Anda di halaman 1dari 205

Classical Mechanics Review

(Louisiana State University Qualifier Exam)


Jeff Kissel
May 18, 2007
A particle is dropped into a hole drilled straight through the center
of the earth. Neglecting rotational effects, show that the particles
motion is simple harmonic. Compute the period and give an estimate
in minutes. Compare your result with the period of a satellite orbiting
near the surface of the earth.

Figure 1: A particle dropped from R .


Only the mass inside the shell of radius r contributes to the gravitational
force. If we assume the Earth have uniform density, then the mass contributing
to Fg is
M
4
3

r3
M

=
=

M
4
3

3
R

M 3
3 r
R

Thus, the force experienced by the particle m from mass M is


GmM
Fg =
r2 

Gm M 3

=
3 r
R
r2


GM m
r
Fg =
3
R
= kr
1

(1)

(2)

Jeff Kissel May 18, 2007

Classical Mechanics

where the last step absorbs all the constants G, M , R , and m into the constant
k, which shows that Eq. 2 is linearly proportional to the radius. Similarly,
Hookes force law for simple harmonic motion states that the force of a simple
harmonic oscillator is proportional to the displacement from the equilibrium position. Since the center of the earth would be the particles equilibrium position,
r is the displacement, and thus by direct comparison, the particle obeys simple
harmonic motion, with spring constant k.
To computepthe period , we use what we know for a spring with resonant
frequency = k/m,
r
2
m
1
=
= 2
=
f

k
s 

3
R
= 2 
m
m
GM
s
3
R
= 2
GM
2

3
4 2 R
GM

(3)

which is Keplers 3rd law for planetary motion (for a small particle), again
indicating that motion is periodic.
Finally, one can find an estimate of this period by using Earths gravitational

acceleration of a small particle at the surface, g = GM


9.8 m/s 2 .
R2

3
4 2 R
GM


R
= 4 2
g

4R

(4)

= 4(6.4 10 m) [s]
= 2.57 107 s

= 5069 s = 84 min

(5)

which is approximately equivalent to a satellite in low-Earth orbit. (At 185 km,


i.e. when low-Earth orbit becomes stable, the correct period is 88.19 min.)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
Two identical bodies of mass m are attached by identical springs of
spring constant k as shown in the figure.

Figure 1: A system of two masses attached by springs (A) in equilibrium, and


(B) after some oscillation.
a) Find the frequencies of free oscillation of this system.
The first goal will be to set up the systems characteristic equation,
V 2T = 0

(1)

where V is the potential energy matrix, T is the kinetic energy matrix, and
are the desired eigenfrequencies (frequencies of free oscillation) of the system.
The potential on the particles is the sum of each individual springs potentials,
V =

1 2 1 2 1 2
kd + kd + kd
2 1 2 2 2 3
1

(2)

Jeff Kissel October 11, 2006

Classical Mechanics

If we define the displacements between the masses in terms of generalized coordinates ,


(0)

d1 =
d2 =

(x1 x1 ) + (0)
(x2

d3 =

(0)
x2 )

(0) +

(0)
+ (x1
(0)
(x2 x2 )

= 1
x1 ) = 2 1
= 2

we can then write the potential as,


V

=
=
=

1
1
1
k(1 )2 + k(2 1 )2 + k(2 )2
2
2
2
1
1 2 1
k + k(22 + 12 1 2 2 1 ) + k22
2 1 2
2

1
k 222 + 212 1 2 2 1
2

(3)

which in the desired matrix form is,


V

1
Vij i j =
2

2k
k

k
2k

(4)

The kinetic energy matrix is much easier to find because it is neatly diagonal:
1
1
mv 2 + mv 2
2 1 2 1

1
m 0
T Tij i j =
0 m
2
T =

(5)

Once weve got these matrices, we want to find the eigenfrequencies, of


the charateristic equation
V 2T = 0
(6)
Notice however, that Eq. 4 and Eq. 5 are symmetric, 2 2 matrices, so we
can use some trickery and write them as multiples of the identity matrix or the
SU(2) spinor matrices:
V
T

= 2k I k
x
= mI

Now we can just diagonalize Eq. 6 instead of the usual determinant method,
since we know the eigenvalues of I are {1,1}, and for any
i are {1,-1}. Ill also
use the fact that a diagonalized
x is just
z
0
D1 (m 2 I)D
2 (D1 ID)

= V 2T

= 2k I k
x m 2 I
= D1 (2k I k
x )D
k
2k
(D1 ID) (D1
x D)
=
m
m
2

Jeff Kissel October 11, 2006

Classical Mechanics

2 I =


1 0
2
=
0 1
2

2k
k
I

z
m m 

2k
k
1 0
1

0 1
0
m
m
2k k
m

(7)
0
1

So the eigen frequencies for this two mass system are


r
r
k
3k
1 =
or
=
m
m

(8)

(9)

Well need the eigenvectors for part (b), so well find them now. A big
advantage to using this SU(2) group argument is that since Eq. 7 only involes
I and
z , we know immediately that the two (normalized) eigenvectors for the
system are the (normalized) eigenvectors of those two matrices,
 
1
1
(10)
for I u
1 =
1
2


1
1
for
z u
2 =
(11)
1
2
where u
1 corresponds to when the two masses are sloshing back and forth in
phase with each other, and u
1 is when they oscillate out of phase, i.e. if M1 is
moving to the left, M2 is moving to the right, or vice versa.
b) M1 is displaced from its position by a small distance A1 to the
right while M2 is not moved from its position. If the two masses are
released with zero velocity, what is the subsequent motion of M2 ?
With the (very safe) assumption that the system is periodic in time, we know
from Fouriers theorem that the positions of each particle as a function of time
may be written as a sum of sines and cosines, whose phase is the eigenfrequencies. So,
x1 (t) =
x2 (t) =

A sin 1 t + B cos 1 t + C sin 2 t + D cos 2 t


E sin 1 t + F cos 1 t + G sin 2 t + H cos 2 t

(12)
(13)

However, we can narrow this down at least a little because of the eigenvectors.
r


k
1
1
u
1 = 2
A = E, B = F
1 =
1
m
r


3k
1
2 =
u
2 = 12
C = G, D = H
1
m
which means Eqs. 12 and 13 (and there derivatives) simplify a lil bit to
x1 (t)

= A sin 1 t + B cos 1 t + C sin 2 t + D cos 2 t


3

Jeff Kissel October 11, 2006

x 1 (t)

x2 (t)
x 2 (t)

Classical Mechanics

= A1 cos 1 t B1 sin 1 t + C2 cos 2 t D2 sin 2 t

= A sin 1 t + B cos 1 t C sin 2 t D cos 2 t


= A1 cos 1 t B1 sin 1 t C2 cos 2 t + D2 sin 2 t

Remember the problem statement said M1 is diplaced by a small distance


A1 [at t = 0]? That means the initial conditions (where t = 0, so the sine
terms are zero, and cosine terms are one) are
x1 (0) = A1
x2 (0) = 0

=
=

B =

x 1 (0) = 0

x 2 (0) = 0

A = C

B+D
BD
1
A1
=
2

A+C 3

AC 3
= 0

So then holy moly with extra cannoli, the positions of M1 and M2 at a funcion
of time are
!
!
r
r
1
1
k
3k
x1 (t) =
A1 cos
t + A1 cos
t
(14)
2
m
2
m
!
!
r
r
1
k
k
1
A1 cos
t A1 cos
t
(15)
x2 (t) =
2
m
2
m

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
X
X
#3

Notes/Changes

Added to bank.

A planet is in circular motion about a much more massive star. The


star undergoes an explosion where three percent of its mass is ejected
far away, equally in all directions. Find the eccentricity of the new
orbit for the planet.
Initially the planet is in a circular Keplerian orbit, where we know several
things. The force and potential are
F =

GM m
r2

, V =

GM m
r

(1)

For any circular orbit, the radius is given by


F
GM
m
rc2

GM
rc
rc3
rc

= mac
v2
= 
m t
rc


For circular orbits, r
vt2 = r 2 + r2 2

=
=

0
2 2
rc

= 2 rc2
GM
2

1/3
GM
=
2

(2)

Also, from Eq 3.57 on p94 of Goldstein, any mass m orbiting a r2 central

Jeff Kissel May 18, 2007

Classical Mechanics

force will have eccentricity,


e =

1+

2E2
mk 2

(3)

with E as the total energy, = mr2 as the angular momentum, and k is this
case as GM m. Since the orbit is circular, the eccentricity is zero, so the initial
total energy is
r
2E2
1+
0 =
mk 2
2E2
0 = 1+
mk 2
2
2E
1 =
mk 2
mk 2
(4)
E = 2
2
And finally, the total energy can also be written as the sum of the kinetic
an potential energies,
E

=
=

 GM m
1  2
2 2

m 

r
+
r
c
c
2
rc

2/3
GM
1
GM m
m

1/3

2
GM

=
=
E

2/3

2/3
1 
m GM
m GM
2

 
2/3
1
1 m GM
2
2/3
1 
m GM
2

(5)

Phew! Now, just after the explosion, the star is at the same radius, rc .
However the star has a new mass, M = (1 )M where Ive denoted as
the percentage of mass loss. The only portion of the energy that is effects is
the potential, in which k = (1 )k. The angular momentum = mrc2 is
independent of M . So, the new energy is
E

=
=

GM m
1
mrc2 2
2
rc
2/3

1
G ((1 )M ) m
GM

m

1/3
2

GM

2/3

2/3
1 
m GM
(1 ) m GM
2
2

Jeff Kissel May 18, 2007

=
E

Classical Mechanics

 
2/3
1
(1 ) m GM
2

2/3
1
( + ) m GM
2

The ratio of initial energy to final energy is then,



2/3
1


+
)
m
GM
(


2

E

2/3
=

1
E

2 
mGM


1
E
= 2 +
E
2

E = (1 2) E
This in turn affects the eccentricity, which means Eq. 3 becomes,
r
2E 2

1+
e =
mk 2
s
2(1 2) E2
1+
=
2
m ((1 ) k)
s
(1 2) 22
1+
E
=
(1 )2 mk 2


mk 2
Eq. 4: E =
22
s


2
(1 2) 22 mk
1
=
(1 )2 mk 2 
2
s
1 2
e =
1
(1 )2

(6)

(7)

(8)

For = 0.03, the new eccentricity is


e

= 0.030928

(9)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A homogeneous cube each edge of which has a length , is initially
in a position of unstable equilibrium with one edge in contact with
a horizontal plane. The cube is then given a small displacement and
allowed to fall. Find the angular velocity of the cube when one face
strikes the plane if:
a) the edge cannot slip on the plane (friction).
b) sliding can occur (no friction).

Figure 1: A cube falling with friction holding one corner stationary, and the
same cube sliding on the corner without friction.
For both scenarios we can use a conservation of energy arguement to solve
for the angular velocity.
a) The initial kinetic energy, Ti is zero, because the the cube is in (unstable)
equilibrium. The potential we can say is simply that of a point mass at the
cubes center of mass, a height h above the ground,
0
Ei = T
i + Vi = mgh
1

Jeff Kissel October 11, 2006

Classical Mechanics

1
Vi = mg
2

(1)

In this case, the edge that is stationary acts as a pivot around which the
cube will rotate. Thus, the kinetic energy of the cube while in motion is only
rotational. We can use the parallel axis theorem to find the moment of inertia rotating about an edge of the cube, noting that the moment of inertia for
spinning along the center of mass, ICoM = 16 m2 , and the displacement, d is
h = 12 ;
I||

=
=
=

1
m2 + mh2
ICoM + md2 =
6
2

1

m + m
6
2


1 1
2
2
m
=
+
m2
6 2
3

(2)

The final kinetic energy as the cube hits the plane is then
=

Tf

=
=

1
I|| 2
2 f 
1 2
m2 f2
2 3
1
m2 f2
3

The final potential is as a point mass at a height,


Vf =

(3)
1
2

1
mg
2

(4)

So conservation of energy dictates that


Vi
1
mg
2
1
g
2
1
2
3 f
f

=
=
=
=
=

Tf + Vf
1
1
m2 f2 + mg
3
2
1
1
2 + g
3 f 2


1
1
g
2 2
r 

3g
21
2

(5)

Jeff Kissel October 11, 2006

Classical Mechanics

b) In this case, the kinetic energy will be slightly different. Now, because the
plane is frictionless, the cube rotates about its center of mass, which is moving
(only) in the direction of gravity. The position and velocity of the center of
mass become
~rCoM

=
=

~vCoM

=
=
=

yCoM


h sin +
4
y CoM



h cos +
4



cos +
4
2

taking to be .
Though the initial kinetic energy is still zero, the final kinetic energy changes
to
Tf

=
=
=
=

1
1
m~vf2 + ICoM f2
2
2
!



2
0
1 1
f
1
2
+
m
m2 f2
cos 
f +
2
4
2 6
2
 
1
1
1
+
m2 f2
m2 f2
4
2
12
5
m2 f2
24

(6)

Notice that the initial and final potential energies will remain the same,
because the cube starts in the same position as in case a), and ends in the same
orientation, just displaced by 21 .
So, using Eq. 1 and 4, the conservation of energy equation is
Vi =
1
mg =
2
5
2 =
24 f
f2

Tf + Vf
5
1
m2 f2 + mg
24
2


1
1
g
2 2
!

21
24g
5
2
r

12g 
21
5

(7)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A chain of linear density ( [g/cm]) is hanging vertically above a
table. Its lowest point is at a height h above the table. The chain is
released and allowed to fall. Calculate the force exerted on the table
by the chain when a length x of chain has fallen onto the surface.

Figure 1: A chain suspended a height h above the table at t0 , and then x amount
fallen onto the table after time t
The force exerted on the table will be equivalent to the sum of the gravitational (from whatever part of the chain has already landed on the table) and
impulse (from the part of the chain that has just hit the table) forces. The
gravitational force from the amount of mass m = x already on the table is
simply
Fg = mg = xg
(1)
1

Jeff Kissel October 11, 2006

Classical Mechanics

During a time interval dt, the mass of the rope equal to (v dt) is hitting the
table. The change in momentum imparted onto the table is then
dp

dm v

[(v dt)] v

v 2 dt

(2)

The impulse force is then


Fimpulse =

dp
= v 2
dt

(3)

However we would like to know how velocity v is related to x(t). Some one
dimensional kinematics should shed light on the matter:
v 2 vi2

2ax

v2

2g(x + h)

(4)

So Eq. 3 becomes
Fimpulse = v 2 = 2g(x + h)

(5)

Finally, the total force of the table from the falling chain is
F

= Fg + Fimpulse
= xg + 2(x + h)g
= 3xg + 2hg
= g(3x + 2h)

(6)

which is equivalent to the weight of the length 3x of the rope, plus the correction
for the initial height.

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A thin circular ring of radius R and mass m is constrained to rotate
about a horizontal axis passing through two points on the circumference. The perpendicular distance from the axis to the center of the
ring is h.

Figure 1: A thin hoop confined to swing from an off-diameter axis of rotation.


a) Find a Lagrangian for this object.
Looking at the hoop from the side few, its apparent that this system is akin
to a simple rigid pendulum of length h. Since we know the pendulums bob (i.e.
the hoops center of mass) is confined the plane of the page (or in and out of
the page for the front view) a convenient generalized coordinate for this system
is the angle as seen in Figure 1. The kinetic energy of the hoop is then
T

=
=
=

1 2
I
2

1
mR2 + mh2 2
2

1
m R2 + h2 2
2
1

(1)

Jeff Kissel October 11, 2006

Classical Mechanics

where the moment of inertia I was found using the parallel axis thereom,
I||

= ICoM + md2

(2)

noting that d is the distance from the center of mass to the parallel axis, and
ICoM = 21 mr2 for a thin hoop.
The potential energy will be similar a simple pendulum:
V

m~g ~r

mgh cos

(3)

So using Eq. 1 and 3 the Lagrangian for a thin hoop hung from an off diameter
axis of rotation is then
L

T V

1
m R2 + h2 2 + mgh cos
2

(4)

b) Find the period of small oscillations about this axis.


Retrieving the equation of motion from the usual Lagrangian formalism will
yield the frequency of small oscillations , from which we can find the period.


d L
L
0

dt

 
d 
m R2 + h2 + mgh sin = 0
dt

m R2 + h2 =



mgh sin
gh
2
sin
R + h2

(5)

From here, we make the approximation that sin so that Eq. 5 becomes
a simplified 2nd order differential equation of the form x = 2 x, whose general
in which is the frequency. Hence,
solution is x(t) = x0 cos (t)

(t)

gh

R 2 + h2
r

0 cos

gh
t
2
R + h2

(6)

The period of small oscillations is the time it takes Eq. 6 to repeat the initial
2) = cos (0). From this we arrive at
conditions at t = 0, such that cos (
the familiar expression for the period of a pendulum as expected:

= 2

(7)


gh
R 2 + h2

 12

(8)

Jeff Kissel October 11, 2006

Classical Mechanics

Hoop! There it is!


c) For what value of h is the period minimum?
We wish to minimize with respect to h, i.e. set the derivative of Eq. 8
equal to zero,

h
0

 (gh) 2 (R2 + h2 ) 2
= 
2
1

1
1
3
1
1
1
(gh) 2 (g) (R2 + h2 ) 2 + (gh) 2 (R2 + h2 ) 2 (2h)
2
2
1
g (R2 + h2 ) 2
2h
+
3
1
1
2 (gh) 2
2 (gh) 2 (R2 + h2 ) 2

=
=

g (R2 + h2 ) 2 (R2 + h2 ) 2 + 2h (gh)


((
3
(2(+(h2 ) 12
2 (gh)
((2((R
(
= g(R2 + h2 ) + 2gh2
= gR2 gh2 + 2gh2
=

R2
h

= g (h2 R2 )
= h2
= R

(9)

Only the non-trivial, positive value for h is physical, so the shortest period
would result if the axis of rotation which has a distance
h=R

(10)

from the center of mass of the hoop.


Sanity check: if one wants to minimize the hoops period, Eq. 7 says that
one could maximize the frequency. The axis at which the hoop wont oscillate
at all is zero is through center, and an axis with the maximum frequency is that
which has the longest pendulum arm, i.e. at the edge of the hoop, or h = R.

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
An object is dropped from a tower of height h. The tower is located
at the equator of the Earth. The rotational speed of the earth is .

~ =
Figure 1: Da Erf. Note that +
ex points North, and +
ey points West, so
ex .
(a) If the Earth is treated as perfectly round and uniform and the
acceleration due to gravity on a non-rotating earth is g, what is the
acceleration of gravity at the tower?
With Earth rotating, we must not only consider the acceleration due to
gravity, but the centripetal and Coriolis acceleration. At the top of the tower,
~agrav = g ez , the radial vector is ~r = +h ez , and the velocity is ~v = vr ez .
The effective acceleration (in the rotating frame) felt at the tower is then given
as
~aeff

= ~agrav + ~acentrif ugal + ~aCoriolis


~
~ ~r 2 ~v
= ~agrav
=

(g
ez ) (
ex ) (
ex ) (h
ez ) 2(
ex ) (vr ez )

g ez h2 (
ex ex ez ) 2vr (
ex
ez )
1

Jeff Kissel May 18, 2007

~aeff

Classical Mechanics

g ez h2 (
ex
ey ) 2vr (
ey )

=
=

g ez h2 (
ez ) + 2vr ey
2vr ey (g h2 ) ez

(1)

Yet, at the top of the tower, we assume the object to have no radial velocity,
vr = 0. Thus, the effective acceleration is
~geff

(g h2 ) ez

(2)

(b) Even though it is released from rest, this object will not land
directly below the point from which it was dropped. Calculate the
amount and direction (N, E, S, W, or elsewhere) of the horizontal
deflection of the object. You may assume the deflection is small.
The Coriolis effect is the only term that will have an effect in the horizontal
(
ey , East or West) direction, so pulling the first term from Eq. 1 if vr is now
~geff t,
FCoriolis

m aCoriolis

2mvr ey

=
m y =

2m((g h2 ) t) ey
2m(g h2 ) t

2(g h2 ) t


1 2
2
y = 2(g h )
t
2


1 3
2
y = 2(g h )
t
6

y = (g h2 ) t3
(3)
3
This quantity y is the horizontal deflect, get we can be more explicit. The
time spent falling t can be found from the free-fall kinematics of a particle
released from rest:
1
1
(g h2 ) t2
y = 
v0t geff t2 h =
2
2
1/2

2h
2h
2
t =
(4)
t =
(g h2 )
(g h2 )
y =

which means the deflection becomes




(g h2 )
3

(2h)3/2
3(g h2 )1/2

2h
(g h2 )

3/2

The direction of this displacement is in the


ey direction, or to the East.
2

(5)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
Consider a pendulum consisting of a uniform disk of radius, R and
mass, m suspended from a massless rod that allows it to swing in the
plane of the disk.

Figure 1: A circular disk, rotating in its plane on an axis a distance, d away


from the center of mass.
(a) Using the parallel axis theorem, or calculating it directly, find
the moment of inertia I for the pendulum about an axis a distance
d(0 < d < R) from the center of the disk
The moment of inertia through the center of a disk of radius R and mass m
is

1
mR2
(1)
2
Using the parallel axis theorem, which states that the moment of inertia around
a point parallel to the center of mass a distance d away is
ICoM =

Id = ICoM + md2

(2)

So, using Eq. 1 the moment of inertia displaced by |~r| = d is


Id

Id

1
mR2 + md2
2
1
m (R2 + 2d2 )
2
1

(3)

Jeff Kissel October 11, 2006

Classical Mechanics

(b) Find the gravitational torque on the pendulum when displaced


by and angle .
From the definition of rotational torque from a given force:
= ~r F~g
= |~r| |F~ | sin
= (d)(mg) sin

= mgd sin

(4)

(c) Find the equation of motion for small oscillations and give the frequency . Further, find the value of d corresponding to the maximum
frequency for fixed R and m.
If we note that = I, then we know the equation of motion immediately
from Eq. 4(using the small angle approximation that sin = ),
Id

= mgd
mgd
=

Id
(5)

which is a 2nd order differential equation of the form = 2 from which we


can pull out . Subbing in Eq. 3,
2

2
mgd
m (R2 + 2d2 )

2

2gd
=
R2 + 2d2
=

(6)

To find the value of d for which is maximized, we shall differentiate with


respect to d and set to zero,
d
dd

d
dd


(2gd)(R2 + 2d2 )1

 21

2gd
1
2g (R2 + 2d2 )1 (2gd)(R2 + 2d2 )2 (4d)
2 R2 + 2d2

1 

8gd2
1 R2 + 2d2 2
2g
2
2
2gd
R2 + 2d2
(R + 2d2 )2
!
1
 2

1 R + 2d2 2
2g(R2 + 2d2 ) 8gd


2 2

2gd
2
(R
+ 2d )2



1
R2 + 2d2 2 2gR2 + 4gd2 8gd2
1
R2 + 2d2 2 (2g(R2 2d2 ))
1

R2 + 2d2 2 R2 2d2
2

Jeff Kissel October 11, 2006

Classical Mechanics

R2 + 2d2 = 0
iR
d=
2

R2 2d2 = 0
R
d =
2

But we know the left set is unphysical and R2 is equivalent to


R
dmax =
2

(7)
R

,
2

so
(8)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A heavy particle of mass m is placed close to the top of a frictionless
vertical hoop with radius R and allowed to slide down the hoop. Find
the angle at which the particle falls off.
We know the point at which the particle falls off the hoop is when the normal
force N acting on the particle goes to zero. So, we need an expression for N .

Figure 1: Ball of mass m on a hoop with radius R.


Before the particle begins to move, it has only potential energy Vi . At the
instant it leaves the hoop, it has both potential Vf and kinetic Tf energies. The
particle is allowed to slide, so there assumed to be no friction and therefore no
rolling, thus we need no worry about the moment of inertial of the ball. The
energy conservation equation is then
Ei
Vi

=
=

mgR =

Ef
Tf + Vf
1
mv 2 + mgR cos
2

(1)

However, Eq. 1 does not contain the normal force N , which we need. It can be
found by assuming the hoop is circular, so that the normal force acting on the

Jeff Kissel October 11, 2006

Classical Mechanics

particle is just its centripetal acceleration subtracted from its weight, or


=

mg cos mac
v2
mg cos m
R

(2)

Multiplying this by 21 R will yield a more useful equation for the kinetic energy,
1
R (N
2
1
RN
2
1
mv 2
2

=
=
=

mg cos mac )
1
mgR cos
2
1
mgR cos
2

1
mv 2
2
1
RN
2

(3)

which we can now plug into the conservation equation (Eq. 1) to arrive at
explicit equation for N ,
mgR =
1
RN
2
N

=
=

1
1
mgR cos RN + mgR cos
2
2
3
mgR cos mgR
2
3mg cos 2mg

(4)

Since weve established that the particle falls off the hoop when N = 0, Eq. 4
becomes,
0 =
=
cos

mg(3 cos 2)
3 cos 2
2
3

(5)

So the angle at which a heavy particle of mass M slides off a circular hoop or
radius R is
48.2
(6)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A solid sphere of radius R and mass M rolls without slipping down a
rough inclined plane of angle .

Figure 1: A sphere rolling down a rough ramp.


Take the coefficient of static friction to be and calculate the linear acceleration of the sphere down the plane, assuming that it rolls
without slipping.
In this problem, we can use conservation of energy between the top and
bottom of the ramp. At the top of the ramp, the ball begins rolling from rest
and thus has no initial kinetic energy. This means the initial energy is
Ei

0
= T
i + Vi
= mCoM g x sin
1

(1)

Jeff Kissel October 11, 2006

Classical Mechanics

where the origin is set at the top of the ramp.


The final energy has only kinetic terms, and is a sum of the motion of the
center of mass, and the rotation of the sphere,
=

Ef

0
Tf + V
f
1
1
mCoM v 2 + ICoM 2
2
2

(2)

Setting Eqs. 1 and 2 equal to each other, plugging in for the moment of
x
and = R
,
inertia of a sphere, ICoM = 52 mCoM R2 and noting that v = x,
Ei

 g x sin
mCoM


x 2

dx
dt

Ef
1
5
 x 2 +
mCoM
10 
2
10
g x sin
7

  2
2
x
2



R
m
CoM 
5
R

(3)

x 2

q
where Ive defined a constant = 10
7 g sin . At this point we integrate both
sides to find x as a function of time, which we can then differentiate twice with
respect to time to get the acceleration, x
= aCoM ,
1

x 2 dx

1
2x 2

1 2 2
t
x =
4

= dt
= t

x = 21 2 t
q
2
10
x
= 12
7 g sin
aCoM =

7
5

g sin

x =

1 2

(4)

Calculate the maximum value of for which the sphere will not slip.
The sphere will begin to slip once the x-component of the weight becomes
greater than the maximum frictional force providable by . The angle at which
(
x)
this occurs, c can be found by setting the normal force, Fg equal to the force
of friction Ff = N ,
mg sin c
 sin c
mg


= N

 cos c
= 
mg
sin c
= tan c
=
cos c
c = tan1 ()

(5)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A rocket is filled with fuel and is initially at rest. It starts moving by
burning fuel and expelling gases with the velocity u, constant relative
to the rocket. Determine the speed of the rocket at the moment when
its kinetic energy is largest.

Figure 1: A free rocket traveling in an inertial frame, without the force of gravity.
At some time t the rocket, traveling at a velocity v, has some mass m. An
infinitesimal time dt later, the rocket has lost a mass dm, and gained a speed
dv. At this time, an infinitessimal amount of fuel, dm is ejected at a velocity
u with respect to the rocket. In an inertial frame however, that fuels velocity
is v u. Thus, using conservation of momentum, we can get an expression for
the mass of the rocket as a function of time,
procket

= procket + pf uel
1

Jeff Kissel October 11, 2006

mv
mv
m dv

Classical Mechanics

= (m dm)(v + dv) + dm (v u)
= m v + m dv v dm du dv + v dm u dm
= u dm du dv

(1)

From here, we note that to first order, the term du dv is extremely small, so it
is dropped, leaving
1
dv
u

1
dm
m

(2)

Noting that dm = dm and integrating both sides give us an express for


the mass of the rocket as a function of time:
Z m
Z v
1
1

dv =
dm
m m
0 u
 0 
v
m

= ln
u
m0
m

= m0 e u

(3)

Plugging into the expression of the rockets kinetic energy,


T

1
2

mv 2

v
1
m0 v 2 e u
2

which we then maximize with respect to v to find the largest speed vmax ,


T
1
v
2 u
=
m0 v e
v
v 2
1 v
v
1
e u
0 = m0 v e u m0 v 2
2
u
v
v
v




m0
v
e u
m0
v
e u =


2u
vmax = 2u

(4)

(5)

Thus, the kinetic energy is maximum when the velocity of the rocket is twice
that of the ejected fuel.

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 19, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
#12
#12
#12

Notes/Changes

Added damping.

A small lead ball of mass m is attached to one end of a vertical spring


with the spring constant k. The other end of the spring oscillates up
and down with amplitude A and frequency . Determine the motion
of the ball after a long period of time. You may assume that the
ball is subject to a small amount of damping, with the damping force
being given by F = v.

Figure 1: A driven harmonic oscillator sitting in a viscous damping fluid.


Hopefully, one can assume that all motion is in the y direction.

Jeff Kissel May 19, 2007

Classical Mechanics

The kinetic energy of the system is


1
my 2
2

(1)

and the potential, as long as we define the origin to be at the equilibrium point
such that the displacement y y ye = y, is as follows,
=

1
ky 2 + mgy
2

(2)

This makes the Lagrangian


L

T V
1
1
my 2 ky 2 mgy
2
2

L =

(3)

We can use the form of the Euler-Lagrange equations which has forces not
derivable from a potential, Q, to determine the equation of motion. The forces
not derivable from potential are the driving and damping forces,
Fo

= A cos (t)

(4)

Fd

= y

(5)

So, the equation of motion




d L
L

=
dt y
y
d
(my)
(ky mg) =
dt
m
y + ky + mg =
m
y + y + ky

k
y +
y +
y
m
m

y + y + 02 y

x + x +

02



g
=
x+ 2
0

x
+ x + 02 x + g =
x
+ x + 02 x =

is
Qi
A cos (t) y
A cos (t) y
A cos (t) mg
A
cos (t) g
m
k
A

, 02 =
, and B =
Let =
m
m
m
B cos (t) g
Let x = g2 + y y = x +
0
x = y
x
= y

g
02

B cos (t) + g
B cos (t) + g
B cos (t)

(6)

Yes folks, its the driven harmonic oscillator with damping. Your favorite
differential equation to solve.
2

Jeff Kissel May 19, 2007

Classical Mechanics

Well start with the homogeneous solution.


