Lectures
Lectures
Contents
1 Math Basics 1
1.1 Scalars, Vectors and Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2 Newton’s Laws 9
2.1 Air Resistance in One Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Projectile Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3 Motion in a Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4 Conservation of Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.5 Solving an Integral Numerically . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.6 Conservation of Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.7 Conservation of Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.8 Symmetries and Conservation Laws . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3 Oscillations 26
3.1 Harmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.2 Damped Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3 Sinusoidally Driven Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.4 Alternative Derivation for Driven Oscillators . . . . . . . . . . . . . . . . . . . . . . 31
3.5 Resonance Widths; the Q factor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3.6 Principal of Superposition and Periodic Forces (Fourier Transforms) . . . . . . . . . . 33
3.7 Response to Transient Force . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.8 Solving Equations of Motion Numerically . . . . . . . . . . . . . . . . . . . . . . . 38
3.9 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
1 Math Basics
Unit Vectors
Also known as basis vectors, unit vectors point in the direction of the coordinate axes, have unit norm,
and are orthogonal to one another. Sometimes this is referred to as an orthonormal basis,
1 0 0
êi · êj = δij = 0 1 0 . (1.1)
0 0 1
Here, δij is unity when i = j and is zero otherwise. This is called the unit matrix, because you
can multiply it with any other matrix and not change the matrix. The “dot” denotes the dot product,
A⃗ ·B ⃗ = A1 B1 + A2 B2 + A3 B3 = |A||B| cos θAB . Sometimes the unit vectors are called x̂, ŷ and
ẑ . Vectors can be decomposed in terms of unit vectors,
⃗
r = r1 ê1 + r2 ê2 + r3 ê3 . (1.2)
The vector components r1 , r2 and r3 might be called x, y and z for a displacement of vx , vy and vz
for a velocity.
Rotations
Here, we use rotations as an example of matrices and their operations. One can consider a different
orthonormal basis ê′1 , ê′2 and ê′3 . The same vector ⃗ r mentioned above can also be expressed in the new
basis,
r = r1′ ê′1 + r2′ ê′2 + r3′ ê′3 .
⃗ (1.3)
1
PHY 321 Lecture Notes 1 MATH BASICS
Even though it is the same vector, the components have changed. Each new unit vector ê′i can be
expressed as a linear sum of the previous vectors,
∑
ê′i = Uij êj , (1.4)
j
and the matrix U can be found by taking the dot product of both sides with êk ,
∑
êk · ê′i = Uij êk · êj
j
∑
êk · ê′i = Uij δjk = Uik . (1.5)
j
Thus, the matrix lambda has components Uij that are equal to the cosine of the angle between new
unit vector ê′i and the old unit vector êj .
ê′1 · ê1 ê′1 · ê2 ê′1 · ê3
U = ê′2 · ê1 ê′2 · ê2 ê′2 · ê3 , Uij = ê′i · êj = cos θij . (1.6)
ê′3 · ê1 ê′3 · ê2 ê′3 · ê3
Note that the matrix is not symmetric, Uij ̸= Uji . One can also look at the inverse transformation, by
switching the primed and unprimed coordinates,
∑
êi = Uij−1 ê′j , (1.7)
j
U −1 = U t . (1.8)
A tensor obeying Eq. (1.8) defines what is known as a unitary, or orthogonal, transformation.
The matrix U can be used to transform any vector to the new basis. Consider a vector
⃗
r = r1 ê1 + r2 ê2 + r3 ê3 (1.9)
= r1′ ê′1 + r2′ ê′2 + r3′ ê′3 .
This is the same vector expressed as a sum over two different sets of basis vectors. The coefficients ri
and ri′ represent components of the same vector. The relation between them can be found by taking
the dot product of each side with one of the unit vectors, êi , which gives
∑
ri = êi · ê′j rj′ . (1.10)
j
Using Eq. (1.7) one can see that the transformation of r can be also written in terms of U ,
∑
ri = Uij−1 rj′ . (1.11)
j
2
PHY 321 Lecture Notes 1 MATH BASICS
Thus, the matrix that transforms the coordinates of the unit vectors, Eq. (1.7) is the same one that
transforms the coordinates of a vector, Eq. (1.11).
Example 1.1:
Find the rotation matrix U for finding the components in the primed coordinate system given from
those in the unprimed system, given that the unit vectors in the new system are found by rotating the
coordinate system by and angle ϕ about the z axis.
Solution:
In this case
Under a unitary transformation U (or basis transformation) scalars are unchanged, whereas vectors ⃗
r
and matrices M change as
Physical quantities with no spatial indices are scalars (or pseudoscalars if they depend on right-handed
vs. left-handed coordinate systems), and are unchanged by unitary transformations. This includes
quantities like the trace of a matrix – the matrix itself had indices but none remain after performing
the trace.
Because there are no remaining indices, one expects it to be a scalar. Indeed one can see this,
−1
TrM ′ = Uij Mjm Umi (1.14)
−1
= Mjm Umi Uij
= Mjm δmj
= Mjj = TrM.
⃗ and B
• Scalar Product (or dot product): For vectors A ⃗,
⃗·B
A ⃗ = Ai Bi = |A||B| cos θAB ,
√
|A| ≡ A⃗ · A.
⃗
3
PHY 321 Lecture Notes 1 MATH BASICS
Note that the summation sign is inferred any time there are repeated indices. For example,
∑
Ai B i = A i Bi . (1.15)
⃗:
• Multiplying a matrix C times a vector A
(CA)i = Cij Aj .
⃗.
For the ith element one takes the scalar product of the ith row of C with the vector A
• Mutiplying matrices C and D :
(CD)ij = Cik Dkj ,
This means the one obtains the ij element by taking the scalar product of the ith of C with the
j th column of D .
⃗ and B
• Vector Product (or cross product) of vectors A ⃗:
C ⃗ × B,
⃗ = A ⃗
Ci = ϵijk Aj Bk .
Here ϵ is the third-rank anti-symmetric tensor, also known as the Levi-Civita symbol. It is ±1
only if all three indices are different, and is zero otherwise. The choice of ±1 depends on
whether the indices are an even or odd permutation of the original symbols. The permutation
xyz or 123 is considered to be +1. For the 27 elements,
ϵijk = −ϵikj = −ϵjik = −ϵkji (1.16)
ϵ123 = ϵ231 = ϵ312 = 1,
ϵ213 = ϵ132 = ϵ321 = −1,
ϵiij = ϵiji = ϵjii = 0.
You used cross products extensively when studying magnetic fields. Because the matrix is anti-
symmetric, switching the x and y axes (or any two axes) flips the sign. If the coordinate system
is right-handed, meaning the xyz axes satisfy x̂ × ŷ = ẑ , where you can point along the x axis
with your extended right index finger, the y axis with your contracted middle finger and the z
axis with your extended thumb. Switching to a left-handed system flips the sign of the vector
C⃗ =A ⃗×B ⃗×B
⃗ . Note that A ⃗ = −B ⃗ ×A ⃗ . The vector C ⃗ is perpendicular to both A
⃗ and B⃗
and the magnitude of C ⃗ is given by
4
PHY 321 Lecture Notes 1 MATH BASICS
For taking the dot product of the gradient with a vector, sometimes called a divergence,
⃗ ∇·A
div A, ⃗ = ∂i Ai .
For taking the vector product with another vector, sometimes called curl, ∇ × A
⃗,
⃗ (∇ × A)
curl A, ⃗ i = ϵijk ∂j Ak (⃗
r , t).
∂2 ∂2 ∂2
∇ =∇·∇=
2
+ + .
∂x2 ∂y 2 ∂z 2
Some identities
Here we simply state these, but you may wish to prove a few. They are useful for this class and will
be essential when you study E&M.
⃗ · (B
A ⃗ × C) ⃗ · (C
⃗ = B ⃗ × A)
⃗ =C ⃗ · (A
⃗ × B)
⃗ (1.17)
⃗ × (B
A ⃗ × C) ⃗ · C)
⃗ = (A ⃗ B⃗ − (A⃗ · B)
⃗ C⃗
⃗ × B)
(A ⃗ · (C⃗ × D) ⃗ · C)(
⃗ = (A ⃗ B ⃗ · D)
⃗ − (A ⃗ · D)(
⃗ B ⃗ · C)
⃗
Example 1.2:
The height of a hill is given by the formula
Here z is the height in meters and x and y are the east-west and north-south coordinates. Find the
position x, y where the hill is the highest, and give its height.
Solution: The maxima or minima, or inflection points, are given when ∂x z = 0 and ∂y z = 0.
∂x z(x, y) = 2y − 6x − 18 = 0,
∂y z(x, y) = 2x − 8y + 28 = 0.
Solving for two equations and two unknowns gives one solution x = −2, y = 3. This then gives
z = 72. Although this procedure could have given a minimum or an inflection point you can look at
5
PHY 321 Lecture Notes 1 MATH BASICS
the form for z and see that for x = y = 0 the height is lower, therefore it is not a minimum. You can
also look and see that the quadratic contributions are negative so the height falls off to −∞ far away
in any direction. Thus, there must be a maximum, and this must be it. Deciding between maximum or
minimum or inflection point for a general problem would involve looking at the matrix ∂i ∂j z at the
specific point, then finding the eigenvalues of the matrix and seeing whether they were both positive
(minimum) both negative (maximum) or one positive and one negative (saddle point).
Stoke’s law can be understood by considering a small rectangle, −∆x < x < ∆x, −∆y < y <
∆y . The path integral around the edges is
∫
ℓ·A
d⃗ ⃗ = 2∆y[Ay (∆x, 0) − Ay (−∆x, 0)] − 2∆x[Ax (0, ∆y) − Ax (0, −∆y)] (1.18)
C
{ }
Ay (∆x, 0) − Ay (−∆x, 0) Ax (0, ∆y) − Ax (0, −∆y)
= 4∆x∆y −
2∆x 2∆y
{ }
∂Ay ∂Ax
= 4∆x∆y −
∂x ∂y
= ∆S · ∇ × A.
⃗ (1.19)
Some Notation
From here on in, we may use bold face to denote vectors, e.g. v instead of ⃗
v . We also might use dots
over quantities to represent time derivatives, e.g. v̇ = dv/dt. As mentioned above, repeated indices
infer sums, e.g.,
∑
xi yi = x·y
xi yi = ⃗ ⃗, (1.20)
i
∑
xi Aij yj = xi Aij yj ,
ij
1.2 Exercises
1. All physicists must become comfortable with thinking of oscillatory and wave mechanics in
terms of expressions that include the form eiωt .
6
PHY 321 Lecture Notes 1 MATH BASICS
(a) Perform Taylor expansions in powers of ωt of the functions cos ωt and sin ωt.
(b) Perform a Taylor expansion of eiωt .
(c) Using parts (a) and (b) show that eiωt = cos ωt + i sin ωt.
(d) Show that ln(−1) = iπ .
2. One of the many uses of the scalar product is to find the angle between two given vectors. Find
the angle between the vectors ⃗
b = (1, 2, 4) and ⃗
c = (4, 2, 1) by evaluating their scalar product.
d d⃗
r d⃗
s
r ·⃗
(⃗ s) = ·⃗ r·
s+⃗ .
dt dt dt
4. Multiply the rotation matrix in Example 1.1 by its transpose to show that the matrix is unitary
or orthogonal, i.e. you get the unit matrix.
5. Find the matrix for rotating a coordinate system by 90 degrees about the x axis.
6. Consider a parity transformation which reflects about the x = 0 plane. Find the matrix that
performs the transformation. Find the matrix that performs the inverse transformation.
7. Show that the scalar product of two vectors is unchanged if both undergo the same rotation. Use
−1
the fact that the rotation matrix is unitary, Uab = Uba .
9. Show that ∑
ϵijk ϵklm = δil δjm − δim δjl .
k
10. Consider a cubic volume V = L3 defined by 0 < x < L, 0 < y < L and 0 < z < L.
⃗ that depends arbitrarily on x, y, z . Show how Gauss’s law,
Consider a vector A
∫ ∫
dv∇ · A
⃗= ⃗ · A,
dS ⃗
V S
∫b
is satisfied by direct integration. I.e., you should use the fact that a
dx (d/dx)f (x) = f (b) −
f (a).
11. Consider the function z = 3x2 − 4y 2 + 12xy − 6x + 24. Find any maxima or minima and
determine whether it is a maximum or a minimum or an inflection point.
12. A real n−dimensional symmetric matrix λ can always be diagonalized by a unitary transfor-
mation, i.e. there exists some unitary matrix U such that,
λ̃11 0 · · · 0
0 λ̃
−1 22 · · · 0
Uij λjk Ukm = λ̃im = . . .. . (1.21)
.. .. .
0 · · · · · · λ̃nn
7
PHY 321 Lecture Notes 1 MATH BASICS
The values λ̃ii are referred to as eigenvalues. The set of n eigenvalues are unique, but their
ordering is not – there exists a unitary transformation that permutes the indices.
Consider a function f (x1 , · · · , xn ) that has the property,
∂i f (⃗
x)|⃗x=0 = 0, (1.22)
for all i. Show that if this function is a minimum, and not a maximum or an inflection point,
that the n eigenvalues of the matrix
λij ≡ ∂i ∂j f (⃗
x)|⃗x=0 , (1.23)
must be positive.
13. (a) For the unitary matrix U that diagonalizes λ as shown in the previous problem. Show that
each row of the unitary matrix represents an orthogonal unit vector by using the definition
of a unitary matrix.
(b) Show that the vector
(k)
xi ≡ Uki = (Uk1 , Uk2 , · · · , Ukn ), (1.24)
These vectors are known as eigenvectors, as they have the property that when multiplied
by λ the resulting vector is proportional ( same direction) as the original vector. Because
one can transform to a basis, using U , where λ is diagonalized, in the new basis the eigen-
vectors are simply the unit vectors.
8
PHY 321 Lecture Notes 2 NEWTON’S LAWS
2 Newton’s Laws
In this chapter we mainly consider the motion of a single particle moving under the influence of some
set of forces. For Sec.s 2.1-2.3 we will consider some problems where the force does not depend on the
position. In that case Newton’s law m⃗ v˙ = F⃗ (⃗
v ) is a first-order differential equation and one solves
for ⃗
v (t), then moves on to integrate ⃗ v to get the position. In section 2.4 we consider the case where
the force depends on the position and we show how to use energy conservation to derive trajectories,
e.g. x(t), though as shown in Sec. 2.5, this may require a numerical solution. The role of momentum
conservation in solving certain classes of problems is presented in Sec. 2.6 and the role of angular
momentum is shown in Sec. 2.7.
dV = Avdt, (2.2)
where A is the cross-sectional area and vdt is the distance the object moved in time dt. Plugging this
into the expression above,
dP
= −ρm Av 2 . (2.3)
dt
This is the force felt by the particle, and is opposite to its direction of motion. Now, because air doesn’t
stop when it hits an object, but flows around the best it can, the actual force is reduced by a dimen-
sionless factor cW , called the drag coefficient.
9
PHY 321 Lecture Notes 2 NEWTON’S LAWS
For a particle with initial velocity v0 , one can separate the dt to one side of the equation, and move
everything with v s to the other side of Eq. (2.5), then integrate,
dv cW ρm A
2
= −Bdt, B ≡ . (2.6)
∫ vf v ∫ tf m
dv
2
= −B dt = −Btf ,
v0 v
( ) t0
1 1
− = −B(tf − t0 ),
v0 vf
v = [1/v0 + B(t − t0 )]−1 .
We dropped the subscript f , i.e. tf → t and vf = v . One can see that v(t = t0 ) = v0 , and that
v(t = ∞) = 0, as expected because the drag force should ultimately stop an object unless there is
an additional force such as gravity.
Example 2.1:
Consider a ball dropped from rest. Let B = cW ρm A/m represent the effect of the drag force as
above.
1. Solve for the terminal velocity Solution: This is where the acceleration is zero. The equations
of motion are
dv
= −g + Bv 2 ,
dt
with the initial velocity set to zero. Because the motion is downward, the drag force is positive.
The terminal velocity is that for which the acceleration is zero,
√
vt = g/B.
dv
−Bdt = ,
(g/B) − v 2
∫
1 v/vt 1 1
−Bt = dx = tanh−1 (v/vt ),
vt 0 1 − x2 vt
v = −vt tanh(Bvt t) = −vt tanh(gt/vt ).
as expected. In addition to dimensional analysis, such tests provide useful checks for answers.
For many systems, e.g. an automobile, there are multiple sources of resistance. In addition to wind
resistance, where the force is proportional to v 2 , there are dissipative effects of the tires on the pave-
ment, and in the axel and drive train. These other forces can have components that scale proportional
to v , and components that are independent of v . Those independent of v , e.g. the usual f = µK N
10
PHY 321 Lecture Notes 2 NEWTON’S LAWS
frictional force you consider in Physics I, only set in once the object is actually moving. As speeds be-
come higher, the v 2 components begin to dominate relative to the others. For automobiles at freeway
speeds, the v 2 terms are largely responsible for the loss of efficiency. To travel a distance L at fixed
speed v , the energy/work required to overcome the dissipative forces are f L, which for a force of the
form f = αv n becomes
∫
W = dx f = αv n L. (2.7)
For n = 0 the work is independent of speed, but for the wind resistance, where n = 2, slowing down
is essential if one wishes to reduce fuel consumption. It is also important to consider that engines are
designed to be most efficient at a chosen range of power output. Thus, some cars will get better mileage
at higher speeds (They perform better at 50 mph than at 5 mph) despite the considerations mentioned
above.
