Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

October 14th, 2010

Class 11

Crystal Vibrations: Monoatomic basis

Consider a family of planes in the crystal and assume one of these planes is moved out of its
equilibrium position. Assuming all the planes are connected to each other by elastic interactions,
this action will produce a force in all other planes. Assuming a simple harmonic model, a model
where a force is a linear function of the displacement, the force on plane s is proportional to the
change in the distance between this and another plane located p planes from it (remember that for
a spring F=-k(x-xo)) and E= ½k(x-xo)2). Considering only consecutive planes, p=±1

d 2 us
Fs  M  C us 1  us   C us 1  us   C us 1  us 1  2us 
dt 2

where C is the elastic constant. Since in the equation above the second derivative of us is
proportional to us, the function itself is a sine, a cosine or a complex exponential. It is easier to
elaborate on this if we think that the above is the force on 1 atom in the plane rather than on the
entire plane.

Assuming an exponential function of time as a solution in time:


us a ,t   us a  exp  it  where us was used for both the space-time function and the space
function of the displacement to follow the book convention.

By using the above definition


 M 2us  C us1  us1  2us  where now, u is the oscillation amplitude.
This is a difference equation in the displacement and has traveling waves solutions of the form

us  0,1 a   u exp is  0,1Ka 

so  M 2u exp is Ka   uCexp is  1Ka   exp is  1Ka   2 exp is Ka 

What gives, after some simple algebra and canceling u exp is Ka  at both ends of the equal

 M 2  Cexp iKa   exp  iKa   2  2C1  cosKa  or using that 1  cos  2 sin 2  1  
2 

 Ka   Ka 
what leads to the dispersion relation  K  
4C 4C
2  sin 2   sin 
M  2  M  2 

We can sample the limits of this equation. For the long wavelength limit (small K),

1
  

cosKa   1 
1
Ka 2   and  2  2C 1  cos Ka  C Ka 2
2 M M

Ca 2
or   K
M

On the other end, the wave is stationary for K=±/a. These are the limits of the first Brillouin
zone.

for Kmax=/a  2/min=/a or min=2a and


  
us  u exp   is a   u exp  is   u 1
s

 a 
Thus, consecutive planes move in opposite phase.

All the physics of the problem is already in the first Brillouin zone, second and others are
redundant and do not produce new information.

To understand this, notice that us are actually the amplitude of the oscillation at each atom, thus
any wave for which the amplitude at each atomic position is NOT a new solution since there are
not atoms to see intermediate amplitudes. Notice in Figure 5 (page 93) that the two waves predict
the same amplitude at each point.

2
October 19th, 2010
Class 12

Crystal Vibrations: Monoatomic basis


We found that the dispersion relation for monoatomic basis was
 Ka   Ka 
 what leads to the dispersion relation  K  
4C 4C
2  sin 2  sin 
M  2  M  2 
That for the long wavelength limit (small K) leads to,

  

cosKa   1 
1
Ka 2   and  2  2C 1  cos Ka  C Ka 2
2 M M

Ca 2
or   K
M

and for K=±/a leads to stationary waves (the derivative is zero at those points). These are the
limits of the first Brillouin zone.

for Kmax=/a  2/min=/a or min=2a and


  
us  u exp   is a   u exp  is   u 1
s

 a 
Thus, consecutive planes move in opposite phase.

All the physics of the problem is already in the first Brillouin zone, second and others are
redundant and do not produce new information.

To understand this, notice that us are actually the amplitude of the oscillation at each atom, thus
any wave for which the amplitude at each atomic position is the same and NOT a new solution
since there are no atoms to see intermediate amplitudes. Notice in Figure 5 (page 93) that the two
waves predict the same amplitude at each point.

Notice that the ratio of the amplitude of the atomic displacement between consecutive planes is

us 1 u exp i s  1Ka 


  exp iKa 
us u exp i sKa 
All independent values of this ratio are obtained for Ka between - and . Given any value of K
outside the mentioned [- ,] interval, there exist another wavevector K’=K-2n/a (with n and
integer) that it is in the interval. For this wavevector K’ in the first Brillouin zone

  2n    2n 
 exp iKa   exp  i  K ' a   exp iKa exp  i a   exp iK ' a 
us 1

us   a    a 

Thus the relative amplitude is exactly the same for K as it is for a K’ in the first Brillouin zone.

1
Group Velocity
When a packet of waves with different wave lengths (and frequency) travel together, it is not the
speed of each of the components what matters, but the velocity of the group (imagine a pulse
sent through space). In non-dispersive mediums, frequency and wavevector are proportional and,
as we will see, phase velocity (the one of individual components) is equal to group velocity (like
light in vacuum), all the components travel at the same speed. But in general that is not the case,
each component travels at a different speed, and the actual package travels at a speed giving by:

d
vg 
dK
To see this, consider two waves with a slightly different frequency and wavevector (and just for
illustration purpose, the same amplitude too). The composed waves will be:

ue i t Kx  ue i ' t K' x 

This equation can be re-arranged to give

  K 
2ue i  t  Kx  cos  t x  Where  and K are the average frequency and wave vector.
 2 2 
That is a high frequency wave with a low frequency envelope. The velocity at which the
package moves is the velocity of the envelope, thus


vg  thus the vg indeed give the speed of the package.
K

If for a particular medium, the frequency and the wavevector are proportional, then group and
phase velocity are the same, and equal to the speed of each component. In this case the medium
is non-dispersive and all the components move at the same speed, thus that package does not
change shape. If instead different components travel at different speeds, the package will spread,
or disperse as it moves on (thus the relationship between  and K is known as the dispersion
relationship)
d
For a continuum of frequencies, v g 
dK

In the case of phonons, using the dispersion relation obtained above.