0

x
H + x H + 02 xH
The characteristic equation:
0 = am2 + bm + c

0 = (1)m2 + ()m + (02 )


has roots
m = p iq

r
b2
b
c
with p = and q =
2
4
r

2
02
p = and q =
2
4
Such that the general solution is
ept (C cos (q t) + D sin (q t))

xH (t)

21 t

C cos

02

t
4

+ D cos

02

t
4

!!

(7)

For the inhomogeneous solution, we can generalize the equation making it


complex,

x
eI + x
e I + 02 x
eI

= Beit

and then take the real part at the end. This way we can assume a complex
e ei(t) , which knocks out the differentiation because
exponential solution, x
eI = G
e eit
x
eI = G

e eit
x
e I = i G

eI = 2 G
e eit
x

(8)

e
which means all we have to do is find the complex constant G,
Beit
B
B
e
G
e
G

e eit
e eit + i G
e eit + 2 G
= 2 G
0
e + i G
e + 02 G
e
= 2 G

e (02 2 ) i
= G
B
=
2
(0 2 ) i
 2

B
(0 2 ) + i
=
(02 2 ) i (02 2 ) + i
B(02 2 ) + iB
=
((02 2 ) + 2 2 )
3

(9)

Jeff Kissel May 19, 2007

Classical Mechanics

So the inhomogeneous solution is


x
eI (t)

xI (t)

xI (t)

B(02 2 )
B
eit +
ieit
2
2
2
2
2
((0 ) + )
((0 2 ) + 2 2 )


B(02 2 )
B
it
it
= e
e
+
ie
((02 2 ) + 2 2 )
((02 2 ) + 2 2 )
 it 


e e
 = cos (t)
e ieit = sin (t)
=

B
B(02 2 )
cos (t) +
sin (t) (10)
2
2
2
2
2
((0 ) + )
((0 2 ) + 2 2 )

Yeah. The complete solution for x(t) is


!
!!
r
r
2
2
21 t
2
2
0
0
t + D cos
t
x(t) = e
C cos
4
4
+

B
B(02 2 )
cos (t) +
sin (t)
2) + 22)
((02 2 ) + 2 2 )
(11)

((02

Oh but wait, we have to plug back in for y(t) = x(t) +


1
t
= e 2

y(t)

C cos

02

t
4

+ D cos

g
,
02

02

t
4

!!

B(02 2 )
B
g
cos (t) +
sin (t) 2
2
2
2
2
2
2
2
2
((0 ) + )
((0 ) + )
0
k
A

, 02 =
, and B =
And remember: =
m
m
m
!
!!
r
r
t

k
k
2
2
= e 2m C cos
t + D cos
t

m 4m2
m 4m2
+

y(t)

A(k m 2 )
A
mg
cos (t) +
sin (t)
(k m 2 )2 + 2 2
(k m 2 )2 + 2 2
k
(12)

Which is the position as a function of time, where C and D are constants


determined by initial conditions.
OK so what happens after a long time? The exponential damping in the
homogeneous term kills off all resonances that would be found from 0 , and the
spring just oscillates according to the inhomogeneous equation. In other words,
the spring just sloshes along with the driving frequency, according to
lim y(t) =

A
mg
A(k m 2 )
cos (t) +
sin (t)
2
2
2
2
2
2
2
2
(k m ) +
(k m ) +
k
(13)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 17, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
#13
#13
#13

Notes/Changes

Last sentence changed

In movie cameras and projectors film speed is 24 f rames/sec. You


see on the screen a car that moves without skidding, and know that
the real-life diameter of its wheels is 1 m. The wheels on the screen
make 3 turns/sec. What is the speed of the car, assuming it is not
moving in excess of 200 mi/hr?

Figure 1: A wheel that appears to rotate one eighth of its circumference.


In one frame, it appears to the viewer that the wheel has turned and angular
velocity

3
24

[turns]

[sec]

[f rames]

[sec]


1 turns
8 f rame

(1)

Jeff Kissel May 17, 2007

Classical Mechanics

However, appearances can be deceiving. To the viewer, measuring the position of the wheel discretely, the wheel has turned only 1/8 of its circumference.
The car could be moving much faster, such that the wheel makes any integer
number N of extra turns before reaching the final 1/8 (for example, N=1 in
Figure 1. The angular velocity should then technically be


1
turns
=
N+
(2)
8 f rame
Finally, to find the (linear) speed v, we must convert Eq. 2 into the correct
units,



:

[f rames]
1
[turns]
[radians]




24
=
N+
2
:



[f rame]
[turn]
8 
[sec]



1
rads/s
(3)
= 48 N +
8
and then convert angular to linear speed, noting that the wheel has a 1 m
diameter, and thus a 0.5 m radius.
v

=


r 
1
rads/s (0.5 m)
=
48 N +
8


1
= 24 N +
m/s
8

(4)

Thus for N = 0, v = 3 m/s 6.7 mi/hr, and N = 1, v = 27 m/s


189.7 mi/hr. (1 m/s = 2.23693629 mi/hr).
Though this problem may seem pretty lame, it brings up an important concepts in signal processing: digital sampling of analog signals, which may result
in aliasing. Check out the Wikipedia article on it!

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A ball of mass m collides with another ball, of mass M = am, initially
at rest. The collision is elastic and central, i.e. the initial velocity of
the ball is along the line connecting the centers of the two objects.

Figure 1: Two balls collide with mass ratio a.


(a) Determine how the energy lost by the moving ball depends on
the mass ratio a, and find the value of a for which the energy loss is
largest. Describe what happens at that value of the mass ratio.
We can solve this problem using conservation of energy and momentum.
The problem does not mention any potential, so we can find the velocity of the
second ball in terms of the first starting with kinetic energy conservation.
=

Ei
>

1 2 1 
ui 2
mvi + M
2

2
m vi2

vi2

Ef

1 2 1
mvf + M u2f
2
2
= 
m vf2 + a
m u2f
=

vf2 + au2f

(1)

We can now bring in conservation of momentum to knock out one of the


unknowns,
pi = pf
i = mvf + M uf
mvi + 
M
u
1

Jeff Kissel October 11, 2006

Classical Mechanics

m vi

vf

= 
m vf + a
muf
= vi auf

(2)

Now, pluggind Eq. 1 into Eq. 2,


vi2

(vi auf )2 + au2f

vi2 = 
vi2 + a2 u2f 2avi uf + avf2

2
0 = a
u
f
 (a + 1) uf 2 a
 vi 
(a + 1)uf = 2 vi
2vi
uf =
(a + 1)

(3)

Finally, the kinetic energy lost by the first ball is just that that is gained by
the second,
E

=
=

1
M u2
2
2

2vi
1
am
2
a+1
2amvi2
(a + 1)2

(4)

To find the value for which the energy lost is largest, we can just maximize
with respect to a,
E
a
0
1
2

2mv
i

(a + 1)2
(a + 1)3
a+1
a=1

2mvi2 a (a + 1)2
a
= 2mvi2 (a + 1)2 4mvi2 a (a + 1)3
2a
2

=
2mv
(a + 1)3  i

= 2a 
(a
+
1)2
=

= 2a
(5)

So, the change in energy is the greatest when the balls have the same exact
mass: the first balls kinetic energy is split evenly between them during the
collision such that vf and uf are equal, but in opposite direcetions.
(b) Investigate the limiting cases of heavy and light balls and comment on your result.
In the limiting case that a 1, the first ball will be much more massive
than the second. In this case, the first ball will lose virtually none of its energy,
but because the second ball is so less massive, it will propel off with uf = 2vi .
lim E

a0

2amvi2
2 =
(a
 + 1)
2

2M vi2

Jeff Kissel October 11, 2006

Classical Mechanics

lim vf

au
= vi 
f =

lim uf

a0
a0

2vi
=
(a
 + 1)

vi
2vi

In the case where a 1, the energy imparted onto the second ball will be
neglegible. In addition, since so little velocity is transferred to the second ball,
the first will be forced in the opposite direction with a final velocity equal to its
initial.
lim E

lim vf

lim uf

2amvi2
=
:

a+
1)2
(
av
i

= vi (a2
 =
 + 1)
i
= 2v
=
:
(
a+
1)
=

0
vi
0

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
Two particles of mass m and m + 2 move in circular orbit around each
other under the influence of gravity. The period of the motion is T .
They are suddenly stopped and then released, after which they fall
towards each other. If the objects are treated as mass points, find
the time they collide in terms of T .

Figure 1: (Left) Two masses orbiting about each other suddenly stopped, as a
two body problem. (Right) The same system reduced to a one-body problem
using the reduced mass .
As noted in the picture, the problem can be simplified to a one body problem
using the reduced mass,
=
=
=

m1 m2
m1 + m2
m2 + 2m
2m + 2
1 m+2
2 (1 + m1 )

(1)

Now a one body problem, we can find the Lagrangian of this more simple system,
V

k
r
1

Jeff Kissel October 11, 2006

T
L
L

Classical Mechanics

1
1
r 2 + r2 2
2
2
= T V
1
k
=
(r 2 + r2 2 ) +
2
r

where Ive defined the gravitational constant of


here we crank the usual Lagrangian formalism,


L
d L

dt r
r
d
k
(r)
(r 2 )
dt
r
k

r r + 2
r


d L
L

dt



d
r2 (0)
dt
r2

(2)

the system k = Gm1 m2 . From

=
=

0
0

(3)

(4)

It appears as though the sooner will be a little more useful than the latter,
so lets roll with it. We know for circular motion, not only is Eq. 4 true, but
r = 0 is also true. Using this fact, we can distill Eq. 3 a bit to get a familiar
law from my buddy Kepler:
k

r r + 2


r
k
r2

r3

k
2

r3

kT 2
4 2

( = =

2
)
T
(5)

Now that weve naively re-derived Keplers 3rd, lets stop the masses. Taking
Eq. 5 as our starting radius, r03 , we know stop the angular motion of the masses,

i.e. set theta,


so the Eq. 3 instead becomes
0 =
r =
( r =

r 
r
+ 2
r
k

r2
r
r r
r
=
= r
)
t
t r
r
2

Jeff Kissel October 11, 2006

Classical Mechanics

r dr

1 2
r
2

r 2

k
dr
r2

k
+C
r
2k
+ 2C
r

We can use the initial condition that at r = r0 , r = 0, so that C = rk0 , and


finally we can solve for the linear merger time,
s
2k
2k

r =
r r0
 
 12
2k 1
1
dr =
dt

r
r0
dr
dt =  
 12
2k 1
1

r r0
 21
  21 Z 0  1
1

t =
dr
(6)
2k
r
r0
r0
which is a horribly disgusting integral, that can be solved analytically using a
few u substitutions, (let u = r1 , and then u = u0 sec2 ) but Ill happily just
leave it as Eq. 6. Numerically evaluated this turns out to be
T
t
8 2

(7)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
Three points masses of identical mass are located at (a, 0, 0), (0, a, 2a)
and (0, 2a, a). Find the moment of inertia tensor around the origin,
the principal moments of inertia, and a set of principal axes.

Figure 1: A system of three identical masses.


The moment of inertia tensor for a discrete mass system is given by
X

mi jk ri2 (xj xk )i
I = Ijk =
i

So for
~r1 = a x

r12 = a2

~r2 = a y + 2a z
~r3 = 2a y + a z

r22 = a2 + 4a2 = 5a2


r32 = 4a2 + a2 = 5a2

(1)

Jeff Kissel October 11, 2006

Classical Mechanics

the diagonal elements of I are


X

mi ri2 xi xi )
I11 =
i

= m(r12 x1 x1 ) + m(r22 x2 x2 ) + m(r32 x3 x3 )

I22

= m(a2 a2 ) + m(5a2 0) + m(5a2 0)


= 10ma2
X

mi ri2 yi yi )
=

(2)

= m(r12 y1 y1 ) + m(r22 y2 y2 ) + m(r32 y3 y3 )

I33

= m(a2 0) + m(5a2 a2 ) + m(5a2 4a2 )


= 6ma2
X

mi ri2 zi zi )
=

(3)

= m(r12 z1 z1 ) + m(r22 z2 z2 ) + m(r32 z3 z3 )


= m(a2 0) + m(5a2 4a2 ) + m(5a2 a2 )
= 6ma2

(4)

And the off-diagonal elements are


X
mi (xi yi )
I12 = I21 =
i

I13 = I31

=
=

m(x1 y1 ) + m(x2 y2 ) + m(x3 y3 )


m(0) + m(0) + m(0)

0
X

(5)
mi (xi zi )

I23 = I32

=
=

m(x1 z1 ) + m(x2 z2 ) + m(x3 z3 )


m(0) + m(0) + m(0)

0
X

(6)
mi (yi zi )

=
=

m(y1 z1 ) + m(y2 z2 ) + m(y3 z3 )


m(0) m(2a2 ) m(2a2 )

4ma2

(7)

So the moment of inertia tensor for this system is

10ma2
0
0
0
6ma2 4ma2
I=
0
4ma2 6ma2

(8)

Jeff Kissel October 11, 2006

Classical Mechanics

The principle moments of inertia and principle axes are the eigenvalues and
eigenvectors (respectively) of the moment of inertia tensor. So to find eigenvalues, we find the characteristic equation of I



10ma2
0
0
h
i


0
6ma2
4ma2
det I b
1 =

0
4ma2
6ma2

0 = (10ma2 ) (6ma2 )(6ma2 ) (4ma2 )2 0 + 0

= (10ma2 ) 36m2 a4 12ma2 + 2 16m2 a4

= (10ma2 ) 2 12ma2 + 20m2 a4
= (10ma2 )(10ma2 )(2ma2 )
1 = 10ma2

3 = 2ma2

2 = 10ma2

(9)

so there is a two-fold degeneracy in the moments of inertia which means the


system is akin to a symmetrical top. To find the eigenvectors (principle axes),
we demand that the principle moments of inertia obey the eigenvalue equation
(I b
1) ~x = 0. So for 1 = 2 = 10ma2 ,


0
0
0
x
0
0 4ma2 4ma2 y = 0
0 4ma2 4ma2
z
0
4ma2 y 4ma2 z

y = z

= 0

so, picking the simplest eigenvectors, the principle axes for 1 & 2 are


0
1
1 = 1 and 2 = 0
1
0
And for 3 = 2ma2 ,

8ma2
0
0

0
4ma2
4ma2

0
x
4ma2 y
4ma2
z
8ma2 x
2

4ma y 4ma z
4ma2 y + 4ma2 z

x = 0,

0
= 0
0
= 0
= 0
= 0

y = z

and again picking the simplest eigenvector, the principle axis for 3 is

0
3 = 1
1

(10)

(11)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
Consider a double pendulum consisting of a mass m suspended on
a massless rod of length , to which is attached by a pivot another
identical rod with an identical mass m attached at the end, as shown
in the Figure 1.

Figure 1: A double pendulum, with identical masses and string lengths.


Using the angles 1 and 2 as generalized coordinates,
a) Find a Lagrangian for the system.
Using 1 and 2 as generalized coordinates, the position (~r1 , ~r2 ) and velocity
(~v1 , ~v2 ) vectors become
~r1
~v1 = ~r 1
~r2

= ( sin 1 ) x
+ ( cos 1 ) y




1 sin 1 y
=
1 cos 1 x

= ( sin 1 + sin 2 ) x
+ ( cos 1 + cos 2 ) y
1

(1)
(2)

Jeff Kissel October 11, 2006

v~2 = ~r 2

Classical Mechanics

= (sin 1 + sin 2 ) x + (cos 1 + cos 2 ) y






=
1 cos 1 + 2 cos 2 x 1 sin 1 + 2 sin 2 y




1 sin 1 + 2 sin 2 y
= 1 cos 1 + 2 cos 2 x

(3)

(4)

In the kinetic energy term of the Lagrangian, well need the squares of the
velocities, which are
v12 = v~1 v~1

v22 = v~2 v~2

2

+ 1 sin 1

cos2 1 + sin2 1

1 cos 1

2

= 2 12
(5)
= 2 12
2
2 

=
1 cos 1 + 2 cos 2 + 1 sin 1 + 2 sin 2
= 2 12 cos2 1 + 2 22 cos2 2 + 22 1 2 cos 1 cos 2
+ 2 12 sin2 1 + 2 22 sin2 2 + 22 1 2 sin 1 sin 2


= 2 2 cos2 1 + sin2 1 + 2 2 cos2 2 + sin2 2
1

+ 22 1 2 (cos 1 cos 2 + sin 1 sin 2 )


= 2 12 + 2 22 + 22 1 2 cos (2 1 )

(6)

Finally, noting that ~g = g y and m1 = m2 = m we can combine the results


of Eqs. 1, 3, 5, and 6 to find the kinetic and potential energies:
T

=
=

1
1
m~v12 + m~v12
2
2

1 
1
2 2
m 1 + m 2 12 + 2 22 + 22 1 2 cos (2 1 )
2
2
1
2 2
m 1 + m2 22 + m2 1 2 cos (2 1 )
2
m~g ~r1 + m~g ~r2
mg cos 1 mg (cos 1 + cos 2 )

2mg cos 1 mg cos 2

=
=
=

(7)

(8)

which we combine to form a Lagrangian for a double pendulum with identical


masses and string lengths:
L

=
=

T V
1
m2 12 + m2 22 + m2 1 2 cos (2 1 ) + 2mg cos 1 + mg cos 2
2

1 2  2 2
(9)
m 21 + 2 + 21 2 cos (2 1 ) + mg (2 cos 1 + cos 2 )
2

Jeff Kissel October 11, 2006

Classical Mechanics

b) Find an approximate Lagrangian that is appropriate for small oscillations and obtain from it the equations of motions when 1 & 2 1.
For small oscillations, we can approximate cos x 1 12 x2 + O(x4 ), and
drop those terms which are of higher order than quadratic in i and i so that
Eq. 9 is


1
1
2
(2
)
L m2 12 + m2 22 + m2 1 2 1 
1
2
2




1
1
+ 2mg 1 12 + mg 1 22
2
2
where what is canceled above will turn out to be quartic in i and i so it is
dropped. Simplified, the approximate Lagrangian is then
L

1
1
m2 12 + m2 22 + m2 1 2 + 2mg mg12 + mg mg22
2
2
 1

1 2  2 2
(10)

m 21 + 2 + 21 2 mg 212 + 22 + 3mg
2
2

c) Assuming that each angle varies as 1,2 = A1,2 eit , find the frequencies for small oscillations.
Performing the usual Lagrange processes using Eq. 10 to find the equations
of motion, we can then use these to solve for the frequencies of small oscillations.
In general, Lagranges equations of motion are defined by


L
d L

=0
(11)
dt qi
qi
where qi are the generalized coordinates of the system, and qi are their respective
time derivatives.
So the equations of motion for 1 and 2 are
L
= 2m2 1 + m2 2
1
2m2 1 + m2 2


21 + 2

L
= m2 2 + m2 1
2
m2 2 + m2 1


1 + 2
3

L
= 2mg 1
1
+

2mg 1 = 0

2g1 = 0

(12)

L
= mg 2
2
+

mg 2 = 0

g 2 = 0

(13)

Jeff Kissel October 11, 2006

Classical Mechanics

If we assume j = Aj eit , then j 2 j , and the equations of motion yield


four oscillation frequencies:

+ 2g 1 = 0
2 2 1 2 2

it
2
it
= 0


+ 2gA 
eit
2A 
e +A 
e
1

2 (2A1 + A2 )
2

= 2gA1
2gA1
=
(2A1 + A2 )
s
2gA1
=
(2A1 + A2 )


2 1 2 2




2 A1 
eit + A2 
eit

+ g 2 = 0

+ gA2 
eit = 0

gA2
(A1 + A2 )
s
gA2
=
(A1 + A2 )

2 (A1 + A2 )

So der you go.

(14)

= gA2
=

(15)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 22, 2006
A particle of mass m. at rest initially, slides without friction on a
wedge of angle and and mass M that can move without friction on
a smooth horizontal surface.

Figure 1: A mass m slides in the +


x direction down a frictionless wedge M ,
which can slide across a frictionless plane in the x
direction. Note, this picture
does not represent the actual motion; realistically X is in the
x direction as
will be shown later.
a) What is the Hamiltonian?
First we must find the Lagrangian (L = T V ), to determine whether it is
explicitly independent of time. If this is so, then the Hamiltonian is simply the
total energy (H = T + V ).
We can define a coordinate system where the origin is the top of the wedge
at some initial time as in Figure 1, such that the position vector for the top
corner of the wedge at time, t will be
~ = X x + (0) y
R

(1)

and particles position vector will be


~r
~r

=
=

~ + (x x
R
+ y y)
(X + cos ) x
+ ( sin ) y
1

(2)

Jeff Kissel October 22, 2006

Classical Mechanics

where is the distance the mass has traveled down the ramp. While the particle
is on the ramp, it is constrained to move along , or
y
x
yY
=
xX
= (y (0)) cos

tan

sin
cos
(x X) sin
f : (x X) sin

y cos = 0

(3)

which is an equation of constraint related to the normal force.


The kinetic energy of the system is that of two free particles, since there is
no friction:
1
1
(4)
T = m(x 2 + y 2 ) + M X 2
2
2
The only potential in the system is gravitational, with ~g = g y. We can ignore
the potential on the wedge because it is constant, so the systems potential is
just the gravitational potential of the particle,
Vg

m~g ~r

= mgy

(5)

The Lagrangian is then potential subtracted from the kinetic energy,


L

T V
1
1
=
m(x 2 + y 2 ) + M X 2 + mgy
2
2
1
1
1
=
mx 2 + my 2 + M X 2 + mgy
2
2
2

(6)

which does not explicitly depend on time, so the Hamiltonian is the sum of the
kinetic and potential energies,
H=

1
1
1
mx 2 + my 2 + M X 2 mgy
2
2
2

(7)

b) Derive the equation of motion from the Lagrangian.


With Lagrange multipliers, The Euler-Lagrange equation in general is


L
f
f
d L

=
+
dt qi
qi
qi
qi

(8)

yet the second term on the right-hand side is zero because Eq. 3 is not a function
of any qi . From Eq. 8, the equations of motion for each coordinate are,
d
(mx)

=
dt
2

(sin )

Jeff Kissel October 22, 2006

Classical Mechanics

m
x =
d
(my)
mg
dt
m
y mg

sin

(9)

( cos )

cos

d  
mX
=
dt
=
mX

(10)

( sin )
sin

(11)

Immediately, if we add Eq. 11 to Eq. 9, we see why Figure 1 has the direction
of motion for the wedge incorrect,

m
x + MX

x
=

M
X
m

(12)

Also, this tells us that the x


position of the systems center of mass stays stationary because the only external force on the system is gravity, which is in the
y direction. Weve assumed that the particles initial position is at the origin,
but if we also assume that the initial velocity of the center of mass is also zero
then from Eq. 12,
x
t t
x
t
t
Z x
x

=
=
=

M
x =
X
m

M
m
M

m
M

X
t t
X
t
t
Z X
X
0

or X =

m
x
M

(13)

giving us an equation relating X in terms of x. We can use the equation of


constraint (Eq. 3) to get y in terms of x,
(x X) sin
m
(x +
x) sin
M

= y cos

= x(1 +

= y cos
m
) tan
M

(14)

from which only y and x are a function of time, so


m
) tan
M
m
) tan
y = x(1 +
M
y

= x(1
+

(15)

Jeff Kissel October 22, 2006

Classical Mechanics

OK, OK, lets finally get the equations of motion. Starting with Eq. 9,
m
x = sin

m(
y g) 
Eq. 10: m
y mg = cos =
cos
m
x = m(g y) tan
(16)


m
Eq. 15: y = x
(1 +
) tan
M
m
) tan2
m
x = mg tan m
x(1 +
M
m
m
x + m
x tan2 + m
x
tan2 = mg tan2
M
m
tan2 )
x = 
mg tan
m(1 + tan2 +

M
cos2
sin2
m sin2
sin
( 2 +
+
)
x = g
2
2
cos cos M cos
cos
1
:



2  2
(
cos
+ sin ) + (m/M ) sin2 x
= g sin cos
g sin cos
x
=
1 + (m/M ) sin2
M sin cos
x
= g
(17)
M + m sin2
Since the acceleration in the x
direction is a constant in time, ax Eq. 17 has
the usual 1-D kinematic solution,
1
x0 + v0,x t + ax t2
2


set x0 0, and v0,x 0


M sin cos
1
g
t2
x(t) =
2
M + m sin2
x(t)

(18)

For the y equation of motion, we plug Eq. 17 back into Eq. 15,
y =
=
y =

m
x
(1 +
) tan
 M


 sin 
sin
M
cos

1

(M + m) 
g


cos

M
M + m sin2 
2
(M + m) sin
g
M + m sin2

(19)

which is also constant, so with the same initial conditions (y0 0, y0,y 0)
we get a similar free-fall equation,


(M + m) sin2
1
t2
(20)
y= g
2
M + m sin2
4

Jeff Kissel October 22, 2006

Classical Mechanics

Finally we get the solution for X plugging Eq. 17 into Eq. 12,
m

X
= x

M
 sin cos
M
m

g
=

M
M
+ m sin2

m sin cos

X
= g
M + m sin2


X0 0, and v0,X 0


1
m sin cos
X = g
t2
2
M + m sin2

(21)

One could also solve for the normal force via Eq. 9, but the problem does
not ask for it, so it is left as an exercise to the reader. (Ew. I feel so dirty saying
that.)
c) What are the constants of motion?
To find the constants of motion, i.e. conserved quantities, we need to express
the Lagrangian (Eq. 6), in terms of independent coordinates. We can use the
constraint equation (Eq. 3) to solve for X,
(x

x
X

L =
L =

X) sin y cos = 0
y
= 0
X
tan
y
x
tan
y
x
tan
1
1
1
mx 2 + my 2 + M X 2 + mgy
2
2
2
2

1
1
1
y
+ mgy
mx 2 + my 2 + M x
2
2
2
tan

(22)

Because this Lagrangian is explicitly independent of x, then the L


x term in
the Euler-Lagrange equation will be zero. Thus, the canonical momentum in
the x
direction is a conserved quantity.
px

L
x

px


x

y
tan
M y
= (m + M )x
tan

= mx + M

(1)
(23)

General Rule: If the Lagrangian is independent of any generalized coordinate, that coordinate is cyclic, and therefore its respective canonical momentum (pi L/ qi ) is conserved.
5

Jeff Kissel October 22, 2006

Classical Mechanics

Also, the Lagrangian does not explicitly depend on time, so as previously


stated, the total energy is conserved (which is why we could write the Hamiltonian as the sum of kinetic and potential energies).
d) Describe the motion of the system.

Figure 2: An alternate view of the particle-wedge system at times t = 1, 2, 3


and 4.
Since friction is not involved in the system, the x
coordinate of the center of
mass between the two objects will remain stationary as has been shown in part
b. The y coordinate will accelerate as though in slow free-fall (a la Eq. 20), until
the particle reaches the surface on which the wedge slides. From then on, it will
simply travel according to Eq. 18. Thus, (as in Figure 2) from a reference point
following the center of mass, the wedge and particle will fly apart from each
0), and the system approximates
other. If M m, it will not move much (X
that of a mass sliding down a fixed incline.

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 26, 2006
A uniform rod slides with its ends inside a smooth (frictionless) vertical circle of radius a. The rod of uniform density and mass m subtends
an angle of 120 at the center of the circle.

Figure 1: A rod, with length L (which is the length of a chord subtending 120 )
is restricted to slide inside a frictionless circular pipe.
(a) Compute the center of mass moment of inertia of the rod ICoM in
terms of m and a (not the length of the rod).
In most general terms, the moment of inertia tensor is calculated as
Z
ICoM =
(r)(ij r2 xi xj ) d3 r

(1)

However, we assume the rod is only has one dimension so we need only
integrate along d instead of d3 r; there are no off-diagonal terms (i.e. xi xj = 0);
and the rod has constant density, so (r) = m/L, which means Eq. 1 becomes
considerable less intimidating,
ICoM

m
L

+L/2

m
d = 2
L
2

L/2

L/2

2 d

Jeff Kissel October 26, 2006

Classical Mechanics

 3
m 3 L/2
m 1 L
=
2

L 3 0
L 3
2
3
2
L
=
m
24
L
1
mL2
(2)
ICoM =
12
Finally, because we know angle which the rod subtends is 120 , it forms an
isosceles triangle whose sides are a : a : L,
and interior angles are 120 : 30 : 30,
which means L/2 = a cos (30 ), or L = 3 a. Thus, in terms of m and a, the
center of mass moment of inertia is
1
ICoM =
ma2
(3)
4
=

(b) Obtain the potential energy, the kinetic energy, and the La
grangian L((t), (t);
m, a, g) for this system, where (t) the dynamical
variable, is the instantaneous angular position relative to its equilibrium position and m, a, and g are constant parameters of the system.
As the question implies later, once we have the center of mass moment of
inertia, we can treat the rod like a pendulum, as long as we displace ICoM to
the center of the circle, a distance d = a sin (30 ). So, following the parallel
axis theorem,
ICoM + md2
1
ma2 + m(a sin (30 ))2
=
4

2
1
1
=
ma2 + m
a
4
2
1
1
ma2 + ma2
=
4
4
1
2
ma
(4)
Id =
2
Now the kinetic energy T is only rotational, the potential V is only gravitational since there is no friction, and the Lagrangian is L = T V ,
Id

=
=

1 2
Id
2

1 1
2
ma 2
2 2
1
ma2 2
4

= mga cos

1
ma2 2 mga cos
4
2

(5)
(6)
(7)

Jeff Kissel October 26, 2006

Classical Mechanics

(c) Compare the dynamics of this system with that of a simple pendulum with mass M and length L (exactly, without any approximations
such as the small oscillation approximation). Compare the parameters L and M to a and m.
With out making any approximations (I promise!) we can find the equation of motion, and compare those to that of a simple pendulum. Lets get
Lagrangimiphysical!


d L
L
= 0

dt



d 1
+ mga sin = 0
ma2
dt 2
1
ma2 = 
mg a
 sin
2
g
= 2 sin
(8)
a
which is indeed comparable to a simple pendulums equation of motion which
is different only by the factor of 2 if L a and M m. For a pendulum of
length L, and mass M ,
T =

1
M L2 2
2

V = M gL cos

L =


d L

dt

d 
+
M L2
dt
L2 =
M

=

1
M L2 2 M gL cos
2
L
= 0

M gL sin = 0
g

M
L sin
g
sin
L

(9)

(c) Find the frequency of small oscillations of the system.