11
PHY 321 Lecture Notes 2 NEWTON’S LAWS
The reason the drag force was chosen proportional to v , was not to be realistic, but to make it simple to
consider the x and y components separately. Whereas vx starts at some value and decays exponentially
to zero, vy decays exponentially to the terminal velocity, vt = −g/γ .
Although this direct integration is simpler than the method we invoke below, the method below will
come in useful for some slightly more difficult differential equations in the future. The differential
equation for vx is straight-forward to solve. Because it is first order there is one arbitrary constant, A,
and by inspection the solution is
vx = Ae−γt . (2.11)
The arbitrary constants for equations of motion are usually determined by the initial conditions, or
more generally boundary conditions. By inspection A = v0x , the initial x component of the velocity.
The differential equation for vy is a bit more complicated due to the presence of g . Differential equa-
tions where all the terms are linearly proportional to a function, in this case vy , or to derivatives of the
function, e.g., vy , dvy /dt, d2 vy /dt2 · · · , are called linear differential equations. If there are terms
proportional to v 2 , as would happen if the drag force were proportional to the square of the velocity,
the differential equation is not longer linear. Because this expression has only one derivative in v it
is a first-order linear differential equation. If a term were added proportional to d2 v/dt2 it would be
a second-order differential equation. In this case we have a term completely independent of v , the
gravitational acceleration g , and the usual strategy is to first rewrite the equation with all the linear
terms on one side of the equal sign,
dvy
+ γvy = −g. (2.12)
dt
Now, the solution to the equation can be broken into two parts. Because this is a first-order differential
equation we know that there will be one arbitrary constant. Physically, the arbitrary constant will be
determined by setting the initial velocity, though it could be determined by setting the velocity at any
given time. Like most differential equations, solutions are not “solved”. Instead, one guesses at a form,
then shows the guess is correct. For these types of equations, one first tries to find a single solution, i.e.
one with no arbitrary constants. This is called the particular solution, yp (t), though it should really be
called “a” particular solution because there are an infinite number of such solutions. One then finds a
solution to the homogenous equation, which is the equation with zero on the right-hand side,
dvy,h
+ γvy,h = 0. (2.13)
dt
Homogenous solutions will have arbitrary constants.
The particular solution will solve the same equation as the original general equation, Eq. (2.12),
dvy,p
+ γvy,p = −g. (2.14)
dt
However, we needn’t find one with arbitrary constants. Hence, it is called a “particular” solution.
The sum of the two,
vy = vy,p + vy,h , (2.15)
is a solution of the total equation because of the linear nature of the differential equation. One has now
found a general solution encompassing all solutions, because it both satisfies the general equation (like
the particular solution), and has an arbitrary constant that can be adjusted to fit any initial condition
(like the homogneous solution). If the equation were not linear, e.g if there were a term such as vy2 or
vy v̇y , this technique would not work.
12
PHY 321 Lecture Notes 2 NEWTON’S LAWS
Returning to the example above, the homogenous solution is the same as that for vx , because there
was no gravitational acceleration in that case,
Note that this is the terminal velocity of a particle falling from a great height. The general solution is
thus,
vy = Be−γt − g/γ, (2.18)
and one can find B from the initial velocity,
Plugging in the expression for B into Eq. (2.18) gives the y motion given the initial velocity,
It is easy to see that this solution has vy = v0y when t = 0 and vy = −g/γ when t → ∞.
One can also integrate the two equations to find the coordinates x and y as functions of t,
∫ t
v0x ( )
x = dt′ v0x (t′ ) =
1 − e−γt , (2.21)
0 γ
∫ t
gt v0y + g/γ ( )
y = dt′ v0y (t′ ) = − + 1 − e−γt .
0 γ γ
If the question was to find the position at a time t, we would be finished. However, the more common
goal in a projectile equation problem is to find the range, i.e. the distance x at which y returns to zero.
For the case without a drag force this was much simpler. The solution for the y coordinate would have
been y = v0y t − gt2 /2. One would solve for t to make y = 0, which would be t = 2v0y /g ,
then plug that value for t into x = v0x t to find x = 2v0x v0y /g = v0 sin(2θ0 )/g . One follows the
same steps here, except that the expression for y(t) is more complicated. Searching for the time where
y = 0, Eq. (2.21) gives
v0y + g/γ (
gt )
0=− + 1 − e−γt . (2.22)
γ γ
This cannot be inverted into a simple expression t = · · · . Such expressions are known as “transcen-
dental equations”, and are not the rare instance, but are the norm. In the days before computers, one
might plot the right-hand side of Eq. (2.22) graphically as a function of time, then find the point where
it crosses zero.
Now, the most common way to solve for an equation of type Eq. (2.22) would be to apply Newton’s
method numerically. This involves the following algorithm for finding solutions of some equation
F (t) = 0.
13
PHY 321 Lecture Notes 2 NEWTON’S LAWS
3. Unless you guessed perfectly, F ̸= 0, and assuming that ∆F ≈ F ′ ∆t, one would choose
If the F (t) were perfectly linear in t, one would find t in one step. Instead, one typically finds a value
of t that is closer to the final answer than tguess . One breaks the loop once one finds F within some
acceptable tolerance of zero. A program to do this might look like:
#include <cstdlib>
#include <cstdio>
void CalcF(double t,double &F,double &Fprime);
using namespace std;
int main(){
int ntries=0;
double tguess,tolerance=1.0E-8,F,Fprime,delt;
printf(”Enter tguess: ”);
scanf(”%lf”,&tguess);
CalcF(tguess,F,Fprime);
do{
delt=-F/Fprime;
if(fabs(delt)>0.5*fabs(tguess))
delt=0.5*tguess*delt/fabs(delt);
tguess=tguess+delt;
CalcF(tguess,F,Fprime);
ntries+=1;
}while(fabs(F)>tolerance && ntries<20);
if(ntries<20)
printf(”Solution: t=%g\n”,tguess);
else
printf(”Failed to converge after 20 tries\n”);
return 0;
}
IN CLASS EXERCISE
Write a program that solves for the range of a particle’s motion given its initial velocity v0 , initial
direction θ0 (in degrees) and a damping factor γ . Have the program prompt you to enter these three
inputs. The program should use Newton’s method to find the time at which y = 0. As a first guess
for the time, you might try tguess = 2v0y /g . In this case Newton’s method would need the function
F to be replaced by y(t) given in Eq. (2.20) and F ′ would be vy .
14
PHY 321 Lecture Notes 2 NEWTON’S LAWS
Fx = qBvy (2.24)
Fy = −qBvx .
One can solve the equations by taking time derivatives of either equation, then substituting into the
other equation,
One can integrate the equations to find the positions as a function of time,
∫ x ∫ t
x − x0 = dx = dtv(t) (2.28)
x0 0
−A
= cos(ωc t + ϕ),
ωc
A
y − y0 = sin(ωc t + ϕ).
ωc
The trajectory is a circle centered at x0 , y0 with amplitude A rotating in the clockwise direction.
The equations of motion for the z motion are
v˙z = 0, (2.29)
which leads to
z − z0 = Vz t. (2.30)
Added onto the circle, the motion is helical.
Note that the kinetic energy,
1 1
T = m(vx2 + vy2 + vz2 ) = m(ωc2 A2 + Vz2 ), (2.31)
2 2
15
PHY 321 Lecture Notes 2 NEWTON’S LAWS
is constant. This is because the force is perpendicular to the velocity, so that in any differential time
element dt the work done on the particle F ⃗ · dr = dtF⃗ · v = 0.
One should think about the implications of a velocity dependent force. Suppose one had a constant
magnetic field in deep space. If a particle came through with velocity v0 , it would undergo cyclotron
motion with radius R = v0 /ωc . However, if it were still its motion would remain fixed. Now, suppose
an observer looked at the particle in one reference frame where the particle was moving, then changed
their velocity so that the particle’s velocity appeared to be zero. The motion would change from circular
to fixed. Is this possible?
The solution to the puzzle above relies on understanding relativity. Imagine that the first observer
believes B⃗ ̸= 0 and that the electric field E ⃗ = 0. If the observer then changes reference frames
by accelerating to a velocity ⃗ ⃗ and E
v , in the new frame B ⃗ both change. If the observer moved to the
frame where the charge, originally moving with a small velocity v , is now at rest, the new electric
v×B
field is indeed ⃗ ⃗ , which then leads to the same acceleration as one had before. If the velocity is
√
not small compared to the speed of light, additional γ factors come into play, γ = 1/ 1 − (v/c)2 .
Relativistic motion will not be considered in this course.
F r ) = −∇U (⃗
⃗ (⃗ r ). (2.32)
allows one to solve for the one-dimensional trajectory x(t), by finding t(x),
∫ ∫
x
dx′ x
dx′
t= = √ . (2.36)
x0 v(x′ ) x0 2(E − U (x′ ))/m
16
PHY 321 Lecture Notes 2 NEWTON’S LAWS
Note this would be much more difficult in higher dimensions, because you would have to determine
which points, x, y, z , the particles might reach in the trajectory, whereas in one dimension you can
typically tell by simply seeing whether the kinetic energy is positive at every point between the old
position and the new position.
Example 2.2:
Consider a simple harmonic oscillator potential, U (x) = kx2 /2, with a particle emitted from x = 0
with velocity v0 . Solve for the trajectory t(x),
∫ x
dx′
t = √ (2.37)
2(E − kx2 /2)/m
0
√ ∫ x
dx′
= m/k √ , x2max = 2E/k.
0 x 2
max
− x ′2
Here E = mv02 /2 and xmax is defined as the maximum displacement before the particle turns around.
This integral is done by the substitution sin θ = x/xmax .
where ∆x = (x − x0 )/N and xn = x0 + (n − 1/2)∆x. Note that for best accuracy the value of
xn has been placed in the center of the nth interval. The accuracy will improve for higher values of
N , or equivalently, smaller step size ∆x. The pertinent part of a C++ program might look something
like:
#include <cstdlib>
#include <cstdio>
using namespace std;
int main(){
int n,N=100;
double Deltax,x,x_n,t=0.0,E,mass=2.0;
double x0,xf,v0; // intial and final positions and initial velocity
17
PHY 321 Lecture Notes 2 NEWTON’S LAWS
If two bodies exert forces on each other, these forces are equal in magnitude and opposite in
direction.
⃗ij , then
This means that for two bodies i and j , if the force on i due to j is called F
⃗ij = −F
F ⃗ji . (2.40)
⃗ = m⃗
Newton’s second law, F a, can be written for a particle i as
∑
⃗i =
F ⃗ij = mi⃗
F ai , (2.41)
j̸=i
where F⃗i (a single subscript) denotes the net force acting on i. Because the mass of i is fixed, one can
see that
d ∑
⃗i =
F mi⃗
vi = ⃗ij .
F (2.42)
dt j̸=i
Now, one can sum over all the particles and obtain
d ∑ ∑
mi vi = ⃗ij
F (2.43)
dt i ij,i̸=j
= 0.
The last step made use of the fact that for every term ij , there is an equivalent term ji with opposite
force. Because the momentum is defined as m⃗ v , for a system of particles,
d ∑
vi = 0, for isolated particles.
mi ⃗ (2.44)
dt i
18
PHY 321 Lecture Notes 2 NEWTON’S LAWS
By ”isolated” one means that the only force acting on any particle i are those originating from other
particles in the sum, i.e. “no external” forces. Thus, Newton’s third law leads to the conservation of
total momentum,
∑
⃗ =
P mi⃗
vi , (2.45)
i
d
⃗ = 0.
P
dt
Example 2.3:
(Note that in the textbook, the rocket problem is discussed much later, along with collisions) Consider
a rocket of mass M0 that ejects gas from the bottom of the rocket with a speed ve relative to the rocket.
The rocket’s mass reduces as the fuel is ejected by a rate α = −dM /dt. Find the speed of the rocket
as a function of time in terms of the initial mass and α. Neglect gravity.
Consider the rocket of mass M moving with velocity v . After a brief instant, the velocity of the rocket
is v + ∆v and the mass is M − ∆M . Momentum conservation gives
M v = (M − ∆M )(v + ∆v) + ∆M (v − ve )
0 = −∆M v + M ∆v + ∆M (v − ve ),
0 = M ∆v − ∆M ve .
In the second step we ignored the term ∆M ∆v because it is doubly small. The last equation gives
ve
∆v = ∆M, (2.46)
M
dv ve dM
= .
dt M dt
Integrating the expression with lower limits v0 = 0 and M0 , one finds
∫ M
dM ′
v = ve
M′ M0
v = −ve ln(M /M0 )
= −ve ln[(M0 − αt)/M0 ].
Because the total momentum of an isolated system is constant, one can also quickly see that the center
of mass of an isolated system is also constant. The center of mass is the average position of a set of
masses weighted by the mass, ∑
mi xi
x̄ = ∑i . (2.47)
i mi
The rate of change of x̄ is
1 ∑ 1
x̄˙ = mi ẋi = Px . (2.48)
M i
M
Thus if the total momentum is constant the center of mass moves at a constant velocity, and if the total
momentum is zero the center of mass is fixed.
19
PHY 321 Lecture Notes 2 NEWTON’S LAWS
where r̂ is a unit vector pointing outward from the origin. The angular momentum is defined as
L r×p
⃗ =⃗ r×⃗
⃗ = m⃗ v. (2.50)
⃗
dL
v ×⃗
= m⃗ v + m⃗ v˙
r×⃗ (2.51)
dt
v ×⃗
= m⃗ r×F
v+⃗ ⃗ = 0.
For any term i, there are two contributions. For example, for i denoting the x direction, either j denotes
the y direction and k denotes the z direction, or vice versa, so
⃗
dL
r×F
=⃗ ⃗ ≡⃗
τ, (2.54)
dt
where ⃗
τ is the torque.
For a system of isolated particles, one can write
d ∑ ∑
⃗i =
L ri × F
⃗ ⃗ij (2.55)
dt i i̸=j
1∑
= ri × F
⃗ rj × F
⃗ij + ⃗ ⃗ji
2 i̸=j
1∑
= ri − ⃗
(⃗ rj ) × F
⃗ij = 0,
2 i̸=j
where the last step used Newton’s third law, F ⃗ij = −F ⃗ji . If the forces between the particles are
⃗ij || (⃗
radial, i.e. F ri − ⃗
rj ), then each term in the sum is zero and the net angular momentum is fixed.
Otherwise, you could imagine an isolated system that would start spinning spontaneously.
20
PHY 321 Lecture Notes 2 NEWTON’S LAWS
One can write the torque about a given axis, which we will denote as ẑ , in polar coordinates, where
One can use the chain rule to write the partial derivative w.r.t. ϕ (keeping r and θ fixed),
∂x ∂y ∂z
∂ϕ = ∂x + ∂y +
∂z (2.58)
∂ϕ ∂ϕ ∂ϕ
= −r sin θ sin ϕ∂x + sin θ cos ϕ∂y .
Thus, if the potential is independent of the azimuthal angle ϕ, there is no torque about the z axis and
Lz is conserved.
Thus, if the potential is unchanged by a translation of the coordinate system, the total momentum is
conserved. If the potential is translationally invariant in a given direction, defined by a unit vector, ϵ̂
in the ⃗
ϵ direction, one can see that
ϵ̂ · ∇i V (⃗
ri ) = 0. (2.61)
21
PHY 321 Lecture Notes 2 NEWTON’S LAWS
The component of the total momentum along that axis is conserved. This is rather obvious for a single
particle. If V (⃗
r ) does not depend on some coordinate x, then the force in the x direction is Fx =
−∂x V = 0, and momentum along the x direction is constant.
We showed how the total momentum of an isolated system of particle was conserved, even if the
particles feel internal forces in all directions. In that case the potential energy could be written
∑
V = ri − ⃗
Vij (⃗ rj ). (2.62)
i,j≤i
ri → ⃗
In this case, a translation leads to ⃗ ϵ, with the translation equally affecting the coordinates of
ri +⃗
each particle. Because the potential depends only on the relative coordinates, δV is manifestly zero.
If one were to go through the exercise of calculating δV for small ⃗ ϵ in Eq. (2.60), one would find that
the term ∇i V (⃗ ri − ⃗
rj ) would be canceled by the term ∇j V (⃗ ri − ⃗rj ).
The relation between symmetries of the potential and conserved quantities (also called constants of
motion) is one of the most profound concepts one should gain from this course. It plays a critical
role in all fields of physics. This is especially true in quantum mechanics, where a quantity A is
conserved if its operator commutes with the Hamiltonian. For example if the momentum operator
−iℏ∂x commutes with the Hamiltonian, momentum is conserved, and clearly this operator commutes
if the Hamiltonian (which represents the total energy, not just the potential) does not depend on x.
Also in quantum mechanics the angular momentum operator is Lz = −iℏ∂ϕ . In fact, if the potential
is unchanged by rotations about some axis, angular momentum about that axis is conserved. We return
to this concept, from a more formal perspective, later in the course when Lagrangian mechanics is
presented.
Example 2.4:
Consider a particle of mass m moving according to the potential
{ }
x2 + z 2
V (x, y, z) = A exp − .
2a2
Of the quantities, E, px , py , pz , Lx , Ly , Lz are conserved?