 Ca 2   Ka 
 cos  0  Ka  
d  M   2 
vg   vg is discontinuous for K=0 where it goes from -1
dK  Ca 2   Ka 
 cos     Ka  0

 M   2 
to 1 in going from negative values of K to positive values of K. Which is not surprising, it only
means that the wave now moves in the opposite direction.

In the limits of the Brillouin zone, Ka=± and vg=0, consistent with standing waves.

2
As we saw before, for the long wavelength limit (small K),

Ca 2
 K
M

Thus the group velocity is equal to /K. This can be understood by realizing that in the long
wave length limit K~0, >>a and the interaction with the lattice is minimal leading to a non-
dispersive behavior.

Force constant from the interaction with non-consecutive planes


If we consider now the effect of a plane p places from the reference plane s then

 

us p a   u exp i s  p Ka from where the dispersion condition can be easily shown to be

C p 1  cos  pKa  and considering all the planes


2
2 
M

 C 1  cos  pKa 
2
2  p
M p 0

Where Cp, the force constant associated with the interaction between plane s and p is given by:
(See derivations provided)

 /a
 2 cos pKa dK
Ma
Cp   
2  / a

More than one type of atoms per primitive basis


When the basis consist on one atom, there are three independent vibration modes, 1 longitudinal
(the atoms in the plane oscillate in a direction perpendicular to the plane) and 2 transversal where
the atoms oscillates parallel to the plane (the two modes come from the two dimensions of the
plane). When the basis have more than one type of atom, then each type of atom give rise to
three independent vibration modes, thus in a crystal with a p-atoms basis, there are 3p
independent modes. Out of the 3p modes, 3 and called acoustical modes and 3p-3 are called
optical modes for a reason that will be evident soon. Also both optical and acoustical modes
have transversal and longitudinal in the rate of 2-to-1.

In what follows we will consider a cubic crystal with a two atoms base with alternate planes of
atoms 1 (with mass M1) and atoms 2 (with mass M2).

d 2u s
Atom 1 : M1  C vs  vs1  2us 
dt 2
d 2v
Atom 2 : M 2 2s  C us  us 1  2vs 
dt

3
Solutions to the above equations can be proposed to be of the form
u s  u exp it  sKa  and v s  v exp it  sKa 
Where a is the distance between plane of atoms of the same kind

Using these solutions in the equation above,

 M1 2u  Cv  v exp iKa   2u   M 2 2v  C u  u expiKa   2v 

What leads to a set of two homogeneous equations with two unknowns,

2C  M  u  C 1  e   0
1
2  iKa

 C 1  e u  2C  M    0
iKa
1
2

This set of equations has non-trivial solution only if the following determinant is zero

2C  M1 2 
 C1  exp  iKa 
 C1  exp iKa  2C  M   2
2

what leads to
M 1 M 2 4  2C M 1  M 2  2  2C 2 1  cos Ka   0

with solutions

2C M 1  M 2   4C 2 M 1  M 2   8M 1 M 2C 2 1  cos Ka 
2

2 
2M 1M 2

For the limit Ka << 1 cos(Ka)~1- ½ (Ka)2 or 1- cos(Ka)~ ½ (Ka)2

2C M 1  M 2   4C 2 M 1  M 2   4 M 1 M 2C 2 Ka 
2 2

 
2

2M 1M 2

bx
and for small values of Ka we can expand the square root in series a 2  bx ~ a  for small
2a
values of x

 4M 1M 2C 2 Ka  
2
2C M 1  M 2    2C M 1  M 2  
 
 4C M  M  

2  1 2

2 M 1M 2

what provide the two solutions for small Ka

4
4M 1M 2C 2 Ka 
2
4C M 1  M 2  
4C M 1  M 2  4C M 1  M 2   1 1 
2    2C   
2 M 1M 2 2 M 1M 2  1
M M 2 

To the first order in Ka


and

4 M 1 M 2C 2 Ka 
2

4C M 1  M 2 
2  
C
Ka 2
2M 1M 2 2 M 1  M 2 

By using these solutions in the set of equations for u and v and solving for them, it can be seen
u M
that for the first of the solutions in the K~0 limit,   2 , thus the consecutive planes vibrate
v M1
in opposite direction keeping the center of mass fixed. If the two atoms are ions of opposite sign,
this mode can be excited by an electromagnetic field, thus this mode is called optical mode. The
other solution for the K~0 leads to u=v thus the planes move in the same direction as it is for
long wavelengths sound waves, thus this mode is known as acoustical mode.

For the other limit, Ka=± cos(Ka)=-1 and thus the solutions are :

2C M 1  M 2   4C 2 M 1  M 2   16 M 1 M 2C 2 2C M 1  M 2   2C M 1  M 2 
2

 
2

2M 1M 2 2M 1M 2

2C 2C
Or  2  and  2 
M2 M1

This leads to a solution where each type of atom vibrates independently from the other. One
solution correspond to one type of atom vibrating while the other is at rest and viceversa

There is a gap between optical and acoustical mode at the boundary of the first Brillouin zone
where frequencies between (2C/M2)2 and (2C/M1)2 are not allowed.

Figure 8 shows both modes. Also in Figure 8, it can be seen that a 3-D crystal, longitudinal and
transversal vibrations split in frequency.

In a crystal with a p atom base there are 3 acoustical modes (1 longitudinal and 2 transversal) and
3p-3 optical modes.