Oh, NOW its OK to make approximations, huh? Fine. Gosh.
In using the small angle approximation, sin , which means Eq. 8
becomes a second-order differential equation of the form x
= 2 x which has
the general solution x(t) = x0 cos (t), where is the frequency (small) of
oscillation(s).

2g

ra
2g
a

(10)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
X
X
#22

Notes/Changes

Added to bank.

A mass m is attached to a spring of spring constant k that can slide


vertically on a pole without friction, and moves along a frictionless
inclined plane as shown in Fig 1. After an initial displacement along
the plane, the mass is released. Find an expression for the x and y
position of the mass as a function of time. The initial displacement
of the mass is x0 . You may assume that the object never slides down
the ramp so far that it strikes the floor.

Figure 1: A box of undisclosed materials oscillating on a ramp.


If one sets up the coordinates as in Fig 1, where the equilibrium is at the
origin, then two important simplifications occur:
1) y
2) x

= ax + b y = ax
= x xe x = x

With these simplifications, the kinetic energy is



1
m x 2 + y 2
T =
2
1

(1)
(2)

(3)

Jeff Kissel May 18, 2007

Classical Mechanics

and the potential energy is the sum of that from the spring and that from
gravity,
V

1
kx2 + mgy
2
(Because of Eq. 2, x = x)
1
kx2 + mgy
2

(4)

Making the Lagrangian for the system in terms of x and y,


L =
=
L =

T V



1
1
2
2
2
m x + y
kx + mgy
2
2
1
1
1
mx 2 + my 2 kx2 mgy
2
2
2

(5)

However, there is a fixed relationship between x and y, namely the Eq. 1, so


y

ax y = ax

(6)

which makes Eq. 5 a function only of x and x,

L =
=
L =

1
1
1
mx 2 + m(ax)
2 kx2 mg(ax)
2
2
2
1
1
1
mx 2 + ma2 x 2 kx2 mgax
2
2
2
1
1
m(1 + a2 )x 2 kx2 mgax
2
2

(7)

Using the Euler-Lagrange equation, we can then find an expression for x(t),


d L
L
0 =

dt x
x

d
=
m(1 + a2 )x (kx mga)
dt
0 = m(1 + a2 )
x + kx + mga
k
ga
x =
x
m(1 + a2 )
m(1 + a2 )
ga
k
, and R =
Let 2 =
2
m(1 + a )
m(1 + a2 )
x

x
2

= 2 + R
Let x = 2 x + R
x = 2 x
x
= 2 x



x R
+R
= 2
2
2

Jeff Kissel May 18, 2007

x
2
x

x (t)

Classical Mechanics

= x
= 2 x
= A cos (t)

(8)

Where the last step we know comes from the typical harmonic oscillator solution,
with A as some constant. Plugging back in for x, and using the initial conditions
that x(0) = x0 ,
2 x(t) + R =
x(t)

=
=

x(t)

A cos (t)
A
2 cos (t)

x0 = x(0)

x0 + R2
A

R
2

= A2 (1) R2

= A2
2
= x0 R

R
( 2 x0 R)
cos (t) 2

2



R
R
x0 + 2
cos (t) 2





R
ga
m(1 + a2 )
ga
=
=
2
m(1 + a2 )
k
k
s
!
k
=
m(1 + a2 )
s
!

ga 
ga
k
x0
t +
cos
k
m(1 + a2 )
k

(9)

and using Eq. 1, we can get y(t),


y(t) = ax(t)
s
!

ga
k
ga 
cos
t +
y(t) = a x0
k
m(1 + a2 )
k

(10)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 17, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
X
X
#22

Notes/Changes

An astronaut is on the surface of a spherical asteroid of radius r and


mean density similar to that of the earth. On the earth this astronaut
can jump about a height h of 0.5 m. If he jumps on this asteroid, he
can permanently leave the surface.
(a) Taking the radius of the earth as 6.4106 m, find the largest radius
the asteroid can have.
With the initial velocity on earth v , the astronaut jumps a height h. If he
can leave the asteroid with this same velocity, we can treat v as the escape
veloctiy for the asteroid vesc . We can determine his velocity on Earth from 1-D
kinematics (using the usual trick that his final velocity on the way down is equal
to his initial velocity on the the way up),
vf2

0
2

7

vi =
2
v
=
v

2ax
2g h
p
2g h

The escape velocity on the asteroid occurs when the astronauts total kinetic
energy equals the potential felt from the asteroid, i.e. his total energy is zero,
E

0 =
1
2
m
av
2  esc

1
GMA ma
ma v 2
2
rA
1
GM
A ma
2
ma vesc

2
rA
m
GMA
a
rA
1

Jeff Kissel May 17, 2007

Classical Mechanics

2
vesc

vesc

=
=

vesc

p
2g
h =

rA

rA

=
=

2GMA
r
r A
2GMA
rA

A
A = M
VA
MA = A VA

3
MA = 4
r

A A
3
s


2G 4 3
r A
rA
3 A
r
4A
rA 2G
3
!
A =
A = 4M
3
3 R
v
!
u
u
M
4


rA t2G 
4
3
3
3 R
s
2GM
rA
3
R


GM
g
2 , vesc = v
R
s
2g
rA
R
s
2g

rA 
R
p
R h
p
(6.4 106 m)(0.5 m)
1788.85 m

(1)
(2)

b) How fast could the asteroid rotate and not have the astronaut
be flung away from the surface?
Here, the maximum rotation speed would be when the astronauts weight
on the asteroid exactly equals the asteroids centripetal acceleration,
Fa
m
a gA

gA

= m a ac
2
vt,A
m
= 
a
rA
(vt = r)
2
= max
rA
2

Jeff Kissel May 17, 2007

Classical Mechanics

max

gA
rA

(3)

Yet, we dont know what gravity is like on this asteroid. We can find out,
because its density is equal to Earths,
gA

=
=
=

gA

GMA
2
rA


G 4 3
r A
2
rA
3 A
4
 M R 3

GrA

4
3
3


rA GM
2
R
R
rA
g
R

!!

(4)

So the maximum rotation the asteroid can have with out flinging the astronaut
into space is
s


r
1
A

g
max =
rA R
r
g
(5)
max =
R
r
9.8 m/s
=
6.4 106 m
max = 1.23 103 rads/s
(6)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
#X
#24
#24

Notes/Changes

Added to question bank.

In 1956, G. Kuzmin introduced the simple axisymmetric potential


GM
K (R, z) = p
R2 + (a + |z|)2

(1)

as an approximate description of the potential of a disk-like galaxy.


(a) Show that this potential can be though of as the result of two
point masses placed at R = 0, z = a, generating the potential in the
half-space z > 0 and z < 0 respectively.
In cylindrical polar coordinates, a point mass placed at the origin generates
the following potential,
P M (R, z) =

GM

R2 + z 2

(2)

where R2 = x2 + y 2 .
If the point mass is displaced from the origin a positive distance z0 , without
moving radially, then P M (R, z) becomes
P M (R, z + z0 )

GM
GM
= p
= p
2
2
2
R + (z z0 )
R + (z0 z)2

(3)

where Im able to flip the quantity in parentheses because it is squared.


In the region of z < 0, The Kuzmin potential takes the form
GM
K (R, z) = p
2
R + (a z)2
1

(4)

Jeff Kissel May 18, 2007

Classical Mechanics

and hence by direct comparison with Eq. 3, for z < 0, Eq. 4 is the potential of
a point mass centered at z = a, and R = 0.
Similarly, the potential of a point mass shifted from the origin a negative z0
(again keeping R = R) is
P M (R, z z0 )

GM
GM
= p
= p
(5)
R2 + (z (z0 ))2
R2 + (z0 + z)2

In the region of z > 0, the Kuzmin potential is

GM
K (R, z) = p
2
R + (a + z)2

(6)

so, again by direct comparison with 5, Eq. 6 is the potential of a point mass
centered at z = a, and R = 0.
(b) Show that all the mass must be located on the z = 0 plane and
distributed according to the surface density
k (R) =

aM
2(R2 + a2 )3/2

(7)

If (R, z) below the z = 0 plane acts as a point mass at z = a, then above


the z = 0 plane, Poissons Equation (2 = 4G) contains no mass density
. Poissons Equation then becomes Laplaces Equation: 2 = 0. Similarly,
if (R, z) above the z = 0 plane acts as a point mass at z = a, then the
potential also satisfies Laplaces Equation in the region above the z = 0 plane.
If 2 (R, z > 0) = 2 (R, z < 0) = 0, assuming the mass must be somewhere,
then all the mass must lie in an infinitely thin disk at z = 0, of mass density .
Taking a Gaussian pill-box surrounding the sheet of mass , we know the
potential at z = 0 must satisfy Gauss Law of flux,
I
~ K (R, 0) d~a = 4GK A

(8)
pill-box

where the A is the area of the disk. We can simplify the left-hand side a bit,
Z
I
Z





~ K (R, 0) d~a =
K da =
K
da = A K
(9)

z
z
z
z=0
A
pill-box
A

However, the partial derivative is discontinuous across z = 0, so we can take


the difference between the partial derivative coming from the top and from the
bottom,






K (R, +z)
K (R, z)

4GK A = A
z
z
z=0
z=0
!

GM

=
p

2
2
z
z=0
R + (a + z)
2

Jeff Kissel May 18, 2007

Classical Mechanics

GM

z
=

=
4 GK

p
R2 + (a z)2

z=0


 2

(R + (a + z)2 )1/2
GM
z
z=0

 2
2 1/2
+ GM
(R + (a z) )

z=0

 z
1
2
2 3/2
(2(a + z))(1)
GM (R + (a + z) )
2
z=0


1
2
2 3/2
+ GM (R + (a z) )
(2(a z))(1)
2
z=0
!

(a (0))
(a + (0))

+
GM +

3/2
3/2
2
2
2
2
z=0
(R + (a + (0)) )
(R + (a (0)) )
2a
GM
3/2
(R2 + a2 )

Ma
(10)
3/2
2
2
2 (R + a )

(c) What is the total mass generating K ?


We can integrate the surface density over the disk to find the total mass:
Z

da

=
Z

da

1
2

2Z
0

Ma
(R2

3/2

+ a2 )

R dR d

1
Z

2
Ma

R dR
2
2
 0 (R + a2 )3/2

= 2R dR
du
1
let u = R2 + a2
du =
R dR
2

0 a2 ,
Z
i
Ma h
M a 3/2
u
du =
2u1/2 2
2 a2
a
2
0
h
1 i
1
Ma +

a2
(11)

Thus, the total mass gererating K is M .

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 19, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
X
X
#25

Notes/Changes

Added to bank.

A particle of mass m moves in a force field given by a potential energy


V = Krs where both K and s may be positive or negative.
(a) For what values of K and s do stable, circular orbits exist?
(These results are a product of Goldstein Section 3.6 pp89-90.)
Circular orbits occur when the effective potential,
Veff

V +

2
2mr2

(1)

has an extremum, i.e. when


 2 
V


=
+


r r=R r 2mr2 r=R


V
~
~
|F (r)| |V (r)| =
r


2

0 = (F (R)) +
mR3
2
F (R) =
mR3

mr2

Veff

r r=R

F (R) = m 2 R

(2)

(3)

where Ive denoted R as the circular radius, of which R = 0.


However, these extrema can be either a maxima or minima. Circular orbits
are only stable when the the extrema of the effective potential is a minima, i.e.
1

Jeff Kissel May 19, 2007

Classical Mechanics

when the second derivative of the effective potential is positive. This restricts
both K and s,


Veff
2 Veff
=

r2 r=R
r
r


0 <
(Feff )
r 
r=R

2


F+
<

r
mr3 r=R


32
F

<

r r=R mR4
F
2
0 <
+3

r r=R
mR4

2
F
1
< 3

r
R
mR3
r=R

V


= (sKrs1 )
=
2
F
r
Eq. 2: F (R) =
F

mR3
=
s(s 1)Krs2
r
F (R)
s(s 1)KRs2 < 3
R


V
= (sKrs1 )
F =
r

sKRs1
s2
s(s 1)KR
< 3
R




s2

s2
s(s 1)
KR
< +3 s
KR
(s 1) < +3
s1
s

> 3
> 2

(4)

where canceling out K demands that its positive so


K

<

(5)

(b) What is the relation between the period P of the orbit and the
radius R of the orbit?
Weve already shown that
F

V
= (sKrs1 )
r

(6)

We can set this equal to Eq. 3, using r = R and = 2/P because the orbit is
circular,
sKRs1

m 2 R

=
2

Jeff Kissel May 19, 2007

Classical Mechanics

1
2

P2
(2)2

P2

mR
sKRs1
m
sKRs2
4 2 m 2s
R
sK
4 2 m (2s)/2
R
sK

(7)

For a quick sanity check, lets make sure we get Keplers 3rd Law when
K = GM m and s = 1,
PKepler

=
=

PKepler

4 2 m
R(2(1))/2
(1)(GM m)
4 2
m (2+1)/2
R
GM
m
4 2 3/2
R
GM

Lchayim!

(8)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
November 1, 2006
All parts of this question call for quick and brief answers.
(a) A bead of mass m slides along a wire bent into a parabolic shape.
The wire is pivoted at the origin and is spinning around the vertical.
Identify suitable generalized coordinates and constraints to describe
the beads motion.

Figure 1: A bead confined to a parabolic wire that is spinning about the z axis.
A suitable set of generalized coordinates would be cylindrical coordinates,
r, , and z, (as shown in Figure 1) where the equation of constraint would be
that z = r2 .
(b) A system with generalized coordinates q1 , q2 and q3 is described by
the Lagrangian L(q1 , q2 , q3 , q3 ); that is, the Lagrangian is independent
of q1 , q2 and time (explicitly). What are the conserved quantities of
this motion?
Short answer: The conserved quantities are energy, p1 , and p2 .
Thorough answer: The canonical momentums are defined as
pi

L
qi

(1)

Jeff Kissel November 1, 2006

Classical Mechanics

Which means if the Lagrangian is independent of any generalized coordinate


qi , then its Euler-Lagrange equation will yield a constant of motion,
d
dt


L
qi
d
(pi )
dt
pi

0
7
L

 =0
qi
= 0
= constant

(2)

Thus, since the Lagrangian is independent of q1 and q2 , p1 and p2 are constants of motion.
Also, if the Lagrangian is independent of time then the energy function h is
constant of motion, i.e. energy is conserved, and the Hamiltonian is H = T + V .
The proof is as follows,
dL
dt

dL
dt

Constant =

0
X L
X L
7
L
dqi +
dqi + 
q

i
i
t
i
i


X L
X d L
dqi +
dqi
dt qi
qi
i
i
 




L
d
d L
L dqi
qi
=
dqi +
dt
qi
dt qi
qi dt


L
d
qi
dt
qi
X d  L  dL
qi

dt
qi
dt
i
!
d X L
qi
L
dt
qi
i
X L
L
qi
h =
qi
i

(3)

(c) IfP
the kinetic energy isP
a quadratic function of generalized velocities, k, ak qk q , express k qk (T / qk ) in terms of T .

Eulers theorem states if a function f is a an homogeneous function of degree


n in variables xi , then
X
f
xi
= nf
(4)
x
i
i

Since T is quadratic in terms of q,


T has degree 2, so
X
k

qk

T
= 2T
qk
2

(5)

Jeff Kissel November 1, 2006

Classical Mechanics

(d) A hollow and a solid sphere, made of different materials so as to


have the same masses M and radii R, roll down an inclined plane,
starting from rest at the top. Which one will reach the bottom first?
Short Answer: The solid sphere will reach the bottom faster because it has
a smaller moment of rotational inertia, i.e. less energy is put into rotating the
solid sphere so more energy can by used in linear velocity of the center of mass.
Thorough answer: The moments of inertia for a solid and hollow sphere
(abbreviated as SS and HS, respectively) are
ISS =

2
M R2
5

IHS =

2
M R2
3

(6)

Remember v = R, so each spheres kinetic energy is


TSS

=
=
=

TSS

THS

=
=
=

THS

1
1
2
M vCoM,SS
+ ISS 2
2
2


1
vCoM,SS 2
1 2
2
2
M vCoM,SS
+
M
R
2
2 5
R
1
1
2
2
M vCoM,SS
+ M vCoM,SS
2
5
7
2
M vCoM,SS
10
1
1
2
M vCoM,HS
+ IHS 2
2
2


vCoM,HS 2
1 2
1
2
2
M vCoM,HS
+
M
R
2
2 3
R
1
1
2
2
M vCoM,HS
+ M vCoM,HS
2
3
5
2
M vCoM,HS
6

(7)

(8)

Energy is conserved in this system, so the gravitational potential at the top


of the ramp will equal the kinetic energy at the bottom, so
Vi = Tf
7
2
M vCoM,SS
M gh =
10
10
2
gh
vCoM,SS
=
7

M gh =
2
vCoM,HS

5
2
M vCoM,HS
6
6
=
gh
5

(9)

which means the solid sphere will arrive first, since it is traveling at a faster
linear velocity (10/7 > 6/5) down the ramp.

Jeff Kissel November 1, 2006

Classical Mechanics

(e) Write down the Hamiltonian for a free particle in three-dimensional


spherical-polar coordinates.
Short Answer: In (physics) spherical-polar coordinates (where is the angle
from the z axis, and is the angle from the x
axis in the xy plane),
H=

p2
p2
p2r
+
+
2m 2mr2
2mr2 sin2

(10)

Thorough Answer: The Lagrangian in free space (in spherical coordinates)


is
L

1
m(r 2 + r2 2 + r2 sin2 2 )
2

(11)

and Hamiltonian can be found from (and is in fact defined as) a Legendre
transformation between qi and pi ,
H

qi pi L

where pi

L
qi

(12)

So, turning the crank on Eq. 11, and plugging back into Eq. 12,
pr
L
= mr
r =
r
m
L
p
2
p
= mr
=
2

mr

L
p
p
= mr2 sin2 =
2

mr sin2

pr

the Lagrangian as a function of qi and pi is then


L


2 !
 p 2
 p 2
p
1

r
2
2
2
+r
+ r sin
m
2
m
mr2
mr2 sin2
p2
p2r
p2
+
+
2m 2mr2
2mr2 sin2

and the Hamiltonian is




 p 
p 
p

r
pr +
p +
H =
p
m
mr2
mr2 sin2
p2
p2
p2
r

2m
2mr2
2mr2 sin2
2
p
p2
p2
p2
p2r
p2r

+ 2 +

m
mr
mr2 sin2 2m 2mr2
2mr2 sin2
2
2
p
p
p2r
+
+
H =
2
2
2m 2mr
2mr sin2

(13)

(14)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
November 1, 2006
All parts of this question call for quick and brief answers.
(a) In a bound Keplerian orbit, what is the ratio of the average kinetic
energy to the average potential energy?
Short Answer: The only force experienced by a particle in a bound Keplerian
orbit is the central, gravitational inverse square force (a power law of order
n = 2). This gravitational force is derivable from a scalar is derivable from
a scalar potential V . Any system satisfying these conditions follows the Virial
Theorem: the average kinetic energy is twice that of the average potential, or
1
T = V
2

(1)

Thorough Answer: An useful scalar quantity known as the Virial, or the


Virial of Clausius, is defined as
X
~pi ~ri
(2)
G(t) =
i

where the summation is over all particles in a given system. Newtons equations
for each particle in the system are
Fi =

d~
pi
dt

(3)

which gives us the inspiration to take the full time derivative of the Virial,
dG(t)
dt

=
=

X d~
X
d~ri
pi
~pi
~ri +
dt
dt
i
i
X
X
mi~r i ~r i
F~i ~ri +
i

X
i

dG(t)
dt

X
i

F~i ~ri +

X
i

F~i ~ri + 2T
1

mri2

(4)

Jeff Kissel November 1, 2006

Classical Mechanics

Since were interested in the (time) average(d) kinetic energy, T , well integrate Eq. 4 over time, from 0 to some time and normalize by the same time
,
Z
X
1 dG(t)
F~i ~ri + 2T
dt =

0
dt

i
X
1
(5)
F~i ~ri + 2T
(G( ) G(0)) =

i
If we chose to be the period and motion is cyclic (as in a Keplerian orbit),
the left-hand side vanishes (G( ) G(0) = 0) leaving
2T

F~i ~ri

1X ~
Fi ~ri
2 i

(6)

which is the general form of what is known as the Virial Theorem.


If the forces are derivable from a potential, (as in a Keplerian orbit) i.e.
~
~ then the theorem becomes
F = V
T

1X ~
V ~ri
2 i

(7)

and for a single particle moving under a central force (as in a particle in a
Keplerian orbit), this reduces to
T

1 V
r
2 r

(8)

If V is a power law (as in a Keplerian orbit), where the exponent is chosen


so that the force law goes as rn , then
V

= arn+1

V
r = (n + 1)V
r
T

1
(n + 1)V
2

(9)

which means for a Keplerian orbit (whose force follows and inverse square law,
n = 2), the average kinetic energy and average potential have the following
constant of proportionality
T

1
V
2

(10)

Jeff Kissel November 1, 2006

Classical Mechanics

(b) Write down the Lagrangian and the Hamiltonian of a Free Particle
of mass m in three dimensional polar coordinates.
Short Answer: The Lagrangian and Hamiltonian in polar spherical coordinates are
L

1
m(r 2 + r2 2 + r2 sin2 2 )
2
p2
p2r
p2
+
+
2m 2mr2
2mr2 sin2

(11)
(12)

Thorough Answer: A free particle is that which is traveling at some velocity


~v = ~r , without the the influence of any potentials. Thus, the Lagrangian (L =
T V ) will only contain the kinetic term T .
The Cartesian components of the particles position in spherical-polar coordinates (where we use the physics definitions of and : is the angle from the
z axis, and is the angle from the x
axis in the xy plane) are,
x

= r sin cos

y
z

= r sin sin
= r cos

so the components of velocities are,


x =
y =
z

r sin cos + r cos cos r sin sin


r sin sin + r cos sin + r sin cos
r cos r sin

The kinetic energy T of any given particle can always be written as


T

1
m (x 2 + y 2 + z 2 )
2

So, we can plug in the spherical-polar versions of x,


y,
and z from above,
2T
m

d
r 2 sin2g
cos2 + r2 cos2 cos2 2 + r2 sin2 sin2 2

}|
{
z

2

2
 sin cos
2r
r sin
+2rr sin cos cos 
(
(((
((
2

(
(
sin

cos

sin

cos

2r(
(
( (

2
+ r 2 sing
sin2 + r2 cos2 sin2 2 + r2 sin2 d
cos2 2

z
}|
{

2
2
 sin cos
+2rr sin cos sin + 
2r
r sin
(
(((
((
2

(
(
sin

cos

sin

cos

+2r(
(
( (
i
+ r 2 cos2 + r2 sin2 2 2rr sin cos
3

(13)

Jeff Kissel November 1, 2006

=
2T
m

Classical Mechanics

r 2 sin2g
(sin2 + cos2 ) + r2 cos2 (sin2 + cos2 )2
d
+ r2 sin2 (sin2 + cos2 ) 2
2rr sin cos (sin2 + cos2 ) + r 2 cos2
+ r2 sin2 2 2rr sin cos


1 = sin2 + cos2

2 2
2 2
2
r 2 sin2 + r|2 cos
{z } +r sin
(
(
((
((
2 2
((
((
2r(
r sin
cos + r 2 cos2 + r|2 sin
+(
2r(
r sin
cos
{z } (

r 2 (sin2 + cos2 ) + r2 (sin2 + cos2 )2 +r2 sin2 2


|
{z
}
r 2 + r2 2 + r2 sin2 2

1  2
m r + r2 2 + r2 sin2 2
2

(14)

Now as previously mentioned, the potential V for a free particle is zero,


so the Lagrangian (L = T V ) as a function of qi and qi is
L


1  2
m r + r2 2 + r2 sin2 2
2

(15)

The Hamiltonian can be found from (and is in fact defined as) a Legendre
transformation between qi and pi ,
H

X
i

qi pi L

where pi

L
qi

(16)

So, turning the crank on Eq. 15, and plugging back into Eq. 16,
pr
L
= mr
r =
r
m
L
p
2
p
= mr
=
mr2

p
L
= mr2 sin2 =
p
2

mr sin2

pr

the Lagrangian as a function of qi and pi is then


L


2 !
 p 2
 p 2
1
p

r
2
2
2
+r
+ r sin
m
2
m
mr2
mr2 sin2
p2
p2r
p2
+
+
2m 2mr2
2mr2 sin2
4

(17)

Jeff Kissel November 1, 2006

Classical Mechanics

and the Hamiltonian is




p 
 p 
p
r

p
H =
pr +
p
+

m
mr2
mr2 sin2
p2
p2
p2
r

2m
2mr2
2mr2 sin2
p2
p2
p2
p2
p2r
p2r

+ 2 +

2
2
m
mr
mr2 sin 2m 2mr
2mr2 sin2
2
2
p
p
p2r
+
+
H =
2
2m 2mr
2mr2 sin2

(18)

(c) A system with generalized coordinates q1 , q2 and q3 is described by


the Lagrangian L(q1 , q2 , q3 , q3 ); that is, the Lagrangian is independent
of q1 , q2 and time (explicitly). What are the conserved quantities of
this motion?
Short answer: The conserved quantities are energy, p1 , and p2 .
Thorough answer: The canonical momentums are defined as
pi

L
qi

(19)

Which means if the Lagrangian is independent of any generalized coordinate


qi , then its Euler-Lagrange equation will yield a constant of motion,
d
dt


L
qi
d
(pi )
dt
pi

0
7
L

 =0
qi
= 0
= constant

(20)

Thus, since the Lagrangian is independent of q1 and q2 , p1 and p2 are constants of motion.
Also, if the Lagrangian is independent of time then the energy function h is
constant of motion, i.e. energy is conserved, and the Hamiltonian is H = T + V .
The proof is as follows,
dL
dt

0
X L
X L
7
L
dqi +
dqi + 
q

i
i
t
i
i


X L
X d L
dqi +
dqi
dt qi
qi
i
i
 




L
d
d L
L dqi
qi
=
dqi +
dt
qi
dt qi
qi dt
5

Jeff Kissel November 1, 2006

dL
dt

Constant =

Classical Mechanics



d
L
qi
dt
qi
X d  L  dL
qi

dt
qi
dt
i
!
d X L
qi
L
dt
qi
i
X L
L
qi
h =
qi
i

(21)

To recap, weve decided that h (energy), p1 , and p2 are the conserved quantities of this system.
(d) A hollow and a solid sphere, made of different materials so as to
have the same masses M and radii R, roll down an inclined plane,
starting from rest at the top. Which one will reach the bottom first?
Short Answer: The solid sphere will reach the bottom faster because it has
a smaller moment of rotational inertia, i.e. less energy is put into rotating the
solid sphere so more energy can by used in linear velocity of the center of mass.
Thorough answer: The moments of inertia for a solid and hollow sphere
(abbreviated as SS and HS, respectively) are
ISS =

2
M R2
5

IHS =

2
M R2
3

(22)

Remember v = R, so each spheres kinetic energy is


TSS

=
=
=

TSS

THS

=
=
=

THS

1
1
2
M vCoM,SS
+ ISS 2
2
2


vCoM,SS 2
1 2
1
2
2
M vCoM,SS
+
M
R
2
2 5
R
1
1
2
2
M vCoM,SS
+ M vCoM,SS
2
5
7
2
M vCoM,SS
10
1
1
2
M vCoM,HS
+ IHS 2
2
2


1
vCoM,HS 2
1 2
2
2
M vCoM,HS
+
M
R
2
2 3
R
1
1
2
2
M vCoM,HS
+ M vCoM,HS
2
3
5
2
M vCoM,HS
6
6

(23)

(24)

Jeff Kissel November 1, 2006

Classical Mechanics

Energy is conserved in this system, so the gravitational potential at the top


of the ramp will equal the kinetic energy at the bottom, so
Vi = Tf
7
2
M vCoM,SS
10
10
2
gh
vCoM,SS
=
7

M gh =

M gh =
2
vCoM,HS

5
2
M vCoM,HS
6
6
=
gh
5

(25)

which means the solid sphere will arrive first, since it is traveling at a faster
linear velocity (10/7 > 6/5) down the ramp.
(e) A satellite initially at a distance RE (= radius of the Earth) above
the Earths surface is moved out to a distance 2RE . How does its
orbital time period change?
A satellite orbiting around the Earth can be approximated quite accurately
by a Keplerian orbit. Keplers 3rd law says that such an orbits period P is
proportional to the radius of orbit a taken to the 3/2s power, so
P2
P
P

a3

8P 2

(2a)3
8a3

The satellites period increases by a factor of


doubled.

(26)

8, if the radius of orbit is

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
A particle of mass m moves under the influence of an attractive
force F = k r, where r is the distance from the force center.
What are the conserved quantities of this three dimensional motion?
If the Lagrangian of the system is independent of any generalized coordinate qi (i.e. cyclic in qi , the its Euler-Lagrange equation demands that the
respective canonical momentum pi is a constant of motion,
0

0
pi

0



d L
L7
=

dt qi
qi


L
pi
qi
d
=
(pi )
dt
= constant

(1)

In order to determine if this system has any of its canonical momentums


conserved, we must find the Lagrangian, L = T V . The force of this system
(F = k r) is derivable from a potential V in which we need,
F

~ V

Z r

(k r ) dr
0

1 2
V =
kr
(2)
2
The kinetic energy T is that of any moving particle, which spherical polar coordinates (where is the angle from the z axis and is the angle from the x
to
the project of the vector onto the xy plane) is

1  2
(3)
T =
m r + r2 2 + r2 sin2 2
2
Combining Eqs. 2 and 3 yields the Lagrangian,
 1
1  2
m r + r2 2 + r2 sin2 2 kr2
(4)
L =
2
2
1

Jeff Kissel May 18, 2007

Classical Mechanics

Since Eq. 4 is independent of , p is a constant of motion.


Also, if the Lagrangian is independent of time then the energy function h is
a constant of motion. Because of this energy is conserved, and the total energy
E of the system is the Hamiltonian E = H = T + V . The proof is as follows,
dL
dt

=
=

dL
dt

0 =
=

0 =
constant =

0
X L dqi X L dqi

L7
+
+ 
qi dt
qi dt t
i
i
X d  L  dqi X L dqi
+
dt qi
dt
qi dt
i
i

X  d  L  dqi
L dqi
+
dt qi
dt
qi dt
i
 




L
d
d L
dqi
L dqi
qi
=
+
dt
qi
dt qi
dt
qi dt


X d
L
qi
dt

qi
i


X d
L
dL
qi

dt

dt
i
i
!
d X L
qi
L
dt
qi
i


L
pi
qi
!
d X
qi pi L
dt
i
X
qi pi L h

(5)

Write down an expression for the total energy.


