CP1 Notes
CP1 Notes
CP1 Notes
Toby Adkins
1 Mechanics I 3
1.1 Newton's Laws and Energy Conservation 4
1.1.1 Newton's Laws 4
1.1.2 Energy Conservation 4
1.2 Collisions 5
1.2.1 Centre of Mass frame 5
1.2.2 Oblique Scattering 7
1.2.3 Max Deection Angle 7
1.3 Resisted Motion 9
1.3.1 Resisted Projectile Motion 9
1.4 Variable Mass Problems 11
1.4.1 Rocket Motion 11
1.4.2 Relativistic Rocket 12
2 Mechanics II 14
2.1 Dimensional Analysis 15
2.2 Dynamical Coordinate Systems 16
2.2.1 Rotating Reference Frame 16
2.3 Rotational Dynamics 18
2.3.1 Angular Momentum and Torque 18
2.3.2 Moment of Inertia 18
2.3.3 Axis Theorems 20
2.3.4 Example Problems 21
2.4 Central Forces and Orbits 25
2.4.1 Kepler's Laws 25
2.4.2 Energy of Orbital Motion 26
2.4.3 The Eective Potential 27
2.4.4 The Reduced Mass System 28
2.4.5 Central Force Scattering 29
2.4.6 Example Problems 30
2.5 Lagrangian Mechanics 35
2.5.1 Denitions 35
2.5.2 The Lagrangian 36
2.5.3 The Hamiltonian 36
2.5.4 Example Problems 37
3 Special Relativity 45
3.1 The Michaelson-Morley Experiment and Einstein's Postulates 46
3.1.1 The Aether and The Michaelson-Morley Experiment 46
3.1.2 Einstein's Postulates 47
3.2 The Lorentz Transformations 49
1
Toby Adkins CP2
2
1. Mechanics I
• Collisions
• Resisted Motion
Most of the material covered here will not be unfamiliar to students who have studied
physics in high-school, though it may be expressed in a slightly dierent way, such as using
vector notation in collisions. The mathematics required is quite simple, only requiring
basic knowledge of integration and linear algebra.
3
Toby Adkins CP2
3. When a body exerts a force on a second body, the second body simultaneously exerts
a force equal and opposite to the second body.
Most students may not have seen (1.1) in the form shown above; generally, this will reduce
to F = m dv
dt for problems that do not involved variable mass.. However, (1.1) does lead to
an interesting result: in an isolated system, total momentum is conserved.
X
F =0
dp
→ =0
dt
This means that pr (the linear momentum) is a conserved quantity.
In general, we only use Newton's Laws for simple systems, as it is much more eective to
use other methods when the system becomes more complicated.
The law of the conservation of energy states that in any closed system, subject to no external
non-conservative forces, energy is conserved. This can be used as a powerful analytic tool,
even when the system is subject to external forces. For example, one can nd the about
of energy dissipated by a resistive force by equating the energy before and after; we know
that the dierence must be equal to the work done by the non-conservative force.
4
Toby Adkins CP2
1.2 Collisions
These should be the types of problems that most people will already be familiar with at
this stage, but we would recommend reading this section in any case as there may be some
elements not previously encountered. There are a couple of crucial points to remember for
collisions:
• In all collisions subject to no external forces, momentum is conserved
m1 u1 + m2 u2 = m1 v1 + m2 v2
m1 (v1 − u1 ) = m2 (u2 − v2 )
m1 u1 2 + m2 u2 2 = m1 v1 2 + m2 v2 2
m1 (v1 2 − u1 2 ) = m2 (u2 2 − v2 2 )
m1 (v1 − u1 ) · (v1 + u1 ) = m2 (u2 − v2 ) · (u2 + v2 )
Hence, we obtain:
v1 − v2 = u2 − u1 (1.6)
This means that in a elastic collision, relative velocity is maintained throughout.
5
Toby Adkins CP2
• The speeds of the particles before and after the collision are the same.
m1 u01 + m2 u02 = 0
m1 v 01 + m2 v 02 = 0
u02 − u01 = v 01 − v 02
m1 0 m1 0
− u − u01 = v 01 + v
m2 1 m 1
2
0 m1 0 m1
−u1 1 + = v1 1 +
m2 m2
→ |ui | = |v i | (1.8)
• The particles leave an oblique collision at the same angle in the centre of mass frame.
This is because angles are transformed in the CM frame; we need to use trigonometry
to convert back to the lab frame.
Consider the following example. A particle of mass 2m travelling at an initial speed u col-
lides inelastically with a stationary particle of mass m. What is the velocity of the second
particle after the collision?
6
Toby Adkins CP2
This is the nal velocity of the particle of mass m in the lab frame.
→ 2v 1 v 2 · cos(θ) = 0 (1.9)
This leads to the interesting result that either v 1 = 0 or θ = π2 after the collision. In the
second case, u1 · v 1 > 0, and hence the rst particle is scattered through an angle of less
than π2 .