5
Quantization of Elastic Waves
By describing vibration in the lattice by a quantum harmonic oscillator model, it is observed that
the lattice energy is quantized, in other words, the energy of a lattice vibrating with a single
frequency  is:

 1
   n  
 2

That means that no just any value of energy is allowed but the ones in that equation. For n=0 the
lowest energy state is obtained and it is known as zero point energy. Other values of n represent
exited states of the crystal vibration where n photons populate the state. Phonons are not
subjected to the Pauli Exclusion Principle (as photons are not) thus any number of phonons can
occupy a given state, however you add phonons one by one, a fraction of a phonon would not do.

The quantization of the energy has the consequence to produce quantization on the amplitude of
the oscillation.

To see this consider a vibration described by

u  uo cos Kx cos t

As for a harmonic oscillator, the temporal average of the energy is half kinetic and half potential.

1  u 
2

Lets consider the kinetic energy density K      , where  is the mass density
2  t 

The volume integral of the kinetic energy density gives (notice that the integral of the cos2 is Lx/2
while the integral in y and z gives Ly and Lz), thus the volume integral is V/2

¼ V2uo2sin2t

with a time average of 1 V2uo2 thus (since <sin2t>=1/2). Making this average kinetic
8
energy equal to ½ of the total energy

1 V2uo2= ½(n+½) ħ
8

we obtain that

uo2= 4(n+½) ħ/V= 4(n+½) ħ/M

Thus the vibration amplitude is related to the occupancy n of the mode.

The solution of the motion’s equation for atoms in the crystal gives 2 so, what is the sign of ?
Conventionally  is assumed positive. For unstable structures, 2 is negative and  is
imaginary.

6
Phonon Momentum
A phonon of wavevector K can interact with photons, neutrons, electrons, and other particles as
if they had a momentum ħK. Looking at this in a different way, if a photon (or other quantum
particle) collides with the crystal, without losing energy, the selection rule for the incoming
particle wave vector imposes that

k’=k+G

No energy is transferred to the crystal but the crystal as a whole recoils with a momentum -ħG.

If inelastic scattering occurs, some energy is indeed lost by the incoming particle to the crystal
that will excite phonon modes, besides the fact that the transferred energy should be at least ħ,
the energy of a phonon of frequency, there is a generalized selection rule that can be written as

k’+K=k+G

where K is the phonon’s wave vector, thus the wavevector of the scattered particle is conditioned
to the wavevector of the created phonon such that the above equation is satisfied. In the same
way, if a phonon is annihilated in the interaction, the equivalent selection rule is:

k’=K+k+G

Neutrons are often used to determine the dispersion relations for the phonon. After scattering,
the energy and direction of scattered neutrons is recorded thus k and k’ are determined.

The kinetic energy of the neutrons is p2/2Mn and the momentum is p= ħk, so conservation of
energy imposes

 2 k 2  2 k' 2
   Where ħ is the energy of the phonon created (+) or absorbed (-) from
2M n 2M n
where  is obtained.

By knowing the crystal lattice (G) and from the formula k’±K=k+G, K is determined, where G
is chosen such that K is in the first Brillouin zone.

The ± sign applies to phonon created or absorbed.

7
October 21th, 2010
Class 13

Phonon Heat Capacity


The heat capacity of a material is the ratio between the heat (energy) provided to the material and
the change in its temperature. In other words, the amount of heat necessary to increase the
temperature of said material in 1 degree. A good analogy is a liquid container, the “heat” is the
water poured into the container, the temperature is the liquid level on the container, the heat
capacity is a relationship between the two, how much water you need to add so the level raisee 1
unit. A good analogy for a large heat capacity material in comparison to a small heat capacity is
a wide and a narrow container, in a narrow container (small heat capacity), a small amount of
water will be enough to raise the level significantly, while for a wide container (large heat
capacity), more water is needed to produce the same water level.

When heat is added to a material, the energy can be used for many things, one is to excite
phonon modes and that is what we will be concerned about here. The heat capacity can be
defined at a constant volume or a constant pressure, while experimentally heat capacity is often
measured in a constant pressure condition, the heat capacity at a constant volume is more
fundamental (since it only involves changes in internal energy as opposed to the case at constant
pressure where some of the energy is used to change the lattice volume involving mechanical
work). The two heat capacities CV and Cp are related such that Cp-CV=f(T) thus knowing one,
the other can be obtained.

 U 
CV    Where U is the energy. The different contributions to the heat capacity can be
 T V
obtained by taking the derivative with respect to temperature of the corresponding energy.
Particularly, the contribution to the heat capacity of the phonon modes, is known as the lattice
heat capacity Clat. The total phonon energy (ignoring the zero point energy) is:

U lat   nK ,p K ,p
K p

<nK,p> is the average occupancy of each of the phonon modes at thermal equilibrium, K is the
wave number related with  through the dispersion relation, and p in the polarization (acoustical,
optical, transversal, longitudinal)
The average occupancy of a given mode is given by the Plank distribution. (See book for
the steps leading to this equation)

1
n 
  
exp    1
 B 
k T
Combining the two

1
K ,p
U lat  
 K ,p 
exp    1
K p

 k BT 

At this point it is important to remember that the K’s are discrete but too close to each other and
can be considered a continuum, but they form a continuum and so an integral is more appropriate
than a summation. If the number of modes in the frequency range +d with polarization p is
given by Dp()d, then


U lat    dD p  
  
exp    1
p

 k BT 

Where of course the dispersion relation allows us to integrate in  instead of K. Thus

U lat x 2e x
Clat   k B   dD p  
T p e x  12

Where x 
k BT

Now the heat capacity can be obtained if the density of modes is known. Coming up next, we
revise some models to determine the density of states and with that to solve the integral and
obtain the density of states.