As shown in part (a), the total energy E is conserved, thus it is the Hamiltonian E = H = T + V . From Eqs. 2 and 3, the total energy as a function of
qi and qi is then
E = H

 1
1  2
m r + r2 2 + r2 sin2 2 + kr2
2
2

(6)

Either from (b) or otherwise, derive an orbit equation relating r to .


As stated previously, the Lagrangian for the system cyclic of , so the system
is spherically symmetric. The total angular momentum vector
~j

= ~r p~
2

(7)

Jeff Kissel May 18, 2007

Classical Mechanics

is conserved. The definition of the cross product requires then that ~j is perpendicular to ~r (and p~). This can only be true if ~r lies in a plane whose normal
is parallel to ~j. Thus, motion is restricted to a plane, and must not only be
constant but zero. If = 0, then we can reduce the Lagrangian (Eq. 4) to just

a function of r, r,
and ,
 1
1  2
m r + r2 2 + kr2
(8)
E =
2
2
and p is now a constant of motion because Eq. 8 is cyclic in . We can find

this constant (denoted as p ), and therefore ,


p

= mr2 =
2

mr

(9)

This means the total energy (after killing the term) becomes

1  2
m r + r2 2 + V
E =
2

2 !
1

2
2
=
m r + r
2
mr2
E

1
2
+V
mr 2 +
2
2mr2

From here, we can solve Eq. 10 for r,


and then r(),
1
mr 2
2

r 2

mr2 d
dr

=
=
=

2
2mr2
2V
2
2E

2 2
m
m
m r

1/2
2E
2V
2

2 2
m
m
m r


dr
dr
dr d
=
=
r =
d dt
d
mr2 d


1/2
2V
2
2E

2 2
m
m
m r

1/2

2E
2V
2

2 2
dr
mr2 m
m
m r


1/2
kr2
2
1 m2 2E

dr
r 2 2
m
m
m2 r 2
1/2

mkr2
1
1
2mE

dr
2
2
r2
r2

du = r2 dr
let u = r1
r2 =
u2

EV

(10)

Jeff Kissel May 18, 2007

=
d

2 d

Classical Mechanics

1/2
2mE
mk
2

2 2 u
du
2
u

2u dv
dv =
u =
v 1/2
let v = u2

1 1/2
du = 2 v
dv

1/2
mk
2mE
1 1/2

2 v
u
dv
2
v
2

1/2
mk
1 2mvE
2

v
dv

2
2
2
2mE
let =
and = mk
2

1/2

v 2 v 2
dv

2
2
v 2 v 2 =
v 2 v 2 + 4 4 
2
2

= 4 2 v 2 v + 4


 2
2

v 2
=

2
4

1/2
 
 2
2

2 v
dv

2


dy
= v 2
2

= 4 2

1/2
dy
a2 y 2
Z
1
p
dy
a2 y 2
y
+2
sin1
a

let y
let a2

2 d
Z
2 d

2 2

v 2
q
2

4 2

mE
u 2

1
r2

= dv

v 2

sin1 q
2

4
2
sin (2 2)

1/2
mk
m2 E 2
2
sin (2( ))
4

(sin (A B) = 2 sin A cos B)


mE
2p
+ 2 m2 E 2 mk2 sin ( ) cos ( )
2

(11)

where is an arbitrary integration/phase constant. *Phew*. Thats my final


answer. The orbit equation for a simple harmonic oscillator potential.

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
November 5, 2006
All parts of this question call for quick and brief answers.
(a) Positronium is a bound state of an electron and a positron. What
is the effective mass of positronium?
Short Answer: For classical mechanics purposes, the effective mass of a
system of particles is its reduced mass. An electron and a positron differ in
charge but have the same mass, so positroniums effective mass is
mP s

=
=
=

mP s

P s
me me+
me + me+
m2e
2me
1
me
2

(1)

Thorough Answer:
Effective mass is defined by analogy with Newtons second law F~ = m~a.
Using quantum mechanics it can be shown that for an electron in an external
electric field E,
1 d2 (k)
qE
(2)
a= 2
dk 2
h
where a is acceleration, h is reduced Plancks constant, h
= h/2, k is the wave
number (often loosely called momentum since k = p/h), (k) is the energy as a
function of k, or the dispersion relation as it is often called. From the external
~ where q
electric field alone, the electron would experience a force of F~ = q E,
is the charge. Hence under the model that only the external electric field acts,
effective mass m becomes:
1
 2
d (k)
2

m =h

(3)
dk 2
For a free particle, the dispersion relation is a quadratic, and so the effective
mass would be constant (and equal to the real mass). In a crystal, the situation is
1

Jeff Kissel November 5, 2006

Classical Mechanics

far more complex. The dispersion relation is not even approximately quadratic,
in the large scale. However, wherever a minimum occurs in the dispersion
relation, the minimum can be approximated by a quadratic curve in the small
region around that minimum. Hence, for electrons which have energy close to
a minimum, effective mass is a useful concept.
In energy regions far away from a minimum, effective mass can be negative or
even approach infinity. Effective mass, being generally dependent on direction
(with respect to the crystal axes), is a tensor. However, for most calculations
the various directions can be averaged out.
Effective mass should not be confused with reduced mass, which is a concept from Newtonian mechanics. Effective mass can only be understood with
quantum mechanics.
http://en.wikipedia.org/wiki/Effective_mass, 2006
Regardless, since this question is asked in the classical mechanics portion
of the exam, (and every other classical mechanics reference I looked at refers
positroniums reduced mass) I assume effective mass is equivalent to the reduced mass.
mP s

mP s

= P s
me me+
=
me + me+
m2e
=
2me
1
m
=
2 e

(For classical treatment)

(4)

(b) At what point in a bound Kepler orbit is the speed of the satellite
at its minimum?

Figure 1: A particle in bound Kepler Orbit about a more massive object. Point
P is the periapsis and point A is the apsis.
Short Answer: Keplers Second Law (of Planetary Motion) states that the
radius vector of an object caught in an inverse square potential sweeps out
equal areas in equal times (as in Figure 1). In order to adhere to this law, the
object must travel fastest at the closest point of orbit, known as periapsis (think
2

Jeff Kissel November 5, 2006

Classical Mechanics

perigee, perihelion, or periastron), and slowest at the farthest point, known as


apsis (apogee, aphelion, apasteron, etc.). In other words, point A in Figure 1 is
the point at which the satellites speed in minimum.
Thorough Answer: The Lagrangian for a particle in a Keplerian orbit (in
physics spherical-polar coordinates; is the angle from the z axis, is that from
the +
x axis) is
L=

1
k
m(r 2 + r2 2 + r2 sin 2 ) +
2
r

(5)

Because Eq. 5 is explicitly independent of and we know p and p are


conserved quantities. The later fact restricts motion to a plane, i.e. we can
ignore the third term in the Lagrangian. The former yields Keplers Second
Law:
k
1
m(r 2 + r2 2 ) +
2
r
0


7
d L
L
 = 0

dt



d 1 2
= 0
mr
dt 2
1
(6)
r2 = constant
2
The factor 12 is inserted because 12 r2 is the areal velocity the triangular area
swept out by the radius vector per unit time.
L =

Figure 2: The area swept out but the radius vector in a time dt.
The interpretation follows from Figure 2 above, the differential area swept
out in a time dt being
dA

=
3


1
r r d
2

(7)

Jeff Kissel November 5, 2006

Classical Mechanics

dA
1 2 d
=
r
(8)
dt
2 dt
which from Eq. 6 we know is constant.
So, for a particle in Keplerian orbit to sweep out the same area per unit time,
the object must travel fastest at the closest point of orbit, known as periapsis
(think perigee, perihelion, periastron, etc.), and slowest at the farthest point,
known as apsis (apogee, aphelion, apasteron, etc.). In other words, point A in
Figure 1 is the point at which the satellites speed in minimum.

(c) If you have two spheres of the same mass and radius, describe a
non-destructive test by which you could distinguish between the one
that is solid and the one that is hollow.
Short Answer: If one rolls both spheres down a ramp, the solid sphere will
reach the bottom faster. The solid sphere has a smaller moment of rotational
inertia, i.e. less energy is put into rotating the solid sphere so more energy can
by used in linear velocity of the center of mass. Thus, the solid sphere will travel
faster down the ramp.
Thorough answer: The moments of inertia for a solid and hollow sphere
(abbreviated as SS and HS, respectively) are
2
2
M R2
IHS = M R2
5
3
Remember v = R, so each spheres kinetic energy is
ISS =

TSS

=
=
=

TSS

1
1
2
M vCoM,SS
+ ISS 2
2
2


vCoM,SS 2
1 2
1
2
2
M vCoM,SS
+
M
R
2
2 5
R
1
1
2
2
M vCoM,SS
+ M vCoM,SS
2
5
7
2
M vCoM,SS
10

(9)

(10)

1
1
2
M vCoM,HS
+ IHS 2
2
2


1 2
vCoM,HS 2
1
2
2
M vCoM,HS
+
M
R
=
2
2 3
R
1
1
2
2
=
M vCoM,HS
+ M vCoM,HS
2
3
5
2
M vCoM,HS
(11)
THS =
6
Energy is conserved in this system, so the gravitational potential at the top
of the ramp will equal the kinetic energy at the bottom, so
THS

Vi = Tf
4

Jeff Kissel November 5, 2006

Classical Mechanics

7
2
M vCoM,SS
10
10
2
vCoM,SS
=
gh
7

M gh =

M gh =

2
vCoM,HS

5
2
M vCoM,HS
6
6
=
gh
5

(12)

which means the solid sphere will arrive first, since it is traveling at a faster
linear velocity (10/7 > 6/5) down the ramp.
(d) Which way is a particle moving from east to west in the southern
hemisphere deflected by the Coriolis force arising from the Earths
rotation?

Figure 3: A particle moves from East to West in the Southern hemisphere, along
the vector ~r.
The Coriolis Force is defined as
FCor

2m(~ ~vr )

(13)

Thus, as in Figure 3, if ~vr points west, is perpendicular to Earth, then


2m(~
~vr ) points South.
(Theres not much to this problem, so I wont bother with a thorough
answer.If youre wondering where the definition of the Coriolis force comes from,
check out Goldstein pp 174 - 180.)
(e) In a bound Keplerian orbit, what is the ratio of the average kinetic
energy to the average potential energy?
Short Answer: The only force experienced by a particle in a bound Keplerian
orbit is the central, gravitational inverse square force (a power law of order
n = 2). This gravitational force is derivable from a scalar is derivable from
a scalar potential V . Any system satisfying these conditions follows the Virial
Theorem: the average kinetic energy is twice that of the average potential, or
1
T = V
2
5

(14)

Jeff Kissel November 5, 2006

Classical Mechanics

Thorough Answer: An useful scalar quantity known as the Virial, or the


Virial of Clausius, is defined as
X
~pi ~ri
(15)
G(t) =
i

where the summation is over all particles in a given system. Newtons equations
for each particle in the system are
Fi =

d~
pi
dt

(16)

which gives us the inspiration to take the full time derivative of the Virial,
dG(t)
dt

=
=

X d~
X
d~ri
pi
~pi
~ri +
dt
dt
i
i
X
X
mi~r i ~r i
F~i ~ri +
i

F~i ~ri +

F~i ~ri + 2T

mri2

dG(t)
dt

(17)

Since were interested in the (time) average(d) kinetic energy, T , well integrate Eq. 17 over time, from 0 to some time and normalize by the same time
,
Z
X
1 dG(t)

dt =
F~i ~ri + 2T
dt
0

i
X
1
(18)
F~i ~ri + 2T
(G( ) G(0)) =

i
If we chose to be the period and motion is cyclic (as in a Keplerian orbit),
the left-hand side vanishes (G( ) G(0) = 0) leaving
2T

F~i ~ri

1X ~
Fi ~ri
=
2 i

(19)

which is the general form of what is known as the Virial Theorem.


If the forces are derivable from a potential, (as in a Keplerian orbit) i.e.
~ then the theorem becomes
F~ = V
T

1X ~
V ~ri
2 i
6

(20)

Jeff Kissel November 5, 2006

Classical Mechanics

and for a single particle moving under a central force (as in a particle in a
Keplerian orbit), this reduces to
T

1 V
r
2 r

(21)

If V is a power law (as in a Keplerian orbit), where the exponent is chosen


so that the force law goes as rn , then
V

arn+1

V
r
r

(n + 1)V

1
(n + 1)V
2

(22)

which means for a Keplerian orbit (whose force follows and inverse square law,
n = 2), the average kinetic energy and average potential have the following
constant of proportionality
T

1
V
2

(23)

(f ) Write down the Hamiltonian for a free particle in three-dimensional


spherical-polar coordinates.
Short Answer: In (physics) spherical-polar coordinates (where is the angle
from the z axis, and is the angle from the x
axis in the xy plane),
H=

p2
p2
p2r
+
+
2m 2mr2
2mr2 sin2

(24)

Thorough Answer: The Lagrangian in free space (in spherical coordinates)


is
L

1
m(r 2 + r2 2 + r2 sin2 2 )
2

(25)

and Hamiltonian can be found from (and is in fact defined as) a Legendre
transformation between qi and pi ,
X
L
qi pi L where pi
H
(26)

qi
i
So, turning the crank on Eq. 25, and plugging back into Eq. 26,
L
pr
= mr
r =
r
m
L
p
p
=
= mr2
2

mr

p
L
= mr2 sin2 =
p
2

mr sin2

pr

Jeff Kissel November 5, 2006

Classical Mechanics

the Lagrangian as a function of qi and pi is then


L


2 !
 p 2
 p 2
1
p

r
2
2
2
+r
+ r sin
m
2
m
mr2
mr2 sin2
2

p
p2r
p2
+
+
2m 2mr2
2mr2 sin2

and the Hamiltonian is




 p 
p 
p

r
pr +
p
p
+
H =

m
mr2
mr2 sin2
p2
p2
p2
r

2m
2mr2
2mr2 sin2
2
2
2
p
p2
p
p2
pr
p2r

+ 2 +

2
m
mr
mr2 sin 2m 2mr2
2mr2 sin2
p2
p2
p2r
+
+
H =
2m 2mr2
2mr2 sin2

(27)

(28)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
November 15, 2006
A pendulum, free to swing in a vertical plane, has its point of suspension on the rim of a hoop rotating with constant angular speed .
Write the Lagrangian and the resulting equations of motion in some
suitable coordinates for this system.

Figure 1: A pendulum suspended from a hoop rotating with frequency .


Assuming the hoop to be massless, one can solve this problem in a similar
manner to that of a double pendulum whose first bob has zero mass. If the hoop
does has some mass (i.e. it is the canonical thin hoop), there will simply be
a constant kinetic energy term, 21 I 2 , that is coordinate independent and thus
will not effect the equations of motion. In other words, the center of mass of
the hoop is motionless, and its mass is balanced out at the (assumed perfect)
circular edges, such that it rotation is only apparent because of the attached
pendulum. So we proceed assuming the double pendulum scenario.
The components of position and velocity vectors for each bob are
x1 = a sin 1
y1 = a cos 1

x 1 = a cos 1
y 1 = a sin 1

(1)
(2)

Jeff Kissel November 15, 2006

x2 = a sin 1 + sin 1
y2 = a cos 1 + cos 2

Classical Mechanics

x 2 = a cos 1 + 2 cos 2
y 2 = a sin 1 2 sin 2

(3)
(4)

where Ive used 1 = , which is constant. Using these, we can arrive at the
potential and kinetic energies,
V

m
~g 
~r

| 1{z }1

m2~g ~r2

ignored for massless hoop

m2 g(a cos 1 + cos 2 )

(5)

( 1
1((
1
(
2 ((
m
(x(
+ y 12 )(
+ Ih 2 + m2 (x 22 + y 22 )
1(
1
(
(
2
{z
} 2
|2

ignored for massless hoop



1
m2 (a cos 1 + 2 cos 2 )2 + (a cos 1 2 cos 2 )2
2
(a2 2 cos2 1 + 2 22 cos2 2 + 2a2 cos 1 cos 2 )

=
2T /m2

+ (a2 2 sin2 1 + 2 2 sin2 2 + 2a2 sin 1 sin 2 )


a2 2 (sin2 1 + cos2 1 ) + 2 22 (sin2 2 + cos2 2 )
+ 2a2(cos 1 cos 2 + sin 1 sin 2 )


cos cos + sin sin = cos( )
sin2 + cos2
=
1
a2 2 + 2 22 + 2a2 cos(1 2 )


1
m2 a2 2 + 2 22 + 2a2 cos(1 2 )
2

2T /m2

(6)

Finally, from Eqs. 5 and 6, the Lagrangian for this hoop-pendulum system is
L
L

= T V
1
1
=
m2 a2 2 + m2 2 22 + m2 a2 cos(1 2 )
2
2
+ m2 ga cos 1 + m2 g cos 2

(7)

from which we can arrive at the equations of motion. For 1 , because 1 =


and = 0,


L
d L

0 =
dt
1

d 
=
m2 a2 + m2 a2 cos (1 2 )
dt


m2 a2 sin (1 2 )(1) m2 ga sin 1
=

h
i
0 + m2 a2 cos (1 2 ) m2 a2 sin (1 2 )( 2 )
+ m2 a2 sin (1 2 ) + m2 ga sin 1
2

Jeff Kissel November 15, 2006

Classical Mechanics

(
(((
m2(
a
2 (
sin((1 2 ) m2 a22 sin (1 2 )
m2 a2 cos (1 2 ) (
((
(
(((
m2(
a
2 (
sin((1 2 ) + m2 ga sin 1
+ (
((
m2
a 2 cos (1 2 ) = 
m2
a 22 sin (1 2 ) 
m2
a g sin 1

2

22 sin (1 2 ) g sin 1
cos (1 2 )

(8)

Equivalently, for 2 ,
0

=
=



L
d L

dt 2
2

d 
m2 2 2 + m2 a cos (1 2 )
dt


m2 a2 sin (1 2 )(1) m2 g sin 2

= m2 2 2 m2 a sin (1 2 )( 2 )
m2 a2 sin (1 2 ) + m2 g sin 2

m
2 2

2

(
(((
m2(
a
2 (
sin((1 2 )
= m2 2 2 m2 a 2 sin (1 2 ) + (
((
(
(((
m2(
a
2 (
sin((1 2 ) + m2 g sin 2
(
((
2
= 
m
m
2 a sin (1 2 ) 
2 g sin 2
= a 2 sin (1 2 ) g sin 2

(9)

Eqs. 8 and 9 are alternate versions the same the equation of motion of the
system. This means we can set Eqs. 8 and 9 to find an expression for 2 in
terms of , 1 , and 2 . Abbreviating 1 2 as
a 2 sin g sin 2

22 sin
cos

22

22 sin g sin 1

cos
cos
g sin 1
a 2 sin g sin 2
cos
g sin 2 cos
g sin 1
a 2 cos
sin
sin
1

2
g sin 1 a sin cos g sin 2 cos 2
sin

 21
a 2
g sin 2
g sin 1
csc
cos
cot

(10)

Remember, thought = 1 is constant, both 1 and 2 are still functions of


time, so this does not necessarily mean that 2 is a constant.
3

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
Consider a system of N non-interacting particles in which the energy
of each particle can assume two and only two distinct values: 0 and
. The total energy of the system is E.
(a) Find the entropy, S of the system as a function of E and N.
Determine S in the thermodynamic limit where N 1 and E .
In this two level system, the number of particles with energy , n is the
the total amount of energy E, dispersed among each particle. The number of
particles with no energy, n0 is the number of particles with energy subtracted
from the total number N . Summing these statements in equations:
E

n0

= N

(1)
E

(2)

With these numbers we can establish how the system behaves. This two level
energy system will follow a binomial distribution, such that the number of microstates in the system is


N
(N, n ) =
n
N!
=
n ! n0 !
N!


(E, N ) =
(3)
E
E
!
N
!

where Ive substituted in Eqs. 1 and 2. Now that we have the number of
microstates, the entropy is
S
S
kB

kB ln( (E, N ) )




E
E
= ln(N !) ln
! ln (N ) !

(4)

Jeff Kissel October 11, 2006

Statistical Mechanics

where Ive used the properties of the natural logarithm to break up the multiplication and division present in Eq. 3. From here, we can use Sterlings
approximation (ln(x!) = x ln(x) x) to simplify the expression further,
 
E
S
E
E

+ 
= N ln(N ) 
N
ln
kB





E
E
E
(N ) ln N
+ (N )



 

E
E
E
E
(N ) ln N
ln
= N ln(N )



 

E
E
E
E
S(E, N ) = kB N ln(N )
(N ) ln N
(5)
ln

This formula is already in the thermodynamic limit that N 1, because


we used Sterlings approximation, and if we take E , i.e. E/ 1,
then as long as E/ < N , Eq. 5 holds (note the opposite case is unphysical).
If N E/, then one can ignore the second term, and in the third term,
(N E/) N , so the entropy approximately zero.
(b) Find the temperature as function of E.
One can find the temperature from the entropy as it is defined,
1

S
(6)
T
E

1

= kB
ln()
T
E


 

E
E
1

E
E
N ln(N )
(N ) ln N
=
ln
kB T
E

"
 
 1  #
1
E
E
1
E
=
ln
+

"
 

1 
#

E
E
1
E
1
+ N
N

ln N

    
 

1
1
1
1
E
E
= ln
+

+
ln N

 

 
E
E
1
ln N
ln
=




1
E

=
N
ln



1
N
1
=
ln
1
kB T


(7)
T =
kB ln NE 1
2

Jeff Kissel October 11, 2006

Statistical Mechanics

(b) For what range of energy values is the temperature negative?


From Eq. 7, the temperature of this system will be negative when
 


E
E
ln
> ln N

(8)

So, from this expression we can exponentiate both sides and solve for the
energy condition under which the temperatures are negative:
E

2E

>

>

>

N
2

(9)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
X
X
#32

Notes/Changes

Added to question bank.

A crude estimate of the surface temperature of the Earth is to assume that clouds reflect a fraction ( 0.3) of all sunlight, the rest
being absorbed by the Earth and re-radiated. Treating the Sun as a
blackbody at temp TS = 5800 K, find the surface temperature of the
Earth. You may assume the earth is an ideal absorber and that the
rotation of the earth allows it to emit in all directions. The radius of
the Sun is RS = 6.96 105 km and that of the Earth is 6, 400 km.
(R)

The flux of sunlight recieved by the Earth S will be the power radiated
by the sun, PS reduced by the reflection factor , and the distance the Earth is
from the Sun d,
(r)

(1 )
PS
4d2

(1)

From the Stefan-Boltzman Law for blackbody radiators, we know the sun will
radiate a power,
PS

AS TS4

(2)

where AS is the total surface area of the Sun, is the Stefan Boltzmann constant, and TS is the temperature of the Sun. So, Eq. 1 becomes
(r)

=
=

(r)

(1 )
PS
4d2

(1 )
PS 4RS2 TS4
2
4d
R2
(1 ) 2S TS4
d
1

(3)

Jeff Kissel May 18, 2007

Statistical Mechanics

At any given time however, only half of the Earth is facing the sun, so the power
(a)
(r)
absorbed by the Earth PE will be the flux from the sun S , gathered by the
cross-section of the Earth E (i.e. the area of a circle with radius RE ),
(a)

PE
E

(r)

(a)

= E S

(r)

PE

2
= RE
(1 )

RS2
TS4
d2

(4)

Now since were treating the Earth as a perfect (blackbody) radiator, we can
set the power absorbed equal to the power emitted, and solve for temperature.
(e)

PE

2

TE4
4
R
E

TE4
TE
TE

(a)

= PE

RS2
2

= 
R
TS4
E (1 ) 2 
d
R2
= (1 ) S2 TS4
4d

1/4
R2
=
(1 ) S2
TS
4d
r
RS
1/4
= (1 )
TS
2d

(5)

Plugging in numerical values, (note that the distance from the Sun to Earth d
is not given),
r
RS
1/4
TS
TE = (1 )
2d

= 0.3
RS = 6.96 1010 cm

d = 1.496 1013 cm
TS = 5800 K
TE

= 255.874 K

(6)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A simplified model of diffusion consists of a one-dimensional lattice,
with lattice spacing a, in which an impurity makes a random walk
from one lattice site to an adjacent one, making jumps at time intervals t.

Figure 1: An impurity taking a random walk either right (r) or left () with
N = 14 steps, resulting in n = 4 steps to the right.

(a) After N jumps have been made, find the probability that the atom
has moved a distance R = na from its starting point, in the limit of
large N , and n.
As can be seen in the picture, the particle can either move a step right, r or
left, (or a distance a), and since the total number of steps N = r + , and
the net number of steps n = |r | = R / a, the distance traveled is
R

=
=
=
=

na
|r | a

|r (N r)| a
(2r N ) a

(1)

Since the probability that the particle will travel right, PN (r) is just as
probable as it traveling left, PN (), the total probability of it traveling anywhere
is PN (n) = PN (r) + PN () = 2PN (r).

Jeff Kissel October 11, 2006

Statistical Mechanics

Note that from Eq. 1, n = 2r N r = N +n


2 . This is a two-level system,
so it will follow a (not necessarily normalized) binomial distribution, or


N
PN (n) = 2 PN (r) = 2
r
N!
2 N !

= 2
= N +n 
r! !
! N N +n
!
2
2
2 N !
 N n 
(2)
PN (n) =
N +n
!
!
2
2

From here, because were operating with large N and n, well use the natural
logarithm, and take advantage of Sterlings approximation (x! = x ln(x) x)
and the binomial expansion (ln(1 + x) = x 21 + O(x3 )).






N n
N +n
! ln
!
ln PN (n)
= ln(2) + ln(N !) ln
2
2




N +n
N +n
N +n

ln
= ln(2) + N ln(N ) N
2
2
2




N n
N n
N n

ln
2
2
2


N +n
N +n
ln
= ln(2) + N ln(N )
2
2


N n
N n

ln
2
2
N +n N n
N+
+
2
2



n
1
1
N 1+
= ln(2) + ln() + N ln(N ) (N + n) ln
2
2
N



n
1
1
(N n) ln
N 1

2
2
N
1
(
(
(+ N + 

(N
+
2N(+
n
n)
((
2
= ln(2) + ln() + N ln(N )
h

1
n i
(N + n) ln(2) + ln(N ) + ln 1 +
2 h
N

n i
ln(2) + ln(N ) + ln 1
N
N
:


 
1
1(N n) ln(2) + ln()
+
=
1 + (N+n)
2
2



((((
1
(1((N
(
(
+ N ((N
+
n)

n)
ln(N )
(
(
2
((( 2


n 1
n
1
(N n) ln 1
(N + n) ln 1 +
2
N
2
N
2

Jeff Kissel October 11, 2006

Statistical Mechanics

(1 + N ) ln(2) + ln()
 4 

n3
n2
1
n
n
+
+
O

(N + n)
2
3
2
N
2N
3N
4N 4
 4 

1
n3
n
n2
n
+ (N n)

2
N
2N 2
3N 3
4N 4
= (1 + N ) ln(2) + ln()

 4 
1
n2
n2
n3
n3
n

n
+
+ O
+

2
2
2
2N
3N
N
2N
4N 3

 4 
2
2
3
3
n
n
n
n
n
1
n
+
+ O

2
2
2
2N
3N
N
2N
4N 3
1

= (1 + N ) ln(2) + ln()
n
n)
(
2
 2

 3 
3
1  n3 n
n
n
n2
1

O

+
+


2
2
2 2N
2N
2
6N
6N
4N 4
 3 
2
n
n
= (1 + N ) ln(2) + ln()
+O
2N
4N 4

 3 




2
n
n
ln PN (n)
= ln 2(N +1)
+ O  4
2N  4N
 3 
n
where Ive dropped the terms O 4N
because N is huge. Phew! We can then
4
exponentiate both sides to get,
PN (n)


R = na
PN (R)

n2

2(N +1) e 2N

n2 = R2 /a2
=

R2

2(N +1) e 2N a2

(3)

Normalizing to one will yield the constant ,


Z
1 =
PN (R) dR
0
Z
R2
= 2(N +1)
e 2N a2
0

let x =
=
1

dR
dx =
a 2N dx = dR
a 2N
Z

2
ex dx
2(N +1) a 2N
0

 a 2N
2(N +1)
2
1

N
a2
2N
R

a 2N

(4)

Jeff Kissel October 11, 2006

Statistical Mechanics

Thus, for the limit in which N and n are large, the probability of an impurity
has moved a distance R is
PN (R) =
=
PN (R) =

R2
1
(
N+1) 2N
a2
2
e

N
2N
a
2
R2
2

e 2N a2
a 2N

 21
R2
2
e 2N a2
2
N a

(5)

(b) The diffusion coefficient is defined by the differential equation


n
2n
=D
t
x2

(6)

where n is the impurity concentration. Find an expression for D in


the model described above.
The number of steps that impurity takes, N should be equivalent to the
total time traveled, t divided by the time it takes to travel the distance of one
lattice structure, . Thus the probability (Eq. 5) as a function of distance, x
and time, t should be
PN (x, t)

2
a2 t

 21

x2

e 2ta2

(7)

The impurity concentration is a function of the probability of each of the


M particles in the system. In fact, it would be the sum over all the particles
probability that over a time t theyve moved a distance x (from some starting
point x0 ), or
n(x, t)

M
X
i=1

PN (xi , t)

(8)

If this is the case, then we can argue that Eq. 6 must be satisfied for
PN (x, t) in place of n(x, t). From this fact we can get a value for D. Before
we do that, lets clean up PN (x, t) a little bit to get it ready for some serious
differenchumacationating.
PN (x, t)

let

PN (x, t)

2
a2
2
a2

 12
 12

and

= t 2 ex

t 2 e 2a2 x

2 1

2 1

2a2
(9)

Jeff Kissel October 11, 2006

Statistical Mechanics

Now, we can sub this into Eq. 6,


PN
=
t


2 1
1

=
t 2 ex
|t
{z
}
I
I =

=
=
II =
=
=
=
=
=
I =

2



2 x
t1


(t 
2x )
e
=
+
5
2 t2
D

=
=

2 PN
x2
2  1 x2 t1 
D
t 2 e
2
| x
{z
}
II

2 1
2 1
1
3
1
x2 t2
t 2 ex t + t 2 ex t
2
2 1
2 1
ex t
x2 ex t

+
3
5
2t 2
t2
2 1
(t 2x2 )ex t

5
2t 2


2 1
1

D
2xt1
t 2 ex t
x
!
2 1

2xex t
D

3
x
t2
!
2 1
2 1 
2ex t
2xex t
2x
D

3
3
t
t2
t2
!
2 1
2 1
2ex t
42 x2 ex t
D

3
5
t2
t2
!
2 1
2 1
2tex t + 42 x2 ex t
D
5
t2
D

D
II
+D

2(t 2x2 )ex


t

2 1

5
2


2

1

2 x
2
(t 

2x
)
e t
5

t2

1
4
1 2a2
4
a2
2

(10)

Huh! After all that work, the diffusion coefficient is gratifyingly elegant, depending only on the size of the lattice structure!