Solution:
One has both rotational and translational invariance about the y axis, so Ly and py are conserved.
Because the potential does not depend explicitly on time, energy is also conserved.
2.9 Exercises
1. Consider a bicyclist with air resistance proportional to v 2 and rolling resistance proportional to
v , so that
dv
= −Bv 2 − Cv.
dt
If the cyclist has initial velocity v0 and is coasting on a flat course, a) find her velocity as a func-
tion of time, and b) find her position as a function of time.
2. For Eq. (2.22) show that in the limit where γ → 0 one finds t = 2v0y /g .
22
PHY 321 Lecture Notes 2 NEWTON’S LAWS
3. The motion of a charged particle in an electromagnetic field can be obtained from the Lorentz
equation. If the electric field vector is E and the magnetic field is B, the force on a particle of
mass m that carries a charge q and has a velocity v
⃗ = qE
F v×B
⃗ + q⃗ ⃗
(a) If there is no electric field and if the particle enters the magnetic field in a direction per-
pendicular to the lines of magnetic flux, show that the trajectory is a circle with radius
mv v
r= = ,
qB ωc
qEz
z(t) = z0 + ż0 t + t2 ,
2m
where
z(0) ≡ z0 and ż(0) ≡ ż0 .
(c) Continue the calculation and obtain expressions for ẋ(t) and ẏ(t). Show that the time
averages of these velocity components are
Ey
⟨ẋ⟩ = , ⟨ẏ⟩ = 0.
B
I.e. show that the motion is periodic and then average over one complete period. Hint:
write equations of motion for vy and for ṽx ≡ vx − Ey /B .
(d) Integrate the velocity equations found in (c) and show (with the initial conditions x(0) =
−A/ωc , ẋ(0) = Ey /B, y(0) = 0, ẏ(0) = A that
−A Ey A
x(t) = cos ωc t + t, y(t) = sin ωc t.
ωc B ωc
These are the parametric equations of a trochoid. Sketch the projections of the trajectory
on the xy -plane for the cases (i) A > |Ey /B|, (ii) A < |Ey /B|, and (iii) A = |Ey /B|.
(The last case yields a cycloid.)
4. A particle of mass m has velocity v = α/x, where x is its displacement. Find the force F (x)
responsible for the motion.
5. A particle is under the influence of a force F = −kx + kx3 /α2 , where k and α are con-
stants and k is positive. Determine U (x) and discuss the motion. What happens when E =
(1/4)kα2 ?
23
PHY 321 Lecture Notes 2 NEWTON’S LAWS
6. Using Eq. (2.36) find the position as a function of a time by numerical √integration for the case
where a particle of mass m moves under the potential U (x) = U0 x/L and for an initial
velocity v0 . Use the following values: m = 2.5kg , v0 = 75.0, L = 10m, U0 =15 J. Solve
for time until the particle returns to the origin given that it started at the origin moving with
velocity v0 . Make a graph of x vs t for the entire trajectory from your computer output. Turn in
the graph and a printout of your program.
7. Consider a rocket with initial mass M0 at rest in deep space. It fires its engines which eject
mass with an exhaust speed ve relative to the rocket. The rocket loses mass at a constant rate
α = dM /dt. Find the speed of the rocket as a function of time.
8. Imagine that a rocket can be built so that the best percentage of fuel to overall mass is 0.9. Explain
the advantage of having stages.
9. Ted and his iceboat have a combined mass of 200 kg. Ted’s boat slides without friction on the
top of a frozen lake. Ted’s boat has a winch and he wishes to wind up a long heavy rope of mass
300 kg and length 100 m that is laid out in a straight line on the ice. Ted’s boat starts at rest at one
end of the rope, then brings the rope on board the ice boat at a constant rate of 0.25 m/s. After
400 seconds the rope is all aboard the iceboat.
(a) Before Ted turns on the winch, what is the position of the center of mass relative to the
boat?
(b) Immediately after Ted starts the winch, what is his speed?
(c) Immediately after the rope is entirely on the boat, what is Ted’s speed?
(d) Immediately after the rope is entirely on board, what is Ted’s displacement relative to his
original position?
(e) Immediately after the rope is entirely on board, where is the center of mass compared to
Ted’s original position?
(f) Find Ted’s velocity as a function of time.
10. Two disks are initially at rest, each of mass M , connected by a string between their centers. The
disks slide on low-friction ice as the center of the string is pulled by a string with a constant
force F through a distance d. The disks collide and stick together, having moved a distance b
horizontally. Determine the final speed of the disks just after they collide. You may want to use
the proof from No. (12).
24
PHY 321 Lecture Notes 2 NEWTON’S LAWS
11. Santa Claus is skating on the magic ice near the North Pole, which is frictionless. A massless
rope sticks out from the pole horizontally along a straight line. The rope’s original length is L0 .
Santa approaches the rope moving perpendicular to the direction of the rope and grabs the end
of the rope. The diameter of the North Pole is 2a. The rope then winds around the thin pole,
a << L0 , until Santa is half the original distance, L0 /2, from the pole. If Santa’s original
speed was v0 , what is his new speed? Assume there is no energy wasted in curving the rope
around the pole.
12. Prove that the work on the center of mass of a system of particles during a small time interval
δt, which is defined by
δWcm ≡ F
⃗tot · δ⃗
rcm ,
is equal to the change of the kinetic energy of the center of mass,
1
Tcm ≡ 2
Mtot vcm , vcm · δ⃗
δTcm = Mtot⃗ vcm .
2
Use the following definitions:
∑
⃗tot =
F ⃗i ,
F
i
∑
Mtot = mi ,
i
1 ∑
⃗
rcm = mi⃗
ri ,
Mtot i
1 ∑
⃗
vcm = mi⃗
vi .
Mtot i
25
PHY 321 Lecture Notes 3 OSCILLATIONS
3 Oscillations
1 1 3
V (x) = V (x0 ) + (x − x0 ) ∂x V (x)|x0 + (x − x0 )2 ∂x2 V (x)x0 + ∂x V (x)x0 + · · · (3.1)
2 3!
If the position x0 is at the minimum of the resonance, the first two non-zero terms of the potential are
1
V (x) ≈ V (x0 ) + (x − x0 )2 ∂x2 V (x)x0 , (3.2)
2
1
= V (x0 ) + k(x − x0 )2 , k ≡ ∂x2 V (x)x0 ,
2
F = −∂x V (x) = −k(x − x0 ).
Here A and ϕ are arbitrary. Equivalently, one could have written this as A cos(ω0 t) + B sin(ω0 t),
or as the real part of Aeiω0 t . In this last case A could be an arbitrary complex constant. Thus, there
are 2 arbitrary constants (either A and B or A and ϕ, or the real and imaginary part of one complex
constant. This is the expectation for a second order differential equation, and also agrees with the
physical expectation that if you know a particle’s initial velocity and position you should be able
to define its future motion, and that those two arbitrary conditions should translate to two arbitrary
constants.
A key feature of harmonic motion is that the system repeats itself after a time T = 1/f , where f is
the frequency, and ω = 2πf is the angular frequency. The period of the motion is independent of
the amplitude. However, this independence is only exact when one can neglect higher terms of the
potential, x3 , x4 · · · . Once can neglect these terms for sufficiently small amplitudes, and for larger
amplitudes the motion is no longer purely sinusoidal, and even though the motion repeats itself, the
time for repeating the motion is no longer independent of the amplitude.
26
PHY 321 Lecture Notes 3 OSCILLATIONS
One can also calculate the velocity and the kinetic energy as a function of time,
Example 3.1:
A pendulum is an example of a harmonic oscillator. By expanding the kinetic and potential energies
for small angles find the frequency for a pendulum of length L with all the mass m centered at the
end by writing the eq.s of motion in the form of a harmonic oscillator.
Solution: The potential energy and kinetic energies are (for x being the displacement)
x2
U = mgL(1 − cos θ) ≈ mgL ,
2L2
1 m
T = mL2 θ̇ 2 ≈ ẋ2 .
2 2
For small x Newton’s 2nd law becomes
mg
mẍ = − x,
L
and
√ the spring constant would appear to be k = mg/L, which makes the frequency equal to ω0 =
g/L. Note that the frequency is independent of the mass.
Just to make the solution a bit less messy, we rewrite this equation as
√
ẍ + 2β ẋ + ω02 x = 0, β ≡ b/2m, ω0 ≡ k/m. (3.8)
27
PHY 321 Lecture Notes 3 OSCILLATIONS
Both β and ω have dimensions of inverse time. To find solutions (see appendix C in the text) you must
make an educated guess at the form of the solution. To do this, first realize that the solution will need
an arbitrary normalization A because the equation is linear. Secondly, realize that if the form is
x = Aert (3.9)
that each derivative simply brings out an extra power of r . This means that the Aert factors out and
one can simply solve for an equation for r . Plugging this form into Eq. (3.8),
We refer to the two solutions as r1 and r2 corresponding to the + and − roots. As expected, there
should be two arbitrary constants involved in the solution,
1. Underdamped: β < ω0
√
−βt iω ′ t −βt −iω ′ t ′
x = A1 e e + A2 e e , ω ≡ ω02 − β 2 (3.13)
= (A1 + A2 )e−βt cos ω ′ t + i(A1 − A2 )e−βt sin ω ′ t.
′
Here we have made use of the identity eiω t = cos ω ′ t + i sin ω ′ t. Because the constants
are arbitrary, and because the real and imaginary parts are both solutions individually, we can
simply consider the real part of the solution alone:
2. Critical dampling: β = ω0
In this case the two terms involving r1 and r2 are identical because ω ′ = 0. Because we need
to arbitrary constants, there needs to be another solution. This is found by simply guessing, or
by taking the limit of ω ′ → 0 from the underdamped solution. The solution is then
The critically damped solution is interesting because the solution approaches zero quickly, but
does not oscillate. For a problem with zero initial velocity, the solution never crosses zero. This
is a good choice for designing shock absorbers or swinging doors.
28
PHY 321 Lecture Notes 3 OSCILLATIONS
3. Overdamped: β > ω0
√ √
x = A1 e−(β+ β 2 −ω02 )t
+ A2 e−(β− β 2 −ω02 )t
(3.16)
This solution will also never pass the origin more than once, and then only if the initial velocity
is strong and initially toward zero.
Example 3.2:
Given b, m and ω0 , find x(t) for a particle whose initial position is x = 0 and has initial velocity v0
(assuming an underdamped solution).
Solution: The solution is of the form,
Consider a single solution with no arbitrary constants, which we will call a particular solution, xp (t).
It should be emphasized that this is A particular solution, because there exists an infinite number of
such solutions because the general solution should have two arbitrary constants. Now consider solu-
tions to the same equation without the driving term, which include two arbitrary constants. These are
called either homogenous solutions or complementary solutions, and were given in the previous sec-
tion, e.g. Eq. (3.14) for the underdamped case. The homogenous solution already incorporates the two
arbitrary constants, so any sum of a homogenous solution and a particular solution will represent the
general solution of the equation. The general solution incorporates the two arbitrary constants A and
B to accommodate the two initial conditions. One could have picked a different particular solution,
i.e. the original particular solution plus any homogenous solution with the arbitrary constants Ap and
Bp chosen at will. When one adds in the homogenous solution, which has adjustable constants with
arbitrary constants A′ and B ′ , to the new particular solution, one can get the same general solution
by simply adjusting the new constants such that A′ + Ap = A and B ′ + Bp = B . Thus, the choice
of Ap and Bp are irrelevant, and when choosing the particular solution it is best to make the simplest
choice possible.
29
PHY 321 Lecture Notes 3 OSCILLATIONS
{ } F0
D −ω 2 cos(ωt − δ) − 2βω sin(ωt − δ) + ω02 cos(ωt − δ) = cos(ωt). (3.20)
m
One can now use angle addition formulas to get
{
D (−ω 2 cos δ + 2βω sin δ + ω02 cos δ) cos(ωt) (3.21)
} F0
+(−ω 2 sin δ − 2βω cos δ + ω02 sin δ) sin(ωt) = cos(ωt).
m
Both the cos and sin terms need to equate if the expression is to hold at all times. Thus, this becomes
two equations
{ } F0
D −ω 2 cos δ + 2βω sin δ + ω02 cos δ = (3.22)
m
−ω 2 sin δ − 2βω cos δ + ω02 sin δ = 0.
2βω
tan δ = . (3.23)
ω02 − ω 2
Using the identities tan2 +1 = csc2 and sin2 + cos 2 = 1, one can also express sin δ and cos δ ,
2βω
sin δ = √ , (3.24)
(ω02 − ω 2 )2 + 4ω 2 β 2
(ω02 − ω 2 )
cos δ = √
(ω02 − ω 2 )2 + 4ω 2 β 2
Inserting the expressions for cos δ and sin δ into the expression for D ,
F0 /m
D= √ . (3.25)
(ω02 − ω 2 )2 + 4ω 2 β 2
For a given initial condition, e.g. initial displacement and velocity, one must add the homogenous
solution then solve for the two arbitrary constants. However, because the homogenous solutions decay
with time as e−βt , the particular solution is all that remains at large times, and is therefore the steady
state solution. Because the arbitrary constants are all in the homogenous solution, all memory of the
initial conditions are lost at large times, t >> 1/β .
The amplitude of the motion, D , is linearly proportional to the driving force (F0 /m), but also depends
on the driving frequency ω . For small β the maximum will occur at ω = ω0 . This is referred to as a
resonance. In the limit β → 0 the amplitude at resonance approaches infinity.
30
PHY 321 Lecture Notes 3 OSCILLATIONS
rather than as F0 cos ωt. The real part of F is the same as before. For the differential equation,
F0
ẍ + 2β ẋ + ω02 x = eiωt , (3.27)
m
one can treat x(t) as an imaginary function. Because the operations d2 /dt2 and d/dt are real and
thus do not mix the real and imaginary parts of x(t), Eq. (3.27) is effectively 2 equations. Because
eωt = cos ωt + i sin ωt, the real part of the solution for x(t) gives the solution for a driving force
F0 cos ωt, and the imaginary part of x corresponds to the case where the driving force is F0 sin ωt. It
is rather easy to solve for the complex x in this case, and by taking the real part of the solution, one
finds the answer for the cos ωt driving force.
We assume a simple form for the particular solution
xp = Deiωt , (3.28)
F0 /m 2βω
|D| = √ , tan δ = . (3.30)
(ω02 − ω 2 )2 + 4β 2 ω 2 ω02 − ω 2
This is the same expression for δ as before. One then finds xp (t),
(F0 /m)eiωt−iδ
xp (t) = ℜ √ (3.31)
(ω02 − ω 2 )2 + 4β 2 ω 2
(F0 /m) cos(ωt − δ)
= √ .
(ω02 − ω 2 )2 + 4β 2 ω 2
This is the same answer as before. If one wished to solve for the case where F (t) = F0 sin ωt, the
imaginary part of the solution would work
(F0 /m)eiωt−iδ
xp (t) = ℑ √ (3.32)
(ω02 − ω 2 )2 + 4β 2 ω 2
(F0 /m) sin(ωt − δ)
= √ .
(ω02 − ω 2 )2 + 4β 2 ω 2
31
PHY 321 Lecture Notes 3 OSCILLATIONS
Example 3.3:
Consider the damped and driven harmonic oscillator worked out above. Given F0 , m, β and ω0 , solve
for the complete solution x(t) for the case where F = F0 sin ωt with initial conditions x(t = 0) = 0
and v(t = 0) = 0. Assume the underdamped case.
Solution: The general solution including the arbitrary constants includes both the homogenous and
particular solutions,
F0 sin(ωt − δ)
x(t) = √ + A cos ω ′ te−βt + B sin ω ′ te−βt .
m (ω02 − ω 2 )2 + 4β 2 ω 2
√
The quantities δ and ω ′ are given earlier in the section, ω ′ = ω02 − β 2 , δ = tan−1 (2βω/(ω02 −
2
ω ). Here, solving the problem means finding the arbitrary constants A and B . Satisfying the initial
conditions for the initial position and velocity:
x(t = 0) = 0 = −η sin δ + A,
v(t = 0) = 0 = ωη cos δ − βA + ω ′ B,
F0 1
η ≡ √ .
m (ω02 − ω 2 )2 + 4β 2 ω 2
The problem is now reduced to 2 equations and 2 unknowns, A and B . The solution is
F0 /m
xp (t) = √ cos(ωt − δ), (3.34)
(ω02 − ω 2 )2 + 4ω 2 β 2
( )
−1
2βω
δ = tan .
ω02 − ω 2
√
If one fixes the driving frequency ω and adjusts the fundamental frequency ω0 = k/m, the
maximum amplitude occurs when ω0 = ω because that is when the term from the denominator
(ω02 − ω 2 )2 + 4ω 2 β 2 is at a minimum. This is akin to dialing into a radio station. However, if one
fixes ω0 and adjusts the driving frequency one minimize with respect to ω , e.g. set
d [ ]
(ω02 − ω 2 )2 + 4ω 2 β 2 = 0, (3.35)
dω
√
and one finds that the maximum amplitude occurs when ω = ω02 − 2β 2 . If β is small relative to
ω0 , one can simply state that the maximum amplitude is
F0
xmax ≈ . (3.36)
2mβω0
32
PHY 321 Lecture Notes 3 OSCILLATIONS
xmax
FWHM
2
ω
Figure 3.1: The maximum amplitude squared of the steady-state motion of a sinusoidally driven harmonic
oscillator is shown as a function of the driving frequency ω . This peaks near the fundamental frequency ω0
and the full-width-half-maximum, F W HM , is given by the damping, 2β .
