25 here would
also be sensible). Using 1 angstrom for r we thus obtain
χ = −µ0 dE 2 /dB 2 = −µ0

χ = −4.3 × 10−9
which is much much smaller than the paramagnetic contribution, even
at 2000K!

‡(19.5) Diamagnetism
(a) Argon is a noble gas with atomic number 18 and
atomic radius of about .188 nm. At low temperature it
forms an fcc crystal. Estimate the magnetic susceptibility
of solid argon.
(b) The wavefunction of an electron bound to an impurity in n-type silicon is hydrogenic in form. Estimate

the impurity contribution to the diamagnetic susceptibility of a Si crystal containing 1020 m−3 donors given that
the electron effective mass m∗ = 0.4me and the relative
permittivity is ǫr = 12. How does this value compare to
the diamagnetism of the underlying silicon atoms? Si has
atomic number 14, atomic weight 28.09, and density 2.33
g/cm3 .

(a) First we establish the density of Argon. The radius of the avrgon
atom is r√= .188 nm. For an fcc crystal, the nearest neighbor distance is
r = a/(2 2) We then have four atoms
√ per conventional unit cell for an
overall atomic density of 4/a3 = 1/(4 2r3 ). The number of electrons is
18 times the number of atoms, so we have (see 19.4 for derivation)
χ = −µ0

3 e2
e2
ρe2 2
hr2 i = −µ0 √
hr i = −µ0 18 √
= −1 × 10−4
6m
(4 2r3 )6m
4 2 mr

This number is actually a bit too big because the average hr2 i is smaller
than this number since most of the electrons are further inside the atom
than the full atomic radius. Measurement
gives us a suscpetability about
p
10−5 , which would correspond to hr2 i ≈ r/3.
(b) For a hydrogenic orbital (recall ψ ∼ e−r/a0 ) it is easy to calcualte
that hr2 i = 3a20 . We need only calculate the Bohr radius of an impurity
in a silicon atom. Recalling that the expression for the Bohr radius is
4πǫ0 ~2
me2
we see that the Bohr radius in silicon should be rescaled by
a0 =

∗
aSi
0 = a0 ǫr m/m = a0 ∗ 12/.4 = 1.6nm.

Again using the result of the prior problem (and using m∗ rather than
m)
ρe2 2
hr i ≈ −10−11
χ = −µ0
6m∗

164 Magnetic Properties of Atoms: Para- and Dia-Magnetism

Finally we would like to compare this to the diamagnetism of the pure
silicon. First let us calculate the density of atoms in the system
2.33g mol 6.02 × 1023 atoms (100cm)3
atoms
= 5 × 1028
cm3 28.09g
mol
m3
m3
Now we need to determine the atomic radius of silicon from the data
given. If we didn’t know the crystal structure, we might guess roughly
that
1
r = n−1/3
2
which would be exact for a simple cubic lattice. However, we know that
Si is diamond structure with 8 atoms per conventional unit cell, so the
density is n = 8/a3 with a the lattice constant so a = 2n−1/3 . The
nearest
0, 0] to [a/4, a/4, a/4] or a distance
√ neighbor distance is from [0, √
of a 3/4, so the atomic radius is a 3/8 = 41 n−1/3 = 1.17 Angstrom.
We then have
ρe2 2
χ = −µ0
hr i ≈ −4 × 10−6
6m
This is much much greater than the diamagnetism of the few impurities.
(Note: I think the table value is almost exactly this, however, this is
fortuitous as the table value is usually given in cgs which differs from
the SI version by a factor of 4 pi. Again the error is in the estimate of
r2 . )
n=

‡(19.6) Paramagnetism
Consider a gas of monatomic atoms with spin S = 1/2
(and L = 0) in a magnetic field B. The gas has density
n.
(a) Calculate the magnetization as a function of B and

T . Determine the susceptibility.
(b) Calculate the contribution to the specific heat of
this gas due to the spins. Sketch this contribution as a
function of µB B/kB T .

“Monatomic atoms” ? What was I smoking when I wrote this? It
should say monovalent. Doh! I should have also specified the g-factor
(as usual we can take it to be 2).
(a) Same calculation as we have done a million times.
1
1
Z = e+βgµB 2 B + e−βgµB B
2
F = −kB T log Z
1
1
m = −∂F/∂B = gµB tanh(βgµB B)
2
2
M = mn
1
χ = lim µ0 ∂M/∂B = µ0 n( gµB )2 β
B→0
2
(b) Similarly, we have
1
1
U = ∂ log Z/∂β = gµB B tanh(βgµB B)
2
2
1 2
1
1
2
(gµB B) sech (βgµB B)
C = ∂U/∂T =
2
kB T
2
2

165

(b) Show that the susceptibility is given by

(19.7) Spin J Paramagnet*
Given the Hamiltonian for a system of non-interacting
spin-J atoms
H = g̃µB B · J
(a)* Determine the magnetization as a function of B and
T.

χ=

nµ0 (g̃µB )2 J(J + 1)
3
kB T

where n is the density of spins. (You can do this part
of the exercise without having a complete closed-form expression for part a!)

(a) The eigenenergies of a single spin are
Em = g̃µB Bm

m = −J, −J + 1 , −J + 2 , . . . J − 1 , J

So the canonical partition function of the single spin is
Z

=

J
X

e−βgµB Bm = eβg̃µb BJ

=

e−βg̃µB p

p=0

m=−J

=

2J
X

1 − e−βg̃µB B(2J+1)
eβg̃µB B(2J+1)/2 − e−βg̃µB B(2J+1)/2
eβg̃µb BJ
=
−βg̃µ
B
B
1−e
eβg̃µB B/2 − e−βg̃µB B/2
sinh(βg̃µB B(2J + 1)/2)
sinh(βg̃µB B/2)

we construct the free energy per spin F/N = −kB T log Z and then get
the magnetic moment per spin
m=−

∂(F/N )
= (g̃µB /2) [(2J + 1) coth(βg̃µB B(2J + 1)/2) − coth(βg̃µB B/2)]
∂B

and the total magnetizaton is then the moment per spin times the density
of spins n, so
M = n(g̃µB /2) [(2J + 1) coth(βg̃µB B(2J + 1)/2) − coth(βg̃µB B/2)]
(b) To obtain the susceptibility, we want
χ = lim µ0
B→0

∂M
∂B

To take this limit we take the argument of the coth to be small and we
use the expansion
1 x
lim coth x = +
x→0
x
3
Note that the two 1/x terms cancel when subtracted leaving


M ∼ n(g̃µB /2) (2J + 1)2 βg̃µB B/6 − βg̃µB B/6
So we have

χ=
as required.

nµ0 (g̃µB )2 J(J + 1)
3
kB T

166 Magnetic Properties of Atoms: Para- and Dia-Magnetism

Note, the question claims that part (b) may be achieved without completing part (a). To do this, note that we only concerned with small B
so we can expand the partition function directly for small B.
Z

=

J
X

e−βg̃µB Bm =

m=−J

J
X

m=−J

1
1 − βg̃µB Bm + (βg̃µB B)2 m2 + . . .
2

The first term in the sum gives (2J + 1) and the second term gives 0 by
symmetry. The third term is the only hard one. I claim that
G[J] =

J
X

m=−J

=

1
(2J + 1)(J + 1)J
3

(19.1)

I will derive this below, but for now let us assume it is correct. So we
have
1
Z = (2J + 1)[1 + (βg̃µB B)2 J(J + 1) + . . .]
6
Using the free energy per spin is F/N = −kB T log Z, since B is small
we have
F/N = −kB T log(2J + 1) −

1
(g̃µB B)2 J(J + 1)
6kB T

And the magnetic moment per site is
m=−

(g̃µB )2 BJ(J + 1)
∂(F/N )
=
∂B
3kB T

which multiplying by the density of spins n to give the magnetization,
gives the same result as above.
Finally we turn to derive Eqn. 19.1. The result of this sum must be
some polynomial in J. Further, approximating it as an integral, it must
have a maximum power of J 3 , and the coefficient of the J 0 term must
be zero since G[0] = 0. Thus we propose
G[J] = aJ 3 + bJ 2 + cJ
Then we can also write the difference of two successive sums as just the
new ad
G[J + 1] − G[J] = 2(J + 1)2
which we multiply out to give
a(3J 2 + 3J + 1) + b(2J + 1) + c = 2(J + 1)2
matching coefficients then gives the values of a, b, c which proves Eq. 19.1.