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
Show that

T V 2
T
where CP and CV are the heat capacities at constant pressure and constant volume, respectively, is the coefficient of thermal expansion,
and T is the isothermal compressibility.
CP = CV +

The important quantities dealt with in this problem defined as




S
CV T
T V


S
CP T
T P


V
1

V
T P


1 V
T
V P T

(1)
(2)
(3)
(4)

where the subscript on the parenthesis indicates what is being held constant in
the partial derivative. Note also that it is assumed that the number of particles
in the system N is constant, so that dN = 0 and the subscript N is suppressed.
We start by expanding the full differential entropy into partial differentials
with respect to T and V ,




S
S
dS =
dT +
dV
T V
V T
Plugging in Eq. 1, this becomes
CV
dS =
dT +
T

S
V

dV

From here we can expand dV in a similar manner to dS,


 





V
V
S
CV
dT +
dP
dT +
dS =
T
V T
T P
P T
1

Jeff Kissel October 11, 2006




CV
+
T

S
V

Statistical Mechanics
 
T

V
T

 

dT +

 

S
V

V
P

dP

(5)

With the goal to bring CP into the expression we take the pressure to be
constant so the dP = 0 and the second term in Eq. 5 vanishes to get
 
 


V
CV
S
T
(S)P =
+
T
V T T P


 


S
CV
V
S
=
+
T P
T
V T T P
where using Eq. 2 yields
CP
T

CP

=
=

 

V
S
V T T P

 

S
V
CV + T
V T T P


S
(V )
CV + T
V T
CV
+
T

(6)

by applying a rearranged Eq. 3.


Now, a useful Maxwell relation is that




S
P
=
V T
T V

(7)

So Eq. 6 is then
CP

CV + T V

P
T

Finally, if we pull apart the partial derivative


tiply by one, well find the desired result:

 
P
CP = CV + T V V
V
|
{z T}
1
T

CP

CV +

T V 2
T

using Eqs. 3 and 4.

(8)

using the chain rule and mul1


V
|

V
T
{z

 
P

}
(9)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
November 6, 2006
Lead has a molar mass of M = 207.2 g mole1 . At 25 C and 1 atm of
pressure, it has an isothermal bulk modulus T of 1.6 1010 P a, a mass
density = 11.4 g cm3 , and a coefficient of thermal expansion, of 87
106 C 1 . Its specific heat at constant pressure CP = 128 J kg 1 C 1 .
(a) How big is the difference between CP and its constant volume
specific heat CV ?
One can prove that, (see Classical S06-#23, F06-#34 for detailed proof)
CP CV =

T V 2
T

(1)

and the only thing in the problem were not given is the volume, for one mole
is just V = M/. So
CV CV

= 8.7 107 K 1

T = 298 K
M =
CV CV

T M2
T

(2)

= 1.14 104 kg m3

T = 1.6 1010 P a
2.072 105 kg mole1

2.56224 1019 J kg 1 K 1

(3)

Apparently really small!


(b) The Law of Dulong and Petit states that the heat capacity of any
solid at room temperature arises from the vibrations of the atoms
(3N degrees of freedom), which can be calculated by treating the
vibrations as a set of 3N classical harmonic oscillators. Does the Law
of Dulong and Petit describe CP or CV ? What would you predict the
heat capacity of lead to be if this law is correct?

Jeff Kissel November 6, 2006

Statistical Mechanics

The Law of Dulong and Petit (describing the heat capacity at constant
volume, CV ) is the formal name for the following identity,
CV = 3R

(4)

for a solid, where R is the gas constant equivalent to 8.314472 J K 1 mole1 .


This can be shown as follows:
For classical harmonic oscillators, whose only degrees of freedom, f are
quadratic, we can use the equipartition theorem which states that
E=

1
nf RT
2

where n is the number of moles in the system. Yet CV is defined as




E
CV
T V,N
1
CV =
f nR
2

(5)

(6)

So for one mole of a solid, for which there are


(3 dimensions) (1kinetic f + 1 potential f ) = 6
degrees of freedom,
CV = 3R

(7)

Thus, by the power of Dulong and Petit, I predict the heat capacity at
constant volume to be
CV

=
=

3(8.314472 J K 1 mole1 )
24.9434 J g 1 K 1

(8)

(c) Find the Debye temperature of lead. How does the specific heat
of lead vary with temperature for temperatures well below the Debye
temperature?
For the first part, Ill just explicitly state the definition of the Debye Temperature, TD and everything one needs to know. For the second part, Ill derive
the Debye Temperature from scratch, and then take the low temperature limit
(i.e. take whats important to the problem from pgs 308-311 of Schroeder, and
added some steps he didnt do).
So, the Debye Temperature is defined as,
1

hcs 6N 3
TD
(9)
2kB V
where cs is the speed of sound in a solid. Converting to what weve been given
in the problem,
r
N
NA
T
=
cs

V
M
2

Jeff Kissel November 6, 2006

Statistical Mechanics

TD

h
2kB

h = 6.626 1034 J s

 12 

6NA
M

 31

(10)

kB = 1.381 1023 J K 1

T = 1.6 1010 P a
M = 2.072 105 kg mole1

= 1.14 104 kg m3
NA = 6.02 1023 particles mole1

TD

113.242 K

(11)

Aside from a few subtle differences such as slower speeds, the addition of
longitudinal polarization, and a lower limit on wavelength, sounds waves in a
solid are quite similar to light waves. Thus, in a solid, each mode of simple
harmonic oscillation has a set of equally spaced energy levels with unit energy
equal to

hf =

hcs n
2L

hcs

(12)

where h is Plancks constant, cs is the speed of sound, L is the length of the


crystal, and n = |~n| is the magnitude of the vector in n-space signifying the
shape of the wave. At a temperature, T the number of units of energy the solid
contains follows the Planck Distribution,
n
P l

1
e 1

(13)

(note n
P l is entirely different from n from Eq. 12). To find the total thermal
energy, E of the crystal, we add up the energies of all dimensions (or allowed
modes),
XXX
E=3
n
P l ()
(14)
nx

ny

nz

where the factor of 3 comes from that a phonon has three possible polarization
states (one longitudinal and two transverse). From here we would normally (i.e.
for electromagnetic waves) convert to integrals, but we first must worry about
the limits of integration.
As weve said, in a crystal the atomic spacing puts a lower limit on the
wavelength. Consider the following one-dimensional lattice of atoms (Figure 1).
Each mode of oscillation has its own distinct shape, with number of bumps
equal to n. Because each bump has to contain at least one atom, n cannot
exceed the number of atoms in a row.
If a three-dimensional
crystal is a perfect cube, then the number of atoms

along any direction in 3 N , so each sum in Eq. 14 should go from 1 to 3 N .


If the crystal isnt a perfect cube, then neither is the corresponding volume in
3

Jeff Kissel November 6, 2006

Statistical Mechanics

Figure 1: Modes of oscillation of a row of atoms in a crystal lattice.


n-space. However, the sum will still run over a region of n-space whose total
volume is still n.
Though the sums (integrals) in Eq. 14 depends on nx , ny , and nz in a nontrivial way, the function depends on the magnitude of ~n in a simpler way; it
doesnt depend on angles at all. Thus, the Debye model approximates the cube
with the first octant of a sphere, with total volume N (preserving the degrees
of freedom) and radius


1 4n3max
N =
8
3
3
nmax
=
6
1

6N 3
nmax =
(15)

Figure 2: The
sum in in Eq. 14 is technically over a cube in n-space, with a
side length of 3 N . Debye makes the approximation that the cube is a sphere
of radius nmax
4

Jeff Kissel November 6, 2006

Statistical Mechanics

Remarkably, it is experimentally confirmed that this approximation is exact


in both the high- and low- temperature limits, and intermediate temperatures,
though not exact are surprisingly good. So, now in Eq. 14 we can convert sums
to integrals, then move to spherical coordinates,
XXX
E = 3
n
P l ()
nx

ny

nz

nmax

n2 sin dn d d
/kB T 1
e
0
0
0
 Z 2 Z 2
hcs n 

, =
sin dd =
2
2L
0
0
Z nmax
3
3
n
hcs
dn
2 0
2L ehcs n/2LkB T 1

n2

(16)

which looks nasty at first, but if we choose our variables such that
x
dx =

hcs
dn
2LkB T
xmax

hcs
n
2LkB T

kB T dx =
=
=

hcs
dn
2L

hcs nmax
2LkB T
1 hcs
T 2LkB
TD
T

6N
V

 31

(17)
where Ive finally introduced the Debye temperature,
1

hcs 6N 3
TD
2kB V

(18)

then E becomes,
E

=
=

=
=

nmax

hcs
3 3
1
n hc n/2Lk T
dn
s
B
2
2L
e

1
0

3
Z xmax
1
3 2LkB T
x
k T dx
x1 B
2
hc
e
s
0
3

  T 3

2LkB T
=
TD
hcs
6N
3

Z TD /T
x3
T
3 6N
kB T
dx
2
TD
ex 1
0
Z TD /T
x3
9N kB T 4
dx
3
x
TD
e 1
0
Z

(19)

Jeff Kissel November 6, 2006

Statistical Mechanics

At which point you plug the integral into your favorite computer program to
find E for any temperature if you like. BET, the question ask for limits. First,
we can check the Law of Dulong and Petit by taking the limit when T TD ,
the upper limit of the integral is much less than 1, so x is always very small and
we can approximate ex 1 + x in the denominator,
E

9N kB T 4
3
TD

TD /T

TD /T

9N kB T
3
TD

x3
dx
1 + x 1
x2 dx

CV

3
 kB T 4 TD
9N

3
3
3T

TD
3N kB T
(when T TD )



E
3N kB 3nR
T V,N

(20)
(21)

When T TD , as the question asks, the upper limit on the integral is so


large that by the time we get to it, the integrand is dead (due to the ex in
the denominator). So, we can just replace the upper limit with infinity; the
extra modes were adding dont contribute anyways. In this approximation, the
integral evaluates to 4 /15, so the total energy is
E

CV

3 4 N kB T 4
3
5
TD

3
12 4 N kB
T
5
TD

which is the second part of the question.

(when T TD )

(22)
(23)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 12, 2006
Consider a monatomic ideal gas of mass density, at temperature,
T whose atoms have mass m. The number of atoms with velocities
~v in the velocity space volume element d3~v is given by the MaxwellBoltzmann distribution
!

 23
m vx2 + vy2 + vz2

m2
3
(1)
n(~v ) d ~v =
exp
m 2kB T
2mkB T
The equation given is wrong. I shall prove otherwise.
For the canonical ensemble, partition function for N indistinguishible particles in an ideal gas is
Z

VN
N!

2mkB T
h2

 32

(2)

So for a single (N = 1) free (E = p2 /2m) particle in an ideal gas (where


= 1/kB T ), the phase space probability distribution is
1 eE(~p,~q) 3 3
d p~ d ~q
h3
Z

2
1 e 2m p~
3
3
=
 23 d p~ d ~q

h3
V 2m
h2
Z

P (1) (~
p) d3 p~ =
P(~
p, ~q) d3 ~q d3 p~

P (1) (~
p, ~
q ) d3 ~p d3 ~q =

P (1) (~
p) d3 p~ =

1
h3

P (1) (~
p) d3 p~ =

e 2m p~

 32 V

2m

2m
h2

 32

e 2m

p
~2

d3 p~

(3)

Jeff Kissel October 12, 2006

Statistical Mechanics

Yet, both this momentum distribution, and some velocity distribution should
be normalized to unity, so
Z
Z
(1)
3

1 =
P (~
p)d p~ =
n(~v ) d3~v

n(~v ) d3~v

n(~v ) d3~v

= P (1) (~
p) d3 p~

 32

2 2

e 2m m ~v m3 d3~v
=
2m
3

m 2 1 m~v2 3
=
e 2
d ~v
2

(4)

which differs from the equation given in that there is no /m factor out front,
and the argument of the exponential should have either have an extra m in the
numerator, or no m in the denominator.
Hok, now lets get onto the problem.
(a) What is the average velocity h~v i?
The average velocity is defined as
Z
h~v i =
v n(v) d3 v

(5)

However, n(~v ) is a gaussian distribution, i.e. even over the interval (, ),


and v, begin linear, is odd. The integration of the product of an even and odd
function from to will equal 0. Thus,
h~v i = 0

(6)

(b) Derive the distribution of speeds P (v) dv.


This can be done by converting the velocity distribution (Eq. 4) into spherical coordinates, and integrating over the angular terms.
P (v) dv

n(~v ) dvx dvy dvz

n(v) v 2 sin dvdd


Z 2 Z

=
sin dd v 2 n(v) dv

P (v) dv

= 4v 2 n(v)dv
3

m 2 1
2
e 2
= 4v
2

where Ive exploited the fact that ~v 2 = v v = v 2 .

mv 2

dv

(7)

Jeff Kissel October 12, 2006

Statistical Mechanics

(c) What is the most probable speed, v?


The most probable speed will be the maximum of the speed distribution, so
Ill denote v as vmax . So, well maximize Eq. 7 with respect to v.


P (v)
v

= 4

v = vmax

m(vmax ) e

21 mv 2

(vmax )2
v = vmax

m
2
1

2v e 2

 23

 2 1
v e 2
v

mv 2

+ v e 2

= 2(vmax )e

12 mv 2

= 2(vmax )e
2
=
m
r
2kB T
=
m

12 mv 2

mv 2

mv 2

i
(mv)

v = vmax

m(vmax ) e

12 mv 2

(8)

(d) Obtain expressions for vavg , root mean square speed, vrms , and v,
then rank them in increasing order.
q
We can obtain the root mean square velocity, vrms = (v 2 ) by doing the
integral
Z
(v 2 ) =
v 2 P (v) dv
(9)
0

but we can get it quickly from the Equipartition theorem (E = 21 f kB T ) for a


monatomic gas (where the degrees of freedom, f is 3), since the energy is only
kinetic for a ideal (free) particle,
E

3
kb T
2

1
m (v 2 )
2
3kB T
rm
3kB T
m

(v 2 ) =
vrms

1
mv 2
2

(10)

For the average velocity, vavg = v, well actually have to do the integral.
Z
v =
v P (v) dv
0

m
2

 23 Z

v 3 e 2

mv 2

dv

Jeff Kissel October 12, 2006

=
=

Statistical Mechanics

1
let = m
2
Z
3
2
4 2

v 3 ev dv
0
Z

3
4 2
v 2

ve
dv

0

let u = v 2

=
=
=

du = 2v dv
1
(0 0, )
du = v dv
2
 Z

3
1
4 2

eu du
pi 2 0
!
3
eu
2 2




0


3
2
2
1

=
=
v

2 2

2

1
2 2
2

m
r
8kB T
m

(11)

Thus, since Eq. 8, 10, and 11 all three have the factor
and only differ by 2 < 8/ < 3,

p
1/m in common,

vmax < vavg < vrms

Figure 1: The Maxwell-Boltzmann distribution for T 100 K.

(12)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
October 11, 2006
A certain material is completely specified by its volume V and temperature T . It has an equation of state P = AT 4 , where A is a constant
independent of the volume. The heat capacity at fixed volume is measured to be BV T 3 .
a) From dimensional analysis, B and A have the same units. Show
that B = 12A.
The heat capacity is defined as


S
CV T
T V


S

T V

BV T 3

BV T 2

(1)

Integrating both sides, well get an expression for the entropy,


Z T
Z S
T 2 dT
dS = BV
0

1
BV T 3
3

(2)

However, since wed like to bring A into the equation we have to use the Maxwell
relation




S
P
=
(3)
V T
T V
where the subscript denotes what is being held constant in the derivative. Plugging in our equation of state and Eq. 2,



1

3
AT 4 V
=
BV T
V 3
T
T
1
BT 3 = 4AT 3
3

B = 12A
(4)
1

Jeff Kissel October 11, 2006

Statistical Mechanics

b) Find the entropy of this material as a function of V and T .


Phh, we dun already been up in there. Peep Eq. 2:
S =

1
BV T 3
3

(2)

c) If this material is cooled adiabatically and reversibly from 20K to


10K, by how much does the volume change?
In a reversible adiabatic process, entropy is conserved, and thus is called an
isentropic process. This means we can set up a conservation equation from
Eq. 2,
1
BVi Ti3
3
Vi Ti3

=
=

Vi
Vf

Vi
Vf

Vf

1
BVf Tf3
3
Vf Tf3
 3
Tf
Ti
3

1
(10 K)
=
(20 K)
8
8Vi

So the volume increases eight-fold by doubling the temperature.

(5)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
#27
#38
#38

Notes/Changes

Number Shift
Part (d) removed

Consider an ideal Fermi gas of spin-1/2 particles in a 3-dimensional


box. The number of particles per unit volume is n, the mass of the
particles is m, and the energy of the particles is the usual = p2 /2m.
Assume that the temperature is quite low.
(a) Find formulas for the Fermi energy F , Fermi wavevector kF , and
Fermi temperature TF in term of m, n and constants such as h and
kB .
To find the Fermi energy (the chemical potential of spin 1/2 particles at zero
temperature), we must find the total number of particles N as T approaches
zero.
For any temperature T , one normally finds N via summing over the FermiDirac probability distribution h nj () iF D (because the particles are of spin
1/2),
h nj () iF D

1
e(j ) + 1

(1)

1
where = kT
, j is the kinetic energy of a state j, and is the chemical potential
of the particles. However, at the limit of zero temperature, h nj () iF D decays
to either 0 (when j > ) or 1 (when j < ). The energy, at T = 0, at which
the now-step-function distribution flips from 1 to 0 is the Fermi energy, whose
value is determined (again) by the number of particles.
To perform the calculation of the number of particles, we first find the density
of states: the Jacobian of transformation between summing over energy states
j , and integrating over kinetic energies. First, to account the spin degeneracy

Jeff Kissel May 18, 2007

Statistical Mechanics

of +1/2 and 1/2, we tack on a factor of 2,


X
X
= 2
j

X
j

Z Z
2
=
d3 p~d3 ~q
h3
Z
2
=
d3 p~
V
h3

Z
2V
(4)
p2 dp
=
h3
0

=
2m 1/2
2 p

p
1/2
let =
dp = 12 2m
d

2m 2
2m
p =
Z 
 1
8V
2m
2m1/2 d
=
h3
2
0
Z
8V 1
3/2
(2m)
=
1/2 d
h3 2
0
3/2

Z
Z
2m
1/2

g() d

4V
h2
0
0

(2)

Now we can find the number of particles N , and invert for the Fermi energy
F .
N (T = 0) =

X
j

h nj () iF D

h n() iF D gF D () d
Z
Z F
(0) g() d
(1) g() d +
0

=
N

=
=

4V


2m
h2
3/2

3/2

2m
2 23

2
h
3 F

3/2
3
8
2m
V
F2
2
3
h

2/3
2
13N
h
2m 8 V

2/3
h2 3 N
8m V
4V

2 d

(3)

Jeff Kissel May 18, 2007

Statistical Mechanics

2
h
2m


2/3
2N
3
V

(4)

In this form its nice and quick to pull out the Fermi wave-vector kF , since
the quantum mechanical kinetic energy is
E

h2 2
p2
=
k
2m
2m

(5)

So by direct comparison between Eqs. 4 and 5,


kF

 31

2 N
3
V

(6)

The Fermi terperature TF is similarly quick, as the thermal kinetic energy


is,
E

kB T

(7)

Inverting for temperature and subbing in 4,


TF

2
h
2mkB

2

N 3
3 2
V

(8)

(b) Find the total energy of the gas at zero temperature.


This calculation can be done by summing over possible energy states, weighted
by their probabilities (determined by the Fermi-Dirac Distribution),
X
j h nj (j ) i
hEi
j

hEiT =0

g() d

h n iF D gF D () d
Z
(0) g() d
(1) g() d +

V
2 2

V
2 2

V
5 2


3
V
2m 2 1

2 d
2 2 h2
3 Z

2m 2 F 3
2 d
h2
0
3

2m 2 2 52

5 F
h2
3

2m 2 52
F
h2
3

Jeff Kissel May 18, 2007

Statistical Mechanics

V
Eq. 3: N =
3 2
hEiT =0

2m
h2

 32

3
N F
5

3
2

!
(9)

(c) Show that the heat capacity at low temperature is proportional


to T .
The exact proof of this problem unfortunately gets rather lengthly, but is
known as the Sommerfield expansion.
If I denote the constants in Eq. 2 to be CF , and use < n > as shorthand for
h n() iF D then the ensemble average energy (as seen in part (b)) is
Z
Z
3
hEi =
h n() iF D gF D () d = CF
2 < n > d (10)
0

However, now we need energies at low temperatures (kT F ), not just


at T = 0. Although the integral in Eq. 10 goes from 0 to , i.e. all positive
energies, the most interesting region is near = , where h nj (j ) i falls off
sharply. So, well isolate this region using integration by parts,

hEikT F

let u =
hni
v = 52 5/2
d<n>
du = d d
let dv = 3/2
"

 #

:0 2 Z


d<n>
2 5/2 
5/2

d


<n> +

CF
5 
5 0
d

0

(11)

where the surface term vanishes because at = 0, 5/2 = 0, and at = ,


< n >= (e() + 1)1 = 0. This leaves,


Z
2
d<n>
5/2
hEi =
d

CF

5
d
0




1
d
d<n>
e(j )
=

=
d
d e(j ) + 1
(e(j ) + 1)2
Z
e(j )
2
CF
5/2 ( )
=
5
(e j
+ 1)2
0

dx
=
d
let x = ( )
0 ,

Z
2
ex
hEi =
CF
dx
(12)
5/2 x
5
(e + 1)

Because, the integrand of this expression falls off exponentially when x =


( ) 1, we can now make two approximations:
1. Extend the lower limit of the integral () to (); this makes the
integral symmetric, and its harmless because the integrand is negligible at
negative s because of the shape of the distribution at these temperatures.
4

Jeff Kissel May 18, 2007

Statistical Mechanics

2. Taylor expand 5/2 around = ,



 2 5/2 
 5/2 

1
d ( )
d( )
2
5/2
5/2
+ ( )

+ ( )

+ ...
d
2!
d2

5
1 15
5/2 + ( )3/2 +
( )2 1/2 + ...
2
2 4
5
15
5/2 + (xkT )3/2 + (xkT )2 1/2 + ...
2
8

(13)

So that
hEi

"
Z
2
ex
dx
CF 5/2
x
2
5
(e + 1)
|
{z
}
(I)
Z
ex
5
kT 3/2
dx
x x
+
2
(e + 1)2

|
{z
}
(II)
#
Z
15
ex
2 1/2
2
(kT )
dx +...
+
x
8
(ex + 1)2

|
{z
}
(III)
Z

(I)

ex
dx
x
(e + 1)



d<n>
d

d<n>

"

#
:1
:0




 >

<
n()
> +
<
n()


=
Z

(II)

ex
dx
(ex + 1)

= 1
Z

(14)

ex
x ex
dx
x
+ 1)(e + 1) ex

Z
x
dx
x
(e + 1)(1 + ex )

{z
}
| {z } |

x ex
dx =
x
(e + 1)2
=

(ex

even
bounds

x ex
dx =
(ex + 1)2

odd
function

(15)

Jeff Kissel May 18, 2007

(III)

|
{z }

even
bounds

Statistical Mechanics

x2 ex
dx
(ex + 1)2
| {z }

= 2

v
let dv

x2
2x dx

Z

xm
dx
ex + 1

=
=
=
=

hEi

x2 ex
dx
(ex + 1)2

1
ex +1
ex
(ex +1)2



5 
2
2
CF 5/2 (1) + 
CF
kT 3/2 (0)
5
5
2

 2
2
15

2 1/2
+
+ ...
CF
(kT )
5
8
3
2
2
CF 5/2 +
CF (kT )2 1/2 + ...
5
4

 32
h
2
Eq. 4:
F
= 2m
3 2 N
V2

3
2m

F
=
3 2 N
V

h
2 3


3/2
2m 2
=
3 2 N
F

2
V
h

3

N
V
2m 2

=
3
2
2
3/2

3
2m 2
V
3 N
CF =
=
2 2
2 3/2
h
2
F
!
!
2

2
3 N
3 N
5/2 +
(kT )2 1/2 + ...
5
2 3/2
4 2 3/2
F

=
=

0

Z
>


x2 

2x
 +
2
dx

ex

x
+1 0
e +1
0

Z
x
dx
4
x
e +1
0

1
(1 m) (m + 1) (m + 1)
2
1
4(1 ) (2) (2)
2
 2
 

1
(1)
4
2
6
2

3
2
3

x2 ex
dx
(ex + 1)2

even
function

let u =
du =

3
5/2
1/2
3 2
N 3/2 +
N 3/2 + ...
5
8
F
F
6

(16)

(17)

Jeff Kissel May 18, 2007

hEi

Statistical Mechanics

3
N F
5

5/2

(kT )2
3 2
N
+
8
F

1/2

+ ...

(18)

Performing a similar Taylor expansion process for N (see Schroeder pp 282284) yields the expansion that,

= 1

2
12

kT
F

2

+ ...

(19)

which we can exploit for our purposes,

5/2

1/2

2
1
12

kT
F

2

!5/2

+ ...

(Binomial Expansion: (1 + x)p 1 + px (For x 1))



2
5 2 kT
+ ...
1
2 12 F

2
5 2 kT
1
+ ...
(20)
24
F
!5/2

2
2 kT
1
+ ...
12 F

2
1 2 kT
+ ...
1
2 12 F

2
2 kT
+ ...
(21)
1
24 F

and plug back into the ensemble average energy (Eq. 18),
!

2
5 2 kT
3
hEi =
N F 1
+ ...
5
24
F
!

2
2 kT
(kT )2
3 2
1
N
+ ... + ...
+
8
F
24 F

(22)

and then keeping only terms of order (kT /F )2 or less,


hEi =
=
hEi =



3
3 2
3 2
(kT )2
(kT )2
(kT )2
+
+O
N F
N
N
5
24
F
8
F
3F


3 2
2
(kT )2
(kT )2
(kT )2
3
+
+O
N F
N
N
5
8
F
8
F
3F
2
(kT )2
3
+ ...
N F +
N
5
4
F

(23)

Jeff Kissel May 18, 2007

Statistical Mechanics

Now we can finally get to the question. Remember we are trying to show
that CV T for low temperatures. This is quick attainable from Eq. 23, and
the definition of the heat capacity at constant volume,


hEi
CV
T V,N



3
2
(kT )2
=
+ ...
N F +
N
T 5
4
F
2
2

k
CV =
T T
(24)
N
2
F

Jeff Kissel May 18, 2007

Statistical Mechanics

OLD VERSIONS
(d) If the magnetic moment of the particles is e , show that the paramagnetic susceptibility for low fields in the limit of zero temperature
is given by

3 n2e
2 F

(25)

This part deals with the quantum mechanical property of fermions known as
the Zeeman splitting. When a weak magnetic field is applied to a fermi gas, the
~ |i) and down (anti-parallel to B,
~ |+i) energy levels of
up (parallel to B,
the particles are perturbed such that the energy states of the up particles are
increased, and the down particles are decreased, causing a splitting of what
initially was a single set of quantized energy states. The Zeeman effect is a 1st
order perturbation theory problem in quantum mechanics (Check p245-246 of
Griffiths, or Wikipedias article for more detialed explanations. Merzbachers is
pretty weak sauce, but its on p472-473). Anyways, this is a Stat-Mech problem,
so we really only care about the resulting energies, so Ill only highlight the
details.
The perturbing hamiltonian is
b
H

b B
~
~
e

(26)

b is the applied, weak magnetic field, and


where B
be is the magnetic moment of
the fermions (spin-1/2 particles), given by

~be

b
H

b
e B

b
H

b
~
gJ B S
h

(27)

where gJ is the Lande g-factor, B = (eh)/(2m) is the Bohr magneton, and


Sb is the spin operator. The Lande g-factor is approximately 2 for (spin-1/2)
electrons (See Wikipedias Lande g-factor for details), so I assume gJ = 2.
b (Eq. 26, assuming the field |B|
~ = B applied in the same direction as
Thus, H
e ) becomes,
=

2 B Sb

h


h
b

bi
Si =
2

B
bz B

(28)

where
bz is the Pauli spin matrix, and Ive assumed that the field was applied
in the z-direction.
9

Jeff Kissel May 18, 2007

Statistical Mechanics

The total energy of the anti-parallel (down, |+i, or the positive eigenvalue
of
bz ) state is
T

= + B B

in which, as before, = p2 /(2m) is the kinetic energy of the particles. The


parallel (up, |i, negative eigenvalue of
bz ) energies are
T

= B B

Which means the Fermi-Dirac distribution (1) becomes


h nj () iF D

1
e[(j B B)) + 1

(29)

i.e. is splits into two distributions,


1
h nj ( B B) iF D
2

and

1
h nj ( B B) iF D
2

(30)

where the factors of 1/2 insures the distribution remains normalized. The left
distribution corresponds to the up spins, i.e. those energy levels displaced by
a positive B B and therefore the probability range over which it extends is
from B B < T < . The right distribution are the down spins, displaced by
negative B B, so the range is B B < T < .
From here, we can calculate the net magnetization per unit volume M , the
derivative of which (with respect to B) is . The net magnetization would be
the total number of particles anti-align subtracted from those aligned (in a given
volume), or
M

(N N )
V

(31)

We know how to find the number of particles: integrate the probability distribution corresponding the particles over the correct region. Thus, the magnetization
is then
Z
B
1
M =
h n( + B B) iF D gF D () d
V
B B 2

Z
1

h n( B B) iF D gF D () d
B B 2
Z
B
[h nj ( + B B)i h nj ( B B)i] g() d
=
2V 0
!
f (x)
f (x+h)f (xh)

x
2h
(x)
f (x h) f (x + h) =
2 h fx

Z 
B
h n() i
=

g() d
2 B B
2V

0

Z 
2B B
h n() i
M =
g() d
(32)

0
10

Jeff Kissel May 18, 2007

Statistical Mechanics

At low temperatures (i.e kT F ), the Fermi-Dirac distribution becomes


asymptotically close to a step function with the barrier at the Fermi energy
(as discussed in part (a)). Thus, the derivative of the distribution at low temperatures is well approximated by a delta function about F , or

Z
Z 
h n() i
( F ) d
(33)
d =

lim
T 0 0

0
which makes Eq. 32 a lot prettier,
M

=
=
=

Z
2B B
( F ) g() d
V
0
2B B
g(F )
V
1/2 B
2B CF F
V
!
3 N
Eq. 17: CF =
2 3/2
F
!
3
N
B
1/2
2B
F
2 3/2
V
F

3 2B N
B
2 F V

(34)

Finally, if I convert to the notation the proof desires (N/V n), the magnetic susceptibility is

M
B 

3 2B

nB
B 2 F
3 n 2B
2 F

which is sometimes referred to as the Pauli Susceptibility.