Figure 3.1 displays the maximum amplitude squared as a function of ω , and one can see the peak at
ω ≈ ω0 . The squared amplitude is usually more the quantity of interest than the amplitude because
the power being absorbed by the oscillator is proportional to the square of the amplitude, not the
amplitude. The width of the peak is quantified by the full-width-half-maximum, sometimes called
F W HM and can be found by finding the frequency difference, ω − ω0 , such that the maximum
response falls by a factor of two,
4ω 2 β 2 1
= . (3.37)
(ω02 − ω 2 )2 + 4ω 2 β 2 2
For small damping this occurs when ω = ω0 ± β , so the F W HM ≈ 2β . For the purposes of tuning
to a specific frequency, one wants the width to be as small as possible. The ratio of ω0 to F W HM is
known as the quality factor, or Q factor,
ω0
Q≡ . (3.38)
2β
This is known as the principal of superposition. It only applies when the homogenous equation is
linear. If there were an anharmonic term such as x3 in the homogenous equation, then when one
∑
summed various solutions, x = ( n xn )2 , one would get cross terms. Superposition is especially
useful when F (t) can be written as a sum of sinusoidal terms, because the solutions for each sinu-
soidal term is analytic, and are given in the previous two subsections.
Driving forces are often periodic, even when they are not sinusoidal. Periodicity implies that for some
time τ
F (t + τ ) = F (t). (3.40)
33
PHY 321 Lecture Notes 3 OSCILLATIONS
One example of a non-sinusoidal periodic force is a square wave. Many components in electric circuits
are non-linear, e.g. diodes, which makes many wave forms non-sinusoidal even when the circuits are
being driven by purely sinusoidal sources.
For the sinusoidal example studied in the previous subsections the period is τ = 2π/ω . However,
higher harmonics can also satisfy the periodicity requirement. In general, any force that satisfies the
periodicity requirement can be expressed as a sum over harmonics,
f0 ∑
F (t) = + fn cos(2nπt/τ ) + gn sin(2nπt/τ ). (3.41)
2 n>0
From the previous subsection, one can write down the answer for xpn (t), by substituting fn /m or
gn /m for F0 /m into Eq.s (3.31) or (3.32) respectively. By writing each factor 2nπt/τ as nωt, with
ω ≡ 2π/τ ,
f0 ∑
F (t) = + fn cos(nωt) + gn sin(nωt). (3.42)
2 n>0
The solutions for x(t) then come from replacing ω with nω for each term in the particular solution
in Eq.s (3.19-3.25),
f0 ∑
xp (t) = + αn cos(nωt − δn ) + βn sin(nωt − δn ), (3.43)
2k n>0
fn /m
αn = √ ,
((nω)2 − ω02 ) + 4β 2 n2 ω 2
gn /m
βn = √ ,
((nω)2 − ω02 ) + 4β 2 n2 ω 2
( )
−1
2βnω
δn = tan .
ω02 − n2 ω 2
Because the forces have been applied for a long time, any non-zero damping eliminates the homoge-
nous parts of the solution, so one need only consider the particular solution for each n.
The problem will considered solved if one can find expressions for the coefficients fn and gn , even
though the solutions are expressed as an infinite sum. The coefficients can be extracted from the func-
tion F (t) by
∫ τ /2
2
fn = dt F (t) cos(2nπt/τ ), (3.44)
τ −τ /2
∫
2 τ /2
gn = dt F (t) sin(2nπt/τ ).
τ −τ /2
To check the consistency of these expressions and to verify Eq. (3.44), one can insert the expansion of
F (t) in Eq. (3.42) into the expression for the coefficients in Eq. (3.44) and see whether
∫ { }
2 τ /2 f0 ∑
fn =? dt + fm cos(mωt) + gm sin(mωt) cos(nωt). (3.45)
τ −τ /2 2 m>0
Immediately, one can throw away all the terms with gm because they convolute an even and an odd
function. The term with f0 /2 disappears because cos(nωt) is equally positive and negative over the
34
PHY 321 Lecture Notes 3 OSCILLATIONS
interval and will integrate to zero. For all the terms fm cos(mωt) appearing in the sum, one can use
angle addition formulas to see that cos(mωt) cos(nωt) = (1/2)(cos[(m+n)ωt]+ cos[(m−n)ωt].
This will integrate to zero unless m = n. In that case the m = n term gives
∫ τ /2
τ
dt cos2 (mωt) = , (3.46)
−τ /2 2
and
∫ τ /2
2
fn =? dt fn /2 (3.47)
τ −τ /2
= fn ✓.
Example 3.4:
Consider the driving force:
Find the Fourier coefficients fn and gn for all n using Eq. (3.44).
Solution: Only the odd coefficients enter by symmetry, i.e. fn = 0. One can find gn integrating by
parts,
∫ τ /2
2 At
gn = dt sin(nωt) (3.49)
τ −τ /2 τ
u = t, dv = sin(nωt)dt, v = − cos(nωt)/(nω),
∫
−2A τ /2 −t cos(nωt) τ /2
gn = dt cos(nωt) + 2A .
nωτ 2 −τ /2 nωτ 2 −τ /2
The first term is zero because cos(nωt) will be equally positive and negative over the interval. Using
the fact that ωτ = 2π ,
2A
gn = − cos(nωτ /2) (3.50)
2nπ
A
= − cos(nπ)
nπ
A
= (−1)n+1 .
nπ
The true function is compared to the expansion in Fig. 3.2 where the sum is cut off at finite n.
35
PHY 321 Lecture Notes 3 OSCILLATIONS
� ���
� ����
��������
� �
�����
����
�� � � ��
���
Figure 3.2: The periodic function F (t) = t/τ, |t| < τ /2 (black) is compared to the Fourier expansion de-
scribed in Eq. (3.49) with 10 terms (red) or 100 terms (green).
by an amount v0 = I/m while not changing the position. One can then solve the trajectory by solving
Eq. (3.14) with initial conditions v0 = I/m and x0 = 0. This gives
I
x(t) = e−βt sin ω ′ t, t > 0. (3.51)
mω ′
√
Here, ω ′ = ω02 − β 2 . For an impulse Ii that occurs at time ti the trajectory would be
Ii
x(t) = e−β(t−ti ) sin[ω ′ (t − ti )]Θ(t − ti ), (3.52)
mω ′
where Θ(t − ti ) is a step function, i.e. Θ(x) is zero for x < 0 and unity for x > 0. If there were
several impulses linear superposition tells us that we can sum over each contribution,
∑ Ii
x(t) = e−β(t−ti ) sin[ω ′ (t − ti )]Θ(t − ti ) (3.53)
i
mω ′
Now one can consider a series of impulses at times separated by ∆t, where each impulse is given by
Fi ∆t. The sum above now becomes an integral,
∫ ′
∞
′ ′
e−β(t−t ) sin[ω ′ (t − t′ )]
x(t) = dt F (t ) Θ(t − t′ ) (3.54)
mω ′
∫−∞
∞
= dt′ F (t′ )G(t − t′ ),
−∞
−β∆t
e sin[ω ′ ∆t]
G(∆t) = Θ(∆t)
mω ′
′
The quantity e−β(t−t ) sin[ω ′ (t − t′ )]/mω ′ Θ(t − t′ ) is called a Green’s function, G(t − t′ ). It
describes the response at t due to a force applied at a time t′ , and is a function of t − t′ . The step
function ensures that the response does not occur before the force is applied. One should remember
that the form for G would change if the oscillator were either critically- or over-damped.
36
PHY 321 Lecture Notes 3 OSCILLATIONS
When performing the integral in Eq. (3.54) one can use angle addition formulas to factor out the part
with the t′ dependence in the integrand,
1
x(t) = ′
e−βt [Ic (t) sin(ω ′ t) − Is (t) cos(ω ′ t)] , (3.55)
∫mω
t
′
Ic (t) ≡ dt′ F (t′ )eβt cos(ω ′ t′ ),
−∞
∫ t
′
Is (t) ≡ dt′ F (t′ )eβt sin(ω ′ t′ ).
−∞
If the time t is beyond any time at which the force acts, F (t′ > t) = 0, the coefficients Ic and Is
become independent of t.
Example 3.5:
Consider an undamped oscillator (β → 0), with characteristic frequency ω0 and mass m, that is at
rest until it feels a force described by a Gaussian form,
{ }
−t2
F (t) = F0 exp .
2τ 2
For large times (t >> τ ), where the force has died off, find x(t).
Solution: Solve for the coefficients Ic and Is in Eq. (3.55). Because the Gaussian is an even function,
Is = 0, and one need only solve for Ic ,
∫ ∞
′2 /(2τ 2 )
Ic = F0 dt′ e−t cos(ω0 t′ )
−∞
∫ ∞
′2 /(2τ 2 ) ′
= ℜF0 dt′ e−t eiω0 t
∫−∞
∞
′
dt′ e−(t −iω0 τ e−ω0 τ
2 )2 /(2τ 2 ) 2 2 /2
= ℜF0
−∞
√
= F0 τ 2πe−ω0 τ /2 .
2 2
The third step involved completing the square, and the final step used the fact that the integral
∫ ∞ √
dx e−x
2 /2
= 2π.
−∞
To see that this integral is true, consider the square of the integral, which you can change to polar
coordinates,
∫ ∞
dx e−x
2 /2
I =
∫−∞
∞
dxdy e−(x
2 +y 2 )/2
I2 =
−∞
∫ ∞
rdr e−r
2 /2
= 2π
0
= 2π.
37
PHY 321 Lecture Notes 3 OSCILLATIONS
xn = x(n∆t). (3.56)
Assume that one knows xn for all i <≤ j . One can express the velocities and accelerations in New-
ton’s equations of motion for a mass m as:
xn+1 − xn−1
vn = , (3.57)
2∆t
vn+1/2 − vn−1/2
an =
∆t
xn+1 − 2xn + xn−1
= .
(∆t)2
The equations of motion at tn = n∆t become
If one knows xn−1 and xn the only unknown in the equation is xn+1 . One can solve for it, then move
onto the next n. Note that to proceed one needs to know x at two points. This can be tricky depending
on how the initial conditions are specified.
Example 3.6:
Write a numerical program to solve the evolution of a particle of mass m feeling a spring force, −kx
and a drag force, −bv . Assume there is also some external force, F (t) = F0 sin ωt. Let the initial
conditions be x(t = 0) = 0 and v(t = 0) = v0 .
Solution: The differential equation,
mẍ + bẋ + kx = 0,
becomes
m b
(xn+1 − 2xn + xn−1 ) + (xn+1 − xn−1 ) + kxn = F (n∆t).
(∆t)2 2∆t
Solving for xn+1 ,
{ } [ ] [ ]
m b 2m m b
+ xn+1 = − k xn + F (n∆t) + − + xn−1 ,
(∆t)2 2∆t (∆t)2 (∆t)2 2∆t
[ ] [ ]
2m
(∆t)2
− k xn + F (n∆t) + − (∆t)
m
2 + b
2∆t
xn−1
xn+1 = m b
.
(∆t)2 + 2∆t
38
PHY 321 Lecture Notes 3 OSCILLATIONS
In order to iterate forward to find xn+1 , one must know xn−1 and xn . However, to get started, one
only knows one value of x, along with the velocity being zero. To get x at two time steps, one must
can use the above two pieces of information,
x1 − x−1
v0 = ,
2∆t
x1 − 2x0 + x−1
−bv0 − kx0 = m ,
∆t2
then combine the equations to eliminate x−1 to solve for x1 . The fact that x0 = 0 makes the algebra
simpler,
bv0 ∆t2
x1 = v0 ∆t − .
2m
The program might read:
3.9 Exercises
1. A floating body of uniform cross-sectional area A and of mass density ρ and at equilibrium
displaces a volume V . Show that the period of small oscillations about the equilibrium position
is given by √
τ = 2π V /gA
2. Show that the critically damped solution, Eq. (3.15), is indeed the solution to the differential
equation.
39
PHY 321 Lecture Notes 3 OSCILLATIONS
x2 x4
V (x) = −k +α .
2 4
(a) What is the angular frequency for small vibrations about the minimum of the potential?
What is the effective spring constant?
(b) If you add a small force F = F0 cos(ωt−ϕ), and if the particle is initially at the minimum
with zero initial velocity, find its position as a function of time.
(c) If there is a small drag force −bv , repeat (b).
6. A “delta” function is a function that is zero everywhere except where the argument is zero. At
this point the function is infinite so that the area under the curve is unity. The delta function
obeys the relations
∫ b
dt′ δ(t′ − t0 ) = 1,
a
∫ b
dt′ f (t′ )δ(t′ − t0 ) = f (t0 ),
a
as long as the time t0 lies between the limits a and b. Otherwise, the integrals are zero.
i.e. show that it is zero everywhere except the origin and that it integrates to unity.
(b) A step function, Θ(t), a.k.a. the “Theta” function or the Heaviside function, is zero for
negative arguments and is unity for positive arguments. Show that
d
Θ(x − x0 ) = δ(x − x0 ).
dx
(c) Using the definition of Fourier coefficients in Eq.s (3.42) and (3.44), show that
2∑
∞
1
δ(t − t0 ) = − + cos(ωn (t − t0 )), ωn = 2nπ/τ.
τ τ n=0
2∑
∞
1
f (t) = − + einω(t−t0 ) , ω = 2π/τ.
τ τ n=0
40
PHY 321 Lecture Notes 3 OSCILLATIONS
∫
(a) Using the fact that if one integrates over the interval, −τ /2 < t < τ /2, that dt einωt =
0 for n ̸= 0, show that ∫
dtf (t) = 1.
∑
(b) Using the fact that n xn = 1/(1 − x), show that
1 2/τ
f (t) = − + .
τ 1 − eiω(t−t0 )
(c) From the expression in (b), show that the real part of f (t) obeys
ℜf (t) = 0, for t ̸= t0
This shows that ℜf is a delta function and validates the result of the previous problem.
9. Consider a particle of mass m in a harmonic oscillator with angular frequency ω0 and no damp-
ing. It experiences an external force,
F (t) = f0 Θ(t)e−γt .
A “Theta” function is a step function, and is zero for negative arguments and unity for positive
arguments.
41
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
4.1 Gravity
The gravitational potential energy and forces involving two masses a and b are
Gma mb
Uab = − , (4.1)
|⃗
ra − ⃗
rb |
Gma mb
Fba = − r̂ab ,
|⃗
ra − ⃗
rb |2
rb − ⃗
⃗ ra
r̂ab = .
|⃗
ra − ⃗rb |
Here G = 6.67 × 10−11 Nm2 /kg2 , and Fba is the force on b due to a. By inspection, one can see that
the force on b due to a and the force on a due to b are equal and opposite. The net potential energy
for a large number of masses would be
∑ 1∑
U = Uab = Uab . (4.2)
a<b
2 a̸=b
Just like electrodynamics, one can define ”fields”, which for a small additional mass m are the force
per mass and the additional potential energy per mass. The gravitational field related to the force has
dimensions of force per mass, or acceleration, and can be labeled ⃗ r ). The potential energy per mass
g (⃗
has dimensions of energy per mass. This is analogous to the electromagnetic potential, which is the
potential energy per charge, and the electric field which is the force per charge.
Because the field ⃗g obeys the same inverse square law for a point mass as the electric field does for a
point charge, the gravitational field also satisfies a version of Gauss’s law,
I
⃗ ·⃗
dA g = −4πGMinside . (4.3)
B
F (r) = . (4.4)
4πr 2
Now, let F also be assigned a direction, so that it becomes a vector pointing along the direction of the
flying paint. For any surface that surrounds the nozzle, not necessarily a sphere, one can state that
I
⃗ ·F
dA ⃗ = B, (4.5)
regardless of the shape of the surface. This follows because the rate at which paint is deposited on
the surface should equal the rate at which it leaves the nozzle. The dot product ensures that only the
42
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
Example 4.1:
Consider Earth to have its mass M uniformly distributed in a sphere of radius R. Find the magnitude
of the gravitational acceleration as a function of the radius r in terms of the acceleration of gravity at
the surface g(R). Assume r < R, i.e. you are inside the surface.
Solution: Take the ratio of Eq. (4.3) for two radii, R and r < R,
The potential energy per mass is similar conceptually to the voltage, or electric potential energy per
charge, that was studied in electromagnetism, if V ≡ U /m, ⃗ g = −∇V .
GM δm GM δm
F =− +2 ∆D + · · · (4.6)
D2 D3
If the z direction points toward the large object, ∆D can be referred to as z . In the accelerating frame
of an observer at the center of the planet,
d2 z
δm = F − δma′ + other forces acting on δm, (4.7)
dt2
where a′ is the acceleration of the observer. Because δma′ equals the gravitational force on δm if it
were located at the planet’s center, one can write
d2 z GM δm
m =2 z + other forces acting on δm. (4.8)
dt2 D3
Here the other forces could represent the forces acting on δm from the spherical planet such as the
gravitational force or the contact force with the surface. If θ is the angle w.r.t. the z axis, the effective
force acting on δm is
GM δm
Feff ≈ 2 r cos θ ẑ + other forces acting on δm. (4.9)
D3
43
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
This first force is the ”tidal” force. It pulls objects outward from the center of the object. If the ob-
ject were covered with water, it would distort the objects shape so that the shape would be elliptical,
stretched out along the axis pointing toward the large mass M . The force is always along (either
parallel or antiparallel to) the ẑ direction.