Spontaneous Magnetic
Order: Ferro-, Antiferro-,
and Ferri-Magnetism
(20.1) Ferromagnetic vs Antiferromagnetic
States
Consider the Heisenberg Hamiltonian
H=−

X
1X
gµB B · Si
J Si · Sj +
2
i

(20.1)

hi,ji

and for this exercise set B = 0.
(a) For J > 0, i.e., for the case of a ferromagnet, intuition tells us that the ground state of this Hamiltonian
should simply have all spins aligned. Consider such a
state. Show that this is an eigenstate of the Hamiltonian
Eq. 20.1 and find its energy.

20

(b) For J < 0, the case of an antiferromagnet on a cubic lattice, one might expect that (at least for B = 0) the
state where spins on alternating sites point in opposite
directions might be an eigenstate. Unfortunately, this is
not precisely true. Consider such a state of the system.
Show that the state in question is not an eigenstate of the
Hamiltonian.
Although the intuition of alternating spins on alternating sites is not perfect, it becomes reasonable for systems with large spins S. For smaller spins (like spin 1/2)
one needs to consider so-called “quantum fluctuations”
(which is much more advanced, so we will not do that
here).

It is useful here to recall that
1
Si · Sj = (Si+ Sj− + Si− Sj+ ) + Siz Sjz
2
(Indeed, students often need to be reminded of this! Maybe it is worth
giving this as a hint!)
(a) If each spin is aligned in the ẑ direction (it has Sz = 1), then the
energy is −gµb B per spin and for each bond we have energy −JSi · Sj =
−JSiz Sjz = −JS 2 since S + on the spins all give zero. The system is in
an energy eigenstate with energy
E = −N gµB |B| − N zJS 2 /2
with z the number of neighbors of each site (=6 for a cubic lattice).
(b) Assuming an antiferromagnetic configuration, the key here is to
note that there are terms Si+ Sj− which do not vanish (where S + is applied to a down spin, and S − is applied to an up spin). This means
that when the hamiltonian is applied to the proposed antiferromagnetic
ground states, it generates other spin configurations. Hence this is not
an eigenstate.

168 Spontaneous Magnetic Order: Ferro-, Antiferro-, and Ferri-Magnetism

(20.2) Frustration
Consider the Heisenberg Hamiltonian as in Exercise 20
with J < 0, and treat the spins as classical vectors.
(a) If the system consists of only three spins arranged

in a triangle (as in Fig. 20.2), show that the ground state
has each spin oriented 120◦ from its neighbor.
(b) For an infinite triangular lattice, what does the
ground state look like?

Probably here I should have stated explicitly that all three spins have
the same |S| (might be interesting to consider a case where they don’t
all have the same spin!).
(a) The Hamiltonian is
H = J(S1 · S2 + S1 · S3 + S2 · S3 )
Since Si · Si = S 2 is a constant we can write
H = (J/2)(S1 + S2 + S3 )2 + constant
To minimize the energy, we must have
S1 + S2 + S3 = 0
which implies that the spins are at 120 degree angles from each other –
but can lie in any plane.
(b) For a triangular lattice, each triangle must have three spins each
at 120 degree angles from its neighbors. So choose three directions all
at 120 degree angles from each other in any given plane. Call these
three directions A, B, C. Now we must assign each site on the triangular
lattice one of the three values A, B, or C in such a way that all triangles
contain one site of type A, one of type B, and one of type C. One can
think of this as being now a crystal whose unit cell has three times the
area of the original unit cell, and now contains one spin of each type A,
B, C.

(20.3) Spin Waves*
For the spin-S ferromagnet particularly for large S, our
“classical” intuition is fairly good and we can use simple
approximations to examine the excitation spectrum above
the ground state.
First recall the Heisenberg equations of motion for any
operator
i~

dÔ
= [Ô, H]
dt

with H the Hamiltonian (Eq. 20.1 with Si being a spin S
operator).
(a) Derive equations of motion for the spins in the

Hamiltonian Eq. 20.1. Show that one obtains
dSi
~
= Si ×
dt

J

X
j

Sj − gµb B

!

(20.2)

where the sum is over sites j that neighbor i.
In the ferromagnetic case, particularly if S is large, we
can treat the spins as not being operators, but rather as
being classical variables. In the ground state, we can set
all Si = ẑS (Assuming B is in the −ẑ direction so the
ground state has spins aligned in the ẑ direction). Then
to consider excited states, we can perturb around this

169
solution by writing
Siz
Six
Siy

=
=
=

2

S − O((δS) /S)
δSix

δSiy

where we can assume δS x and δS y are small compared to
S. Expand the equations of motion (Eq. 20.2) for small
perturbation to obtain equations of motion that are linear
in δSx and δSy
(b) Further assume wavelike solutions
δSix

=

Ax eiωt−ik·r

δSiy

=

Ay eiωt−ik·r

This ansatz should look very familiar from our prior consideration of phonons.

Plug this form into your derived equations of motion.
 Show that Six and Siy are out of phase by π/2.
What does this mean?
 Show that the dispersion curve for “spin-waves” of
a ferromagnet is given by ~ω = |F (k)| where
F (k) = gµb |B|

+ JS(6 − 2[cos(kx a) + cos(ky a) + cos(kz a)])

where we assume a cubic lattice.
 How might these spin waves be detected in an experiment?
(c) Assume the external magnetic field is zero. Given
the spectrum you just derived, show that the specific heat
due to spin wave excitations is proportional to T 3/2 .

(a) We need the angular momentum algebra
[S x , S y ] = iS z
and cyclic permutations of this. We then have the Heisenberg equations
i~

dSix
= [Six , H]
dt

The only terms that this does not commute with are those containing
Siy and Siz . Thus we have
i~

X y y
dSix
(Sj Si + Sjz Siz )]
= [Six , gµb (By Siy + Bz Siz ) − J
dt
j

with the sum over j being over neighbors of i. Thus we obtain


X
dSix
i~
(Sjy Siz − Sjz Siy )
= i (gµb )(By Siz − Bz Siy ) − J
dt
j

and similar for the other two components of the spin. Thus we conclude


X
dSi
Sj − gµb B
= Si × J
~
dt
j
Writing

Si = ẑS + δSi
we obtain
~

X
dδSi
δSj + δSi × (JZS ẑ + gµb |B|)
= JS ẑ ×
dt
j

where Z is the number of neighbors of a site.