Its cool. Youll remember all of this during the qualifier. I promise.

11

(35)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
#28
#39
#39

Notes/Changes

Number Shift

A collection of N spin-1/2 atoms are fixed in a solid. The atoms do


not interact with each other. The magnetic moment of each atom is
0 . If a magnetic field H is applied to the solid, each atom has an
energy of 0 H.
a) Find the mean energy hEi in a magnetic field.
For a single spin-1/2 particle, there are only two possible energy states, as
stated in the problem:
E

0 H

(1)

where the or up state is aligned with the magnetic field H, and the + or
down state is anti-aligned. The grand canonical partition function for the
particle is then,
X
Zi
eE(s)
=
=

Zi

s
E

+ eE+
e
0 H
e
+ e0 H


1
cosh (e + e )
2
2 cosh (0 H)

(2)

where = kB1T as usual.


For the entire system of N non-interacting indistinguishable particles, the
partition function becomes
Z

1
(Zi )N
N!
1

Jeff Kissel May 18, 2007

Statistical Mechanics

2N
coshN (0 H)
N!

(3)

where taking the partition function to the N th power accounts for all the particles, and dividing by N ! corrects for double counting present because of indistinguishability.
Once we have the partition function of a system, we can find any other
property of the system desired, as long as we remember a few relations. The
relation relevant to the ensemble mean energy hEi is
hEi

1 Z

=
(ln (Z))
Z

(4)

From Eq. 4 we see well need the natural log of Eq. 3, so lets get that out
of the way,
 N

2
N
ln (Z) = ln
cosh (0 H)
N!
 N


2
= ln
+ ln coshN (0 H)
N!
 N
2
+ N ln (cosh (0 H))
(5)
ln (Z) = ln
N!
Happily, the first term will vanish when plugged into Eq. 4, because its independent of , so we need not use Sterlings approximation or anything fancy
like that. So, the ensemble mean energy is
!

 N
2 

+ N ln (cosh (0 H))
ln
hEi =
N !


N
=
sinh (0 H) 0 H
cosh (0 H)
hEi

= 0 N H tanh (0 H)

(6)

b) Find the entropy of this collection.


Again, because we have the partition function of the system, we need only
to remember the relation between that and the Helmholtz Free Energy F ,
F

kT ln (Z)

from which well find the entropy using the following Maxwell relation,


F
S =
T V,N

(7)

(8)

Jeff Kissel May 18, 2007

Statistical Mechanics

So lets plug and chug!



 N
2
coshN (0 H)
F = kT ln
N!

 N


2
N
S =
kT ln
cosh (0 H)
T
N!
"
!#

 N
2N

2
N
ln
cosh (0 H) kT
+ N ln cosh (0 H)
= k ln
N!
T
N!


N

= k ln (Z) kT
sinh (0 H) 0 H
cosh (0 H)
T




1 1
1
1 1

=
=
=
T
T kT
k T2
kT T

N
H
sinh
(
H)
kT
0
 0
= k ln (Z) + 
kT T cosh (0 H)

 N

2
0 N H
S = k ln
coshN (0 H) +
tanh (0 H)
(9)
N!
T
c) The magnetization m of a solid is defined as the net magnetic
moment per unit volume. The average magnetic moment is defined
via hEi = hM iH. For noninteracting moments, the magnetization
typically obeys a Curie Law where m = 0 H/T for vanishingly small
H, find the value of the constant 0 for this problem.
If (as given in the problem statement),
hEi = hM iH

(10)

then the average magnetic moment hM i, (which is effectively equivalent to magnetization m mentioned in the problem for large systems) using Eq. 6 is
hM i =
m

hEi
H
 tanh (0 H)
0 N
H

H

0 N tanh (0 H)

(11)

Now, in the Curie Regime, either for vanishingly small H (i.e. H 1


or for high temperatures (i.e T 1), the argument of the hyperbolic tangent
0 H is also very small. However, in the case where |x| 1,
tanh x

(12)

So Eq. 11, yields whats known as the Curie Constant 0 ,


lim or lim m

H0

0 N

0 H
kT

Jeff Kissel May 18, 2007

Statistical Mechanics

m(T ; H 0) =
0
Doneski.

20 N H
k T
20 N
k

(13)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
#29
#40
#40

Notes/Changes

Number Shift
Part (d) removed.

An ideal non-relativistic gas of N spin-1/2 particles of mass m is


confined to move in a two dimensional area of size A = L2 . Assume
the energy levels are close enough to be described as a continuous
density of states.
(a) Find the density of states per unit area g().
The ensemble average of any observable O is defined as
X
hOi =
Oj Pj

(1)

The density of modes g() is the Jacobian of transformation when one


changes this sum to an integral. This conversion for a 2-D system of fermions
(where we add a factor of 2 to account for the spin degeneracy) is as follows
X
X
= 2
j

=
=
=

Z Z
2
d2 p~d2 ~q
h2
Z
2
A
d2 ~p
h2

Z
2A
(2)
p dp
h3
0

p2
p =
let =
dp =
2m
1

1
2

1/2
2m 1/2
2m
d

Jeff Kissel May 18, 2007

X
j

Statistical Mechanics

Z 

4A

 1 2m1/2
1/2

2m

=
 d
2
h
2
0
Z
4A 1
=
d
2m
h2 2
0
Z
Z
4mA 2

g() d

h2
0
0

(2)

(b) Find the Fermi temperature in terms of the above quantities.


Here, we evaluate Eq. 1 for the total number of particles N , by summing
over the probability distribution of fermions (as per Eq. 1),
1

nF D

e()

+1

(3)

which at the limit of zero temperature splits into two regions (i.e. has values
1 or 0). The non-vanishing portion of the distribution is in the energy regime
0 < < F , (the upper limit is the Fermi Energy, or chemical potential at
zero temperature), because the Pauli Exclusion Principle dictates that fermions
(spin-1/2 particles) cannot occupy the same energy state. Thus instead of all
collapsing to some ground state, they stack up in energy to F when temperatures very small.
The result of the integral will yield N (F ), which we can invert to get F (N )
X
nF D
N =
j

NT =0

=
=

lim
T 0
Z F
0

g() d
+1
Z
(0) g() d
(1) g() d +
e()

(4)

4mA
=
d
h2
0
4mA
N =
F
h2
N h2
(5)
F =
4mA
Finally, we make use of the equipartition theorem which states that every kinetic
degree of freedom gets an energy of 12 kB T . Since there are two degrees of
freedom in this 2-D system,
Z

kB T

So,
TF

f
N h2
=
kB
4mkB A

(6)

Jeff Kissel May 18, 2007

Statistical Mechanics

(c) Find the total kinetic energy of the gas at zero temperature.
Again, well use Eq. 1 for ensemble average energy, hEi. Because we want the
value at T = 0, the same limits apply as in part (b). Ready, steady, integrate.
Z F
g() d
hEiT =0 =
0
Z F
4mA

hEiT =0 =
d
h2
0
1 2 4mA
=

2 F h2
2mA 2
F
(7)
hEiT =0 =
h2
and from Eq. 5,
hEiT =0

=
=
=

hEiT =0


2
2mA
h2
N
h2
4mA
N 2 h4
2mA
2
h
16 2 m2 L4
2 2
N h
8mA
N 2 h2
8mA

(8)

(d) Find the chemical potential at all temperatures.


To find the chemical potential at all temperatures (instead of at zero temperature, which is the Fermi energy), well find the total number of particles
(Eq. 4), as in part (b). However, we can no longer divide the integral into two
regions, so we must solve the integral fully from 0 to . Once we have N (),
we can invert the expression for (N ) as we did for the Fermi energy (which
was the chemical potential at T = 0).
Z
nF D () g() d
N =
Z0
4mA
1
d
=
() + 1
h2
e
0
Z
4mA
1
=
d
() + 1
h2
e
0

dx
=
d

d
=
kT x
let x = ( )

0 ,

Z
Z
4mkT
1
1
4mA
A
kT
dx
=
dx
=
2
x
2
x
h
h
e + 1
e + 1
3

Jeff Kissel May 18, 2007

Statistical Mechanics

1
b
dx = [x ln (ex + 1)]a ;
ex + 1

h2
2mkT

1/2 !

[x ln (ex + 1)]
2


h
i

lim x ln (ex + 1) = x ln (ex ) = x x = 0
x

i
A h
()
=
(0)

()

ln
(e
+
1)
2

= + ln (e + 1)
=

N
N 2
A
e

N 2
A

e
ln (e

= e eln (e

= e

N 2
A

= ln (e
= ln (e

+1)

= e (e + 1) = (1 + e )

N 2
A

N 2
A

= kT ln (e

1)
1)

N 2
A

1)

(9)

Jeff Kissel May 18, 2007

Statistical Mechanics

OLD VERSIONS
(d) Find the pressure exerted by the gas at zero temperature. dF =
pdA.
Im pretty sure that the reason the equation dF = pdA is to just to establish the sign convention (some texts like to define work, W = dF as positive
when leaving the system, i.e. W = P dV , and others like W = P dV , i.e. positive work is done when entering the system. This problem chooses the latter,
but since were dealing with a 2-D system, dV dA).
Anyways, because of the convention chosen, the thermodynamic identity is
dhEi =

T dS P dA + dN

(10)

We assume the number of particles stays constant at hN i, so the third term


vanishes, and the question desires the pressure at T = 0, so were left with a
relation for pressure in terms of what we know from previous parts,


hEi
dhEi = P A|N P =
(11)
A N
So from Eq. 8,
P

=
=

 
1
N 2 h2
8m A A


N 2 h2
1

2
8m
A
2 2
N h
8mA2

(12)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
In the early universe, a chemical equilibrium between photons and
e+ , e particles was achieved via the conversion process
e+ + e

(1)

The energy of an electron or a positron is given by


= mc2 +

1
mv 2
2

(2)

Using the fact that photons have zero chemical potential, derive an
equation describing the concentrations n+ and n of positrons and
electrons as related to each other at a given temperature.
The second law of thermodynamics demands that the entropy S of a given
system and its surroundings must either remain the same (in equilibrium), or
increase (dStotal 0). Also, because of this fact the Gibbs free energy G either
must either remain the same (in equilibrium) or decrease, i.e. dG 0. For a
pure system, with only one type of particle, this differential Gibbs free energy
is
G
dG
dG

E TS + PV

(3)

= dE T dS SdT + P dV + V dP
(Thermodynamic ID: dE = T dS P dV + dN )

= SdT + V dP + dN

(4)
(5)

At constant temperature T and pressure P (a safe assumption for the early


universe), the first two terms vanish, leaving only
dG = dN

(6)

For a system of mixed particles (like in this problem) this generalizes to


X
j dNj
dG = 1 dN1 + 2 dN2 + ... =
j

(7)

Jeff Kissel May 18, 2007

Statistical Mechanics

when there are j types of particles in the system.


OK. Now that weve got the basics outta the way, lets get to the problem.
The question established that the positions e+ , electrons e , and photons are
in equilibrium, so dG = 0. Using the coefficients of Eq. 1 as our dNj s, we find
X
j dNj
0 =
j

=
=

e+ Ne+ + e Ne + N
e+ (1) + e (1) + (2)

e+ + e + 2

(8)

However, the chemical potential for photons is zero (either believe the problem
statement, or check out Schroeder pp 289-290), so
0 =
e+ =

e+ + e
e

(9)

But wait! The mass of positrons and electrons are the same, which means
their energies are the same (via Eq. 2). From Eq. 4, at constant S (were
in equilibrium, so its OK) and at constant V (were taking the whole early
universe as our system; also OK),


E
=
(10)
N S,V
which means the chemical potentials must be equivalent. Hmm... the only
number thats both equivalent and opposite to itself is zero. So,
e+ = e =

= 0

(11)

Good to know. Now, the only thing left to do is find the concentrations Ne+
and Ne .
Positrons and electrons are both fermions, which means they obey the FermiDirac distribution,
h nj (Ej ) iF D

1
e(Ej ) + 1

(12)
2

p
1
where = kT
, Ej is the kinetic energy = 21 mv 2 = 2m
plus the rest mass
2
mc (as in Eq. 2) of a state j. In the early universe, where we have to consider
relativistic energies Ej kT , the exponential will be huge, so we can ignore the
additive factor of 1. Also, weve already shown that the chemical potentials of
the particles were working with are zero. This leaves a probability distribution
of

h nj (Ej ) i =

1
 + 1

e(Ej )

1
= eEj
eEj

(13)

Jeff Kissel May 18, 2007

Statistical Mechanics

The total number of each particle, N defined by summing over the probability distribution of states (i.e. Eq. 13),
X
h nj (E) i
(14)
N
j

but as weve already established, there are lots of energy states so we can safely
convert this to an integral, as long as weve got the conversion factor: the density
of states g(). Scuse me while I grab this ever-so-vital Jacobian,
X
X
= 2
j

=
=
=

d3 ~p d3 ~q


Z
3

V
2V
d p~ =
(2)3

Z
2V
(4)
k 2 dk
(2)3
0
2 2

p2
h k
=
2m
2m

let =

=
=
=
=

d3~k

=
=
=

h k
m dk
q
2m

h
2
q
2m
h
2
dk
m
h
2


m h2 1/2
2m

d
h2
h2 2m
0

Z
2 3/2
2V
m
1/2 d
(4)
(2)3
h3
0

Z
V 2 2 3/2
m
1/2 d
2 2
h3
0

3/2
Z
Z
2m
V
1/2

d
=
g() d
2 2 h2
0
0
2V
(4)
(2)3

Aaaand back to business. The total number of each particle is


X
h nj (E) i
N
j

h n(E) ig()d
e


E

E
eE

V
2 2

2m
h2

3/2

1/2

= mc2 + 1/2 mv 2
2
p

=
mc2 + 2m

=
mc +
mc2
=
e
e
3

(15)

Jeff Kissel May 18, 2007

Statistical Mechanics

3/2

Z
V
2m
mc2
=
e
e 1/2 d
2 2
h2
0
R
(n+1)
xn ex dx
=
n+1
R0 1/2 x

(3/2)
dx =
0 x e
3/2

(n + 1)
=
n!

(3/2)
= (1/2)! = 2

3/2
 

2m
V
mc2
e
=
2
3/2
2 2
2
h
3/2

2
2m
V
emc
=
4 h2

3/2
2
V 22 2mkT
=
emc
2
4
h
3/2

2
2mkT
emc
= 2V
2
h
1/2

h2
let =
2mkT
2
V
= 2 3 emc

(16)

Phew. OK, thankfully were done: because energies are the same, and s
are same, this calculation for N is valid for both positrons and electrons. If the
total number of each particle is the same, then the concentrations in the early
universe are equivalent.
The more subtle purpose of this question was to show the reason behind
one of the big unsolved questions in physics: we live in a real universe full of
electrons (as opposed to an anti-universe full of positrons), i.e. in the early
universe there must have been slightly more electrons than positrons. But we
just just proved that the concentrations were the same... W T F mate? Whelp,
show how theyre different, and Ill give ya a Nobel Prize.

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
A collection of N bosons is contained in a volume V . The spin of
the particles is 0.
a) Find the temperature at which Bose condensation occurs.
First, well find the total number of particles N in terms of the critical
temperature and volume, and then invert for the critical temperature Tc .
To find N , we sum the Bose-Einstein distribution,
h nj iBE

(j )

(1)

p2

j
is the kinetic energy of each state
over all j possible states. Here, j = 2m
1
(independent of position, q), is the chemical potential, and = kT
.

j=0

X
j=0

h nj iBE

(2)

1
e

(j )

1
e 1

X
j=1

1
e

(j )

(3)

Bose-Einstein condensation occurs when the chemical potential is approximately zero. Thus, we exclude the first term in Eq. 3 for now, as its limit goes
to zero as approaches zero. Then we can change the second term to an integral (valid for large N ), and proceed finding expression for Tc . The conversion
involves finding the density of states for bosons, so lets take a detour for a bit:
Z Z
X
=
d3 ~p d3 ~q

d3 p~ =

V
(2)3

d3~k

Jeff Kissel May 18, 2007

Statistical Mechanics

V
(4)
(2)3

=
=
=

k 2 dk

p2
~2 k 2
=
2m
2m

let =

=
=
=

h k
m dk
q
2m
~2
q
~2
2m
m
~2 dk

2m
m ~2 12
d

2
~
~2 2m
0

Z
2 3
1
V
2
2 d
(4)
m
(2)3
~3
0
Z

3

V
3
1
(2)

2
2m
(4) 3
2 d

3

h
(2)
0


 23
Z
Z

2V 2m

gBE () d
d

h2
0
0
V
(4)
(2)3

(4)

Ok. Back on course, remember were finding N at the critical temperature Tc ,


where 0.
N (Tc )

=
=

h nj ( 0) iBE

j=1
Z
0

=
=

h n( 0) iBE gBE () d

2V

2m
h2

2V

2m
h2

 32 Z

 32 Z

 1

()
1

2
d

e 1

d
dx =
d =
kT dx
let x =
1
1
12
= (kT ) 2 x 2


3 Z
1
1
2V 2m 2
(kTc ) 2 x 2

kTc d
=

h2
ex 1
0

3
Z
1
3
x2
2V 2m 2
2
(kT
)
=
c

h2
ex 1
0
R

xn1
=
(n)(n)
x 1 dx
0
e

( 32 ) = 2
( 32 ) 2.612
3


2 2mk 2
 3
2
Tc  ( )
= V 
2

h
2
2


2

(5)

Jeff Kissel May 18, 2007

Statistical Mechanics

3
3
2mk 2
Tc2
2
h
 2  32
h
1 N
3 V
2mk
( 2 )

3
= ( ) V
2

N
3

Tc2

Tc

h2
2mk

1 N
( 32 ) V

 23

(6)

Hooray! The critical temperature for Bose-Einstein condensation.


b) Find how the number of particles in the lowest energy state varies
with temperature below the condensation temperature.
If the particles are in a 3 dimensional box, the particles deBroglie wavelength
must be half integer multiples of the length L of a side. Hence,
L =

1
nn
2

n =

2L
n

(7)

and therefore the momentum in any given direction is


pi =

h
n

hni
2L

(8)

The individual kinetic energies in this 3-D box define a sphere in n-space,

=
=

|~
p|2
2m

2 
2 
2 !
hnx
hny
hnz
1
+
+
2m
2L
2L
2L
h2
(n2 + n2y + n2z )
8mL2 x

(9)

where n is the radius of the sphere in n-space. The lowest energy state will then
be
0 =

h2
2
8mL2 (1

+ 12 + 12 ) =

3h2
8mL2

(10)

which is a very small energy as long as L is macroscopic. The number of particles


in this ground state is given by Eq. 1, at j = 0 ,
N0

1
e

(0 )

(11)

When temperatures are below the critical temperature, N0 gets quite large.
Now look back at the integrand
of Eq. 5: as goes to zero, the density of
states (the part proportional to ) goes to zero while the Bose-Einstein distribution blows up (in proportion to 1/). Although the product is an integrable
3

Jeff Kissel May 18, 2007

Statistical Mechanics

function, its not clear that this infinite spike at = 0 correctly corresponds
to the sum over all states (Eq. 3). In fact, weve already seen from Eq. 11
that the number of atoms in the ground state can be enormous when 0,
and this number was not included in the calculation of the critical temperature
in part (a). On the other hand, the integral should correctly approximate the
number of exited particles not in the ground state with energies far away from
this divergence, i.e. 0 . If one imagines cutting off the integral at a lower
limit that is somewhat greater than 0 but much less than kT , one still gets
approximately the same answer,
Nexcited

3
( ) V
2

2mkT
h2

T
Tc

 23

(For T < Tc )

(12)

or more simply,
Nexcited

 32

(T < Tc )

(13)

If these are all the excited particles, the rest must be in the ground state, and
thus it follows that
N
N0

=
=
=

N0

N0 + Nexcited
N Nexcited
  23
T
N
N
Tc
  32 !
T
1
N
Tc

Props to Schroeder (2000), pgs. 315-318.

(T < Tc )

(14)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
The average energy of a system in thermal equilibrium is hEi.

a) Prove that the mean square deviation of the energy h (E hEi) i


is given by
h (E hEi)2 i =

kB T 2 CV

(1)

where CV is the heat capacity of the system at constant volume and


kB is Boltzmanns constant.
Ill proceed by attacking the problem from both sides of Eq. 1.
The left-hand side, can by simplified to something a little more reasonable
to decipher, and perhaps more recognizable (note I use some shorthand and call
the mean square deviation of the energy as h 2 i):
h 2 i =
=
=
=
h 2 i =

h (E hEi) i
h E 2 2 E hEi + hEi2 i


Ensemble average is linear: h A + B i = hAi + hBi

hE 2 i 2hEihhEii + hEi2

Ensemble average is a constant hhEii = hEi
hE 2 i 2hEi2 + hEi2

hE 2 i hEi2

(2)

which looks a lot like the definition of the variance that were used to from
statistics.
Now lets figure out some alternate definitions of each term in Eq. 2. The
ensemble average of any operator O is defined as
hOi

X
j

Oj Pj =

1 X
Oj eEj
Z j

(3)

P
where P is the probability distribution of states, Z j eEj is the partition
function, = (kT )1 , and Ej is the energy of the jth state.
1

Jeff Kissel May 18, 2007

Statistical Mechanics

The ensemble average energy, using Eq. 3, can therefore be written as


1 X
Ej eEj
Z j


X
X Ej
( Ej ) eEj
=
e
j
j


X
1

eEj
Z
j


X
Ej
e
Z

hEi

hEi

1 Z
Z

(4)

2
2


1 Z
1 Z
= 2
=

Z
Z

(5)

which means
2

hEi

Similarly, the ensemble average of energy squared is


hE 2 i =

=
hE 2 i =

1 X 2 Ej
Ej e
Z j
!
P
P
Ej
Ej
=
j ( Ej ) e
je

P
P Ej
2
2 Ej
=
j Ej e
je
2


1 2 X Ej
e
Z 2
j
1 2Z
Z 2 2

(6)

So, from Eqs. 5 and 6, Eq. 2 becomes


h 2 i =

hE 2 i hEi2 =

1
1 2Z
2
2
2
Z
Z

2

(7)

On to the right-hand side of Eq. 1. The heat capacity CV is defined as




hEi
(8)
CV
T V,N
We can convert the derivative with respect to T to one with respect to by
noting the following,



k
1
1

= (kT )2 (k) =
=
=
T
T kT
(kT )2
kT 2
2

Jeff Kissel May 18, 2007

Statistical Mechanics

CV is then
CV

=
kT 2 CV

=
=

kT 2 CV

hEi
hEi
1 hEi
=
=
T
T
kT 2


1 Z
Eq. 4: hEi =
Z


1
1 Z
+
kT 2 Z

 



Z
1 Z
1
+
Z
Z


2
1 Z
1 Z Z
+

Z 2
Z 2
2

1 Z
1 2Z
2
Z 2
Z

(9)

which, from Eq. 7 is exactly the MS deviation,


kT 2 CV

= h 2 i = hE 2 i hEi2

(10)

b) Use this result to show that the energy of a macroscopic system


may ordinarily be considered constant when the system is in thermal
equilibrium.
Ill show this be proving that the standard deviation in energy rms is very
small compared with the energy hEi.
We only have a definition for h 2 i, so Ill just prove the deviation is small
using that.
1/2
 2 1/2
 2
rms
h i
hE i hEi2
=
=
hEi
hEi2
hEi2

Eq. 1
: hE 2 i hEi2 = kT 2 CV
Equipartition Thm. : hEi = f N kT

=
rms
hEi

= f2 N k
Eq. 8
: CV = hEi
T

 1/2

!1/2
1/2

2 f
2
N
k
kT
4 kT 2 f2 (N k)
2
kT
2

=
=

2
2
f 2 (N k)2 T 2
kT
fN 
f
N
kT
2
1/2

2
(11)
fN

For macroscopic systems, in which N is huge, this ratio is really, really, really,
really, ridiculously small. For example, for 1 mole (N = NA = 6.02 1023 ) of
monatomic gas (f = 3),
rms
= 1.05234 1012
hEi
3

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
Consider a container of volume 100 cm3 containing a classical ideal
gas at 1 atm pressure and 350 K.
a) Find the number of particles.
Yo, wassup ideal gas law! (P is pressure, V is volume, N is total number of
particles, k is Boltzmanns constant, and T is temperature.)
P V = N kT
N

(1)
=

=
N

PV
kT

P
V

=
=

1 atm
100 cm3
k

=
=
=

1.013 105 P a

1 104 m3
1.381 1023 J K 1

(1.013 105 P a)(1 104 m3 )


(1.381 1023 J K 1 )(350 K)

= 2.0958 1021 particles

(2)

b) Compute the mean kinetic energy of a particle in the gas.


Holler at your Equipartition Theorem! (f is the number of degrees of freedom
for the gas particle.)
hEi =

=
hEi =

f
N kT
2
(Assume the gas is monatomic, i.e. f = 3)
3
(2.0958 1021 particles)(1.381 1023 J K 1 )(350 K)
2
15.159 J

(3)

(4)

c) Suppose one counted the number of particles in a small sub-volume


of size 0.1 micron on a side. What is the probability of finding no
particles in this volume?

Jeff Kissel May 18, 2007

Statistical Mechanics

If the small sub-volume v has sides of length L, then the sub-box has volume
v = L3 . Thus, there must be certain number Nv = V /v sub-volumes in the total
volume V . Now imagine emptying box V of particles. If we toss one particle in
the box, the probability that it will not land in a given sub-box will be




Nv 1
one particle
(5)
P
=
does not land in v
Nv
The next particle we toss in will have the exact same probability, and the one
after that, and so on. Thus, if we toss N particles in V , the probability that no
particles will be found in the sub-volume v is
P

none of N particles
are in v

=
=
P

none of N particles
are in v

N
Nv 1
Nv
!N
V
v 1

(6)

V
v

v
V

V
1
v

v N
1
V

N
(7)

For V = 1 104 m3 and v = 0.1 m = 1 1021 m3 , the ratio of v/V is


1 1017 . This means whats in parentheses is (1 [retarded small]) 1, and
therefore P = 1N = 1. (Graphing calculators and even Mathematica, after
taking the natural logarithm, i.e. computing N ln (1 v/V ), still say that its
one).
No particles will be found in a 0.1 m sub volume.

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 17, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
#35
#46
#46

Notes/Changes

Question number shift


Comment added to part (c)

The pressure in a vacuum system is 103 mm Hg. The external pressure is 1 atm at 300 K. This is a pinhole in the vacuum system of
area 1010 cm2 . Assume that any molecule entering the pinhole goes
through. Use an average molecular weight of air as 29 amu.
(a) How many molecules enter the vacuum system each hour?
The process in which gas at high pressure is moving freely through a pinhole
into a much lower pressure is known as effusion. Effusion problems are usually
defined such that the flow of particles is from a confined system, at high pressure,
outward into an unconfined vacuum. This problem, which has particles leaking
into a confined vacuum, is very similar. The resulting equation of particle flow
rate dN/dt will differ only by a negative sign to account for the directional
difference. Thus, Ill derive the flow rate for outward leaking gas, and change
the sign at the end.
The pressure P on an area A of the container wall caused by a single air
particle of mass ma over a time period t is


1
F
ma v
=
(1)
P =
A
t
A
If we (arbitrarily) chose a reference frame so the particle moves in the x
direction, with average velocity vx , then v as it bounces of the wall in time t
should be 2 vx . This velocity can be approximated by the root-mean-squared
velocity vx = vxrms defined by the temperature and mass of the particle via
the Equipartition Theorem,
hEiKin

hEiT herm
1

Jeff Kissel May 17, 2007

Classical Mechanics

1
ma vx2
2
1
ma vx2
2

=
=

vx2

q
vx2

1
kT
2
1
kT
2
kT
ma
r
kT
vxrms
ma

(2)

This turns Eq. 1 into


P

v 1
t A
2 vx 1
= ma
t A
2 vxrms 1
ma
t A
r
2ma
kT
=
At
ma

= ma

(3)

For an infinitesimal amount of (non-interacting) air particles dN , the pressure of the gas trying to escape becomes
r
r
2ma
2(ma dN )
kT
kT dN
=
(4)
P =
A dt
ma
A
ma dt
This pressure should be approximately equivalent to the pressure on a wall
with a pinhole poke in it, as the area of the hole (A = 1010 cm2 ) is approximately the size of the air molecule, allowing barely more than one particle
through at a time.
We can now re-arrange for the flow rate of air particles dN/dt from a pressurized contain outward into a vacuum,
r
dN
P A ma
=
(5)
dt
2ma kT
As Ive mentioned earlier, Ill now flip the sign, because the direction of flow
is opposite from that which weve used to derive the flow rate. Thus, the flow
rate of particles into a confined vacuum from a pressurized exterior is
r
1
dN
1
=
P A
(6)
dt
2
ma kT
(Note that Ive combined the ma s).
Finally, we plug in the numbers given in the problem statement, converting
to mks units. For Pa = 1 atm = 1.013 105 P a, Apinhole = 1010 cm2 = 1

Jeff Kissel May 17, 2007

Classical Mechanics

1014 m2 , ma = 29 amu = (29 amu)(1.6611027 kg/amu) = 4.81026 kg,


k = 1.381 1023 J K 1 , and T = 300 K, the flow rate is
dN
dt

1
(1.013 105 P a) (1 1014 m2 )
2
s

dN
dt

1
(4.8
kg)(1.381 1023 J K 1 )(300 K)

particles s1 3600 s hr1
1026

3.5854 1013

1.3 1017 particles hr1

(7)

(b) If the volume of the system is 2 liters, by how much does the
pressure rise in 1 hour?
Assuming the air obeys the ideal gas law (quite safe at room temperature),
we can find the change in pressure by computing
V P
P

= kT N
kT
N
=
V

(8)

assuming that effusion happens slow enough such that the system remains in
thermal equilibrium, i.e. T is constant.
Having computed the flow rate of particles into the vacuum (Eq. 7), over
one hour the number of particles in the chamber should increase by
N = Nf Ni
N

dN
t
dt
= (1.3 1017 particles hr1 ) (1 hr)
= 1.3 1017 particles

(9)

which if volume remains a constant V = 2 L = (2 L)(0.001 L/m3 ) = 0.002 m2


at room temperature (T = 300 K), the change in pressure after the same hour
should be
P

=
=

kT
N
V
(1.381 1023 J K 1 )(300 K)
(1.3 1017 particles)
(0.002 m3 )
0.267 P a = 0.002005 mm Hg

(10)

(c) How long does it take for the pressure to rise to 750 mm Hg? Note:
this is close to the pressure outside the vacuum tank.
We now have a rate for pressure increase from part (b), (assuming the pressure increases linearly with time as the ideal gas law dictates),
dP
dt

= 2.005 103 mm Hg hr1


3

(11)

Jeff Kissel May 17, 2007

Classical Mechanics

which means the it will take


dP
P

=
=

dt
t
(Reset clock such that ti = 0)
t
(750 mm Hg)
P
=

(2.005 103 mm Hg hr1 )


3.74 105 hr ( = 42 yrs !)

to increase the pressure 750 mm Hg.