7
Toby Adkins CP2
Eliminating v2 :
2
m1 m2 u1 2 − v 1 2 = m1 2 u1 2 − v 1 2
2
= m1 2 u1 2 + v1 2 − 2u1 v1 cos θ
u1 m2 v1 m2
cos(θ) = 1− + 1+
2v1 m1 2u1 m1
Hence, the maximum deection angle for the larger mass is given by:
m2
θ = sin −1
(1.10)
m1
8
Toby Adkins CP2
9
Toby Adkins CP2
Now, for max range we want y = 0. We can use the Taylor expansion of the logarithm and
ignore terms of O(k3 ) and above for small resistive forces:
1 k2 2 1 k3 3
vo g g k
0= x+ x+ 2 − x− x − x
uo kuo k uo 2 uo 2 2 uo 2
vo 1 g 1 gk 2
= − 2
x− x
uo 2 uo 3 uo 2
s
2
3uo 1 g 1 g 2 4 vo gk
→x= − 2
± 4
+
2gk 2 uo 4 uo 3 uo 4
10
Toby Adkins CP2
For variable mass problems where the mass is being acquired or lost vertically at zero
velocity, the net force on the system is just equal to the weight of the mass. For a rate of
mass ejection α:
X dv
F =m
dt
= mo g − (mo − αt)g
= αgt
This just represents the loss of momentum from the system that results from the body
losing mass.
By (1.1):
δp = (m − δm)(v + δv) + δm(v − u) − mv
= mδv − δmu − δmδv
= mδv − δmu
11
Toby Adkins CP2
Suppose we have a rocket that converts mass directly into energy and ejects it from the
back of the rocket in the form of photons. Initially, let us consider the problem in the rest
frame of the rocket. By the conservation of momentum:
mdv 0 + vγdmf = 0
By the conservation of energy:
dm − γdmf = 0
mdv 0 = −vdm
dv 0
m +v =0
dm
We now need to transform to the rest frame of the stationary observer. Using (3.9):
v + dv 0
v + dv = 0
1 + v dv
c2
v
≈ (v + dv 0 ) 1 − 2 dv 0
c
v 2 v 2
= v + dv 0 − 2 dv 0 − 2 dv 0
c c
v2
= v + dv 0 − 2 dv 0
c
2
v
→ dv = dv 0 1 − 2
c
12
Toby Adkins CP2
This separable equation can then be solved using partial fractions to nd the subsequent
evolution of the system.
13
2. Mechanics II
This section aims to cover some of the more advanced concepts in Mechanics, including:
• Dimensional Analysis
• Rotational Dynamics
• Lagrangian Mechanics
In all likelihood, most of the concepts covered in this section will be new to most readers.
Quite a few examples have been included to help facilitate further familiarity with said
concepts. This chapter assumes that readers are already very competent with using the
concepts elucidated in Chapter (1).
14
Toby Adkins CP2
We denote the dimensions of some quantity Q by the notation [Q]. The 'fundamental'
dimensions are:
[T ime] = T
[M etres] = L
[M ass] = M
[Charge] = Q
15
Toby Adkins CP2
Consider the example of ring a projectile from a tower of height h on the equator of the
Earth. Let the Earth have angular velocity ω . The distance taken for the projectile to
reach the ground is most easily found by considering (2.4). We know that ω̇ = 0 as the
speed of rotation of the earth is constant, and we can ignore ω × (ω × r0 ) as ω is small.
mr¨0 = F − 2mω × r˙0
ẍ 0 0 ẋ
ÿ = − −g − 2 0 × ẏ
z̈ 0 ω ż
0 ẏ
= − −g − 2ω −ẋ
0 0
16
Toby Adkins CP2
This is assuming that the radius of curvature of the earth is suciently large that the
surface of the earth can be considered to be locally at.
17
Toby Adkins CP2
From the cross product, it becomes clear that the direction of angular momentum is per-
pendicular to the plane of the instantaneous motion.
dL
= r × F + ṙ × p
dt
= r × F + ṙ × mṙ
=r×F
The rst and third terms will disappear by the denition of the centre of mass. Hence, we
obtain:
L = L0 + r cm × M v cm (2.7)
As previously stated, this means we can characterise motion as the translation of the CM
and a rotation around it.
18
Toby Adkins CP2
2
→ I = M R2
5
Figure 2.1: Coordinate system for deriving the Moment of Inertia of a Cylinder
19
Toby Adkins CP2
Ix :
M
dm = · (2πr · drdx)
πR2 l
1
→ Ix = M R 2
2
Iy :
1
dIx = dmR2
2
= 2dIy
1 1
dIz = dmR2 + dmx2
4 4
1 1
→ Iz = M R2 + M L2
4 12
Some integration steps have not been included; it is recommended that the reader works
through each of these examples on their own to get a handle on deriving moments of inertia.
Let d be a constant vector that gives the position of the parallel axis relative to the
axis through the centre of mass.
r0 = r + d
Z
2
I = r0 · dm
Z Z Z
= r2 · dm + d2 · dm + 2d · rdm
However, the third term will again disappear by the denition of the centre of mass.
Hence we obtain the Parallel Axis Theorem :
I = I cm + M d2 (2.9)
20
Toby Adkins CP2
→ Iz = Ix + Iy (2.10)
Note that this only holds true for a planar body. One consequence of this theorem
is if the body is uniform (e.g a regular pentagonal laminar), the moment of inertia
through any two orthogonal axes must be the same.
1. Imagine that a ball of radius a rolls without slipping down a ramp inclined at an
angle θ to the horizontal. Let x be the distance moved by the ball down the ramp,
and h be the vertical distance of the ball below the top of the ramp. What is the
acceleration of the ball down the ramp?