170 Spontaneous Magnetic Order: Ferro-, Antiferro-, and Ferri-Magnetism

(b) Plugging in the wave ansatz (for a cubic lattice) we obtain
i~ωAx

=

F (k)Ay

i~ωAy

=

−F (k)Ax

where
F (k) = JS(6 − 2[cos(kx a) + cos(ky a) + cos(kz a)]) + gµb |B|
This system of equations can be solved immediately to give dispersion
~ω = |F (k)|
Spin waves are typically detected by inelastic neutron scattering. This
is like scattering from phonons except that one uses (spin-polarized)
neutrons in order to couple to the spin.
(c) For small k = |k|, and hence small energy, the energy spectrum is
quadratic
E = JSa2 k 2
We will also need
k = (E/JSa2 )1/2
and
dk = dE/(2(EJSa2 )1/2 )
Calculating the total energy stored in spin waves, we have
Z
Z ∞
dk
V
U =V
E(k)nB (E(k)) ≈ 2
dkk 2 E(k)nB (E(k))
3
2π 0
BZ (2π)
with nB the usual Bose factor and the approximation accurate for low
temperatures. We rewrite this as
Z ∞
V
U =
dEE 3/2 nB (E)
4π 2 (JSa2 )3/2 0
Z ∞
x5/2
V (kb T )5/2
dx
=
ex − 1
4π 2 (JSa2 )3/2 0
which we differentiate to get the heat capacity
Z ∞
x5/2
5 V (kb T )3/2
dx
C = dU/dT = kb
2 4π 2 (JSa2 )3/2 0
ex − 1
(one does not really need to carry all of the constants to see how it
scales!). For those who are interested though, the integral can be evaluated in the usual way
Z ∞
Z ∞
∞ Z ∞
∞
X
X
1
x5/2
5/2 −nx
dyy 5/2 e−y = Γ(7/2)ζ(7/2)
=
dxx e
=
dx x
7/2
e − 1 n=1 0
n
0
0
n=1

√
where Γ(7/2) = 15 π/8 and ζ(7/2) ≈ 1.1267 . . ..

171

(20.4) Small Heisenberg Models
(a) Consider a Heisenberg model containing a chain of
only two spins, so that
H = −JS1 · S2

.

Supposing these spins have S = 1/2, calculate the energy spectrum of this system. Hint: Write 2S1 · S2 =
(S1 + S2 )2 − S1 2 − S1 2 .

(b) Now consider three spins forming a triangle (as
shown in Fig. 20.2). Again assuming these spins are
S = 1/2, calculate the spectrum of the system. Hint:
Use the same trick as in part (a)!
(c) Now consider four spins forming a tetrahedron.
Again assuming these spins are S = 1/2, calculate the
spectrum of the system.

Note the obvious typo, it should read 2S1 ·S2 = (S1 +S2 )2 −S1 2 −S2 2 .
(a)
H = −(J/2)[(S1 + S2 )2 − S21 − S22 ]

Since (S)2 = S(S + 1) for spin-1/2 we have (S)2 = 3/4. Further, when
two spin 1/2’s are added they can form either a spin-0 singlet or a spin1 triplet (three Sz states). So (S1 + S2 )2 takes the values 0 for the
singlet or S(S + 1) = 2 for the S = 1 triplet. Thus the Hamiltonian has
eigenstates 3J/4 for the singlet (one eigensate) and −J/4 for the triplet
(three eigenstates).
(b) Similarly
H = −(J/2)[(S1 + S2 + S3 )2 − S21 − S22 − S23 ]
Here, again for spin 1/2 we have (S)2 = 3/4. When adding three spin1/2s, we can obtain spin-1/2 in two ways and spin-3/2 in one way. (To
see this, think about adding the first two spin-1/2 to get spin-0 or spin1. Now adding a spin-1/2 the spin-0 gives spin-1/2 and adding spin-1/2
to the spin-1 gives either spin-1/2 or spin-3/2. Note that counting the
total nbumber of eigenstates we should get 8 since each spin-1/2 has
two possible Sz states. Each of the two possible spin-1/2’s can take two
possible Sz states and the spin-3/2 can take 4 possible Sz states, which
gives a total of 8 possible Sz states.
In the case that the three spins add to spin-1/2, we obtain energy
3J/4 (four eigenstates) whereas if the three spins add to spin-3/2, S2 =
S(S + 1) = 15/4, so the energy is −3J/4 (four eigenstates).
(c) Same story
H = −(J/2)[(S1 + S2 + S3 + S4 )2 − S21 − S22 − S23 − S24 ]
The sum of the four spins can give spin 0 in two ways, spin 1 in 3 ways,
and spin 2 in one way. Again we should add up the total number of
eigenstates to check that it is 24 = 16. We have 2 spin 0’s + 3 states for
spin 1 in 3 ways + 5 states in spin 2. So we have 2+9+5 = 16. The spin0 singlets (two eigenstates) have energy (−J/2)[0 − 4(3/4)] = 3J/2. The
three spin-1 triplets (9 eigenstates) have energy (−J/2)[2 − 4(3/4)] =
J/2 and the spin-2 fiveplets (5 eigenstates) have energy (−J/2)[2 × 3 −
4(3/4)] = −3J/2

172 Spontaneous Magnetic Order: Ferro-, Antiferro-, and Ferri-Magnetism

(20.5) One-Dimensional Ising Model with B = 0
(a) Consider the one-dimensional Ising model with spin
S = 1. We write the Hamiltonian for a chain of N spins
in zero magnetic field as
H = −J

N−1
X

σi σi+1

i=1

where each σi takes the value ±1. The partition function
can be written as
X
Z=
e−βH .
σ1 ,σ2 ,...σN

Using the transformation Ri = σi σi+1 rewrite the partition function as a sum over the R variables, and hence
evaluate the partition function.
 Show that the free energy has no cusp or discontinuity at any temperature, and hence conclude that there
is no phase transition in the one-dimensional Ising model.
(b) *At a given temperature T , calculate an expression for the probability that M consecutive spins will be
pointing in the same direction. How does this probability
decay with M for large M ? What happens as T becomes
small? You may assume N ≫ M .

(a) The first spin can be either in the spin-up or spin down state, so
we leave σ1 as a variable to be summed over. The remaining spins are
defined by Ri for i = 1, . . . N − 1 where each R can take the values ±1.
The Hamitonian in terms of the R variables is
H = −J

N
−1
X

Ri

i=1

So the partition function is
Z = 2(e−βJ + eβJ )N −1 = 2(2 cosh(βJ))N −1
with the factor of 2 out front being the sum over the first spin. The free
energy is thus
F = −kB T log Z = −kB T N log 2 − kb T N log cosh(βJ)
which is a completely continuous function with no cusps or discontinuities at finite β.
(b) The probability that a given R is in the +1 state is
P (R = +1) =

eβJ
1
= −2βJ
βJ
−βJ
e +e
e
+1

Having R in the +1 state tells us that two consecutive spins are pointing
in the same direction. Looking at M − 1 consecutive values of R, the
probability that all of these are +1 is then

M−1
1
P (M ) =
e−2βJ + 1
which would put M consecutive spins in the same direction. This probability decays exponentially with M . To see this, rewrite
P (M ) = exp[−(M − 1) log(e−2βJ + 1)]

173

so the decay length is 1/ log(e−2βJ + 1). In the low tempearture limit
P (R + 1) becomes close to unity, and the decay length becomes long.
Actually, we can see that for very large β we have log(e−2βJ +1) ≈ e−2βJ ,
so the decay length is e2βJ lattice sites and is exponentially long.

(20.6) One-Dimensional Ising Model with B 6= 0*
Consider the one-dimensional Ising model with spin
S = 1. We write the Hamiltonian (Eq. ??) for a chain of
N spins in magnetic field B as
H=
where

N
X
i=1

Hi

H1 = hσ1

Hi = −Jσi σi−1 + hσi

(20.3)

for i > 1

(a) Show that these partial partition functions satisfy
a recursion relation
Z(M, σM ) =

X

σM −1

TσM ,σM −1 Z(M − 1, σM −1 )

where T is a 2 by 2 matrix, and find the matrix T . (T is
known as a “transfer matrix”).
(b) Write the full partition function in terms of the
matrix T raised to the (N − 1)th power.
(c) Show that the free energy per spin, in the large N
limit, can be written as

where each σi takes the value ±1 and we have defined
h = gµB B for simplicity of notation.
Let us define a partial partition function for the first M
spins (the first M terms in the Hamiltonian sum Eq. 20.3)
given that the M th spin is in a particular state. I.e.,
PM
X
Z(M, σM ) =
e−β i=1 Hi

where λ+ is the larger of the two eigenvalues of the matrix
T.
(d) From this free energy, derive the magnetization,
and show that the susceptibility per spin is given by

so that the full partition function is Z = Z(N, +1) +
Z(N, −1).

which matches the Curie form at high T .