Density of States in One Dimension


The values of K that are possible in a given lattice would be a continuum unless appropriate
boundary conditions are applied. One common type is the periodic boundary condition, in one
dimension that simply means that the first and last point in the chain are entirely equivalent (thus
starting the sequence again). A condition like that implies

u(sa)= u(sa+L)

By imposing this condition to the phonon wave function in a line of N+1 sites, such that site 1
and site N+1 are equivalent by the boundary conditions (N independent sites)

uoei(Ksa-(K)t)= uoei(K(sa+L)-(K)t)=uoei(Ksa-(K)t) eiKL

What implies that eiKL=1 or KL=2n or

2
2 4 6 N
K  0,  , , , ,  Notice that for ±N/L (where L/N=a) we have e±i where +
L L L L
and - are actually the same angle and thus the same solution, so there are exactly N independent
solutions. For n > N the same solutions are obtained.

2 4 6 N
K  0,  , , , ,
L L L L

Thus for a chain of N+1 atoms with periodic boundary conditions, there are exactly N states, 1
per each atom moving independently in the chain.

Also notice that K=±n/L actually refers to a wave with the same frequency only propagating in
opposite directions.

The distance between two points in the K-space is 2/L. The number of states per unit of length
in the K space is then L/2 and thus the number of states of absolute value in the range [0,K] is
then given by

L
NK  2  K
2

(where the 2 account for the two waves with the same K propagating in opposite direction ±n/a).

The density of states, “number of states per unit of length” can be defined as dN/dK for N large
enough that the states can be consider a continuum. So

L d
D d  dN K 
dN K dK L dK
d  2 d 
dK d 2 d  d
dK

The dependence of the density of states in terms of ω has been forced, but now the density of
states is expressed in terms of the group velocity which can be obtained if the dispersion relation
is known.
Density of States in Three Dimensions
By applying periodic boundary conditions in the same way that we did for 1-D, the number of
states inside a sphere of radius K is:
 L   4K 
3 3
N K      where L3=V
 2   3 

Thus the density of states is then given by:

VK 2 d
D d 
dN K dN K dK
d  d 
d dK d 2 2 d
dK

3
The problem is now reduced to finding the group velocity (or the dispersion relation) for a
particular crystal and to apply the equations above.

Debye Model for the Density of States

In this model, the velocity of sound is taken as a constant for each polarization type, thus
VK 2 1 V 2
   p K and D   2 
2 d 2 3p
dK
Now we have all the ingredients and we can calculate the lattice energy (under the Debye
approximation).
d  d  V 2  
U lat    dD p       d  2 3  
0 0  2   kBT
p e kBT  1 p  p e 1
Where D is the Debye frequency and correspond to that frequency associated with the
maximum value of K, KD. Now, if we assume that the velocity is independent of the polarization
mode and we consider three polarization modes (3 acoustical, 1 per dimension)
 3V   d 3  3Vk 4T 4  xd x3
U lat   2 3   d    2 B 3 3   dx x
 2   0 e k BT  1  2   
0 e 1

Where the following variable substitutions has been made

from where

Here it is convenient to introduce another definition, the Debye temperature 

what leads to

The Debye temperature is then the temperature at which xD=1 or the temperature at which the
maximum phonon energy is equal to kBT.
 L   4 
3 3
Now using that N K     3 
 to obtain an expression for .
 2   3 

Where N is the total number of occupied states


With these definitions, now
3 T  
3 3
 3V  d T  x3 4
xd
U lat   2 3   d   9 Nk BT    d x  Nk BT  
 2   0 e kBT  1   0 e 1    15

What leads to

4
U lat 233.783 3
CV   T
T 3

This is known as the Debye T3 approximation. See figure 9, page 115 for the heat capacity of
Argon, where this law is in very good agreement with the experiment. For Argon, the Debye
temperature is 92K. The agreement with this law worsens as T approaches the Debye
temperature, the law is most approximately met for T within 1-2% of the Debye temperature.
Thus the Debye approximation is good for very low temperature (compare to the Debye
temperature).

Einstein Model of the Density of States

This model considers N oscillators all vibrating at the same frequency o and thus
D()=Nδ(-o)
 
U lat    dN   o    3N o o
p e kBT  1 e kBT  1
   o
CV  3Nk B  2 o  2

 k BT   o k BT


 e 
 1
 
For large T, CV~3NkB, the Dulong and Petit Value.

In summary, the Einstein model accounts for the high temperature limit while the Debye model
accounts for the low temperature regime.

5
October 25th, 2010
Class 14

Phonon Heat Capacity


 U 
CV    Where U is the energy. The different contributions to the heat capacity can be
 T V
obtained by taking the derivative with respect to temperature of the corresponding energy.
Particularly, the contribution to the heat capacity of the phonon modes, is known as the lattice
heat capacity Clat.


U lat    dD p  
  
exp    1
p

 k BT 

Density of States in One Dimension


We found that,

L
NK  2  K
2

Thus,

L d
D d  dN K 
dN K dK L dK
d  2 d 
dK d 2 d  d
dK
Density of States in Three Dimensions
By applying periodic boundary conditions in the same way that we did for 1-D, the number of
states inside a sphere of radius K is:
 L   4K 
3 3
N K      where L3=V
 2   3 

Thus the density of states is then given by:

VK 2 d
D d 
dN K dN K dK
d  d 
d dK d 2 2 d
dK
The problem is now reduced to finding the group velocity (or the dispersion relation) for a
particular crystal and to apply the equations above.