Firstly, the condition for a ball to roll without slipping is given by v = ωr, as the edge
has to move at the same velocity as the CM. Furthermore, if the ball rolls without
slipping, we know that no energy is dissipated by frictional forces. This means that
energy is conserved. We can thus write:
1
mgh = m v 2 + Iω 2
2
1 2 I 2
= m v + v
2 ma2
1
= mv 2 (k + 1)
2
I
f or k =
ma2
Re-arranging and substituting for h:
2gh
v2 =
k+1
2gx sin(θ)
=
k+1
Dierentiating with respect to time:
dv 2g sin(θ)
2v · = ·v
dt k+1
dv g sin(θ)
→ =
dt k+1
Evidently, it is clear that does not only hold for a sphere sphere, but in fact holds
for any object able to roll without slipping.
21
Toby Adkins CP2
2. Imagine that a snooker ball of radius a sits on a rough, horizontal table. At what
height does one have to hit the ball with the cue such that it moves o without slip-
ping?
22
Toby Adkins CP2
We have imposed the rolling without slipping condition. The centre of mass acceler-
ation is given by r̈ = −µg .
µmgr2 t
= vo − µgt
I
µmgr2
t µg + = vo
I
−1
µmgr2
→ to = vo · µg +
I
Thus, the distance travelled by the CM is:
1
d = vo · t − · r̈ · t2
2
1
= vo · to − · (µmg) · to 2
2
And then solve the equation numerically for d.
3. Imagine that a rod of mass m and length l is released horizontally under gravity
without resistive forces. An force of magnitude F is delivered instantaneously to
one end of the rod that causes the rod to rotate with angular frequency ωo . What
magnitude of force is required such that the rod returns exactly to its starting position?
The subsequent motion of the rod involve the motion of the centre of mass upwards
and a rotation around the centre of mass, as shown in Figure (2.4).
We want to nd the time to max height by considering the change in angular mo-
mentum on the rod:
l
F δt = Iωo
2
F δt = mvo
l
mvo = Iωo
2
ωo l
vo =
6
ωo l
→ t max height =
6g
Using the fact that I = 12 1
ml2 about the CM for a rod. Now, for it to return exactly
to it's starting position, we require it to execute an integer number of full revolutions
23
Toby Adkins CP2
in the time it takes to rise and then fall again. Mathematically, we can express this
as:
ωo l
2· ωo = 2πn
6g
6πng
ωo 2 =
l
2Iωo
F =
l
1
= mlωo
6
1 p
→ F = m 6πngl
6
Thus, there are actually an innite number of forces that can be applied as n can
take any integer value; the only thing that changes is the maximum height and
subsequently the number of rotations.
4. Consider a rod of length l and mass m lying on a at horizontal table. Suppose that
an impulse is delivered at a distance x from one end of the rod. What is the value
of x that is required such that the end of the rod experiences no initial acceleration?
This is a typical 'centre of percussion' problem.
24
Toby Adkins CP2
What is a central force? Well, a central force is one that only acts in the radial direction,
i.e F = F (r)br. Readers will already be familiar with a number of central forces, such as
the force due to Coulomb's Law or Newton's Law of Universal Gravitation :
Gm1 m2
Fg = · rb (2.11)
r2
By it's denition, a central force will always pull an object inwards toward the origin of
the central force, and the motion of objects under the inuence of these central forces is
what we typically think of as orbits. For a very simple circular orbit, the central force is
analogous to the centripetal force that keeps the object in circular orbit.
Central forces have two main properties when applied to orbital mechanics:
• Under a central force, angular momentum is conserved
L=r×p
dL
= r × ṗ + ṙ × p
dt
= r × ṗ + ṙ × mṙ
=r×F
25
Toby Adkins CP2
3. For a body orbiting a larger body of mass M with radius of orbit r and period T :
r3 GM
2
= (2.13)
T 4π 2
This can be derived by equating the gravitational (2.11) and centripetal forces for
an object performing circular motion.
mv 2 GmM
=
r r2
For a circle,
2πr
v=
T
Substituting this result in:
4π 2 r2 GM
= 2
rT 2 r
r 3 GM
→ 2 =
T 4π 2
Despite deriving this equation assuming a circular orbit, it does actually hold for any
periodic orbit (not hyperbolic orbits), though the proof for this is signicantly more
complicated.
26
Toby Adkins CP2
Note that this assumes that M m, as we have assumed the centre of mass of the two
bodies is located at the centre of the body of mass M . If not, we have to use the Reduced
Mass System, as in Section (2.4.4).
The total energy of the system is thus given by the Orbit Equation:
1 L2 GmM
E total = mṙ2 + 2
− (2.14)
2 2mr r
We have three distinct cases for the type of orbit based on the value of the energy:
• Eo - Circular motion with r = ro
27
Toby Adkins CP2
For an impulse delivered to an orbiting body, this can have an eect on the orbit in one of
two ways, depending on the direction in which the impulse is applied:
• Tangentially, L −→ L + ∆L - This causes a shift in the eective potential curve on
the graph (though the total energy remains the same), causing the radii to change.
• Radially, ṙ −→ ṙ + ∆ṙ - This causes a shift in the total energy, but the eective
potential curve remains the same. For example, the energy may shift from Eo to E1 .
When in doubt, always sketch the eective potential; seeing the energy represented graph-
ically can sometimes be the key to solving a tricky problem!
By (1.1),
m1 · r̈1 = F 12
m2 · r̈2 = F 21 = −F 12
F
r̈1 = 12
m1
F
r̈2 = − 12
m2
28
Toby Adkins CP2
Let M = m1 + m2 and µ = m1 m2
m1 +m2 (the reduced mass). Then:
r̈ = r̈2 − r̈1
1
= − · F 12
µ
GµM
→ µ · r̈ = − 2 · rb
r
We can use this equation of motion to nd the period of the orbit of the system.