F/N ≈ −kB T log λ+

σ1 ,...,σM −1

(a) The transfer matrix takes the form
 −βH(+,+)
  −β(−J+h)
e
e−βH(+,−)
e
T =
=
e−βH(−,+) e−βH(−,−)
e−β(J+h)

χ ∝ βe2βJ

e−β(J−h)
e−β(−J−h)



(b) The full partition function is then
Z = vT N −1 v T
where v = (1, 1). Although T is not hermitian, it can still be brought to
diagonal form via a Jordon decomposition
T = JΛJ −1
where Λ is a diagonal matrix of the eigenvalues of T and J is a matrix
made of the eigenvectors of T (these are no longer orthonormal!). (c)
Assuming these two eigenvalues are not degenerate, we then have for
large N that
−1
Z ∼ λN
+

174 Spontaneous Magnetic Order: Ferro-, Antiferro-, and Ferri-Magnetism

where λ+ is the larger of the two eigenvalues. Thus the free energy per
site is
F/N ∼ −kB T log λ+
For the record, the value of the larger eigenvalue is
λ+ =


p
e−β(h+J) 
1 + e2βh + 1 − 2e2βh + e4βh + 4e4β(h+J)
2

Just to check that this agrees with the previous problem, note that
setting h = 0 we correctly obtain
λ+ = e−βJ + eβJ
(d) Probably we want to derive the magnetic moment per spin which is
m=−

∂(F/N )
∂ log λ+
= gµB kb T
∂B
∂h

The result is a bit of a mess. Then once we have this, we want the
susceptibility per spin
χ = µ0 ∂m/∂B
which one evaluates at B = 0. It is much easier to make expansions for
small h since this is all we will need. In the end we will obtain
χ = g 2 µ2B µ0 kb T
as claimed.

∂ 2 log λ+
= g 2 µ2B µ0 βe−2βJ
∂h2

Domains and Hysteresis
(21.1) Domain Walls and Geometry
Suppose a ferromagnet is made up of a density ρ of
spins each with moment µB .
(a) Suppose a piece of this material forms a long circular rod of radius r and length L ≫ r. In zero external
magnetic field, if all of the moments are aligned along the
L-direction of the rod, calculate the magnetic energy of
this ferromagnet. (Hint: a volume of aligned magnetic
dipoles is equivalent to a density of magnetic monopoles
on its surface.)
(b) Suppose now the material is shaped such that
r ≫ L. What is the magnetic energy now?
(c) If a domain wall is introduced into the material,
where might it go to minimize the magnetic energy in the
two different geometries. Estimate how much magnetic
energy is saved by the introduction of the domain wall.

(d) Suppose the spins in this material are arranged in
a cubic lattice, and the exchange energy between nearest
neighbors is denoted J and the anisotropy energy is very
large. How much energy does the domain wall cost? Comparing this energy to the magnetic energy, what should
we conclude about which samples should have domain
walls?
(e) Note that factors of the lattice constant a are often
introduced in quoting exchange and anisotropy energies
to make them into energies per unit length or unit volume. For magnetite, a common magnetic material, the
exchange energy is JS 2 /a = 1.33 × 10−11 J/m and the
anisotropy energy is κS 2 /a3 = 1.35×104 J/m3 . Estimate
the width of the domain wall and its energy per unit area.
Make sure you know the derivation of any formulas you
use!

(a) Following the hint, the magnetic energy between two monopoles
′
of charge qm and qm
separated by distance r is given by
E=

′
µ0 qm qm
4π r

Here we have charge of magnitude qm = µB Aρ where A = πr2 is the
area of the end. Check that this has the right dimensions, recall that
a dipole is a charge times a length, so qm correctly has dimension of a
charge.
Thus the energy is
µ0 (µB ρ(πr2 ))2
4π
L
(b) This problem, on the other hand is analogous to the energy of a
capacitor. The energy stored in a capacitor is q 2 /(2C) where electrically
the capacitance is ǫ0 A/d with A the area and d the spacing. The analogy
here is thus A/(dµ0 ). Thus the total energy stored is
2
E = qm
µ0 d/(2A) =

21

µ0 (µB ρ)2 (πr2 )L
2

(c) First, let us consider a domain wall that cuts L in half (i.e., a plane
parallel to the A surface). This puts two magnetic charges of the same

176 Domains and Hysteresis

sign right next to each other and is therefore energetically unfavorable.
Now consider a domain wall perpendicular to this plane — i.e., that runs
along the L axis. In the (a) case we can view this as, instead of having a
single charge at each end of the rod, now we have two opposite charges
along each end of the rod (each with charge ±qm /2). This completely
kills the leading energy cost and now we have instead some sort of dipolar
interaction between the two ends — with energy typically on the much
smaller order
µ0 (µB ρ(πr2 /2))2 r2
E=
4π
L3
Similarly with the capacitor configuration of part (b), we can convert
“charge” energy to “dipole” energy by placing a domain wall in the
diraction along the axis (in the direction of L). However, here we will
not be able to consider the system “dipolar” until the size of the domains
d is smaller than L. Once we do this, the energy will again drop as in
the (a) case proportional to d6 /L3
(d) Since the lattice spacing will be a = ρ−1/3 , the energy of introducing a domain wall of area A is AJρ2/3 = 2LrJρ2/3 . In the (a) case,
for large enough L, the magnetic energy is not very large to begin with,
so introducing an energy cost proportional to L is not a very good idea.
Now let us consider the (b) case. Let us consider introducing a domain
wall network spaced by d. The number of domains is πr2 /d2 and the
total aread of the walls will be πr2 L/d and their total domain wall energy will be πr2 LJρ2/3 /d. As a function of d, as discussed in part (c)
then magnetic energy drops strongly when d < L. Thus, there will be
some happy medium which optimizes the total energy.
p(e) As mentioned in the text, the width of a domain wall should be
J/κ√which here is 3100 lattice spacings. The energy per unit area is
then Jκ = 4 × 107 Joules per a2 area.
(21.2) Critical Field for Crystallite
(a) Given that the energy of a crystallite in a magnetic
field is given by
E/V = E0 − |M ||B| cos θ − κ′ |M |2 (cos θ)2
show that for |B| < Bcrit there is a local energy minimum
where the magnetization points opposite the applied field,
and find Bcrit .
(b)* In part (a) we have assumed B is aligned with the
anisotropy direction of the magnetization. Describe what

can occur if these directions are not aligned.
(c) For small B, roughly how large (in energy per unit
volume) is the activation barrier for the system to get
from the local minimum to the global minimum.
(d) Can you make an estimate (in terms of actual numbers) of how big this activation barrier would be for a ferromagnetic crystal of magnetite that is a sphere of radius
1 nm ? You may use the parameters given in Exercise
31.1.e (you may need to estimate some other parameters
as well).

(a) Let cos θ = z. The function E(z) is an upside-down parabola
with maxium at z = −B/(2κ′ |M |) < 0. If this parameter is z > −1
then there are two minima, one at θ = 0 and the other at θ = π. The
absolute minium is always at θ = 0.
(b) Let us consider B again pointing in the ẑ direction and the anisotropy
direction oriented at an angle φ from ẑ. In this case we have an energy

177

functional
E/V = E0 − |M ||B| cos θ − κ′ |M |2 [cos(θ − φ)]2
If y = κ|M |/(2B) is small, then the minimum will always be close to 0. If
this parameter is greater than 1/2, then there may be a metastable minimum as well, although for y close to 1/2 then there is only a metastable
minimum if φ is close to zero. If y is large, then the metastable minima
always exist near to φ and φ + π.
(c) Consider small B, so z is close to zero. In this case, Emin =
E0 − κ|M |2 and Emax = E0 . So the activation barrier is κ|M |2 .
(d) The activation barrier is just the total anisotropy energy. This is
1.35 × 104 J/m3 and the size is (4/3)π × 1nm3 ≈ 10−26 m3 , so we have a
total energy of about 10−22 J ≈ 10Kelvin activation energy.