Debye Model for the Density of States


In this model, valid for small values of T, the velocity of sound is taken as a constant for each
polarization type, thus
V 2
   p K and D  
2 3p

1
T  
4 3

U lat  Nk BT   where
   15

What leads to
U lat 233.783 3
CV   T
T 3

This is known as the Debye T3 approximation. The agreement with this law worsens as T
approaches the Debye temperature, the law is most approximately met for T within 1-2% of the
Debye temperature. Thus the Debye approximation is good for very low temperature (compare
to the Debye temperature).

Einstein Model of the Density of States


This model considers N oscillators all vibrating at the same frequency o and thus
D()=Nδ(-o)
 
U lat    dN   o    3N o o
p e kBT  1 e kBT  1
   o
CV  3Nk B  2 o2  2

 k BT   o k BT


 e  1
 
For large T, CV~3NkB, the Dulong and Petit Value.

In summary, the Einstein model accounts for the high temperature limit while the Debye model
accounts for the low temperature regime.

Anharmonic Crystal Interactions


The harmonic approximation is just that, an approximation and many of the disagreement
between the theoretical predictions (within the harmonic approximation) and the experiment can
be tracked down to the limitations of this approach. For instance, the thermal expansion
predicted by the harmonic model is zero. To see this, the average displacement is
2
  cx

k BT
xe dx
x  
2
  cx

k BT
e dx

Where the Boltzmann distribution was used to weight the displacements according to their
thermodynamical probability. However, the integral on the top is the integral of an odd
functions (a function for which f(x)=- f(-x)) integrated is a symmetrical interval, and thus it is
zero. The denominator is the intergral of a non-zero, positive defined function what warranties
that the denominator does not go to zero. So the average displacement is zero at ALL
temperatures and the material is NOT predicted to expand with temperature.

2
If anharmonic terms are added to the energy (terms of power larger than 2 of the displacement),
this prediction changes. For instance

U x   cx 2  gx3  fx 4 then

cx 2  gx3  fx 4
 


k BT
xe dx
x  
cx  gx  fx
2 3 4
 


k BT
e dx


If we assume the cubic and quartic terms are very small compared to kBT, we can expand in
series to get

cx 2  gx3  fx 4 cx 2
     gx 4 fx 5 
dx   1 / 2 5 / 2 kBT 3 / 2
3 g
 dx   e  x  
k BT k BT
xe
 
 kBT kBT  4 c
and
cx 2  gx3  fx 4 cx 2
 gx3 fx 4   
1/ 2
   

dx  kBT 1 / 2   1 / 2 5 / 2 kBT 3 / 2   kBT 
3 f
 dx   e 1  
k BT k BT
e
 
 kBT kBT  c 4 c c 

Leading to
3g
x  2 kBT
4c

Thermal Conductivity
Without anharmonicity there is no interaction between phonons, without this interaction thermal
conductivity would be infinite as shown below.

The steady state flow of heat is given by


dT
ju   K
dx
This diffusion process requires a tortuous transport as opposed to ballistic, otherwise the flux
would only depend on the difference in temperature between the ends of the sample and not on
the gradient, (like electrons accelerated by an electric field, their kinetic energy change only
depends on the voltage difference and no on how far they move). So this behavior requires “heat
carriers” to often collide and scatter.

Using kinetic theory we can find a model for K.

Consider a gas of particles (like phonons) with n particles per unit of volume moving with
average velocity in the x direction of vx , in equilibrium, half will move in the positive

3
1
direction and half in the negative direction. Thus, the flux of particle is j p  n vx where
2
there is a flux of equal magnitude on each direction.

If c is the heat capacity per particle, each particle will lose an amount of energy of cΔT when it
moves from a place at temperature T+ΔT to a place at temperature T. If lx is the distance
between collisions, then
dT dT
T  lx  vx where τ is the average time between collision
dx dx

So, the flux of energy jU


dT 1 dT 1 dT
jU   j pcT  n vx2 c   nv 2c   Cvl
dx 3 dx 3 dx
Where nc=C (heat capacity per unit volume), and vτ=l
1
So K  Cvl
3

Thus, if there is no interaction with the lattice, l is infinite and the thermal conductivity is infinite.

The interaction occurs because of the third order


anharmonic term presence of one phonon causes
the modulation in space and time of the lattice
elastic constant, a second phonon sees that strain
and it is scattered producing then a third phonon in
a three-phonon process.

K1+ K2 =K3
Picture from wikipedia

This process, although it explains phonon collision and finite thermal conductivity, it is still not
enough to account for phonon local thermal equilibrium as the total momentum of the phonon
gas remains unchanged (thus one end of a metal cannot get in equilibrium with a temperature
source) any momentum lost by one phonon goes to another. An interaction with the lattice as a
whole is necessary, thus the process must be of the form

K1+ K2 =K3+G

These processes are known as Umklapp processes. However this is not a forced phenomena,
remember that all physically meaningful values for the phonon wave number K must be in the
first brillouin zone, it is possible that the K1+ K2 leads to a phonon with wave number K3 outside
the zone, but that is not a physically meaningful state, thus a G vector must be used to bring the
phonon back into the first Brillouin zone. Looking at this issues from a different point of view,
when two colliding phonons that have a wave number so large that the sum will lead to a
momentum outside the first Brillouin zone, the process is not possible as a phonon with such a
momentum cannot exist and the lattice as a whole must absorb some momentum, thus the total
momentum of the phonon gas changes and the phonons can thermalize. At high T, the

4
population of high momentum states is very large and Umklapp processes dominate. In this
regime the number of phonon is proportional to T (See Einstein model) and the number of
collision is proportional to the population, the mean free path is proportional to 1/T, thus thermal
conductivity is proportional to 1/T. At low T, most phonons have low enough momentum that it
is very unlikely that in a collision a phonon is created with momentum outside the first Brillouin
zone and thus heat conduction is dominated by normal processes.
Free Electron Theory
Many of the properties of metals can be understood if we consider the electrons in the metal as
particles that are free to move in a box. This model it is clearly an approximation since it
assumes two completely wrong hypothesis, 1) electrons do not interact with each other, 2)
electron do not interact with the lattice. Both are totally wrong even in classical terms since
electrons and the lattice are charged, electrons are negative and the lattice bears a net positive
charge. Surprisingly, this theory explain acceptably well some of the properties of metals as we
will see below.