Example 4.2:
Consider the Earth to be a sphere of radius R covered with water, with the gravitational acceleration at
the surface noted by g . Now assume that a distant body provides an additional constant gravitational
acceleration ⃗
a pointed along the z axis. Find the distortion of the radius as a function of θ . Ignore
planetary rotation and assume a << g .
Solution: Because Earth would then accelerate with a, the field a would seem invisible in the acceler-
ating frame. A tidal force would only appear if a depended on position, i.e. ∇⃗
a ̸= 0.
Example 4.3:
Now consider that the field is no longer constant, but that instead a = −kz with |kR| << g .
Solution: The surface of the planet needs to be at constant potential (if the planet is not accelerating).
The force per mass, −kz is like a spring, and the potential per mass is kz 2 /2. Otherwise water would
move to a point of lower potential. Thus, the potential energy for a sample mass δm is
δm
V (R) + δmgh(θ) − kr 2 cos2 θ = Constant
2
δm δm
V (R) + δmgh(θ) − kR2 cos2 θ − δmkRh(θ) cos2 θ − kh2 (θ) cos2 θ = Constant.
2 2
Here, the potential due to the external field is (1/2)kz 2 so that −∇U = −kz . One now needs to
solve for h(θ). Absorbing all the constant terms from both sides of the equation into one constant C ,
and because both h and kR are small, we can through away terms of order h2 or kRh. This gives
1
gh(θ) − kR2 cos2 θ = C,
2
C 1
h(θ) = + kR2 cos2 θ,
g 2g
1
h(θ) = kR2 (cos2 θ − 1/3).
2g
The term with the factor of 1/3 replaced the constant and was chosen so that the average height of the
water would be zero.
Example 4.4:
The Sun’s mass is 27 × 106 the Moon’s mass, but the Sun is 390 times further away from Earth as the
Sun. What is ratio of the tidal force of the Sun to that of the Moon.
Solution: The gravitational force due to an object M a distance D away goes as M /D 2 , but the tidal
force is only the difference of that force over a distance R,
M
Ftidal ∝ R.
D3
44
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
1. Find equations of motion for r and t with no angle (θ ) mentioned, i.e. d2 r/dt2 = · · · . Angular
momentum conservation will be used, and the equation will involve the angular momentum L.
2. Use angular momentum conservation to find an expression for θ̇ in terms of r .
3. Use the chain rule to convert the equations of motions for r , an expression involving r, ṙ and r̈ ,
to one involving r, dr/dθ and d2 r/dθ 2 . This is quitecomplicated because the expressions will
also involve a substitution u = 1/r so that one finds an expression in terms of u and θ .
4. Once u(θ) is found, you need to show that this can be converted to the familiar form for an
ellipse.
d d
r2 = (x2 + y 2 ) = 2xẋ + 2y ẏ = 2r ṙ, (4.10)
dt dt
x y
ṙ = ẋ + ẏ,
r r
x y ẋ2 + ẏ 2 ṙ 2
r̈ = ẍ + ÿ + − .
r r r r
Recognizing that the numerator of the third term is the velocity squared, and that it can be written in
polar coordinates,
v 2 = ẋ2 + ẏ 2 = ṙ 2 + r 2 θ̇ 2 , (4.11)
one can write r̈ as
Fx cos θ + Fy sin θ ṙ 2 + r 2 θ̇ 2 ṙ 2
r̈ = + − (4.12)
m r r
2 2
F r θ̇
= +
m r
L2
mr̈ = F + .
mr 3
45
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
This derivation used the fact that the force was radial, F = Fr = Fx cos θ+Fy sin θ , and that angular
momentum is L = mrvθ = mr 2 θ̇ . The term L2 /mr 3 = mv 2 /r behaves like an additional force.
Sometimes this is referred to as a centrifugal force, but it is not a force. Instead, it is the consequence
of considering the motion in a rotating (and therefore accelerating) frame.
Now, we switch to the particular case of an attractive inverse square force, F = −α/r 2 , and show
that the trajectory, r(θ), is an ellipse. To do this we transform derivatives w.r.t. time to derivatives
w.r.t. θ using the chain rule combined with angular momentum conservation, θ̇ = L/mr 2 .
dr dr L
ṙ = θ̇ = , (4.13)
dθ dθ mr 2
( )
d2 r
2
dr d L
r̈ = θ̇ + ṙ
dθ 2 dθ dr mr 2
( )
d2 r L 2 dr L
= − 2 ṙ
dθ 2 mr 2 dθ mr 3
( ) ( ) ( )
d2 r L 2 2 dr 2 L 2
= −
dθ 2 mr 2 r dθ mr 2
Equating the two expressions for r̈ in Eq.s (4.12) and (4.13) eliminates all the derivatives w.r.t. time,
and provides a differential equation with only derivatives w.r.t. θ ,
( )2 ( )2 ( )2
d2 r L 2 dr L F L2
− = + , (4.14)
dθ 2 mr 2 r dθ mr 2 m m2 r 3
that when solved yields the trajectory, i.e. r(θ). Up to this point the expressions work for any radial
force, not just forces that fall as 1/r 2 .
The trick to simplifying this differential equation for the inverse square problems is to make a sub-
stitution, u ≡ 1/r , and rewrite the differential equation for u(θ).
r = 1/u, (4.15)
dr 1 du
= − 2 ,
dθ u dθ
( )
d2 r 2 du 2 1 d2 u
= − .
dθ 2 u3 dθ u2 dθ 2
Plugging these expressions into Eq. (4.14) gives an expression in terms of u, du/dθ , and d2 u/dθ 2 .
After some tedious algebra,
d2 u Fm
= −u − . (4.16)
dθ 2 L2 u2
For the attractive inverse square law force, F = −αu2 ,
d2 u mα
.= −u + (4.17)
dθ 2 L2
The solution has two arbitrary constants, A and θ0 ,
mα
u = + A cos(θ − θ0 ), (4.18)
L2
1
r = .
(mα/L2 ) + A cos(θ − θ0 )
46
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
The radius will be at a minimum when θ = θ0 and at a maximum when θ = θ0 + π . The constant
A is related to the eccentricity of the orbit. When A = 0 the radius is a constant r = L2 /(mα), and
the motion is circular. If one solved the expression mv 2 /r = −α/r 2 for a circular orbit, using the
substitution v = L/(mr), one would reproduce the expression r = L2 /(mα).
The form describing the elliptical trajectory in Eq. (4.18) can be identified as an ellipse with one focus
being the center of the ellipse by considering the definition of an ellipse as being the points such
that the sum of the two distances between the two foci are a constant. Making that distance 2D , the
distance between the two foci as 2a, and putting one focus at the origin,
√
2D = r + (r cos θ − 2a)2 + r 2 sin2 θ, (4.19)
4D + r − 4Dr = r + 4a − 4ar cos θ,
2 2 2 2
D 2 − a2 1
r = = .
D + a cos θ D/(D − a ) − a cos θ/(D 2 − a2 )
2 2
By inspection, this is the same form as Eq. (4.18) with D/(D 2 − a2 ) = mα/L2 and a/(D 2 − a2 ) =
A.
Example 4.5:
Consider a particle of mass m in a 2-dimensional harmonic oscillator with potential
1 1
U = kr 2 = k(x2 + y 2 ).
2 2
47
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
1 2
L2
Ueff = kr +
2 2mr 2
The effective potential looks like that of a harmonic oscillator for large r , but for small r , the centrifu-
gal potential repels the particle from the origin. The combination of the two potentials has a minimum
for at some radius rmin .
L2
0 = krmin − 3
,
mrmin
( )1/4
L2
rmin = ,
mk
L √
θ̇ = 2
= k/m.
mrmin
For particles at rmin with ṙ = 0, the particle does not accelerate and r stays constant, i.e. a circular
orbit. The radius of the circular orbit can be adjusted by changing the angular momentum L.
b) Now consider small vibrations about rmin . The effective spring constant is the curvature of the
effective potential.
d2 3L2
keff = U
eff (r) =k+
dr 2 4
mrmin
r=rmin
= 4k,
√ √
ω = keff /m = 2 k/m = 2θ̇.
Here, the second step used the result of the last step from part (a). Because the radius oscillates with
twice the angular frequency, the orbit has two places where r reaches a minimum in one cycle. This
differs from the inverse-square force where there is one minimum in an orbit. One can show that the
orbit for the harmonic oscillator is also elliptical, but in this case the center of the potential is at the
center of the ellipse, not at one of the foci.
The solution is also simple to write down exactly in Cartesian coordinates. The x and y equations of
motion separate,
ẍ = −kx,
ÿ = −ky.
x = A cos ω0 t + B sin ω0 t,
y = C cos ω0 t + D sin ω0 t.
48
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
With some work using double angle formulas, one can calculate
r 2 = x2 + y 2
= (A2 + C 2 ) cos2 (ω0 t) + (B 2 + D 2 ) sin2 ω0 t + (AB + CD) cos(ω0 t) sin(ω0 t)
= α + β cos 2ω0 t + γ sin 2ω0 t,
A2 + B 2 + C 2 + D 2 A2 − B 2 + C 2 − D 2
α = , β= , γ = AB + CD,
2 2
r 2 = α + (β 2 + γ 2 )1/2 cos(2ω0 t − δ), δ = arctan(γ/β),
and see that the radius oscillates with frequency 2ω0 . The factor of two comes because the oscillation
x = A cos ω0 t has two maxima for x2 , one at t = 0 and one a half period later.
L2
mr̈ = F + (4.23)
mr 3
= Feff
= −∂r Ueff ,
[ ]
Feff = −∂r U (r) + (L2 /2mr 2 ) .
For a circular orbit, the radius must be fixed as a function of time, so one must be at a maximum or
a minimum of the effective potential. However, if one is at a maximum of the effective potential the
radius will be unstable. For the attractive Coulomb force the effective potential will be dominated
by the −α/r term for large r because the centrifugal part falls off more quickly, ∼ 1/r 2 . At low
r the centrifugal piece wins and the effective potential is repulsive. Thus, the potential must have a
minimum somewhere with negative potential. The circular orbits are then stable to perturbation.
If one considers a potential that falls as 1/r 3 , the situation is reversed and the point where ∂r U dis-
appears will be a local maximum rather than a local minimum – see Fig. 4.1. The repulsive centrifugal
piece dominates at large r and the attractive Coulomb piece wins out at small r . The circular orbit
is then at a maximum of the effective potential and the orbits are unstable. It is the clear that for
potentials that fall as r n , that one must have n > −2 for the orbits to be stable.
Example 4.6:
Consider a potential U (r) = βr . For a particle of mass m with angular momentum L, find the
angular frequency of a circular orbit. Then find the angular frequency for small radial perturbations.
Solution:
49
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
Veff(r)
U(r) = −α2 /r3
0
U(r) = −α1 /r
r
Figure 4.1: The effective potential is sketched for two cases, a 1/r attractive potential and a 1/r 3 attractive
potential. The 1/r case has a stable minimum, whereas the circular orbit in the 1/r 3 case is unstable.
For the circular orbit you search for the position rmin where the effective potential is minimized,
{ }
L2
∂r βr + = 0,
2mr 2
L2
β = 3
,
mrmin
( 2 )1/3
L
rmin = ,
βm
L β 2/3
θ̇ = 2
=
mrmin (mL)1/3
Now, we can find the angular frequency of small perturbations about the circular orbit. To do this we
find the effective spring constant for the effective potential,
If the two frequencies, θ̇ and ω , differ by an integer factor, the orbit’s trajectory will repeat itself each
time around. This is the case√ for the inverse-square force, ω = θ̇ , and for the harmonic oscillator,
ω = 2θ̇ . In this case, ω = 3θ̇ , and the angles at which the maxima and minima occur change with
each orbit.
50
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
d2 σ
. (4.25)
d cos θs dϕ
Usually, the literatures expresses differential cross sections as
dσ 1 dσ
dσ/dΩ = = , (4.26)
d cos θdϕ 2π d cos θ
where the last equivalency is true when the scattering does not depend on the azimuthal angle ϕ, as
is the case for spherically symmetric potentials.
The differential solid angle dΩ can be thought of as the area subtended by a measurement, dAd ,
divided by r 2 , where r is the distance to the detector,
51
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
With this definition dσ/dΩ is independent of the distance from which one places the detector, or the
size of the detector (as long as it is small).
Differential scattering cross sections are calculated by assuming a random distribution of impact pa-
rameters b. These represent the distance in the xy plane for particles moving in the z direction relative
to the scattering center. An impact parameter b = 0 refers to being aimed directly at the target’s center.
The impact parameter describes the transverse distance from the z = 0 axis for the trajectory when it
is still far away from the scattering center and has not yet passed it. The differential cross section can
be expressed in terms of the impact parameter,
dσ = 2πbdb, (4.28)
which is the area of a thin ring of radius b and thickness db. In classical physics, one can calculate the
trajectory given the incoming kinetic energy E and the impact parameter if one knows the mass and
potential. From the trajectory, one then finds the scattering angle θs (b). The differential cross section
is then
dσ 1 dσ db b
= =b = . (4.29)
dΩ 2π d cos θs d cos θs (d/db) cos θs (b)
Typically, one would calculate cos θs and (d/db) cos θs as functions of b. This is sufficient to plot the
differential cross section as a function of θs .
The total cross section is ∫ ∫
dσ dσ
σtot = dΩ = 2π d cos θs . (4.30)
dΩ dΩ
Even if the total cross section is infinite, e.g. Coulomb forces, one can still have a finite differential
cross section as we will see later on.
Example 4.7:
An asteroid of mass m and kinetic energy E approaches a planet of radius R and mass M . What is
the cross section for the asteroid to impact the planet?
Solution:
Calculate the maximum impact parameter, bmax , for which the asteroid will hit the planet. The total
cross section for impact is σimpact = πb2max . The maximum cross-section can be found with the help
√
of angular momentum conservation. The asteroid’s incoming momentum is p0 = 2mE and the
angular momentum is L = p0 b. If the asteroid just grazes the planet, it is moving with zero radial
kinetic energy at impact. Combining energy and angular momentum conservation and having pf refer
to the momentum of the asteroid at a distance R,
p2f GM m
− = E,
2m R
pf R = p0 bmax ,
allows one to solve for bmax ,
pf
bmax = R
p
√0
2m(E + GM m/R)
= R √
2mE
E + GM m/R
σimpact = πR2 .
E
52
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
r¨ = F
µ⃗ ⃗12 , (4.34)
1 1 1 m1 m2
= + , µ= . (4.35)
µ m1 m2 m1 + m2
Thus, one can treat the trajectory as a one-body problem where the reduced mass is µ, and a second
trivial problem for the center of mass. The reduced mass is especially convenient when one is consid-
ering gravitational problems because then
Gm1 m2
µr̈ = − r̂ (4.36)
r2
GM µ
= − r̂, M ≡ m1 + m2 .
r2
For the gravitational problem, the reduced mass then falls out and the trajectory depends only on the
total mass M .
The kinetic energy and momenta also have analogues in center-of-mass coordinates. The total and
relative momenta are
⃗ ≡ p
P ⃗1 + p ⃗˙ cm ,
⃗2 = M R (4.37)
⃗ ≡ µ⃗
q r˙.
53
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
With these definitions, a little algebra shows that the kinetic energy becomes
1 1
T = m1 |⃗
v1 |2 + m2 |⃗
v2 |2 (4.38)
2 2
1 ˙ 1 ˙2
= M |R
⃗ cm |2 + µ|⃗
r|
2 2
P2 q2
= + .
2M 2µ
The standard strategy is to transform into the center of mass frame, then treat the problem as one of
a single particle of mass µ undergoing a force F ⃗12 . Scattering angles can also be expressed in this
frame, then transformed into the lab frame. In practice, one sees examples in the literature where
dσ/dΩ expressed in both the “center-of-mass” and in the “laboratory” frame.
θ'
θs
Figure 4.2: The incoming and outgoing angles of the trajectory are at ±θ ′ . They are related to the scattering
angle by 2θ ′ = π + θs .
In order to calculate differential cross section, we must find how the impact parameter is related to the
scattering angle. This requires analysis of the trajectory. We consider our previous expression for the
trajectory where we derived the elliptic form for the trajectory, Eq. (4.18). For that case we considered
an attractive force with the particle’s energy being negative, i.e. it was bound. However, the same form
54
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
will work for positive energy, and repulsive forces can be considered by simple flipping the sign of α.
For positive energies, the trajectories will be hyperbolas, rather than ellipses, with the asymptotes of
the trajectories representing the directions of the incoming and outgoing tracks. Rewriting Eq. (4.18),
1
r= mα . (4.39)
L2
+ A cos θ
Once A is large enough, which will happen when the energy is positive, the denominator will become
negative for a range of θ . This is because the scattered particle will never reach certain angles. The
asymptotic angles θ ′ are those for which the denominator goes to zero,
mα
cos θ ′ = − . (4.40)
AL2
The trajectory’s point of closest approach is at θ = 0 and the two angles θ ′ , which have this value
of cos θ ′ , are the angles of the incoming and outgoing particles. From Fig. 4.2, one can see that the
scattering angle θs is given by,
π θs
2θ ′ − π = θs , θ ′ = + , (4.41)
2 2
sin(θs /2) = − cos θ ′
mα
= .