(21.3) Exact Domain Wall Solution*
The approximation used in Section 21.1.1 of the energy of the anisotropy (κ) term is annoyingly crude. To
be more precise, we should instead write κS 2 (cos θi )2 and
then sum over all spins i. Although this makes things
more complicated, it is still possible to solve the problem
so long as the spin twists slowly so that we can replace
the finite difference δθ with a derivative, and replace the
sum over sites with an integral. In this case, we can write
the energy of the domain wall as
(
)

2
Z
dx JS 2 a2 dθ(x)
2
2
E=
− κS [cos θ(x)]
a
2
dx

with a the lattice constant.
(a) Using calculus of variations show that this energy
is minimized when
(Ja2 /κ)d2 θ/dx2 − sin(2θ) = 0
(b) Verify that this differential equation has the solution
r 


√
κ
−1
2(x/a)
θ(x) = 2 tan
exp
J
p
thus demonstrating the same L ∼ J/κ scaling.
(c) Show that the total
of the domain wall be√
√ energy
comes Etot /(A/a2 ) = 2 2S 2 Jκ.

It is better to write the first equation as + sin2 instead of − cos2 .
These differ only by a constant, but in the former case, the background
energy is zero whereas as written the background energy is finite.
(a) This is a trivial exercise in calculus of variations. To clarify it,
write θ = θ0 + δθ and expand to linear order in δθ. Isolating the terms
inside the integral linear in δθ, and then integrating by parts to remove
the derivatives gives


2 2

−JS a



d2 θ(x)
dx2




+ 2κS cos θ(x) sin θ(x) δθ(x)
2

Setting this variation to zero immediately gives the desired√result. p
(b) This is an exercise in nasty algebra. First write b = 2(1/a) Jκ
so that we have
θ(x) = 2 tan−1 (ebx )

178 Domains and Hysteresis

From this we obtain on the left of the equation
dθ/dx

=

d2 θ/dx2

=

ebx
b
=
2bx
1+e
cosh(bx)
b2 sinh(bx)
−
cosh2 (bx)

2b

(21.1)
(21.2)

On the right of the equation we need to evaluate
sin 2θ = 2 sin θ cos θ = 4 sin

θ
θ
θ
θ
cos (cos2 − sin2 )
2
2
2
2

(21.3)

We have θ/2 = tan−1 (ebx ) so ebx = tan 2θ so we have 1 + e2bx = sec2
so
θ
2
θ
sin
2

cos

=
=

θ
2

1
√
1 + e2bx
eb x
√
1 + e2bx

Plugging these results into the prior expression for sin 2θ the gives
sin 2θ = 4

−2 sinh(bx)
ebx 1 − e2bx
=
2bx
2bx
1+e 1+e
cosh2 (bx)

Comparing this result to Eq. 21.2 along with the expression for b immediately confirms the desired solution.
(c) Starting with
(
)

2
Z
dx JS 2 a2 dθ(x)
2
2
E=
+ κS [sin θ(x)]
a
2
dx
we have dθ/dx =

b
cosh(bx)

and

sin θ = 2 cos θ/2 sin θ/2 =

2ebx
= 1/ cosh(bx)
1 + e2bx

This then naturally gives
)
(

2
Z
b
1
dx JS 2 a2
2
+ κS
E=
a
2
cosh(bx)
cosh2 (bx)
Plugging in the value of b this simplifies to
Z
Z
1
1
dx
2κS 2
dx
E = 2κS 2
=
a cosh2 bx
ab
cosh x
√
plug in the value of b to get a prefactor of S 2 2κJ and the integral of
sech2 is tanh thus giving a factor of 2, yielding the desired result.

22

Mean Field Theory
(22.1) ‡ Weiss Mean Field Theory of a Ferromagnet
Consider the spin-1/2 ferromagnetic Heisenberg Hamiltonian on the cubic lattice:
X
J X
H=−
Si · Sj + gµB B
Si
(22.1)
2
i
hi,ji

Here, J > 0, with the sum indicated with hi, ji means
summing over i and j being neighboring sites of the cubic lattice, and B is the externally applied magnetic field,
which we will assume is in the ẑ direction for simplicity.
The factor of 1/2 out front is included so that each pair of
spins is counted only once. Each site i is assumed to have
a spin Si of spin S = 1/2. Here µB is the conventional
Bohr magneton defined to be positive. The fact that the
final term has a + sign out front is from the fact that the
electron charge is negative, therefore the magnetic moment opposes the spin direction. If one were to assume
that these were nuclear spins the sign would be reversed
(and the magnitude would be much smaller due to the
larger nuclear mass).

(a) Focus your attention on one particular spin Si , and
write down an effective Hamiltonian for this spin, treating all other variables Sj with j 6= i as expectations hSj i
rather than operators.
(b) Calculate hSi i in terms of the temperature and the
fixed variables hSj i to obtain a mean-field self-consistency
equation. Write the magnetization M = |M| in terms of
hSi and the density of spins.
(c) At high temperature, find the susceptibility χ =
dM/dH = µ0 dM/dB in this approximation.
(d) Find the critical temperature in this approximation.
 Write the susceptibility in terms of this critical temperature.
(e) Show graphically that in zero external field (B = 0),
below the critical temperature, there are solutions of the
self-consistency equation with M 6= 0.
(f) Repeat parts (a)–(d) but now assuming there is an
S = 1 spin on each site (meaning that Sz takes the values
−1, 0, +1).

(a) The effective Hamiltonian for one spin-1/2 is


X
hSj i + gµb B = +Si · gµb Bi,ef f
Hi = Si · −J
j

where

gµb Bi,ef f = gµb B − J

X
j

hSj i

with the sums being over sites j neighboring site i. (Note the factor of
1/2 out front is now missing).
(b)Assuming B aligned with ẑ tells us that hSk i should be aligned
with −ẑ if nonzero. Thus we can treat these quantities as scalars rather
than vectors. We then have the usual calculation
Zi

=

hSiz i =

exp(βgµb Bi,ef f (1/2)) + exp(−βgµb Bi,ef f (1/2)) (22.2)
1
− tanh[βg µ̃b Bef f (1/2)]
2

180 Mean Field Theory

which gives us the self consistency equation
1
hSz i = − tanh[β(gµb B − JzhSz i)(1/2)]
2

(22.3)

where z is the number of neighboring spins for each site (which is z = 6
for a cubic lattice).
(c) At high T the magnetization is zero in the absence of a field. Thus
for small field we can expand the tanh for small argument to obtain.
1
hSz i = − [β(gµb B − JzhSz i)(1/2)]
2
or
hSz i =
The moment per site is thus

(−1/4)β(gµb B)
1 − βJz/4

m = −gµb hSz i =

(1/4)(gµb )2 B
kb T − Jz/4

giving the susceptibility
χ=

(1/4)(gµb )2 N
(µ0 /4)(gµb )2 N
=
kb T − Jz/4
kb (T − Tc )

(22.4)

refWeissMeanFerro where N is the density of spins.
(d) the critical temperature is the point where the susceptibility diverges, or kb Tc = Jz/4.
(e) Graphically we plot the right and left sides of Eq. 22.3 as shown
roughly in Fig. 22.1. The horizontal axis is −hSi. The left side of the
equation is a straight line, the right side is a tanh. At high enough T ,
these intersect only at hSi = 0, however at T < Tc they intersect as well
at a finite value of hSi. Note that the point at which the curves are
tangent is the critical temperature. We will show in the next problem
that below Tc the solution with hSi = 0 is unstable.
(f) The procedure is the same for S = 1. In this case, the partition
function Eq. 22.2 is replaced by
Zi

= exp(βgµb Bi,ef f ) + 1 + exp(−βgµb Bi,ef f )

leading to
hSiz i =

(22.5)

−2 sinh(x)
2 cosh(x) + 1

with x = βgµb Bef f = β(gµb B −JzhSiz i). At high T we expand for small
x we obtain the equation
hS z i =

−2
−2x
=
β(gµb B − JzhS z i)
3
3

which then gives us
hS z i =

(−2/3)βgµb B
1 − 23 βJz

181

1
2
01

02

2

1

02
01

Fig. 22.1 graphical solution of Eq. 22.3

or a susceptibility of
χ=

(2/3)µ0 (gµb )2 N
kb T − 23 Jz

and a critical temperature of kb Tc = (2/3)Jz (where N is the density of
spins).