Consider alkali metals, Li, Na, K, Rb, Cs. All of them have the structure of an inert gas
plus 1 electron on a s level. For these metals it is not hard to share that electron and in treating
these system theoretically, it is assumed each atom contribute with 1 electron to the electron gas.
All those electrons now navigate freely around the entire metal.

Classical theory of metals, produced long before quantum mechanics (Drude’s model)
surprisingly predict fairly well conduction properties, ohms law, thermal conduction and the
relationship between them (although the proportionality constant was a little bit off compared to
the experiment). However it overestimates by about 2 orders of magnitudes the electronic
contribution of the heat capacity. How can this be possible? The answer is “cancellation of
errors”, the heat capacity is proportional to the number of electrons that are able to absorb
thermal energy, while conduction properties is proportional to that and to the average kinetic
energy of those electrons. The classical theory (as we will see later) overestimate the number of
electrons that can actually absorb thermal energy in about 2 orders of magnitude (thus the heat
capacity is overestimated) but it underestimates, by about the same factor, the energy of those
electrons and thus, by coincidence, conductive properties are very well predicted.

Electrons in a one-dimensional infinite box


The closest problem to the one in hand is that of electrons in a box, a typical problem in quantum
mechanics.

Any particle is characterized by a kinetic energy and a potential energy. Classically that means
that the total energy of the system is given by:
p2
H   V Where H is known as “the Hamiltonian” or energy operator, p is the particle’s
2m
momentum, m is its mass and V is the potential acting on the particle.

In quantum mechanics, p can be represented by the operator p= –iħ d/dx where the action of this
“operator” p is to take the derivative of the function and then multiply that by the constant –iħ.
p2 means applying this operator twice, thus

5
 2  2 n
H n    V n   n n
2m x 2

The above is known as the Schrödinger equation and all that there is to know about the electron
can be obtained from the equation above (except that this equation does not have an exact
solution for most potential functions)

For the electron in a box case, the potential energy is defined as V=0 inside, V=infinite outside
What lead to the equation:

 2  2
  E
2m x 2
With boundary conditions that force the wave function to be zero at z=0 and z=L (where L is the
box width)

 The condition for that the wave function needs to be zero for z=0 is satisfied proposing a
solution of the form:
  n  A sinkx
 The condition that the wave function must be zero at L is satisfied if k=nπ/L or,
 nx 
 n  A sin  n  1, 2, 3,... => k=2/n
 L 
Thus the solutions are quantized and the allowed energy values are given by

 2 k 2  2  n 
2

n    
2m 2m  L 

 Normalization of the wave function

L L L  nz  2 L
  *dz    dz  1  An2  sin 2  dz  An
2
0 0 0
 L  2
2
An 
L

Energy difference between two consecutive levels increases as n increases

6
E  h 
h2
2

ni2  n 2f 
8mL
Transition between levels leads to absorption or emission of a photon with energy equal to the
difference between levels.

The above solutions correspond to a single electron in an infinite well. If we consider now that
the metal have N electrons (assuming 1 per atom), and invoke the Pauli exclusion principle, then
we can only accommodate 2 electrons per energy level and thus only N/2 levels will be filled
with electrons. The energy level with the maximum energy that holds electrons is known as the
Fermi energy and it is given by

2  n f  
2
 2  N 
2

f      
2m  L  2m  2 L 
Class activity 5.
Effect of Temperature
The picture described above, where energy levels are being filled one by one (with two electrons
each) until we run out of electrons, is only valid at zero Kelvin. As the temperature increases,
some electrons with energies of the order of kBT around the Fermi energy can be promoted to
higher energy levels and thus some levels above the Fermi level are occupied while some levels
below the Fermi level are empty. The Fermi-Dirac distribution accounts for that effect:
f   
1
    
exp   1
 k BT 
where  is the energy and  is the chemical potential (we will define it in a little bit). Notice that
as T tends to zero, the equation above tends to two different values, 0 or 1, depending on the
relative values of  and , for < the exponential goes to zero while for > the exponential
goes to infinity. As the temperature increase, this function decreases smoothly as  increases.

The Fermi-Dirac distribution is a probability distribution; it indicates the probability of a given


energy level to be occupied. At 0K, the probability of occupation of energy levels below the
chemical potential is 1 (all are occupied) while the probability of occupation of energy levels
above the chemical potential is zero (all empty). The chemical potential is an empirical
parameter defined as the energy where the probability of occupation is exactly ½. For very low
temperature kBT<<, f

We can already see here the reasons for the discrepancy between the experiment and the Drudes
model. Due to the Pauli Exclusion Principle, not all the electrons can increase their energy due
to temperature, only those near the Fermi level, this account for the 2 orders of magnitude in the
heat capacity. By virtue of the Pauli Exclusion Principle also, those electrons that can absorb
energy have a much higher energy than expected by the classical theory since they are forced to
be in higher energy levels, both effects, small number of electrons with higher energy
approximately cancel each other given conduction properties in reasonable agreement with the
experiment

7
October 28th, 2010
Class 15

Electrons in a three-dimensional infinite box


For the case of electrons in a box, the finite size box may be not the best way to go considering
the macroscopic dimensions of a metal that can be regarded, without much approximation, as
infinite in size. For a case like that, periodic boundary conditions may be more appropriate.
Thus for the three dimensional cases, periodic boundary conditions will be used.