AL2
Now that we have θs in terms of m, α, L and A, we wish to re-express L and A in terms of the
impact parameter b and the energy E . This will set us up to calculate the differential cross section,
which requires knowing db/dθs . It is easy to write the angular momentum as
Finding A is more complicated. To accomplish this we realize that the point of closest approach occurs
at θ = 0, so from Eq. (4.39)
1 mα
= + A, (4.43)
rmin L2
1 mα
A = − .
rmin L2
Next, rmin can be found in terms of the energy because at the point of closest approach the kinetic
energy is due purely to the motion perpendicular to r̂ and
α L2
E=− + 2
. (4.44)
rmin 2mrmin
1 mα √
= + (mα/L2 )2 + 2mE/L2 . (4.45)
rmin L2
We can plug the expression for rmin into the expression for A, Eq. (4.43),
√ √
A= (mα/L2 )2 + 2mE/L2 = (α2 /(4E 2 b4 ) + 1/b2 (4.46)
55
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
Finally, we insert the expression for A into that for the scattering angle, Eq. (4.41),
mα
sin(θs /2) = (4.47)
AL2
a α
= √ , a≡
a2 + b2 2E
The differential cross section can now be found by differentiating the expression for θs with b,
1 ab db bdb
cos(θs /2)dθs = = sin3 (θs /2), (4.48)
2 (a2 + b2 )3/2 a2
πa2
dσ = 2πbdb = cos(θs /2)dθs
sin3 (θs /2)
πa2
= sin θs dθs
2 sin4 (θs /2)
dσ πa2
= ,
d cos θs 2 sin4 (θs /2)
dσ a2
= .
dΩ 4 sin4 (θs /2)
where a = α/2E . This the Rutherford formula for the differential cross section. It diverges as θs →
0 because scatterings with arbitrarily large impact parameters still scatter to arbitrarily small scattering
angles. The expression for dσ/dΩ is the same whether the interaction is positive or negative.
Example 4.8:
Consider a particle of mass m and charge z with kinetic energy E (Let it be the center-of-mass energy)
incident on a heavy nucleus of mass M and charge Z and radius R. Find the angle at which the
Rutherford scattering formula breaks down.
Solution:
Let α = Zze2 /(4πϵ0 ). The scattering angle in Eq. (4.47) is
a α
sin(θs /2) = √ , a≡ .
a2 + b2 2E
The impact parameter b for which the point of closest approach equals R can be found by using an-
gular momentum conservation,
√ √
p0 b = b 2mE = Rpf = R 2m(E − α/R),
√
2m(E − α/R)
b = R √
2mE
√
α
= R 1− .
ER
Putting these together
{ }
a α
θs = 2 sin−1 √ , a= .
a2 + R2 (1 − α/(RE)) 2E
56
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
It was from this departure of the experimentally measured dσ/dΩ from the Rutherford formula that
allowed Rutherford to infer the radius of the gold nucleus, R.
4.9 Exercises
1. Approximate Earth as a solid sphere of uniform density and radius R = 6360 km. Suppose
you drill a tunnel from the north pole directly to another point on the surface described by a
polar angle θ relative to the north pole. Drop a mass into the hole and let it slide through tunnel
without friction. Find the frequency f with which the mass oscillates back and forth. Ignore
Earth’s rotation. Compare this to the frequency of a low-lying circular orbit.
gmoon = g0 + kz + · · · ,
where z is measured relative to Earth’s center and is measured along the axis connecting
the Earth and moon. Give your answer in terms of the distance between the moon and the
earth, Rm and the mass of the moon Mm .
(b) Calculate the difference between the height of the oceans at maximum and minimum tides.
Express your answer in terms of the quantities above, plus Earth’s radius, Re . Then give
you answer in meters.
3. Consider an ellipse defined by the sum of the distances from the two foci being 2D , which
expressed in a Cartesian coordinates with the middle of the ellipse being at the origin becomes
√ √
(x − a)2 + y 2 + (x + a)2 + y 2 = 2D.
Here the two foci are at (a, 0) and (−a, 0). Show that this form is can be written as
x2 y2
+ = 1.
D2 D 2 − a2
4. Consider a particle in an attractive inverse-square potential, U (r) = −α/r , where the point of
closest approach is rmin and the total energy of the particle is E . Find the parameter A describing
the trajectory in Eq. (4.18). Hint: Use the fact that at rmin there is no radial kinetic energy and
E = −α/rmin + L2 /2mrmin 2
.
5. Consider the effective potential for an attractive inverse-square-law force, F = −α/r 2 . Con-
sider a particle of mass m with angular momentum L.
(a) Find the radius of a circular orbit by solving for the position of the minimum of the effective
potential.
(b) What is the angular frequency, θ̇ , of the orbit? Solve this by setting F = mθ̇ 2 r .
(c) Find the effective spring constant for the particle at the minimum.
57
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
(d) What is the angular frequency for small vibrations about the minimum? How does this
compare with the answer to (b)?
U = α ln(r/a).
(a) If the particle is moving in a circular orbit of radius R, find the angular frequency θ̇ . Solve
this by setting F = −mθ̇ 2 r (force and acceleration point inward).
(b) Express the angular momentum L in terms of α, m and R. Also express R in terms of L,
α and m.
(c) Sketch the effective radial potential, Veff (r), for a particle with angular momentum L. (No
longer necessarily moving in a circular orbit.)
(d) Find the position of the minimum of Veff in terms of L, α and m, then compare to the result
of (b).
(e) What is the effective spring constant for a particle at the minimum of Veff ? Express your
answer in terms of L, m and α.
(f) What is the angular frequency, ω , for small oscillations of r about the Rmin ? Express your
answer in terms of θ̇ from part (a).
7. Consider a particle of mass m in an attractive potential, U (r) = −α/r , with angular momen-
tum L with just the right energy so that
A = mα/L2
1
r= .
(mα/L2 ) + A cos θ
The trajectory can then be rewritten as
2r0 L2
r= , r0 = .
1 + cos θ 2mα
(a) Show that for this case the total energy E approaches zero.
(b) Write this trajectory in a more recognizable parabolic form,
y2
x = x0 − .
R
I.e., express x0 and R in terms of r0 .
(c) Explain how a particle with zero energy can have its trajectory not go through the origin.
(d) What is the scattering angle for this trajectory?
8. Show that if one transforms to a reference frame where the total momentum is zero, p⃗1 = −⃗ p2 ,
that the relative momentum q ⃗1 or −⃗
⃗ corresponds to either p p2 . This means that in this frame
the magnitude of q ⃗1 − p
⃗ is one half the magnitude of p ⃗2 .
58
PHY 321 Lecture Notes 4 GRAVITY AND CENTRAL FORCES
10. Consider two particles of identical mass scattering at an angle θcm in the center of mass.
(a) In a frame where one is the target (initially at rest) and one is the projectile, find the scat-
tering angle in the lab frame, θ , in terms of θcm .
(b) Express dσ/d cos θ in terms of dσ/d cos θcm . I.e., find the Jacobian, d cos θcm /d cos θ .
11. Assume you are scattering alpha particles (He-4 nuclei Z = 2, A = 4) off of a gold target
(Z = 79, A = 197). If the radius of the nucleus is 7.5 × 10−15 meters, and if the energy of
the beam is 38 MeV,
(a) What is the total cross section for having a nuclear collision? Give the answer in millibarns,
1 mb= 10−31 m2 .
(b) Find the scattering angle (in degrees) at which the Rutherford differential cross section
formula breaks down?
59
PHY 321 Lecture Notes 5 ROTATING COORDINATE SYSTEMS
r′
d2⃗
m ⃗.
=F (5.1)
dt2
Now, if we have a second coordinate system,
1
⃗ r′ − ⃗
r=⃗ r0 , r⃗0 = a0 t2 .
⃗ (5.2)
2
We would see that
d2⃗
r
m ⃗ − m⃗
=F a0 . (5.3)
dt2
Here ⃗a0 is the acceleration of the coordinate system. The last term acts like an additional apparent
force. In fact, it acts like a contribution to the gravitational force which alters the acceleration of gravity
by δ⃗g = −⃗ a0 .
velocity of Earth’s rotation. The acceleration is inward toward the axis of rotation, so the additional
contribution to the apparent acceleration of gravity is outward in the x−y plane. In contrast the usual
g is radially inward pointing toward the origin.
⃗
In the rotating coordinate system (not an inertial frame), motion is determined by the apparent force
and one can define effective potentials. In addition to the normal gravitational potential energy, there
is a contribution to the effective potential,
m m
δUeff = − ω 2 r⊥
2
=− r 2 ω 2 sin2 θ, (5.4)
2 2
where θ is the polar angle, measured from the north pole. If the true gravitational force can be consid-
ered as originating from a point in Earth’s center, the net effective potential for a mass m near Earth’s
surface could be
1
Ueff = mgh − m ω 2 (R + h)2 sin2 θ. (5.5)
2
Example 5.1:
How much wider is Earth at the equator than the north-south distance between the poles assuming
that the gravitational field above the surface can be approximated by that of a point mass at Earth’s
center.
60
PHY 321 Lecture Notes 5 ROTATING COORDINATE SYSTEMS
Solution: The surface of the ocean must be at constant effective potential for a sample mass m. This
means that if h now refers to the height of the water
m
mg[h(θ = π/2) − h(θ = 0)] = ω 2 (R + h)2 .
2
Because R >> h, one can approximate R + h → R on the right-hand side, thus
ω 2 R2
h(θ = π) − h(θ = 0) = .
2g
This come out a bit less than 11 km, or a difference of near 22 km for the diameter of the Earth in the
equatorial plane compared to a diameter between the poles. In reality, the difference is approximately
41 km. The discrepancy comes from the assumption that the true gravitational force can be treated as
if it came from a point at Earth’s center. This would be true if the distribution of mass was radially
symmetric. However, Earth’s center is molten and the rotation distorts the mass distribution. Remark-
ably this effect nearly doubles the elliptic distortion of Earth’s shape. Due to this distortion, the top
of Mount Everest is not the furthest point from the center of the Earth. That belongs to the top of a
volcano, Chimborazo, in Equador, which is one degree in latitude below the Equator. Chimborazo is
about 8500 ft lower than Everest when measured relative to sea level, but is 7700 feet further from the
center of the Earth.
If one includes the fact that A, the vector measured in the rotating frame, might be changing as a
function of time,
∆A ⃗ ′ = ∆A ⃗ × A∆t,
⃗+ω ⃗ (5.7)
d ′ d
⃗ =
A A ⃗ × A.
⃗+ω ⃗
dt dt
If the vector happens to be the position ⃗
r,
r˙ ′ = ⃗
⃗ v′ = ⃗ ⃗ ×⃗
v+ω r. (5.8)
61
PHY 321 Lecture Notes 5 ROTATING COORDINATE SYSTEMS
v˙ ′ is F
Because ⃗ ⃗ /m,
⃗ = m {⃗
F ω ×⃗
a + 2⃗ v+ω ⃗ × (⃗
ω×⃗r )} , (5.10)
m⃗ ⃗ − 2m⃗
a = F ω×⃗ v − m⃗ω × (⃗
ω×⃗r ).
The extra terms on the right behave like additional forces. Like gravitational forces, they are propor-
tional to the mass, so the mass cancels for many problems.
The last term, −m⃗ω × (⃗
ω ×⃗ ⃗ × (B
r ), represents the centrifugal force. Using the vector identity, A ⃗×
⃗ = B(
C) ⃗ A ⃗ · C)
⃗ − C(
⃗ A⃗ · B)
⃗ ,
−⃗
ω × (⃗
ω×⃗ r + (ω · ⃗
r ) = ω 2⃗ r )⃗
ω. (5.11)
If ω
⃗ is in the z direction,
−⃗
ω × (⃗
ω×⃗
r ) = ω 2 (xx̂ + y ŷ). (5.12)
√ centrifugal force points outward in the x − y plane, and its magnitude is mω r⊥ , where r⊥ =
2
The
x2 + y 2 .
The second term is Eq. (5.10) represents the Coriolis force. It does not enter problems like the shape of
the Earth above because in that case the water was not moving relative to the rotating frame. Once an
object is moving in a rotating frame, the particle is no longer being described in a single accelerating
frame because at each point the acceleration is −ω 2⃗ r.
Example 5.2:
A ball is dropped from a height h = 500m above Minneapolis. Due to the Coriolis force, it is deflected
by an amount δx and δy . Find the deflection. Ignore the centrifugal terms.
Solution: The equations of motion are:
dvx
= −2(ωy vz − ωz vy ),
dt
dvy
= −2(ωz vx − ωx vz ),
dt
dvz
= −g − 2(ωx vy − ωy vx ),
dt
ωz = ω cos θ, ωy = ω sin θ, ωx = 0.
Here the coordinate system is x̂ points east, ŷ points north and ẑ points upward.
One can now ignore all the Coriolis terms on the right-hand sides except for those with vz . The other
terms will all be doubly small. One can also throw out terms with ωx . This gives
dvx
≈ −2ωvz sin θ,
dt
dvy
≈ 0,
dt
dvz
≈ −g.
dt
62
PHY 321 Lecture Notes 5 ROTATING COORDINATE SYSTEMS
There will be no significant deflection in the y direction, δy = 0, but in the x direction one can
substitute vz = −gt above,
∫ t
vx ≈ dt′ 2ωgt′ sin θ = ωgt2 sin θ,
0
∫ t
gω sin θt3
δx ≈ dt′ vx (t′ ) = .
0 3
One can find the deflections by using h = 12 gt2 , to find the time, and using the all-knowing internet
to see that the latitude of Minneapolis is 44.6◦ or θ = 45.4◦ .
√
t = 2h/g = 10.1 s,
2π
ω = = 7.27 × 10−5 s−1 ,
3600 · 24 s
δx = 17.4 cm (east).
r¨ = T
m⃗ g − 2mΩ
⃗ + m⃗ ⃗ ×⃗
v,
Tx = −mgx/L, Ty = −mgy/L.
ẍ = −gx/L + 2ẏΩz ,
ÿ = −gy/L − 2ẋΩz .
Here we
√ have used the fact that the oscillations are sufficiently small so we can ignore vz . Using
ω0 ≡ k/m,
ẍ − 2Ωz ẏ + ω02 x = 0
ÿ + 2Ωz ẋ + ω02 y = 0,
where Ωz = |Ω| ⃗ cos θ , with θ being the polar angle (zero at the north pole). The terms linear in time
derivatives are what make life difficult. This will be solved with a trick. We will incorporate both
differential equations into a single complex equation where the first/second are the real/imaginary
parts.
η ≡ x + iy,
η̈ + 2iΩz η̇ + ω02 η = 0.
63
PHY 321 Lecture Notes 5 ROTATING COORDINATE SYSTEMS
Now, we guess at a form for the solutions, η(t) = e−iαt , which turns the differential equation into
−α2 + 2Ωz α + ω02 = 0,
√
α = Ωz ± Ω2z + ω02 ,
≈ Ωz ± ω 0 .
The solution with two arbitrary constants is then
[ ]
η = e−iΩz t C1 eiω0 t + C2 e−iω0 t .
Here, C1 and C2 are complex, so they actually represent four arbitrary numbers. These four numbers
should be fixed by the four initial conditions, i.e. x(t = 0), ẋ(t = 0), y(t = 0) and ẏ(t = 0). With
some lengthy algebra, one can rewrite the expression as
η = e−iΩz t [A cos(ω0 t + ϕA ) + iB cos(ω0 t + ϕB )] .
Here, the four coefficients are represented by the two real arbitrary real amplitudes, A and B , and two
arbitrary phases, ϕA and ϕB . For an initial condition where y = 0 at t = 0, one can see that B = 0.
This then gives
η(t) = Ae−iΩz t cos(ω0 t + γ)
= A cos Ωz t cos(ω0 t + γ) + iA sin Ωz t cos(ω0 t + γ).
Translating into x and y ,
x = A cos Ωz t cos(ω0 t + γ), (5.13)
y = A sin Ωz t cos(ω0 t + γ).
Assuming the pendulum’s frequency is much higher than Earth’s rotational frequency, ω0 >> Ωz ,
one can see that the plane of the pendulum simply precesses with angular velocity Ωz . This means
that in this limit the pendulum oscillates only in the x-direction with frequency many times before
the phase Ωz t becomes noticeable. Eventually, when Ωz t = π/2, the motion is along the y -direction.
If you were at the north pole, the motion would switch from the x-direction to the y direction every
6 hours. Away from the north pole, Ωz ̸= |Ω| ⃗ and the precession frequency is less. At the equator it
does not precess at all. If one were to repeat for the solutions where A = 0 and B ̸= 0 in Eq. (5.13),
one would look at motions that started in the y -direction, then precessed toward the −x direction.
Linear combinations of the two sets of solutions give pendulum motions that resemble ellipses rather
than simple back-and-forth motion.
5.5 Exercises
1. Consider a pail of water spinning about a vertical axis at the center of the pail with frequency
ω . Find the height of the water (within a constant) as a function the radius r⊥ from the axis of
rotation. Use the concept of a centrifugal potential in the rotating frame.