(22.2) Bragg-Williams Approximation
This exercise provides a different approach to obtaining the Weiss mean-field equations. For simplicity we will
again assume spin 1/2 variables on each site.
Assume there are N lattice sites in the system. Let
the average spin value be hSiz i = s. Thus the probability
of a spin being an up spin is P↑ = 1/2 + s whereas the
probability of a spin being a down spin is P↓ = 1/2 − s.
The total number of up spins or down spins is then N P↑
and N P↓ respectively where there are N total lattice sites
in the system.
(a) Consider first a case where sites do not interact
with each other. In the micro-canonical ensemble, we can
count the number of configurations (microstates) which
have the given number of spin-ups and spin-downs (determined by s). Using S = kB ln Ω, calculate the entropy
of the system in the large N limit.
(b) Assuming all sites have independent probabilities
P↑ and P↓ of pointing up and down respectively, calculate
the probability that two neighboring sites will point in the

same direction and the probability that two neighboring
sites will point in opposite directions.
 Use this result to calculate an approximation to the
expectation of the Hamiltonian. Note: This is not an exact result, as in reality, sites that are next to each other
will have a tendency to have the same spin because that
will lower their energies, but we have ignored this effect
here.
(c) Putting together the results of (a) and (b) above,
derive the approximation to the free energy
F

=
=

E − T S

1
1
+ s) log( + s)
2
2

1
1
+ ( − s) log( − s)
2
2
N kB T (

+gµB Bz N s − JN zs2 /2

where z is the number of neighbors each spin has, and we
have assumed the external field B to be in the ẑ direction. (Again we assume the spin is electron spin so that

182 Mean Field Theory
the energy of a spin interacting with the external field is
+gµb B · S.)
(d) Extremize this expression with respect to the variable s to obtain the Weiss mean field equations.
 Below the critical temperature note that there are
three solutions of the mean field equations.

 By examining the second derivative of F with respect to s, show that the s = 0 solution is actually a
maximum of the free energy rather than a minimum.
 Sketch F (s) both above and below the critical temperature for B = 0. At non-zero B?

(a) The number of configurations is


N!
N
=
Ω=
N↑
(N (1/2 + s))!(N (1/2 − s))!
giving the entropy (using Stirling’s approximation)


1
1
1
1
S = kb ln Ω = −kb N ( + s) log( + s) + ( − s) log( − s)
2
2
2
2
(b) The expected energy per bond is −J(1/2)2 times the probability of
having like spins on neighboring sites plus J(1/2)2 times the probability
of having unlike spins. We thus have
E/bond = −(J/4)[P↑ P↑ +P↓ P↓ ]+2(J/4)P↑ P↓ = −(J/4)(P↑ −P↓ )2 = −Js2
If there are N z/2 bonds in the whole system (with z being the number
of neighbors of each site) we thus obtain a total energy
E = −JN zs2 /2
We must add to this the coupling to the external field which is simply
N gµB s. Adding these terms together F = E − T S gives us the desired
result.
(c)


1
1
1 dF
= gµb B − sJz + kb T log( + s) − log( − s)
N ds
2
2
Setting this to zero we obtain


1
1
β(−gµb B + sJz) = log( + s) − log( − s)
2
2
defining the left to be x, we exponentiate both sides to get
ex =

1/2 + s
1/2 − s

which we rearrange to
s=

1 ex − 1
1
1
= tanh(x/2) = tanh(β(−gµb B + sJz)/2)
x
2e +1
2
2

which is the same self-consistency condition.

183

(d) Look at the second derivative


1
1
1 d2 F
=
−Jz
+
k
T
+
b
N ds2
1/2 + s 1/2 − s
Now examine this at s = 0, we obtain
1 d2 F
N ds2

m=0

= −Jz + 4kb T

Note that this changes sign exactly at the critical temperature! Thus
below Tc the s = 0 solution becomes a maximum of the free energy rather
than a minimum as it is as above Tc . The free energy as a function of s
is plotted in Fig. 22.2.

9
01263
01268
01267
0123

0124

124

5

123

9
01264
01263
01268
01267
0123


0124

124

123

5

Fig. 22.2 Free energy (vertical axis) as a function of s (horizontal axis), at, and
below Tc . Left: in zero applied B. Right: In gµb B = 0.02Tc

(22.3) Spin S Mean Field Theory
Using the result of Exercise 19.7 use mean field theory to calculate the critical temperature for a spin S ferromagnet with a given g-factor g, having coordination

number z and nearest-neighbor exchange coupling Jex .
(It may be useful to re-solve Exercise 19.7 if you don’t
remember how this is done.)

184 Mean Field Theory

From Exercise 19.7, in a field B the resulting hSi per site (for small
perturbation B) is given by
S(S + 1)
βgµB B
3

hSi =

And in mean field theory, the effective field seen from neighboring spins
is
gµB Bef f = Jex zhSi
Thus we have at the crtical point
hSi =

S(S + 1)
βJex zhSi
3

which has the solution for the critical temperature
kb Tc =

(22.4) Low-Temperature Mean Field Theory
Consider the S = 1/2 ferromagnet mean field calculation from Exercise 22.1. At zero temperature, the magnet
is fully polarized.
(a) Calculate the magnetization in the very low temperature limit. Show that the deviation from fully polarized
becomes exponentially small as T goes to zero.
(b)* Now consider a spin S ferromagnet. Determine
the magnetization in the low T limit. You can express

S(S + 1)
Jex z
3

your result conveniently in terms of the result of Exercise
22.3.
(c)* In fact this exponential behavior is not observed
experimentally! The reason for this has to do with spinwaves, which are explored in Exercise 20.3, but are not
included in mean field theory. Using some results from
that exercise, determine (roughly) the low-temperature
behavior of the magnetization of a ferromagnet.