GM
r̈ + ·r =0
r3 s
r3
→ T = 2π
G(m1 + m2 )
Gm2 M ϕ
=2· · cos
L 2
29
Toby Adkins CP2
We now need another expression for ∆p. Again looking at Figure (2.8), it becomes
clear that this is equal to:
ϕ
∆p = 2 · mvo · sin
2
Equating these two expressions, we arrive at the nal expression for the scattering
angle of:
ϕ Lvo bvo 2
cot = = (2.16)
2 GmM GM
Note that this is independent of the mass of the body.
1. Suppose that NASA scientists wanted to put a satellite into a circular orbit of radius
ro with a velocity vo , but instead accidentally directed the impulse at an angle θ to
the tangent of the desired orbit. What are the maximum and minimum radii of the
new orbit in terms of ro ?
In this case, the total energy of the orbit is the same, but the angular momentum is
changed. Initially, let us equate the gravitational and centripetal forces for circular
motion:
mvo 2 GmM
=
ro ro 2
GM
→ vo 2 =
ro
This is quite a useful result for simplifying expressions involving circular orbits, so
try to remember it!
30
Toby Adkins CP2
As we know the energy in each orbit is the same, we need to nd expressions for the
energy at points A and B .
1 GmM
EA = mvo 2 −
2 ro
1 GmM
=−
2 ro
1 2 L2 GmM
EB = mṙ + 2
−
2 2mr r
Equating these two expressions and letting ṙ = 0 :
1 GmM L2 GmM
− = −
2 ro 2mr2 r
1 GmM 2 L2
0= r − (GmM )r +
2 ro 2m
2 2 2
= r − (2ro )r + ro cos θ
using the results for L and vo 2 . Thus, we arrive at the quite simple expression of:
→ r = ro (1 ± sin θ)
2. A comet in a circular orbit around the Sun has speed vo and radius r0 = αR, where
R is the radius of the Earth's orbit, and α is a constant satisfying 0 < α < 1. The
comet has its velocity reduced by ∆v in a collision that does not change its initial
direction. What is the minimum ∆v required to move the comet into an orbit which
intersects that of the Earth?
31
Toby Adkins CP2
At the instant where it's velocity is reduced, it is still in circular motion. We require
that Eo = E R .
L2 GmM 1 2 L2 GmM
2
− = mṙ + 2
−
2mro ro 2 2mR R
1 1 GmM GmM
L2 2
− = −
2mro 2mR2 ro R
2 2 ro R
L = 2Gm M
ro + R
s
2
αR
L = 2Gm M
1+α
How does the expression change if the impulse ∆v is delivered in the radial direction
towards the sun?
32
Toby Adkins CP2
Notice how the results are markedly dierent depending on the way in which the
impulse is applied. In general, it is much more 'energy-ecient' to transfer between
orbits by changing the angular momentum, rather than giving the body some radial
velocity.
3. A probe of mass m is initially in circular orbit with radius r1 . Show that with two
appropriately placed impulses, the probe is able to reach a second circular orbit of
radius r2 > r1 .
This question introduced the concept of a Hohmann transfer orbit; it is one of the
lowest energy methods for transferring an object between two circular orbits.
Before we tackle this question, we need to derive a useful result. Consider the energy
of a circular orbit ro :
1 GmM
|Ec | =
2 ro
1 GmM
= mv 2 −
2 r
Rearranging for v , we arrive at the vis-viva equation:
s
2 1
v= GM − (2.17)
r a
for a semi-major axis a. Note that despite deriving it by assuming circular motion,
the result does hold true for any closed orbit. Now we are ready to tackle the problem
at hand.
We can use the impulses ∆v1 and ∆v2 , as shown in Figure (2.10), to achieve the
desired result.
33
Toby Adkins CP2
Let us calculate each of the successive impulses. The initial orbital velocity is given
by:
r
GM
v1 =
r1
At perigee, the elliptical transfer orbit has velocity:
s
2 2
v10 = GM −
r1 r1 + r2
s
2r2
= GM
r1 (r1 + r2 )
34
Toby Adkins CP2
2.5.1 Denitions
First, we need to dene a few key terms and concepts that are going to be referred to
throughout the course of this section. They are as follows:
• Generalised Coordinates - A set of parameters qk that specify the conguration of
the system. These do not have to be orthogonal
• Degrees of Freedom - The number of independent coordinates that is sucient to
uniquely describe a system
• Holonomic Constraints - Constraints on a system are described as holonomic assum-
ing that:
1. The constraints are independent
2. The system can be described by relations between general coordinate variables
and time
3. The number of generalised coordinates is reduced to the degrees of freedom of
the system
• Noether's Theorem - Every dierentiable symmetry of the action of a physical system
has a corresponding conservation law
35
Toby Adkins CP2
• Cylindrical:
1
L = m ṙ2 + r2 θ̇2 + ż 2 − U (r, θ, z)
2
• Spherical:
1
L = m ṙ2 + r2 θ̇2 + r2 sin2 θ · φ̇2 − U (r, θ, φ)
2
Now, how do we use this? The hardest part about using the Lagrangian is usually nding
the Lagrangian itself; the rest is simply handle-turning. Once we have managed to nd L,
we can nd the equations of motion of the system using the Euler-Lagrange Equations :
d ∂L ∂L
= (2.19)
dt ∂ q˙k ∂qk
Note that for each coordinate and it's time derivative, we have one Euler-Lagrange (EL)
equation. Note that the time derivative on the left-hand side is total; that is, which dif-
ferentiate the coordinates contained in the expression with respect to time, rather than
dierentiate explicitly with respect to time. Don't worry if this seems a little confusing
right now; have a read through some of the examples to becomes familiar with the process.