2. A high-speed cannon shoots a projectile with an initial velocity of 1000 m/s in the east direction.
The cannon is situated in Minneapolis (latitude of 45 degrees) The projectile velocity is nearly
horizontal and it hits the ground after a distance x = 3000 m. Find the alteration of the point
of impact in the north-south (y ) direction due to the Coriolis force. Assume the effect is small
so that you can approximate the eastward (x) component of the velocity as being constant. Be
sure to indicate whether the deflection is north or south.
64
PHY 321 Lecture Notes 5 ROTATING COORDINATE SYSTEMS
3. Someone wishes to use a Foucault pendulum as a crude clock. If the person lives in Minneapolis
(latitude of 45◦ ), how much time will pass between having the pendulum swinging in the east-
west direction until it swings in the north-south direction.
65
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
Thus, we are minimizing J with respect to an infinite number of values of y(xi ) at points xi . As an
additional criteria, we will assume that y(x1 ) and y(x2 ) are fixed, and that that we will only consider
variations of y between the boundaries. The dependence on the derivative, y ′ = dy/dx, is crucial
because otherwise the solution would involve simply finding the one value of y that minimized f ,
and y(x) would equal a constant if there were no explicit x dependence. Furthermore, y wouldn’t
need to be continuous at the boundary.
The Euler equation is a differential equation for y(x), that when solved, provides the required solu-
tion. For an extrema
∫ x2 { }
∂f ∂f ′
δJ = dx δy(x) + δy (x) = 0. (6.1)
∂y ∂y ′
x1
For ANY small δy(x) at any point x, the change should be zero if one is at an optimum function y(x).
Integrating the second term by parts,
∫ x2 { ( )}
∂f d ∂f
δJ = dx − δy(x) (6.2)
∂y dx ∂y ′
x1
∂f ∂f
+ δy − δy
∂y ′ x2 ∂y ′ x1
= 0.
Because y is not allowed to vary at the endpoints, δy(x1 ) = δy(x2 ) = 0, so the middle line can be
ignored. Also, this relation must hold for ANY δy , so one can write the Euler equations (or sometimes
called the Euler Lagrange equations)
∂f d ∂f
=
. (6.3)
∂y dx ∂y ′
This will yield a differential equations for y . Combined with the boundary conditions,
66
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
one can now solve the differential equations for y . Because f is only a function of f ′ , not f ′′ or f ′′′ ,
the d/dx in Euler Lagrange equation can lead to terms involving f ′′ , but not f ′′′ . Thus, the equation
will be a second-order differential equation (not a linear equation usually), and the two boundary
conditions are sufficient to determine the entire equation.
Example 6.1:
Consider a particle constrained to move along a path (like a bead moving without friction on a wire)
and you need to design a path from x = y = 0 to some final point xf , yf . Assume there is a constant
force in the x direction, Fx = mg . Design the path so that the time the bead travels is a minimum.
Solution: The net time is
∫ ∫ √
dℓ xf
1 + y ′2
T = = dx √ = minimum.
v 0 2gx
√
Here we made use of the fact that dℓ = dx2 + dy 2 and that the velocity is determined by KE =
mv 2 /2 = mgx. The Euler equations can be applied if you first define the function as
√
′
1 + y ′2
f (y, y ; x) = √ .
x
The equations are then
d ∂f
= 0.
dx ∂y ′
The simplification ensued from f not having any dependence on y . This yields the differential equa-
tion
y′
= (2a)−1/2 , (6.5)
x1/2 (1 + y ′2 )1/2
because ∂f /∂y ′ must be a constant, which with some foresight we label (2a)−1/2 . One can now
solve for y ′ ,
(y ′ )2 = 2ax(1 + y ′2 )
√
′
x
y = ,
2a − x
∫ x √ ∫ x
′
x′ dx′ ′
x′ dx′
y(x) = dx √ = dx √
0 2a − x′ 0 2ax′ − x′2
∫ x ∫
1 (2x′ − 2a)dx′ x
dx′
= + a √
2 0 (2ax′ − x′2 )1/2 0 2ax′ − x′2
∫ ∫ x
−1 2ax−x du dx′
2
= √ +a √
2 0 u 0 a2 − (x′ − a)2
√
= − 2ax − x2 + a cos−1 (1 − x/a).
This turns out to be the equation for a cycloid or a brachiostone. If you rolled a wheel of radius a down
the y axis and followed a point on the rim, it would trace out a cycloid. Here, the constant a must be
chosen to match the boundary condition, y2 = y(x2 ). You can see the textbook for more details, plus
you get a chance to work with cycloids in the exercises at the end of this chapter.
67
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
C(x1 · · · xn ) = 0 (6.6)
∇f · ⃗
ϵ = 0, (6.8)
δC = ∇C · ⃗
ϵ = 0. (6.9)
That is to say if I take a small step in a direction that doesn’t change the constraint, then f must not
change if it is an extrema. Not changing the constraint implies the step is orthogonal to ∇C . As there
are n dimensions of x, the vector ∇C defines one direction, and ⃗ ϵ can be in any of the n−1 directions
orthogonal to ∇C . If ∇f · ⃗ ϵ = 0 for ANY of the n − 1 directions of ⃗ ϵ orthogonal to ∇C , then
∇f || ∇C. (6.10)
Because the two vectors are parallel you can say there must exist some constant λ such that
Here, λ is known as a Lagrange multiplier. Satisfying Eq. (6.11) is a necessary, but not a sufficient
condition. One could add a constant to the constraint and the gradient would not change. One must
find the correct value of λ that satisfies the constraint C = 0, rather than C = some other con-
stant. The strategy is then to solve Eq. (6.11) then adjust λ until one finds the x1 · · · xn that gives
C(x1 · · · xn ) = 0.
The method of Lagrange multipliers is counter-intuitive to one’s intuition to use the constraint to re-
duce the dimensionality of the problem. Normally, minimizing a function of n variables, leads to n
equations and n unknowns. A constraint could be used, by substitution, to replace the n variables
with n − 1 variables. Instead, we add an unknown parameter, λ, and change the equation to n + 1
equations with n + 1 unknowns, with the extra unknown being the Lagrange multiplier λ. Often, it
68
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
is rather easy to solve for x1 · · · xn . Then one is left with the usually difficult problem of finding λ,
often requiring the solution of a transcendental equation.
Example 6.2:
As an example of using Lagrange multipliers for a standard optimization formula we attempt to max-
imize the following function,
∑
n
F (x1 · · · xn ) = − xi ln(xi ),
i=1
with respect to the n variables xi . With no constraints, each xi would maximize the function for
[ ]
d ∑
− xi ln(xi ) = 0
dxj i
− ln(xj ) − 1 = 0, xj = e−1 .
Here, ϵi and E are fixed constants. We go forward by finding the extrema for
∑ ∑ ∑
G(x1 · · · xn ) = F − α xi − β ϵi x i = {−xi ln(xi ) − αxi − βϵi xi } .
i i i
There are two Lagrange multipliers, α and β , corresponding to the two constraints. One then solves
for the extrema
d
G = 0
dxj
= − ln(xj ) − 1 − α − βϵj ,
xj = exp {−1 − α − βϵj } .
∑ ∑
For any given α and β this provides a solution for constraining i xi and i ϵi xi to some values,
just not the values of unity and E that you wish. One would then have to search for the correct values
by adjusting α and β until the constraint are actually matched by solving a transcendental equation.
Although this can be complicated, it is certainly less expensive than searching over all N values of xi .
∑
This particular example corresponds to maximizing the entropy for a system, S = − i xi ln(xi ),
where xi is the probability of the system being in a particular discrete level i that has energy ϵi . One
wishes to maximize the entropy subject to the constraints that the probabilities sum to unity and the
average energy has some given value. The result that xi ∼ e−βϵi demonstrates the origin of the
Boltzmann factor, with the inverse temperature β = 1/T .
Lagrange multipliers also assist with the Euler-Lagrange equation. If one breaks an interval x1 <
x < x2 into a large number n → ∞ points ∑separated
{ by ′dx, the Euler-Lagrange equation
} involves
finding the n values yi at each point so that i dxf yi , yi = (yi+1 − yi−1 )/(2dx) is maximized
69
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
for some given function f . If an additional auxiliary constraint is added, also some function of the
n values yi , one can use the method of Lagrange multipliers. In the constraint can also be written as
some function of C(yi , yi′ ), then one simply adds a term λC(y, y ′ ) to the function f and uses the
Euler-Lagrange equation to find the extrema of.
∫ x2
J = dx f {y(x), y ′ (x), x} − λC {y(x), y ′ (x), x} , (6.12)
x1
in Eq. (6.3).
Example 6.3:
Consider a chain of length L and mass per unit length κ that hangs from point x = 0, y = 0 to
point xf , yf . The shape must minimize the potential energy. Find general expressions for the shape
in terms of three constants which must be chosen to match y(0) = 0, y(xf ) = yf and the fixed
length. Equivalently, one finds the function y(x) that provides an extrema for the integral,
Solution: One must minimize
∫ ∫ ∫ xf √ ∫ xf √
dℓ κgy − λ dℓ = dx 1+ y ′2 κgy −λ dx 1 + y ′2 .
0 0
Here λ is the Lagrange multiplier associated with constraining the length of the chain. The constrained
length L appears nowhere in the expression. Instead, one solves for form of the answer, then adjusts
λ to give the correct length. For the purposes of the Euler-Lagrange minimization one considers the
function
√ √
f (y, y ′ ; x) = κgy 1 + y ′2 − λ 1 + y ′2 . (6.14)
Because λ is an unknown constant and because minimizing a function multiplied by a constant is the
same as minimizing the function, we can equivlently minimize the integral using the function
√ √
f˜(y, y ′ ; x) = y 1 + y ′2 − λ̃ 1 + y ′2 , (6.15)
λ
λ̃ ≡ .
κg
The Euler-Lagrange equations then become
{ }
d y′ y′ √
√ y − λ̃ √ = 1 + y ′2 .
dx 1 + y ′2 1 + y ′2
Here, we will guess at the form of the solution,
y = λ̃ + a cosh[(x − x0 )/a].
One must find λ̃, x0 and a to satisfy three conditions, y(x = 0) = 0, y(x = xf ) = yf and that the
length is L. For a hanging chain a is positive. A solution with negative a would represent a maximum
of the potential energy. A remarkable property of the solution is that once you define the length and
the end-point positions y1 and y2 , the solution does not depend on κ or g . Thus, the shape of the
chain would be the same if you took it to the moon. These solutions are known as catenaries,
http://en.wikipedia.org/wiki/Catenary.
6.3 Lagrangians
Lagrangians represent a powerful method for solving problems that would be nearly impossible by
direct application of Newton’s third law, F⃗ = m⃗ a. The method works well for problems where a
system is well described by a few generalized coordinates. A generalized coordinate might be the angle
describing the position of a pendulum. This one angle takes the place of using x and y to describe
the position of the pendulum, then applying a clumsy constraint.
The Lagrangian equations of motion can be derived from a principle of least action, where the action
S is defined as ∫
S= dt L(q, q̇, t), (6.16)
where q is some coordinate that describes the orientation of a system and the Lagrangian L is defined
as
L = T − U, (6.17)
the difference of the kinetic and potential energies. Minimizing the action through the Euler-Lagrange
equations gives the Lagrangian equations of motion,
d ∂L ∂L
= . (6.18)
dt ∂ q̇ ∂q
We begin with two simple examples, neither of which gains from the Lagrangian approach.
Example 6.4:
Consider a particle of mass m connected to a spring with stiffness k. Derive the Lagrangian equations
of motion.
Solution:
1 1
L = mẋ2 − kx2 ,
2 2
d ∂L ∂L
= ,
dt ∂ ẋ ∂x
mẍ = −kx.
Example 6.5:
71
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
Derive the Lagrangian equations of motion for a pendulum of mass m and length ℓ.
m
L = ℓ2 θ̇ 2 − mgℓ(1 − cos θ),
2
d ∂L ∂L
= ,
dt ∂ θ̇ ∂θ
mℓ2 θ̈ = −mgℓ sin θ,
g
θ̈ = − sin θ,
ℓ
g
θ̈ ≈ − θ.
ℓ
1. The potential energy is a function of the generalized coordinates qi , but not of q̇i .
2. The relation between the original coordinates x, y, z · · · and the generalized coordinates does
not depend on q̇i , e.g. x(q, t) not x(q, q̇, t).
3. Any constraints used to reduce the number of degrees of freedom are functions of q
⃗, but not of
˙
⃗.
q
Going forward with the proof, consider xi (q1 , q2 · · · , t) and look at the l.h.s. of Lagrange’s equations
of motion.
∂T ∑ ∂T ∂ ẋi ∑ ∂T ∂xi
= + (6.19)
∂ q̇j i
∂ ẋi ∂ q˙j i
∂xi ∂ q˙j
∑ ∂ ẋi
= mẋi
i
∂ q˙j
∑ (δxi /δt)|fixed qj ′ ̸=j
= mẋi
i
δqj /δt
∑ δxi |fixed qj ′ ̸=j
= mẋi
i
δqj
∑ ∂xi
= mẋi .
i
∂qj
In the first line we used the fact that T does not depend on x. Continuing with taking the derivative
of U ,
∂U ∑ ∂U ∂xi
− = − = 0. (6.20)
∂ q̇j i
∂xi ∂ q̇j
72
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
In the first line above we used the fact that U does not depend on ẋ then we used the second condition
that x does not depend on q̇ . Adding the two pieces together, then taking the derivative w.r.t. time,
d ∂ ∑ ∂xi ∑ ∂ ẋi
(T − U ) = mẍi + mẋi .
dt ∂ q̇ i
∂qj i
∂qj
Now, we consider the r.h.s. of Lagrange’s equations. Because the kinetic energy depends only on ẋ
and not x, and because the potential depends on x but not ẋ,
∂ ∑ ∂T ∂ ẋi ∑ ∂U ∂xi
(T − U ) = − (6.21)
∂qj i
∂ ẋi ∂qj i
∂xi ∂qj
∑ ∂ ẋi ∑ ∂U ∂xi
= mẋi −
i
∂qj i
∂xi ∂qj
Using the fact that mẍi = −(∂/∂xi )U , one can see that the bottom expressions in Eq.s (6.19) and
(6.21) are identical,
d ∂ ∂
(T − U ) = (T − U ). (6.22)
dt ∂ q̇i ∂qi
Example 6.6:
Consider a cone of half angle α standing on its tip at the origin. The surface of the cone is defined as
√
r= x2 + y 2 = z tan α.
Find the equations of motion for a particle of mass m moving along the surface under the influence
of a constant gravitational force, −mg ẑ . For generalized coordinates use the azimuthal angle ϕ and
r.
Solution:
The kinetic energy is
1 1
T = mr 2 θ̇ 2 + m(ṙ 2 + ż 2 )
2 2
1 1 ( )
= mr 2 θ̇ 2 + mṙ 2 1 + cot2 α
2 2
1 1
= mr 2 θ̇ 2 + mṙ 2 csc2 α.
2 2
The potential energy is
U = mgr cot α,
73
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
d ( )
mr 2 θ̇ = 0,
dt
d ( )
m csc2 αṙ = mr θ̇ 2 − mg cot α,
dt
r̈ = r θ̇ 2 sin2 α − g cos α sin α
The first equation is a statement of the conservation of angular momentum with L = mr 2 θ̇ , so the
second equation can also be expressed as
L2 sin2 α
r̈ = − g sin α cos α.
m2 r 3
Example 6.7:
A bead slides along a wire bent in the shape of a parabola,
1
z= kr 2 , r 2 = x2 + y 2 .
2
Also, the parabolic wire is rotating about the z axis with angular velocity ω . Derive the equations of
motion. Are there any stable configurations?
Solution:
Using the fact that
∂z
ż = ṙ = kr ṙ,
∂r
the kinetic and potential energies are
1 ( )
T = m ṙ 2 + ż 2 + r 2 ω 2
2
1 ( )
= m ṙ 2 + (kr ṙ)2 + r 2 ω 2 ,
2
U = mgkr 2 /2.
d { }
mṙ(1 + k2 r 2 ) = −mgkr + mk2 ṙ 2 r + mω 2 r,
dt
−gkr + ω 2 r − k2 ṙ 2 r
r̈ =
1 + k2 r 2
For a stable configuration, there needs to be a solution with ṙ = 0 and r̈ = 0. This can only happen at
r = 0, and then for the acceleration to be inward for small deviations of r one needs to have gk > ω 2 .
If ω 2 > gk the bead will move outward indefinitely.
74
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
k k
m 2m m
x1
x2
x3
Figure 6.1: For Example 6.8, three masses connected by two springs. The center-of-mass motion is unaffected
by the springs.
Example 6.8:
Consider two springs, whose relaxed lengths are ℓ, connected to three masses as depicted in Fig. 6.1.
Describe the two normal modes of the motion. We can write the Lagrangian as
m m k k
L = ẋ21 + mẋ22 + ẋ23 − (x2 − x1 − ℓ)2 − (x3 − x2 − ℓ)2 .
2 2 2 2
75
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
There are three coordinates, thus there are three equations of motion,
mẍ1 = −k(x1 − x2 + ℓ)
2mẍ2 = −k(x2 − x1 − ℓ) − k(x2 − x3 + ℓ)
= −k(2x2 − x1 − x3 )
mẍ3 = −k(x3 − x2 + ℓ).