(a) In absence of external field, using the results of 22.1, the self
consistency equation is
1
tanh(βJzSz /2)
2

Sz =
For large argument x we have
tanh x =

1 − e−2x
≈ 1 − 2e−2x
1 + e−2x

So at low temperature we have
Sz ≈

1
− e−βJzSz
2

Since at low temperature, Sz is very close to 1/2, this becomes
Sz ≈

1
− e−βJz/2
2

185

Another way to see this is to note that, if all the neighbors are aligned,
then gµBef f = Jz 21 . Thus the energy for flipping a given spin in this
effective field is Jz/2. So the activated deviation from fully aligned is
e−βJz/2 .
(b) Probably the easiest way to do this one is to use the final comment
of part (a). If all the spins are aligned to extremize Sz , the effective field
is gµBef f = JzS. The low energy excitations excite Sz by a single step
at activation energy JzS, thus we obtain
Sz = S − e−βJzS
I have no idea what the comment about expressing the answer in terms
of 22.3!
(c) As noted in problem 20.3 the low energy spectrum of spin waves
(in absence of external field) is given by
~ω = JSa2 |k|2
Each excitation reduces the magnetization by one step. The total number of spins is V /(a3 ) and the number of excitations (the number of lost
steps) at finite T is
Z
dk
V
nB (~ωk )
(2π)3

with nB the bose factor. Thus, for each spin the reduction in magnetization is

3/2 Z ∞
Z ∞
kb T
x2
1
1
4π
2
dx x
dkk
=
a3
2 a2 k 2
3
2
2
βJS
(2π) 0
2π
JS
e −1
e
−1
0
and the final integral gives 2ζ(3) ≈ 2.4.
(22.5) Mean Field Theory for the Antiferromagnet
For this Exercise we use the molecular field (Weiss
mean field) approximation for the spin-1/2 antiferromagnetic model on a three-dimensional cubic lattice. The
full Hamiltonian is exactly that of Eq. 22.1, except that
now we have J < 0, so neighboring spins want to point
in opposite directions (compared to a ferromagnet where
J > 0 and neighboring spins want to point in the same
direction). For simplicity let us assume that the external
field points in the ẑ direction.
At mean field level, the ordered ground state of this
Hamiltonian will have alternating spins pointing up and
down respectively. Let us call the sublattices of alternating sites, sublattice A and sublattice B respectively (i.e,
A sites have lattice coordinates (i, j, k) with i + j + k odd
whereas B sites have lattice coordinates with i + j + k
even).

In mean field theory the interaction between neighboring spins is replaced by an interaction with an average
spin. Let sA = hS z iA be the average value of the spins
on sublattice A, and sB = hS z iB be the average value of
the spins on sublattice B (we assume that these are also
oriented in the ±ẑ direction).
(a) Write the mean field Hamiltonian for a single site
on sublattice A and the mean field Hamiltonian for a single site on sublattice B.
(b) Derive the mean-field self-consistency equations
sA

=

sB

=

1
tanh(β[JZsB − gµB B]/2)
2
1
tanh(β[JZsA − gµB B]/2)
2

with β = 1/(kB T ). Recall that J < 0.
(c) Let B = 0. Reduce the two self-consistency equations to a single self-consistency equation. (Hint: Use
symmetry to simplify! Try plotting sA versus sB .)

186 Mean Field Theory
(d) Assume sA,B are small near the critical point and
expand the self-consistency equations. Derive the critical
temperature Tc below which the system is antiferromagnetic (i.e., sA,B become non-zero).
(e) How does one detect antiferromagnetism experimentally?
(f) In this mean-field approximation, the magnetic susceptibility can be written as
∂(sA + sB )
χ = −(N/2)gµ0 µB lim
B→0
∂B
(why the factor of 1/2 out front?).
 Derive this susceptibility for T > Tc and write it in
terms of Tc .

 Compare your result with the analogous result for
a ferromagnet (Exercise 22). In fact, it was this type of
measurement that first suggested the existence of antiferromagnets!
(g)* For T < Tc show that
χ=

(N/4)µ0 (gµb )2 (1 − (2s)2 )
kB T + kB Tc (1 − (2s)2 )

with s the staggered moment (ie, s(T ) = |sA (T )| =
|sB (T )|).
 Compare this low T result with that of part f.
 Give a sketch of the susceptibility at all T .

This follows very much problem 22 (a)


X
hSj B i + gµb B
Hi,A = Si · −J
Hi,B

=



Si · −J

j

X
j



hSj A i + gµb B

(b) solving as above gives the desired self-consistency equations.
(c) For B = 0 we have
sA

=

sB

=

1
tanh(βJZsB /2)
2
1
tanh(βJZsA /2)
2

These are solved by
sA = −sB =

1
1
tanh(βJZsB /2) = − tanh(βJZsA /2)
2
2

Recall that J < 0. (Since tanh is an odd function these are the only
possible solutions for J < 0, which one can check graphically by plotting
sA versus sB ).
Thus the problem is reduced to
1
1
sA = − tanh(βJZsA /2) = tanh(β|J|ZsA /2)
2
2
which is identical to the relation we obtained in the ferromagnetic case.
(d) As in the ferromagnet kb Tc = z|J|/4.
(e) This staggered moment is most easily observed with neutron scattering where the scattering will be spin-dependent. Going into the antiferromagnetic phase there is a new unit cell of lattice constant 2a, thus
one sees new diffraction peaks at k = n2π/(2a) (with n odd) which are
not present above Tc .

187

(f)The factor of 1/2 is because we have only N/2 A sites or B sites.
Expanding the self consistency equations for high T we obtain
sA
sB

=
=

β[JZsB − gµb B]/4)
β[JZsA − gµb B]/4)

One can solve this system of two equations in two unknowns to obtain
sA = sB =
Thus giving

−βgµb B
4 − βJz

(1/4)µ0 (gµb )2 N
kb (T + Tc )
compare to the T − Tc factor in Eq. 22.4.
(g) This required implicit differentiation. Differentiating our two selfconsistency equations gives
χ=

∂sA
∂B
∂sB
∂B

B=0

B=0



βJZ
4

βJZ
=
4
=

∂sB
∂B
∂sA
∂B


βgµb
sech2 (βJZsB /4)
4
B=0

βgµb
sech2 (βJZsA /4)
−
4
B=0
−

Noting that sA = −sB = s so both sech terms are the same, we add
both equations together to get


βgµb
βJZ ∂(sA + sB )
∂(sA + sB )
=
−
sech2 (βJZs/4)
∂B
4
∂B
2
h=0
B=0
Making note that sech2 + tanh2 = 1, we then have
sech2 (βJZs/4) = 1 − tanh2 (βJZm/4) = 1 − (2s)2
where we have used the self-consistency equation to replace the tanh by
s. We thus have


βgµb
βJZ ∂(sA + sB )
∂(sA + sB )
(1 − (2s)2 )
=
−
∂B
4
∂B
2
B=0
B=0
This then can be rearranged into



∂(sA + sB )
1 − βJZ(1 − (2s)2 )/4 = −(βgµb /2)(1 − (2s)2 )
∂B
B=0
or

(N/4)µ0 (gµb )2 (1 − (2s)2 )
kb T + kb Tc (1 − (2s)2 )
with s the staggered moment (with N the density of spins). Above Tc ,
we have s = 0 and we have the same behavior as part f above. Below
Tc , the factor 1 − (2s)2 goes quickly down to zero, and the susceptibility
drops rapidly. At Tc there is a cusp. This behavior is illustrated in the
figure 22.3
χ=

188 Mean Field Theory

789 89999 9

99 !

012
016
015
014
013
00

012

3   3912  4


412

Fig. 22.3 Susceptibility of an Antiferromagnet.

(22.6) Correction to Mean Field*
Consider the spin-1/2 Ising ferromagnet on a cubic lattice in d dimensions. When we consider mean field theory,
we treat exactly a single spin σi and the z = 2d neighbors on each side will be considered to have an average
spin hσi. The critical temperature you calculate should
be kB Tc = Jz/4.
To improve on mean field theory, we can instead treat

a block of two connected spins σi and σi′ where the neighbors outside of this block are assumed to have the average
spin hσi. Each of the spins in the block has 2d − 1 such
averaged neighbors. Use this improved mean field theory to write a new equation for the critical temperature
(it will be a transcendental equation). Is this improved
estimate of the critical temperature higher or lower than
that calculated in the more simple mean-field model?

We have to consider four configurations in our partition function:
(↑, ↑), (↑, ↓), (↓, ↑), (↓, ↓)
Correspondingly, the partition function is
Z = e−β(Jhσi(z−1)+gµB B)−βJ/4 + 2eβJ/4 + eβ(Jhσi(z−1)+gµB B)−βJ/4
From this we calculate the magnetization per site (note that this partition function represents two sites), and then setting B to zero we obtain
m = hσi =

e−βJ/4 sinh(β(Jhσi(z − 1))
2e−βJ/4 cosh(β(Jhσi(z − 1)) + 2eβJ/4

5

189

Expanding for small hσi we find equality (which finds Tc ) when


e−βJ/4
J(z − 1)
kB T =
4
cosh(βJ/4)
The final factor in brackets is always less than unity, so this expression
always gives a lower prediction for Tc than our prior mean field theory.