The EL equations actually have some interesting consequences due to Noether's Theorem
(2.5.1). Essentially, if ∂q
∂L
= 0, then the quantity ∂∂L
q˙k is conserved. This is know as
the conjugate momentum, denoted by pk , and can be things such as linear or angular
k
36
Toby Adkins CP2
Like with the Lagrangian, the Hamiltonian has a set of equations that we use to derive the
equations of motion, know as the Hamiltonian Equations :
∂H
p˙k = − (2.21)
∂qk
∂H
q˙k = (2.22)
∂pk
Note that the Hamiltonian itself is a characteristic of the motion of the system. Consider
its total derivative:
∂H ∂H ∂pk ∂H ∂qk dH
= · + · +
∂t ∂pk ∂t ∂qk ∂t dt
Using (2.20):
∂H dH
=
∂t dt
This means that if H has no explicit time dependence, it is a conserved quantity for the
system.
For small variations about a given minimum, we can either substitute qk = qo + δq and
expand (recalling that δq
qo 1 ) or we can use the concept of an eective potential:
∂Uo ∂ 2 U o qk 2
U ef f = Uo + · qk + · + ···
∂qk ∂qk 2 2!
∂ 2 Uo
q¨k = − · (qk − qo )
∂qk 2
around a minimum qo with a value of potential Uo at that point.
Let the pendulum have total mass m and moment of inertia I about the axis of
rotation. Let us nd the Lagrangian:
U = −mgl cos θ
1
KE = I θ̇2
2
1 2
→ L = I θ̇ + mgl cos θ
2
Applying the EL equation:
d
I θ̇ = −mgl sin θ
dt
I θ̈ = −mgl sin θ
37
Toby Adkins CP2
We can then substitute the moment of inertia of an appropriate body into this for-
mula.
2. Suppose that two equal masses are joined by a string that runs over two pulleys. One
of the masses is constrained to move in the vertical direction, while the other is free
to swing freely, as shown in Figure (2.12). Describe the evolution of the system for
small angle oscillations of the right-hand mass of magnitude .
Let the distance from the wheel of the pulley to the right hand mass be r. Dene
the zero of potential of the masses as the equilibrium position when both masses are
38
Toby Adkins CP2
For small oscillations, cos θ ≈ 1− θ2 . Substituting this into the r coordinate equation
2
of motion:
θ2
2
2r̈ = rθ̇ − g 1 − 1 −
2
θ 2
= rθ̇2 − g
2
g
= r (ω sin(ωt + φ))2 − ( cos(ωt + φ))2
2
2 2 1 2
= g sin (ωt + φ) − cos (ωt + φ)
2
Averaging over a couple of periods:
2 g
2r̈ =
4
2
g
r̈ =
8
As > 0, g > 0, this is a positive quantity, and so the left hand mass gradually
climbs while the right hand mass gradually falls. This is a second order eect.
3. Consider a ball of mass m rolling without slipping on the surface of a hemisphere
that is able to deform slightly. Find the normal force. At what angle will the ball lose
contact with the surface?
As the surface is able to deform, we need to take account of the potential energy of
this by a 'step' potential function V (r).
1 1
KE = mr2 θ̇2 + mṙ2
2 2
U = mgr cos θ + V (r)
1 1
→ L = mr2 θ̇2 + mṙ2 − mgr cos θ − V (r)
2 2
39
Toby Adkins CP2
The ball will lose contact with the surface the instant that Fr = 0.
0 = mg cos θ − mrθ̇2
mg cos θ = mRθ̇2
R v 2
cos θ =
g R
v2
=
Rg
By the conservation of energy from when the ball is initially at rest at the top of the
hemisphere:
1
mv 2 = mgR(1 − cos θ)
2
v 2 = 2gR(1 − cos θ)
R cos θ = 2R(1 − cos θ)
3R cos θ = 2R
2
cos θ =
3
Hence, the particle will leave the surface at θ max = cos−1 2
.
3
40
Toby Adkins CP2
4. Imagine that a point mass of mass m is attached to the rim of a disk of mass M .
What is the value of the period for small angle oscillations amount the minimum of
the motion?
Let the zero of potential be at the centre of the wheel. Now, assuming that the
wheel rolls without slipping, the position x of the centre of the wheel must satisfy
the condition x = Rθ. We can thus write the position of the smaller mass as:
r = (x − R sin θ, R(1 − cos θ))
= (R(θ sin θ), R(1 − cos θ))
ṙ = R(θ̇ − θ̇ cos θ, θ̇ sin θ)
41
Toby Adkins CP2
and above:
θ̈(M R2 ) + mgR sin(θ) = 0
mg
θ̈ + θ=0
MR
Hence the frequency of small angle oscillations about equilibrium is:
r
mg
Ω=
MR
5. A bead of mass m slides around inside a cone of semi-vertical angle α. The bead is
constrained to move on the surface of the cone. Find the frequency of small oscilla-
tions about a stable equilibrium, and nd the value for α for which this frequency is
equal to the frequency at stable equilibrium.
42
Toby Adkins CP2
To nd the frequency of stable circular motion, we need to impose the conditions
that r̈ = 0 , ṙ = 0 , r = ro .