By following the same mathematical procedure than the one we followed to determine the
number of vibrational modes in a crystal, we found that the solutions for an electron in a three
dimensional box with periodic boundary conditions are:

2 2
     n
2m
 2   2  2  2 
       n
2m  x 2 y 2 z 2 
Since  k r   expik  r 
d d 2
 ik x expik x  x and  k x expik x  x  k x 
2 2
2
dx dx
d d 2
 
 ik y exp ik y   y  and  
 k y exp ik y   y   k y 
2 2

dy dy2
d d 2
 ik z expik z  z  and  k z expik z  z   k z 
2 2
2
dz dz
       2 
2 2 2
Thus,        n becomes
2m  x 2 y 2 z 2 


2
2m
 
 k x2  k y2  k z2    n
or
 
2

n  k x2  k y2  k z2
2m

n 
2 2
2m
k  

where k 2  k x2  k y2  k z2 
Our relationship between wavevector and wavelength is such that,
2nx 2n y 2nz
kx  , ky  , kz  .
Lx Ly Lz
The quantization comes from the boundary conditions  n 0,0,0   n Lx , Ly , Lz 

Using the quantum mechanical definition of p


 
p k  i k  k  k

1
So the linear momentum of an electron with a wavevector k is ħk and the speed of such an
electron is v= ħk/m

The energy at the Fermi level (now in 3D called Fermi surface), is given by:
F 
2
2m
 
kF
2

where kF is the wavevector at the Fermi surface.

In the same way we did with the vibrational modes, we can count the number of electronic states
and since the solutions are the same, we can see that in the unit of volume there are
3
 L  4 3
  electronic states in the k space. Thus in a sphere of volume k F (volume of the Fermi
 2  3
3
 L  4 V 3
sphere) there are 2     3 k f  2 kF states (counting both spin states). If in the same
3

 2  3
volume there are N atoms and each contributes with 1 electron, then that number above must be
equal to N from where
1 2 1
 3 2 N  3
 2  3 2 N  3
k    3 N 
2 3
k F    F    and  F  F    
 V  2m  V  m  m  V 

N/V is the electron density (equal to the atomic density assuming each atom contributes with 1
electron).

From the above equation it is useful to define the Fermi temperature TF as the temperature for
which kBTF=F. This temperature has nothing to do with the thermodynamical temperature,
TF~104 is the temperature Tg an electron gas need to be for its thermal energy be kBTg=F.

Density of States

2
 2  3 2 N  3
The equation     , is valid for any value of energy not only for the Fermi energy,
2m  V 
in that case Nε represents the number of electrons up to that energy. So inverting the equation we
obtain
3
V  2m  2
N  2  2 
3   

where N is the number of one electron states (with or without electrons) considering each energy
level as 2 states (one for spin up and the other for spin down) with energy between 0 and  for
any  larger or smaller than the Fermi energy.

so the density of states is

2
3  V  2m  2  3N
3 3
V  2m  2
D  
dN 1
    2
   
d 2 2   2  2  3 2   2   2
 

Heat Capacity
As said before, only a few electrons near the Fermi level contribute to the heat capacity,
electrons with energy much smaller that the Fermi level cannot take advantage of the thermal
energy because energy levels above them are occupied and the Pauli principle does not allow
them to also occupy those energy levels.

Classically, the energy of an electron gas is 3/2kBT per electron, thus each electron contributes to
the heat capacity by a quantity 3/2kB (½ per degree of freedom). In the Drudes model (no Pauli),
all the N electrons contribute to the heat capacity and thus the electronic contribution to the heat
capacity was predicted to be 3/2NkB, or the specific heat capacity 3/2 N/V kB where N/V is the
electron concentration.

Considering the Pauli principle, electrons with energy within kBT of the Fermi energy can absorb
energy, thus only a fraction T/TF can actually contribute to the heat capacity and then
qualitatively the total thermal kinetic energy of electrons is

Uel≈(NT/TF)kBT

leading to a heat capacity of

Cel≈NkB(T/TF)

which is not the exact value but retains the T dependence found in the experiment. At room
temperature, T/TF~0.01 for TF~3x104K which is in the order of a typical value for the Fermi
temperature

For a more quantitative calculation consider the change in energy of the metal when the
temperature increases to a certain temperature T is given by

 F
U    f  , T D d   D d
0 0

Multiplying the identity

 F
N   f  , T D d   D d
0 0

by F

 F F 

  F f  , T D d    F D d


0 0
what leads to   F D d    F f  , T D d  0
0 0

3
Adding this zero to the equation above and reorganizing we get
 F
U      F  f  , T D d    F   D d 
0 0
F  F
     F  f  , T D d      F  f  , T D d    F   D d
0 F 0

or
 F
U       f  , T D d   
F F   1  f  , T D d
F 0

The first term is the product of the energy gained by one electron when it goes from the Fermi
level to the energy level , times the number of states at the energy  times the probability that
the state has been occupied integrated in all the level above the Fermi level. This is equal to the
total amount of energy invested to populate energy levels about the Fermi level. The second
term is the product of the energy gained by one electron when it goes from the energy level 
below the Fermi level to the Fermi level times the number of states at the energy  times the
probability that the state has indeed been vacated integrated in all the level below the Fermi level,
that is the total amount of energy invested to de-populate energy levels below the Fermi level.