Magnetism from
Interactions: The Hubbard
Model

(23.1) Itinerant Ferromagnetism
(a.i) Review 1: For a three-dimensional tight binding
model on a cubic lattice, calculate the effective mass in
terms of the hopping matrix element t between nearest
neighbors and the lattice constant a.
(a.ii) Review 2: Assuming the density n of electrons
in this tight binding band is very low, one can view the
electrons as being free electrons with this effective mass
m∗ . For a system of spinless electrons show that the total
energy per unit volume (at zero temperature) is given by
5/3

E/V = nEmin + Cn

where Emin is the energy of the bottom of the band.
 Calculate the constant C.
(b) Let the density of spin-up electrons be n↑ and the
density of spin-down electrons be n↓ . We can write these
as
n↑

=

(n/2)(1 + α)

(23.1)

n↓

=

(n/2)(1 − α)

(23.2)

where the total net magnetization of the system is given
by
M = −µb nαB.
Using the result of part (a), fixing the total density of
electrons in the system n,

 calculate the total energy of the system per unit
volume as a function of α.
 Expand your result to fourth order in α.
 Show that α = 0 gives the lowest possible energy.
 Argue that this remains true to all orders in α
(c) Now consider adding a Hubbard interaction term
X i i
HHubbard = U
N↑ N↓
i

E = E0 − 2t [cos(kx a) + cos(ky a) + cos(kz a)]
E = Emin + ta2 |k|2

Nσi

with U ≥ 0 where
is the number of electrons of spin
σ on site i.
Calculate the expectation value of this interaction term
given that the up and down electrons form Fermi seas
with densities n↑ and n↓ as given by Eqns. 23.1 and 23.2.
 Write this energy in terms of α.
(d) Adding together the kinetic energy calculated in
part b with the interaction energy calculated in part c,
determine the value of U for which it is favorable for α to
become non-zero.
 For values of U not too much bigger than this value,
calculate the magnetization as a function of U .
 Explain why this calculation is only an approximation.
(e) Consider now a two-dimensional tight binding
model on a square lattice with a Hubbard interaction.
How does this alter the result of part (d)?

(a.i) The tight binding spectrum is

which we expand to get

23

192 Magnetism from Interactions: The Hubbard Model

where Emin = E0 − 6t is the bottom of the band. We thus have
ta2 |k|2 = ~2 |k|2 /(2m∗ )
or

~2
2ta2
(a.ii) As we have calculated many times (note we are considering spin
polarized electrons here, so there is no factor of 2 out front)
Z kf
Z kf
3
V
V
V 4πkf
2
N=
dk
=
4πk
dk
=
(2π)3 0
(2π)3 0
(2π)3 3
m∗ =

or

kf = (6π 2 n)1/3
The total energy is given by
Z kf 2 2
Z kf 2 2
V
V
~ k
~ k
E − Emin N =
dk
=
4πk 2 dk
(2π)3 0 2m∗
(2π)3 0 2m∗
~2 4πkf5
V ~2 5
V
k
=
=
3
∗
(2π) 2m
5
20π 2 m∗ f
Thus we have
E/V = Emin n + Cn5/3
with

ta2
1 ~2
2 5/3
(6π
)
=
(6π 2 )5/3
20π 2 m∗
10π 2
(b) Note that n remains fixed. So we have
h
i
5/3
5/3
E/V − Emin n = +C n↑ + n↓
i
h
= C(n/2)5/3 (1 + α)5/3 + (1 − α)5/3
C=

Taylor expanding here, note that the odd terms of the expansion cancel
leaving only even terms

E/V − Emin n = 2C(n/2)5/3 1 + (1/2!)(5/3)(2/3)α2+

(1/4!)(5/3)(2/3)(−1/3)(−4/3)α4 + . . .


= 2C(n/2)5/3 1 + (5/9)α2 + (5/243)α4 + . . .

Note that successive terms of the expansion always have positive coefficient.
(c) The expected number of electrons per unit site is na3 and similarly n↑ a3 and n↓ a3 are the expected number of spin up and spin down
electrons per site. Thus, the expectation of the hubbard interaction per
unit volume is
Ehubbard /V

=
=

(U/a3 )(n↑ a3 )(n↓ a3 ) = U a3 n↑ n↓
U a3 (n/2)2 (1 + α)(1 − α) = U a3 (n/2)2 (1 − α2 )

193

(d) We thus obtain a total energy given by
h
i
i
h
Etotal /V = constant+α2 2C(5/9)(n/2)5/3 − U a3 (n/2)2 +α4 2C(n/2)5/3 (5/243) +. . .
(23.3)
And taking the limit of small α we see that α = 0 is the solution only
for
U ≤ 2C(5/9)(n/2)−1/3
For U not too much bigger than this, we can use the quartic form of the
energy given in Eq. 23.3. Minimizing with respect to α gives
s
−2C(5/9)(n/2)5/3 + U a3 (n/2)2
α=
4C(n/2)5/3 (5/243)

and M = −µb nα.
(e) In the 2d case, the key here is that N ∼ kf2 and ǫ ∼ EF2 , so
E/V ∼ n2 . As a result, we have the kinetic term given by
E = C̃[(1 + α)2 + (1 − α2 )] = 2C̃[1 + α2 ]
and there is no quartic term. Thus, we have a total energy of the form
Etotal = const + α2 (C̃ − K̃U )
Once U becomes large enough that α = 0 is not the lowest energy
solution, then α immediately goes to its maximum possible value 1.
Thus, the transition is discontinuous going suddenly from unpolarized
spins to fully polarized spins.

(23.2) Antiferromagnetism in the Hubbard
Model
Consider a tight binding model with hopping t and a
strong Hubbard interaction
X i i
N↑ N↓ .
HHubbard = U
i

(a) If there is one electron per site, if the interaction term

U is very strong, explain qualitatively why the system
must be an insulator.
(b) On a square lattice, with one electron per site,
and large U , use second-order perturbation theory to determine the energy difference between the ferromagnetic
state and the antiferromagnetic state. Which one is lower
energy?

(a) If U is strong enough, there must always be one electron per site.
This makes a traffic jam of electrons where no one can move (so long as
there is no doping).
(b) For the antiferromagnet, each spin can make a virtual excursion
to each of the neighboring sites, at an energy cost of U . If the hopping
is t at 2nd order in perturbation theory, this gives a reduction in the
ground state energy, per site, of
−2zt2 /U
where z is the number of neighbors. For a ferromagnet, no excursion is
allowed by the Pauli principle, so it is higher energy.



Source Exif Data:
File Type                       : PDF
File Type Extension             : pdf
MIME Type                       : application/pdf
PDF Version                     : 1.4
Linearized                      : No
Page Count                      : 199
XMP Toolkit                     : XMP toolkit 2.9.1-13, framework 1.6
About                           : uuid:0d5bcb74-7142-11f0-0000-e7b0455fc944
Producer                        : GPL Ghostscript 9.10
Modify Date                     : 2015:08:02 15:48:14+01:00
Create Date                     : Sat -Ja-n T 4: 1:7:00::2
Creator Tool                    : gnuplot 4.4 patchlevel 3
Document ID                     : uuid:0d5bcb74-7142-11f0-0000-e7b0455fc944
Format                          : application/pdf
Title                           : BookSol.dvi
Creator                         : simon
Description                     : gnuplot plot
Subject                         : gnuplot plot
Author                          : simon
EXIF Metadata provided by EXIF.tools

Navigation menu