0 = ro θ̇2 sin2 α − g cos α sin(α)
θ̇2 ro sin2 α = g cos α sin α
r
g
→ θ̇c =
ro tan α
Now for small oscillations about equilibrium, we let r = ro + δ .
L2 sin2 α
δ̈ = − g cos α sin α
m2 (ro + δ)3
L2 sin2 α
= 3 − g cos α sin α
m2 ro 3 1 + rδo
43
Toby Adkins CP2
44
3. Special Relativity
• Space-Time Diagrams
• Relativistic Dynamics
This turns out to be one of the more interesting topics of the mechanics course, as it has
consequences that not many students will have encountered or considered before. However,
some of the concepts can be dicult to grasp, so it is worth working through problems fully
to get to grips with the material. Note that for the entirety of this chapter, the symbol γ
is given by:
1
γ=q
v 2
1− c
where c is the speed of light in a vacuum, and v is the velocity of the moving body.
Sometimes, we use β = vc to simplify some expressions.
45
Toby Adkins CP2
The Michaelson-Morley experiment was designed to test this theory, using the apparatus
shown in Figure (3.1). We assume that the apparatus is moving at a speed v with respect
to the aether.
Consider the time taken for the light to travel along each of the perpendicular paths. First,
t1 along the horizontal path:
l l
t1 = +
c+v c−v
2l
= 2
c 1 − vc2
v2
2l
≈ 1+ 2
c c
46
Toby Adkins CP2
The M-M experiment thus predicted that there would be a time dierence in the arrival of
light at the detector that would cause a shift in the background interference pattern present.
However, no such shift was observed. This null result contributed to the destruction of the
aether model, and it's eventual replacement by Einstein's Postulates.
47
Toby Adkins CP2
One such consequence (more to follow in later sections) is that events that appear simul-
taneous in one frame may not be so in another. Consider the 'thought experiment' of two
observers; one located equidistant from each ends of a train moving at a relativistic speed
v , and the other on the stationary platform.
Figure 3.2: A diagram showing the lengths involved in the Train thought experiment
Lights at either end of the train (A and B) are turned on. First, consider the time taken
for the light to travel along OA (t1 ) and OB (t2 ) i.e in the rest frame of the observer inside
the train.
OA L
t1 = =
c c
t1 = t2
This means that for the observer on the train observes the lights at A and B to turn on
simultaneously. What about the observer on the platform?
L
OA0 L−v
t1 0 = = c
c c
L
OB 0 L + v
t2 0 = = c
c c
t1 6= t2
This means that the observer on the platform observes the lights turning on at dierent
times (B before A). This, it is clear that whether or not two events are simultaneous will
depend on the frame of reference from which they are being observed.
48
Toby Adkins CP2
As we have seen, our perception of length and time can change depending on the frame
that we nd ourselves in. We can relate these quantities in two frames by The Lorentz
Transformations (LT's). An event in frame S dened by (ct, x, y, z) is related to an event
(ct0 , x0 , y 0 , z 0 ) in frame S 0 by:
These can also be represented as a matrix, which is useful if we want to perform successive
LT's. It also implies that the Lorentz transformations are in some sense a 'rotation' in
space-time (remark how the matrix is symmetric).
γ −βγ 0 0
−βγ γ 0 0
Λ=
0
(3.5)
0 1 0
0 0 0 1
We have included the zero entries to impress upon the reader the importance of the fact that
the LT's apply to all directions, even though we are dealing with only the one-dimensional
case at this stage.
This to be expected, as the eects of special relativity are negligible at low velocities.
49
Toby Adkins CP2
that is an invariant quantity ; a quantity that does not change regardless of one's frame of
reference.
Two observers may nd dierent space and time separations of two events, but they will both
agree on the interval that they calcuate. The use of such invariant quantities becomes more
relevant later on when we start tackling relativistic dynamics, but we have introduced it
here so the reader becomes more familiar with it reading through the subsequent sections.
Consider the proper-time interval ∆t0 = t02 − t01 between two events as measured by a clock
that is stationary in frame S 0 , which itself is moving at a velocity v relative to frame S .
Using (3.2), the time interval between these two events as perceived in frame S is given
by:
∆t = t2 − t1
vx02 vx01
0 0
= γ t2 − 2 − γ t1 − 2
c c
v
= γ(t02 − t01 ) − γ 2 (x02 − x01 )
c
However, these two events are measured by a clock that is stationary in frame S 0 . This
means that x02 = x01 . Hence, we nd that:
∆t = γ∆t0 (3.7)
Remember when trying to apply this equation that time is dilated on objects moving relative
to your frame of reference.
Consider the proper length interval ∆x = x2 − x1 at rest in frame S . Using (3.2), the
length interval between these two events as perceived in frame S 0 is given by:
∆x = x2 − x1
= γ(x02 − vt02 ) − γ(x01 − vt01 )
= γ(x02 − x01 ) − γ(t02 − t01 )
50
Toby Adkins CP2
However, these measurements are taken at the same time in frame S 0 , meaning that t02 = t01 .
Hence we nd that:
∆x
∆x0 = (3.8)
γ
Remember that when trying to apply this equation that length is contracted on objects
moving relative to your frame of reference.
In Special Relativity, u0rel is known as the mutual velocity of A and B; this does not actu-
ally violate the limitation of the speed of light as no information travels at this velocity. If
we wanted to beam a signal between the two ships and measure their relative velocity, we
would not obtain v rel . So the question becomes: how do we add velocities in relativity?