(12)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
Two phases of a pure material coexist along a line Pcoex (T ). Use mA as
the molar mass of the substance. The latent heat of transformation
(in J/mole) is L, and the molar volume of each phase are v1 and v2 respectively. Assume that phase 1 is the stable phase for temperatures
lower than the temperature where they coexist.
a) What is known about the chemical potentials of each phase in the
low temperature region where phase 1 exists, the high temperature
region where phase 2 exists, and on the coexistence line?
Chemical potential is defined as the Gibbs free energy per particle. The
proof is as follows,
G
dG
dG

E TS + PV

(1)

= dE T dS SdT + P dV + V dP
(Thermodynamic ID: dE = T dS P dV + dN )

(2)

= SdT + V dP + dN

(3)

Taking the partial derivative of both sides of Eq. 3 with respect to N , holding
T and P constant,


G
=
(4)
N T,P
or more simply, since weve assumed the number of particles is fixed in addition
to temperature and pressure,
G =

(5)

Phase transitions are determined when two phases are in chemical equilibrium with each other (for example a solid phase in equilibrium with a liquid
phase is the melting/freezing phase transition). In order to achieve an overall
system (system A and the reservoir) increase in entropy (essential for a spontaneous process), the Gibb free energy is minimized in equilibrium:
dStotal

dSA + dSR 0
1

(6)

Jeff Kissel May 18, 2007

Statistical Mechanics

For each part of the system, Eq. 2 applies, (used here in the form dS =

P
T dV T dN ) such that for fixed T , P , and N
dStotal

=
=
=
=

dE
T

0
>

1

P
dSA +
dER +
dVR 
dN
T
T
T

(dER = dER ; dVR = dVA )
dEA
P
dSA

dVA
T
T
1
(T dSA dEA P dVA )
T
1
(dEA T dS + P dVA )
T
1
(dN )T,P
T


0 
0
:

:


(dG)
= 
SdT 
V dP + dN
T,P

dStotal

1
dG
T

(7)

Thus, the temperature and pressure line where the Gibbs free energy, and
(via Eq. 5) the chemical potential of each phase is equivalent defines the phase
transition. Therefore, on the coexistence linePcoex (T ), the chemical potentials
are equivalent.
By inspection, one can then infer that at low temperatures where phase 1
exists, phase 1 has lower chemical potential than phase 2, i.e. 1 < 2 (low
chemical potential corresponds to stronger attraction). At high temperatures
where phase 2 exists, 2 < 1 .
b) Derive the Clausius-Clapeyron equation for the slope of the coexistence curve.
dPcoex
dT

L
T (v1 v2 )

(8)

As established in part (a), the Gibbs free energies along the coexistence
curve for two phases are equivalent. So, for constant N , using Eq. 3,
G1 = G2
: 0= S dT + V dP + 
:0


S1 dT + V1 dP + 
1
dN
2 dN
2
2
(V1 V2 ) dP = (S1 S2 ) dT
dP
(S1 S2 )
=
dT
(V1 V2 )

(9)

The molar latent heat L is defined as L Q/m where Q is the heat required
to complete a particular phase transition and m is the molar mass. During this
phase transition, temperature is constant, and for such an isothermal process
2

Jeff Kissel May 18, 2007

Statistical Mechanics

S = S1 S2 = Q/T . Thus Eq. 9 becomes the Clausius-Clapeyron (C-C)


relation,
dP
dT

=
=
=

dP
dT

(S1 S2 )
(V1 V2 )
m
Q
T (V1 V2 ) m
Q/m
T (V1 /m V2 /m)
L
T (v1 v2 )

(10)

c) If the higher temperature phase 2 can be treated as an ideal gas,


and molar volume of phase 1 is so small compared to v2 as to be
neglected, find the dependence of the saturated vapor pressure on
temperature.
For high temperatures (as the question dictates), v2 can be neglected, so the
C-C relation becomes
dP
L
(11)
=
dT
T v1
From the ideal gas law, the molar volume is
P V = N kT

or P V = nRT

Pv

RT
RT
P

(12)

which means Eq. 11 yields


dP
dT
dP
P

=
=

1
dP =
P
P0
 
P
=
ln
P0
P
=
P0

Psvp (T ) =

LP
RT 2
L
dT
RT 2
Z T
L
1
dT
R T0 T 2


L 1
1

R T
T0
L

eR

( T1 T1 )
0

P0 e R ( T0 T )

(13)

which is the saturated vapor pressure as a function of temperature. P0 is the


experimental value of the pressure at temperature T0 (usually room temperature), used as calibration. This is called the saturated vapor pressure because
we ignored v2 , i.e. claiming that the system is saturated with phase 1.
3

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
Consider a collection of N particles of spin-1/2 in a volume V . Compute the magnetic spin susceptibility of these particles if they can be
treated as a non-degenerate gas, but
The order in which well attack both parts of this problem is to find (1)
the partition function Z for the system, (2) the ensemble average magnetic
moment hi, (3) the magnetization per unit volume M , and (4) the magnetic
susceptibility m . In equation form, thats
X

Z1
hi

ej

X
j

M =

j Pj
hi
V

1
ZN
N! 1

Z =

hi =
m =

ln(Z)
(B)

M
B

(1)
(2)
(3)

a) the spin is still treated quantum mechanically.


To treat the spin quantum mechanically implies that the magnetic portion
of energy states are discrete, i.e. in the presence of an external magnetic field of
strength B, the fermions (spin-1/2 particles) have only the spin up (aligned
with B, and energy = B) and spin down (anti-aligned with B, =
+B). The total energy of a given state is then the usual continuous, freeparticle kinetic portion (independent of position) and the discrete, spin-magnetic
field interaction portion (independent of position and momentum),

p2
B
2m

(4)

This means the partition function for one particle Z1 is


X
ej
Z1
j

eB

p2

e 2m + eB

X
j

p2

e 2m

Jeff Kissel May 18, 2007

Statistical Mechanics

eB + eB

X

p2

e 2m

=
=
=
=

=
=
Z1




1 x
x
e +e
cosh x =
2
 Z Z

p2
1
3
3
2m
(2 cosh (B))
d ~p d ~q
e
h3
Z
p2
V
e 2m d3 p~
2 cosh (B) 3
h
Z
p2
V
2 cosh (B) 3
e 2m d3 ~p
h
3
Z
p2
V
e 2m dp
2 cosh (B) 3
h

r 
Z

x2
e
dx =


3/2
V 2m

2 cosh (B) 3
h

3/2

h2
2mkT
let

=
2V cosh (B)
h2
2mkT
2V
cosh (B)
3

(5)

For a system of N such particles, which are indistinguishable, the partition


function (from Eq. 1) is

N
1
2V
Z =
coshN (B)
(6)
N!
3
Onto the average magnetic moment. Any generic (ensemble) average of an
observable is defined as
X
1 X
O(j)P(j) =
hOi =
(7)
Oj ej
Z
j
j
which means the average magnetic moment is
hi

1 X
j ej
Z j

p2

e 2m e
j

p2
2m

e
P

(B)

p2

e 2m e
p2

e 2m e

Jeff Kissel May 18, 2007

Statistical Mechanics

1
(Z)
Z (B)

(ln(Z))
(B)

=
hi

(8)

and plugging in Eq. 6


hi

=
=
=

=
hi

"

#)
N
1 2V
N
ln
cosh ( B)
N ! 3
" 
#
N


1 2V

N
ln
+ ln cosh B
(B)
N ! 3
" 
#
N
2V

1 


ln
+
[N ln (cosh ( B))]


3

(B)
N
!

(B)



d
cosh u = sinh u du
du
N sinh (B) ()
cosh (B)
N tanh (B)

(B)

(9)

The magnetization per unit volume is then


M

hi
V
N
tanh (B)
V

(10)

Finally, the magnetic susceptibility for a collection of spin-1/2 particles,


whose spin is treated quantum mechanically is
m

M
N
=
tanh (B)
B
VB

(11)

b) theirmagnetic moments are treated as a classical magnetic moment


of = 3B (recall that s2 = 3/4 for a spin-1/2 systems), but that
the moment can point in any direction, so = B cos(), with the
component of the magnetization along B being given by cos().
Again, we begin by establishing the energies of each state. As before,
there remains the continuous free-particle, kinetic energy portion, yet the spinmagnetic field interaction term is now continuous as well and no longer independent of position, i.e.

p2
B cos
2m
3

(12)

Jeff Kissel May 18, 2007

Statistical Mechanics

which means we cant pull that portion out of the (single particle) partition
function as we did in part (a). So,
X
ej
Z1 =
j

=
=

Z1



Z Z
p2
B cos
2m
1
d3 p~ d3 ~q
e
h3
Z Z
p2
1
e 2m e+E cos d3 p~ d3 ~q
3
h

3
h
|

p2

e 2m

{z
Zp

3
e+B cos d3 ~q
d p~

}
{z
}
|

(13)

Zq

The momentum portion Zp computes very similar to part (a), just with out
the volume,
Z
p2
1
2m
Zp =
d3 p~
e
h3
Z
p2
1
2m
d3 ~p
e
=
h3
Z
3
p2
1
2m
e
=
dp
h3

3/2
1 2m

=
h3

3/2

1
2mkT
= 3
(14)
Zp =
2
h

The position portion of the partition function Zq (independent of momentum) is a little trickier, but still doable,
Z
Zq =
e+B cos d3 ~q
=

2 Z

e+B cos r2 sin drdd

R 3
if
d ~q = R R R
V
2
then
V
=
r
sin

R 2 drdd

V
=
4
r dr
R 2

r dr =
V /4
Z 2 Z
V
e+B cos sin dd
4 0
0

Jeff Kissel May 18, 2007

=
=
=
=

Zq

Statistical Mechanics

V
(2)
4

e+B cos sin d

= sin d
du
du =
sin d
let u = cos

01
1
 Z 1

V
Bu

e
du
2
1
Z 1
V
eBu du
2 1

1
V
1
Bu
e
2 B
1

V 1 B
e
eB
B 2



1 x
x
sinh(x) =
e e
2
V
sinh (B)
B
0

(15)

Combining Eqs. 14 and 15, the partition function for a single particle, and
thus for the whole system is
Z1

V sinh (B)
sinh (B)
3
B

N
N 
V
sinh ( B)
2
B

(16)

From here, the rest is just plug and chug, in the same manner as in part (a):
find hi, then M , and finally m . Ere we go!
hi =
=
=

=
=
hi =

(ln (Z))
(B)
( "
N #)
N 
V
sinh ( B)

ln
(B)
2
B

 

i


h
 V
ln
N
N
ln
(sinh
(
B))

N
ln
B
+

(B)
2
(B)




d
sinh(u) = cosh(u)du
du


N
N cosh (B)()

sinh (B)
B
1
N coth (B) N
B


1
N coth (B)
(17)
B
5

Jeff Kissel May 18, 2007

=
=

Statistical Mechanics



N
1
hi
=
coth (B)
V
V
B
M
B 

N
1
coth (B)
VB
B

(18)

(19)

and finally, plugging in = 3B as the problem desires, the magnetic susceptibility for a collection of spin-1/2 particles, whose spin is treated classically
is




3B N
1
3B B
m =
(20)
coth
VB
3B B

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
An object of heat capacity C is used as the cold thermal reservoir
by a Carnot engine. The hot reservoir has an infinite heat capacity.
During the operation of the engine to produce work, the temperature
of the cold reservoir will slowly rise. Assume the starting temperature
of the cold reservoir is Tc and the constant temperature of the hot
reservoir is Th .
a) To what temperature will the cold reservoir before the engine
ceases to produce work?
To cease producing work W , the engines efficiency must go to zero. Efficiency is defined as
benefit
cost

W
Qh

(1)

where Qh is the heat cost, taken from the hot reservoir. For a closed cycle
(Carnot) engine, the first law of thermodynamics tells us that energy in the
cycle must be conserved. Thus, the heat absorbed must be equal to the heat
released Qc plus the work produced, or
Qh = Qc + W

or W = Qh Qc

(2)

We can plug this into Eq. 1 to get

Qc
Qh Qc
= 1
Qh
Qh

(3)

The second law of thermodynamics states that the entropy S Q/T of


any system plus its surroundings cant decrease. Here, since the engine cycle is
closed, this means the entropy expelled to the cold reservoir must be at least as
much as that absorbed from the hot reservoir, i.e.
Qc
Qh

Tc
Th

Qc
Tc

Qh
Th

(4)

Now, or equation for efficient becomes

1
1

Tc
Th

(5)

Jeff Kissel May 18, 2007

Statistical Mechanics

The engines efficiency will drop to zero once the second term in Eq. 5 goes
to unity, in other words when Tc = Th . So, when the cold reservoir (originaly
as Tc ) reaches that of the hot reservoir Th , the engine will cease to do work.
b) How much heat flows into the cold reservoir until the engine stops
producing work?
When one asks how much heat can flow into an object, were concerned
about its heat capacity, C, defined as
C

Q
T

(6)

The engine will stop doing work once the maximum amount heat has flowed
into the cold reservoir, i.e. when the engine has increased the temperature from
its current Tc to Th , and described in part (a). To find the amount of heat this
requires, we re-arrange Eq. 6, using this temperature change as T ,
Qch

CTch = C (Th Tc )

(7)

c) What is the total amount of work the engine can produce before
the process stops?
As the engine does work, it will steadily increase the temperature of the
cold reservoir to Th , at which point work will cease. As stated in part (a), the
entropy released to the cold reservoir must either remain constant or increase
during a cycle. We want the maximum amount of work, so well assume that
no entropy escapes during a given cycle such that
dQc
Tc

dQh
Th

(8)

Because the temperature of the cold reservoir is increasing every cycle, the
amount of heat that can be given decreases over time. Incrementally this heat
is dQc = CdTc , from Eq. 6. So, Eq. 8 becomes,
CdTc
Tc

dQh

dQh
Th
T
C h dTc
Tc

(9)

The amount of work produced by this incremental cycle in terms of temperature, from re-arranging Eq. 1, plugging in Eq. 5 for , and Eq. 9 for
dQh
dW

=
=

dW

dQh



T
T
C h dTc
1 c
T
T
 h  c
Th
C
1 dTc
Tc
2

(10)

Jeff Kissel May 18, 2007

Statistical Mechanics

The total work produced by the engine as the temperature of the cold reservoir increases from some initial Tc = Tc to Th when work ceases is
W

Th

Tc

=
=
=
W

Th
1
Tc

dTc

Z Th
1
dTc
dTc C
CTh
Tc
Tc
T
 hc
iTh   
T

C Th ln (Tc )
Tc Thc
Tc

 

Th
C Th ln
Th + Tc
Tc
   


Th
C Th ln
1 + Tc
Tc
Z

Th

(11)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
X
#52
#52

Notes/Changes

Added to question bank


Almost entirely reworded.

A certain gas has an equation of state


P

= b+

N RT
V

(1)

where N is the mole number, V is the volume, T is the absolute


temperature and b is a constant.
(a) Find the value of CP CV . where CP and CV are the heat capacities
at fixed pressure and volume.
The equation we need for the difference in heat capacities is




P
V
CP CV = T
T V,N T P,N

(2)

which Ill take a few lines to prove. The variables used will be entropy S, temperature T , pressure P , volume V , chemical potential , total internal energy E,
and total number of particles N . Subscripts on parentheses of partial derivatives
indicate what is being held constant.

S
T
S
T

dS

P,N

P,N

S(T, V, N )






S
S
S
dT +
dV +
dN
T V,N
V T,N
N T,V










V
S
S
S
(0)
(1) +
+
T V,N
V T,N T P,N
N
 T,V







V
S
S
+
T V,N
V T,N T P,N
1

Jeff Kissel May 18, 2007

Classical Mechanics

!


S
S
1
T T
C

=
P
T P,N
T
P,N
(3)
S
S
T T

= T1 CV
T V,N
V,N




V
1
S
CV +
T
V T,N T P,N




V
S
T
V T,N T P,N
!




S
P
Maxwell relation:
(4)
=
V T,N
T V,N




V
P
T
T V,N T P,N
CP
CV

1
CP
T

CP CV

CP CV

(5)

OK, now that you believe me on that one, we can re-arrange our equation
of state (Eq. 1) to find the partial derivatives in Eq. 2.
N RT
V
N RT
(P b) =
V
(P b)V = N RT
N RT
N RT
P = b+
V =
P b
V




V
NR
P
NR
=
=
T P,N
P b
T V,N
V
P

= b+

(6)

which means Eq. 2 becomes


CP CV

=
=
=

CP CV



NR
NR
T
V
P b



NR
NR
T
V
(N RT /V )
  
T
V
NR
V
T
NR

(7)

(b) Show that constant volume heat capacity is not dependent on the
volume. Specify how the heat capacity can depend on N and T .
We can kill two birds with one stone by showing that
CV
=0
V

(8)

We can do so be using the definition of CV as seen from Eqs. 3, and Eq. 4


2

Jeff Kissel May 18, 2007

Classical Mechanics

again,


CV
V

#

S
T
T P,N
T,N
"
#


S
T
V
T P,N
T,N


x y f (x, y) = y x f (x, y)
"
#


S
T
T
V T,N
P,N
"
#

P

T
T
T V,N
P,N
!




N RT
NR

P
b+
=
=
T V,N
T
V
V



NR
T
T
V

T,N

=
CV
V

"

(9)

(c) This gas undergoes a process at fixed temperature where the


volume is changed from Vi to Vf . Find the change in entropy in terms
of the variables N, R, T, Vf and Vi .
The change in entropy for an isothermal process is given by
S

Q
T

(10)

where Q is the heat leaving the system. The amount heat leaving the system
will be given by the first law of thermodynamics,
U

Q =

Q =

Q+W
U W

Equipartion Thm.: U
Isothermal process: T
U

=
=

N RT

0
0

(11)

The work done to decrease the volume from Vi to Vf can be found by integrating
the pressure,
W

= P V
Z Vf
P dV
=
Vi

Jeff Kissel May 18, 2007

Classical Mechanics



N RT
Eq. 1: P = b +
V
Z Vf
Z Vf
N RT
b dV
=
dV
V
Vi
Vi
 
Vf
= b(Vf Vi ) N RT ln
Vi

(12)

which means the total heat lost is


Q

= W = b(Vf Vi ) + N RT ln

Vf
Vi

(13)

Vf
Vi

(14)

and the change in entropy is


S

Q
b
=
(Vf Vi ) + N R ln
T
T

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
Consider a classical gas composed of N particles of mass m that posses
a permanent electric dipole moment . The gas is subject to a uniform
electric field of size E in the the x
-direction and is enclosed in a
container of size V .
a) Find the partition function.
The partition function for a single particle is defined by
X
b]
ej = Tr[e H
Z1

(1)

b is the systems Hamiltonian, and


where j is the energy of the jth state, H
1
b
b is defined as the sum over all
= (kT ) . The trace Tr[O], of some operator O
states, or the integral over all phase space
Z Z Z
b = 1
~
Tr[O]
O d3 ~p d3 ~q d
(2)
h3

where h is Plancks constant, p & q are the generalized momentum and position,
and is the solid angle corresponding to internal rotations about and . The
factor of h3 appears to be sure the trace is dimensionless.
Because these particles have a permanent dipole moment , the Hamiltonian
is not just the usual free-particle kinetic energy component, but also the electric
energy component generated by the interaction between the electric field E and
the dipole moment. In equation form, thats
b
H

b
H

2
b
p~
b~ E
~

2m
2
b
b
p~
=
p~ b
p~
b
~

~ E = E
b cos

pb2
E
b cos
2m

= pb2

!
(3)

where is the angle of the dipole from the z axis (in spherical polar coordinates),
~ is assumed to point (, the angle from the x
along which the electric field E

Jeff Kissel May 18, 2007

Statistical Mechanics

axis, is free to rotate from 0 to 2). Using Eqs. 1, 2, and 3 the partition function
for this system integral form is then


Z Z Z
p2
E cos
2m
1
~
d3 ~p d3 ~q d
Z1 =
e
h3
Z Z
p2
1
~
=
e 2m e+E cos d3 ~p d3 ~q d
3
h

Z1

1 Z
p2

d3 ~q
e 2m d3 ~p
= 3

h
{z
}
| {z }
|
Zp

Zq

~
e+E cos d

(4)

{z
}
|
Z

The momentum portion of the partition function Zp (independent of position


and internal angle) can be cleaned up significantly,
Z
p2
1
2m
Zp =
d3 p~
e
h3
3
Z
p2
1
2m
dp
e
=
h3
r 
Z
2

ex dx =


3/2
1 2m
=

h3

3/2

h2
2mkT
let

=
=
h2
2mkT
1
Zp =
(5)
3
where Ive defined a variable which has dimensions of length, typically called
the quantum length.
Because position portion of the partition function Zq (independent of momentum and internal angle) has no integrand, we just evaluate it to be the
volume,
Z
Zq =
d3 ~q = V
(6)

Finally, the internal angle portion of the partition function Z (independent


of poistion and momentum) is a little trickier, but still doable,
Z
~
Z =
e+E cos d
=

2 Z

e+E cos sin d d

Jeff Kissel May 18, 2007

Statistical Mechanics
Z

e+E cos sin d

=
du
du =
let u = cos

01
 Z 1

Eu
= 2
e
du

= 2

= 2

sin d
sin d
1

eEu du

Zq

1
1
Eu
= 2
e
E
1

2
eE eE
=
E


ex ex = 2 sinh(x)
=

4
sinh (E)
E

(7)

Combining Eqs. 5, 6, and 7 yields the partition function for one particle,
Z1 =

4V sinh (E)
3
E

(8)

For a system of N such particles, which are indistinguishable, the partition


function is

N
N 
sinh (E)
1 4V
(9)
Z = N1 ! Z1N =
N!
3
E
b) Find the net polarization of the sample as a function of the filed
strength, pressure, and temperature. The polarization P is defined
as the net dipole moment per unit volume.
As stated in the question, the polarization P is defined as the average dipole
moment hi per unit volume V . In equation form, thats
P

hi
V

(10)

So, lets find hi. Any generic (ensemble) average of an observable is defined as
X
1 X
(11)
Oj ej
O(j)P(j) =
hOi =
Z
j
j
which means the average dipole moment is
1 X
j ej
hi =
Z j
3

Jeff Kissel May 18, 2007

Statistical Mechanics

=
=
hi

p2

e 2m e+(E)
p2

e 2m e+(E)

P
p2

+(E)
2m
e
e
j
(E)
P p2 +(E)
2m e
je
j

1
(Z)
Z (E)

(ln(Z))
(E)

(12)

and plugging in Eq. 9


( "

N #)
N 
1 4V
sinh (E)

ln
hi =
(E)
N!
3
E
#
"

N !




1 4V
N
N
=
ln
+ ln sinh ( E) ln (E)
(E)
N ! 3
"



 !#
N
1 
4V


ln 
=
(E) N ! 3


i
h 
+
ln sinhN ( E) ln (E)N
(E)

[N ln (sinh ( E)) N ln (E)]


=
(E)


d
sinh(u) = cosh(u)du
du


N
N cosh ( E)()

=
sinh ( E)
E
1
hi = N coth ( E) N kT
(13)
E
From which we find the polarization P (using Eq. 10) in terms of electric field
strength E, pressure P , and temperature T ,
hi
V
N
N kT 1
=
coth ( E)
V
V E

P V = N kT
P
N
V = kT = P and
1
= P coth ( E) P
E


1
P(E, P, T ) = P coth ( E)
E
P

N kT
V

= P

(14)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
X
#54
#54

Notes/Changes

Added to question bank.


Part (d) changed.

The energy spectrum of neutrinos (spin 1/2 massless particles) is


= pc. For a collection of N neutrinos confined to volume V , calculate
Ill do part (b) first, and then (a).
(b) The Fermi energy F .
Ill find the Fermi energy by first finding the density of modes g() for the
massless, spin 1/2 neutrinos, then the total number of particles N () at temperature T = 0, which I can then invert for .
The ensemble average of any observable O is defined as
X
Oj Pj
(1)
hOi =
j

The density of modes g() is the Jacobian of transformation when one


changes this sum to an integral. This conversion for massless fermions (where
we add a factor of 2 to account for the spin degeneracy) is as follows
X
X
= 2
j

=
=
=

Z Z
2
d3 p~d3 ~q
h3
Z
2
d3 ~p
V
h3

Z
2V
(4)
p2 dp
h3
0
1

Jeff Kissel May 18, 2007

X
j

Classical Mechanics

p =
c
1
dp = c d
let = pc
2
2
p =
c2
Z  2
1
8V

=
d
2
h3
c
c
0
Z 2
8V

=
d
3
h3
c
Z 0
Z
8V 2

g() d
(hc)3
0
0

(2)

On to the total number of particles, which is the sum over the probabilities
of each state (i.e. Eq. 1 without the observable). The probability distribution
for fermions is the Fermi-Dirac distribution,
nF D

e()

+1

(3)

which at the limit of zero temperature splits into two regions (i.e. has values
1 or 0). The non-vanishing portion of the distribution is in the energy regime
0 < < F , (the upper limit is the Fermi Energy, or chemical potential at
zero temperature), because the Pauli Exclusion Principle dictates that fermions
(spin-1/2 particles) cannot occupy the same energy state. Thus instead of all
collapsing to some ground state, they stack up in energy to F when temperatures very small.
Thus, at zero temperature, the total number of particles is
X
PF D
N (T = 0) = lim
T 0

=
=
=
=
N

1
g() d
() + 1
e
0
Z
(0) g() d
(1) g() d +
F
0
Z F
Z F
8V 2
8V

d
=
2 d
3
3
(hc)
(hc)
0
0
8V 3F
(hc)3 3
8 V 2

(hc)3 3 F

lim
T 0
Z F

(4)

which we can solve for the Fermi energy of massless, spin 1/2 neutrinos,
3F

(hc)3 3 N
8 V

1/3
hc 3 N
2 V
2

(5)

Jeff Kissel May 18, 2007

Classical Mechanics

(a) The Fermi wave vector.


The Fermi wave vector is quickly obtainable from the Fermi energy, because
for photons,
E

= pc
(p = h
k)
hk
=
c
2
2
=
cE
h

E
k

(6)

So,
kF

kF

2
c F
h

1/3 !
hc 3 N
2
c
=
h
2 V

1/3
3N
= c2
V

1/3
2N
2
= c 3
V
=

(7)

(c) The total energy at T = 0.


The total ensemble averaged energy of the system at T = 0 will be given by
X
hEiT =0 =
PF D
k

PF D g() d


Z F
8V 2
(1)
d

(hc)3
0
Z F
8V
3 d
(hc)3 0
8V 4F
(hc)3 4

1/3 !4
hc 3 N
2V
(hc)3
2 V

4/3 !
2V
(hc)4 3 N
(hc)3
16
V

1/3

3N
hc  3 N
V

V
8
V

0

UT =0

=
=
=
=
=
=

Jeff Kissel May 18, 2007

Classical Mechanics

hEiT =0

hc
2

3
N
4

3
N F
4

3N
V

1/3 !
(8)

(d) The compressibility of the neutrino gas.


We can treat the neutrino gas as an ideal (non-interacting) gas, such that
V

N kB T
P

(9)

The (isothermal) compressibility, in terms of the pressure is then






1 V
1
N kB T 1
N kB T
1
1
T
=
=
= (P ) 2 =
2
V P
V P
P
V P
P
P
which is true for any ideal gas. So we need the pressure of the nuetrino gas.
The pressure P exerted by neutrinos should in general be
1 X
(10)
Pj ej
P =
Z j
where Pj is the pressure of any given energy state j .
The neutrinos temperature is a constant T = 0, so there is no heat transfer,
Q = SdT = 0. We can derive an expression for the pressure of a given state Pj
using the thermodynamic identity (assuming the number of particles remains
constant)
dj
dj
Pj

0
*

= 
dQ
j + dWj
(Definition of Work: dW = P dV )
= Pj dV
j
=
V

(11)

For the free neutrino gas, the system is effectively an infinite 3-D infinite
square well. Tor each dimension, the particles deBroglie wavelength must then
be half integer multiples of the length L of a side. Hence,
L =

1
nn
2

n =

2L
n

(12)

The momentum in any given state is then


pj =

h
n

hnj
2L

(13)

1
hcnj

2L
L

(14)

making the energy


j

= pj c =
4

Jeff Kissel May 18, 2007

Classical Mechanics

According to Eq. 20, we need the partial derivative of this energy with
respect to V . Acknowledging that L = V 1/3 ,
Pj

=
=
=
=

Pj

j
V 

hcnj 1

V
2 V 1/3


1 1
hcnj

2
3 V 4/3


11
hcnj 1
+
2 V 1/3 3 V
1 j
3V

(15)

Plugging this result back into Eq. 19,


PT =0

=
=
=

=
=
PT =0

1 X
Pj e
Z j


1 X 1 j
e
Z j
3V

1 11 X
j e
3V Z j


X
1 X
From Eqs. 1:
j Pj = hEi
j e =
Z j
j
1 hEiT =0
3 
V

11 3
N F
3V 4
1N
F
4V

(16)

Which makes the isothermal compressibility


T

V
F
N

1/3 !
V
hc 3 N
= 4
N
2 V
1/3

3 V2
= 2hc
N2
= 4

(17)

Jeff Kissel May 18, 2007

Classical Mechanics

OLD VERSIONS
b) The pressure exerted by the neutrinos at T = 0.
In general, the probability distribution P in Eq. 1 is
P

1 X j
e
Z j

(18)

where Z is the partition function, and j is the energy of the jth state. The
pressure P exerted by neutrinos should then be
P

1 X
Pj e
Z j

(19)

The neutrinos temperature is a constant T = 0, so there is no heat transfer,


Q = SdT = 0. We can derive an expression for the pressure of a given state Pj
using the thermodynamic identity (assuming the number of particles remains
constant)
dj
dj
Pj

0
*

= 
dQ
j + dWj
(Definition of Work: dW = P dV )

= Pj dV
j
=
V

(20)

For the free neutrino gas, the system is effectively an infinite 3-D infinite
square well. Tor each dimension, the particles deBroglie wavelength must then
be half integer multiples of the length L of a side. Hence,
1
nn
2

L =

n =

2L
n

(21)

The momentum in any given state is then


pj =

h
n

hnj
2L

(22)

1
hcnj

2L
L

(23)

making the energy


j

= pj c =

According to Eq. 20, we need the partial derivative of this energy with
respect to V . Acknowledging that L = V 1/3 ,
Pj

=
=

j
V 


hcnj 1

V
2 V 1/3

Jeff Kissel May 18, 2007

Classical Mechanics

=
=
Pj



1 1
hcnj

2
3 V 4/3


hcnj 1
11
+
2 V 1/3 3 V
1 j
3V

(24)

Plugging this result back into Eq. 19,


1 X
Pj e
Z j


1 X 1 j
=
e
Z j
3V

1 11 X
j e
=
3V Z j


X
1 X
j Pj = hEi
j e =
From Eqs. 1 and 18:
Z j
j

PT =0

1 hEiT =0
3 V

(25)

c) The heat capacity for 0 < T F .