This is a bit complicated because the center-of-mass motion does not easily separate from the three
equations. Instead, choose the following coordinates,
x1 + 2x2 + x3
X = ,
4
q1 = x1 − x2 + ℓ,
q3 = x3 − x2 − ℓ.
k
U = (q12 + q32 ).
2
To express the kinetic energy express x1 , x2 and x3 in terms of X , q1 and q3 ,
x1 = (3q1 − q3 − 4ℓ + 4X)/4,
x2 = (4X − q1 − q3 )/4,
x3 = (3q3 − q1 + 4ℓ + 4X)/4.
m 1 1 m 1
T = (3q̇1 − q̇3 + 4Ẋ)2 + m (4Ẋ − q̇1 − q̇3 )2 + (3q̇3 − q̇1 + 4Ẋ)2
2 16 16 2 16
3m 2 m
= (q̇1 + q̇32 ) − q̇1 q̇3 + 2mẊ 2 ,
8 4
3m 2 m k k
L = (q̇1 + q̇32 ) − q̇1 q̇3 + 2mẊ 2 − q12 − q32 .
8 4 2 2
The three equations of motion are then,
3 1
mq̈1 − mq̈3 = −kq1 ,
4 4
3 1
mq̈3 − mq̈1 = −kq3 ,
4 4
4M Ẍ = 0.
The last equation simply states that the center-of-mass velocity is fixed. One could obtain the same
result by summing the equations of motion for x1 , 2x2 and x3 above. The second two equations are
more complicated. To solve them, we assume a form
q1 = Aeiωt ,
q3 = Beiωt ,
76
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
Because this is a linear equation, we can multiply the solution by a constant and it will still be a
solution. Thus, we can set B = 1, then solve for A, effectively solving for A/B . Putting this guess
into the equations of motion,
3A 1 A
− ω2 + ω 2 = −ω02 ,
4B 4 B
3 2 1A 2
− ω + ω = −ω02 .
4 4B
This is two equations and two unknowns, ω 2 and A/B . Substituting for A/B gives a quadratic
equation,
ω 4 − 3ω02 ω 2 + 2ω04 = 0,
ω02 ≡ k/m.
(1) ω = ω0 , A = −B,
√
(2) ω = ω0 2, A = B.
The first solution corresponds to the two outer masses moving in opposite directions, in sync, with
the middle mass fixed. The second solution has both outer masses moving in the same direction, but
with the center mass moving opposite. These two solutions are referred to as normal modes, and are
characterized by their frequency and by the linear combinations of coordinates that oscillate together.
In general, the solution is a linear combination of normal modes, which usually results in a chaotic
looking motion. However, once the solution is expressed in terms of the normal modes, each of which
oscillates independently in a simple manner, one can better understand the motion. Further, the fre-
quencies of these modes represent the natural resonant frequencies of the system. This is important
in the construction of many structures, such as bridges or vehicles.
Example 6.9:
Consider a double pendulum confined to the x − y plane, where y is vertical. A mass m is connected
to the ceiling with a massless string of length ℓ. A second mass m hangs from the first mass with an
identical massless string of the same length. Using θ1 and θ2 to describe the orientations of the strings
relative to the vertical axis, find the Lagrangian and derive the equations of motion, both for arbitrary
angles and in the small-angle approximation. Finally, express the equations of motion in the limit of
small oscillations.
Solution:
The kinetic and potential energies are:
1 1{ }
T = mℓ2 θ̇12 2
+ m (ℓθ̇1 cos θ1 + ℓθ̇2 cos θ2 ) + (ℓθ̇1 sin θ1 + ℓθ̇2 sin θ2 ) 2
2 2
1 { }
= mℓ2 2θ̇12 + θ̇22 + 2θ̇1 θ̇2 cos(θ1 − θ2 ) ,
2
U = mgℓ(1 − cos θ1 ) + mg [ℓ(1 − cos θ1 ) + ℓ(1 − cos θ2 )]
= mgℓ(3 − 2 cos θ1 − cos θ2 )
77
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
d { }
mℓ 2
2θ̇1 + θ̇2 cos(θ1 − θ2 ) = −mℓ2 θ̇1 θ̇2 sin(θ1 − θ2 ) − 2mgℓ sin θ1 ,
dt
2θ̈1 + θ̈2 cos(θ1 − θ2 ) + θ̇22 sin(θ1 − θ2 ) = −2ω02 sin θ1 ,
ω02 ≡ g/ℓ,
d { }
mℓ 2
θ̇2 + θ̇1 cos(θ1 − θ2 ) = mℓ2 θ̇1 θ̇2 sin(θ1 − θ2 ) − mgℓ sin θ2 ,
dt
θ̈2 + θ¨1 cos(θ1 − θ2 ) = −ω02 sin θ2 .
For small oscillations, one can only consider terms linear in θ1 and θ2 or their derivatives,
To find the solutions, assume they are of the form θ1 = Aeiωt , θ2 = Beiωt . Solve for ω and A/B ,
noting that B is arbitrary.
Plug in the desired form and find
We can treat B as arbitrary and set it to unity. When we find A, it is the same as A/B for arbitrary B .
This gives the equations
2ω 2 A + ω 2 = 2ω02 A,
ω 2 A + ω 2 = ω02 .
This is two equations and two unknowns. Solving them leads to a quadratic equation with solutions
1
A/B = ± √ ,
2
ω02
ω2 = √ .
1 ± 1/ 2
Again, these two solutions are the normal modes, and the general solution is a sum of the two solutions,
with two arbitrary constants. For the angles θ1 and θ2 are:
A+
θ1 = √ eiω+ t , θ2 = A+ eiω+ t ,
2
−A− iω− t
θ1 = √ e , θ2 = A− eiω− t ,
2
√
1
ω± = ω0 √ .
1 ± 1/ 2
78
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
One can also express the solution in vector notation, with the vectors having arbitrary amplitudes A+
and A− ,
( 1 )
√
θ+ = 2 A+ eiω+ t ,
1
( −1 )
√
θ− = 2 A− eiω− t .
1
Here, the upper/lower components of the vector describe θ1 /θ2 respectively.
M q̈ = −Kq,
a form that looks like the spring equation. However, q is an n−dimensional vector and M and k are
n × n matrices. In the double pendulum example, the dimensionality is 2 and the q refers to the θ1
and θ2 , and the matrices for M and K can be read off Eq. (6.24),
( ) ( )
2 1 2ω02 0
M = , K= .
1 1 0 ω02
Multiplying both sides of the equation by the inverse matrix M −1 ,
( )
q̈ = − M −1 K q.
Here,
( )
−1 1 −1
M = ,
−1 2
( )
−1 2 −1
M K = ω02 .
−2 2
One can find a transformation, basically a rotation, that transforms to a frame where M −1 K is diago-
nal. In this coordinate system the diagonal components of M −1 K represent the squared frequencies
of the normal modes, ( ) 2
ω+ 0
M −1 K → − 2 ,
0 ω−
and are known as “eigen” frequencies. The corresponding unit vectors,
( ) ( )
1 0
and in the new coordinate system,
0 1
can be rotated back into the original frame, and become the solutions for the normal modes. These
are then called “eigenvectors”, which are the same as the normal modes. Finding the eigenfrequen-
cies is performed by realizing that the determinant of a matrix is unchanged by the rotation between
coordinate systems. Writing the equations of motion as an eigenvalue problem,
[A − λi 1] ui = 0, A ≡ M −1 K, λi ≡ ωi2 . (6.25)
79
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
In the coordinate system where M −1 K is diagonal, and the forms for ui are simple this requires that
in that system, the diagonal elements of M −1 K are the eigenvalues, ωi2 . For each ωi2 , the determinant
|A − λi 1| must vanish. This is then true in any coordinate system,
2 2
One can solve a quadratic equation for λ, which gives two eigenvalues corresponding to ω+ and ω−
found above. Choosing one of the eigenvalues, one can insert one of the eigenvalues λi into Eq. (6.27)
and solve for ui , then choose the other eigenvalue and solve for the other corresponding vector.
If this were a 3-dimensional set of equations, the determinant would include terms like λ3 and would
become a cubic equation with three eigenvalues. One would then solve for three eigenvectors. If one
has a system with dimensionality n > 2, one usually resorts to solving the problem numerically due
to the messiness of the algebra. The main programming languages all have packages which readily
diagonalize matrices and find eigenvectors and eigenvalues.
After showing that H is conserved, i.e. (d/dt)H = 0, we then show that H can be identified with
the total energy, H = T + V .
One can see that H is conserved by applying first using the chain rule for (d/dt)H in Eq. (6.29), then
applying Lagrange’s equations,
d ∑ { ∂L (
d ∂L
)
∂L ∂L
}
H = q̈i + q̇i − q̈i − q̇i (6.30)
dt ∂ q̇i dt ∂ q̇i ∂ q̇i ∂qi
∑ { ∂L }
i
∂L ∂L ∂L
= q̈i + q̇i − q̈i − q̇i
i
∂ q̇i ∂qi ∂ q̇i ∂qi
= 0.
These steps assumed that L had no explicit time dependence, i.e. L is a function of q and q̇ , but not
of t.
Next, we show that L can be identified with the energy. Because V does not depend on q̇ ,
∑ ∂T
H = q̇i − T + V. (6.31)
i
∂ q̇i
80
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
The proof that H equals the energy hinged on the fact that the kinetic energy was quadratic in q̇ . This
can be attributed to time-reversal symmetry. Because the Cartesian coordinates xi do not depend on
q̇i or on time, ẋi = (∂xi /∂qj )q̇j . Thus, the kinetic energy, T = mẋ2i /2, should be proportional to
two powers of q̇ , which validates the assumption in Eq. (6.32).
Here, energy conservation is predicated on the Lagrangian not having an explicit time dependence.
Without an explicit time dependence the equations of motion are unchanged if one translates a fixed
amount in time because the physics does not depend on when the clock starts. In contrast, the absolute
time becomes relevant if there is an explicit time dependence. In fact, conservation laws can usually be
associated with symmetries. In this case the translation symmetry in time leads to energy conservation.
For another example of how symmetry leads to conservation laws, consider a Lagrangian for a particle
of mass m moving in a two-dimensional plane where the generalized coordinates are the radius r and
the angle θ . The kinetic energy would be
1 { }
T = m ṙ 2 + r 2 θ̇ 2 , (6.34)
2
and if the potential energy V (r) depends only on the radius r and not on the angle, Lagrange’s equa-
tions become
d ∂V
(mṙ) = − + mθ̇ 2 r, (6.35)
dt ∂r
d
(mr 2 θ̇) = 0.
dt
The second equation implies that mr 2 θ̇ is a constant. Indeed, it is the angular momentum which
is conserved for a radial force. Here, the conservation of angular momentum is associated with the
independence of the physics to changes in θ , or in other words, rotational invariance. Once one knows
the fact that L = mr 2 θ̇ is conserved, it can be inserted into the equations of motion for ṙ ,
∂V L2
mr̈ = − + . (6.36)
∂r mr 3
This is related to Noether’s theorem http://en.wikipedia.org/wiki/Noether’s_theorem, named after Emmy
Noether, http://en.wikipedia.org/wiki/Emmy_Noether. Simply stated, if the Lagrangian L is indepen-
dent of qi , one can see that the quantity ∂L/∂ q̇i is conserved,
d ∂L
= 0. (6.37)
dt ∂ q̇i
81
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
Another easy example is in Cartesian coordinates where the potential depends only on x and y but not
on z . In that case, there is a translational symmetry. From Eq. (6.37), this translates into conservation
of the momentum in the z direction.
Example 6.10:
Consider a pair of particles of mass m1 and m2 where the potential is of the form
U (⃗
r1 , ⃗
r2 ) = Va (|m1⃗ r2 |/(m1 + m2 )) + Vb (|⃗
r1 + m2⃗ r1 − ⃗
r2 |).
Using symmetry arguments alone, are there any conserved components of the momentum? or the an-
gular momentum??
There is no translational invariance, hence there are no conserved components of the momentum.
However, there is rotational invariance about any axis that goes through the origin. Hence, there is
angular momentum conservation in all three directions. Symmetry arguments are great ways to rec-
ognize the existence of conserved quantities, but actually expressing them in terms of coordinates can
be tricky. For instance, you may need to write the Lagrangian in terms of angles.
Example 6.11:
82
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
given y ′ (0) = 6 m/s, y(0) = 0.3 m, and given m = 0.5 kg, F0 = 400 N, and a spring constant
k = 2000 N/m. Let the driving force have a period of 2.5 s, and the damping rate be β = 0.15 s−1 .
Let’s first solve for y with indices a, b, c = −1, 0, 1, using the differential equation and the boundary
conditions. The BC state that yb = 0.3 m and (yc − ya ) = 2∆tyb′ , where yb′ = 6 m/s. Writing the
differential equation,
yc − 2yb + ya yc − ya
+β + ω02 yb = (F0 /m) cos ωtb ,
∆t2 ∆t
one can substitute for ya because the boundary condition for yb′ = y0′ gives ya = y−1 ,
ya = yc − 2∆tyb′ ,
yc − 2yb + ya yc − ya
+β + ω02 yb = (F0 /m) cos ωtb ,
∆t2 (∆t )
1 β 2yb − ya βyb
yc 2
+ = fb − ω02 yb + 2
+ ,
∆t ∆t ∆t ∆t
b −ya
fb − ω02 yb + 2y∆t 2 + βy
∆t
b
yc = 1 β
,
∆t2
+ ∆t
fb ∆t2 − ω02 yb ∆t2 + 2yb − ya + βyb ∆t
= .
1 + β∆t
Using a = 0, b = 1, c = 2, this gives y2 . Then one can repeat with a = 1, b = 2 and c = 3 to find
y3 , and iterate to find all n.
Solution:
The code might look something like this:
void main(){
const double PI=4.0*atan(1.0);
double m=0.5,k=2000.0;
double omega0=sqrt(k/m), omega=2.0*PI/2.5, beta=0.15;
83
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
6.9 Exercises
1. Consider a hill whose height y is given as a function of the horizontal coordinate x. Consider
a segment of the hill from x = 0 to x = L with initial height y(x = 0) = 0 and whose final
height is y(x = L) = −h. Transforming the last equation in Example 6.1 for a downward
vertical force rather than a horizontal force,
√
x=− −2ay − y 2 + a arccos(1 + y/a).
Consider a wheel of radius a rolling along the bottom of the x axis. Mark a point on the top of
the wheel, which is originally at the origin, x = y = 0, when the top of the wheel touches the
origin. As the wheel rolls by an angle θ the marked point moves due to both the translation and
the rotation of the wheel. The y coordinate of the marked point is
2. Consider a chain of length L that hangs from two supports of equal height stretched from x =
−X to x = +X . The general solution for a catenary is
y = λ + a cosh[(x − x0 )/a],
84
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
3. Consider a mass m connected to a spring with spring constant k. Rather than being fixed, the
other end of the spring oscillates with frequency ω and amplitude A. For a generalized coordi-
nate, use the displacement of the mass from its relaxed position and call it y = x−ℓ−A cos ωt.
In this system the potential energy of the spring is ky 2 /2.
4. Consider a bead of mass m on a circular wire of radius R. Assume a force kx acts on the bead,
where x and y axes run through the center of the circle. Using θ as the generalized coordinate
(measured relative to the x axis),
5. Consider a pendulum of length ℓ with all the mass m at its end. The pendulum is allowed to
swing freely in both directions. Using ϕ to describe the azimuthal angle about the z axis and
θ to measure the angular deviation of the pendulum from the downward direction, address the
following questions:
(a) If the pendulum is initially moving horizontally with velocity v0 and angle θ0 = 90◦
(horizontal), use energy and angular momentum conservation to find the minimum angles
of θmin subtended by the pendulum. (Note that the angle will oscillate between 90◦ and
the minimum angle.
(b) Write the Lagrangian using θ and ϕ as generalized coordinates.
(c) Write the equations of motion for θ and ϕ.
(d) Rewrite the equations of motion for θ using angular momentum conservation to eliminate
and reference to ϕ.
(e) Find the value of L required for the stable orbit to be at θ = 45◦ .
(f) For the steady orbit found in (e) consider small perturbations of the orbit. Find the fre-
quency with which the pendulum oscillates around θ = 45◦ .
6. Consider a mass m that is connected to a wall by a spring with spring constant k. A second
identical mass m is connected to the first mass by an identical spring. Motion is confined to the
x direction.
(a) Write the Lagrangian in terms of the positions of the two masses x1 and x2 .
(b) Solve for the equations of motion.
85
PHY 321 Lecture Notes 6 LAGRANGIANS AND THE CALCULUS OF VARIATIONS
x1 = Aeiωt , x2 = Beiωt .
Solve for A/B and ω . Express your answers in terms of ω02 = k/m.
r1 − ⃗
7. Consider two masses m1 and m2 interacting according to a potential V (⃗ r2 ).
⃗ cm = (m1⃗
(a) Write the Lagrangian in terms of the generalized coordinates R r1 +m2⃗
r2 )/(m1 +
m2 ) and ⃗r=⃗ r1 − ⃗
r2 , and their derivatives.
⃗ cm , find expressions for the
(b) Using the independence of the Lagrangian with respect to R
conserved total momentum.
86