For this, we want to consider these spacecraft U and V moving away from one another
along the x-axis with velocities u and v respectively. To nd the mutual velocity of the two
objects, we want to transform the rate of change of the distance between the two objects
in the rest frame of a stationary observer into the frame of one of the two spacecraft. For
simplicity's sake, we are going to perform this derivation assuming only one-dimensional
motion, but it can be generalised to three dimensions. Let us transform into the frame of
spacecraft U. Considering (3.2) in their dierential form:
dx0
u rel =
dt0
γ(dx − vdt)
=
γ dt − v dx
c2
dx − vdt
=
dt − v dx
c2
dx
dt −v
= v dx
1− c2 dt
But dx
dt = u. Thus, we obtain a formula for relativistic velocity addition of:
u−v
u rel = (3.9)
1 − u·v
c2
We have gone back to vector notation here to remind the reader that u and v are in fact
vector quantities.
Let us go back to our example of the spacecraft A and B moving towards one another. As
before, we found that the mutual velocity was 1.7c. Now using (3.9):
vA + vB
u rel =
1 + vAc2vB
51
Toby Adkins CP2
Note the signs have changed as the objects are moving towards one another in this problem.
0.8c + 0.9c
u rel =
1 + (0.8)(0.9)
≈ 0.988c
As we have shown, (3.9) gives very dierent results to simply adding the velocities; this is
the nature of Special Relativity!
52
Toby Adkins CP2
Event A can only eect events that lie inside it's light cone. Intervals inside the light cone
obey (∆s)2 > 0, and are referred to as 'time-like'. On the light cone itself, (∆s)2 = 0.
This is said to be 'light-like'. Elsewhere, (∆s)2 < 0, and this is said to be 'space-like'.
Objects or events that remain xed in space-time have vertical world lines, but objects or
events that move in space-time have slanted world lines. The faster the speed, the less
steep the world line. By convention, world lines that are at 45◦ represent objects or events
travelling at the speed of light. This means that no world line can be at less than 45◦ .
3.3.1 An Example
A stationary observer A res a light pulse at a mirror located at a distance d away. A
second observer B is moving at a relativistic speed v along the line of projection of the
light pulse. What is the time interval between the transmission and reception of the signal
by A from B's frame of reference?
We can solve this problem quite eectively by the use of space-time diagrams in both A's
and B's rest frame.
Let Event 1 be the point at which the light hits the mirror, and Event 2 the time the light
returns to observer A. For Event 1:
x1 = d
ct1 = cto
53
Toby Adkins CP2
For Event 2:
x2 = 0
ct2 = 2cto
x02 = −2γβcto
This means that the space-time diagram in the rest frame of B becomes:
This is because the world lines of the observer A and the mirror at a distance d appear to
be moving in the rest frame of B. Now the time interval in B's frame of reference is then
54
Toby Adkins CP2
given by:
tB = t1 + t2
γ γ
= (d − cβto ) + · 2βto
c c
d
=γ + βto
c
d d
=γ +β
c c
p
d (1 + β)2
= ·p
c (1 − β)(1 + β)
s
d 1+β
→ tB = ·
c 1−β
As you can see, the calculation is made easier to visualise and understand what is going
on when we use both the LT's and space-time diagrams in conjunction with one another.
55
Toby Adkins CP2
E = γmc2 (3.10)
p = γmv (3.11)
KE = (γ − 1)mc2 (3.12)
Consider the following:
E 2 − p2 c2 = γ 2 m2 c4 − γ 2 m2 v 2 (3.13)
v2
2 2 4
=γ m c 1− 2 (3.14)
c
= m2 c4 (3.15)
This is an invariant quantity because the rest mass m of a body can never change, making
it an important (and useful) result. We can show that it is explicitly frame invariant by
considering (3.4):
vE 2
02 02 2 2 2 2
E − p c = γ (E − vp) − γ pc −
c
2 2
2 2 2 2 v E 2 2
=γ E −p c − +v p
c2
v2
2 2 2 2
=γ E −p c 1− 2
c
= E 2 − p2 c2
= m2 c4
We can actually represent the momentum and energy of a body by four-vector momentum.
Four vector is of the form:
µ E
p = , p, 0, 0
c
Again, we have left the zeroes in to demonstrate that the momentum can be in any di-
rection, and not just along the x-axis. As it contains the information for all the energy
and momentum of a system, four-vector momentum is conserved. This means that we
can equate four-vector momenta if the momenta are written in the same frame. Like the
interval s, the dot product of a four-vector with itself yields an invariant quantity:
E2
pµ ·pµ = − p2
c2
= m2 c2
This means that we can equate the dot product of the sum of four-vector momenta, even
if the momenta are written in dierent frames. Do not worry if this seems a little confus-
ing right now; there are plenty of examples to follow that will help elucidate these concepts.
56
Toby Adkins CP2
Like with the LT's, these results reduce to the Classical results for v c. For example,
consider the kinetic energy:
KE = (γ − 1)mc2
1 v2
≈ 1+ − 1 mc2
2 c2
1
= mv 2
2
Some of you might be wondering at this point about how fast we need to be going for these
relativistic eects to matter? Well, if we again consider the kinetic energy:
KE = (γ − 1)mc2
1 v2 3 v2
≈ 1+ + − 1 mc2
2 c2 8 c4
1 3 v4
= mv 2 + m 2
2 8 c
For us to have a 1% error in kinetic energy, we require:
3 4
8β
0.01 = 1 2
2β
3
= β2
4
β ≈ 0.115
Thus, a particle has to be going at just over 10% of the speed of light for relativistic eects
to become noticeable.