Awwww shit. Its the Sommerfield expansion again. It doesnt matter that
the neutrinos are massless, as long as theyre fermions.
If I denote the constants in the density of modes to be CF , and use < n >
as shorthand for h n() iF D then the ensemble average energy (as seen in part
(b)) is
Z
Z
3
hEi =
h n() iF D gF D () d = CF
2 < n > d (26)
0

However, now we need energies at low temperatures (kT F ), not just


at T = 0. Although the integral in Eq. 26 goes from 0 to , i.e. all positive
energies, the most interesting region is near = , where h nj (j ) i falls off
sharply. So, well isolate this region using integration by parts,

hEikT F

let u =
hni
v = 52 5/2
d<n>
let dv = 3/2
du = d d
"

 #

:0 2 Z


d<n>
2 5/2 
5/2

d


<n> +

CF
5 
5 0
d

0

(27)

where the surface term vanishes because at = 0, 5/2 = 0, and at = ,


7

Jeff Kissel May 18, 2007

Classical Mechanics

< n >= (e() + 1)1 = 0.


Z
2
CF
hEi =
5/2
5
0

d<n>

=
d
Z
2
=
CF
5/2
5
0

hEi

This leaves,


d<n>
d

d



e(j )
1
d
=

d e(j ) + 1
(e(j ) + 1)2

e(j )
(e(j ) + 1)2

dx
=
d
let x = ( )
0 ,

Z
ex
2
CF
dx
5/2 x
5
(e + 1)

(28)

Because, the integrand of this expression falls off exponentially when x =


( ) 1, we can now make two approximations:
1. Extend the lower limit of the integral () to (); this makes the
integral symmetric, and its harmless because the integrand is negligible at
negative s because of the shape of the distribution at these temperatures.
2. Taylor expand 5/2 around = ,

 5/2 
 2 5/2 

d( )
1
d ( )
5/2
5/2
2

+ ( )
+ ( )

+ ...
d
2!
d2

15
1
5
( )2 1/2 + ...
= 5/2 + ( )3/2 +
2
2 4
5
15
= 5/2 + (xkT )3/2 + (xkT )2 1/2 + ...
(29)
2
8
So that
hEi

(I)

"
Z
2
ex
dx
CF 5/2
x
2
5
(e + 1)
|
{z
}
(I)
Z
ex
5
dx
kT 3/2
x x
+
2
(e + 1)2

|
{z
}
(II)
#
Z
x
15
e
+
dx +...
(kT )2 1/2
x2 x
8
(e + 1)2

|
{z
}
(III)
Z

ex
dx
(ex + 1)



d<n>

d
d

Jeff Kissel May 18, 2007

Classical Mechanics
Z

d<n>

"

#
1
0
:

:


>
 > + 
<
n()
<
n()




=
Z

ex
dx
+ 1)

(II)

= 1

(ex

ex
x ex
dx
x
x
x
(e + 1)(e + 1) e
Z
x
dx
x + 1)(1 + ex )
(e

{z
}
| {z } |

x ex
dx =
(ex + 1)2

(30)

odd
function

even
bounds

x ex
dx =
+ 1)2

(III)

| {z }

even
bounds

(ex

x2 ex
dx
(ex + 1)2
| {z }

= 2

v
let dv

x2
2x dx

x2 ex
dx
(ex + 1)2

xm
dx
ex + 1

=
=
=
=

x2 ex
dx
+ 1)2

(ex

=
=

1
ex +1
ex
(ex +1)2

0

Z
>


x2 
2x
 +
2
dx

ex

x
+1 0
e +1
0

Z
x
dx
4
x+1
e
0

1
(1 m) (m + 1) (m + 1)
2
1
4(1 ) (2) (2)
2
 2
 

1
(1)
4
2
6
2

3
2
3

=
Z

even
function

let u =
du =

(31)

(32)

Jeff Kissel May 18, 2007

hEi

Classical Mechanics



5 
2
2
CF 5/2 (1) + 
CF
kT 3/2 (0)
5
5
2

 2
15

2
+ ...
CF
(kT )2 1/2
+
5
8
3
2
2
CF 5/2 +
CF (kT )2 1/2 + ...
5
4

 32
h
2
Eq. 5:
F
= 2m
3 2 N
V2

3
2m

F
=
3 2 N
V
h
2 3

3/2
2m 2
F
=
3 2 N

V
h
2

3

N
2m 2
V

=
3
2
3/2

2
h

3
2m 2
V
3 N
CF =
=
2 2
2 3/2
h
2
F
!
!
2

3 N
3
N
2
5/2 +
(kT )2 1/2 + ...
5
2 3/2
4 2 3/2
F

=
hEi

(33)

3
5/2
1/2
3 2
N 3/2 +
N 3/2 + ...
5
8
F
F
 5/2
 1/2

(kT )2
3 2

3
N F
N
+
+ ...
5
F
8
F
F

(34)

Performing a similar Taylor expansion process for N (see Schroeder pp 282284) yields the expansion that,

2

2 kT
= 1
+ ...
(35)
F
12 F
which we can exploit for our purposes,

5/2

1/2

2
1
12

kT
F

2

!5/2

+ ...

(Binomial Expansion: (1 + x)p 1 + px (For x 1))



2
5 2 kT
+ ...
1
2 12 F

2
5 2 kT
+ ...
(36)
1
24
F
!5/2

2
2 kT
1
+ ...
12 F

2
1 2 kT
+ ...
1
2 12 F

2
2 kT
1
+ ...
(37)
24 F
10

Jeff Kissel May 18, 2007

Classical Mechanics

and plug back into the ensemble average energy (Eq. 34),
!

2
5 2 kT
3
N F 1
hEi =
+ ...
5
24
F
!

2
3 2
2 kT
(kT )2
+
1
N
+ ... + ...
8
F
24 F

(38)

and then keeping only terms of order (kT /F )2 or less,


hEi =
=
hEi =



3 2
3 2
(kT )2
(kT )2
(kT )2
3
+
+O
N F
N
N
5
24
F
8
F
3F


3 2
2
(kT )2
(kT )2
(kT )2
3
+
+O
N F
N
N
5
8
F
8
F
3F
3
2
(kT )2
+ ...
N F +
N
5
4
F

(39)

Now we can finally get to the question. Remember we are trying to show
that CV T for low temperatures. This is quick attainable from Eq. 39, and
the definition of the heat capacity at constant volume,


hEi
CV
T V,N


3
2
(kT )2

+ ...
N F +
N
=
T 5
4
F
2
2

T T
(40)
N
CV =
2
F
d) The rms fluctuations in the mean energy of the neutrinos of 0 <
T F .
As in part (c) it doesnt matter that the neutrinos are massless, only that
theyre fermions.
For any system in equilibrium the mean square fluctuations (variance) is
h 2 i =

hE 2 i hEi2 = kT 2 CV

(41)

(for a detailed proof of this check out statmechS06-32F06-43). From Eq. 40,
this becomes
h 2 i

h 2 i

= kT 2 CV
 2


k2
2
= kT
T
N
2
F
N
= (kT )3
2 F
11

(42)

Jeff Kissel May 18, 2007

Classical Mechanics

which means the rms (root mean square) fluctuations are


rms


1/2
N
(kT )3
2 F

12

(43)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
#27
#38
#55

Notes/Changes

Question number shift.


Part (c) removed.

Consider an ideal Fermi gas of spin-1/2 particles in a 3-dimensional


box. The number of particles per unit volume is n, the mass of the
particles is m, and the energy of the particles is the usual = p2 /2m.
Assume that the temperature is quite low.
a) Find formulas for the Fermi energy F , Fermi wavevector kF , and
Fermi temperature TF in term of m, n and constants such as h and
kB .
To find the Fermi energy (the chemical potential of spin 1/2 particles at zero
temperature), we must find the total number of particles N as T approaches
zero.
For any temperature T , one normally finds N via summing over the FermiDirac probability distribution h nj () iF D (because the particles are of spin
1/2),
h nj () iF D

1
e(j ) + 1

(1)

1
where = kT
, j is the kinetic energy of a state j, and is the chemical potential
of the particles. However, at the limit of zero temperature, h nj () iF D decays
to either 0 (when j > ) or 1 (when j < ). The energy, at T = 0, at which
the now-step-function distribution flips from 1 to 0 is the Fermi energy, whose
value is determined (again) by the number of particles.
To perform the calculation of the number of particles, we first find the density
of states: the Jacobian of transformation between summing over energy states
j , and integrating over kinetic energies. First, to account the spin degeneracy

Jeff Kissel May 18, 2007

Statistical Mechanics

of +1/2 and 1/2, we tack on a factor of 2,


Z
Z Z
X
X
2
2
3
3
= 2
= 3
d3 p~
d p~ d ~q = 3 V
h
h

j
k
Z
2V
(4)
p2 dp
=
h3
0

2m 1/2
=
2 p

p
1
1/2
let =
dp = 2 2m
d

2m 2
p =
2m
Z 
 1
8V
=
2m
2m1/2 d
h3
2
0
Z
8V 1
3/2
(2m)
=
1/2 d
h3 2
0
3/2

Z
Z
X
2m
1/2
g() d

4V

h2
0
0
j

(2)

Now we can find the number of particles N , and invert for the Fermi energy
F .
N (T = 0) =

X
j

h nj () iF D

h n() iF D gF D () d
Z
Z F
(0) g() d
(1) g() d +
0

=
N

=
=

4V


2m
h2
3/2

3/2

2m
2 23

2
h
3 F
3/2

3
8
2m
F2
V
2
3
h

2/3
2
h
13N
2m 8 V

2/3
h2 3 N
8m V

2/3
N
h2
3 2
2m
V
4V

2 d

(3)

(4)

In this form its nice and quick to pull out the Fermi wave-vector kF , since

Jeff Kissel May 18, 2007

Statistical Mechanics

the quantum mechanical kinetic energy is


p2
h2 2
=
k
2m
2m
So by direct comparison between Eqs. 4 and 5,
 31

2 N
kF =
3
V
E

(5)

(6)

The Fermi terperature TF is similarly quick, as the thermal kinetic energy


is,
E

kB T

(7)

Inverting for temperature and subbing in 4,


2

N 3
h2
3 2
TF =
2mkB
V

(8)

b) Find the total energy of the gas at zero temperature.


This calculation can be done by summing over possible energy states, weighted
by their probabilities (determined by the Fermi-Dirac Distribution),
X
j h nj (j ) i
hEi
j

=
hEiT =0

Z0 F
0

h n iF D gF D () d
Z
(0) g() d
(1) g() d +
F

g() d

=
=
=
=

hEiT =0

3

V
2m 2 1
2 d
2 2 h2
0

3 Z
2m 2 F 3
V
2 d
2 2 h2
0
3

2m 2 2 52
V

2 2 h2
5 F
3

V
2m 2 52
F
5 2 h2
!
3

V
2m 2 32
Eq. 3: N =
F
3 2 h2

3
N F
5

(9)

Jeff Kissel May 18, 2007

Statistical Mechanics

c) If the magnetic moment of the particles is e , show that the paramagnetic susceptibility for low fields in the limit of zero temperature
is given by

3 n2e
2 F

(10)

This part deals with the quantum mechanical property of fermions known as
the Zeeman splitting. When a weak magnetic field is applied to a fermi gas, the
~ |i) and down (anti-parallel to B,
~ |+i) energy levels of
up (parallel to B,
the particles are perturbed such that the energy states of the up particles are
increased, and the down particles are decreased, causing a splitting of what
initially was a single set of quantized energy states. The Zeeman effect is a 1st
order perturbation theory problem in quantum mechanics (Check p245-246 of
Griffiths, or Wikipedias article for more detialed explanations. Merzbachers is
pretty weak sauce, but its on p472-473). Anyways, this is a Stat-Mech problem,
so we really only care about the resulting energies, so Ill only highlight the
details.
The perturbing hamiltonian is
b
H

b B
~
~
e

(11)

b is the applied, weak magnetic field, and


where B
be is the magnetic moment of
the fermions (spin-1/2 particles), given by

~be

b
H

b
e B

b
H

b
~
gJ B S
h

(12)

where gJ is the Lande g-factor, B = (eh)/(2m) is the Bohr magneton, and


Sb is the spin operator. The Lande g-factor is approximately 2 for (spin-1/2)
electrons (See Wikipedias Lande g-factor for details), so I assume gJ = 2.
b (Eq. 11, assuming the field |B|
~ = B applied in the same direction as
Thus, H
e ) becomes,
=

2 B Sb

h


h
b
Si =

bi
2

B
bz B

(13)

where
bz is the Pauli spin matrix, and Ive assumed that the field was applied
in the z-direction.
The total energy of the anti-parallel (down, |+i, or the positive eigenvalue
of
bz ) state is
T

= + B B
4

Jeff Kissel May 18, 2007

Statistical Mechanics

in which, as before, = p2 /(2m) is the kinetic energy of the particles. The


parallel (up, |i, negative eigenvalue of
bz ) energies are
T

= B B

Which means the Fermi-Dirac distribution (1) becomes


h nj () iF D

1
e[(j B B))

+1

(14)

i.e. is splits into two distributions,


1
h nj ( B B) iF D
2

and

1
h nj ( B B) iF D
2

(15)

where the factors of 1/2 insures the distribution remains normalized. The left
distribution corresponds to the up spins, i.e. those energy levels displaced by
a positive B B and therefore the probability range over which it extends is
from B B < T < . The right distribution are the down spins, displaced by
negative B B, so the range is B B < T < .
From here, we can calculate the net magnetization per unit volume M , the
derivative of which (with respect to B) is . The net magnetization would be
the total number of particles anti-align subtracted from those aligned (in a given
volume), or
M

(N N )
V

(16)

We know how to find the number of particles: integrate the probability distribution corresponding the particles over the correct region. Thus, the magnetization
is then
Z
B
1
M =
h n( + B B) iF D gF D () d
V
B B 2

Z
1

h n( B B) iF D gF D () d
B B 2
Z
B
[h nj ( + B B)i h nj ( B B)i] g() d
=
2V 0
!
f (x)
f (x+h)f (xh)

x
2h
(x)
f (x h) f (x + h) =
2 h fx

Z 
B
h n() i
=

g() d
2 B B
2V

0

Z 
2B B
h n() i
M =
g() d
(17)

0
At low temperatures (i.e kT F ), the Fermi-Dirac distribution becomes
asymptotically close to a step function with the barrier at the Fermi energy
5

Jeff Kissel May 18, 2007

Statistical Mechanics

(as discussed in part (a)). Thus, the derivative of the distribution at low temperatures is well approximated by a delta function about F , or

Z
Z 
h n() i
d =
( F ) d
(18)
lim

T 0 0

0
which makes Eq. 17 a lot prettier,
M

=
=
=

Z
2B B
( F ) g() d
V
0
2B B
g(F )
V
1/2 B
2B CF F
V
!
3 N
Eq. 28: CF =
2 3/2
F
!
3
N
B
1/2
2B
F
2 3/2
V
F

3 2B N
B
2 F V

(19)

Finally, if I convert to the notation the proof desires (N/V n), the magnetic susceptibility is

M
B 

3 2B

nB
B 2 F
3 n 2B
2 F

which is sometimes referred to as the Pauli Susceptibility.

Its cool. Youll remember all of this during the qualifier. I promise.

(20)

Jeff Kissel May 18, 2007

Statistical Mechanics

OLD VERSIONS
c) Show that the heat capacity at low temperature is proportional to
T.
The exact proof of this problem unfortunately gets rather lengthly, but is
known as the Sommerfield expansion.
If I denote the constants in Eq. 2 to be CF , and use < n > as shorthand for
h n() iF D then the ensemble average energy (as seen in part (b)) is
Z
Z
3
hEi =
h n() iF D gF D () d = CF
2 < n > d (21)
0

However, now we need energies at low temperatures (kT F ), not just


at T = 0. Although the integral in Eq. 21 goes from 0 to , i.e. all positive
energies, the most interesting region is near = , where h nj (j ) i falls off
sharply. So, well isolate this region using integration by parts,

hEikT F

let u =
hni
v = 52 5/2
d<n>
let dv = 3/2
du = d d
"

 #

:0 2 Z


2 5/2 
d
<
n
>
5/2
d
CF

< n > +

5 
5 0
d

0

(22)

where the surface term vanishes because at = 0, 5/2 = 0, and at = ,


< n >= (e() + 1)1 = 0. This leaves,


Z
d<n>
2
5/2
d

CF

hEi =
5
d
0




d<n>
e(j )
1
d

=
=
d
d e(j ) + 1
(e(j ) + 1)2
Z
2
e(j )
=
CF
5/2 ( )
5
(e j
+ 1)2
0

dx
=
d
let x = ( )
0 ,

Z
ex
2
CF
dx
(23)
5/2 x
hEi =
5
(e + 1)

Because, the integrand of this expression falls off exponentially when x =


( ) 1, we can now make two approximations:
1. Extend the lower limit of the integral () to (); this makes the
integral symmetric, and its harmless because the integrand is negligible at
negative s because of the shape of the distribution at these temperatures.
2. Taylor expand 5/2 around = ,

 5/2 
 2 5/2 

d( )
1
d ( )
5/2
5/2
2

+ ( )
+ 2! ( )
+ ...
d
d2

Jeff Kissel May 18, 2007

1 15
5
( )2 1/2 + ...
5/2 + ( )3/2 +
2
2 4
5
15
5/2 + (xkT )3/2 + (xkT )2 1/2 + ...
2
8

=
=
So that
hEi

Statistical Mechanics

(24)

"
Z
2
ex
5/2
dx
CF
x
2
5
(e + 1)
|
{z
}
(I)
Z
ex
5
3/2
dx
kT
x x
+
2
(e + 1)2

|
{z
}
(II)
#
Z
15
ex
2 1/2
2
+
dx +...
(kT )
x
8
(ex + 1)2

|
{z
}
(III)

(I)

ex
dx
x
(e + 1)



d<n>
d

d<n>

"

#
:1
:0




 >

<
n()
> +
<
n()


=
Z

ex
dx
(ex + 1)

(II)

= 1
Z

(25)

ex
x ex
dx
x
x
x
(e + 1)(e + 1) e
Z
x
dx
x
(e + 1)(1 + ex )

{z
}
|
| {z }

x ex
dx =
(ex + 1)2
=

even
bounds

(III)

| {z }

even
bounds

x ex
dx =
+ 1)2

(ex

x2 ex
dx
(ex + 1)2
| {z }

odd
function

= 2

(26)
Z

even
function

x2 ex
dx
(ex + 1)2

Jeff Kissel May 18, 2007

Statistical Mechanics

v
let dv

x2
2x dx

let u =
du =

xm
dx
+1

ex

=
=
=
=

hEi

2 x

x e
dx
+ 1)2

(ex

hEi



5 
2
2
CF 5/2 (1) + 
CF
kT 3/2 (0)
5
5
2

 2
2
15

+
+ ...
CF
(kT )2 1/2
5
8
3
2
2
CF 5/2 +
CF (kT )2 1/2 + ...
5
4

 32
h
2
3 2 N
Eq. 4:
F
= 2m
V2

3
2m

F
=
3 2 N
V
h
2 3

3/2
2m 2
F
=
3 2 N

2
V
h

3

V
2m 2
N

=
3
2
2
3/2

3
V
3 N
2m 2
CF =
=
2 2
2 3/2
h
2
F
!
!
2
2
3 N
3 N

5/2 +
(kT )2 1/2 + ...
5
2 3/2
4 2 3/2
F
5/2

1
ex +1
ex
(ex +1)2

0

Z
>


x2 

2x
 +
2
dx
x+1

ex
+ 1 0
e
0

Z
x
dx
4
x+1
e
0

1
(1 m) (m + 1) (m + 1)
2
1
4(1 ) (2) (2)
 2  2 

1
(1)
4
2
6
2

3
2
3

Z

=
=

(27)

(28)

1/2

3
3
N 3/2 +
N 3/2 + ...
5
8
F
F
 5/2
 1/2
3

(kT )2
3 2

N F
N
+
+ ...
5
F
8
F
F

(29)

Jeff Kissel May 18, 2007

Statistical Mechanics

Performing a similar Taylor expansion process for N (see Schroeder pp 282284) yields the expansion that,

= 1

2
12

kT
F

2

+ ...

(30)

which we can exploit for our purposes,

5/2

1/2

2
1
12

kT
F

2

!5/2

+ ...

(Binomial Expansion: (1 + x)p 1 + px (For x 1))



2
5 2 kT
+ ...
1
2 12 F

2
5 2 kT
+ ...
(31)
1
24
F
!5/2

2
2 kT
1
+ ...
12 F

2
1 2 kT
+ ...
1
2 12 F

2
2 kT
+ ...
(32)
1
24 F

and plug back into the ensemble average energy (Eq. 29),
!

2
5 2 kT
3
N F 1
hEi =
+ ...
5
24
F
!

2
3 2
2 kT
(kT )2
+
1
N
+ ... + ...
8
F
24 F

(33)

and then keeping only terms of order (kT /F )2 or less,


hEi =
=
hEi =



(kT )2
3 2
3 2
(kT )2
(kT )2
3
+
+O
N F
N
N
5
24
F
8
F
3F


2
2
2
2
3
3

(kT )
(kT )
(kT )2
+
+O
N F
N
N
5
8
F
8
F
3F
2
(kT )2
3
+ ...
N F +
N
5
4
F

(34)

Now we can finally get to the question. Remember we are trying to show
that CV T for low temperatures. This is quick attainable from Eq. 34, and

10

Jeff Kissel May 18, 2007

Statistical Mechanics

the definition of the heat capacity at constant volume,




hEi
CV
T V,N



3
2
(kT )2
=
+ ...
N F +
N
T 5
4
F
2
2

k
CV =
T T
N
2
F

11

(35)

Statistical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 18, 2007
Consider a system consisting of two particles, each of which can be
in any one of three quantum states of respective energies 0, , and 3.
The system is in contact with a heat reservoir at temperature T .
a) Write an expression for the partition function Z if the particles
obey classical Maxwell-Boltzmann statistics and are considered distinguishable.
The partition function of any system is defined as
X
ej
Z

(1)

where = (kT )1 , and j is the energy of the jth state. When particles obey
Maxwell-Boltzmann statistics, it means that the two particles are distinguishable from each other. Each particle can occupy all three energy states, as shown
below.

Figure 1: The energy level diagram for a particle in three possible energy states.
The partition function a given particle is
X
ej
Z1 =
j

Z1

=
=

e
+ e() + e(3)
1 + e + e3
(0)

(2)

For N distinguishable particles, each will have the exact same partition function.
So, this system of two distinguishable particles will have the partition function
2
Z = Z1N = 1 + e + e3
(3)
1

Jeff Kissel May 18, 2007

Statistical Mechanics

b) Write a similar expression if the particles obey Bose-Einstein


statistics.
If the particles obey Bose-Einstein statistics, it means there are indistinguishable from each other, unlike Maxwell-Boltzmann particles. The particles
(known as bosons) also have integer spin, which means they can occupy the
same energy state. Because of each of these facts, we must draw the system as
a whole and count the energies. There are nine possible states,

Figure 2: The energy level diagram of a two-boson,, three-energy-level system.


where weve ruled out 4 , 7 , and 8 because theyre indistinguishable from
2 , 3 , and 6 respectively. Each state has a total energy j equal to the total
energy of both particles, so the partition function becomes
X
ej
Z =
j

= e

+ e2 + e3 + e5 + e6 + e9

= e
+ e(0+) + e(0+3) + e(+0) + e(+3) + e(3+3)
= 1 + e + e3 + e2 + e4 + e6
(4)
(0+0)

c) Write a similar expression if the particles obey Fermi-Dirac statistics.


Finally, Fermi-Dirac particles (fermions) are also indistinguishable, but have
half-integer spins. This has two consequences: 1) they obey the Pauli exclusion
principle, which means they cannot occupy the same state like bosons can, and
2) each energy state will have some degeneracy related to the spin s of the
particles. For example, if s = 1/2, then we would have to tack on an extra
factor of 2 to partition function. In general for spin s, the correction factor goes
as 2s + 1. Lets count the possible states.

Jeff Kissel May 18, 2007

Statistical Mechanics

Figure 3: The energy level diagram of a two-fermion,, three-energy-level system.


Again, energy states 3 , 5 , and 6 are rules out because of indistinguishability. So, the following the same technique as before, but accounting for the
2s + 1 spin degeneracy,
X
ej
Z = (2s + 1)
j

=
=
Z


(2s + 1) e1 + e2 + e4


(2s + 1) e(0+) + e(0+3) + e(+3)

(2s + 1) 1 + e + e3 + e4

(5)

Classical Mechanics Review


(Louisiana State University Qualifier Exam)
Jeff Kissel
May 19, 2007

Exam Iteration
Spring 2006
Fall 2006
Spring 2007

Question Number
X
X
#3

Notes/Changes

Added to bank.

A box of volume V hold blackbody radiation at a temperature T .


(a) Show that the amount of energy stored in a small range of frequencies (f, f + df ) is given by
E(f ) df

a f3
df
1

ebf

(1)

and find expressions for a and b in terms of V, T, kB , h, and c.


The total energy will be given by the energy times the probability of any
given energy state, summed over all configuration states. Sounds daunting, but
youve seen it before.
Photons follow the Planck distribution function,
PjP

hEi

1
1

(2)

j PjP

(3)

ej

so the total energy is


X
j

where the sum is over j, the number of polarizations (for photons theres 2:
linearly polarized and circularly polarized), and wave number k = 2
c f.
X
j PjP
hEi =
j

= 2

X
k

k PkP
1

Jeff Kissel May 19, 2007

=
=
=

=
=
hEi

E(f ) df

Classical Mechanics

Z Z
1
2 3
p PpP d3 ~p d3 ~q
h


p R6= p (~q) and PpP 6= PpP (~q)

d3 ~q = V
Z
V
2 3
p PpP d3 p~
h


p = h
~
~k d3 p~ = h
3 d3~k
Z
V 3
2 3 h
k PkP d3~k
h

Z
V
k PkP d3~k
2
(2)3
Z
V
(4)
k PkP k 2 dk
2
(2)3
0

ck =
ck = 2f

k = 2 f
c

dk = 2 df
c
f = hf

2  
Z
2
2
1
V
df
f
hf hf
2
0
e
1 c
c
Z
V (2)3 hf 3
df
2 c3
ehf 1
0
Z
f3
8V h
df
c3 ehf 1
0
8V h
f3
df
c3 ehf 1

(4)
(5)

So,
a

8V h
c3

(6)

and
b

h =

h
kB T

(7)

(b) The Wien displacement law describes the relationship between the
peak intensity in the blackbody energy distribution (Eq. 1) and the
temperature. Derive this relationship. Note: the function x3 /(ex 1)
has a maximum at x = 2.82 approximately.
The intensity at a given frequency for a blackbody is given by
I(f )

1
f3
1 E(f )
2h
c u(f ) =
c
=
4
4
V
c2 ehf 1
2

(8)

Jeff Kissel May 19, 2007

Classical Mechanics

Thus if we find a relationship between the maximum frequency fmax and temperature T , we find the relationship with the maximum intensity Imax and
temperature T .
As the problem suggests, we can define the dimensionless constant x = hf ,
so that f = x/(h) and
I(x)

2h
c2

1
h

3

x3
(kB T )3 x3
=
2
ex 1
(hc)2 ex 1

such that when we take the maximum with respect to x




I(x)
x3
(kB T )3

= 2

x
(hc)2 ex 1 xmax



x3

0 =

x ex 1 xmax

(9)

(10)

we end up with exactly what we need from the hint, that xmax = 2.82. Plugging
back in for x,
xmax
fmax

fmax

= hfmax
kB T
xmax
=
h
kB =

h =
xmax =

5.873 1010

1.38 1023 J/K


6.626 1034 J s
2.82

(K s)1 T

(11)

(c) Find the total energy of the blackbody radiation and how it depends on V, T, h, c and kB .
The total energy is given by Eq. 4, so we just need to carry out the integral,
Z
8V h
f3
hEi =
df
c3 ehf 1
0

1
=
f
h x
1
df
=
Let x = hf
h dx

00 ,
 
Z  3
1
x
1
8V h
dx
=
x1
c3
h
e
h
0
 4 Z
x3
8V h 1
dx
=
3
x
c
h
e 1
0
Z

yn
dy
=
[n
+
1][n
+
1]
ey 1
0
4
8kB
V T 4 [4][4]
=
(hc)3
3

Jeff Kissel May 19, 2007

=
hEi

Classical Mechanics


 4


4
[4][4] =
(6) =
90
15
 5 4 
8 kB
V T4
15(hc)3
V T 4

(12)

where at the last step Ive grouped all constants together into , which is known
as the radiation constant, related to the Stefan-Boltzmann constant in the
same way intensity (which is luminosity for you astronomers) is related to energy,

4
1
2 5 kB
c =
3
4
15h c2
1 V T 4
1 hEi
c
=
c
= T 4
4
V
4
V

(13)

(14)

Anda mungkin juga menyukai