So now the heat capacity can be obtained simply by taken the derivative of the last equation with
respect to T. Noticing that only f() depends on T

df  , T  df  , T  df  , T 
 F 
     F  D d    F    D d      F  D d
dU
Cel 
dT  F dT 0
dT 0
dT
Since we are mostly interested in temperatures around room temperature, then T/TF= kBT/F~0.01
where the density of states can be approximated by the density of states at the Fermi level and
the chemical potential can be considered independent of temperature and ~F

Then

df  , T 

Cel  D F     F  d
0
dT

f  , T  
1
    F 
exp   1
 k BT 
    F 
exp  
df  , T     F   k BT 

dT k BT 2       F  2
 exp    1

  k B T  

4
    F 
exp  
    F 
 2 
 k BT  ex
Cel  k B D F    d  k 2
TD   x 2
dx
0 
k BT 
 exp     F   1
   
2 B F

 F

e x
 
1
2

  k BT
  k BT  
x=(-F)/ kBT

In order to integrate, we can replace the lower limit with -, which is approximately true for
kB<<F.

Thus

ex 2
Cel  kB2TD F   x 2 d  kB2TD F    2 D F kB2T
1
 e x
1  2
3 3

Using that

D F  
3N 3N

2 F 2k B TF

1 3N 1 T
Cel   2 k B2 T   2 Nk B
3 2k B TF 2 TF

Experimentally Determined Heat Capacity


For very low temperatures, the total heat capacity can be considered the sum of the electron and
phonon contributions, thus

C=T+AT3

Where the electronic contribution is proportional to T and the phonon contribution is


proportional to T3. Dividing by T, we obtain a linear relationship between C/T and T2 with
intercept  and slope A. (see figure 9 chapter 6, in page 145)
The observed value of the intercept does not agree very well with the predicted value and
the reason is that the electron seems not to have the classical mass when in the solid but an
effective mass. Usually an effective mass is defined as

mth  measured 

m  calculated 

The reason for which the apparent mass is not the free electron mass involves three separate
effects
1) The interaction of electrons with positive ions. The resulting mass is called the band
effective mass

5
2) The interaction between electrons and phonons what distorts the lattice that in turn
interact with other electrons increasing their effective mass
3) Electron-electron interaction what also increases the effective mass

Electrical Conductivity
The momentum of a charge is given by mv=ħk, thus its equation of motion in an electric field
can be written as:
dv dk
F m   qE (where q=-e for electrons)
dt dt

qEt
What leads to k  the same for ALL electrons, thus the entire Fermi sphere is moving!

What is consistent with a free charge in an electric field.

In a real system, electrons are not really free, they interact with the ions that are hopping in place
(phonon) and they interact with imperfections in the crystal. A more appropriate model consists
in assuming that these electrons are free to accelerate just during a time t= after which they just
stop and start over. The overall effect is that these electrons find a steady state at a higher speed
than before the electric field is applied (sort of like a terminal velocity). So the shift in the
position of the Fermi sphere is qE/ħ while the speed increases an amount v=ħk/m= qE/m.
At equilibrium, the average speed of the electron gas is zero, thus v is the drift velocity.

Consider n electrons per unit of volume moving at a speed v, the number of electrons that will
cross a surface of unit area perpendicular to the direction of motion in a time to is equal to the
number of electrons in a cylinder of unit cross section area and length l=vto, thus, in the unit of
time, the total charge crossing a unit are is:

j=nqv=nq2E/m

j is the current density and the above is known as the Ohm’s law where

= nq2/m

is the conductivity and the inverse

= m/nq2 is the resistivity.

Another important property of charge carriers is their mobility which is defined as μq=v/E where
E is the electric field. Carriers with high mobility will pick up speed at small values of electric
field thus leading to larger currents. With this definition
j=nqv=nq μqE and
= nq μq or μq=q/m
The two main sources for resistivity are the electron-phonon interaction (interaction with the ions
that are hopping in their place) and the interaction with impurities and crystal defects. The
former is more relevant at high temperature while the later is relevant at all temperature. If the

6
electric field is then removed, collisions will make the electrons go back to their unbiased speed.
The rate of these collisions is independent from each other, thus
f=fL+fi

where f is the total number of collisions per unit of time fL is the number of collisions with the
lattice (charge-phonon collisions) per unit of time, and fi is the number of collisions with
impurities per unit of times

thus the characteristic time can be obtained as:


1 1 1
 
 L i

The equation above then indicates that the total number of collisions per unit of time is the sum
of the total number of collisions with the lattice (phonons) per unit of time plus the total number
of collisions with impurities per unit of time) or multiplying by m/ne2 =L+i

Hall Effect
If a magnetic field is applied perpendicular to a current, the electrons will suffer a force
perpendicular to both, the drift velocity and B, of a magnitude
F=qvxB

This force produces an accumulation of charge and an electric field that generates an
electrical force that opposes to the magnetic force. In equilibrium

-eEH=evB or EH=-vB

By defining the Hall coefficient RH as

EH=-RH JB

vB= RH JB or RH=v/J
(notice that RH can be measured by measuring the electric field perpendicular to the current and
it is a property of the material).

Using J=nqv RH=1/nq (where q is negative for electrons)

For Li, monovalent, the measured Hall coefficient at room temperature is -1.7x10-10 m3/C and
the calculated value -1.4x10-10 m3/C. For Al (trivalent) both values are ~ -0.3x10-10 m3/C. But
for Zinc, the calculated value is -0.5x10-10 m3/C
while the measured one is +0.3x10-10 m3/C. The
reason is that the carriers are positive. Although
electrons are always the ones moving, some
material, especially divalent, behave as if the
carriers were positive. We’ll talk about that later.

You might also like