3.4.1 Photons
So what happens when the particle that we are considering is a photon? By denition, a
photon has zero rest mass energy. This means that the square of the four-momentum for
any photon is always zero. Furthermore, this result leads to:
E hν h
p= = = (3.16)
c c λ
where h is Plank's Constant. This means that a photon still has momentum even though
it's rest mass is zero. This does not make sense classically, but if considered relativistically:
lim p = lim γmv
v→c v→c
= (∞)(o)c
3.4.2 Approximations
In relativity, there are a number of useful approximations that we can make to simplify
calculations.
• If a particle is very light, and E mc2 , then by (3.15):
E ≈ pc
57
Toby Adkins CP2
The wording of questions will often make it obvious when these are to be used. If you get
stuck on a question, look to see if any approximations such as these can be made.
Let the original electron have energy Ee , the photon energy Eγ and the energy of
the nal electron be Ee0 .
p
Ee 0 = me 2 c4 + pe 2 c2
Eγ = pγ c
This relationship can only hold if Eγ = 0, meaning that an isolated electron cannot
emit a photon.
2. In a similar way to the previous problem, we will pose the question: Can the collision
of a positron and electron create a single photon?
58
Toby Adkins CP2
Using four vector invariance across the laboratory (left hand side) and the centre of
mass (right hand side) frames:
(9 + E)2 − (pc)2 = (2mB c2 )2
In this case, E me c2 . This means that we can use the rst approximation in
Section (3.4.2). Thus the net momentum is given by:
9 E
p= −
c c
Substituting this result in and re-arranging:
(9 + E)2 − (9 − E)2 = (2mB c2 )2
36E = 4(5.3)2
→ E ≈ 3.12 GeV
The B 0 mesons undergo decay with a mean proper life-time of 1.5 × 10−12 s. Assum-
ing that the particles are produced at threshold energy, what is the mean distance
that the B 0 mesons travel in the lab frame before they decay?
We can substitute our result for E back in to nd the γ factor for the B 0 mesons.
This gives γ ≈ 1.14. From the lab. frame, the life-time of the particles appears much
longer; we thus have to consider the dilated life-time of the particles. Using (3.7):
τ 0 = γτ
= (1.5 × 10−12 )(1.14)
= 1.71 × 10−12 s
59
Toby Adkins CP2
We can ask one further question (last one!): What is the probability that one of the
photons decays before reaching this mean distance?
As the particles decay, they must obey the population decay equation:
dN
= −λN
dt
This has solution:
N t
= e− τ
No
where No is the initial population. The associated that a particle survives/decays
over a time to are:
to
P (Survive) = e− τ (3.17)
− tτo
P (Decay) = 1 − e (3.18)
Using to = 1.71 × 10−12 (time to travel the mean distance) and τ = 1.71 × 10−12 (the
mean life-time), we obtain:
P (Decay) ≈ 63%
This example covers a lot of the basic concepts involved in relativistic collisions that
produce decaying particles.
4. A photon of unknown energy E collides head-on with a photon of energy ε for ε E .
What is the minimum value of E required to produce an electron-positron pair?
We have left out the four-vector momenta for the electrons; we don't have to worry
about this as we can just consider them to be created in the zero-momentum (ZMF)/centre-
of-mass frame, and equate invariants as follows:
2 2
p1µ + p2µ = peµ + peµ
4Eε
4me 2 c2 =
c2
me 2 c4
→E=
ε
What is the speed of the ZMF in this case?
60
Toby Adkins CP2
As the energy of the second photon is so small (ε), we can neglect its energy when
transforming to the ZMF. This means we can easily calculate the energy in the ZMF,
and thus nd it's speed:
s
1−β
EZM F = (E + ε)
1+β
r
2 1−β
2me c ≈ E
2
2
2me c2
→β =1−2
E
61
Toby Adkins CP2
After the collision, both of the electrons have to have the same energy by the con-
servation of energy for both bodies (they have the same mass). We can thus write:
E10 0
0
p3µ= , p cos θ, p sin θ, 0
c
0
µ E2 0 0
p4 = , p cos θ, −p sin θ, 0
c
Re-arranging for θ:
E 2 + 2EEo + Eo 2 − 4Eo 2
sec2 θ =
E 2 − Eo 2
(E + 3Eo )(E − Eo )
=
(E + Eo )(E − Eo )
r
E + Eo
→ cos θ =
E + 3Eo
62
Toby Adkins CP2
3.5.1 Derivation
Let a source emitting a signal at a period τ be moving away from an observer at a speed
v . Using the LT's, the rst pulse of the signal is at t = 0, x = 0, and the second is at
t = γτ, x = vγτ from the frame of reference of the observer. The observed period is thus
given by:
x
τ 0 = γτ +
c
vγτ
= γτ +
c v
= γτ 1 +
c
s
1+β
→ τ0 = τ (3.20)
1−β
Thus, for a receding source, τ 0 > τ (known as a red-shift), and for an approaching source,
τ 0 < τ (known as a blue-shift). This can be used to determine the speed at which planets
are receding, for example.
63
Toby Adkins CP2
as the emitted photon travels at c. The two photons arriving are separated by:
DC = AD − AC
= cδt(1 − β cos θ)
Thus, the projection of the object appears to move faster than c. However, this evidently
does not violate causality as no information is actually transmitted at this speed.
64